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

\UseRawInputEncoding

Functoriality of Bose-Mesner algebras and profinite association schemes

Makoto Matsumoto Mathematics Program
Graduate School of Advanced Science and Engineering
Hiroshima University, 739-8526 Japan
m-mat@math.sci.hiroshima-u.ac.jp
Kento Ogawa Mathematics Program
Graduate School of Advanced Science and Engineering
Hiroshima University, 739-8526 Japan
knt-ogawa@hiroshima-u.ac.jp
 and  Takayuki Okuda Mathematics Program
Graduate School of Advanced Science and Engineering
Hiroshima University, 739-8526 Japan
okudatak@hiroshima-u.ac.jp
Abstract.

We show that taking the set of primitive idempotents of commutative association schemes is a functor from the category of commutative association schemes with surjective morphisms to the category of finite sets with surjective partial functions. We then consider projective systems of commutative association schemes consisting of surjections (which we call profinite association schemes), for which Bose-Mesner algebra is defined, and describe a Delsarte theory on such schemes. This is another method for generalizing association schemes to those on infinite sets, related with the approach by Barg and Skriganov. Relation with (t,m,s)(t,m,s)-nets and (t,s)(t,s)-sequences is studied. We reprove some of the results of Martin-Stinson from this viewpoint.

Key words and phrases:
Association Scheme, Profinite set, Delsarte theory, (t,m,s)(t,m,s)-net, (t,s)(t,s)-sequence, Ordered Hamming scheme, LP-program
2020 Mathematics Subject Classification:
05E30 Association schemes, strongly regular graphs, 20D60 Arithmetic and combinatorial problems, 20E18 Limits, profinite groups
The first author is partially supported by JSPS Grants-in-Aid for Scientific Research JP18K03213. The second author is partially supported by JST SPRING, Grant Number JPMJSP2132. The third author is partially supported by JSPS Grants-in-Aid for Scientific Research JP20K03589, JP20K14310, and JP22H01124.

1. Introduction

Association schemes are central objects in algebraic combinatorics, with many interactions with other areas of mathematics. It is natural to try to take projective limits of a system of association schemes, but there seems to be some obstruction to have a natural Bose-Mesner algebra. We show that a surjective morphism between commutative association schemes behaves well with respect to the two product structures of Bose-Mesner algebras and primitive idempotents, which gives rise to a notion of profinite association schemes and their Delsarte theory. They are closely related with profinite groups. We study kernel schemes and ordered Hamming schemes introduced by Martin-Stinson[16] as examples. We reprove their characterization of (t,m,s)(t,m,s)-net as a design in an ordered Hamming scheme, and give a characterization of (t,s)(t,s)-sequences in terms of profinite association schemes.

2. Surjections among association schemes and functoriality

Let {\mathbb{N}} denote the set of natural numbers including 0, and >0{\mathbb{N}}_{>0} the set of positive integers.

2.1. Category of association schemes

Let us recall the notion of association schemes briefly. See [1][6] for details. We summarize basic terminologies.

Definition 2.1.

Let XX be a set. By #X\#X we denote the cardinality of XX. If XX has a topology, C(X)C(X) denotes the vector space of continuous functions from XX to {\mathbb{C}}. If a topology is not specified, XX is regarded to be a discrete space. For SXS\subset X, we define the indicator function of SS as a function χS:X{0,1}\chi_{S}:X\to\{0,1\} where χ(x)=1\chi(x)=1 if and only if xSx\in S. If SS is a singleton {x}\{x\}, we denote the indicator function by δx\delta_{x}. For a map f:XYf:X\to Y between finite sets, f:C(Y)C(X)f^{\dagger}:C(Y)\to C(X) denotes the pull back: gC(Y)gfg\in C(Y)\mapsto g\circ f. It is injective (respectively surjective) if ff is surjective (respectively injective). By the multiplication of functions, C(X)C(X) is a unital commutative ring. This multiplication is called the Hadamard product. For a finite set XX, the set C(X×X)C(X\times X) is naturally identified with the set of complex valued square matrices of size #(X)\#(X), and the matrix product is given by AB(x,z)=yXA(x,y)B(y,z)AB(x,z)=\sum_{y\in X}A(x,y)B(y,z). The Hadamard product \circ is given by the component wise product, namely, (AB)(x,z)=A(x,z)B(x,z)(A\circ B)(x,z)=A(x,z)B(x,z). (Note that \circ may denote the composition of mappings, but no confusions would occur.)

Definition 2.2.

Let XX, II be finite sets, and R:X×XIR:X\times X\to I a surjection. We call (X,R,I)(X,R,I) an association scheme, if the following properties are satisfied. For each iIi\in I, R1(i)R^{-1}(i) may be regarded as a relation on XX, denoted by RiR_{i}. Let AiA_{i} be the corresponding adjacency matrix in C(X×X)C(X\times X).

  1. (1)

    There is an i0Ii_{0}\in I such that Ai0A_{i_{0}} is an identity matrix.

  2. (2)

    Consider the injection R:C(I)C(X×X)R^{\dagger}:C(I)\to C(X\times X). Its image AXA_{X} is closed by the matrix product.

  3. (3)

    AXA_{X} is closed under the transpose of C(X×X)C(X\times X).

The algebra AXA_{X} with the two multiplications (the Hadamard product and the matrix product) is called the Bose-Mesner algebra of (X,R,I)(X,R,I). We may use the same notation i0i_{0} for distinct association schemes. The number of 11 in a row of AiA_{i} is called the ii-th valency and denoted by kik_{i}. If AXA_{X} is commutative with respect to the matrix product, then (X,R,I)(X,R,I) is said to be commutative.

In the following, we write simply an “association scheme XX” for an association (X,R,I)(X,R,I) by an abuse of language. We follow MacLane [15] for the terminologies of the category theory, in particular, functors, projective systems, and limits. The association schemes form a category, by the following[11][25].

Definition 2.3.

Let (X,R,I)(X,R,I), (X,R,I)(X^{\prime},R^{\prime},I^{\prime}) be association schemes. A morphism of association schemes from XX to XX^{\prime} is a pair of mappings f:XXf:X\to X^{\prime} and g:IIg:I\to I^{\prime} such that the following diagram commutes:

X×XIf×fgX×XI.\begin{array}[]{ccc}X\times X&\to&I\\ f\times f\downarrow\phantom{f\times f}&\circlearrowleft&\phantom{g}\downarrow g\\ X^{\prime}\times X^{\prime}&\to&I^{\prime}.\\ \end{array} (2.1)

It is clear that the surjectivity of ff implies that of gg. The morphism (f,g)(f,g) is said to be surjective if ff is surjective.

2.2. Functoriality of Bose-Mesner algebra

Proposition 2.4.

If (f,g):XX(f,g):X\to X^{\prime} is a surjective morphism of association schemes, then for any xXx^{\prime}\in X^{\prime}, the cardinality of the fiber #f1(x)\#f^{-1}(x^{\prime}) is #X/#X\#X/\#X^{\prime}.

Proof.

The cardinality of the fiber is the summation of the valencies of RiR_{i} for ii with g(i)=i0g(i)=i_{0}^{\prime}, hence independent of the choice of xx^{\prime}, which implies the result. This is well-known [25], and it holds under a weaker condition (for unital regular relation partitions, see [12]). ∎

Theorem 2.5.

  1. (1)

    Let (f,g):(X,R,I)(X,R,I)(f,g):(X,R,I)\to(X^{\prime},R^{\prime},I^{\prime}) be a morphism. It induces an injection C(I)C(I)C(I^{\prime})\to C(I) and hence Ψ:AXAX\Psi:A_{X^{\prime}}\to A_{X}. Then Ψ\Psi is a morphism of unital rings with respect to the Hadamard product.

  2. (2)

    Suppose that ff is surjective. We define the convolution product X\bullet_{X} on AXA_{X} by

    (AXB)(x,z)=1#XyXA(x,y)B(y,z)(A\bullet_{X}B)(x,z)=\frac{1}{\#X}\sum_{y\in X}A(x,y)B(y,z)

    (i.e., the matrix product normalized by the factor of #X\#X). Then, AXA_{X} has a unit EX:=#XAi0E_{X}:=\#XA_{i_{0}}, and Ψ\Psi is a morphism of rings with respect to the convolution product (which may not map the unit of AXA_{X^{\prime}} to the unit of AXA_{X}).

  3. (3)

    The vector space C(X)C(X) is a left AXA_{X}-module by the following action:

    AX×C(X)C(X),(A,h)(Ah)A_{X}\times C(X)\to C(X),(A,h)\mapsto(A\bullet h)

    given by

    (Ah)(x)=1#XyXA(x,y)h(y),(A\bullet h)(x)=\frac{1}{\#X}\sum_{y\in X}A(x,y)h(y),

    and the unit EXE_{X} acts trivially. This module structure is compatible with Ψ\Psi in the sense that the following diagram commutes:

    AX×C(X)C(X)Ψ×ffAX×C(X)C(X).\begin{array}[]{ccc}A_{X}\times C(X)&\to&C(X)\\ \Psi\times f^{\dagger}\uparrow\phantom{f^{\dagger}\times f^{\dagger}}&\circlearrowleft&\phantom{g\dagger}\uparrow f^{\dagger}\\ A_{X^{\prime}}\times C(X^{\prime})&\to&C(X^{\prime}).\\ \end{array} (2.2)

    If we consider C(X)C(X)C(X^{\prime})\subset C(X), then the element Ψ(EX)AX\Psi(E_{X^{\prime}})\in A_{X} acts on C(X)C(X) as a projector C(X)C(X)C(X)\to C(X^{\prime}).

Proof.

For a morphism of finite sets XYX\to Y, the induced map C(Y)C(X)C(Y)\to C(X) is a morphism of unital rings with respect to the Hadamard product. The first statement follows from this and the commutative diagram (2.1). For the second part, take A,BAXA,B\in A_{X}^{\prime}. (In fact, one may take A,BC(X×X)A,B\in C(X^{\prime}\times X^{\prime}); the following arguments depend only on the fact that the cardinality of the fiber of ff is constant.) Then

Ψ(AXB)(x,z)\displaystyle\Psi(A\bullet_{X^{\prime}}B)(x,z) =\displaystyle= (AXB)(f(x),f(z))\displaystyle(A\bullet_{X^{\prime}}B)(f(x),f(z))
=\displaystyle= 1#XyXA(f(x),y)B(y,f(z))\displaystyle\frac{1}{\#X^{\prime}}\sum_{y^{\prime}\in X^{\prime}}A(f(x),y^{\prime})B(y^{\prime},f(z))
=\displaystyle= 1#X1#f1(y)yXA(f(x),f(y))B(f(y),f(z))\displaystyle\frac{1}{\#X^{\prime}}\frac{1}{\#f^{-1}(y^{\prime})}\sum_{y\in X}A(f(x),f(y))B(f(y),f(z))
=\displaystyle= 1#X#X#XyXΨ(A)(x,y))Ψ(B)(y,z)\displaystyle\frac{1}{\#X^{\prime}}\frac{\#X^{\prime}}{\#X}\sum_{y\in X}\Psi(A)(x,y))\Psi(B)(y,z)
=\displaystyle= 1#XyXΨ(A)(x,y))Ψ(B)(y,z)\displaystyle\frac{1}{\#X}\sum_{y\in X}\Psi(A)(x,y))\Psi(B)(y,z)
=\displaystyle= (Ψ(A)XΨ(B))(x,z).\displaystyle(\Psi(A)\bullet_{X}\Psi(B))(x,z).

For the third part, it is a routine calculation to check that AXA_{X} acts on C(X)C(X) as a unital ring. For AAXA\in A_{X^{\prime}} and hC(X)h\in C(X^{\prime}), the compatibility follows from:

f((Ah))(x)\displaystyle f^{\dagger}((A\bullet h))(x) =\displaystyle= (Ah)(f(x))\displaystyle(A\bullet h)(f(x))
=\displaystyle= 1#XyXA(f(x),y)h(y)\displaystyle\frac{1}{\#X^{\prime}}\sum_{y^{\prime}\in X^{\prime}}A(f(x),y^{\prime})h(y^{\prime})
=\displaystyle= 1#X1#f1(y)yf1(y),yXA(f(x),f(y))h(f(y))\displaystyle\frac{1}{\#X^{\prime}}\frac{1}{\#f^{-1}(y^{\prime})}\sum_{y\in f^{-1}(y^{\prime}),y^{\prime}\in X^{\prime}}A(f(x),f(y))h(f(y))
=\displaystyle= 1#X#X#XyYΨ(A)(x,y))(fh)(y)\displaystyle\frac{1}{\#X^{\prime}}\frac{\#X^{\prime}}{\#X}\sum_{y\in Y}\Psi(A)(x,y))(f^{\dagger}h)(y)
=\displaystyle= 1#XyXΨ(A)(x,y))(fh)(y)\displaystyle\frac{1}{\#X}\sum_{y\in X}\Psi(A)(x,y))(f^{\dagger}h)(y)
=\displaystyle= (Ψ(A)(fh))(x).\displaystyle(\Psi(A)\bullet(f^{\dagger}h))(x).

The compatibility implies that Ψ(EX)\Psi(E_{X^{\prime}}) trivially acts on the image of C(X)C(X^{\prime}) in C(X)C(X). For AAXA\in A_{X^{\prime}} and hC(X)h\in C(X),

(Ψ(A)h))(x)\displaystyle(\Psi(A)\bullet h))(x) =\displaystyle= 1#XyXΨ(A)(x,y)h(y)\displaystyle\frac{1}{\#X}\sum_{y\in X}\Psi(A)(x,y)h(y)
=\displaystyle= 1#XyXA(f(x),f(y))h(y),\displaystyle\frac{1}{\#X}\sum_{y\in X}A(f(x),f(y))h(y),

which depends only on f(x)f(x), and hence Ψ(A)h\Psi(A)\bullet h lies in f(C(X))C(X)f^{\dagger}(C(X^{\prime}))\subset C(X). Hence Ψ(EX)\Psi(E_{X^{\prime}}) is a projection C(X)C(X)C(X)\to C(X^{\prime}). ∎

Definition 2.6.

Let ASsurj{\operatorname{AS}}_{\operatorname{surj}} be the category of association schemes with surjective morphisms.

Corollary 2.7.

Let AlgHC{\operatorname{Alg}}_{HC} (HC means Hadamard and convolution) be the category of finite dimensional {\mathbb{C}}-vector spaces AA with:

  1. (1)

    one associative multiplication (Hadamard product) which gives a commutative unital semi-simple {\mathbb{C}}-algebra structure to AA,

  2. (2)

    one (possibly non-commutative) associative multiplication (convolution product) which gives a unital semi-simple {\mathbb{C}}-algebra structure to AA.

(Note that in this case, semi-simplicity is equivalent to that AA is a direct product of a finite number of matrix algebras over {\mathbb{C}}). Morphisms are injective {\mathbb{C}}-linear maps preserving the both two products and the unit for the Hadamard product (we don’t require preservation of unit for the convolution product).

Then, the correspondence XAXX\mapsto A_{X} gives a contravariant functor from ASsurj{\operatorname{AS}}_{\operatorname{surj}} to AlgHC{\operatorname{Alg}}_{HC}.

Proof.

It is proved that AXA_{X} has two products. Semi-simplicity is well-known, c.f. [25]. (It is enough to show that there is no nilpotent ideal, but for any nonzero AAXA\in A_{X}, the product with its unitary conjugate AA=1#XAAA\bullet A^{*}=\frac{1}{\#X}AA^{*} is not nilpotent, hence AXA_{X} is semi-simple.) Functoriality is easy to check, using Theorem 2.5. ∎

Corollary 2.8.

Let us consider the category Mod{\operatorname{Mod}} of RR-modules, whose object is a pair of a unital commutative ring RR and an RR-module MM, and a morphism from (R,M)(R,M) to (R,M)(R^{\prime},M^{\prime}) is a pair of a ring morphism f:RRf:R\to R^{\prime} and a {\mathbb{Z}}-module morphism g:MMg:M\to M^{\prime} which makes the following diagram commute:

R×MMf×ggR×MM.\begin{array}[]{ccc}R\times M&\to&M\\ f\times g\downarrow\phantom{f\times g}&\circlearrowleft&\phantom{g}\downarrow g\\ R^{\prime}\times M^{\prime}&\to&M^{\prime}.\\ \end{array}

Then, the correspondence X(AX,C(X))X\mapsto(A_{X},C(X)) is a contravariant functor from ASsurj{\operatorname{AS}}_{\operatorname{surj}} to Mod{\operatorname{Mod}}.

Proof.

The correspondence is given in Theorem 2.5. It is a contravariant functor, since each of the correspondences XAXX\mapsto A_{X}, XC(X)X\mapsto C(X) is a contravariant functor. ∎

Remark 2.9.

In [8], French constructed a sub-category of association schemes, and a covariant functor from it to the category of algebra. Our construction produces a contravariant functor, which seems to be of different nature.

2.3. Commutative association schemes and primitive idempotents

By semisimplicity and Artin-Wedderburn Theorem, any Bose-Mesner algebra AXA_{X} (with convolution product) is isomorphic to a product of matrix algebras over {\mathbb{C}}. We assume that XX is commutative. Then, AXA_{X} is, as a unital ring, isomorphic to a direct product of copies of {\mathbb{C}}. Namely, AX××=:nA_{X}\cong{\mathbb{C}}\times\cdots\times{\mathbb{C}}=:{\mathbb{C}}^{n} for some nn\in{\mathbb{N}}, and this decomposition is unique as a direct product of rings.

Proposition 2.10.

Let AA be a ring isomorphic to n{\mathbb{C}}^{n} (with componentwise multiplication). An element jAj\in A corresponding to an element of n{\mathbb{C}}^{n} with one coordinate 11 and the other coordinates 0 is called a primitive idempotent of AXA_{X}. It is characterized by the idempotent property j2=jj^{2}=j, j0j\neq 0 (this is equivalent to that the each coordinate is 0 or 11 and at least one 11 exists) and that there is no nonzero idempotent jjj^{\prime}\neq j such that jj=jjj^{\prime}=j^{\prime} holds (this says only one coordinate is 11). Any idempotent is uniquely a sum of primitive idempotents.

Definition 2.11.

Let FinSets{\operatorname{FinSets}} be the category of finite sets. Let FinSetsps{\operatorname{FinSets}}_{\operatorname{ps}} be the category of finite sets and partial surjective functions.

Recall that for sets XX and YY, a partial function ff from XX to YY consists of a pair of a subset dom(f)X{\operatorname{dom}}(f)\subset X (which may be empty) and a function f|dom(f):dom(f)Yf|_{{\operatorname{dom}}(f)}:{\operatorname{dom}}(f)\to Y. Two partial functions are equal if they have the same domain and the same function on the domain. A partial function is denoted by f:XYf:X\rightharpoonup Y. For SYS\subset Y, f1(S):={xXxdom(f),f(x)S}f^{-1}(S):=\{x\in X\mid x\in{\operatorname{dom}}(f),f(x)\in S\}. Composition gfg\circ f of partial functions f:XYf:X\rightharpoonup Y and g:YZg:Y\rightharpoonup Z has domain f1(dom(g))f^{-1}({\operatorname{dom}}(g)). A partial function ff is surjective if Im(f):={f(x)xdom(f)}Y\operatorname{Im}(f):=\{f(x)\mid x\in{\operatorname{dom}}(f)\}\subset Y is YY.

Proposition 2.12.

Let Algcs{\operatorname{Alg}}_{cs} be the category of rings isomorphic to n{\mathbb{C}}^{n} for some nn, with injective {\mathbb{C}}-linear ring morphisms which may not preserve the unit. For AAlgcsA\in{\operatorname{Alg}}_{cs}, we define J(A)J(A) as the finite set of primitive idempotents of AA. For a finite set II, we define C(I)C(I) the set of maps from II to {\mathbb{C}} as defined above. These are contravariant functors, and give contra-equivalence between Algcs{\operatorname{Alg}}_{cs} and FinSetsps{\operatorname{FinSets}}_{\operatorname{ps}}.

Proof.

Note first that the notion of f:XYf:X\rightharpoonup Y is the notion of a family of disjoint subsets SyX(yY)S_{y}\subset X\ \ (y\in Y), where Sy=f1(y)S_{y}=f^{-1}(y), and the surjectivity of ff is equivalent to that SyS_{y}\neq\emptyset for all yYy\in Y. To show that JJ is a functor, let φ:AB\varphi:A\to B be an injective {\mathbb{C}}-algebra homomorphism. For jJ(A)j\in J(A), φ(j)\varphi(j) is a nonzero idempotent. Thus, it is a non-empty sum of elements of J(B)J(B), which gives a non-empty subset SjS_{j} of J(B)J(B) by Proposition 2.10. We construct a partial surjection J(B)J(A)J(B)\rightharpoonup J(A) by: for jJ(A)j\in J(A), elements of SjS_{j} is mapped to jj. Note that if jjJ(A)j\neq j^{\prime}\in J(A), then jj=0jj^{\prime}=0 and hence SjSj=S_{j}\cap S_{j^{\prime}}=\emptyset. Since SjS_{j}\neq\emptyset, the partial function is surjective. To check the functoriality, we consider AφBϕCA\stackrel{{\scriptstyle\varphi}}{{\to}}B\stackrel{{\scriptstyle\phi}}{{\to}}C. The construction of partial surjection for φ\varphi is done by assigning to jJ(A)j\in J(A) the set SjJ(B)S_{j}\subset J(B) such that summation of elements in SjS_{j} is φ(j)\varphi(j). Similarly, for each jSjj^{\prime}\in S_{j}, we have SjJ(C)S_{j^{\prime}}\subset J(C). Thus, J(φ)J(ϕ)J(\varphi)\circ J(\phi) maps jSjSj\coprod_{j^{\prime}\in S_{j}}S_{j^{\prime}} to jj. This is the same with J(ϕφ)J(\phi\circ\varphi).

The functoriality of IC(I)I\to C(I) can be also checked in an elementary manner. It is easy to check that the two functors give contra-equivalence. We omit the detail. ∎

Definition 2.13.

Let AScsurj{\operatorname{AS}}_{\operatorname{csurj}} be a full subcategory of the commutative association schemes of ASsurj{\operatorname{AS}}_{\operatorname{surj}}, and AlgcHC{\operatorname{Alg}}_{cHC} a full subcategory of AlgHC{\operatorname{Alg}}_{HC} consisting of algebras with commutative convolution multiplication. By the remark at the beginning of Section 2.3, we have a contravariant functor AScsurjAlgcHC{\operatorname{AS}}_{\operatorname{csurj}}\to{\operatorname{Alg}}_{cHC}.

Corollary 2.14.

The composite of the three functors AScsurjAlgcHC{\operatorname{AS}}_{\operatorname{csurj}}\to{\operatorname{Alg}}_{cHC}, AlgcHCAlgcs{\operatorname{Alg}}_{cHC}\to{\operatorname{Alg}}_{\operatorname{cs}} forgetting the Hadamard product, and AlgcsFinSetsps{\operatorname{Alg}}_{\operatorname{cs}}\to{\operatorname{FinSets}}_{\operatorname{ps}} gives a covariant functor ASsurjFinSetspsAS_{\operatorname{surj}}\to{\operatorname{FinSets}}_{\operatorname{ps}}. We denote this composition functor by JJ. Thus, (X,R,I)J(X,R,I)(X,R,I)\mapsto J(X,R,I) is a functor, which corresponds XX to the sets of primitive idempotents of AXA_{X} with respect to (the normalized) convolution products.

Definition 2.15.

Let j0j_{0} denote the element of J(X,R,I)J(X,R,I), corresponding to JXJ_{X}, where JXJ_{X} denotes the matrix in C(X×X)C(X\times X) with components all 11. This element acts on C(X)C(X) as a projection to the space of constant functions. We use the symbol j0j_{0} for different association schemes, similarly to i0i_{0}.

2.4. Trivial examples and remarks

In the theory of commutative association schemes, the set II and the set JJ are considered to be “dual.” Thus, it is natural to seek the relation between JJ’s for a morphism of association schemes (while a morphism between II’s is given by definition). The previous result says that there is a natural relation, if the morphism of association schemes is surjective. We couldn’t generalize the result for general (non-surjective) morphisms.

For a surjective morphism between association schemes, the corresponding partial surjection may be a properly partial function, and may be not an injection, as follows.

Definition 2.16.

Let GG be a finite group. The triple (G,R,I)(G,R,I) where I=GI=G and R(x,y)=x1yR(x,y)=x^{-1}y is a (possibly non-commutative) association scheme called the thin scheme of GG.

Remark 2.17.

Let GG be a finite abelian group, and f:GHf:G\to H be a surjective homomorphism of abelian groups. Then we have the thin scheme of GG and the thin scheme of HH, and ff is a surjective morphism of the association schemes. In this case, JJ for GG is naturally isomorphic to the character group GG^{\vee}, and one can check that the functor JJ gives GHG^{\vee}\rightharpoonup H^{\vee}, which comes from the natural inclusion ι:HG\iota:H^{\vee}\to G^{\vee} (xGx\in G^{\vee} maps to yHy\in H^{\vee} if ι(y)=x\iota(y)=x, and xx is outside the domain if there is no such yy). This gives an example that J(X)J(X)J(X)\rightharpoonup J(X^{\prime}) is not a map but a partial function.

Remark 2.18.

For any association scheme (X,R,I)(X,R,I), the mappings f=idXf={\operatorname{id}}_{X} and g:I{0,1}g:I\to\{0,1\}, g(i0)=0g(i_{0})=0 and g(i)=1g(i)=1 if ii0i\neq i_{0}, gives a morphism of association schemes (X,R,I)(X,gR,{0,1})(X,R,I)\to(X,g\circ R,\{0,1\}). In this case, J(X,gR,{0,1})J(X,g\circ R,\{0,1\}) consists of j0j_{0} and EXj0E_{X}-j_{0}. If (X,R,I)(X,R,I) is commutative, then the image of jJ(X,R,I)j\in J(X,R,I) in J(X,gR,{0,1})J(X,g\circ R,\{0,1\}) is j0j_{0} if j=j0j=j_{0}, and EXj0E_{X}-j_{0} if jj0j\neq j_{0}. This gives an example that J(X)J(X)J(X)\rightharpoonup J(X^{\prime}) is not injective.

