ISSN 1855-3966 (printed edn.), ISSN 1855-3974 (electronic edn.) ARS MATHEMATICA CONTEMPORANEA 22 (2022) #P2.01 / 163–205 https://doi.org/10.26493/1855-3974.2352.eaf (Also available at http://amc-journal.eu) Tight relative t-designs on two shells in hypercubes, and Hahn and Hermite polynomials* Eiichi Bannai Faculty of Mathematics, Kyushu University (emeritus), Japan, and National Center for Theoretical Sciences, National Taiwan University, Taiwan Etsuko Bannai National Center for Theoretical Sciences, National Taiwan University, Taiwan Hajime Tanaka † Research Center for Pure and Applied Mathematics, Graduate School of Information Sciences, Tohoku University, Sendai 980-8579, Japan Yan Zhu ‡ College of Science, University of Shanghai for Science and Technology, Shanghai 200093, China Received 3 June 2020, accepted 24 May 2021, published online 14 April 2022 Abstract Relative t-designs in the n-dimensional hypercubeQn are equivalent to weighted regu- lar t-wise balanced designs, which generalize combinatorial t-(n, k, λ) designs by allowing multiple block sizes as well as weights. Partly motivated by the recent study on tight Eu- clidean t-designs on two concentric spheres, in this paper we discuss tight relative t-designs inQn supported on two shells. We show under a mild condition that such a relative t-design induces the structure of a coherent configuration with two fibers. Moreover, from this struc- ture we deduce that a polynomial from the family of the Hahn hypergeometric orthogonal polynomials must have only integral simple zeros. The Terwilliger algebra is the main tool to establish these results. By explicitly evaluating the behavior of the zeros of the Hahn polynomials when they degenerate to the Hermite polynomials under an appropriate limit *This work was also partially supported by the Research Institute for Mathematical Sciences at Kyoto Univer- sity. †Corresponding author. Hajime Tanaka was supported by JSPS KAKENHI Grant Numbers JP25400034 and JP17K05156. ‡Yan Zhu was supported by NSFC Grant No. 11801353. cb This work is licensed under https://creativecommons.org/licenses/by/4.0/ 164 Ars Math. Contemp. 22 (2022) #P2.01 / 163–205 process, we prove a theorem which gives a partial evidence that the non-trivial tight relative t-designs in Qn supported on two shells are rare for large t. Keywords: Relative t-design, association scheme, coherent configuration, Terwilliger algebra, Hahn polynomial, Hermite polynomial. Math. Subj. Class. (2020): 05B30, 05E30, 33C45 1 Introduction This paper is a contribution to the study of relative t-designs in Q-polynomial associa- tion schemes. In the Delsarte theory [16], the concept of t-designs is introduced for arbi- trary Q-polynomial association schemes. For the Johnson scheme J(n, k), the t-designs in the sense of Delsarte are shown to be the same thing as the combinatorial t-(n, k, λ) designs. There are similar interpretations of t-designs in some other important families of Q-polynomial association schemes [16, 17, 19, 34, 41]. The concept of relative t-designs is also due to Delsarte [18], and is a relaxation of that of t-designs. Relative t-designs can again be interpreted in several cases, including J(n, k). For the n-dimensional hypercube Qn (or the binary Hamming scheme H(n, 2)) which will be our central focus in this paper, these are equivalent to the weighted regular t-wise balanced designs, which generalize the combinatorial t-(n, k, λ) designs by allowing multiple block sizes as well as weights. The Delsarte theory has a counterpart for the unit sphere Sn−1 in Rn, established by Delsarte, Goethals, and Seidel [20]. The t-designs in Sn−1 are commonly called the spherical t-designs, and are essentially the equally-weighted cubature formulas of degree t for the spherical integration, a concept studied extensively in numerical analysis. Spher- ical t-designs were later generalized to Euclidean t-designs by Neumaier and Seidel [35] (cf. [21]). Euclidean t-designs are in general supported on multiple concentric spheres in Rn, and it follows that we may think of them as the natural counterpart of relative t-designs in Rn. This point of view was discussed in detail by Bannai and Bannai [3]. See also [7, 8]. The success and the depth of the theory of Euclidean t-designs (cf. [38]) has been one driv- ing force for the recent research activity on relative t-designs in Q-polynomial association schemes; see, e.g., [3, 5, 6, 7, 8, 9, 11, 32, 51, 53, 54]. A relative t-design in a Q-polynomial association scheme (X,R) is often defined as a certain weighted subset of the vertex set X , i.e., a pair (Y, ω) of a subset Y of X and a function ω : Y → (0,∞). We are given in advance a ‘base vertex’ x ∈ X , and (Y, ω) gives a ‘degree-t approximation’ of the shells (or spheres or subconstituents) with respect to x on which Y is supported. See Sections 2 and 3 for formal definitions. Bannai and Bannai [3] proved a Fisher-type lower bound on |Y |, and we call (Y, ω) tight if it attains this bound. We may remark that t must be even in this case. In this paper, we continue the study (cf. [5, 9, 32, 51, 53]) of tight relative t-designs in the hypercubesQn, which are one of the most important families of Q-polynomial association schemes. The Delsarte theory directly applies to the tight relative t-designs in Qn supported on one shell, say, the kth shell, as these are equivalent to the tight combinatorial t-(n, k, λ) designs. (We note that the kth shell induces J(n, k).) Our aim is to extend this structure theory to those supported on two shells. We may view the results of this paper roughly as counterparts to (part of) the E-mail addresses: bannai@math.kyushu-u.ac.jp (Eiichi Bannai), et-ban@rc4.so-net.ne.jp (Etsuko Bannai), htanaka@tohoku.ac.jp (Hajime Tanaka), zhuyan@usst.edu.cn (Yan Zhu) E. Bannai et al.: Tight relative t-designs on two shells in hypercubes, . . . 