3. Profinite association schemes

3.1. Projective system of association schemes

We recall the notion of projective system.

Definition 3.1.

(Projective system)

We work in a fixed category 𝒞{\mathcal{C}}. Let Λ\Lambda be a directed ordered set, namely, Λ\Lambda is non-empty and for any α,βΛ\alpha,\beta\in\Lambda, there is a γΛ\gamma\in\Lambda with αγ\alpha\leq\gamma and βγ\beta\leq\gamma. Let Xλ(λΛ)X_{\lambda}(\lambda\in\Lambda) be a family of objects. For any λλ\lambda\leq\lambda^{\prime}, a morphism pλ,λ:XλXλp_{\lambda^{\prime},\lambda}:X_{\lambda^{\prime}}\to X_{\lambda} is specified, and they satisfy the commutativity condition

pλ,λpλ′′,λ=pλ′′,λp_{\lambda^{\prime},\lambda}p_{\lambda^{\prime\prime},\lambda^{\prime}}=p_{\lambda^{\prime\prime},\lambda}

for any λλλ′′\lambda\leq\lambda^{\prime}\leq\lambda^{\prime\prime}, and pλ,λ=idXλp_{\lambda,\lambda={\operatorname{id}}_{X_{\lambda}}}. Then the system (Xλ,pλ,λ)(X_{\lambda},p_{\lambda,\lambda^{\prime}}) is called a projective system in 𝒞{\mathcal{C}}. The morphisms pλ,λp_{\lambda^{\prime},\lambda} are called the structure morphisms of the projective system. The dual notion (i.e. the direction of morphism is inverted in the definition) is called an inductive system.

Let XλX_{\lambda} be a projective system of association schemes. Then, each of XλX_{\lambda}, IλI_{\lambda} is a projective system of finite sets. If all morphisms are surjective, then by Theorem 2.5, we have an inductive system of Bose-Mesner algebras AXλA_{X_{\lambda}}, and if they are commutative, by Corollary 2.14 a projective system JXλ:=J(Xλ,Rλ,Iλ)J_{X_{\lambda}}:=J(X_{\lambda},R_{\lambda},I_{\lambda}) (with structure morphisms being partially surjective maps).

Definition 3.2.

Let Λ\Lambda be a directed ordered set. A projective system (Xλ,Rλ,Iλ)(X_{\lambda},R_{\lambda},I_{\lambda}) (λΛ)(\lambda\in\Lambda) in ASsurj{\operatorname{AS}}_{\operatorname{surj}} is called a profinite association scheme. (This includes the notion of usual association scheme, as the case where #Λ=1\#\Lambda=1.) We define profinite sets (see Lemma 3.4) X:=limXλ{{{X}^{{\kern-1.04996pt}\wedge}}}:=\varprojlim X_{\lambda} and I:=limIλ{{{I}^{{\kern-1.04996pt}\wedge}}}:=\varprojlim I_{\lambda} with projective limit topologies. A commutative profinite association scheme is a projective system of commutative association schemes. In this case, we define J{{{J}^{{\kern-1.04996pt}\wedge}}} similarly: an element j^J\hat{j}\in{{{J}^{{\kern-1.04996pt}\wedge}}} is the set of pair λΛ\lambda\in\Lambda and compatible family of elements jμJμj_{\mu}\in J_{\mu} for all μλ\mu\geq\lambda, divided by the equivalence relation (λ,jμ)(λ,jμ)(\lambda,j_{\mu})\sim(\lambda^{\prime},j^{\prime}_{\mu^{\prime}}) defined by the existence of λ′′λ\lambda^{\prime\prime}\geq\lambda, λ′′λ\lambda^{\prime\prime}\geq\lambda^{\prime} such that for all μλ′′\mu\geq\lambda^{\prime\prime} jμ=jμj_{\mu}=j^{\prime}_{\mu} holds. An open basis of J{{{J}^{{\kern-1.04996pt}\wedge}}} is given by the inverse image of jλJλj_{\lambda}\in J_{\lambda} through JJλ{{{J}^{{\kern-1.04996pt}\wedge}}}\rightharpoonup J_{\lambda}.

Proposition 3.3.

The natural projections XXλ{{{X}^{{\kern-1.04996pt}\wedge}}}\to X_{\lambda} and IIλ{{{I}^{{\kern-1.04996pt}\wedge}}}\to I_{\lambda} are surjective, the induced map R:X×XI{{{R}^{{\kern-1.04996pt}\wedge}}}:{{{X}^{{\kern-1.04996pt}\wedge}}}\times{{{X}^{{\kern-1.04996pt}\wedge}}}\to{{{I}^{{\kern-1.04996pt}\wedge}}} is surjective, and JJλ{{{J}^{{\kern-1.04996pt}\wedge}}}\rightharpoonup J_{\lambda} is a partial surjection.

The first two statements are well-known for projective systems of compact Hausdorff spaces with continuous morphisms, [22, Proposition 1.1.10, Lemma 1.1.5]. We need to prove the third statement, for partial surjections. It follows from the first statement, but for the self-containedness and for explaining the natural (projective limit) topology, we give a proof for the both.

Lemma 3.4.

  1. (1)

    Let XλX_{\lambda} be a projective system of finite sets and surjections. Define X{{{X}^{{\kern-1.04996pt}\wedge}}} as limXλ\varprojlim X_{\lambda}. Then, the natural map XXλ{{{X}^{{\kern-1.04996pt}\wedge}}}\to X_{\lambda} is surjective. The set X{{{X}^{{\kern-1.04996pt}\wedge}}} is called a profinite set, and has a natural topology, called profinite topology, which is compact Hausdorff. For every xλXλx_{\lambda}\in X_{\lambda}, its inverse image in X{{{X}^{{\kern-1.04996pt}\wedge}}} is clopen, and these form an open basis.

  2. (2)

    Let JλJ_{\lambda} be a projective system of finite sets and partial surjections. Define J{{{J}^{{\kern-1.04996pt}\wedge}}} as above. Then, the natural partial map JJλ{{{J}^{{\kern-1.04996pt}\wedge}}}\rightharpoonup J_{\lambda} is a partial surjection. The set J{{{J}^{{\kern-1.04996pt}\wedge}}} has a natural topology, which is locally compact and Hausdorff. For every jλJλj_{\lambda}\in J_{\lambda}, its inverse image in J{{{J}^{{\kern-1.04996pt}\wedge}}} is clopen, and these form an open basis.

  3. (3)

    Let XλX_{\lambda}, YλY_{\lambda} be projective systems of finite sets over the same directed ordered set Λ\Lambda, and let fλ:XλYλf_{\lambda}:X_{\lambda}\to Y_{\lambda} be a family of maps which commute with the structure morphisms. This induces a continuous map

    f:limXλlimYλ.{{{f}^{{\kern-1.04996pt}\wedge}}}:\varprojlim X_{\lambda}\to\varprojlim Y_{\lambda}.

    If every fλf_{\lambda} is surjective, then f{{{f}^{{\kern-1.04996pt}\wedge}}} is surjective.

Proof.

(1): Recall the construction of the projective limit limXλ\varprojlim X_{\lambda} in the category of set. It is a subset of the direct product λΛXλ\prod_{\lambda\in\Lambda}X_{\lambda} defined as the intersection of

Sμ,μ:={(sλ)λλΛXλsμ=pμ,μ(sμ)}S_{\mu^{\prime},\mu}:=\{(s_{\lambda})_{\lambda}\in\prod_{\lambda\in\Lambda}X_{\lambda}\mid s_{\mu}=p_{\mu^{\prime},\mu}(s_{\mu^{\prime}})\}

for all μ,μΛ\mu^{\prime},\mu\in\Lambda with μμ\mu^{\prime}\geq\mu. (This means that (sλ)λ(s_{\lambda})_{\lambda} belongs to limXλ\varprojlim X_{\lambda} if and only if they satisfy pμ,μ(sμ)=sμp_{\mu^{\prime},\mu}(s_{\mu^{\prime}})=s_{\mu} for every μμ\mu^{\prime}\geq\mu.) Each finite set is equipped with the discrete topology, which is compact and Hausdorff. By Tychonoff’s theorem, the product is compact and Hausdorff. Because each component is Hausdorff, Sμ,μS_{\mu^{\prime},\mu} is a closed subset, and the intersection limXλ\varprojlim X_{\lambda} is compact and Hausdorff. Take any xδXδx_{\delta}\in X_{\delta}. We shall show that this is in the image from limXλ\varprojlim X_{\lambda}. For each αΛ\alpha\in\Lambda, αδ\alpha\geq\delta, we consider the closed set

Tα:={(sλ)λλΛXλsδ=xδ,sβ=pα,β(sα) for all βα}.T_{\alpha}:=\{(s_{\lambda})_{\lambda}\in\prod_{\lambda\in\Lambda}X_{\lambda}\mid s_{\delta}=x_{\delta},s_{\beta}=p_{\alpha,\beta}(s_{\alpha})\mbox{ for all }\beta\leq\alpha\}.

There is an element yαXαy_{\alpha}\in X_{\alpha} such that pα,δ(yα)=xδp_{\alpha,\delta}(y_{\alpha})=x_{\delta}, since pp is surjective. Then, there is a system pα,β(yα)p_{\alpha,\beta}(y_{\alpha}) which is an element of TαT_{\alpha}, and thus TαT_{\alpha} is nonempty. Now we take the intersection of TαT_{\alpha} for all αδ\alpha\geq\delta. For any finite number of αi\alpha_{i}, we may take γ\gamma as an upper bound of all αi\alpha_{i}. Then the finite intersection contains TγT_{\gamma}, which is nonempty. By compactness, the intersection of TαT_{\alpha} for αδ\alpha\geq\delta is nonempty. Take an element (gλ)λ(g_{\lambda})_{\lambda} in the intersection. It lies in limXλ\varprojlim X_{\lambda}, and it projects to xδx_{\delta} via limXλXδ\varprojlim X_{\lambda}\to X_{\delta}. From the definition of the direct product topology, the inverse image of xλXλx_{\lambda}\in X_{\lambda} for various λ\lambda forms an open basis. They are clopen, since xXλx\in X_{\lambda} is clopen.

(2): Choose an element \star which is contained in no JλJ_{\lambda}’s. Let JλJ_{\lambda}^{\star} be Jλ{}J_{\lambda}\cup\{\star\}. For a partial surjection p:JλJμp:J_{\lambda}\rightharpoonup J_{\mu}, we associate a surjection p:JλJμp^{\star}:J_{\lambda}^{\star}\to J_{\mu}^{\star} as follows. For xdom(p)x\in{\operatorname{dom}}(p), p(x)=p(x)p^{\star}(x)=p(x). For xdom(p)x\notin{\operatorname{dom}}(p), p(x)=p^{\star}(x)=\star. In particular, p()=p^{\star}(\star)=\star. Let us take the projective limit as in the previous case to obtain J:=limJλ{{{J}^{{\kern-1.04996pt}\wedge}}}^{\star}:=\varprojlim J_{\lambda}^{\star}. Then for any λ\lambda, by (1), we have a surjection

pλ:JJλ.p^{\star}_{\lambda}:{{{J}^{{\kern-1.04996pt}\wedge}}}^{\star}\to J_{\lambda}^{\star}.

It is clear that there is a unique element ^J\widehat{\star}\in{{{J}^{{\kern-1.04996pt}\wedge}}}^{\star} that is mapped to \star for every JλJ_{\lambda}^{\star}. We define J:=(limJλ){^}{{{J}^{{\kern-1.04996pt}\wedge}}}^{\prime}:=(\varprojlim J_{\lambda}^{\star})\setminus\{\widehat{\star}\}, and a partial surjection pλ:JJλp_{\lambda}:{{{J}^{{\kern-1.04996pt}\wedge}}}^{\prime}\rightharpoonup J_{\lambda} by: for xJx\in{{{J}^{{\kern-1.04996pt}\wedge}}}^{\prime}, xx is outside of the domain of pλp_{\lambda} if pλ(x)=p^{\star}_{\lambda}(x)=\star, and otherwise pλ(x)=pλ(x)p_{\lambda}(x)=p^{\star}_{\lambda}(x). It is easy to check that pλp_{\lambda} is a partial surjection (using the surjectivity of pλp^{\star}_{\lambda}), and pλp_{\lambda} for various λ\lambda is compatible with the structure morphisms pμ,μp_{\mu^{\prime},\mu}.

Since J{{{J}^{{\kern-1.04996pt}\wedge}}}^{\star} is compact Hausdorff, its open subset J{{{J}^{{\kern-1.04996pt}\wedge}}}^{\prime} is locally compact Hausdorff, and has an open basis as in the case (1). We need to prove that J{{{J}^{{\kern-1.04996pt}\wedge}}} in Definition 3.2 is canonically isomorphic to J{{{J}^{{\kern-1.04996pt}\wedge}}}^{\prime}. Let J~\tilde{J} be the set of pairs (λ,(jμ)μ)(\lambda,(j_{\mu})_{\mu}) where (jμ)μ(j_{\mu})_{\mu} is a compatible system for μλ\mu\geq\lambda. For such a pair, we assign an element of J{{{J}^{{\kern-1.04996pt}\wedge}}}^{\star} as follows. For any αΛ\alpha\in\Lambda, take any β\beta with βα\beta\geq\alpha and βλ\beta\geq\lambda. Define jαj_{\alpha} as pβ,α(jμ)p^{\star}_{\beta,\alpha}(j_{\mu}). This is well-defined by the commutativity of the structure morphisms, and gives a map f:J~Jf:\tilde{J}\to{{{J}^{{\kern-1.04996pt}\wedge}}}^{\prime}. This is surjective. Two pairs (λ,(jμ)μ)(\lambda,(j_{\mu})_{\mu}), (λ,(jμ)μ)(\lambda^{\prime},(j^{\prime}_{\mu^{\prime}})_{\mu^{\prime}}) have the same image if and only if there is a λ′′\lambda^{\prime\prime} with λ′′λ\lambda^{\prime\prime}\geq\lambda, λ′′λ\lambda^{\prime\prime}\geq\lambda^{\prime} such that for any μλ′′\mu\geq\lambda^{\prime\prime}, jμ=jμj_{\mu}=j^{\prime}_{\mu} holds. This gives a bijection J:=J~/J{{{J}^{{\kern-1.04996pt}\wedge}}}:=\tilde{J}/\sim\to{{{J}^{{\kern-1.04996pt}\wedge}}}^{\prime}.

(3): See [22, Lemma 1.1.5]. ∎

Remark 3.5.

The above construction gives a category equivalence between the category of sets with partial maps and the category of sets with one base point. Using this, it is not difficult to show that J{{{J}^{{\kern-1.04996pt}\wedge}}} is the projective limit in the category of sets with partial surjections.

Later it will be proved that, for any commutative profinite association scheme, J{{{J}^{{\kern-1.04996pt}\wedge}}} has the discrete topology (Proposition 3.10).

Definition 3.6.

For a profinite association scheme (Xλ,Rλ,Iλ)(X_{\lambda},R_{\lambda},I_{\lambda}), its Bose-Mesner algebra is defined by AX:=limAXλA_{{{{X}^{{\kern-0.74997pt}\wedge}}}}:=\varinjlim A_{X_{\lambda}}. For a commutative profinite association scheme, through the natural isomorphism C(Jλ)AXλC(J_{\lambda})\to A_{X_{\lambda}}, we have an isomorphism limC(Jλ)AX\varinjlim C(J_{\lambda})\to A_{{{{X}^{{\kern-0.74997pt}\wedge}}}} which maps the point-wise product to the convolution product, and an isomorphism limC(Iλ)AX\varinjlim C(I_{\lambda})\to A_{{{{X}^{{\kern-0.74997pt}\wedge}}}} which maps the point-wise product to the Hadamard product.

Let C(X×X)C({{{X}^{{\kern-1.04996pt}\wedge}}}\times{{{X}^{{\kern-1.04996pt}\wedge}}}) denote the set of complex valued continuous functions on X×X{{{X}^{{\kern-1.04996pt}\wedge}}}\times{{{X}^{{\kern-1.04996pt}\wedge}}} (and similarly C(I)C({{{I}^{{\kern-1.04996pt}\wedge}}}) be the set of continuous functions on I{{{I}^{{\kern-1.04996pt}\wedge}}}). Since we have inductive families of injections

AXλC(Xλ×Xλ)C(X×X),A_{X_{\lambda}}\to C(X_{\lambda}\times X_{\lambda})\to C({{{X}^{{\kern-1.04996pt}\wedge}}}\times{{{X}^{{\kern-1.04996pt}\wedge}}}),

by universality, we have a canonical injective morphism AXC(X×X)A_{{{{X}^{{\kern-0.74997pt}\wedge}}}}\to C({{{X}^{{\kern-1.04996pt}\wedge}}}\times{{{X}^{{\kern-1.04996pt}\wedge}}}). similarly, we have an injection limC(Iλ)C(I)\varinjlim C(I_{\lambda})\to C({{{I}^{{\kern-1.04996pt}\wedge}}}).

Definition 3.7.

For jλJλj_{\lambda}\in J_{\lambda}, we denote by EjλAXλE_{j_{\lambda}}\in A_{X_{\lambda}} the primitive idempotent jλj_{\lambda} (the notation introduced to distinguish an element of JJ with an element of AXλA_{X_{\lambda}}). It is an element of C(Xλ×Xλ)C(X_{\lambda}\times X_{\lambda}), and acts on C(Xλ)C(X_{\lambda}) by the normalized convolution defined in Theorem 2.5. Let C(Xλ)jλC(X_{\lambda})_{j_{\lambda}} be the image of EjλE_{j_{\lambda}} by this action.

Proposition 3.8.

Take jλJλj_{\lambda}\in J_{\lambda}. For μλ\mu\geq\lambda, we may regard C(Xλ)C(Xμ)C(X_{\lambda})\subset C(X_{\mu}). Then Ψ(Ejλ)\Psi(E_{j_{\lambda}}) given in Theorem 2.5 acts on gC(Xμ)g\in C(X_{\mu}) by Ψ(Ejλ)g\Psi(E_{j_{\lambda}})\bullet g. The operator Ψ(Ejλ)\Psi(E_{j_{\lambda}}) gives a splitting C(Xμ)C(Xλ)jλC(X_{\mu})\to C(X_{\lambda})_{j_{\lambda}} to C(Xλ)jλC(Xμ)C(X_{\lambda})_{j_{\lambda}}\subset C(X_{\mu}).

Proof.

By Theorem 2.5, Ψ(EXλ)End(C(Xμ))\Psi(E_{X_{\lambda}})\in{{\operatorname{End}}}(C(X_{\mu})) is a projection to C(Xλ)C(X_{\lambda}). Then

Ψ(Ejλ)=Ψ(EjλEXλ)=Ψ(Ejλ)Ψ(EXλ)\Psi(E_{j_{\lambda}})=\Psi(E_{j_{\lambda}}\bullet E_{X_{\lambda}})=\Psi(E_{j_{\lambda}})\bullet\Psi(E_{X_{\lambda}})

is a projection C(Xμ)C(Xλ)C(Xλ)jλC(X_{\mu})\to C(X_{\lambda})\to C(X_{\lambda})_{j_{\lambda}}. ∎

Definition 3.9.

Let jj be an element of J{{{J}^{{\kern-1.04996pt}\wedge}}}. Then jj is said to be isolated at λ\lambda to jλj_{\lambda}, if there is a jλJλj_{\lambda}\in J_{\lambda} such that the inverse image of jλj_{\lambda} is the singleton {j}\{j\}.

Proposition 3.10.

The cardinality of the inverse image of jλj_{\lambda} in JμJ_{\mu} for μλ\mu\geq\lambda does not exceed the dimension of C(Xλ)jλC(X_{\lambda})_{j_{\lambda}}, which is finite. Consequently, for each jJj\in{{{J}^{{\kern-1.04996pt}\wedge}}}, there exists a λ\lambda at which jj is isolated. Thus, J{{{J}^{{\kern-1.04996pt}\wedge}}} has the discrete topology.

Proof.

Suppose that the inverse image of jλj_{\lambda} in JμJ_{\mu} is {j1,j2,,jn}\{j_{1},j_{2},\dots,j_{n}\}. Then the operator Ψ(Ejλ)=i=1nΨ(Eji)\Psi(E_{j_{\lambda}})=\sum_{i=1}^{n}\Psi(E_{j_{i}}) on C(Xμ)C(X_{\mu}) is a projection to C(Xλ)jλC(X_{\lambda})_{j_{\lambda}}. Thus C(Xμ)jiC(X_{\mu})_{j_{i}} (i=1,,n)(i=1,\ldots,n) are nonzero direct summands of C(Xλ)jλC(X_{\lambda})_{j_{\lambda}}, and hence nn does not exceed the dimension of C(Xλ)jλC(X_{\lambda})_{j_{\lambda}}. This means that for any μλ\mu\geq\lambda, the cardinality of the inverse image of jλj_{\lambda} is bounded, and hence there is a μ\mu^{\prime} such that the cardinality is constant for all μμ\mu\geq\mu^{\prime}. Let us fix jJj\in{{{J}^{{\kern-1.04996pt}\wedge}}}. Take any λ0\lambda_{0} such that jj is in the domain of the partial map to Jλ0J_{\lambda_{0}}. We put jλ0j_{\lambda_{0}} to the image of jj. Then by the arguments above, we can find a λ\lambda such that the cardinality of the inverse image of jλ0j_{\lambda_{0}} in JλJ_{\lambda^{\prime}} is constant for all λλ\lambda^{\prime}\geq\lambda. We put jλj_{\lambda} to the image of jj in JλJ_{\lambda}. Then jλj_{\lambda} is in the inverse image of jλ0j_{\lambda_{0}} and the cardinality of the inverse image of jλj_{\lambda} in JλJ_{\lambda^{\prime}} should be one for any λλ\lambda^{\prime}\geq\lambda. This implies that the inverse image of jλj_{\lambda} in J{{{J}^{{\kern-1.04996pt}\wedge}}} is just {j}\{j\}, namely, jj is isolated at λ\lambda. We have proved that {j}\{j\} is a clopen subset of J{{{J}^{{\kern-1.04996pt}\wedge}}} for any jJj\in{{{J}^{{\kern-1.04996pt}\wedge}}}, and hence J{{{J}^{{\kern-1.04996pt}\wedge}}} has the discrete topology. ∎

Corollary 3.11.

We have limC(Jλ)=jJj\varinjlim C(J_{\lambda})=\bigoplus_{j\in{{{J}^{{\kern-0.74997pt}\wedge}}}}{\mathbb{C}}\cdot j. We write EjE_{j} for the image of jJj\in{{{J}^{{\kern-1.04996pt}\wedge}}} in AXA_{{{X}^{{\kern-0.74997pt}\wedge}}}. Then jJEj=AX\bigoplus_{j\in{{{J}^{{\kern-0.74997pt}\wedge}}}}{\mathbb{C}}\cdot E_{j}=A_{{{X}^{{\kern-0.74997pt}\wedge}}}. We may use the notation Cc(J)C_{c}({{{J}^{{\kern-1.04996pt}\wedge}}}) for limC(Jλ)\varinjlim C(J_{\lambda}), since it is the set of functions (automatically continuous) on J{{{J}^{{\kern-1.04996pt}\wedge}}} with finite (==compact) support. Through the isomorphism Cc(J)AXC_{c}({{{J}^{{\kern-1.04996pt}\wedge}}})\to A_{{{X}^{{\kern-0.74997pt}\wedge}}}, EjAXE_{j}\in A_{{{X}^{{\kern-0.74997pt}\wedge}}} corresponds to the indicator function δjCc(J)\delta_{j}\in C_{c}({{{J}^{{\kern-1.04996pt}\wedge}}}).

Remark 3.12.

In the Pontryagin duality, the dual of a compact abelian group is a discrete group. Since I{{{I}^{{\kern-1.04996pt}\wedge}}} is profinite and compact, the above discreteness of J{{{J}^{{\kern-1.04996pt}\wedge}}} is an analogue to this fact.

We summarize the results of this subsection, to compare with the case of finite commutative association schemes.

Theorem 3.13.

Let XλX_{\lambda} be a commutative profinite association scheme. Notations are as in Definition 3.6 and Corollary 3.11. We have the following commutative diagram, where all the vertical arrows are injections:

C(J)C(X×X)C(I)Cc(J)=limC(Jλ)AXlimC(Iλ)=Clc(I),\begin{array}[]{ccccccccc}&&C({{{J}^{{\kern-1.04996pt}\wedge}}})&&C({{{X}^{{\kern-1.04996pt}\wedge}}}\times{{{X}^{{\kern-1.04996pt}\wedge}}})&\leftarrow&C({{{I}^{{\kern-1.04996pt}\wedge}}})&&\\ &&\uparrow&&\uparrow&&\uparrow&&\\ C_{c}({{{J}^{{\kern-1.04996pt}\wedge}}})&=&\varinjlim C(J_{\lambda})&\simeq&A_{{{{X}^{{\kern-0.74997pt}\wedge}}}}&\simeq&\varinjlim C(I_{\lambda})&=&C_{lc}({{{I}^{{\kern-1.04996pt}\wedge}}}),\\ \end{array} (3.1)

where:

  1. (1)

    The set J{{{J}^{{\kern-1.04996pt}\wedge}}} is discrete. The image of the left vertical arrow is the set of compact support functions Cc(J)C_{c}({{{J}^{{\kern-1.04996pt}\wedge}}}) (i.e. takes 0 except but finite points in J{{{J}^{{\kern-1.04996pt}\wedge}}}).

  2. (2)

    The image of the right vertical arrow is the set of locally constant functions Clc(I)C_{lc}({{{I}^{{\kern-1.04996pt}\wedge}}}), which is dense with respect to the supremum norm, see the next lemma.

Lemma 3.14.

Let X{{{X}^{{\kern-1.04996pt}\wedge}}} be the projective limit of a projective system of finite sets XλX_{\lambda} for a directed ordered set Λ\Lambda.