165 results by Bannai and Bannai [2, 4] on tight Euclidean t-designs on two concentric spheres. Let t = 2e be even. In Theorem 5.3, which is our first main result, we show under a mild condition that a tight relative 2e-design in Qn supported on two shells induces the structure of a coherent configuration with two fibers. Moreover, from this structure we de- duce that a certain polynomial of degree e, known as a Hahn polynomial, must have only integral simple zeros. We note that the case e = 1 was handled previously by Bannai, Bannai, and Bannai [5]. The Hahn polynomials are a family of hypergeometric orthogonal polynomials in the Askey scheme [31, Section 1.5], and that their zeros are integral pro- vides quite a strong necessary condition on the existence of such relative 2e-designs. The corresponding necessary condition for the tight combinatorial 2e-(n, k, λ) designs from the Delsarte theory was used successfully by Bannai [1]; that is to say, he showed that, for each given integer e ⩾ 5, there exist only finitely many non-trivial tight 2e-(n, k, λ) de- signs, where n and k (and thus λ) vary. See also [22, 36, 52]. We extend Bannai’s method to prove our second main result, Theorem 7.1, which presents a version of his theorem for our case. The sections other than Sections 5 and 7 are organized as follows. We collect the necessary background material in Sections 2 and 3. Section 3 also includes a few general results on relative t-designs in Q-polynomial association schemes. As in [6, 44], our main tool in the analysis of relative t-designs is the Terwilliger algebra [46, 47, 48], which is a non-commutative semisimple C-algebra containing the adjacency algebra. Section 4 is devoted to detailed descriptions of the Terwilliger algebra ofQn. It is well known (cf. [30, 31]) that the Hahn polynomials (3F2) degenerate to the Hermite polynomials (2F0) by an appropriate limit process, and a key in Bannai’s method above was to evaluate precisely the behavior of the zeros of the Hahn polynomials in this process. In Section 6, we revisit this part of the method in a form suited to our purpose. Our account will also be simpler than that in [1]. In Appendix, we provide a proof of a number-theoretic result (Proposition 7.2) which is a variation of a result of Schur [40, Satz I]. 2 Coherent configurations and association schemes We begin by recalling the concept of coherent configurations. Definition 2.1. The pair (X,R) of a finite set X and a set R of non-empty subsets of X2 is called a coherent configuration on X if it satisfies the following (C1) – (C4): (C1) R is a partition of X2. (C2) There is a subset R0 of R such that⊔ R∈R0 R = {(x, x) : x ∈ X}. (C3) R is invariant under the transposition τ : (x, y) 7→ (y, x) ((x, y) ∈ X2), i.e., Rτ ∈ R for all R ∈ R. (C4) For all R,S, T ∈ R and (x, y) ∈ T , the number pTR,S := ∣∣{z ∈ X : (x, z) ∈ R, (z, y) ∈ S}∣∣ is independent of the choice of (x, y) ∈ T . 166 Ars Math. Contemp. 22 (2022) #P2.01 / 163–205 Moreover, a coherent configuration (X,R) on X is called homogeneous if |R0| = 1, and an association scheme if Rτ = R for all R ∈ R. Remark 2.2. Suppose that a finite group G acts on X , and let R be the set of the orbitals of G, that is to say, the orbits of G in its natural action on X2. Then (X,R) is a coherent configuration. Moreover, (X,R) is homogeneous (resp. an association scheme) if and only if the action of G on X is transitive (resp. generously transitive, i.e., for any x, y ∈ X we have (xg, yg) = (y, x) for some g ∈ G). Let (X,R) be a coherent configuration as above. For every R ∈ R0, let ΦR be the subset of X such that R = {(x, x) : x ∈ ΦR}. Then we have⊔ R∈R0 ΦR = X. We call the ΦR (R ∈ R0) the fibers of (X,R). By setting in (C4) either R ∈ R0 and S = T , or S ∈ R0 and R = T , it follows that for every T ∈ R, we have T ⊂ ΦR × ΦS for some R,S ∈ R0. In particular, (X,R) is homogeneous whenever it is an association scheme. Let γR,S = ∣∣{T ∈ R : T ⊂ ΦR × ΦS}∣∣ (R,S ∈ R0). The matrix [γR,S ]R,S∈R0 , which is symmetric by (C3), is called the type of (X,R). Let MX(C) be the C-algebra of all complex matrices with rows and columns indexed by X , and let V = CX be the C-vector space of complex column vectors with coordinates indexed by X . We endow V with the Hermitian inner product ⟨u, v⟩ = v†u (u, v ∈ V ), where † denotes adjoint. For every R ∈ R, let AR ∈ MX(C) be the adjacency matrix of the graph (X,R) (directed, in general), i.e., (AR)x,y = { 1 if (x, y) ∈ R, 0 otherwise, (x, y ∈ X). Then (C1) – (C4) above are rephrased as follows: (A1) ∑ R∈R AR = J (the all-ones matrix). (A2) ∑ R∈R0 AR = I (the identity matrix). (A3) (AR)† ∈ {AS : S ∈ R} (R ∈ R). (A4) ARAS = ∑ T∈R pTR,SAT (R,S ∈ R). E. Bannai et al.: Tight relative t-designs on two shells in hypercubes, . . . 167 Let A = span{AR : R ∈ R}. Then from (A2) and (A4) it follows that A is a subalgebra ofMX(C), called the adjacency algebra of (X,R). We note that A is semisimple as it is closed under † by virtue of (A3). By (A1), A is also closed under entrywise (or Hadamard or Schur) multiplication, which we denote by ◦. The AR are the (central) primitive idempotents of A with respect to ◦, i.e., AR ◦AS = δR,SAR, ∑ R∈R AR = J. Remark 2.3. If (X,R) arises from a group action as in Remark 2.2, then A coincides with the centralizer algebra (or Hecke algebra or commutant) for the corresponding permutation representation g 7→ Pg (g ∈ G) on V , i.e., A = {B ∈MX(C) : BPg = PgB (g ∈ G)}. A subalgebra of MX(C) is called a coherent algebra if it contains J , and is closed under ◦ and †. We note that the coherent algebras are precisely the adjacency algebras of coherent configurations. It is clear that the intersection of coherent algebras in MX(C) is again a coherent algebra. In particular, for any subset S of MX(C), we can speak of the smallest coherent algebra containing S, which we call the coherent closure of S. From now on, we assume that (X,R) is an association scheme. As is the case for many examples of association schemes, we write R = {R0, R1, . . . , Rn}, where R0 = {R0}, and say that (X,R) has n classes. We will then abbreviate pki,j = p Rk Ri,Rj , Ai = ARi , and so on. The adjacency algebra A is commutative in this case, and hence it has a basis E0, E1, . . . , En consisting of the (central) primitive idempotents, i.e., EiEj = δi,jEi, n∑ i=0 Ei = I. Put differently, E0V,E1V, . . . , EnV are the maximal common eigenspaces (or homoge- neous components or isotypic components) of A, and the Ei are the corresponding orthog- onal projections. Since the Ai are real symmetric matrices, so are the Ei. Note that the matrix |X|−1J ∈ A is an idempotent with rank one, and thus primitive. We will always set E0 = 1 |X| J. For convenience, we let Ai = Ei := O (the zero matrix) if i < 0 or i > n. Though our focus in this paper will be on Q-polynomial association schemes, we first recall the P -polynomial property for completeness. We say that the association scheme (X,R) is P -polynomial (or metric) with respect to the ordering A0, A1, . . . , An if there 168 Ars Math. Contemp. 