  1. (1)

    Let CC be a compact subspace of X{{{X}^{{\kern-1.04996pt}\wedge}}}, and OaO_{a} (aA)(a\in A) be an open covering of CC. Then, there exists a μΛ\mu\in\Lambda such that by putting SS to the image of CC in XμX_{\mu}, the inverse image of sSs\in S in X{{{X}^{{\kern-1.04996pt}\wedge}}} constitutes a (finite and disjoint) open covering of CC, and each of these open sets is contained in one of OaO_{a}.

  2. (2)

    The image of limC(Xλ)C(X)\varinjlim C(X_{\lambda})\to C({{{X}^{{\kern-1.04996pt}\wedge}}}) is the set of locally constant functions on X{{{X}^{{\kern-1.04996pt}\wedge}}}, and is dense with respect to the supremum norm.

Proof.

(1): For xλXλx_{\lambda}\in X_{\lambda}, we denote by BxλB_{x_{\lambda}} the inverse image of xλx_{\lambda} in X{{{X}^{{\kern-1.04996pt}\wedge}}}, which we call a λ\lambda-ball. This gives an open basis, and thus every OaO_{a} is a union of λ\lambda-balls for various λ\lambda. Since CC is compact and covered by these balls, we may find a finite number of these balls which covers CC and each contained in some OaO_{a}. We take an upper bound μ\mu for these finite number of λ\lambda. Let SS be the image of CC in XμX_{\mu}. Then, Bsμ(sμS)B_{s_{\mu}}(s_{\mu}\in S) are finite, clopen, and mutually disjoint balls covering CC, such that each ball is contained in some OaO_{a}.

(2) Let ff be a locally constant function. Then, for any xXx\in{{{X}^{{\kern-1.04996pt}\wedge}}}, there is an open neighborhood OxO_{x} so that ff is constant on OxO_{x}. These constitute an open covering of X{{{X}^{{\kern-1.04996pt}\wedge}}}. Applying (1) for C=XC={{{X}^{{\kern-1.04996pt}\wedge}}} to find μ\mu with S=XμS=X_{\mu}, where each μ\mu-ball is contained in some OxO_{x}, and hence ff is constant on each ball. This means that ff is an image of an element in C(Xμ)C(X_{\mu}). Conversely, any function in C(Xμ)C(X_{\mu}) clearly maps to a locally constant function on X{{{X}^{{\kern-1.04996pt}\wedge}}}.

We shall show the density. For any fC(X)f\in C({{{X}^{{\kern-1.04996pt}\wedge}}}) and ϵ>0\epsilon>0, by continuity, for all xXx\in{{{X}^{{\kern-1.04996pt}\wedge}}} we have an open neighborhood OxO_{x} such that yOxy\in O_{x} implies |f(x)f(y)|<ϵ|f(x)-f(y)|<\epsilon. By applying (1) for C=XC={{{X}^{{\kern-1.04996pt}\wedge}}} to find a μ\mu with S=XμS=X_{\mu}. We see that the μ\mu-balls cover X{{{X}^{{\kern-1.04996pt}\wedge}}}, and for each xμXμx_{\mu}\in X_{\mu}, we define g(xμ)=f(x)g(x_{\mu})=f(x^{\prime}) by choosing any xx^{\prime} in BxμB_{x_{\mu}}. Denote the projection by pr:XXμ{\operatorname{pr}}:{{{X}^{{\kern-1.04996pt}\wedge}}}\to X_{\mu}. Then for any yXy\in{{{X}^{{\kern-1.04996pt}\wedge}}}, g(pr(y))=f(x)g({\operatorname{pr}}(y))=f(x^{\prime}) for some xBpr(y)x^{\prime}\in B_{{\operatorname{pr}}(y)}. Since Bpr(y)OxB_{{\operatorname{pr}}(y)}\subset O_{x} for some xx, the values of ff in this ball differ at most by 2ϵ2\epsilon, namely,

|g(pr(y))f(y)|=|f(x)f(y)|<2ϵ.|g({\operatorname{pr}}(y))-f(y)|=|f(x^{\prime})-f(y)|<2\epsilon.

This shows that

pr(g)fsup<2ϵ,||{\operatorname{pr}}^{\dagger}(g)-f||_{\sup}<2\epsilon,

which is the desired density. ∎

3.2. Inner product and orthogonality

Proposition 3.15.

For a profinite association scheme, X{{{X}^{{\kern-1.04996pt}\wedge}}} has a natural probability regular Borel measure μ\mu, whose push-forward is the normalized (probability) counting measure on XλX_{\lambda} for each λ\lambda. If X{{{X}^{{\kern-1.04996pt}\wedge}}} is a profinite group GG considered as a projective system of G/NλG/N_{\lambda} for (finite index) open normal subgroups, then the above μ\mu coincides with the probability Haar measure.

Proof.

This is a direct consequence of Choksi [5] and a general property of measures (see Halmos [10]), as follows. Each XλX_{\lambda} is finite discrete (hence compact Hausdorff) with normalized counting measure, and the measure is preserved by the uniformness of the cardinality of each fiber as in Proposition 2.4. This means that the system is an inverse family of measured spaces [5, Definition 2], and hence X{{{X}^{{\kern-1.04996pt}\wedge}}} has a weak measure structure: a set MM of ring of sets and a function μ\mu from MM to non-negative real numbers, such that XXλ{{{X}^{{\kern-1.04996pt}\wedge}}}\to X_{\lambda} preserves the measure. In this case, MM is the set of open sets, and μ(X)=1\mu({{{X}^{{\kern-1.04996pt}\wedge}}})=1. Then [5, Theorem 2.2] tells that this μ\mu on X{{{X}^{{\kern-1.04996pt}\wedge}}} is extended to a measure on σ\sigma-ring generated by MM, namely, the Borel sets. There it is also proved that for any measurable set EE with finite measure and any ϵ>0\epsilon>0, there is a compact set CEC\subset E with μ(E)<μ(C)+ϵ\mu(E)<\mu(C)+\epsilon. This proves the inner regularity of μ\mu ([10, §52, P.224]). By taking the complement, for any ϵ>0\epsilon>0 and measurable set EcE^{c}, there is an open UEcU\supset E^{c} with 1μ(Ec)<1μ(U)+ϵ1-\mu(E^{c})<1-\mu(U)+\epsilon, i.e., μ(Ec)>μ(U)ϵ\mu(E^{c})>\mu(U)-\epsilon, which proves the outer regularity of μ\mu, and hence μ\mu is regular.

Let ν\nu be the probability Haar measure of GG. Then, for any (finite index) open normal subgroup NN, ν(N)\nu(N) is 1#(G/N)\frac{1}{\#(G/N)}, since #(G/N)\#(G/N) cosets of NN disjointly covers GG. This gives an inverse family of measured spaces, and the above construction gives a regular Borel measure μ\mu. The Haar measure is regular [10, §58, Theorem B]. For the coset gNgN, μ(gN)=ν(gN)\mu(gN)=\nu(gN). The cosets gNgN for gg and NN varying make an open basis. By the outer regularity, for any compact set EE and any ϵ>0\epsilon>0, there is an open set OEO\supset E with μ(E)>μ(O)ϵ\mu(E)>\mu(O)-\epsilon. Use Lemma 3.14 for C=EC=E and the open covering OO, to find μ\mu, SXμS\subset X_{\mu}. Then Bsμ(sμS)B_{s_{\mu}}(s_{\mu}\in S) constitutes an open cover 𝒞{\mathcal{C}} of EE consisting of a finite number of disjoint cosets (contained in OO) such that the sum of the measure of each coset in 𝒞{\mathcal{C}}, denoted by μ(𝒞)\mu({\mathcal{C}}) (μ(𝒞)μ(O))(\mu({\mathcal{C}})\leq\mu(O)), satisfying

μ(𝒞)μ(E)μ(𝒞)ϵ,\mu({\mathcal{C}})\geq\mu(E)\geq\mu({\mathcal{C}})-\epsilon, (3.2)

but if the same argument simultaneously applied for both μ\mu and ν\nu for the OO satisfying both μ(E)>μ(O)ϵ\mu(E)>\mu(O)-\epsilon and ν(E)>ν(O)ϵ\nu(E)>\nu(O)-\epsilon, we may take the same 𝒞{\mathcal{C}} for the both and the same inequality (3.2) holds for ν\nu, and because consisting of a finite number of disjoint cosets, ν(𝒞)=μ(𝒞)\nu({\mathcal{C}})=\mu({\mathcal{C}}) holds, which implies μ(E)=ν(E)\mu(E)=\nu(E). This means that the restrictions of ν\nu and μ\mu to compact sets give the same regular content, which implies ν=μ\nu=\mu by [10, §54, Theorem B]. ∎

Proposition 3.16.

For a finite set XX, we define a (normalized) Hermitian inner product on C(X)C(X) by

(f,g):=1#XxXf(x)g(x)¯.(f,g):=\frac{1}{\#X}\sum_{x\in X}f(x)\overline{g(x)}.

For a profinite association scheme, it is defined on C(X)C({{{X}^{{\kern-1.04996pt}\wedge}}}) by

(f,g):=xXf(x)g(x)¯𝑑μ(x).(f,g):=\int_{x\in{{{X}^{{\kern-0.74997pt}\wedge}}}}f(x)\overline{g(x)}d\mu(x).

Then C(Xλ)C(X)C(X_{\lambda})\to C({{{X}^{{\kern-1.04996pt}\wedge}}}) preserves the inner products.

This is clear since XXλ{{{X}^{{\kern-1.04996pt}\wedge}}}\to X_{\lambda} preserves the measures.

Definition 3.17.

Let V,WV,W be {\mathbb{C}}-vector spaces with Hermitian inner products. For a linear morphism A:VWA:V\to W, its conjugate A:WVA^{*}:W\to V (which may not exists) is characterized by the property

(Af,g)W=(f,Ag)V(Af,g)_{W}=(f,A^{*}g)_{V}

for any fVf\in V, gWg\in W. It is unique if exists. An endomorphism A:WWA:W\to W is said to be Hermitian if A=AA=A^{*}, as usual.

Let WW be again a {\mathbb{C}}-vector space with Hermitian inner product, and VV is a subspace with restricted inner product. Let VV^{\perp} be the space of vectors which is orthogonal to any vector in VV. If VVV\oplus V^{\perp} is equal to WW, then VV is said to be an orthogonal component of WW.

Definition 3.18.

Let WW be a {\mathbb{C}}-linear space with Hermitian inner product. For a linear subspace VWV\subset W, let p:WWp:W\to W be a projection to VV, i.e., the image of pp is VV and p|V=idVp|_{V}={\operatorname{id}}_{V}. Then by linear algebra WW is decomposed as the direct sum, W=VKerpW=V\oplus\operatorname{Ker}p. We say that pp is an orthogonal projector to VV if VKerpV\oplus\operatorname{Ker}p is an orthogonal direct sum. An orthogonal projector to VV exists uniquely, if VV is an orthogonal component.

Proposition 3.19.

Let WW be a {\mathbb{C}}-linear space with Hermitian inner product, VWV\subset W be an orthogonal component with orthogonal projector pp, and ι:VW\iota:V\hookrightarrow W be the inclusion. We denote by q:WVq:W\to V the restriction of the codomain of pp. Then we have ι=q\iota^{*}=q and q=ιq^{*}=\iota. For A:VVA:V\to V, if AA^{*} exists, then

(ιAq)=ιAq:VV.(\iota\circ A\circ q)^{*}=\iota\circ A^{*}\circ q:V\to V. (3.3)

We sometimes denote ιAq\iota\circ A\circ q simply by AA. We make a remark when this abuse of notation is used.

Proof.

For wWw\in W and vVv\in V, write w=wV+wVw=w_{V}+w_{V}^{\prime} with wVVw_{V}\in V and wVVw_{V}^{\prime}\in V^{\perp}. Then

(qw,v)V=(wV,v)V=(ιwV,ιv)W=(wV+wV,ιv)W=(w,ιv)W,(qw,v)_{V}=(w_{V},v)_{V}=(\iota w_{V},\iota v)_{W}=(w_{V}+w_{V}^{\prime},\iota v)_{W}=(w,\iota v)_{W},

which shows that ι=q\iota^{*}=q and q=ιq^{*}=\iota. Thus (3.3) follows. ∎

Proposition 3.20.

Let WW be a {\mathbb{C}}-vector space with Hermitian inner product, and p:WWp:W\to W an idempotent, namely, p2=pp^{2}=p. By linear algebra, W=Im(p)Ker(p)W=\operatorname{Im}(p)\oplus\operatorname{Ker}(p) and pp is a projector to V:=Im(p)V:=\operatorname{Im}(p). Then, pp is an orthogonal projector if and only if p=pp=p^{*}. If AEnd(V)A\in{{\operatorname{End}}}(V) is an orthogonal projector to UVU\subset V, then ιAq:WW\iota Aq:W\rightarrow W is a projector to UWU\subset W.

Proof.

If p2=pp^{2}=p, then W=VKerpW=V\oplus\operatorname{Ker}p for V=Im(p)V=\operatorname{Im}(p). Thus any element w,wWw,w^{\prime}\in W is written as v+k,v+kv+k,v^{\prime}+k^{\prime}, respectively. If VKerpV\oplus\operatorname{Ker}p is orthogonal,

(p(v+k),v+k)=(pv,v+k)=(v,v)=(v+k,pv)=(v+k,p(v+k)),(p(v+k),v^{\prime}+k^{\prime})=(pv,v^{\prime}+k^{\prime})=(v,v^{\prime})=(v+k,pv^{\prime})=(v+k,p(v^{\prime}+k^{\prime})),

which shows p=pp=p^{*}. For the converse, assume p=pp=p^{*}. Then for vVv\in V, kKerpk\in\operatorname{Ker}p, (v,k)=(pv,k)=(v,pk)=(v,0)=0(v,k)=(pv,k)=(v,pk)=(v,0)=0, and hence WW and Kerp\operatorname{Ker}p are mutually orthogonal. For the last, (3.3) implies that ιAq\iota Aq is an Hermitian operator, and by qι=idVq\circ\iota={\operatorname{id}}_{V} is an idempotent, whose image is UU, hence is an orthogonal projector to UU. ∎

Definition 3.21.

Let (Xλ)(X_{\lambda}) be a profinite association scheme, fix λΛ\lambda\in\Lambda and xλXλx_{\lambda}\in X_{\lambda}. For the surjection pr:XXλ{\operatorname{pr}}:{{{X}^{{\kern-1.04996pt}\wedge}}}\to X_{\lambda}, the fiber pr1(xλ){\operatorname{pr}}^{-1}(x_{\lambda}) at xλx_{\lambda} is a clopen subset of X{{{X}^{{\kern-1.04996pt}\wedge}}}. Thus, we have the restriction of the measure μ\mu on X{{{X}^{{\kern-1.04996pt}\wedge}}} to pr1(xλ){\operatorname{pr}}^{-1}(x_{\lambda}), denoted by the same symbol (pr1(xλ),μ)({\operatorname{pr}}^{-1}(x_{\lambda}),\mu). Note that the volume of the fiber pr1(xλ){\operatorname{pr}}^{-1}(x_{\lambda}) is 1/#Xλ1/\#X_{\lambda}.

Since X=xλXλpr1(xλ){{{X}^{{\kern-1.04996pt}\wedge}}}=\coprod_{x_{\lambda}\in X_{\lambda}}{\operatorname{pr}}^{-1}(x_{\lambda}), the following proposition holds.

Proposition 3.22.

For any λ\lambda and fC(X)f\in C({{{X}^{{\kern-1.04996pt}\wedge}}}), we have

xXf(x)𝑑μ(x)=xλXλypr1(xλ)f(y)𝑑μ(y).\int_{x\in{{{X}^{{\kern-0.74997pt}\wedge}}}}f(x)d\mu(x)=\sum_{x_{\lambda}\in X_{\lambda}}\int_{y\in{\operatorname{pr}}^{-1}(x_{\lambda})}f(y)d\mu(y).
Proposition 3.23.

Let (Xλ)(X_{\lambda}) be a profinite association scheme. Then, by the surjection pr:XXλ{\operatorname{pr}}:{{{X}^{{\kern-1.04996pt}\wedge}}}\to X_{\lambda}, C(Xλ)C(X_{\lambda}) is identified with a subspace of C(X)C({{{X}^{{\kern-1.04996pt}\wedge}}}) with the inner product given by restriction. There is an orthogonal projection q:C(X)C(Xλ)q:C({{{X}^{{\kern-1.04996pt}\wedge}}})\to C(X_{\lambda}) given by averaging over each fiber: for fC(X)f\in C({{{X}^{{\kern-1.04996pt}\wedge}}}) define

q(f)(xλ)=1μ(pr1(xλ))xpr1(xλ)f(x)𝑑μ(x),q(f)(x_{\lambda})=\frac{1}{\mu({\operatorname{pr}}^{-1}(x_{\lambda}))}\int_{x\in{\operatorname{pr}}^{-1}(x_{\lambda})}f(x)d\mu(x),

where μ(pr1(xλ))=1#Xλ\mu({\operatorname{pr}}^{-1}(x_{\lambda}))=\frac{1}{\#X_{\lambda}}. Thus, C(Xλ)C(X_{\lambda}) is an orthogonal component of C(X)C({{{X}^{{\kern-1.04996pt}\wedge}}}).

Proof.

The preservation of the inner products is shown in Proposition 3.16. For any gC(Xλ)g\in C(X_{\lambda}) its image is gprC(X)g\circ{\operatorname{pr}}\in C({{{X}^{{\kern-1.04996pt}\wedge}}}) and we have

q(gpr)(xλ)\displaystyle q(g\circ{\operatorname{pr}})(x_{\lambda}) =\displaystyle= 1μ(pr1(xλ))ypr1(xλ)g(pr(y))𝑑μ(y)\displaystyle\frac{1}{\mu({\operatorname{pr}}^{-1}(x_{\lambda}))}\int_{y\in{\operatorname{pr}}^{-1}(x_{\lambda})}g({\operatorname{pr}}(y))d\mu(y)
=\displaystyle= g(xλ)1μ(pr1(xλ))ypr1(xλ)1𝑑μ(y)\displaystyle g(x_{\lambda})\frac{1}{\mu({\operatorname{pr}}^{-1}(x_{\lambda}))}\int_{y\in{\operatorname{pr}}^{-1}(x_{\lambda})}1\cdot d\mu(y)
=\displaystyle= g(xλ).\displaystyle g(x_{\lambda}).

This implies that qq is a projection to C(Xλ)C(X_{\lambda}).

To show that qq is an orthogonal projection, we take fKerqf\in\operatorname{Ker}q. This means that the average of ff on each fiber of pr{\operatorname{pr}} is zero. To show that qq is orthogonal, it suffices to show that (f,gpr)X=0(f,g\circ{\operatorname{pr}})_{{{X}^{{\kern-0.74997pt}\wedge}}}=0 for gC(Xλ)g\in C(X_{\lambda}) by Definition 3.18, but by applying Proposition 3.22, we have

(f,gpr)X\displaystyle(f,g\circ{\operatorname{pr}})_{{{X}^{{\kern-0.74997pt}\wedge}}} =\displaystyle= xXf(x)g(pr(x))¯𝑑μ(x)\displaystyle\int_{x\in{{{X}^{{\kern-0.74997pt}\wedge}}}}f(x)\overline{g({\operatorname{pr}}(x))}d\mu(x)
=\displaystyle= xXλxpr1(x)f(x)g(pr(x))¯𝑑μ(x)\displaystyle\sum_{x^{\prime}\in X_{\lambda}}\int_{x\in{\operatorname{pr}}^{-1}(x^{\prime})}f(x)\overline{g({\operatorname{pr}}(x))}d\mu(x)
=\displaystyle= xXλg(x)¯xpr1(x)f(x)𝑑μ(x)\displaystyle\sum_{x^{\prime}\in X_{\lambda}}\overline{g(x^{\prime})}\int_{x\in{\operatorname{pr}}^{-1}(x^{\prime})}f(x)d\mu(x)
=\displaystyle= xXλg(x)¯0=0.\displaystyle\sum_{x^{\prime}\in X_{\lambda}}\overline{g(x^{\prime})}\cdot 0=0.

Proposition 3.24.

  1. (1)

    The (convolution) algebra AXA_{{{X}^{{\kern-0.74997pt}\wedge}}} acts on C(X)C({{{X}^{{\kern-1.04996pt}\wedge}}}), which makes C(X)C({{{X}^{{\kern-1.04996pt}\wedge}}}) an AXA_{{{X}^{{\kern-0.74997pt}\wedge}}}-module. This is compatible with the action of AXλA_{X_{\lambda}} on C(Xλ)C(X_{\lambda}), in the sense of Corollary 2.8.

  2. (2)

    The convolution unit EXλAXλAXE_{X_{\lambda}}\in A_{X_{\lambda}}\subset A_{{{{X}^{{\kern-0.74997pt}\wedge}}}} acts as an orthogonal projector to C(Xλ)C(X_{\lambda}).

  3. (3)

    Let jJj\in{{{J}^{{\kern-1.04996pt}\wedge}}}. Take a λ\lambda where jj is isolated to jλj_{\lambda}, and consider EjλAXλE_{j_{\lambda}}\in A_{X_{\lambda}}. By the abuse of notation in Proposition 3.19, EjλE_{j_{\lambda}} extends to the orthogonal projector

    Ej:C(X)C(X)E_{j}:C({{{X}^{{\kern-1.04996pt}\wedge}}})\to C({{{X}^{{\kern-1.04996pt}\wedge}}})

    that projects to the domain of the composition

    C(Xλ)jλC(Xλ)C(X).C(X_{\lambda})_{j_{\lambda}}\to C(X_{\lambda})\to C({{{X}^{{\kern-1.04996pt}\wedge}}}).

    We define

    C(X)j:=C(Xλ)jλ.C({{{X}^{{\kern-1.04996pt}\wedge}}})_{j}:=C(X_{\lambda})_{j_{\lambda}}.

    This is independent of the choice of the λ\lambda.

  4. (4)

    For jjj\neq j^{\prime}, EjEj=0E_{j}\bullet E_{j^{\prime}}=0.

  5. (5)

    We have an orthogonal decomposition

    C(Xλ)=jJC(X)jC(X_{\lambda})=\bigoplus_{j\in J^{\prime}}C({{{X}^{{\kern-1.04996pt}\wedge}}})_{j}

    for the domain JJ^{\prime} of the partial map from J{{{J}^{{\kern-1.04996pt}\wedge}}} to JλJ_{\lambda}. Note that JJ^{\prime} is finite.

Proof.

The action is given by: AAXλA\in A_{X_{\lambda}} acts on fC(X)f\in C({{{X}^{{\kern-1.04996pt}\wedge}}}) by

(Af)(x):=yXA(prλ(x),prλ(y))f(y)𝑑μ(y).(A\bullet f)(x):=\int_{y\in{{{X}^{{\kern-0.74997pt}\wedge}}}}A({\operatorname{pr}}_{\lambda}(x),{\operatorname{pr}}_{\lambda}(y))f(y)d\mu(y).

By Proposition 3.22, this is the same with

yXλA(prλ(x),y)(yprλ1(y)f(y)𝑑μ(y)).\sum_{y^{\prime}\in X_{\lambda}}A({\operatorname{pr}}_{\lambda}(x),y^{\prime})(\int_{y\in{\operatorname{pr}}_{\lambda}^{-1}(y^{\prime})}f(y)d\mu(y)). (3.4)

Comparing with qq in Proposition 3.23, we have

(AXf)(x)\displaystyle(A\bullet_{{{{X}^{{\kern-0.74997pt}\wedge}}}}f)(x) =\displaystyle= yXλA(prλ(x),y)(yprλ1(y)f(y)𝑑μ(y))\displaystyle\sum_{y^{\prime}\in X_{\lambda}}A({\operatorname{pr}}_{\lambda}(x),y^{\prime})(\int_{y\in{\operatorname{pr}}_{\lambda}^{-1}(y^{\prime})}f(y)d\mu(y))
=\displaystyle= yXλA(prλ(x),y)1#Xλq(f)(y)\displaystyle\sum_{y^{\prime}\in X_{\lambda}}A({\operatorname{pr}}_{\lambda}(x),y^{\prime})\frac{1}{\#X_{\lambda}}q(f)(y^{\prime})
=\displaystyle= (AXλq(f))(prλ(x)),\displaystyle(A\bullet_{X_{\lambda}}q(f))({\operatorname{pr}}_{\lambda}(x)),

which shows that

A=ιAq,A=\iota Aq, (3.5)

where ι:C(Xλ)C(X)\iota:C(X_{\lambda})\to C({{{X}^{{\kern-1.04996pt}\wedge}}}) is the injection prλ{\operatorname{pr}}_{\lambda}^{\dagger} and AA in the right hand side is an endomorphism of C(Xλ)C(X_{\lambda}). This and qι=idC(Xλ)q\circ\iota={\operatorname{id}}_{C(X_{\lambda})} show that C(X)C({{{X}^{{\kern-1.04996pt}\wedge}}}) is an AXA_{{{X}^{{\kern-0.74997pt}\wedge}}}-module. The well-definedness (independence of the choice of λ\lambda) follows by taking a sufficiently large μ\mu.

By (3.5) and EXλ=idC(Xλ)E_{X_{\lambda}}={\operatorname{id}}_{C(X_{\lambda})} on C(Xλ)C(X_{\lambda}) show that

EXλ=ιq,E_{X_{\lambda}}=\iota\circ q,

which is the orthogonal projection to C(Xλ)C(X_{\lambda}).

For jJj\in J, take a λ\lambda where jj is isolated to jλJλj_{\lambda}\in J_{\lambda}. By (3.5) we have

Ejλ=ιEjλq,E_{j_{\lambda}}=\iota E_{j_{\lambda}}q,

where EjλE_{j_{\lambda}} in the right hand side is an endomorphism of C(Xλ)C(X_{\lambda}). Hence to prove that EjλE_{j_{\lambda}} is an orthogonal projector C(X)C(Xλ)jλC({{{X}^{{\kern-1.04996pt}\wedge}}})\to C(X_{\lambda})_{j_{\lambda}}, it suffices to show that EjλAXλE_{j_{\lambda}}\in A_{X_{\lambda}} is an orthogonal projector C(Xλ)C(Xλ)jλC(X_{\lambda})\to C(X_{\lambda})_{j_{\lambda}} by Proposition 3.20. The orthogonality of EjλAXλE_{j_{\lambda}}\in A_{X_{\lambda}} is well known [1]. (Since AXλA_{X_{\lambda}} is closed under the anti-isomorphism * of ring, EjλE_{j_{\lambda}}^{*} must be a primitive idempotent, and since EjλEjλ0E_{j_{\lambda}}E_{j_{\lambda}}^{*}\neq 0, they must coincide, hence EjλE_{j_{\lambda}} is Hermitian and an orthogonal projection by Proposition 3.20.) We define Ej:=EjλE_{j}:=E_{j_{\lambda}}, where

Ej:C(X)C(X)j:=C(Xλ)jλ.E_{j}:C({{{X}^{{\kern-1.04996pt}\wedge}}})\to C({{{X}^{{\kern-1.04996pt}\wedge}}})_{j}:=C(X_{\lambda})_{j_{\lambda}}.

For the well-definedness, we take λ\lambda, λ\lambda^{\prime} where jj is isolated. There exists a μ\mu with μλ\mu\geq\lambda, μλ\mu\geq\lambda^{\prime}. Then jj is isolated to jμj_{\mu} at μ\mu, and by (3.5) it is not difficult to see that the EjE_{j} defined using μ\mu is the same as those defined using λ,λ\lambda,\lambda^{\prime}. For EjEj=0E_{j}\bullet E_{j^{\prime}}=0, take a λ\lambda where both jj, jj^{\prime} are isolated to jλj_{\lambda}, jλj^{\prime}_{\lambda}, respectively. Then, jλjλj_{\lambda}\neq j_{\lambda}^{\prime}, Ejλ,EjλAXλE_{j_{\lambda}},E_{j^{\prime}_{\lambda}}\in A_{X_{\lambda}}, and EjλEjλ=0E_{j_{\lambda}}\bullet E_{j^{\prime}_{\lambda}}=0. Thus, the relation holds in the inductive limit AXA_{{{{X}^{{\kern-0.74997pt}\wedge}}}}.

For fixed λ\lambda, it is well-known [1] that we have an orthogonal decomposition

C(Xλ)=jλJλC(Xλ)jλ.C(X_{\lambda})=\bigoplus_{j_{\lambda}\in J_{\lambda}}C(X_{\lambda})_{j_{\lambda}}.

(This follows from the orthogonality of EjλE_{j_{\lambda}}.) Since each C(Xλ)jλC(X_{\lambda})_{j_{\lambda}} is an orthogonal direct sum of C(X)jC({{{X}^{{\kern-1.04996pt}\wedge}}})_{j} over the finite number of jJj\in{{{J}^{{\kern-1.04996pt}\wedge}}} whose image is jλj_{\lambda}, the claim follows. ∎

The following is an analogue to Peter-Weyl Theorem.

Theorem 3.25.

We have an orthogonal direct decomposition

limC(Xλ)=jJC(X)j.\varinjlim C(X_{\lambda})=\bigoplus_{j\in J}C({{{X}^{{\kern-1.04996pt}\wedge}}})_{j}.

Hence the right hand side is dense in C(X)C({{{X}^{{\kern-1.04996pt}\wedge}}}) with respect to the supremum norm.

Proof.

Proposition 3.24 implies that C(X)jC({{{X}^{{\kern-1.04996pt}\wedge}}})_{j} are mutually orthogonal, and their orthogonal sum is a subspace of C(X)C({{{X}^{{\kern-1.04996pt}\wedge}}}). For any λ\lambda, C(Xλ)C(X_{\lambda}) is an orthogonal sum of a finite number of C(X)jC({{{X}^{{\kern-1.04996pt}\wedge}}})_{j}. Thus limC(Xλ)\varinjlim C(X_{\lambda}) is the direct sum of C(X)jC({{{X}^{{\kern-1.04996pt}\wedge}}})_{j} for jJj\in{{{J}^{{\kern-1.04996pt}\wedge}}}. The density follows from Lemma 3.14. ∎

3.3. Measures and Fourier analysis

Since X{{{X}^{{\kern-1.04996pt}\wedge}}} has a natural probability measure, I{{{I}^{{\kern-1.04996pt}\wedge}}} has a pushforward measure via X×XI{{{X}^{{\kern-1.04996pt}\wedge}}}\times{{{X}^{{\kern-1.04996pt}\wedge}}}\to{{{I}^{{\kern-1.04996pt}\wedge}}}. It will be shown that J{{{J}^{{\kern-1.04996pt}\wedge}}} has a natural measure, and these measures are closely related to the classical notions of the multiplicity mjλm_{j_{\lambda}} for JλJ_{\lambda}, and the valency kiλk_{i_{\lambda}} for IλI_{\lambda} (see [1][6]).

Definition 3.26.

Let μI\mu_{{{I}^{{\kern-0.74997pt}\wedge}}} be the pushforward measure on I{{{I}^{{\kern-1.04996pt}\wedge}}} from X×X{{{X}^{{\kern-1.04996pt}\wedge}}}\times{{{X}^{{\kern-1.04996pt}\wedge}}}. This is characterized as follows. For iλIλi_{\lambda}\in I_{\lambda}, let prIλ1(i)I{\operatorname{pr}}_{I_{\lambda}}^{-1}(i)\subset{{{I}^{{\kern-1.04996pt}\wedge}}} be the open set. Then

μI(prIλ1(iλ))=μX×X(RλprXλ×Xλ)1(iλ).\mu_{{{I}^{{\kern-0.74997pt}\wedge}}}({\operatorname{pr}}_{I_{\lambda}}^{-1}(i_{\lambda}))=\mu_{{{{X}^{{\kern-0.74997pt}\wedge}}}\times{{{X}^{{\kern-0.74997pt}\wedge}}}}(R_{\lambda}\circ{\operatorname{pr}}_{X_{\lambda}\times X_{\lambda}})^{-1}(i_{\lambda}).
Proposition 3.27.

In the above definition,

μI(prIλ1(iλ))=kiλ#Xλ\mu_{{{I}^{{\kern-0.74997pt}\wedge}}}({\operatorname{pr}}_{I_{\lambda}}^{-1}(i_{\lambda}))=\frac{k_{i_{\lambda}}}{\#X_{\lambda}}

holds, where kiλk_{i_{\lambda}} is the valency of the adjacency matrix AiλAXλA_{i_{\lambda}}\in A_{X_{\lambda}}.

Proof.

Recall that AiλA_{i_{\lambda}} is the indicator function of Rλ1(iλ)R_{\lambda}^{-1}(i_{\lambda}), and the sum of the rows is an integer kiλk_{i_{\lambda}} called the valency [1]. Thus, the volume μI(prIλ1(iλ))\mu_{{{{I}^{{\kern-0.74997pt}\wedge}}}}({\operatorname{pr}}_{I_{\lambda}}^{-1}(i_{\lambda})) is the volume of μX(prXλ×Xλ)1Rλ1(iλ)\mu_{{{{X}^{{\kern-0.74997pt}\wedge}}}}({\operatorname{pr}}_{X_{\lambda}\times X_{\lambda}})^{-1}R_{\lambda}^{-1}(i_{\lambda}), which is

x,yX×XAiλ(prXλ(x),prXλ(y))𝑑μ(x)𝑑μ(y)\displaystyle\int_{x,y\in{{{X}^{{\kern-0.74997pt}\wedge}}}\times{{{X}^{{\kern-0.74997pt}\wedge}}}}A_{i_{\lambda}}({\operatorname{pr}}_{X_{\lambda}}(x),{\operatorname{pr}}_{X_{\lambda}}(y))d\mu(x)d\mu(y) =\displaystyle= 1#Xλ2x,yXλ×XλAiλ(x,y)\displaystyle\frac{1}{\#X_{\lambda}^{2}}\sum_{x^{\prime},y^{\prime}\in X_{\lambda}\times X_{\lambda}}A_{i_{\lambda}}(x^{\prime},y^{\prime})
=\displaystyle= kiλ#Xλ.\displaystyle\frac{k_{i_{\lambda}}}{\#X_{\lambda}}.

Definition 3.28.

The algebra AXλA_{X_{\lambda}} is closed under the adjoint . We define the Hilbert-Schmidt inner product, which is Hermitian, as follows. For A,BAXλA,B\in A_{X_{\lambda}},

(A,B)HS:=xXλ(AB)(x,x).(A,B)_{{\operatorname{HS}}}:=\sum_{x\in X_{\lambda}}(A\bullet B^{*})(x,x).

For A,BAXA,B\in A_{{{X}^{{\kern-0.74997pt}\wedge}}}, we define its Hilbert-Schmidt inner product by

(A,B)HS:=xX(AB)(x,x)𝑑μ(x).(A,B)_{{\operatorname{HS}}}:=\int_{x\in{{{X}^{{\kern-0.74997pt}\wedge}}}}(A\bullet B^{*})(x,x)d\mu(x).

Then inclusion AXλAXA_{X_{\lambda}}\to A_{{{{X}^{{\kern-0.74997pt}\wedge}}}} preserves the inner products.

Proof.

For BAXλB\in A_{X_{\lambda}}, BAXλB^{*}\in A_{X_{\lambda}} follows because of the definition of association schemes, and the Hilbert-Schmidt inner product is defined. For BAXB\in A_{{{{X}^{{\kern-0.74997pt}\wedge}}}}, we must show that BB^{*} exists in AXA_{{{{X}^{{\kern-0.74997pt}\wedge}}}} and the integration converges. However, BAXλB\in A_{X_{\lambda}} for some λ\lambda, and BAXλB^{*}\in A_{X_{\lambda}}. Since C(Xλ)C(X_{\lambda}) is an orthogonal component of C(X)C({{{X}^{{\kern-1.04996pt}\wedge}}}), B=ιBqB^{*}=\iota B^{*}q is the adjoint as an operator on C(X)C({{{X}^{{\kern-1.04996pt}\wedge}}}), by the abuse of notation in Proposition 3.19. The notation (AB)(x,x)(A\bullet B^{*})(x,x) comes from the middle vertical injection in (3.1). Since A,BAXλC(Xλ×Xλ)A,B^{*}\in A_{X_{\lambda}}\subset C(X_{\lambda}\times X_{\lambda}),

xX(AB)(x,x)𝑑μ(x)=xXλ(AB)(x,x),\int_{x\in{{{X}^{{\kern-0.74997pt}\wedge}}}}(A\bullet B^{*})(x,x)d\mu(x)=\sum_{x^{\prime}\in X_{\lambda}}(A\bullet B^{*})(x^{\prime},x^{\prime}),

which is finite, satisfying the axioms of Hermitian inner product, and compatible with AXλAXA_{X_{\lambda}}\to A_{{{{X}^{{\kern-0.74997pt}\wedge}}}}. ∎

Proposition 3.29.

Through the natural isomorphism

limC(Iλ)AX,\varinjlim C(I_{\lambda})\to A_{{{{X}^{{\kern-0.74997pt}\wedge}}}},

we obtain Hermitian inner product on limC(Iλ)\varinjlim C(I_{\lambda}), which is denoted by (,)HS(-,-)_{{\operatorname{HS}}}. Then for f,glimC(Iλ)f,g\in\varinjlim C(I_{\lambda}), we have

(f,g)HS=iIf(i)g(i)¯𝑑μI(i).(f,g)_{{\operatorname{HS}}}=\int_{i\in{{{I}^{{\kern-0.74997pt}\wedge}}}}f(i)\overline{g(i)}d\mu_{{{I}^{{\kern-0.74997pt}\wedge}}}(i).
Proof.

We may assume f,gC(Iλ)f,g\in C(I_{\lambda}) for some λ\lambda. Then by bilinearity it suffices to check the equality for f=δiλf=\delta_{i_{\lambda}}, g=δiλg=\delta_{i_{\lambda}^{\prime}}. Then, the right hand side is δiλ,iλkiλ/#Xλ\delta_{i_{\lambda},i_{\lambda}^{\prime}}k_{i_{\lambda}}/\#X_{\lambda} by Proposition 3.27. In AXA_{{{X}^{{\kern-0.74997pt}\wedge}}}, these functions correspond to fAiλ,gAiλf\mapsto A_{i_{\lambda}},g\mapsto A_{i_{\lambda}^{\prime}}, and their inner product is

x,yX×XAiλ(prXλ(x),prXλ(y))Aiλ(prXλ(y),prXλ(x))𝑑μ(x)𝑑μ(y)\displaystyle\int_{x,y\in{{{X}^{{\kern-0.74997pt}\wedge}}}\times{{{X}^{{\kern-0.74997pt}\wedge}}}}A_{i_{\lambda}}({\operatorname{pr}}_{X_{\lambda}}(x),{\operatorname{pr}}_{X_{\lambda}}(y))A_{i_{\lambda}^{\prime}}({\operatorname{pr}}_{X_{\lambda}}(y),{\operatorname{pr}}_{X_{\lambda}}(x))^{*}d\mu(x)d\mu(y)
=\displaystyle= 1#Xλ2x,yXλ×XλAiλ(x,y)Aiλ(x,y)\displaystyle\frac{1}{\#X_{\lambda}^{2}}\sum_{x^{\prime},y^{\prime}\in X_{\lambda}\times X_{\lambda}}A_{i_{\lambda}}(x^{\prime},y^{\prime})A_{i_{\lambda}^{\prime}}(x^{\prime},y^{\prime})
=\displaystyle= δiλ,iλkiλ#Xλ.\displaystyle\delta_{i_{\lambda},i_{\lambda}^{\prime}}\frac{k_{i_{\lambda}}}{\#X_{\lambda}}.

This uses a well-known orthogonality of AiλA_{i_{\lambda}}: for any xXλx\in X_{\lambda} and i1,i2Iλi_{1},i_{2}\in I_{\lambda}, the (x,x)(x,x)-component of the matrix product Ai1tAi2A_{i_{1}}\cdot^{t}A_{i_{2}} is the number of yXλy\in X_{\lambda} with R(x,y)=i1R(x,y)=i_{1} and R(x,y)=i2R(x,y)=i_{2}, which is zero unless i1=i2i_{1}=i_{2}, and in the equal case ki1k_{i_{1}}. ∎

Lemma 3.30.

For a commutative association scheme (X,R,I)(X,R,I), we defined JJ as the set of primitive idempotents EjE_{j} with respect to the product \bullet in Theorem 2.5. We put a positive measure mjm_{j} for jJj\in J, so that the inner product with respect to mjm_{j} in C(J)C(J) is isometric to that of AXA_{X} with respect to HS{{\operatorname{HS}}} inner product. Then, mjm_{j} is the multiplicity of jj, defined as dim(C(X)j)\dim(C(X)_{j}) [1].

Proof.

Let Ej,EjAXE_{j},E_{j^{\prime}}\in A_{X} be two primitive idempotents. The computation

(Ej,Ej)HS=xXEjEj(x,x)=δjjtr(Ej)=δjjmj\displaystyle(E_{j},E_{j^{\prime}})_{{{\operatorname{HS}}}}=\sum_{x\in X}E_{j}\bullet E_{j^{\prime}}^{*}(x,x)=\delta_{jj^{\prime}}{{\operatorname{tr}}}(E_{j})=\delta_{jj^{\prime}}m_{j}

implies that if we define μJ(j):=mj\mu_{J}(j):=m_{j}, then the corresponding inner product is compatible with C(J)AXC(J)\to A_{X} with respect to the HS{{\operatorname{HS}}}-inner product on AA. ∎

Proposition 3.31.

For a projective association scheme (Xλ,Rλ,Iλ)(X_{\lambda},R_{\lambda},I_{\lambda}), we define a measure μJ\mu_{{{J}^{{\kern-0.74997pt}\wedge}}} on the discrete set J{{{J}^{{\kern-1.04996pt}\wedge}}} by μJ(j)=dimC(X)j\mu_{{{J}^{{\kern-0.74997pt}\wedge}}}(j)=\dim C({{{X}^{{\kern-1.04996pt}\wedge}}})_{j}. Then, the inner product on limC(Jλ)\varinjlim C(J_{\lambda}) defined by

(f,g)HS=jJf(j)g(j)¯μJ(j)(f,g)_{{\operatorname{HS}}}=\sum_{j\in{{{J}^{{\kern-0.74997pt}\wedge}}}}f(j)\overline{g(j)}\mu_{{{J}^{{\kern-0.74997pt}\wedge}}}(j)

makes limC(Jλ)AX\varinjlim C(J_{\lambda})\cong A_{{{X}^{{\kern-0.74997pt}\wedge}}} isometric with respect to HS{{\operatorname{HS}}} inner product.

Proof.

Any f,glimC(Jλ)f,g\in\varinjlim C(J_{\lambda}) are contained in some C(Jλ)C(J_{\lambda}), where any jj in the support of ff or gg is isolated at λ\lambda. The functions f,gf,g are linear combinations of orthogonal basis EjE_{j} of C(Jλ)C(J_{\lambda}), namely, f=jf(j)Ejf=\sum_{j}f(j)E_{j} and g=jg(j)Ejg=\sum_{j}g(j)E_{j} in AXλA_{X_{\lambda}}. The previous lemma says that (Ej,Ej)HS=δjjdimC(X)j(E_{j},E_{j^{\prime}})_{{\operatorname{HS}}}=\delta_{jj^{\prime}}\dim C({{{X}^{{\kern-1.04996pt}\wedge}}})_{j}, which proves the proposition. ∎

In sum, we have the following form of Fourier transform.

Theorem 3.32.

The isomorphism of vector spaces limC(Jλ)AXlimC(Iλ)\varinjlim C(J_{\lambda})\cong A_{{{{X}^{{\kern-0.74997pt}\wedge}}}}\cong\varinjlim C(I_{\lambda}) is an isometry between the space of compact support functions Cc(J)C_{c}({{{J}^{{\kern-1.04996pt}\wedge}}}) with inner product defined in Proposition 3.31 and the space of locally constant functions Clc(I)C_{lc}({{{I}^{{\kern-1.04996pt}\wedge}}}) with inner product defined in Proposition 3.29, which maps the componentwise product in Cc(J)C_{c}({{{J}^{{\kern-1.04996pt}\wedge}}}) to the convolution product in Clc(I)C_{lc}({{{I}^{{\kern-1.04996pt}\wedge}}}).

Remark 3.33.

For a finite association scheme (X,R,I)(X,R,I), both EjE_{j} (jJ)(j\in J) and AiA_{i} (iI)(i\in I) are orthogonal bases of AXA_{X} with respect to HS{{\operatorname{HS}}}-inner product. The isometry

C(I)C(J)C(I)\to C(J)

is given by

Ai\displaystyle A_{i} \displaystyle\mapsto jJpi(j)Ej.\displaystyle\sum_{j\in J}p_{i}(j)E_{j}.

The isometry shows that

(Ai,Ai)HS=δiiμI(i)=δiiki#X(A_{i},A_{i^{\prime}})_{{\operatorname{HS}}}=\delta_{ii^{\prime}}\mu_{I}(i)=\delta_{ii^{\prime}}\frac{k_{i}}{\#X}

is equal to

jJpi(j)pi(j)¯mj,\sum_{j\in J}p_{i}(j)\overline{p_{i^{\prime}}(j)}m_{j},

which is one of the two orthogonalities of eigenmatrices as classically known [1]. (Here the factor #X\#X appears in a different manner because of the normalization in Theorem 2.5. Our EjE_{j} is #X\#X times the classical primitive idempotent, which makes pi(j)=1#XPi(j)p_{i}(j)=\frac{1}{\#X}P_{i}(j) where Pi(j)P_{i}(j) are the components of the standard eigen matrices.)

The next is an analogue of the Fourier transform for L2L^{2} spaces.

Proposition 3.34.

The isometry limC(Jλ)limC(Iλ)\varinjlim C(J_{\lambda})\to\varinjlim C(I_{\lambda}) extends to a unique isometry

L2(J,μJ)L2(I,μI).L^{2}({{{J}^{{\kern-1.04996pt}\wedge}}},\mu_{{{J}^{{\kern-0.74997pt}\wedge}}})\to L^{2}({{{I}^{{\kern-1.04996pt}\wedge}}},\mu_{{{I}^{{\kern-0.74997pt}\wedge}}}).
Proof.

We consider L2L^{2}-norm for all the spaces. The map limC(Jλ)L2(J)\varinjlim C(J_{\lambda})\to L^{2}({{{J}^{{\kern-1.04996pt}\wedge}}}) is isometric, injective, and has a dense image, by considering the orthogonal basis EjE_{j}. Since L2(J)L^{2}({{{J}^{{\kern-1.04996pt}\wedge}}}) is complete, this is a completion of limC(Jλ)\varinjlim C(J_{\lambda}). The map limC(Iλ)L2(I)\varinjlim C(I_{\lambda})\to L^{2}({{{I}^{{\kern-1.04996pt}\wedge}}}) is also isometric, hence injective, and has a dense image (by Lemma 3.14 it is dense with respect to supremum norm in C(I)C({{{I}^{{\kern-1.04996pt}\wedge}}}), and since I{{{I}^{{\kern-1.04996pt}\wedge}}} is compact, dense with respect to the L2L^{2}-norm in C(I)C({{{I}^{{\kern-1.04996pt}\wedge}}}), and thus dense in L2(I)L^{2}({{{I}^{{\kern-1.04996pt}\wedge}}})), so L2(I)L^{2}({{{I}^{{\kern-1.04996pt}\wedge}}}) is a completion of limC(Iλ)\varinjlim C(I_{\lambda}). By the universality of the completion, they are isometric. ∎

4. Some examples and relation with Barg-Skriganov theory

The results of this paper are closely related to the study by Barg and Skriganov [2]. They defined the notion of association schemes on a set XX with a σ\sigma-additive measure. See [2, Definition 1], where II is denoted by Υ\Upsilon. Association scheme structure is given by a surjection

R:X×XI,R:X\times X\to I,

with axioms similar to the finite case. The inverse image R1(i)X×XR^{-1}(i)\subset X\times X of iIi\in I is required to be measurable in X×XX\times X, and intersection numbers pijkp_{ij}^{k} are defined as the measure of the set

{zXR(x,z)=i,R(z,y)=j}\{z\in X\mid R(x,z)=i,R(z,y)=j\}

where R(x,y)=kR(x,y)=k, which is required to be independent of the choice of x,yx,y. The set II is required to be at most countably infinite. This countability assumption seems to be necessary, for example to avoid the case where pijkp_{ij}^{k} is all zero for any i,j,ki,j,k.

Let χi(x,y)\chi_{i}(x,y) be the indicator function of R1(i)R^{-1}(i). They define adjacency algebras (Bose-Mesner algebras in our terminology) in [2, Section 3] as a set of finite linear combination of χi(x,y)\chi_{i}(x,y) with complex coefficient, with convolution product given by the integration

(ab)(x,y):=Xa(x,z)b(z,y)𝑑μ(z).(a*b)(x,y):=\int_{X}a(x,z)b(z,y)d\mu(z).

However, it is not clear whether the adjacency algebra is closed under the convolution product. In fact, the authors remarked in the last of [2, Section 3.1] “our arguments in this part are of somewhat heuristic nature. We make them fully rigorous for the case of association schemes on zero-dimensional groups; see Section 8.” Later in Section 8, they define the adjacency algebra with an aid of locally compact zero-dimensional abelian groups and their duality. Some of the treated objects (compact cases) are examples of profinite association schemes in this paper.

4.1. Profinite groups

Let GG be a profinite group, that is, a projective limit of finite groups GλG_{\lambda} with projective limit topology. Absolute Galois groups, arithmetic fundamental groups and pp-adic integers are some examples appear in arithmetic geometry [22]. There are significant amounts of studies on analysis on abelian profinite groups, see for example [14].

4.1.1. Schurian profinite association schemes

We describe Schurian association schemes, and then profinite analogues. When we say an action of a group GG, it means the left action, unless otherwise stated.

Lemma 4.1.

Let XX be a topological space and GG a group acting on XX such that every element of GG induces a homeomorphism on XX. Then the space of continuous functions C(X)C(X) is a left GG-module by f(x)f(g1x)f(x)\mapsto f(g^{-1}x). Let G\XG\backslash X denote the quotient space with quotient topology. Let

C(X)GC(X)C(X)^{G}\subset C(X)

denote the subspace of functions invariant by the action of GG. There is a canonical injection

C(G\X)C(X)C(G\backslash X)\hookrightarrow C(X)

associated with the continuous surjection XG\XX\to G\backslash X. Then, there is a canonical isomorphism of {\mathbb{C}}-algebras

C(X)GC(G\X),C(X)^{G}\to C(G\backslash X),

and hence we identify them.

Proof.

An element ff of C(X)GC(X)^{G} is a continuous function which takes one same value for each GG-orbit of XX. Thus, it gives a mapping G\XG\backslash X\to{\mathbb{C}}. By the universality of the quotient topology, this is continuous, hence lies in C(G\X)C(G\backslash X). Conversely, a function in C(G\X)C(G\backslash X) gives a continuous function in C(X)C(X), which is invariant by GG. ∎

Proposition 4.2.

(Schurian association scheme)

Let GG be a finite group, HH a subgroup, and let X:=G/HX:=G/H be the quotient. Then, we have a natural bijection

G\(X×X)H\G/H,G\backslash(X\times X)\to H\backslash G/H, (4.1)

where GG acts diagonally. We define

I:=G\(X×X).I:=G\backslash(X\times X).

The quotient mapping

R:X×XIR:X\times X\to I

is an association scheme, which is called a Schurian scheme.

We give a proof, since descriptions used in the proof are necessary for the profinite case. The equality (4.1) is given by a surjection

(g1H,g2H)Hg11g2H,(g_{1}H,g_{2}H)\mapsto Hg_{1}^{-1}g_{2}H,

which certainly factors through

G\(g1H,g2H)Hg11g2H,G\backslash(g_{1}H,g_{2}H)\mapsto Hg_{1}^{-1}g_{2}H,

and the converse map is given by

HgH(H,gH),HgH\mapsto(H,gH),

where the check of the well-definedness is left to the reader.