22 (2022) #P2.01 / 163–205 are non-negative integers ai, bi, ci (0 ⩽ i ⩽ n) such that bn = c0 = 0, bi−1ci ̸= 0 (1 ⩽ i ⩽ n), and A1Ai = bi−1Ai−1 + aiAi + ci+1Ai+1 (0 ⩽ i ⩽ n), where b−1 and cn+1 are indeterminates. In this case, A1 recursively generates A, and hence has n+ 1 distinct eigenvalues θ0, θ1, . . . , θn ∈ R, where we write A1 = n∑ i=0 θiEi. (2.1) We note that (X,R) is P -polynomial as above precisely when the graph (X,R1) is a distance-regular graph and (X,Ri) is the distance-i graph of (X,R1) (0 ⩽ i ⩽ n). See, e.g., [10, 12, 27, 15] for more information on distance-regular graphs. We say that (X,R) is Q-polynomial (or cometric) with respect to the ordering E0, E1, . . . , En if there are real scalars a∗i , b ∗ i , c ∗ i (0 ⩽ i ⩽ n) such that b ∗ n = c ∗ 0 = 0, b∗i−1c ∗ i ̸= 0 (1 ⩽ i ⩽ n), and E1 ◦ Ei = 1 |X| (b∗i−1Ei−1 + a ∗ iEi + c ∗ i+1Ei+1) (0 ⩽ i ⩽ n), (2.2) where b∗−1 and c ∗ n+1 are indeterminates. In this case, |X|E1 recursively generates A with respect to ◦, and hence has n+ 1 distinct entries θ∗0 , θ∗1 , . . . , θ∗n ∈ R, where we write |X|E1 = n∑ i=0 θ∗iAi. (2.3) We call the θ∗i the dual eigenvalues of |X|E1. We may remark that E1 ◦ Ei, being a principal submatrix of E1 ⊗ Ei, is positive semidefinite, so that the scalars a∗i , b∗i , and c∗i are non-negative (the so-called Krein condition). The Q-polynomial association schemes are an important subject in their own right, and we refer the reader to [23, 29] and the references therein for recent activity. Below we give two fundamental examples ofP - andQ-polynomial association schemes, both of which come from transitive group actions. See [10, 12, 16] for the details. Example 2.4. Let v and k be positive integers with v > k, and letX be the set of k-subsets of {1, 2, . . . , v}. Set n = min{k, v − k}. For x, y ∈ X and 0 ⩽ i ⩽ n, we let (x, y) ∈ Ri if |x ∩ y| = k − i. The Ri are the orbitals of the symmetric group Sv acting on X . We call (X,R) a Johnson scheme and denote it by J(v, k). The eigenvalues of A1 are given in decreasing order by θi = (k − i)(v − k − i)− i (0 ⩽ i ⩽ n), and J(v, k) isQ-polynomial with respect to the corresponding ordering of theEi (cf. (2.1)). Example 2.5. Let q ⩾ 2 be an integer and let X = {0, 1, . . . , q − 1}n. For x, y ∈ X and 0 ⩽ i ⩽ n, we let (x, y) ∈ Ri if x and y differ in exactly i coordinate positions. The Ri are the orbitals of the wreath product Sq ≀Sn of the symmetric groups Sq and Sn acting E. Bannai et al.: Tight relative t-designs on two shells in hypercubes, . . . 169 on X . We call (X,R) a Hamming scheme and denote it by H(n, q). The eigenvalues of A1 are given in decreasing order by θi = n(q − 1)− qi (0 ⩽ i ⩽ n), andH(n, q) isQ-polynomial with respect to the corresponding ordering of theEi (cf. (2.1)). The Hamming scheme H(n, 2) is also known as the n-cube (or n-dimensional hypercube) and is denoted by Qn. Assumption 2.6. For the rest of this section and in Section 3, we assume that (X,R) is an association scheme and is Q-polynomial with respect to the ordering E0, E1, . . . , En of the primitive idempotents. In general, for any positive semidefinite Hermitian matrices B,C ∈ MX(C), we have (cf. [45]) (B ◦ C)V = span(BV ◦ CV ), where BV ◦ CV = {u ◦ v : u ∈ BV, v ∈ CV }. Hence it follows from (2.2) that span(E1V ◦ EiV ) = { Ei−1V + EiV + Ei+1V if a∗i ̸= 0, Ei−1V + Ei+1V if a∗i = 0, (0 ⩽ i ⩽ n), (2.4) from which it follows that h∑ i=0 k∑ j=0 span(EiV ◦ EjV ) = h∑ i=0 k∑ j=0 span(E1V ◦ · · · ◦ E1V︸ ︷︷ ︸ i times ◦EjV ) = h+k∑ i=0 EiV (2.5) for 0 ⩽ h, k ⩽ n. See also [10, Section 2.8]. We now fix a ‘base vertex’ x ∈ X . Let Xi = {y ∈ X : (x, y) ∈ Ri} (0 ⩽ i ⩽ n). We call the Xi the shells (or spheres or subconstituents) of (X,R) with respect to x. For every i (0 ⩽ i ⩽ n), define the diagonal matrix E∗i = E ∗ i (x) ∈MX(C) by (E∗i )y,y = { 1 if y ∈ Xi, 0 otherwise, (y ∈ X). Then we have E∗i E ∗ j = δi,jE ∗ i , n∑ i=0 E∗i = I. We call the E∗i the dual idempotents of (X,R) with respect to x. The subspace A∗ = A∗(x) = span{E∗0 , E∗1 , . . . , E∗n} 170 Ars Math. Contemp. 22 (2022) #P2.01 / 163–205 is then a subalgebra of MX(C), which we call the dual adjacency algebra of (X,R) with respect to x. The Terwilliger algebra (or subconstituent algebra) of (X,R) with respect to x is the subalgebra T = T (x) of MX(C) generated by A and A∗ [46, 47, 48]. We note that T is semisimple as it is closed under †. Remark 2.7. If (X,R) arises from a group action as in Remark 2.2, which we recall is generously transitive in this case, then T is a subalgebra of the centralizer algebra for the action of the stabilizer Gx of x in G. The two algebras are known to be equal, e.g., for J(v, k) and H(n, q); see [25, 43]. For every subset Y of X , let Ŷ ∈ V be the characteristic vector of Y , i.e., (Ŷ )y = { 1 if y ∈ Y, 0 otherwise, (y ∈ X). In particular, X̂ denotes the all-ones vector in V . We will simply write x̂ for the character- istic vector of the singleton {x}. With this notation established, we have X̂i = E ∗ i X̂ = Aix̂ (0 ⩽ i ⩽ n), from which it follows that T x̂ = span{X̂i : 0 ⩽ i ⩽ n} = span{Eix̂ : 0 ⩽ i ⩽ n}. (2.6) The T -module T x̂ is easily seen to be irreducible with dimension n + 1 (cf. [46, Lem- ma 3.6]), and is called the primary T -module. We define the dual adjacency matrix A∗1 = A ∗ 1(x) ∈MX(C) by (cf. (2.3)) A∗1 = |X|diagE1x̂ = n∑ i=0 θ∗iE ∗ i . (2.7) Since the θ∗i are mutually distinct, A ∗ 1 generates A ∗. Moreover, since A∗1v = |X|(E1x̂) ◦ v (v ∈ V ), it follows from (2.4) that EiA ∗ 1Ej = O if |i− j| > 1 (0 ⩽ i, j ⩽ n). (2.8) Let W be an irreducible T -module. We define the dual support W ∗s , the dual endpoint r∗(W ), and the dual diameter d∗(W ) of W by W ∗s = {i : EiW ̸= 0}, r∗(W ) = minW ∗s , d∗(W ) = |W ∗s | − 1, respectively. We call W dual thin if dimEiW ⩽ 1 (0 ⩽ i ⩽ n). We note that the primary T -module T x̂ is dual thin, and that it is a unique irreducible T -module up to isomorphism which has dual endpoint zero or dual diameter n. The following lemma is an easy consequence of (2.8): Lemma 2.8 ([46, Lemma 3.12]). With reference to Assumption 2.6, write A∗1 = A∗1(x), A∗ = A∗(x), T = T (x). Let W be an irreducible T -module and set r∗ = r∗(W ), d∗ = d∗(W ). Then the following hold: E. Bannai et al.: Tight relative t-designs on two shells in hypercubes, . . . 171 1. A∗1EiW ⊂ Ei−1W + EiW + Ei+1W (0 ⩽ i ⩽ n). 2. W ∗s = {r∗, r∗ + 1, . . . , r∗ + d∗}. 3. EiA∗1EjW ̸= 0 if |i− j| = 1 (r∗ ⩽ i, j ⩽ r∗ + d∗). 4. Suppose that W is dual thin. Then i∑ h=0 Er∗+hW = i∑ h=0 (A∗1) hEr∗W (0 ⩽ i ⩽ d ∗). In particular, W = A∗Er∗W . 3 Relative t-designs in Q-polynomial association schemes In this section, we develop some general theory on relative t-designs in Q-polynomial association schemes. Recall Assumption 2.6. Throughout this section, we fix a base vertex x ∈ X , and write E∗i = E ∗ i (x) (0 ⩽ i ⩽ n), A ∗ 1 = A ∗ 1(x), A ∗ = A∗(x), and T = T (x). In Introduction, we meant by a weighted subset of X a pair (Y, ω) of a subset Y of X and a function ω : Y → (0,∞). For convenience, however, we extend the domain of ω to X by setting ω(y) = 0 for every y ∈ X\Y . We will also naturally identify V with the set of complex functions on X , so that ω ∈ V and Y = suppω. In our discussions on relative t-designs, we will often consider the set L = LY = {ℓ : Y ∩Xℓ ̸= ∅}, (3.1) and say that (Y, ω) is supported on ⊔ ℓ∈LXℓ. For comparison, we begin with the algebraic definition of t-designs in (X,R) due to Delsarte [16, 17]. Definition 3.1. A weighted subset (Y, ω) of X is called a t-design in (X,R) if Eiω = 0 for 1 ⩽ i ⩽ t. Delsarte [18] generalized this concept as follows: Definition 3.2. A weighted subset (Y, ω) of X is called a relative t-design in (X,R) (with respect to x) if Eiω ∈ span{Eix̂} for 1 ⩽ i ⩽ t. Remark 3.3. Delsarte introduced the concept of t-designs for subsets Y of X in [16], i.e., when ω = Ŷ , whereas in [17, 18] he mostly considered general (i.e., not necessarily non- negative) non-zero vectors ω ∈ V in the discussions on t-designs and relative t-designs. Some facts/results below, such as Examples 3.4 and 3.5, Proposition 3.6, and Theorem 3.8, are still valid for general ω ∈ V , but the Fisher-type lower bound on |Y | = | suppω| (cf. Theorem 3.9) makes sense only when ω is non-negative. For the Johnson and Hamming schemes, Delsarte [16, 17, 18] showed that these alge- braic concepts indeed have geometric interpretations: Example 3.4. Let (X,R) be the Johnson scheme J(v, k) from Example 2.4. Then (Y, ω) is a t-design if and only if, for every t-subset z of {1, 2, . . . , v}, the sum λz of the values 172 Ars Math. Contemp. 22 (2022) #P2.01 / 163–205 ω(y) over those y ∈ Y such that z ⊂ y, is a constant independent of z. On the other hand, (Y, ω) is a relative t-design if and only if the above λz depends only on |x ∩ z|. We note that (Y, Ŷ ) is a t-design if and only if Y is a t-(v, k, λ) design (cf. [13, Chapter II.4]) for some λ. Example 3.5. Let (X,R) be the Hamming scheme H(n, q) from Example 2.5. Then (Y, ω) is a t-design if and only if, for every t-subset T of {1, 2, . . . , n} and every function f : T → {0, 1, . . . , q − 1}, the sum λT ,f of the values ω(y) over those y = (y1, y2, . . . , yn) ∈ Y such that yi = f(i) (i ∈ T ), is a constant independent of the pair (T , f). On the other hand, (Y, ω) is a relative t-design if and only if the above λT ,f depends only on |{i ∈ T : xi = f(i)}|, where x = (x1, x2, . . . , xn). We note that (Y, Ŷ ) is a t-design if and only if the transpose of the |Y | × n matrix formed by arranging the elements of Y (in any order) is an orthogonal array OA(|Y |, n, q, t) (cf. [13, Chapter III.6]). For the case q = 2, i.e., for Qn, if we choose the base vertex as x = (0, 0, . . . , 0), then (Y, Ŷ ) is a relative t-design if and only if Y is a regular t-wise balanced design of type t-(n,L, λ) (cf. [38, Section 4.4]) for some λ, where L is from (3.1), and where we identify the elements of X = {0, 1}n with their supports. Similar results hold for some other important families of P - and Q-polynomial association schemes; see, e.g., [17, 18, 19, 34, 41]. Proposition 3.6 (cf. [3, Theorem 4.5]). With reference to Assumption 2.6, let (Y, ω) be a weighted subset supported on ⊔ ℓ∈LXℓ. Then we have ω|T x̂ = ∑ ℓ∈L ⟨ω, X̂ℓ⟩ |Xℓ| X̂ℓ, (3.2) where ω|T x̂ denotes the orthogonal projection of ω on the primary T -module T x̂. More- over, (Y, ω) is a relative t-design if and only if ⟨ω, v⟩ = ⟨ω|T x̂, v⟩ = ∑ ℓ∈L ⟨ω, X̂ℓ⟩ |Xℓ| ⟨X̂ℓ, v⟩ for every v ∈ ∑t i=0EiV . Proof. Recall (2.6). The first part follows since the X̂i form an orthogonal basis of T x̂ with ∥X̂i∥2 = |Xi|. The second part is also immediate from Eiω ∈ span{Eix̂} ⇐⇒ Eiω ∈ T x̂ ⇐⇒ Eiω|T x̂ = Eiω. Remark 3.7. It is clear that (Xℓ, X̂ℓ) is a relative n-design for every 0 ⩽ ℓ ⩽ n. Hence, if (Y, ω) is a relative t-design such that Xℓ ⊂ Y for some ℓ, and if ω is constant on Xℓ, then the weighted subset (Y \Xℓ, (I − E∗ℓ )ω) obtained by discarding Xℓ from Y is again a relative t-design. This observation is particularly important when applying Theorem 3.8 below; for example, we can always assume that 0 ̸∈ L. The following is a slight generalization of Delsarte’s Assmus–Mattson theorem for Q- polynomial association schemes [18, Theorem 8.4], and can also be viewed as a variation of [9, Theorem 3.3], which in turn generalizes [28, Proposition 1]. See also [11]. The proof is in fact identical to that of [44, Theorem 4.3], but we include it below because of the potential importance of the result. E. Bannai et al.: Tight relative t-designs on two shells in hypercubes, . . . 