Lemma 4.3.

Let C(X):=Hom(C(X),)C(X)^{\vee}:=\operatorname{Hom}_{\mathbb{C}}(C(X),{\mathbb{C}}) be the dual space, which is regarded as a left GG-module, by letting gGg\in G act on ξC(X)\xi\in C(X)^{\vee} by g(ξ)f=ξ(g1(f))g(\xi)f=\xi(g^{-1}(f)). Then we have an isomorphism of GG-modules

C(X)C(X),δxevxC(X)\stackrel{{\scriptstyle\vee}}{{\to}}C(X)^{\vee},\quad\delta_{x}\mapsto{\operatorname{ev}}_{x}

for every xXx\in X, where δx\delta_{x} is the indicator function of xx and evx{\operatorname{ev}}_{x} is the evaluation at xx.

Proof.

It is well-known that {δxxX}\{\delta_{x}\mid x\in X\} is a linear basis of C(X)C(X), and {evxxX}\{{\operatorname{ev}}_{x}\mid x\in X\} is a linear basis of C(X)C(X)^{\vee}. We compute

g(δx)(y)=δx(g1y)=δgx(y).g(\delta_{x})(y)=\delta_{x}(g^{-1}y)=\delta_{gx}(y).

On the other hand,

g(evx)(f)=evx(g1(f))=(g1(f))(x)=f(gx)=evgx(f).g({\operatorname{ev}}_{x})(f)={\operatorname{ev}}_{x}(g^{-1}(f))=(g^{-1}(f))(x)=f(gx)={\operatorname{ev}}_{gx}(f).

Proposition 4.4.

The identification of a matrix and a linear map is obtained by an isomorphism of GG-modules

C(X×X)=C(X)C(X)idC(X)C(X)Hom(C(X),C(X)).C(X\times X)=C(X)\otimes C(X)\stackrel{{\scriptstyle{\operatorname{id}}\otimes\vee}}{{\to}}C(X)\otimes C(X)^{\vee}\cong\operatorname{Hom}_{\mathbb{C}}(C(X),C(X)). (4.2)
Proof.

An element AC(X×X)A\in C(X\times X) is mapped as

A\displaystyle A \displaystyle\mapsto x,yX2A(x,y)δxδy\displaystyle\sum_{x,y\in X^{2}}A(x,y)\delta_{x}\otimes\delta_{y}
\displaystyle\mapsto x,yX2A(x,y)δxevy.\displaystyle\sum_{x,y\in X^{2}}A(x,y)\delta_{x}\otimes{\operatorname{ev}}_{y}.

The last element is regarded as an element of Hom(C(X),C(X))\operatorname{Hom}_{\mathbb{C}}(C(X),C(X)) by mapping f=yXf(y)δyC(X)f=\sum_{y\in X}f(y)\delta_{y}\in C(X) to

(x,yX2A(x,y)δxevy)(f)=xXyYA(x,y)f(y)δx,(\sum_{x,y\in X^{2}}A(x,y)\delta_{x}\otimes{\operatorname{ev}}_{y})(f)=\sum_{x\in X}\sum_{y\in Y}A(x,y)f(y)\delta_{x},

which is the multiplication of a matrix to a vector. ∎

Proposition 4.5.

By taking the GG-invariant part of (4.2), we have

HomG(C(X),C(X))C(X×X)G=C(G\(X×X))=C(H\G/H),\operatorname{Hom}_{G}(C(X),C(X))\cong C(X\times X)^{G}=C(G\backslash(X\times X))=C(H\backslash G/H), (4.3)

where the left HomG\operatorname{Hom}_{G} denotes the endomorphism as GG-modules.

Proof.

This follows from (4.1), Lemma 4.1 and the definition of GG-homomorphisms. ∎

Corollary 4.6.

The construction in Proposition 4.2 gives an association scheme. The endomorphism ring HomG(C(X),C(X))\operatorname{Hom}_{G}(C(X),C(X)) is isomorphic to the Bose-Mesner algebra AXA_{X}.

Proof.

By the construction of I=G\(X×X)I=G\backslash(X\times X), the set of the adjacency matrices is closed under the transpose and contains the identity matrix. Their linear span is closed by the matrix product, since the adjacency matrices form a linear basis of (4.3) and the matrix product corresponds to the composition of the endomorphism ring HomG(C(X),C(X))\operatorname{Hom}_{G}(C(X),C(X)). This gives an association scheme with Bose-Mesner algebra HomG(C(X),C(X))\operatorname{Hom}_{G}(C(X),C(X)). ∎

Definition 4.7.

Let GG be a finite group acting on a {\mathbb{C}}-linear space VV. By the representation of finite groups, VV is decomposed into irreducible representations. For a fixed irreducible representation WW, the multiplicity of WW in VV is the dimension of the vector space

HomG(W,V).\operatorname{Hom}_{G}(W,V).

We say VV is multiplicity free, if the multiplicity is at most one for any irreducible representation WW.

Corollary 4.8.

The Schurian association scheme is commutative, if and only if C(X)C(X) is multiplicity free. In this case, the set of the primitive idempotents is canonically identified with the set of the irreducible representations of GG in C(X)C(X).

Proof.

If the multiplicity is one for each irreducible representation, Schur’s lemma implies that HomG(C(X),C(X))\operatorname{Hom}_{G}(C(X),C(X)) is a direct sum of copies of {\mathbb{C}} (each copy corresponding to each irreducible representation in C(X)C(X)) as a ring, and hence is commutative. If there is an irreducible component with multiplicity m2m\geq 2, HomG(C(X),C(X))\operatorname{Hom}_{G}(C(X),C(X)) contains a matrix algebra Mm()M_{m}({\mathbb{C}}), and thus is not commutative. ∎

See [22] for the following definition and properties (c.f. Section 3.1 above).

Definition 4.9.

Let Gλ(λΛ)G_{\lambda}(\lambda\in\Lambda) be a projective system of surjections of finite groups. A profinite group is a topological group which is isomorphic to

G:=limGλ,{{{G}^{{\kern-1.04996pt}\wedge}}}:=\varprojlim G_{\lambda},

where each GλG_{\lambda} has the discrete topology. Because the projection prλ:GGλ{\operatorname{pr}}_{\lambda}:{{{G}^{{\kern-1.04996pt}\wedge}}}\to G_{\lambda} is surjective and continuous, we identify GλG_{\lambda} with G/Nλ{{{G}^{{\kern-1.04996pt}\wedge}}}/N_{\lambda}, where NλN_{\lambda} is the kernel of the prλ{\operatorname{pr}}_{\lambda}, which is a clopen normal subgroup of G{{{G}^{{\kern-1.04996pt}\wedge}}}, since GλG_{\lambda} is finite and discrete. By Tychonoff’s theorem, G{{{G}^{{\kern-1.04996pt}\wedge}}} is a compact Hausdorff group, with a basis of open neighborhood of the unit element ee consisting of NλN_{\lambda}. Being Hausdorff, the intersection of NλN_{\lambda} is {e}\{e\}.

We prepare a well-known lemma for describing profinite Schurian association schemes.

Lemma 4.10.

Let G=limGλ=limG/Nλ{{{G}^{{\kern-1.04996pt}\wedge}}}=\varprojlim G_{\lambda}=\varprojlim{{{G}^{{\kern-1.04996pt}\wedge}}}/N_{\lambda} be a profinite group. Let HH and KK be closed subgroups of G{{{G}^{{\kern-1.04996pt}\wedge}}}. Give a quotient topology on H\G/KH\backslash{{{G}^{{\kern-1.04996pt}\wedge}}}/K by

q:GH\G/K.q:{{{G}^{{\kern-1.04996pt}\wedge}}}\to H\backslash{{{G}^{{\kern-1.04996pt}\wedge}}}/K.
  1. (1)

    The image of KK in GλG_{\lambda} is KNλKN_{\lambda} (which is a union of cosets of NλN_{\lambda}, and hence a subset of GλG_{\lambda}). The same statement holds for HH.

  2. (2)

    We have the following commutative diagram of continuous surjections, where α\alpha is a bijection.

    H\G/KHNλ\G/KNλαHNλ\Gλ/KNλ.\begin{array}[]{ccc}H\backslash{{{G}^{{\kern-1.04996pt}\wedge}}}/K&&\\ \downarrow&\searrow&\\ HN_{\lambda}\backslash{{{G}^{{\kern-1.04996pt}\wedge}}}/KN_{\lambda}&\stackrel{{\scriptstyle\alpha}}{{\to}}&HN_{\lambda}\backslash G_{\lambda}/KN_{\lambda}.\end{array} (4.4)
  3. (3)

    The q:GH\G/Kq:{{{G}^{{\kern-1.04996pt}\wedge}}}\to H\backslash{{{G}^{{\kern-1.04996pt}\wedge}}}/K is an open morphism, and H\G/KH\backslash{{{G}^{{\kern-1.04996pt}\wedge}}}/K is compact Hausdorff.

  4. (4)

    The finite sets HNλ\Gλ/KNλHN_{\lambda}\backslash G_{\lambda}/KN_{\lambda} constitute a projective system, with a canonical homeomorphism

    H\G/KlimHNλ\Gλ/KNλ.H\backslash{{{G}^{{\kern-1.04996pt}\wedge}}}/K\stackrel{{\scriptstyle\sim}}{{\to}}\varprojlim HN_{\lambda}\backslash G_{\lambda}/KN_{\lambda}.
Proof.

(1) When we consider GλG_{\lambda} as a quotient of G{{{G}^{{\kern-1.04996pt}\wedge}}}, GλG_{\lambda} is a set of cosets gNλgN_{\lambda}. The image of KK is the union of kNλkN_{\lambda} for kKk\in K, which is nothing but KNλKN_{\lambda}.

(2) Every set in the diagram is a quotient of G{{{G}^{{\kern-1.04996pt}\wedge}}} with various equivalence relation. The vertical arrow is obviously well-defined. The mapping α\alpha is bijective, since as a quotient of G{{{G}^{{\kern-1.04996pt}\wedge}}}, they are the same. The slanting arrow is merely the composition of the two. We recall that GH\G/K{{{G}^{{\kern-1.04996pt}\wedge}}}\to H\backslash{{{G}^{{\kern-1.04996pt}\wedge}}}/K is a quotient map. Thus, every map in the diagram is continuous, since the maps from G{{{G}^{{\kern-1.04996pt}\wedge}}} are continuous (where the mappings to the two sets at the bottom of the diagram factors through G/Nλ{{{G}^{{\kern-1.04996pt}\wedge}}}/N_{\lambda}, which is finite and discrete).

(3) The open sets sNλsN_{\lambda} for sGs\in{{{G}^{{\kern-1.04996pt}\wedge}}} and various λ\lambda constitute an open basis of G{{{G}^{{\kern-1.04996pt}\wedge}}}. The image of sNλsN_{\lambda} in H\G/KH\backslash{{{G}^{{\kern-1.04996pt}\wedge}}}/K is HsKNλHsKN_{\lambda} (since NλN_{\lambda} is normal), whose inverse image in G{{{G}^{{\kern-1.04996pt}\wedge}}} is again HsKNλHsKN_{\lambda}, which is open. Thus, by the definition of quotient topology, HsKNλHsKN_{\lambda} is open in H\G/KH\backslash{{{G}^{{\kern-1.04996pt}\wedge}}}/K, and hence qq is open.

Take two distinct elements HgKHgK, HgKHg^{\prime}K in H\G/KH\backslash{{{G}^{{\kern-1.04996pt}\wedge}}}/K. These are compact disjoint subsets of the Hausdorff space G{{{G}^{{\kern-1.04996pt}\wedge}}}. It is well known that there are an open set OaGO_{a}\subset{{{G}^{{\kern-1.04996pt}\wedge}}} containing HgKHgK and an open set ObHgKO_{b}\supset Hg^{\prime}K that separate HgKHgK and HgKHg^{\prime}K. By (1) of Lemma 3.14 for the union of HgKHgK and HgKHg^{\prime}K, we may take μ\mu such that open sets of the form sNμOasN_{\mu}\subset O_{a} with sNμHgKsN_{\mu}\cap HgK\neq\emptyset cover HgKHgK and open sets sNμObsN_{\mu}\subset O_{b} with sNμHgKsN_{\mu}\cap Hg^{\prime}K\neq\emptyset cover HgKHg^{\prime}K, that is, HgKNμOaHgKN_{\mu}\subset O_{a} and HgKNμObHg^{\prime}KN_{\mu}\subset O_{b} are disjoint. Since qq is open, HgKNμHgKN_{\mu} and HgKNμHg^{\prime}KN_{\mu} are open sets in H\G/KH\backslash{{{G}^{{\kern-1.04996pt}\wedge}}}/K which have no intersection and are neighborhoods of HgKHgK, HgKHg^{\prime}K (which are considered as elements of H\G/KH\backslash{{{G}^{{\kern-1.04996pt}\wedge}}}/K), respectively. Thus H\G/KH\backslash{{{G}^{{\kern-1.04996pt}\wedge}}}/K is Hausdorff.

(4) If we see these sets as quotients of G{{{G}^{{\kern-1.04996pt}\wedge}}} as in (4.4), it is clear that these form a projective system. By the universality of projective limit, we have a continuous map

f:H\G/KlimHNλ\Gλ/KNλf:H\backslash{{{G}^{{\kern-1.04996pt}\wedge}}}/K\to\varprojlim HN_{\lambda}\backslash G_{\lambda}/KN_{\lambda}

by (2). The image of ff is dense, since for any element ξ\xi in the projective limit and its open neighborhood, there is a smaller open neighborhood which has the form prλ1prλ(ξ){\operatorname{pr}}_{\lambda}^{-1}{\operatorname{pr}}_{\lambda}(\xi) for some λ\lambda, since these form a basis of open neighborhoods of ξ\xi. Because

H\G/KHNλ\Gλ/KNλH\backslash{{{G}^{{\kern-1.04996pt}\wedge}}}/K\to HN_{\lambda}\backslash G_{\lambda}/KN_{\lambda}

is surjective, the density of the image follows. Since H\G/KH\backslash{{{G}^{{\kern-1.04996pt}\wedge}}}/K is compact (being a quotient of a compact set G{{{G}^{{\kern-1.04996pt}\wedge}}}), the image is compact and closed, hence ff is surjective. Since H\G/KH\backslash{{{G}^{{\kern-1.04996pt}\wedge}}}/K is Hausdorff, any two distinct points HgKHgK, HgKHg^{\prime}K have disjoint open neighborhoods HgKNλHgKN_{\lambda}, HgKNλHg^{\prime}KN_{\lambda}, respectively, as proved above, hence their images in HNλ\Gλ/KNλHN_{\lambda}\backslash G_{\lambda}/KN_{\lambda} are distinct, which implies the injectivity of ff. By the compactness of H\G/KH\backslash{{{G}^{{\kern-1.04996pt}\wedge}}}/K, ff is closed, hence f1f^{-1} is continuous, thus a homeomorphism. ∎

Proposition 4.11.

(profinite Schurian association scheme) Let G=limGλ=limG/Nλ{{{G}^{{\kern-1.04996pt}\wedge}}}=\varprojlim G_{\lambda}=\varprojlim{{{G}^{{\kern-1.04996pt}\wedge}}}/N_{\lambda} be a profinite group. We denote by GλG_{\lambda} the quotient G/Nλ{{{G}^{{\kern-1.04996pt}\wedge}}}/N_{\lambda}. For a closed subgroup H<GH<G, its image in GλG_{\lambda} is Hλ:=HNλH_{\lambda}:=HN_{\lambda}. Then we have a projective system of Schurian association schemes

Xλ:=Gλ/HλX_{\lambda}:=G_{\lambda}/H_{\lambda}

and

Iλ:=Hλ\Gλ/Hλ.I_{\lambda}:=H_{\lambda}\backslash G_{\lambda}/H_{\lambda}.

These systems give a profinite association scheme, which may be non-commutative. We call this type of profinite association scheme Schurian. We have canonical homeomorphisms

G/HX=limXλ=limG/HNλ{{{G}^{{\kern-1.04996pt}\wedge}}}/H\stackrel{{\scriptstyle\sim}}{{\to}}{{{X}^{{\kern-1.04996pt}\wedge}}}=\varprojlim X_{\lambda}=\varprojlim{{{G}^{{\kern-1.04996pt}\wedge}}}/HN_{\lambda}

and

H\G/HI=limIλ=limHλ\G/Hλ.H\backslash{{{G}^{{\kern-1.04996pt}\wedge}}}/H\stackrel{{\scriptstyle\sim}}{{\to}}{{{I}^{{\kern-1.04996pt}\wedge}}}=\varprojlim I_{\lambda}=\varprojlim H_{\lambda}\backslash{{{G}^{{\kern-1.04996pt}\wedge}}}/H_{\lambda}.
Proof.

Everything is proved in Proposition 4.2 and Lemma 4.10, except that these mappings give surjective morphisms of association schemes. Take λμ\lambda\geq\mu. Since the surjectivity is proved in (2), it suffices to show the commutativity of

Xλ×XλIλXμ×XμIμ,\begin{matrix}X_{\lambda}\times X_{\lambda}&\to&I_{\lambda}\\ \downarrow&&\downarrow\\ X_{\mu}\times X_{\mu}&\to&I_{\mu},\end{matrix}

which is easy to check: an element (gHNλ,gHNλ)(gHN_{\lambda},g^{\prime}HN_{\lambda}) at the left top is mapped to the right Hg1gHNλHg^{-1}g^{\prime}HN_{\lambda}, then to the bottom Hg1gHNμHg^{-1}g^{\prime}HN_{\mu}. This is the same via the left bottom corner. ∎

Proposition 4.12.

Let G/H{{{G}^{{\kern-1.04996pt}\wedge}}}/H be a Schurian profinite association scheme as in 4.11. Suppose that each finite association scheme XλX_{\lambda} is commutative. Then, J{{{J}^{{\kern-1.04996pt}\wedge}}} is the set of irreducible representations of G{{{G}^{{\kern-1.04996pt}\wedge}}} appearing in C(Xλ)C(X_{\lambda}) for some λ\lambda, where the multiplicity is at most one for any C(Xλ)C(X_{\lambda}).

Proof.

We have an inductive system of injections C(Xλ×Xλ)C(X_{\lambda}\times X_{\lambda}), which gives via (4.2) an inductive system of injections

Hom(C(Xμ),C(Xμ))Hom(C(Xλ),C(Xλ))\operatorname{Hom}_{\mathbb{C}}(C(X_{\mu}),C(X_{\mu}))\to\operatorname{Hom}_{\mathbb{C}}(C(X_{\lambda}),C(X_{\lambda}))

for λμ\lambda\geq\mu. To obtain AXμA_{X_{\mu}} and AXλA_{X_{\lambda}}, we take G{{{G}^{{\kern-1.04996pt}\wedge}}}-invariant part (4.3)

HomGμ(C(Xμ),C(Xμ))HomGλ(C(Xλ),C(Xλ))\operatorname{Hom}_{G_{\mu}}(C(X_{\mu}),C(X_{\mu}))\to\operatorname{Hom}_{G_{\lambda}}(C(X_{\lambda}),C(X_{\lambda}))

(note that there is a canonical surjection GGλ{{{G}^{{\kern-1.04996pt}\wedge}}}\to G_{\lambda}, hence a GλG_{\lambda}-module is a G{{{G}^{{\kern-1.04996pt}\wedge}}}-module.) By the surjectivity of GGλ{{{G}^{{\kern-1.04996pt}\wedge}}}\to G_{\lambda}, an irreducible representation in C(Xμ)C(X_{\mu}) of GμG_{\mu} lifts to an irreducible representation in C(Xλ)C(X_{\lambda}) of GλG_{\lambda}.

We assume that every finite association scheme appeared is commutative. Then, the multiplicity of each irreducible representation VC(Xλ)V\subset C(X_{\lambda}) is one by Corollary 4.8 for any λ\lambda. Thus, the partial surjection given in Corollary 2.14 is in this case the inverse to the injection JμJλJ_{\mu}\to J_{\lambda}, induced by GλGμG_{\lambda}\to G_{\mu}, and

J=limJλ{{{J}^{{\kern-1.04996pt}\wedge}}}=\varinjlim J_{\lambda}

is identified with the set of irreducible representations of G{{{G}^{{\kern-1.04996pt}\wedge}}} appeared in some C(Xλ)C(X_{\lambda}). ∎

Indeed, the above irreducible representations are exactly the irreducible representations of G{{{G}^{{\kern-1.04996pt}\wedge}}} appearing in C(G/H)C({{{G}^{{\kern-1.04996pt}\wedge}}}/H). To show this, we recall a well known “no small subgroups” lemma.

Lemma 4.13.

Let GG be a Lie group over {\mathbb{R}} or {\mathbb{C}}. Then, there is an open neighborhood of the unit ee, which contains no subgroup except {e}\{e\}.

Proof.

We consider the Lie algebra 𝔤\mathfrak{g} of GG. Give an arbitrary (Hermitian or positive definite) metric to 𝔤\mathfrak{g}. It is known that

exp:𝔤G\exp:\mathfrak{g}\to G

is homeomorphism when restricted to a small enough neighborhood of 0. We take the homeomorphic neighborhoods UU of eGe\in G and VV of 0𝔤0\in\mathfrak{g}. We may assume that VV is bounded with respect to the metric. Assume that UU contains non-trivial subgroup of GG. Take an egGe\neq g\in G in the subgroup. Consequently, for any integer nn, gnUg^{n}\in U. If h=log(g)V𝔤h=\log(g)\in V\subset\mathfrak{g}, nh=log(gn)nh=\log(g^{n}) corresponds to gng^{n} and hence nhVnh\in V for any nn. However, VV is bounded, which contradicts the assumption. ∎

Corollary 4.14.

Let G{{{G}^{{\kern-1.04996pt}\wedge}}} be a profinite group, and f:GGf:{{{G}^{{\kern-1.04996pt}\wedge}}}\to G a group morphism to a real or complex Lie group GG. Then, Ker(f)\operatorname{Ker}(f) is an open normal subgroup of G{{{G}^{{\kern-1.04996pt}\wedge}}}, and the image f(G)f({{{G}^{{\kern-1.04996pt}\wedge}}}) in GG is a finite subgroup of GG.

Proof.

Take a small enough neighborhood of UU in Lemma 4.13. Consider H=f1(U)GH=f^{-1}(U)\subset{{{G}^{{\kern-1.04996pt}\wedge}}}, which is open in G{{{G}^{{\kern-1.04996pt}\wedge}}}. Thus, there exists λ\lambda such that the open normal subgroup N=NλN=N_{\lambda} is contained in HH. The image f(N)<Gf(N)<G is a subgroup in UU, which must be {e}G\{e\}\subset G by definition of UU. Thus, Ker(f)\operatorname{Ker}(f) contains NN, and is a union of cosets of NN, hence an open normal subgroup of G{{{G}^{{\kern-1.04996pt}\wedge}}}. Since G{{{G}^{{\kern-1.04996pt}\wedge}}} is compact and Kerf\operatorname{Ker}f is open, f(G)G/Kerff({{{G}^{{\kern-1.04996pt}\wedge}}})\cong{{{G}^{{\kern-1.04996pt}\wedge}}}/\operatorname{Ker}f is a finite group. ∎

Proposition 4.15.

Let GG be a compact Hausdorff group, and HH a closed subgroup. Consider the representation of GG on C(G/H)C(G/H). Then, each irreducible component of C(G/H)C(G/H) is finite dimensional.

Proof.

This follows from a variant of the Peter-Weyl theorem, see Takeuchi[24, Section 1]. ∎

Proposition 4.16.

Let G{{{G}^{{\kern-1.04996pt}\wedge}}} be a profinite group and HH a closed subgroup. Consider the profinite association scheme arising from G/H{{{G}^{{\kern-1.04996pt}\wedge}}}/H. Let J(G/H)J({{{G}^{{\kern-1.04996pt}\wedge}}}/H) be the set of the irreducible representations of G{{{G}^{{\kern-1.04996pt}\wedge}}} on C(G/H)C({{{G}^{{\kern-1.04996pt}\wedge}}}/H). Then, all finite association schemes XλX_{\lambda} in Proposition 4.11 are commutative, if and only if the representation C(G/H)C({{{G}^{{\kern-1.04996pt}\wedge}}}/H) is multiplicity free. In this case, there is a canonical bijection

f:J(G/H)J=limJλ.f:J({{{G}^{{\kern-1.04996pt}\wedge}}}/H)\stackrel{{\scriptstyle\sim}}{{\to}}{{{J}^{{\kern-1.04996pt}\wedge}}}=\varinjlim J_{\lambda}. (4.5)
Proof.

Put X:=G/H{{{X}^{{\kern-1.04996pt}\wedge}}}:={{{G}^{{\kern-1.04996pt}\wedge}}}/H. Let VC(X)V\subset C({{{X}^{{\kern-1.04996pt}\wedge}}}) be an irreducible representation of G{{{G}^{{\kern-1.04996pt}\wedge}}}. By Proposition 4.15, VV is finite dimensional. By Lemma 4.13, the representation factors through

GG/NλGL(V)G\to G/N_{\lambda}\to{\mathrm{GL}}(V)

for some λ\lambda. By Lemma 4.1, this means that VC(X)V\subset C({{{X}^{{\kern-1.04996pt}\wedge}}}) lies in

VC(Nλ\X)=C(Xλ)C(X).V\subset C(N_{\lambda}\backslash{{{X}^{{\kern-1.04996pt}\wedge}}})=C(X_{\lambda})\subset C({{{X}^{{\kern-1.04996pt}\wedge}}}).