173 Theorem 3.8. With reference to Assumption 2.6, let (Y, ω) be a relative t-design supported on ⊔ ℓ∈LXℓ. Then (Y ∩Xℓ, E∗ℓω) is a relative (t− |L|+ 1)-design for every ℓ ∈ L. Proof. Let U = (T x̂)⊥ be the orthogonal complement of T x̂ in V , which we recall is the sum of all the non-primary irreducible T -modules in V . On the one hand, we have ω|U ∈ ∑ ℓ∈L E∗ℓU. Since A∗1 generates A ∗ and has at most |L| distinct eigenvalues on this subspace (cf. (2.7)), it follows that A∗ω|U = span { ω|U , A∗1ω|U , . . . , (A∗1)|L|−1ω|U } . (3.3) On the other hand, since E0U = 0, that (Y, ω) is a relative t-design is rephrased as ω|U ∈ n∑ i=t+1 EiU. Hence it follows from (2.8) and (3.3) that A∗ω|U ⊂ |L|−1∑ k=0 (A∗1) k n∑ i=t+1 EiU ⊂ n∑ i=t−|L|+2 EiU. In particular, for every ℓ ∈ L we have E∗ℓω|U ∈ n∑ i=t−|L|+2 EiU. In other words, (Y ∩Xℓ, E∗ℓω) is a relative (t− |L|+ 1)-design, as desired. Bannai and Bannai [3, Theorem 4.8] established the following Fisher-type lower bound on the size of a relative t-design with t even: Theorem 3.9. With reference to Assumption 2.6, let (Y, ω) be a relative 2e-design (e ∈ N) supported on ⊔ ℓ∈LXℓ. Then |Y | ⩾ dim (∑ ℓ∈L E∗ℓ )( e∑ i=0 EiV ) . Definition 3.10. A relative 2e-design (Y, ω) is called tight if equality holds above. Recall from Example 3.5 that the relative t-designs in the hypercubes are equivalent to the weighted regular t-wise balanced designs. Example 3.11. Let (X,R) be the n-cube Qn from Example 2.5. Xiang [51] showed that if e ⩽ ℓ ⩽ n− e for every ℓ ∈ L, then dim (∑ ℓ∈L E∗ℓ )( e∑ i=0 EiV ) = min{|L|−1,e}∑ i=0 ( n e− i ) . (3.4) 174 Ars Math. Contemp. 22 (2022) #P2.01 / 163–205 We may remark that (cf. [12, Theorem 9.2.1]) dimEiV = ( n i ) (0 ⩽ i ⩽ n). (3.5) See also [32] and [6, Theorem 2.7, Example 2.9]. Example 3.12. Consider a symmetric 2-(n + 1, k, λ) design (cf. [13, Chapter II.6]). Ob- serve that removing a point yields a tight relative 2-design in Qn with L = {k − 1, k}. Likewise, taking the complement of every block which contains a given point followed by removing that point gives rise to a tight relative 2-design in Qn with L = {k, n+ 1− k}. The complement of this is yet another example1 such that L = {k − 1, n − k}. See [32, Section 3] and [50, Theorem 8]. Note that the weights are constant for these three exam- ples. On the other hand, Bannai, Bannai, and Bannai [5, Theorem 2.2] showed that there is a tight relative 2-design in Qn with L = {2, n/2} for n ≡ 6 (mod 8), provided that a Hadamard matrix of order n/2 + 1 exists. This construction provides examples in which the weights take two distinct values depending on the shells. See also [53]. Example 3.13. Working with the tight 4-(23, 7, 1) and 4-(23, 16, 52) designs instead of a symmetric 2-(n+1, k, λ) design as in Example 3.12, we obtain four tight relative 4-designs in Q22 with constant weight such that L ∈ { {6, 7}, {6, 15}, {7, 16}, {15, 16} } . See [9, Theorem 6.3] and [32, Section 3]. Let (Y, ω) be a tight relative 2e-design supported on ⊔ ℓ∈LXℓ. Bannai, Bannai, and Bannai [5, Theorem 2.1] showed that if the stabilizer of x in the automorphism group of (X,R) acts transitively on each of the shells Xi then ω is constant on Y ∩ Xℓ for every ℓ ∈ L. The next theorem generalizes this result by replacing group actions by combinatorial regularity. Observe that the fibers of the coherent closure of T are in general finer than the shells Xi. Theorem 3.14. With reference to Assumption 2.6, let (Y, ω) be a tight relative 2e-design (e ∈ N) supported on ⊔ ℓ∈LXℓ. For every ℓ ∈ L, the weight ω is constant on Y ∩ Xℓ provided that Xℓ remains a fiber of the coherent closure of T . Proof. Let (cf. (3.2)) D = diagω, D̃ = diagω|T x̂ = ∑ ℓ∈L ⟨ω, X̂ℓ⟩ |X̂ℓ| E∗ℓ . Note that D̃ ∈ T . Let F be the orthogonal projection onto BV , where B = √ D̃ e∑ i=0 Ei ∈ T . Observe that BV = (BB†)V, 1It seems that this construction is missing in [50, Theorem 8]. E. Bannai et al.: Tight relative t-designs on two shells in hypercubes, . . . 175 and that F is written as a polynomial in the Hermitian (in fact, real and symmetric) matrix BB†. In particular, F ∈ T . Since (Y, ω) is tight, we have dimBV = dim √ D̃ (∑ ℓ∈L E∗ℓ )( e∑ i=0 EiV ) = |Y |. Let u1, u2, . . . , u|Y | be an orthonormal basis of BV , and let G = [ u1 u2 · · · u|Y | ] . Then we have F = GG†. (3.6) Let D′ = D|Y×Y , D̃′ = D̃|Y×Y , F ′ = F |Y×Y , G′ = G|Y×{1,2,...,|Y |}, where |Y×Y etc. mean taking corresponding submatrices. Note that these are square matri- ces, and that D′ and D̃′ are invertible. Then it follows that (G′)†D′(D̃′)−1G′ = I|Y |. (3.7) Indeed, since we may write ui = √ D̃ vi, where vi ∈ e∑ r=0 ErV (1 ⩽ i ⩽ |Y |), it follows from (2.5) (applied to h = k = e) and Proposition 3.6 that the (i, j)-entry of the LHS in (3.7) is equal to (vi) †Dvj = ⟨ω, vi ◦ vj⟩ = ⟨ω|T x̂, vi ◦ vj⟩ = (vi)†D̃vj = ⟨uj , ui⟩ = δi,j , where means complex conjugate. By (3.6) and (3.7), we have I|Y | = D ′(D̃′)−1G′(G′)† = D′(D̃′)−1F ′, so that (D′)−1 = (D̃′)−1F ′. (3.8) In particular, F ′ is a diagonal matrix. Now, let ℓ ∈ L and suppose that Xℓ remains a fiber of the coherent closure of T . Then the (y, y)-entry of F ∈ T is constant for y ∈ Xℓ (cf. (A1) and (A2)), and the same is true for D̃. Hence it follows from (3.8) that ω(y) = Dy,y must be constant for y ∈ Y ∩ Xℓ. This completes the proof. 176 Ars Math. Contemp. 22 (2022) #P2.01 / 163–205 4 The Terwilliger algebra of Qn For the rest of this paper, we will focus on relative t-designs in the n-cube Qn from Ex- ample 2.5. We will need detailed descriptions of the Terwilliger algebra ofQn and its irre- ducible modules, and we collect these in this section. Thus, we assume that (X,R) = Qn, where X = {0, 1}n. We again fix a base vertex x ∈ X , and write E∗i = E∗i (x) (0 ⩽ i ⩽ n), A∗1 = A ∗ 1(x), and T = T (x). The Q-polynomial ordering we consider is the one given in Example 2.5.2 Proposition 4.1 (cf. [39, Section I.C]). We have T = span{E∗i AjE∗k : 0 ⩽ i, j, k ⩽ n}. (4.1) In particular, T is a coherent algebra. Proof. The RHS in (4.1) is a subspace of T . Recall from Example 2.5 that Qn admits the action of G = S2 ≀ Sn. The stabilizer Gx of x in G is isomorphic to Sn, and it is immediate to see that every orbital of Gx is of the form {(y, z) ∈ X ×X : (x, y) ∈ Ri, (y, z) ∈ Rj , (z, x) ∈ Rk} for some i, j, and k, where the corresponding adjacency matrix is E∗i AjE ∗ k . Hence the RHS in (4.1) agrees with the centralizer algebra for the action of Gx on X , which is a coherent algebra; cf. Remark 2.3. Since T is generated by the Ai and the E∗i , the result follows. Lemma 4.2. For 0 ⩽ i, j, k ⩽ n, we have E∗i AjE∗k ̸= O if and only if j ∈ { |i− k|, |i− k|+ 2, |i− k|+ 4, . . . ,min{i+ k, 2n− i− k} } . Proof. Routine. Next we recall basic facts about the irreducible T -modules. Let W be an irreducible T -module. We define the support Ws, the endpoint r(W ), and the diameter d(W ) of W by Ws = {i : E∗iW ̸= 0}, r(W ) = minWs, d(W ) = |Ws| − 1, respectively. We call W thin if dimE∗iW ⩽ 1 (0 ⩽ i ⩽ n). Theorem 4.3 (cf. [26]). Let W be an irreducible T -module and set r = r(W ), r∗ = r∗(W ), d = d(W ), and d∗ = d∗(W ). Then W is thin, dual thin, and we have r = r∗, d = d∗ = n− 2r, Ws =W ∗s = {r, r + 1, . . . , n− r}. Moreover, the isomorphism class of W is determined by r. Remark 4.4. Recall that the universal enveloping algebra U(sl2(C)) is defined by the generators x, y, h and the relations xy− yx = h, hx− xh = 2x, hy− yh = −2y. 2If n is even then Qn has another Q-polynomial ordering E0, En−1, E2, En−3, . . . in terms of the natural ordering; cf. [10, p. 305]. E. Bannai et al.: Tight relative t-designs on two shells in hypercubes, . . . 177 There is a surjective homomorphism U(sl2(C))→ T such that (cf. [26, Lemma 7.5]) x 7→ n∑ i=1 Ei−1A ∗ 1Ei, y 7→ n−1∑ i=0 Ei+1A ∗ 1Ei, h 7→ A1. Every irreducible T -module is then irreducible as an sl2(C)-module. We also obtain an- other surjective homomorphism U(sl2(C)) → T by interchanging A1 and A∗1 and replac- ing the Ei by the E∗i above; cf. [26, Lemma 5.3]. From now on, we fix an orthogonal irreducible decomposition V = ⊕ W∈Λ W (4.2) of the standard module V . In view of Theorem 4.3, let Λr = {W ∈ Λ : r(W ) = r∗(W ) = r} (0 ⩽ r ⩽ ⌊n/2⌋), (4.3) and fix a unit vector vW ∈ ErW for each W ∈ Λr. Since dimEiV = ∑ W∈Λ dimEiW = i∑ r=0 |Λr| (0 ⩽ i ⩽ ⌊n/2⌋) (4.4) by Theorem 4.3, it follows from (3.5) that |Λr| = ( n r ) − ( n r − 1 ) (0 ⩽ r ⩽ ⌊n/2⌋). It is known (cf. [26, Theorem 9.2]) that if W ∈ Λr then the vectors E∗r vW , E ∗ r+1vW , . . . , E ∗ n−rvW (4.5) form an orthogonal basis of W , called a standard basis of W . By [26, Lemma 6.6], we also have ∥E∗i vW ∥2 = ( n− 2r i− r ) ∥E∗r vW ∥2 (r ⩽ i ⩽ n− r). (4.6) We note that 1 = ∥vW ∥2 = n−r∑ i=r ∥E∗i vW ∥2 = 2n−2r∥E∗r vW ∥2. (4.7) For W,W ′ ∈ Λr, we observe that the linear map W →W ′ defined by E∗i vW 7→ E∗i vW ′ (r ⩽ i ⩽ n− r) is an isometric isomorphism of T -modules. Let Ĕi,jr = 2n−2r√( n−2r i−r )( n−2r j−r ) ∑ W∈Λr (E∗i vW )(E ∗ j vW ) † (r ⩽ i, j ⩽ n− r). (4.8) 178 Ars Math. Contemp. 22 (2022) #P2.01 / 163–205 Then we have (Ĕi,jr ) † = Ĕj,ir (r ⩽ i, j ⩽ n− r), (4.9) and from (4.6) and (4.7) it follows that Ĕi,jr Ĕ i′,j′ r′ = δr,r′δj,i′Ĕ i,j′ r for 0 ⩽ r, r′ ⩽ ⌊n/2⌋, r ⩽ i, j ⩽ n − r, and r′ ⩽ i′, j′ ⩽ n − r′. By Theorem 4.3 and Wedderburn’s theorem (cf. [14, Section 3]), T is isomorphic to the direct sum of full matrix algebras T ∼= ⌊n/2⌋⊕ r=0 Mn−2r+1(C), and the Ĕi,jr form an orthogonal basis of T . See also [24, Section 2]. We note that E∗i TE ∗ j = span { Ĕi,jr : 0 ⩽ r ⩽ min{i, j, n− i, n− j} } (0 ⩽ i, j ⩽ n). (4.10) We now recall the Hahn polynomials [31, Section 1.5] Qr(ξ;α, β,N) = 3F2 ( −ξ,−r, r + α+ β + 1 α+ 1,−N ∣∣∣∣ 1) ∈ R[ξ] (0 ⩽ r ⩽ N), (4.11) where sFt ( a1, . . . , as b1, . . . , bt ∣∣∣∣ c) = ∞∑ i=0 (a1)i · · · (as)i (b1)i · · · (bt)i ci i! , and (a)i = a(a+ 1) · · · (a+ i− 1). For α, β > −1, or for α, β < −N , we have N∑ ξ=0 ( α+ ξ ξ )( β +N − ξ N − ξ ) Qr(ξ;α, β,N)Qr′(ξ;α, β,N) = δr,r′ (−1)r(r + α+ β + 1)N+1(β + 1)rr! (2r + α+ β + 1)(α+ 1)r(−N)rN ! . (4.12) Our aim is to describe the entries of the Ĕi,jr . In view of (4.9), we will assume for the rest of this section that 0 ⩽ i ⩽ j ⩽ n. By Proposition 4.1 and Lemma 4.2, we have E∗i TE ∗ j = span { E∗i A2ξ+j−iE ∗ j : 0 ⩽ ξ ⩽ min{i, n− j} } . Moreover, it follows that (cf. (4.10)) E∗i A2ξ+j−iE ∗ j = min{i,n−j}∑ r=0 3F2 ( −ξ,−r, r − n− 1 j − n,−i ∣∣∣∣ 1) ( j i−ξ )( n−j ξ )( j−r j−i )√( n−2r j−r ) ( j i )√( n−2r i−r ) Ĕi,jr . (4.13) E. Bannai et al.: Tight relative t-designs on two shells in hypercubes, . . . 179 This formula can be found in [33, Section 10]. See also [39, 49] for similar calculations. If i ⩽ n− j then 3F2 ( −ξ,−r, r − n− 1 j − n,−i ∣∣∣∣ 1) = Qr(ξ; j − n− 1,−j − 1, i). Since ( j i− ξ )( n− j ξ ) = (−1)i ( j − n− 1 + ξ ξ )( −j − 1 + i− ξ i− ξ ) , it follows from (4.12) (applied to α = j − n− 1, β = −j − 1, N = i) and (4.13) that, for 0 ⩽ r ⩽ i, i∑ ξ=0 3F2 ( −ξ,−r, r − n− 1 j − n,−i ∣∣∣∣ 1)E∗i A2ξ+j−iE∗j = (−1)r(r − n− 1)i+1(−j)rr! (2r − n− 1)(j − n)r(−i)ri! · (−1)i ( j−r j−i )√( n−2r j−r ) ( j i )√( n−2r i−r ) Ĕi,jr = ( n i )( n−i r )√( n−2r j−r )(( n r ) − ( n r−1 ))( n−j r )√( n−2r i−r ) Ĕi,jr . Likewise, if n− j ⩽ i then 3F2 ( −ξ,−r, r − n− 1 j − n,−i ∣∣∣∣ 1) = Qr(ξ;−i− 1, i− n− 1, n− j). In this case, since( j i− ξ )( n− j ξ )( j i )−1 = (−1)n−j ( −i− 1 + ξ ξ )( i− 1− j − ξ n− j − ξ )( n− i n− j )−1 , again it follows from (4.12) (applied to α = −i− 1, β = i−n− 1, N = n− j) and (4.13) that, for 0 ⩽ r ⩽ n− j, n−j∑ ξ=0 3F2 ( −ξ,−r, r − n− 1 j − n,−i ∣∣∣∣ 1)E∗i A2ξ+j−iE∗j = (−1)r(r − n− 1)n−j+1(i− n)rr! (2r − n− 1)(−i)r(j − n)r(n− j)! · (−1)n−j ( j−r j−i )√( n−2r j−r ) ( n−i n−j )√( n−2r i−r ) Ĕi,jr = ( n i )( n−i r )√( n−2r j−r )(( n r ) − ( n r−1 ))( n−j r )√( n−2r i−r ) Ĕi,jr . In either case, it follows that Ĕi,jr = (( n r ) − ( n r−1 ))( n−j r )√( n−2r i−r ) ( n i )( n−i r )√( n−2r j−r ) 180 Ars Math. Contemp. 22 (2022) #P2.01 / 163–205 × min{i,n−j}∑ ξ=0 3F2 ( −ξ,−r, r − n− 1 j − n,−i ∣∣∣∣ 1)E∗i A2ξ+j−iE∗j (4.14) for 0 ⩽ i ⩽ j ⩽ n and 0 ⩽ r ⩽ min{i, n− j}. 5 Tight relative 2e-designs on two shells in Qn We retain the notation of the previous sections. In this section, we discuss tight relative 2e-designs (Y, ω) in Qn supported on two shells Xℓ ⊔ Xm, i.e., L = {ℓ,m} (cf. (3.1)). Recall from (3.4) that we have in this case |Y | = ( n e ) + ( n e− 1 ) , but recall also that this is valid under the additional condition that e ⩽ ℓ,m ⩽ n− e. How- ever, both (Y ∩Xℓ, E∗ℓω) and (Y ∩Xm, E∗mω) are relative (2e− 1)-designs by Theorem 3.8, so that if ℓ < 2e or ℓ > n − 2e for example, then (Y ∩ Xℓ, E∗ℓω) must be trivial in view of Example 3.5, i.e., Xℓ ⊂ Y and ω is constant on Xℓ, and hence (Y ∩Xm, E∗mω) is by itself a relative 2e-design; cf. Remark 3.7. This shows that the above condition is not a restrictive one. We also note that Lemma 5.1. Let (Y, ω) be a relative t-design inQn supported on ⊔ ℓ∈LXℓ. Then (Y ′, Anω) is a relative t-design supported on ⊔ ℓ∈LXn−ℓ, where Y ′ = {y′ : y ∈ Y }, and for every y ∈ X , y′ denotes the unique vertex such that (y, y′) ∈ Rn. Proof. Immediate from EiAn ∈ span{Ei} (0 ⩽ i ⩽ n). In view of the above comments, we now make the following assumption: Assumption 5.2. In this section, let (Y, ω) be a tight relative 2e-design (e ∈ N) in Qn supported on two shells Xℓ ⊔Xm, where e ⩽ ℓ < m ⩽ n− ℓ (⩽ n− e). Our aim is to show that Y then induces the structure of a coherent configuration with two fibers, and to obtain a necessary condition on the existence of such (Y, ω) akin to Delsarte’s theorem on tight 2e-designs. To this end, we first recall the proof of (3.4) given in [6, Theorem 2.7, Example 2.9] under the above assumption. For convenience, set E∗L = E ∗ ℓ + E ∗ m. By (4.2) and (4.3), we have E∗L ( e∑ i=0 EiV ) = e∑ r=0 ∑ W∈Λr E∗L ( e∑ i=r EiW ) . (5.1) Let W ∈ Λr, where 0 ⩽ r ⩽ e. Recall Theorem 4.3 and also the standard basis (4.5) of W . If r = e then EeW is spanned by vW , and hence we have E∗LEeW = span{E∗LvW }. E. Bannai et al.: Tight relative t-designs on two shells in hypercubes, . . . 