Thus, C(X)C({{{X}^{{\kern-1.04996pt}\wedge}}}) is multiplicity free, if and only if C(Xλ)C(X_{\lambda}) is multiplicity free for every λΛ\lambda\in\Lambda, in other words, if and only if every XλX_{\lambda} is a commutative association scheme, by Proposition 4.12. In the commutative case, the following observation shows that every irreducible component of C(X)C({{{X}^{{\kern-1.04996pt}\wedge}}}) appears in C(Xλ)C(X_{\lambda}) for some λ\lambda. Because the multiplicity is one, for any VV above, there is a unique representation isomorphic to VV in limC(Xλ)\varinjlim C(X_{\lambda}). This gives a map ff in (4.5). The injectivity follows from that C(X)C({{{X}^{{\kern-1.04996pt}\wedge}}}) is multiplicity-free. The surjectivity follows from C(Xλ)C(X)C(X_{\lambda})\subset C({{{X}^{{\kern-1.04996pt}\wedge}}}) is a G{{{G}^{{\kern-1.04996pt}\wedge}}}-submodule, and any GλG_{\lambda}-irreducible component of C(Xλ)C(X_{\lambda}) is a G{{{G}^{{\kern-1.04996pt}\wedge}}}-irreducible component since GGλ{{{G}^{{\kern-1.04996pt}\wedge}}}\to G_{\lambda} is surjective. ∎

Remark 4.17.

Kurihara-Okuda[13] gives the notion of Bose-Mesner algebra for general homogeneous space G/HG/H, where GG is compact Hausdorff and HH is a closed subgroup. In the case where GG is a profinite group and C(G/H)C(G/H) is multiplicity free, their construction coincides with our definition, by Proposition 4.16.

4.1.2. Profinite abelian groups

For any finite abelian group XX, we consider its thin scheme. Namely, R:X×XI=XR:X\times X\to I=X with (x,y)yx(x,y)\mapsto y-x. Then, the ii-th adjacency matrix R1(i)R^{-1}(i) is the representation matrix of addition of ii on C(X)C(X), and the Bose-Mesner algebra is isomorphic to the group ring C(X)C(X), with usual Hadamard product and the convolution product

(fg)(x)=1#Xyf(xy)g(y)(f*g)(x)=\frac{1}{\#X}\sum_{y}f(x-y)g(y)

(the factor 1#X\frac{1}{\#X} is not the standard, incorporated to be compatible with Theorem 2.5). Primitive idempotents are exactly the characters ξXˇ\xi\in{{\check{X}}}, where XˇC(X){{\check{X}}}\subset C(X) is the group of characters of XX.

For a profinite abelian group X{{{X}^{{\kern-1.04996pt}\wedge}}}, it is a projective limit of finite abelian groups XλX_{\lambda}. We have an inductive system of the character groups Jλ:=XλˇJ_{\lambda}:=\check{X_{\lambda}}. For λμ\lambda\geq\mu, the partial map pλ,μ:JλJμp_{\lambda,\mu}:J_{\lambda}\rightharpoonup J_{\mu} is given by the injection iμ,λ:JμJλi_{\mu,\lambda}:J_{\mu}\to J_{\lambda}, where the domain of pλ,μp_{\lambda,\mu} is the image of iμ,λi_{\mu,\lambda}, and in the domain pλ,μp_{\lambda,\mu} is the inverse of iμ,λi_{\mu,\lambda}. Thus, I{{{I}^{{\kern-1.04996pt}\wedge}}} is isomorphic to X{{{X}^{{\kern-1.04996pt}\wedge}}}, J{{{J}^{{\kern-1.04996pt}\wedge}}} is the inductive limit of the character groups Xλˇ\check{X_{\lambda}}. The Bose-Mesner algebra is isomorphic to Cc(J)C_{c}({{{J}^{{\kern-1.04996pt}\wedge}}}), the space of finite linear combinations of characters of X{{{X}^{{\kern-1.04996pt}\wedge}}}, which is isomorphic to Clc(I)C_{lc}({{{I}^{{\kern-1.04996pt}\wedge}}}), the space of locally constant functions on IX{{{I}^{{\kern-1.04996pt}\wedge}}}\simeq{{{X}^{{\kern-1.04996pt}\wedge}}}. This example is outside of the scope of Barg-Skriganov theory, if IX{{{I}^{{\kern-1.04996pt}\wedge}}}\simeq{{{X}^{{\kern-1.04996pt}\wedge}}} is uncountable.

Remark 4.18.

Since J{{{J}^{{\kern-1.04996pt}\wedge}}} is discrete, Cc(J)C_{c}({{{J}^{{\kern-1.04996pt}\wedge}}}) has a natural (and orthogonal) basis consisting of EjE_{j} (jJj\in{{{J}^{{\kern-1.04996pt}\wedge}}}), but for the canonically isomorphic vector space Clc(I)C_{lc({{{I}^{{\kern-0.74997pt}\wedge}}})}, we do not have a natural basis, since I{{{I}^{{\kern-1.04996pt}\wedge}}} is compact with (possibly) uncountable cardinality. This makes a definition of eigenmatrices difficult.

4.2. Kernel schemes

Another example of profinite association schemes is a projective system of the kernel schemes. The kernel schemes are finite association schemes defined in Martin-Stinson[16] to study (t,m,s)(t,m,s)-nets. The (t,m,s)(t,m,s)-nets are introduced by Niederreiter for quasi-Monte Carlo integrations, see [18].

We recall the kernel scheme, with a slight modification on the notation for II, to make a description as a projective system easier.

Definition 4.19.

Let nn be a positive integer, and VV be a finite set of alphabet with cardinality v2v\geq 2. Let XnX_{n} be VnV^{n}, and In:={1,2,,n}{}I_{n}:=\{1,2,\ldots,n\}\cup\{\infty\}. Define Rn:Xn×XnInR_{n}:X_{n}\times X_{n}\to I_{n} as follows. Let x=(x1,x2,,xn)x=(x_{1},x_{2},\ldots,x_{n}) and y=(y1,y2,,yn)y=(y_{1},y_{2},\ldots,y_{n}) be elements of XnX_{n}. Let R(x,y)R(x,y) be the smallest index ii for which xiyix_{i}\neq y_{i}. If x=yx=y, then R(x,y)=R(x,y)=\infty. This is a symmetric (and hence commutative) association scheme, with R1()R^{-1}(\infty) being the identity relation. This is called a kernel scheme, and denoted by k(n,v)\overrightarrow{k(n,v)}.

A direct computation of the intersection numbers shows that this is a commutative association scheme, as shown in [16]. We shall give an indirect proof in Corollary 4.27. (This is only for the self-containedness.)

Proposition 4.20.

The kernel schemes k(n,v)\overrightarrow{k(n,v)} forms a projective system, where Xn+1XnX_{n+1}\to X_{n} is given by deleting the right most component, and In+1InI_{n+1}\to I_{n} is given by mmm\mapsto m for mnm\leq n, and the both n+1,n+1,\infty are mapped to \infty. Then X=V>0{{{X}^{{\kern-1.04996pt}\wedge}}}=V^{{\mathbb{N}}_{>0}} holds, and prn:XXn{\operatorname{pr}}_{n}:{{{X}^{{\kern-1.04996pt}\wedge}}}\to X_{n} is obtained by taking the left (first) nn components, while I{{{I}^{{\kern-1.04996pt}\wedge}}} is >0{}{\mathbb{N}}_{>0}\cup\{\infty\}, with topology obtained by: each n>0n\in{\mathbb{N}}_{>0} is clopen, and an open set containing \infty is a complement of a finite subset of >0{\mathbb{N}}_{>0}, that is, I{{{I}^{{\kern-1.04996pt}\wedge}}} is the one-point compactification of >0{\mathbb{N}}_{>0}. We call this a pro-kernel scheme and denote by k(,v)\overrightarrow{k(\infty,v)}.

Proof.

It is easy to check that these make a projective system of association schemes. The topology of I{{{I}^{{\kern-1.04996pt}\wedge}}} comes from the definition of projective limit topology. ∎

Remark 4.21.

In this case I{{{I}^{{\kern-1.04996pt}\wedge}}} is countable, and this is a special case of infinite association schemes examined in detail, named metric schemes, by Barg-Skriganov [2, Section 8], for which the adjacency algebra is defined. In particular, in (8.33) they showed that it is an inductive limit of finite dimensional adjacency algebra, which indeed coincides with our definition.

Remark 4.22.

This is merely an observation. For the set of pp-adic integers p{\mathbb{Z}}_{p}, we have a valuation v:p{}v:{\mathbb{Z}}_{p}\to{\mathbb{N}}\cup\{\infty\} giving an ultrametric. The above pro-kernel scheme is isomorphic to this metric, up to the difference by one on the metric, i.e.,

R:p×p{},(x,y)v(xy).R:{\mathbb{Z}}_{p}\times{\mathbb{Z}}_{p}\to{\mathbb{N}}\cup\{\infty\},(x,y)\mapsto v(x-y).

The same holds for a ring of formal power series 𝔽q[[t]]{\mathbb{F}}_{q}[[t]], where 𝔽q{\mathbb{F}}_{q} denotes the finite field of qq elements. The both yield the isomorphic pro-kernel schemes if p=qp=q.

To see J{{{J}^{{\kern-1.04996pt}\wedge}}} of the pro-kernel scheme, we need to work with primitive idempotents. All necessary results are in Martin-Stinson[16], but because of the choice of the index II and normalization of the products necessary for making the projective system, the obtained constants slightly differ. To help the readers’ understandings, we recall the methods of Martin-Stinson. For the chosen integer v2v\geq 2, let V:=ZvV:=Z_{v} be the additive group /v{\mathbb{Z}}/v. Then X=ZvnX=Z_{v}^{n} is an additive group, and the kernel scheme is a translation scheme [4, §2.10].

Definition 4.23.

Let XX be a finite abelian group. A translation scheme is an association scheme R:X×XIR:X\times X\to I, which factors through X×XX,(x,y)yxX\times X\to X,(x,y)\mapsto y-x, namely, there is an S:XIS:X\to I such that R(x,y)=S(yx)R(x,y)=S(y-x).

Definition 4.24.

Return to the kernel scheme. We define bot:XnIn{{\operatorname{bot}}}:X_{n}\to I_{n} by

bot(x1,,xn):=min{ixi0},{{\operatorname{bot}}}(x_{1},\ldots,x_{n}):=\min\{i\mid x_{i}\neq 0\},

and bot(0,,0)={{\operatorname{bot}}}(0,\cdots,0)=\infty. Then we have

Rn:Xn×XnIn,(x,y)bot(xy).R_{n}:X_{n}\times X_{n}\to I_{n},\quad(x,y)\mapsto{{\operatorname{bot}}}(x-y).

By S:=botS:={{\operatorname{bot}}}, the kernel scheme is a translation scheme.

Let (X,R,I)(X,R,I) be a translation scheme. For iIi\in I, let SiXS_{i}\subset X be S1(i)S^{-1}(i) (SS for the sphere). Let Xˇ{{\check{X}}} be the dual group of XX. Then, ξXˇ\xi\in{{\check{X}}} is an orthonomal basis of C(X)C(X) under the normalized inner product in Definition 3.16. Let AiA_{i} denote the ii-th adjacency matrix, and χi:X{0,1}\chi_{i}:X\to\{0,1\} be the indicator function for SiS_{i}. Our action in Theorem 2.5 shows that for xXx\in X

(Aiξ)(x)\displaystyle(A_{i}\bullet\xi)(x) =\displaystyle= 1#XyXAi(x,y)ξ(y)\displaystyle\frac{1}{\#X}\sum_{y\in X}A_{i}(x,y)\xi(y)
=\displaystyle= 1#XyXχi(yx)ξ(y)\displaystyle\frac{1}{\#X}\sum_{y\in X}\chi_{i}(y-x)\xi(y)
=\displaystyle= 1#Xyx+Siξ(y)\displaystyle\frac{1}{\#X}\sum_{y\in x+S_{i}}\xi(y)
=\displaystyle= 1#XaSiξ(x+a)\displaystyle\frac{1}{\#X}\sum_{a\in S_{i}}\xi(x+a)
=\displaystyle= 1#X(aSiξ(a))ξ(x).\displaystyle\frac{1}{\#X}(\sum_{a\in S_{i}}\xi(a))\xi(x).

Thus ξ\xi is an eigenvector with eigenvalue

1#X(aSiξ(a)).\frac{1}{\#X}(\sum_{a\in S_{i}}\xi(a)). (4.6)

The primitive idempotents are projectors to the common eigen spaces of AiA_{i} for all iIi\in I. Let us assume X=ZvnX=Z_{v}^{n}, and then Xˇ=Zvˇn{{\check{X}}}={{\check{Z_{v}}}}^{n}, where the product of Zvˇ{{\check{Z_{v}}}} is written multiplicatively with unit 11. Thus,

(ξ1,,ξn)(x1,,xn)=i=1nξi(xi).(\xi_{1},\ldots,\xi_{n})(x_{1},\ldots,x_{n})=\prod_{i=1}^{n}\xi_{i}(x_{i}). (4.7)
Definition 4.25.

For ξ=(ξ1,,ξn)Xˇ\xi=(\xi_{1},\ldots,\xi_{n})\in{{\check{X}}}, define top(ξ){{\operatorname{top}}}(\xi) as the max ii such that ξi1\xi_{i}\neq 1. If ξ=1\xi=1, then top(ξ)=0{{\operatorname{top}}}(\xi)=0. Thus top:Xˇ{0,1,,n}{{\operatorname{top}}}:{{\check{X}}}\to\{0,1,\ldots,n\}. We shall denote Jn:={0,1,,n}J_{n}:=\{0,1,\ldots,n\}, because we shall soon show a natural correspondence between the set of primitive idempotents and JnJ_{n}.

Lemma 4.26.

For ξ\xi with j=top(ξ)j={{\operatorname{top}}}(\xi), we denote by pi(j)p_{i}(j) the eigenvalue of AiA_{i} for ξ\xi. Then

pi(j)={vn if i=,(v1)vi if j<i<,vi if j=i,0 if j>i.p_{i}(j)=\begin{cases}v^{-n}&\mbox{ if }i=\infty,\\ (v-1)v^{-i}&\mbox{ if }j<i<\infty,\\ -v^{-i}&\mbox{ if }j=i,\\ 0&\mbox{ if }j>i.\\ \end{cases}

Thus, for any AiA_{i}, the eigenvalue of ξXˇ\xi\in{{\check{X}}} depends only on top(ξ){{\operatorname{top}}}(\xi), and if top(ξ)top(ξ){{\operatorname{top}}}(\xi)\neq{{\operatorname{top}}}(\xi^{\prime}), then there is an ii such that AiA_{i} has different eigenvalues for ξ\xi and ξ\xi^{\prime}.

Proof.

Recall that I={1,2,,n}{}I=\{1,2,\ldots,n\}\cup\{\infty\}. Put

Bi:=iS.B_{i}:=\bigcup_{i\leq\ell\leq\infty}S_{\ell}.

(BiB_{i} is the ball consisting of all the elements (x1,,xn)(x_{1},\ldots,x_{n}) with x=0x_{\ell}=0 for all <i\ell<i, with the cardinality vni+1v^{n-i+1} for i<i<\infty and 11 for i=i=\infty.) Fix ii and ξXˇ\xi\in{{\check{X}}} with top(ξ)=j{{\operatorname{top}}}(\xi)=j, then by (4.6) we have

τ(i,j)ξ\displaystyle\tau(i,j)\xi :=\displaystyle:= iAiξ\displaystyle\sum_{i\leq\ell\leq\infty}A_{i}\bullet\xi
=\displaystyle= 1vn(aBiξ(a))ξ\displaystyle\frac{1}{v^{n}}(\sum_{a\in B_{i}}\xi(a))\xi
=\displaystyle= {vnξ if i=,vi+1ξ if j<i<,0 if ji.\displaystyle\begin{cases}v^{-n}\xi&\mbox{ if }i=\infty,\\ v^{-i+1}\xi&\mbox{ if }j<i<\infty,\\ 0&\mbox{ if }j\geq i.\\ \end{cases}

The last equality follows from that for i=i=\infty, B={0}B_{\infty}=\{0\} and hence the summation is aBiξ(a)=1\sum_{a\in B_{i}}\xi(a)=1, and if j<i<j<i<\infty, then (4.7) is always 11 and the cardinality of BiB_{i} is vni+1v^{n-i+1}, while if jij\geq i, ξ(a)\xi(a) depends on aa and the summation over aBia\in B_{i} is zero, since ξ\xi is a non-trivial character on the abelian group BiB_{i}. Thus, the eigenvalue of AiA_{i} for ξ\xi is the summation over Si=BiBi+1S_{i}=B_{i}\setminus B_{i+1}, that is,

{vn if i=,(τ(i,j)τ(i+1,j))=(v1)vi if j<i<,(τ(i,j)τ(i+1,j))=vi if i=j,0 if j>i,\begin{cases}v^{-n}&\mbox{ if }i=\infty,\\ (\tau(i,j)-\tau(i+1,j))=(v-1)v^{-i}&\mbox{ if }j<i<\infty,\\ (\tau(i,j)-\tau(i+1,j))=-v^{-i}&\mbox{ if }i=j,\\ 0&\mbox{ if }j>i,\end{cases}

as desired. ∎

Corollary 4.27.

The kernel scheme k(n,v)\overrightarrow{k(n,v)} is a commutative association scheme. A primitive idempotent is the projection to the subspace C(X)jC(X)_{j} of C(X)C(X) spanned by {ξtop(ξ)=j}\{\xi\mid{{\operatorname{top}}}(\xi)=j\} for each jJnj\in J_{n}, which identifies the set of primitive idempotents with JnJ_{n}.

Proof.

Recall that we did not use the property of association schemes for the kernel schemes so far. Every AiA_{i} has C(X)jC(X)_{j} as an eigenspace with an eigenvalue. Since C(X)C(X) is a direct sum of C(X)jC(X)_{j} for jJnj\in J_{n}, this implies that the matrices AiA_{i} mutually commute. We consider the algebra generated by AiA_{i} by the matrix product. Each element of this algebra is determined by the n+1n+1 eigenvalues corresponding to JnJ_{n}. Thus, the dimension of the algebra as a vector space does not exceed n+1n+1, but AiA_{i}’s are n+1n+1 linearly independent matrices, hence their linear combination is closed under the matrix multiplication. The rest of the axioms of association schemes are easy to verify. The description of primitive idempotents immediately follows. ∎

Corollary 4.28.

The multiplicity mjm_{j} is 11 for j=0j=0 and mj=(v1)vj1m_{j}=(v-1)v^{j-1} for j>0j>0. The valency kik_{i} is 11 for i=i=\infty and ki=(v1)vnik_{i}=(v-1)v^{n-i} for i<i<\infty.

Proof.

The multiplicity is the number of ξ\xi with top(ξ)=j{{\operatorname{top}}}(\xi)=j, which implies the first statement. The valency is the cardinality of the sphere SiS_{i}, which implies the second. ∎

Corollary 4.29.

For jJnj\in J_{n}, the projector EjE_{j} is given by

Ej(x,y)\displaystyle E_{j}(x,y) =\displaystyle= ξXˇ,top(ξ)=jξ(x)ξ(y)¯\displaystyle\sum_{\xi\in{{\check{X}}},{{\operatorname{top}}}(\xi)=j}\xi(x)\overline{\xi(y)}
=\displaystyle= ξXˇ,top(ξ)=jξ(xy).\displaystyle\sum_{\xi\in{{\check{X}}},{{\operatorname{top}}}(\xi)=j}\xi(x-y).

We have

Ai=jJnpi(j)EjA_{i}=\sum_{j\in J_{n}}p_{i}(j)E_{j}

where pi(j)p_{i}(j) is defined in Lemma 4.26.

Proof.

The first half describes the projection to the eigenspace corresponding to jj by an orthonomal basis. The second is because AiA_{i} has the eigenvalue pi(j)p_{i}(j) for that eigenspace. ∎

For xySix-y\in S_{i}, we define

qj(i)\displaystyle q_{j}(i) :=\displaystyle:= Ej(x,y)=ξXˇ,top(ξ)=jξ(xy)\displaystyle E_{j}(x,y)=\sum_{\xi\in{{\check{X}}},{{\operatorname{top}}}(\xi)=j}\xi(x-y)
=\displaystyle= {1 if 0=j(v1)vj1 if 0<j<ivj1 if j=i0 if j>i.\displaystyle\begin{cases}1&\mbox{ if }0=j\\ (v-1)v^{j-1}&\mbox{ if }0<j<i\leq\infty\\ -v^{j-1}&\mbox{ if }j=i\\ 0&\mbox{ if }j>i.\end{cases}

This computation is dual to that for pi(j)p_{i}(j), so omitted.

Corollary 4.30.
Ej=iIqj(i)Ai.E_{j}=\sum_{i\in I}q_{j}(i)A_{i}.
Proof.

We evaluate the both sides at (x,y)(x,y). The left hand side is qj(i)q_{j}(i) for xySix-y\in S_{i}, which is equal to the right hand side, since Ai(x,y)=χi(xy)A_{i}(x,y)=\chi_{i}(x-y). ∎

Corollary 4.31.

In k(,v)\overrightarrow{k(\infty,v)}, the set J{{{J}^{{\kern-1.04996pt}\wedge}}} is the inductive limit of JnJn+1J_{n}\to J_{n+1}, and hence naturally identified with {\mathbb{N}}.

Now we pass to the pro-kernel scheme, i.e., take the inductive limit of Ak(n,v)A_{\overrightarrow{k(n,v)}} to obtain AX=Ak(,v)A_{{{{X}^{{\kern-0.74997pt}\wedge}}}}=A_{\overrightarrow{k(\infty,v)}}.

Proposition 4.32.

Let XnX_{n} denote the kernel scheme k(n,v)\overrightarrow{k(n,v)} with InI_{n} and primitive idempotents JnJ_{n}, and X{{{X}^{{\kern-1.04996pt}\wedge}}} the pro-kernel scheme k(,v)\overrightarrow{k(\infty,v)} with I{{{I}^{{\kern-1.04996pt}\wedge}}} and the primitive idempotents J{{{J}^{{\kern-1.04996pt}\wedge}}}. For iI{}i\in{{{I}^{{\kern-1.04996pt}\wedge}}}\setminus\{\infty\}, we denote by AiA_{i} the image of δiClc(I)\delta_{i}\in C_{lc}({{{I}^{{\kern-1.04996pt}\wedge}}}) by the isomorphism Clc(I)AXC_{lc}({{{I}^{{\kern-1.04996pt}\wedge}}})\to A_{{{X}^{{\kern-0.74997pt}\wedge}}}. Note that δClc(I)\delta_{\infty}\notin C_{lc({{{I}^{{\kern-0.74997pt}\wedge}}})}. Through AXnAXA_{X_{n}}\to A_{{{{X}^{{\kern-0.74997pt}\wedge}}}}, EjE_{j} is mapped to EjE_{j}, and AiA_{i} for i{1,2,,n}i\in\{1,2,\ldots,n\} is mapped to AiA_{i}, while AAXnA_{\infty}\in A_{X_{n}} is mapped to an element corresponding to the indicator function χ{n+1,,}Clc(I)\chi_{\{n+1,\ldots,\infty\}}\in C_{lc}({{{I}^{{\kern-1.04996pt}\wedge}}}). It is convenient to denote the corresponding element in AXA_{{{X}^{{\kern-0.74997pt}\wedge}}} by the symbol

i{n+1,,}Ai,\sum_{i\in\{n+1,\ldots,\infty\}}A_{i},

since

χ{n+1,,}=i{n+1,,}δi\chi_{\{n+1,\ldots,\infty\}}=\sum_{i\in\{n+1,\ldots,\infty\}}\delta_{i}

holds. (Note that AA_{\infty} does not exist in AXA_{{{X}^{{\kern-0.74997pt}\wedge}}}.)

Proof.

According to Proposition 4.20 we have I=>0{}{{{I}^{{\kern-1.04996pt}\wedge}}}={\mathbb{N}}_{>0}\cup\{\infty\} and J={{{J}^{{\kern-1.04996pt}\wedge}}}={\mathbb{N}} by Corollary 4.31. Since JnJJ_{n}\to{{{J}^{{\kern-1.04996pt}\wedge}}} is an injection, EjAXnE_{j}\in A_{X_{n}} is mapped to EjAXE_{j}\in A_{{{X}^{{\kern-0.74997pt}\wedge}}} (see Proposition 3.24). The projection IIn{{{I}^{{\kern-1.04996pt}\wedge}}}\to I_{n} is obtained by mapping iii\mapsto i for 1in1\leq i\leq n, and ii\mapsto\infty for i>ni>n. Since the indicator function δiC(In)\delta_{i}\in C(I_{n}) is mapped to δiClc(I)\delta_{i}\in C_{lc}({{{I}^{{\kern-1.04996pt}\wedge}}}) for 1in1\leq i\leq n, AiA_{i} is mapped to AiA_{i}. The description of the image of AA_{\infty} is by the definition of AXnAXA_{X_{n}}\to A_{{{{X}^{{\kern-0.74997pt}\wedge}}}}, which comes from IIn{{{I}^{{\kern-1.04996pt}\wedge}}}\to I_{n}. ∎

Everything is well-defined, except i0=Ii_{0}=\infty\in{{{I}^{{\kern-1.04996pt}\wedge}}}. In fact, the operators AiA_{i}, EjE_{j}, the values pi(j)p_{i}(j), qj(i)q_{j}(i), μJ(j)=mj\mu_{{{J}^{{\kern-0.74997pt}\wedge}}}(j)=m_{j}, μI(i)=ki/#X=(v1)vi\mu_{{{I}^{{\kern-0.74997pt}\wedge}}}(i)=k_{i}/\#X=(v-1)v^{-i} are all well-defined, except that AA_{\infty} and p(j)p_{\infty}(j) are not well-defined, since there is no corresponding element to AA_{\infty} in AXA_{{{X}^{{\kern-0.74997pt}\wedge}}}. Indeed, AA_{\infty} would correspond to the indicator function of I\infty\in{{{I}^{{\kern-1.04996pt}\wedge}}}, which is not continuous and hence is outside AXA_{{{X}^{{\kern-0.74997pt}\wedge}}}. This forces to choose something which replaces AA_{\infty}. One possible way is to replace it with E0E_{0} because

E0=iIAiE_{0}=\sum_{i\in{{{I}^{{\kern-0.74997pt}\wedge}}}}A_{i}

holds as a function (constant value 11) on I{{{I}^{{\kern-1.04996pt}\wedge}}}. then

{Aii>0}{E0}\{A_{i}\mid i\in{\mathbb{N}}_{>0}\}\cup\{E_{0}\}

is a linear basis of Clc(I)C_{lc}({{{I}^{{\kern-1.04996pt}\wedge}}}), although this is not an orthogonal basis. This is the method chosen in [2, (8.3), (8.9)] to replace an infinite sum with a finite sum (α0\alpha_{0} in (8.8) in their paper is nothing but E0E_{0}), which gives a precise definition of the adjacency algebra. In our interpretation in Proposition 4.32, the expression in Corollary 4.29 becomes

Ai=jJpi(j)EjA_{i}=\sum_{j\in{{{J}^{{\kern-0.74997pt}\wedge}}}}p_{i}(j)E_{j}

which is a finite sum (see Lemma 4.26) and holds for ii\neq\infty (since the equality holds in AXiA_{X_{i}}), except the problem that for i=i=\infty, the value p(j)=vnp_{\infty}(j)=v^{-n} converges to 0, while the summation is over infinitely many EjE_{j}. In Corollary 4.30 the expression

Ej=iIqj(i)AiE_{j}=\sum_{i\in{{{I}^{{\kern-0.74997pt}\wedge}}}}q_{j}(i)A_{i}

seems an infinite sum, but qj(i)=(v1)vj1q_{j}(i)=(v-1)v^{j-1} is constant for i>ji>j, and thus the right hand side lies in Clc(I)C_{lc}({{{I}^{{\kern-1.04996pt}\wedge}}}), or more precisely, the seemingly infinite sum is a finite sum by the usage of the symbol in Proposition 4.32, and the equality does hold (since it holds in AXjA_{X_{j}}.) This may be translated into a finite summation formula for j1j\geq 1:

Ej\displaystyle E_{j} =\displaystyle= ij(qj(i)(v1)vj1)Ai+(v1)vj1E0\displaystyle\sum_{i\leq j}(q_{j}(i)-(v-1)v^{j-1})A_{i}+(v-1)v^{j-1}E_{0}
=\displaystyle= (i<j(v1)vj1Ai)vjAj+(v1)vj1E0.\displaystyle(\sum_{i<j}-(v-1)v^{j-1}A_{i})-v^{j}A_{j}+(v-1)v^{j-1}E_{0}.