181 Note that E∗LvW is non-zero by Assumption 5.2, and hence dimE∗LEeW = 1 in this case. Suppose next that 0 ⩽ r < e. On the one hand, since E∗L ( e∑ i=r EiW ) ⊂ E∗LW = E∗ℓW + E∗mW, we have dimE∗L ( e∑ i=r EiW ) ⩽ 2. On the other hand, it follows from (2.8) that vW , A ∗ 1vW ∈ ErW + Er+1W ⊂ e∑ i=r EiW, (5.2) and hence E∗LvW , E ∗ LA ∗ 1vW ∈ E∗L ( e∑ i=r EiW ) . Moreover, we have (cf. (2.7)) E∗LvW = E ∗ ℓ vW + E ∗ mvW , E ∗ LA ∗ 1vW = θ ∗ ℓE ∗ ℓ vW + θ ∗ mE ∗ mvW , so that these two vectors are non-zero and are linearly independent by Assumption 5.2 and since θ∗ℓ ̸= θ∗m. It follows that dimE∗L ( e∑ i=r EiW ) = 2. Note that in this case we in fact have E∗L ( e∑ i=r EiW ) = span{E∗ℓ vW , E∗mvW }, as E∗ℓ vW = E ∗ L θ∗mI −A∗1 θ∗m − θ∗ℓ vW , E ∗ mvW = E ∗ L θ∗ℓ I −A∗1 θ∗ℓ − θ∗m vW . (5.3) Combining these comments, we now obtain (3.4) as follows: dimE∗L ( e∑ i=0 EiV ) = e∑ r=0 ∑ W∈Λr dimE∗L ( e∑ i=r EiW ) = |Λe|+ e−1∑ r=0 2|Λr| = dimEeV + dimEe−1V 182 Ars Math. Contemp. 22 (2022) #P2.01 / 163–205 = ( n e ) + ( n e− 1 ) , where we have used (3.5) and (4.4). By the above discussions, the set of vectors below forms an orthogonal basis of the subspace (5.1): ( e−1⊔ r=0 ⊔ W∈Λr {E∗ℓ vW , E∗mvW } )⊔( ⊔ W∈Λe {E∗LvW } ) . As in the proof of Theorem 3.14, let D = diagω. We next apply √ D to the above basis vectors and compute their inner products. First, let W,W ′ ∈ ⊔e−1 r=0 Λr. It is clear that〈√ DE∗ℓ vW , √ DE∗mvW ′ 〉 = 〈√ DE∗mvW , √ DE∗ℓ vW ′ 〉 = 0. (5.4) By (5.3), we have (E∗ℓ vW ) ◦ (E∗ℓ vW ′) = E∗Lu, where means complex conjugate, and u = ( θ∗mI −A∗1 θ∗m − θ∗ℓ vW ) ◦ ( θ∗mI −A∗1 θ∗m − θ∗ℓ vW ′ ) . Observe that u belongs to ∑2e i=0EiV by (2.5) (applied to h = k = e) and (5.2). Hence, by Proposition 3.6 we have〈√ DE∗ℓ vW , √ DE∗ℓ vW ′ 〉 = ⟨ω,E∗Lu⟩ = ⟨ω, u⟩ = ⟨ω, X̂ℓ⟩ |Xℓ| ⟨X̂ℓ, u⟩+ ⟨ω, X̂m⟩ |Xm| ⟨X̂m, u⟩ = ⟨ω, X̂ℓ⟩ |Xℓ| ⟨X̂ℓ, E∗Lu⟩ = ⟨ω, X̂ℓ⟩ |Xℓ| ⟨E∗ℓ vW , E∗ℓ vW ′⟩ = δW,W ′ ⟨ω, X̂ℓ⟩ |Xℓ| ∥E∗ℓ vW ∥2. (5.5) Likewise, we have〈√ DE∗mvW , √ DE∗mvW ′ 〉 = δW,W ′ ⟨ω, X̂m⟩ |Xm| ∥E∗mvW ∥2. (5.6) Next, let W ∈ ⊔e−1 r=0 Λr and W ′ ∈ Λe. Then, by the same argument we have〈√ DE∗ℓ vW , √ DE∗LvW ′ 〉 = 〈√ DE∗mvW , √ DE∗LvW ′ 〉 = 0. (5.7) E. Bannai et al.: Tight relative t-designs on two shells in hypercubes, . . . 183 Finally, let W,W ′ ∈ Λe. In this case, we have〈√ DE∗LvW , √ DE∗LvW ′ 〉 = δW,W ′ ( ⟨ω, X̂ℓ⟩ |X̂ℓ| ∥E∗ℓ vW ∥2 + ⟨ω, X̂m⟩ |X̂m| ∥E∗mvW ∥2 ) . (5.8) Since (Y, ω) is a tight relative 2e-design, it follows from (5.4) – (5.8) that the set of vectors below is an orthogonal basis of the subspace √ DV = span{ŷ : y ∈ Y } of dimension |Y | = ( n e ) + ( n e−1 ) :( e−1⊔ r=0 ⊔ W∈Λr {√ DE∗ℓ vW , √ DE∗mvW })⊔( ⊔ W∈Λe {√ DE∗LvW }) . For convenience, set Yℓ = Y ∩Xℓ, Ym = Y ∩Xm. We will naturally make the following identification by discarding irrelevant entries: √ DE∗ℓ V = span{ŷ : y ∈ Yℓ} ←→ CYℓ ,√ DE∗mV = span{ŷ : y ∈ Ym} ←→ CYm . We write Λr = { W 1r ,W 2 r , . . . ,W |Λr| r } (0 ⩽ r ⩽ e). For 0 ⩽ r ⩽ e, define a |Yℓ| × |Λr| matrix Hℓr and a |Ym| × |Λr| matrix Hmr by Hℓr = [√ DE∗ℓ vW 1r · · · √ DE∗ℓ vW |Λr|r ] , Hmr = [√ DE∗mvW 1r · · · √ DE∗mvW |Λr|r ] . We then define a characteristic matrix H of (Y, ω) by H = [ Hℓ0 · · · Hℓe−1 O · · · O Hℓe O · · · O Hm0 · · · Hme−1 Hme ] . We note that H is a square matrix of size |Y | = ( n e ) + ( n e−1 ) . By (4.6), (4.7), and (5.4) – (5.8), and since |Xi| = ( n i ) (0 ⩽ i ⩽ n), we have H†H = ( e−1 ⊕ r=0 κℓrI|Λr| ) ⊕ ( e−1 ⊕ r=0 κmr I|Λr| ) ⊕ κeI|Λe|, (5.9) where κℓr = ωℓ ( n−2r ℓ−r ) 2n−2r ( n ℓ ) , κmr = ωm ( n−2r m−r ) 2n−2r ( n m ) (0 ⩽ r < e), 184 Ars Math. Contemp. 22 (2022) #P2.01 / 163–205 κe = ωℓ ( n−2e ℓ−e ) 2n−2e ( n ℓ ) + ωm(n−2em−e) 2n−2e ( n m ) , and we abbreviate ωℓ = ⟨ω, X̂ℓ⟩, ωm = ⟨ω, X̂m⟩. Let K denote the diagonal matrix on the RHS in (5.9). Then it follows that I|Y | = HK −1H† = [ ∑e r=0 1 κℓr Hℓr(H ℓ r) † 1 κe Hℓe(H m e ) † 1 κe Hme (H ℓ e) † ∑e r=0 1 κmr Hmr (H m r ) † ] , (5.10) where we write κℓe = κ m e := κe (5.11) for brevity. In particular, we have 1 κe Hℓe(H m e ) † = O. (5.12) Moreover, from (5.9) and (5.10) it follows that( 1 κℓr Hℓr(H ℓ r) † )( 1 κℓr′ Hℓr′(H ℓ r′) † ) = δr,r′ 1 κℓr Hℓr(H ℓ r) † (0 ⩽ r, r′ < e), (5.13) 1 κe Hℓe(H ℓ e) † = I|Yℓ| − e−1∑ r=0 1 κℓr Hℓr(H ℓ r) †, (5.14)( 1 κmr Hmr (H m r ) † )( 1 κmr′ Hmr′ (H m r′ ) † ) = δr,r′ 1 κmr Hmr (H m r ) † (0 ⩽ r, r′ < e), (5.15) 1 κe Hme (H m e ) † = I|Ym| − e−1∑ r=0 1 κmr Hmr (H m r ) †. (5.16) Note that the matrices (κℓr) −1Hℓr(H ℓ r) †, (κmr ) −1Hmr (H m r ) † (0 ⩽ r < e) are non-zero since Hℓr , H m r are non-zero. Likewise, by setting κr = √ κℓrκ m r (0 ⩽ r < e) for brevity, we have( 1 κℓr Hℓr(H ℓ r) † )( 1 κr′ Hℓr′(H m r′ ) † ) = δr,r′ 1 κr Hℓr(H m r ) † (0 ⩽ r, r′ < e), (5.17)( 1 κr Hℓr(H m r ) † )( 1 κmr′ Hmr′ (H m r′ ) † ) = δr,r′ 1 κr Hℓr(H m r ) † (0 ⩽ r, r′ < e), (5.18)( 1 κr Hℓr(H m r ) † )( 1 κr′ Hmr′ (H ℓ r′) † ) = δr,r′ 1 κℓr Hℓr(H ℓ r) † (0 ⩽ r, r′ < e), (5.19)( 1 κr Hmr (H ℓ r) † )( 1 κr′ Hℓr′(H m r′ ) † ) = δr,r′ 1 κmr Hmr (H m r ) † (0 ⩽ r, r′ < e). (5.20) E. Bannai et al.: Tight relative t-designs on two shells in hypercubes, . . . 