4.3. Ordered Hamming schemes

Martin and Stinson [16, Section 1.3] introduced ordered Hamming schemes as Delsarte’s extension of length ss [6, Section 2.5] of the kernel schemes.

Proposition 4.33.

(Extension of length ss)

Let R:X×XIR:X\times X\to I be a commutative association scheme. Let ss be a positive integer, and SsS_{s} a symmetric group of degree ss, acting on the direct product IsI^{s}. We obtain a quotient map IsIs/SsI^{s}\to I^{s}/S_{s}. Then, the composition

Xs×XsIsIs/SsX^{s}\times X^{s}\to I^{s}\to I^{s}/S_{s}

is a commutative association scheme. For i¯Is/Ss\bar{i}\in I^{s}/S_{s}, the corresponding Hadamard primitive idempotent is the sum of

Ai1AisA_{i_{1}}\otimes\cdots\otimes A_{i_{s}}

over (i1,,is)Is(i_{1},\ldots,i_{s})\in I^{s} which maps to i¯\bar{i}. The set of primitive idempotents are naturally identified with Js/SsJ^{s}/S_{s}, as follows. For j¯Js/Ss\bar{j}\in J^{s}/S_{s}, the corresponding primitive idempotent is the sum of

Ej1EjsE_{j_{1}}\otimes\cdots\otimes E_{j_{s}}

over (j1,,js)Js(j_{1},\ldots,j_{s})\in J^{s} which maps to j¯\bar{j}. The corresponding eigenspace is the direct sum of the tensors of the eigenspaces

C(X)j1C(X)jsC(X)_{j_{1}}\otimes\cdots\otimes C(X)_{j_{s}}

over (j1,,js)Js(j_{1},\ldots,j_{s})\in J^{s} which maps to j¯\bar{j}.

A proof is not given in the paper by Delsarte, but found in an unpublished (but reachable) paper by Godsil [9, 3.2 Corollary]. Because he did not mention the eigenspaces, which we need later, we shall recall a proof.

Proof.

Let

Xs×XsIsX^{s}\times X^{s}\to I^{s}

be the direct product association scheme. Its Bose-Mesner algebra AXs=AXsA_{X^{s}}=A_{X}^{\otimes s} is generated by the ss-fold kronecker products of AiA_{i}’s for iIi\in I, which satisfies the axiom of association scheme. The symmetric group SsS_{s} acts on AXsA_{X^{s}} by permutation. We take the fixed part AXsSsA_{X^{s}}^{S_{s}} of AXsA_{X^{s}}. Since SsS_{s} preserves the two products and the transpose, fix the i0i_{0} and the j0j_{0}, AXsSsA_{X^{s}}^{S_{s}} is a Bose-Mesner algebra. As a vector space, this is canonically isomorphic to C(Is)SsC(Is/Ss)C(I^{s})^{S_{s}}\cong C(I^{s}/S_{s}), induced by IsIs/SsI^{s}\to I^{s}/S_{s}. For an i¯Is/Ss\bar{i}\in I^{s}/S_{s}, its inverse image is the set of (i1,,is)(i_{1},\ldots,i_{s}) mapped to i¯\bar{i}, and the set of Ai1AisA_{i_{1}}\otimes\cdots\otimes A_{i_{s}} are mutually disjoint Hadamard idempotent, whose sum is an Hadamard idempotent Ai¯A_{\bar{i}}. This must be primitive in AXsSsA_{X^{s}}^{S_{s}}, since the Hadamard product satisfies Ai¯Ai¯=δi¯i¯A_{\bar{i}}\circ A_{\bar{i}^{\prime}}=\delta_{\bar{i}\ \bar{i}^{\prime}}, and their sum is the Hadamard unit JAXsJ_{A_{X^{s}}}. The same argument applies for the primitive idempotents Ej¯E_{\bar{j}}. The description of the corresponding eigenspace follows. ∎

This construction, when applied to the kernel schemes XnX_{n}, In={1,2,,n}{}I_{n}=\{1,2,\ldots,n\}\cup\{\infty\}, and Rn:Xn×XnInR_{n}:X_{n}\times X_{n}\to I_{n} given Definition 4.19), yields an ordered Hamming scheme.

Definition 4.34.

The ordered Hamming scheme, denoted by H(s,n,v)\overrightarrow{H}(s,n,v), is defined as the extension of length ss of the kernel scheme k(n,v)\overrightarrow{k(n,v)}.

The following is easy to check.

Proposition 4.35.

The projections Xn+1sXnsX_{n+1}^{s}\to X_{n}^{s} and In+1s/SsIns/SsI_{n+1}^{s}/S_{s}\to I_{n}^{s}/S_{s} coming from Proposition 4.20 give a projective system H(s,n+1,v)H(s,n,v)\overrightarrow{H}(s,n+1,v)\to\overrightarrow{H}(s,n,v) of commutative association schemes, namely, a profinite association scheme, which we call the pro-ordered Hamming scheme and denote by H(s,,v)\overrightarrow{H}(s,\infty,v). The I{{{I}^{{\kern-1.04996pt}\wedge}}} of this projective association schemes is (>0{})s/Ss({\mathbb{N}}_{>0}\cup\{\infty\})^{s}/S_{s}, and J{{{J}^{{\kern-1.04996pt}\wedge}}} is s/Ss{\mathbb{N}}^{s}/S_{s}.

We shall mention on this profinite scheme later in the last of Section 5.3.

5. Delsarte theory for profinite association schemes

Here we extend Delsarte theory introduced in [6]. The method here follows Kurihara-Okuda [13], which generalizes Delsarte theory to compact homogeneous spaces. In this section, let (Xλ,Rλ,Iλ)(X_{\lambda},R_{\lambda},I_{\lambda}) be a profinite association scheme, and X{{{X}^{{\kern-1.04996pt}\wedge}}}, I{{{I}^{{\kern-1.04996pt}\wedge}}}, J{{{J}^{{\kern-1.04996pt}\wedge}}}, AXA_{{{{X}^{{\kern-0.74997pt}\wedge}}}} be those defined in Section 3.1.

5.1. Multiset and Averaging functional

We consider a finite multi-subset YY of X{{{X}^{{\kern-1.04996pt}\wedge}}}, which means that YY is a set in which finite multiplicity of elements is allowed and taken into account. To make the notion rigorous, we consider a map from a finite set ZZ to a set XX

g:ZX.g:Z\to X.

Then the image Y:=g(Z)Y:=g(Z) in XX has a natural finite multiset structure, where the multiplicity of yYy\in Y is the cardinality of the fiber #g1(y)\#g^{-1}(y). For SXS\subset X, we use the notation

#(YS):=#(Zg1(S)).\#(Y\cap S):=\#(Z\cap g^{-1}(S)).

This merely means to count the number of elements in YSY\cap S with taking the multiplicity into account. Thus, we call YY a “multi-subset” of XX, and use the notation

a finite multiset YXY\subset X

by an abuse of language.

For any non-empty finite multiset YXY\subset{{{X}^{{\kern-1.04996pt}\wedge}}}, we would like to describe the notions of codes and designs. We begin with defining the averaging functional.

Definition 5.1.

Let YXY\subset{{{X}^{{\kern-1.04996pt}\wedge}}} be a finite multi-subset in the sense above. Define the averaging functional

avgY:C(X),f1#YxYf(x):=1#ZzZf(g(z)).\operatorname{avg}_{Y}:C({{{X}^{{\kern-1.04996pt}\wedge}}})\to{\mathbb{C}},\quad f\mapsto\frac{1}{\#Y}\sum_{x\in Y}f(x):=\frac{1}{\#Z}\sum_{z\in Z}f(g(z)).

By (3.1), Cc(J)AXClc(I)C_{c}({{{J}^{{\kern-1.04996pt}\wedge}}})\cong A_{{{X}^{{\kern-0.74997pt}\wedge}}}\cong C_{lc}({{{I}^{{\kern-1.04996pt}\wedge}}}) holds, and by

AXC(X×X)avgYavgY,A_{{{X}^{{\kern-0.74997pt}\wedge}}}\to C({{{X}^{{\kern-1.04996pt}\wedge}}}\times{{{X}^{{\kern-1.04996pt}\wedge}}})\stackrel{{\scriptstyle\operatorname{avg}_{Y}\otimes\operatorname{avg}_{Y}}}{{\to}}{\mathbb{C}},

avgY2:=avgYavgY\operatorname{avg}_{Y}^{2}:=\operatorname{avg}_{Y}\otimes\operatorname{avg}_{Y} defines a functional on AXA_{{{X}^{{\kern-0.74997pt}\wedge}}}, on Cc(J)C_{c}({{{J}^{{\kern-1.04996pt}\wedge}}}), and on Clc(I)C_{lc}({{{I}^{{\kern-1.04996pt}\wedge}}}). For example, for fClc(I)f\in C_{lc}({{{I}^{{\kern-1.04996pt}\wedge}}}), we have

avgY2(f)\displaystyle\operatorname{avg}_{Y}^{2}(f) =\displaystyle= 1#Y2x,yYfR(x,y)\displaystyle\frac{1}{\#Y^{2}}\sum_{x,y\in Y}f\circ{{{R}^{{\kern-1.04996pt}\wedge}}}(x,y)
=\displaystyle= 1#Y2iIx,yY,R(x,y)=if(i)\displaystyle\frac{1}{\#Y^{2}}\sum_{i\in{{{I}^{{\kern-0.74997pt}\wedge}}}}\sum_{x,y\in Y,{{{R}^{{\kern-0.74997pt}\wedge}}}(x,y)=i}f(i)
=\displaystyle= 1#Y2iI#((Y×Y)R1(i))f(i).\displaystyle\frac{1}{\#Y^{2}}\sum_{i\in{{{I}^{{\kern-0.74997pt}\wedge}}}}\#((Y\times Y)\cap{{{R}^{{\kern-1.04996pt}\wedge}}}^{-1}(i))f(i).

Note that these coefficients are the inner-distribution of YY (Delsarte [6, Section 3.1]) multiplied by 1#Y\frac{1}{\#Y}. Let us denote by I{\mathbb{C}}^{\oplus{{{I}^{{\kern-0.74997pt}\wedge}}}} the vector space whose basis is I{{{I}^{{\kern-1.04996pt}\wedge}}}, namely, the space of finite linear combinations of the elements of I{{{I}^{{\kern-1.04996pt}\wedge}}}. We have a mapping

IClc(I),{\mathbb{C}}^{\oplus{{{I}^{{\kern-0.74997pt}\wedge}}}}\to C_{lc}({{{I}^{{\kern-1.04996pt}\wedge}}})^{\vee}, (5.1)

where \vee denotes the dual (i.e. Hom(Clc(I),)\operatorname{Hom}(C_{lc}({{{I}^{{\kern-1.04996pt}\wedge}}}),{\mathbb{C}})), by the evaluation at ii: i(ff(i))i\mapsto(f\mapsto f(i)), which is injective since only a finite number of linear combinations appear in I{\mathbb{C}}^{\oplus{{{I}^{{\kern-0.74997pt}\wedge}}}}, and their support can be separated by clopen subsets. The above computation shows that avgY2Clc(I)\operatorname{avg}_{Y}^{2}\in C_{lc}({{{I}^{{\kern-1.04996pt}\wedge}}})^{\vee} lies in I{\mathbb{C}}^{\oplus{{{I}^{{\kern-0.74997pt}\wedge}}}}, namely,

avgY2=iI1#Y2(#((Y×Y)R1(i)))iI.\operatorname{avg}_{Y}^{2}=\sum_{i\in I}\frac{1}{\#Y^{2}}(\#((Y\times Y)\cap{{{R}^{{\kern-1.04996pt}\wedge}}}^{-1}(i)))\cdot i\in{\mathbb{C}}^{\oplus{{{I}^{{\kern-0.74997pt}\wedge}}}}. (5.2)

Note that every coefficient is non-negative. Next we compute avgY2\operatorname{avg}_{Y}^{2} on Cc(J)C_{c}({{{J}^{{\kern-1.04996pt}\wedge}}}). For jJj\in{{{J}^{{\kern-1.04996pt}\wedge}}}, we have the orthogonal projector Ej:C(X)C(X)jE_{j}:C({{{X}^{{\kern-1.04996pt}\wedge}}})\to C({{{X}^{{\kern-1.04996pt}\wedge}}})_{j}. For fC(X)jf\in C({{{X}^{{\kern-1.04996pt}\wedge}}})_{j},

f(x)=yXEj(x,y)f(y)𝑑μ(y)=(f,Ej(x,)¯)HS.f(x)=\int_{y\in{{{X}^{{\kern-0.74997pt}\wedge}}}}E_{j}(x,y)f(y)d\mu(y)=(f,\overline{E_{j}(x,-)})_{{\operatorname{HS}}}.

Thus, we define avgYjC(X)j\operatorname{avg}_{Y}^{j}\in C({{{X}^{{\kern-1.04996pt}\wedge}}})_{j} by

avgYj():=1#YxYEj(x,),\operatorname{avg}_{Y}^{j}(-):=\frac{1}{\#Y}\sum_{x\in Y}E_{j}(x,-), (5.3)

which represents the averaging functional in C(X)jC({{{X}^{{\kern-1.04996pt}\wedge}}})_{j}:

(f,avgYj¯)HS=avgY(f).(f,\overline{\operatorname{avg}_{Y}^{j}})_{{\operatorname{HS}}}=\operatorname{avg}_{Y}(f). (5.4)

Then

avgY2(Ej)\displaystyle\operatorname{avg}_{Y}^{2}(E_{j}) =\displaystyle= 1#Y2x,yYEj(x,y)\displaystyle\frac{1}{\#Y^{2}}\sum_{x,y\in Y}E_{j}(x,y) (5.5)
=\displaystyle= 1#YyYavgYj(y)\displaystyle\frac{1}{\#Y}\sum_{y\in Y}\operatorname{avg}_{Y}^{j}(y)
=\displaystyle= (avgYj,avgYj¯)=avgYjHS20.\displaystyle(\operatorname{avg}_{Y}^{j},\overline{\operatorname{avg}_{Y}^{j}})=||\operatorname{avg}_{Y}^{j}||_{{\operatorname{HS}}}^{2}\geq 0.

These positivities make the LP method by Delsarte possible [6, Thorem 3.3].

Definition 5.2.

Let us denote by QjClc(I)Q_{j}\in C_{lc}({{{I}^{{\kern-1.04996pt}\wedge}}}) the image of EjAXE_{j}\in A_{{{{X}^{{\kern-0.74997pt}\wedge}}}}. This may be considered as a description of the canonical isomorphism

Q:Cc(J)Clc(I),δjQj,Q:C_{c}({{{J}^{{\kern-1.04996pt}\wedge}}})\to C_{lc}({{{I}^{{\kern-1.04996pt}\wedge}}}),\quad\delta_{j}\mapsto Q_{j},

where δj\delta_{j} means the indicator function on J{{{J}^{{\kern-1.04996pt}\wedge}}} at jJj\in{{{J}^{{\kern-1.04996pt}\wedge}}}. (Note that EjAXE_{j}\in A_{{{{X}^{{\kern-0.74997pt}\wedge}}}} corresponds to δjCc(J)\delta_{j}\in C_{c}({{{J}^{{\kern-1.04996pt}\wedge}}}) in Theorem 3.13, see Corollary 3.11.) Then we have an injection

IClc(I)QCc(J),{\mathbb{C}}^{\oplus{{{I}^{{\kern-0.74997pt}\wedge}}}}\to C_{lc}({{{I}^{{\kern-1.04996pt}\wedge}}})^{\vee}\stackrel{{\scriptstyle Q^{\vee}}}{{\to}}C_{c}({{{J}^{{\kern-1.04996pt}\wedge}}})^{\vee},

obtained by

i(ff(i))(δjQj(i)).i\mapsto(f\mapsto f(i))\mapsto(\delta_{j}\mapsto Q_{j}(i)).
Definition 5.3.

The above morphism Q|I:ICc(J)Q^{\vee}|_{{{I}^{{\kern-0.74997pt}\wedge}}}:{\mathbb{C}}^{\oplus{{{I}^{{\kern-0.74997pt}\wedge}}}}\to C_{c}({{{J}^{{\kern-1.04996pt}\wedge}}})^{\vee} is called the Mac-Williams transform. In concrete, it maps

Q|I:iIaiijJiIaiQj(i)evj,Q^{\vee}|_{{{I}^{{\kern-0.74997pt}\wedge}}}:\sum_{i\in I}a_{i}\cdot i\mapsto\sum_{j\in{{{J}^{{\kern-0.74997pt}\wedge}}}}\sum_{i\in{{{I}^{{\kern-0.74997pt}\wedge}}}}a_{i}Q_{j}(i){\operatorname{ev}}_{j},

where evj{\operatorname{ev}}_{j} is the dual basis, i.e., evj(δj)=δjj{\operatorname{ev}}_{j}(\delta_{j^{\prime}})=\delta_{jj^{\prime}}, or equivalently,

evj:Cc(J),ff(j).{\operatorname{ev}}_{j}:C_{c}({{{J}^{{\kern-1.04996pt}\wedge}}})\to{\mathbb{C}},f\mapsto f(j).

Note that the left is a finite sum, but the right may be an infinite sum (which causes no problem, since the dual of a direct sum is the direct product).

The following is a formal consequence:

Q|I:avgY2IavgY2Cc(J).Q^{\vee}|_{{{I}^{{\kern-0.74997pt}\wedge}}}:\operatorname{avg}_{Y}^{2}\in{\mathbb{C}}^{\oplus{{{I}^{{\kern-0.74997pt}\wedge}}}}\mapsto\operatorname{avg}_{Y}^{2}\in C_{c}({{{J}^{{\kern-1.04996pt}\wedge}}})^{\vee}. (5.6)

5.2. Codes and designs

We define codes and designs. Let ICII_{C}\subset{{{I}^{{\kern-1.04996pt}\wedge}}} be a subset which does not contain i0i_{0} (CC for code). Let JDJJ_{D}\subset{{{J}^{{\kern-1.04996pt}\wedge}}} be a subset which does not contain j0j_{0} (DD for design). We define a convex cone (IC;JD)I(I_{C};J_{D})\subset{\mathbb{C}}^{\oplus{{{I}^{{\kern-0.74997pt}\wedge}}}} by

(IC;JD)\displaystyle(I_{C};J_{D}) :=\displaystyle:=
{\displaystyle\{ iIaiiai=0 for all but finite i ,\displaystyle\sum_{i\in I}a_{i}\cdot i\mid a_{i}=0\mbox{ for all but finite $i$ },
ai0 for all iI,\displaystyle a_{i}\geq 0\mbox{ for all }i\in{{{I}^{{\kern-1.04996pt}\wedge}}},
ai=0 for all iIC,\displaystyle a_{i}=0\mbox{ for all }i\in I_{C},
iIaiQj(i)0 for all jJ,\displaystyle\sum_{i\in{{{I}^{{\kern-0.74997pt}\wedge}}}}a_{i}Q_{j}(i)\geq 0\mbox{ for all }j\in{{{J}^{{\kern-1.04996pt}\wedge}}},
iIaiQj(i)=0 for all jJD}.\displaystyle\sum_{i\in{{{I}^{{\kern-0.74997pt}\wedge}}}}a_{i}Q_{j}(i)=0\mbox{ for all }j\in J_{D}\ \ \}.
Definition 5.4.

A non-empty finite multi-subset YXY\subset{{{X}^{{\kern-1.04996pt}\wedge}}} is called an ICI_{C}-free-code-JDJ_{D}-design if avgY2\operatorname{avg}_{Y}^{2} lies in the cone (IC;JD)(I_{C};J_{D}). Furthermore, an ICI_{C}-free-code-\emptyset-design [resp. an \emptyset-free-code-JDJ_{D}-design] is simply called an ICI_{C}-free-code [resp. a JDJ_{D}-design].

More explicitly, the ii-component of avgY2\operatorname{avg}_{Y}^{2} in I{\mathbb{C}}^{\oplus{{{I}^{{\kern-0.74997pt}\wedge}}}} is

ai(Y):=1#Y2#((Y×Y)R1(i)),a_{i}(Y):=\frac{1}{\#Y^{2}}\#((Y\times Y)\cap{{{R}^{{\kern-1.04996pt}\wedge}}}^{-1}(i)), (5.7)

which is non-negative (5.2) and required to be 0 for iICi\in I_{C} (namely, there is no pair (x,y)Y×Y(x,y)\in Y\times Y with relation R(x,y)IC{{{R}^{{\kern-1.04996pt}\wedge}}}(x,y)\in I_{C}), and its evj{\operatorname{ev}}_{j}-component in Cc(J)C_{c}({{{J}^{{\kern-1.04996pt}\wedge}}})^{\vee} is

bj(Y):=avgY2(Ej)=avgYjHS2,b_{j}(Y):=\operatorname{avg}_{Y}^{2}(E_{j})=||\operatorname{avg}_{Y}^{j}||_{{\operatorname{HS}}}^{2}, (5.8)

which is non-negative (5.5) and required to be 0 for jJDj\in J_{D} (namely, the jj-component of avgY\operatorname{avg}_{Y} is zero for jJDj\in J_{D}, or equivalently, for any fC(X)jf\in C({{{X}^{{\kern-1.04996pt}\wedge}}})_{j}, yYf(y)=0\sum_{y\in Y}f(y)=0 for jJDj\in J_{D}). Note that i0i_{0} is removed from ICI_{C}, since

1#Y2#((Y×Y)R1(i0))=1#Y\frac{1}{\#Y^{2}}\#((Y\times Y)\cap{{{R}^{{\kern-1.04996pt}\wedge}}}^{-1}(i_{0}))=\frac{1}{\#Y}

is positive, and j0j_{0} is removed since avgYj0HS=1||\operatorname{avg}_{Y}^{j_{0}}||_{{\operatorname{HS}}}=1 (being the operator norm of the averaging for constant functions). Since

#YavgY2(i0)=1 and iI#YavgY2(i)=#Y,\#Y\operatorname{avg}_{Y}^{2}(i_{0})=1\mbox{ and }\sum_{i\in{{{I}^{{\kern-0.74997pt}\wedge}}}}\#Y\operatorname{avg}_{Y}^{2}(i)=\#Y,

we may consider an LP problem: under the constraint that (ai)iI(a_{i})_{i\in{{{I}^{{\kern-0.74997pt}\wedge}}}} lies in the cone (IC;JD)(I_{C};J_{D}) and ai0=1a_{i_{0}}=1, maximize/minimize iIai\sum_{i\in{{{I}^{{\kern-0.74997pt}\wedge}}}}a_{i}, which gives an upper/lower bound on the cardinality of YY which is an ICI_{C}-free-code-JDJ_{D}-design. The following is a formal consequence of (5.6), stating the relation with classic theory [6, Section 3].

Theorem 5.5.

The Mac-Williams transformation Q|IQ^{\vee}|_{{{I}^{{\kern-0.74997pt}\wedge}}} maps (5.7) to (5.8):

iIai(Y)ijJbj(Y)evj.\sum_{i\in{{{I}^{{\kern-0.74997pt}\wedge}}}}a_{i}(Y)\cdot i\mapsto\sum_{j\in{{{J}^{{\kern-0.74997pt}\wedge}}}}b_{j}(Y){\operatorname{ev}}_{j}.

In concrete,

bj(Y)=iIai(Y)Qj(i).b_{j}(Y)=\sum_{i\in{{{I}^{{\kern-0.74997pt}\wedge}}}}a_{i}(Y)Q_{j}(i).
Remark 5.6.

The meaning of the values (5.7) is clear, and called the inner distribution [6, Section 3.1] (up to a factor of #Y\#Y), but there, the meaning of the values bj(Y)=iIai(Y)Qj(i)b_{j}(Y)=\sum_{i\in{{{I}^{{\kern-0.74997pt}\wedge}}}}a_{i}(Y)Q_{j}(i) is not clear, just the semi-positivity is proved. Now the value bj(Y)=avgYjHS2b_{j}(Y)=||\operatorname{avg}_{Y}^{j}||_{{\operatorname{HS}}}^{2} has a clear interpretation. On the space C(X)jC({{{X}^{{\kern-1.04996pt}\wedge}}})_{j}, the operator norm of avgYj\operatorname{avg}_{Y}^{j} is avgYjHS||\operatorname{avg}_{Y}^{j}||_{{\operatorname{HS}}}, since

supfC(X)j{0}|avgYj(f)|f=supfC(X)j{0}(f,avg¯Yj)ffavgYjf=avgYj,\sup_{f\in C({{{X}^{{\kern-0.74997pt}\wedge}}})_{j}\setminus\{0\}}\frac{|\operatorname{avg}_{Y}^{j}(f)|}{||f||}=\sup_{f\in C({{{X}^{{\kern-0.74997pt}\wedge}}})_{j}\setminus\{0\}}\frac{(f,\overline{\operatorname{avg}}_{Y}^{j})}{||f||}\leq\frac{||f||\cdot||\operatorname{avg}_{Y}^{j}||}{||f||}=||\operatorname{avg}_{Y}^{j}||,

and the equality holds for f=avgYjf=\operatorname{avg}_{Y}^{j}. This may be interpreted as the worst-case error in approximating the integration for C(X)jC(X)_{j} by avgY\operatorname{avg}_{Y} (where the true integration value is zero for jj0j\neq j_{0}, because of the orthogonality to the constant functions C(X)j0C(X)_{j_{0}}), see a comprehensive book by Dick-Pillichshammer [7, Definition 2.10] on quasi-Monte Carlo integration. Another remark: the injectivity (5.1) implies that YY and YY^{\prime} have the same inner distribution

#((Y×Y)R1(i))=#((Y×Y)R1(i)) for all iI\#((Y\times Y)\cap{{{R}^{{\kern-1.04996pt}\wedge}}}^{-1}(i))=\#((Y^{\prime}\times Y^{\prime})\cap{{{R}^{{\kern-1.04996pt}\wedge}}}^{-1}(i))\mbox{ for all $i\in{{{I}^{{\kern-1.04996pt}\wedge}}}$}

if and only if

avgYjHS=avgYjHS for all jJ.||\operatorname{avg}_{Y}^{j}||_{{\operatorname{HS}}}=||\operatorname{avg}_{Y^{\prime}}^{j}||_{{\operatorname{HS}}}\mbox{ for all $j\in{{{J}^{{\kern-1.04996pt}\wedge}}}$}.

5.3. (t,m,s)(t,m,s)-nets and (t,s)(t,s)-sequences

The (t,m,s)(t,m,s)-nets are point sets for quasi-Monte Carlo integration, well-studied, see [18]. For (t,m,s)(t,m,s)-nets, the LP bound by ordered Hamming schemes introduced by Martin-Stinson [16] yields strong lower bounds on the cardinality of the point set PP, see [16] and Bierbrauer [3] for example. We do not have such an application, but profinite association schemes may be used to formalize the notion of (t,s)(t,s)-sequence [18], as mentioned at the last of this section.

The purpose of this section is to reprove some results of Marin-Stinson, and see a relation to pro-ordered Hamming schemes. The main result Corollary 5.19 of this section is proved in Martin-Stinson[16, Theorem 3.4] by using the notion of dual schemes and weight enumerator polynomials. Here we shall give a self-contained proof avoiding the use of these tools. Our proof matches to the formalization in Section 5.2, relying on the average functional avgY\operatorname{avg}_{Y} introduced in Definition 5.1.

Recall the kernel scheme in Definition 4.19, where the set of alphabets is Zv:=/vZ_{v}:={\mathbb{Z}}/v, and Xn=ZvnX_{n}=Z_{v}^{n}, In={1,2,,n}{}I_{n}=\{1,2,\ldots,n\}\cup\{\infty\},

bot:XnIn{{\operatorname{bot}}}:X_{n}\to I_{n}

in Definition 4.24 give the kernel scheme as a translation scheme. We shall identify a point in XnX_{n} with a point in the real interval [0,1][0,1], by

(x1,,xn)0.x1x2xn[0,1),(x_{1},\ldots,x_{n})\mapsto 0.x_{1}x_{2}\cdots x_{n}\in[0,1), (5.9)

where the right expression means the vv-adic decimal expansion. Then, we have a map

Xns[0,1]s.X_{n}^{s}\to[0,1]^{s}.

We shall define (t,m,s)(t,m,s)-nets following Niederreiter [17, Definition 4.1]. An elementary interval of type (d1,,ds)(d_{1},\ldots,d_{s}) is a subset of [0,1]s[0,1]^{s} of the form

i=1s[aivdi,(ai+1)vdi),\prod_{i=1}^{s}[a_{i}v^{-d_{i}},(a_{i}+1)v^{-d_{i}}),

where aia_{i} and did_{i} are non-negative integers such that ai<vdia_{i}<v^{d_{i}}. Its volume is

isvdi=vi=1sdi.\prod_{i}^{s}v^{-d_{i}}=v^{-\sum_{i=1}^{s}d_{i}}.
Definition 5.7.

Let 0tm0\leq t\leq m be integers. A (t,m,s)(t,m,s)-net in base vv is a multi-subset P[0,1]sP\subset[0,1]^{s} consisting of vmv^{m} points, such that every elementary interval of type (d1,,ds)(d_{1},\ldots,d_{s}) with d1++ds=mtd_{1}+\cdots+d_{s}=m-t (hence volume vtmv^{t-m}) contains exactly vtv^{t} points of PP.

Our purpose is to state these conditions in terms of designs YXnsY\subset X_{n}^{s} for a finite multi-subset YY.

Definition 5.8.

A finite multi-subset YY in XX is said to be uniform on XX if #(Y{x})\#(Y\cap\{x\}) is constant for xXx\in X, that is, #g1(x)\#g^{-1}(x) is constant for xXx\in X if YY is defined by a map g:ZXg:Z\rightarrow X.

Definition 5.9.

For an ss-tuple of non-negative integers 𝐝=(d1,,ds){\mathbf{d}}=(d_{1},\ldots,d_{s}) with each component less than or equal to nn, define

X𝐝:=i=1sZvdi,X_{\mathbf{d}}:=\prod_{i=1}^{s}Z_{v}^{d_{i}},

and the projection

pr𝐝:XnsX𝐝{\operatorname{pr}}_{\mathbf{d}}:X_{n}^{s}\to X_{\mathbf{d}}

by taking the left most did_{i} components of the ii-th XnX_{n} in XnsX_{n}^{s}, for i=1,,si=1,\ldots,s.

Definition 5.10.

A non-empty multi-subset YXnsY\subset X_{n}^{s} is 𝐝{\mathbf{d}}-balanced, if the image of YY in X𝐝X_{\mathbf{d}} by the map XnsX𝐝X_{n}^{s}\to X_{\mathbf{d}} is uniform on X𝐝X_{\mathbf{d}}.

Proposition 5.11.

A multi-subset YXnsY\subset X_{n}^{s} of cardinality vmv^{m} gives a (t,m,s)(t,m,s)-net if and only if it is 𝐝{\mathbf{d}}-balanced for any non-negative tuples 𝐝=(d1,,ds){\mathbf{d}}=(d_{1},\ldots,d_{s}) with d1++ds=mtd_{1}+\cdots+d_{s}=m-t.

The proof is immediate when we consider the mapping (5.9), since then there is a natural one-to-one correspondence between X𝐝X_{\mathbf{d}} and the set of elementary intervals of type (d1,,ds)(d_{1},\ldots,d_{s}). We may state the conditions in terms of the dual group.

Lemma 5.12.

Let XX be a finite abelian group, and YY a finite multi-subset in XX. For fC(X)f\in C(X), define

I(f):=avgX(f)=1#XxXf(x).I(f):=\operatorname{avg}_{X}(f)=\frac{1}{\#X}\sum_{x\in X}f(x).

(This is the true integral of ff over the finite set XX). The following are equivalent.

  1. (1)

    I(f)=avgY(f)I(f)=\operatorname{avg}_{Y}(f) holds for any fC(X)f\in C(X).

  2. (2)

    The finite multi-subset YY is uniform on XX.

  3. (3)

    avgY(ξ)=0\operatorname{avg}_{Y}(\xi)=0 holds for any non-trivial ξXˇ\xi\in{{\check{X}}}.

Proof.

Suppose that the first condition holds. Let χx\chi_{x} be the indicator function of xXx\in X. The condition I(χx)=avgY(χx)I(\chi_{x})=\operatorname{avg}_{Y}(\chi_{x}) implies 1#X=#(Y{x})#Y\frac{1}{\#X}=\frac{\#(Y\cap\{x\})}{\#Y}, hence YY is uniform on XX. The converse is obvious.

Suppose again the first condition. For any non-trivial ξ\xi, I(ξ)=0I(\xi)=0. This implies that avgY(ξ)=0\operatorname{avg}_{Y}(\xi)=0. For the converse, we assume the third condition. Note that Xˇ{{\check{X}}} is a base of C(X)C(X), so it suffices to show that I(ξ)=avgY(ξ)I(\xi^{\prime})=\operatorname{avg}_{Y}(\xi^{\prime}) holds for any ξXˇ\xi^{\prime}\in{{\check{X}}}. This holds for non-trivial ξ\xi, and for the trivial ξ=1\xi=1, I(1)=1=avgY(1)I(1)=1=\operatorname{avg}_{Y}(1). ∎

Lemma 5.13.

The dual Xˇ𝐝{{\check{X}}}_{\mathbf{d}} of X𝐝X_{\mathbf{d}} can be identified with the set of characters

{ξ:=(ξ1,,ξs)Xnˇstop(ξi)di for i=1,,s}\{\xi:=(\xi_{1},\ldots,\xi_{s})\in\check{X_{n}}^{s}\mid{{\operatorname{top}}}(\xi_{i})\leq d_{i}\mbox{ for }i=1,\ldots,s\}

(see Definition 4.25 for the notation top{{\operatorname{top}}}).

Corollary 5.14.

For a multi-subset YXnsY\subset X_{n}^{s}, the composition YXnsX𝐝Y\to X_{n}^{s}\to X_{\mathbf{d}} is uniform if and only if

avgY(ξ)=0\operatorname{avg}_{Y}(\xi)=0

holds for any non-trivial ξXˇ𝐝\xi\in{{\check{X}}}_{\mathbf{d}}.

Corollary 5.15.

A multi-subset YXnsY\subset X_{n}^{s} of cardinality vmv^{m} is a (t,m,s)(t,m,s)-net if and only if

avgY(ξ)=0\operatorname{avg}_{Y}(\xi)=0

holds for any ξXˇ𝐝{1}\xi\in{{\check{X}}}_{\mathbf{d}}\setminus\{1\} and any non-negative 𝐝=(d1,,ds){\mathbf{d}}=(d_{1},\ldots,d_{s}) satisfying i=1sdi=mt\sum_{i=1}^{s}d_{i}=m-t.

Niederreiter[17] and Rosenbloom-Tsfasman[23] defined the notion of NRT-weight, see also Niederreiter and Pirsic[20]:

Definition 5.16.

Define the NRT-weight wt\operatorname{wt} by the composition

wt:XˇnstopsJnssum,\operatorname{wt}:{{\check{X}}}_{n}^{s}\stackrel{{\scriptstyle{{\operatorname{top}}}^{s}}}{{\to}}J_{n}^{s}\stackrel{{\scriptstyle{{\operatorname{sum}}}}}{{\to}}{\mathbb{N}},

where sum{{\operatorname{sum}}} is a function taking the sum of the ss coordinates and top{{\operatorname{top}}} is defined in Definition 4.25.

Theorem 5.17.

Let YXnsY\subset X_{n}^{s} be a multi-subset of cardinality vmv^{m}. Then it is a (t,m,s)(t,m,s)-net if and only if for any ξXˇns\xi\in{{\check{X}}}_{n}^{s} with wt(ξ)mt\operatorname{wt}(\xi)\leq m-t and ξ1\xi\neq 1, avgY(ξ)=0\operatorname{avg}_{Y}(\xi)=0.

This follows from Corollary 5.15. Following Martin-Stinson[16], let us define

shape:JnsJns/Ss.{{\operatorname{shape}}}:J_{n}^{s}\to J_{n}^{s}/S_{s}.

In Proposition 4.33 and Lemma 4.26, we observed that for j¯Jns/Ss\bar{j}\in J_{n}^{s}/S_{s}, the corresponding eigenspace C(Xns)j¯C(X_{n}^{s})_{\bar{j}} is spanned by the characters in the inverse image of j¯\bar{j} by

XˇnstopsJnsJns/Ss.{{\check{X}}}_{n}^{s}\stackrel{{\scriptstyle{{\operatorname{top}}}^{s}}}{{\to}}J_{n}^{s}\to J_{n}^{s}/S_{s}.

It is easy to see that wt\operatorname{wt} factors as

XˇnstopsJnsshapeJns/Ss0pt,{{\check{X}}}_{n}^{s}\stackrel{{\scriptstyle{{\operatorname{top}}}^{s}}}{{\to}}J_{n}^{s}\stackrel{{\scriptstyle{{\operatorname{shape}}}}}{{\to}}J_{n}^{s}/S_{s}\stackrel{{\scriptstyle 0pt}}{{\to}}{\mathbb{N}},

where the definition of 0pt0pt is clear from the context. (We remark that 0pt0pt is denoted by “height” in Martin-Stinson[16, P.337], but we changed the notation to avoid the collision with a standard use of “height” in the context of the Young diagrams.) Thus,

{ξXˇnswt(ξ)mt}\displaystyle\{\xi\in{{\check{X}}}_{n}^{s}\mid\operatorname{wt}(\xi)\leq m-t\} =\displaystyle= {ξXˇns0pt(shape(tops(ξ)))mt}.\displaystyle\{\xi\in{{\check{X}}}_{n}^{s}\mid 0pt({{\operatorname{shape}}}({{\operatorname{top}}}^{s}(\xi)))\leq m-t\}.

The condition for YXnsY\subset X_{n}^{s} to be a (t,m,s)(t,m,s)-net is that avgY(ξ)\operatorname{avg}_{Y}(\xi) vanishes for any nontrivial ξ\xi in the left hand side (Theorem 5.17), which is equivalent to the vanishing in the right hand side, namely, avgY(ξ)=0\operatorname{avg}_{Y}(\xi)=0 for each of ξC(Xns)j¯\xi\in C(X_{n}^{s})_{\bar{j}} for j¯Jns/Ss{j0}\bar{j}\in J_{n}^{s}/S_{s}\setminus\{j_{0}\}, 0pt(j¯)mt0pt(\bar{j})\leq m-t. This is equivalent to avgYj¯=0\operatorname{avg}_{Y}^{\bar{j}}=0 (the left hand side is the representation of avgY\operatorname{avg}_{Y} in the subspace C(Xns)j¯C(X_{n}^{s})_{\bar{j}} defined in (5.3) with property (5.4)) for these j¯\bar{j}.

Now we proved [16, Theorem 3.4]:

Theorem 5.18.

Let YXnsY\subset X_{n}^{s} be a multi-subset of cardinality vmv^{m}. Then, YY is a (t,m,s)(t,m,s)-net if and only if avgYj¯=0\operatorname{avg}_{Y}^{\bar{j}}=0 for any j¯j0\bar{j}\neq j_{0}, 0pt(j¯)mt0pt(\bar{j})\leq m-t.

In terms of the designs (see Section 5.2), we may state

Corollary 5.19.

Define

JD:={j¯Jns/Ssj¯j0, 0pt(j¯)mt}.J_{D}:=\{\bar{j}\in J_{n}^{s}/S_{s}\mid\bar{j}\neq j_{0},\ 0pt(\bar{j})\leq m-t\}.

Then, YXnsY\subset X_{n}^{s} is a JDJ_{D}-design if and only if it is a (t,m,s)(t,m,s)-net in XnsX_{n}^{s}.

By this, Delsarte’s LP method works, see [16]. The next is an example where we may use a profinite association scheme to make a concept of (t,s)(t,s)-sequence inside the design theory, (see [18, Definition 4.2] and [19, Remark 2.2]).

Let us consider the pro-ordered Hamming scheme H(s,,v)=(X,R)\overrightarrow{H}(s,\infty,v)=({{{X}^{{\kern-1.04996pt}\wedge}}},{{{R}^{{\kern-1.04996pt}\wedge}}}) defined in Proposition 4.35. Note that X=(limXn)s=lim(Xns)=(Zv>0)s{{{X}^{{\kern-1.04996pt}\wedge}}}=(\varprojlim X_{n})^{s}=\varprojlim(X_{n}^{s})=(Z_{v}^{{\mathbb{N}}_{>0}})^{s}. For each nn, we denote by πn\pi_{n} the surjection from X{{{X}^{{\kern-1.04996pt}\wedge}}} onto XnsX_{n}^{s}. That is,

πn(x1,,xs):=([x1]n,,[xs]n)\pi_{n}(x^{1},\dots,x^{s}):=([x^{1}]_{n},\dots,[x^{s}]_{n}) (5.10)

for each (x1,,xs)X=(Zv>0)s(x^{1},\dots,x^{s})\in{{{X}^{{\kern-1.04996pt}\wedge}}}=(Z_{v}^{{\mathbb{N}}_{>0}})^{s}, where we put [(x1,)]n:=(x1,,xn)[(x_{1},\dots)]_{n}:=(x_{1},\dots,x_{n}) for each (x1,)Zv>0(x_{1},\dots)\in Z_{v}^{{\mathbb{N}}_{>0}}.

Definition 5.20.

A sequence of points p0,p1,Xp_{0},p_{1},\ldots\in{{{X}^{{\kern-1.04996pt}\wedge}}} is a (t,s)(t,s)-sequence in base vv, if for any integers kk and m>tm>t, the point set Yk,mXms[0,1)sY_{k,m}\subset X_{m}^{s}\subset[0,1)^{s} consisting of the πm(pj)\pi_{m}(p_{j}) with kvmj<(k+1)vmkv^{m}\leq j<(k+1)v^{m} (considered as a multi-set) is a (t,m,s)(t,m,s)-net in base vv.

For a comparison with a standard Definition 5.22, see Remark 5.23. We may state the condition of (t,s)(t,s)-sequences in terms of the designs in the pro-ordered Hamming scheme H(s,,v)\overrightarrow{H}(s,\infty,v). Recall that J=s/Ss{{{J}^{{\kern-1.04996pt}\wedge}}}={\mathbb{N}}^{s}/S_{s}, and we may define 0pt0pt so that the composition

sJ=(s/Ss)0pt{\mathbb{N}}^{s}\to{{{J}^{{\kern-1.04996pt}\wedge}}}=({\mathbb{N}}^{s}/S_{s})\stackrel{{\scriptstyle 0pt}}{{\to}}{\mathbb{N}}

maps (j1,,js)i=1sji(j_{1},\ldots,j_{s})\mapsto\sum_{i=1}^{s}j_{i}.

Theorem 5.21.

In the pro-ordered Hamming scheme, we define

JD:={h(s/Ss)=J0pt(h),hj0}.J_{D_{\ell}}:=\{h\in({\mathbb{N}}^{s}/S_{s})={{{J}^{{\kern-1.04996pt}\wedge}}}\mid 0pt(h)\leq\ell,h\neq j_{0}\}.

A sequence of points p0,p1,Xp_{0},p_{1},\ldots\in{{{X}^{{\kern-1.04996pt}\wedge}}} is a (t,s)(t,s)-sequence in base vv, if and only if for any integers kk and m>tm>t, the point set Yk,mY_{k,m} consisting of the πm(pj)\pi_{m}(p_{j}) with kvmj<(k+1)vmkv^{m}\leq j<(k+1)v^{m} is a JDmtJ_{D_{m-t}}-design.

In the original definition, the sequence is taken in an infinite set [0,1]s[0,1]^{s}, where infinite precision is necessary. Because of this infiniteness, a fixed finite scheme would not be able to describe the notion of (t,s)(t,s)-sequence. Our profinite association scheme is a tool to deal with the infiniteness.

We close our paper with a discussion on the definitions of (t,s)(t,s)-sequences. The following is a definition by Niederreiter-Özbudak[19, Definition 2.2] (to be precise, its special case), which is a slight variant of the original Niederreiter’s one: [18, Definition 4.2].

Definition 5.22.

Let p0,p1,[0,1]sp_{0},p_{1},\ldots\in[0,1]^{s} be a sequence of points, with prescribed vv-adic expansions. (In other words, each pjp_{j} is considered as an element in X{{{X}^{{\kern-1.04996pt}\wedge}}}.) It is a (t,s)(t,s)-sequence in base vv, if for any integers kk and m>tm>t, the point set Pk,mP_{k,m} consisting of the πm(pj)\pi_{m}(p_{j}) with kvmj<(k+1)vmkv^{m}\leq j<(k+1)v^{m} is a (t,m,s)(t,m,s)-net in base vv.

Remark 5.23.

It is easy to see that Definitions 5.20 and 5.22 are the same. In Definition 5.22, we use πm\pi_{m} in (5.10) for pjp_{j}, where pjp_{j} is an element of X{{{X}^{{\kern-1.04996pt}\wedge}}}. In fact, πm\pi_{m} is not well-defined for [0,1]s[0,1]^{s}, because of the non-injectivity of

(Zv>0)[0,1],(Z_{v}^{{\mathbb{N}}_{>0}})\to[0,1],

namely, for base v=2v=2, for example, 0.0111=0.10000.0111\cdots=0.1000\cdots. One way to avoid this subtle problem may be to require the vv-adic expansion of a real number to be of the latter type, namely, to avoid expansions consisting of all (v1)(v-1)’s after some digit. However, in important constructions such as in Niederreiter-Xing[21], points with the former type of expansions may appear. Thus, to define the notion of (t,s)(t,s)-sequences as a sequence in [0,1]s[0,1]^{s} has a subtle problem. Definition 5.20 would be a natural definition of a (t,s)(t,s)-sequence. In other words, the notion would be better defined in terms of X{{{X}^{{\kern-1.04996pt}\wedge}}} rather than [0,1]s[0,1]^{s}. This is essentially stated in [19, Remark 2.2]. Of course, a large part of researchers would prefer to work in [0,1]s[0,1]^{s}.

References

  • [1] E. Bannai and T. Ito. Algebraic Combinatorics I: Association Schemes. Benjamin / Cummings, Calfornia, 1984.
  • [2] A. Barg and M. Skriganov. Association schemes on general measure spaces and zero-dimensional abelian groups. Adv. Math., 281:142–247, 2015.
  • [3] J. Bierbrauer. A direct approach to linear programming bounds for codes and tms-nets. Designs, Codes and Cryptography, 42(2):127–143, 2007.
  • [4] A. W. Brouwer, A. M. Cohen, and A. Neumaier. Distance-regular graphs, volume MATHE3, 18 of Ergebnisse der Mathematik und ihrer Grenzgebiete. 3. Folge / A Series of Modern Surveys in Mathematics. Springer-Verlag, New York, Heidelberg, Berlin, 1989.
  • [5] J. R. Choksi. Inverse limits of measure spaces. London Mathematical Society, s3-8:321–342, 1958.
  • [6] P. Delsarte. An algebraic approach to the association schemes of coding theory. Philips Res. Rep. Suppl., 10:i–vi and 1–97, 1973.
  • [7] J. Dick and F. Pillichshammer. Digital Nets and Sequences: Discrepancy Theory and Quasi-Monte Carlo Integration. Cambridge University Press, Cambridge, 2010.
  • [8] C. French. Functors from association schemes. J. Combin. Theory Ser. A, 120:1141 – 1165, 2013.
  • [9] C. Godsil. Generalized hamming schemes. arXiv:1011.1044.
  • [10] P.R. Halmos. Measure Theory, volume 18 of Graduate Texts in Mathematics. Springer-Verlag, New York, Heidelberg, Berlin, 1970.
  • [11] A. Hanaki. A category of association schemes. J. Combin. Theory Ser. A, 117:1207 – 1217, 2010.
  • [12] H. Kajiura, M. Matsumoto, and T. Okuda. Non-existence and construction of pre-difference sets, and equi-distributed subsets in association schemes. Graphs and Combinatorics, 37:1531–1544, 2021.
  • [13] H. Kurihara and T. Okuda. A MacWilliams type theorem and Delsarte’s theory for codes and designs on compact homogeneous spaces. in preparation.
  • [14] W.C. Lang. Orthogonal wavelets on the cantor dyadic group. SIAM J. Math. Anal., 27:305–312, 1996.
  • [15] S. MacLane. Categories for the Working Mathematician, volume 5 of Graduate Texts in Mathematics. Springer-Verlag, New York, Heidelberg, Berlin, 2nd edition, 1998.
  • [16] W.J. Martin and D.R. Stinson. Association schemes for ordered orthogonal arrays and (t, m, s)-nets. Canadian Journal of Mathematics, 51:326–346, 1999.
  • [17] H. Niederreiter. Low-discrepancy point sets. Monatsh. Math., 102:155–167, 1986.
  • [18] H. Niederreiter. Random Number Generation and Quasi-Monte Carlo Methods. CBMS-NSF, Philadelphia, Pennsylvania, 1992.
  • [19] H. Niederreiter and F. Özbudak. Low-discrepancy sequences using duality and global function fields. Acta Arith., 130:79–97, 2007.
  • [20] H. Niederreiter and G. Pirsic. Duality for digital nets and its applications. Acta Arith., 97:173–182, 2001.
  • [21] H. Niederreiter and C. P. Xing. Low-discrepancy sequences and global function fields with many rational places. Finite Fields and Their Applications, 2:241–273, 1996.
  • [22] L. Ribes and P. Zalesskii. Profinite Groups. Springer-Verlag, Berlin Heidelberg, 2000.
  • [23] M. Yu. Rosenbloom and M. A. Tsfasman. Codes for the m-metric. Problems of Information Transmission, 33:55–63, 1997.
  • [24] M. Takeuchi. Modern spherical functions, volume 135 of Translations of Mathematical Monographs. American Mathematical Society, Providence, 1994.
  • [25] P. H. Zieschang. An Algebraic Approach to Association Schemes, volume 1628 of Lecture Notes in Mathematics. Springer-Verlag, Berlin Heidelberg, 1996.