185 Since the matrices (κℓr) −1Hℓr(H ℓ r) †, (κmr ) −1Hmr (H m r ) † (0 ⩽ r < e) are non-zero, it fol- lows from (5.17) – (5.20) that the matrices (κr)−1Hℓr(H m r ) † (0 ⩽ r < e) are non-zero and are linearly independent. It follows from Theorem 3.14 and Proposition 4.1 that ω is constant on each of Yℓ and Ym, from which it follows that Dy,y = ω(y) =  ωℓ |Yℓ| if y ∈ Yℓ, ωm |Ym| if y ∈ Ym. (5.21) Hence, by comparing with the formula (4.8) for the matrices Ĕi,jr , we have 1 κℓr Hℓr(H ℓ r) † = ( n ℓ ) |Yℓ| Ĕℓ,ℓr |Yℓ×Yℓ (0 ⩽ r < e), (5.22) 1 κe Hℓe(H ℓ e) † = ωℓ ( n−2e ℓ−e ) 2n−2eκe|Yℓ| Ĕℓ,ℓe |Yℓ×Yℓ , (5.23) 1 κmr Hmr (H m r ) † = ( n m ) |Ym| Ĕm,mr |Ym×Ym (0 ⩽ r < e), (5.24) 1 κe Hme (H m e ) † = ωm ( n−2e m−e ) 2n−2eκe|Ym| Ĕm,me |Ym×Ym , (5.25) 1 κr Hℓr(H m r ) † = √( n ℓ )( n m )√ |Yℓ||Ym| Ĕℓ,mr |Yℓ×Ym (0 ⩽ r < e), (5.26) 1 κe Hℓe(H m e ) † = √ ωℓωm ( n−2e ℓ−e )( n−2e m−e ) 2n−2eκe √ |Yℓ||Ym| Ĕℓ,me |Yℓ×Ym , (5.27) where |Yℓ×Yℓ etc. mean taking corresponding submatrices. From (5.23) and (5.25) it fol- lows that the matrices (κe)−1Hℓe(H ℓ e) †, (κe) −1Hme (H m e ) † are also non-zero, since each of Ĕℓ,ℓe |Yℓ×Yℓ , Ĕm,me |Ym×Ym has non-zero constant diagonal entries by (4.14). Let H ′ be the set consisting of the |Y | × |Y | matrices of the form[ ∑e r=0 a ℓ,ℓ r 1 κℓr Hℓr(H ℓ r) † ∑e−1 r=0 a ℓ,m r 1 κr Hℓr(H m r ) †∑e−1 r=0 a m,ℓ r 1 κr Hmr (H ℓ r) † ∑e r=0 a m,m r 1 κmr Hmr (H m r ) † ] , where aℓ,ℓr etc. are in C, and we are again using the notation (5.11). By (5.13) – (5.20) and the above comments, H ′ is a C-algebra with dimH ′ = 4e+ 2. (5.28) Define Sℓ,ℓ(Y ) = { j : Rj ∩ (Yℓ × Yℓ) ̸= ∅ } , and define Sℓ,m(Y )(= Sm,ℓ(Y )) and Sm,m(Y ) in the same manner. Let H be the set consisting of the |Y | × |Y | matrices of the form[ ∑ j∈Sℓ,ℓ(Y ) b ℓ,ℓ j Aj |Yℓ×Yℓ ∑ j∈Sℓ,m(Y ) b ℓ,m j Aj |Yℓ×Ym∑ j∈Sm,ℓ(Y ) b m,ℓ j Aj |Ym×Yℓ ∑ j∈Sm,m(Y ) b m,m j Aj |Ym×Ym ] , (5.29) 186 Ars Math. Contemp. 22 (2022) #P2.01 / 163–205 where bℓ,ℓj etc. are in C. Then H is a C-vector space with dimH = |Sℓ,ℓ(Y )|+ |Sℓ,m(Y )|+ |Sm,ℓ(Y )|+ |Sm,m(Y )|. (5.30) Note that H is closed under ◦. By (5.22) – (5.26) and Proposition 4.1 (or (4.14)), H ′ is a subspace of H . By (4.14), (5.14), (5.22), and (5.23), we have I|Yℓ| = e−1∑ r=0 ( n ℓ ) |Yℓ| Ĕℓ,ℓr |Yℓ×Yℓ + ωℓ ( n−2e ℓ−e ) 2n−2eκe|Yℓ| Ĕℓ,ℓe |Yℓ×Yℓ = 1 |Yℓ| min{ℓ,n−ℓ}∑ ξ=0  e−1∑ r=0 (( n r ) − ( n r − 1 )) 3F2 ( −ξ,−r, r − n− 1 ℓ− n,−ℓ ∣∣∣∣ 1) + ωℓ ( n−2e ℓ−e )(( n e ) − ( n e−1 )) 2n−2eκe ( n ℓ ) 3F2(−ξ,−e, e− n− 1ℓ− n,−ℓ ∣∣∣∣ 1) A2ξ|Yℓ×Yℓ . (5.31) Hence it follows that {ξ ̸= 0 : 2ξ ∈ Sℓ,ℓ(Y )} is a set of zeros of the polynomial ψℓ,ℓe (ξ) = e−1∑ r=0 (( n r ) − ( n r − 1 )) 3F2 ( −ξ,−r, r − n− 1 ℓ− n,−ℓ ∣∣∣∣ 1) + ωℓ ( n−2e ℓ−e )(( n e ) − ( n e−1 )) 2n−2eκe ( n ℓ ) 3F2(−ξ,−e, e− n− 1ℓ− n,−ℓ ∣∣∣∣ 1) ∈ R[ξ]. (5.32) Note that ψℓ,ℓe (ξ) has degree exactly e, from which it follows that |Sℓ,ℓ(Y )| ⩽ e+ 1. (5.33) Likewise, we find that {ξ ̸= 0 : 2ξ ∈ Sm,m(Y )} is a set of zeros of the polynomial ψm,me (ξ) = e−1∑ r=0 (( n r ) − ( n r − 1 )) 3F2 ( −ξ,−r, r − n− 1 m− n,−m ∣∣∣∣ 1) + ωm ( n−2e m−e )(( n e ) − ( n e−1 )) 2n−2eκe ( n m ) 3F2(−ξ,−e, e− n− 1m− n,−m ∣∣∣∣ 1) ∈ R[ξ], (5.34) and hence that |Sm,m(Y )| ⩽ e+ 1. (5.35) Finally, by (4.14), (5.12), and (5.27), we have O = √ ωℓωm ( n−2e ℓ−e )( n−2e m−e ) 2n−2eκe √ |Yℓ||Ym| Ĕℓ,me |Yℓ×Ym = √ ωℓωm ( n−m e )( n−2e ℓ−e )(( n e ) − ( n e−1 )) 2n−2eκe √ |Yℓ||Ym| ( n ℓ )( n−ℓ e ) E. Bannai et al.: Tight relative t-designs on two shells in hypercubes, . . . 187 × min{ℓ,n−m}∑ ξ=0 3F2 ( −ξ,−e, e− n− 1 m− n,−ℓ ∣∣∣∣ 1)A2ξ+m−ℓ|Yℓ×Ym . Hence it follows that {ξ : 2ξ +m− ℓ ∈ Sℓ,m(Y )} is a set of zeros of the polynomial ψℓ,me (ξ) = 3F2 ( −ξ,−e, e− n− 1 m− n,−ℓ ∣∣∣∣ 1) ∈ R[ξ], (5.36) and that |Sℓ,m(Y )| = |Sm,ℓ(Y )| ⩽ e. (5.37) By (5.30), (5.33), (5.35), and (5.37), we have dimH ⩽ 4e+ 2. Since H ′ is a subspace of H , it follows from (5.28) that H = H ′. In particular, H is a C-algebra. It is also clear that H is closed under † and contains J|Y |. We now conclude that H is a coherent algebra. Note also that equality holds in each of (5.33), (5.35), and (5.37). To summarize: Theorem 5.3. Recall Assumption 5.2. With the above notation, the following hold: (i) The set H from (5.29) is a coherent algebra of type [ e+1 e e e+1 ] . (ii) The sets of zeros of the polynomials ψℓ,ℓe (ξ), ψ m,m e (ξ), and ψ ℓ,m e (ξ) from (5.32), (5.34), and (5.36) are given respectively by {ξ ̸= 0 : 2ξ ∈ Sℓ,ℓ(Y )}, {ξ ̸= 0 : 2ξ ∈ Sm,m(Y )}, and {ξ : 2ξ +m− ℓ ∈ Sℓ,m(Y )}. In particular, the zeros of these polynomials are integral. Concerning the scalars ωℓ and ωm appearing in the polynomials ψℓ,ℓe (ξ) and ψ m,m e (ξ), it follows that Proposition 5.4. Recall Assumption 5.2. The scalars ωℓ and ωm satisfies ωm ωℓ = ( n m )( n−2e ℓ−e )( n ℓ )( n−2e m−e ) · |Ym| − ( ne−1) |Yℓ| − ( n e−1 ) . In particular, the weight function ω is unique up to a scalar multiple. Proof. By comparing the diagonal entries of both sides in (5.31), we have 1 = ψℓ,ℓe (0) |Yℓ| = 1 |Yℓ| ( n e− 1 ) + ωℓ ( n−2e ℓ−e )(( n e ) − ( n e−1 )) 2n−2eκe ( n ℓ )  . Likewise, 1 = ψm,me (0) |Ym| = 1 |Ym| ( n e− 1 ) + ωm ( n−2e m−e )(( n e ) − ( n e−1 )) 2n−2eκe ( n m )  . By eliminating κe, we obtain the formula for ωm(ωℓ)−1. The uniqueness of ω follows from this and (5.21). 188 Ars Math. Contemp. 22 (2022) #P2.01 / 163–205 Example 5.5. Suppose that e = 1. In this case, Theorem 5.3(i) was previously obtained by Bannai, Bannai, and Bannai [5, Theorem 2.2 (i)]. Moreover, Theorem 5.3(ii) and Proposi- tion 5.4 are together equivalent to [5, Proposition 4.3]. Example 5.6. Suppose that e = 2. Then we have ψℓ,me (ξ) = 1 + (−ξ)(−2)(1− n) (m− n)(−ℓ) + (−ξ)(1− ξ)(−2)(−1)(1− n)(2− n) (m− n)(m− n+ 1)(−ℓ)(1− ℓ)2 = 1− 2(n− 1)ξ (n−m)ℓ + (n− 1)(n− 2)ξ(ξ − 1) (n−m)(n−m− 1)ℓ(ℓ− 1) . From Example 3.13 we find two parameter sets satisfying Assumption 5.2: n ℓ m ξ 22 6 7 3, 5 22 6 15 1, 3 The zeros ξ given in the last column are indeed integers. Note that the other two parameter sets in Example 3.13 correspond to the complements of these two; cf. Lemma 5.1. On the other hand, the existence of tight relative 4-designs with the following feasible parameter sets was left open in [9, Section 6]: n ℓ m ξ 37 9 16 114 (71± √ 337) 37 9 21 114 (55± √ 337) 41 15 16 126 (237± √ 1569) 41 15 25 126 (153± √ 1569) Here, we are again taking Lemma 5.1 into account. Observe that the zeros ξ are irrational, thus proving the non-existence. We end this section with a comment on the expressions of the polynomials ψℓ,ℓe (ξ) and ψm,me (ξ). We first invoke the following identity which agrees with the formula of the backward shift operator on the dual Hahn polynomials (cf. [31, Section 1.6]): α(N + 1)(α+ β + 2r)Qr(ξ;α− 1, β,N + 1) = (α+ r)(α+ β + r)(N + 1− r)Qr(ξ − 1;α, β,N) − r(α+ β +N + 1 + r)(β + r)Qr−1(ξ − 1;α, β,N). (5.38) This can be routinely verified by writing the LHS as a linear combination of the polynomi- als (1− ξ)i (0 ⩽ i ⩽ r) using (−ξ)i = (1− ξ)i − i(1− ξ)i−1, and then comparing the coefficients of both sides. Setting α = ℓ − n, β = −ℓ − 1, and N = ℓ − 1 in (5.38), it follows that the first term of the RHS in (5.32) is rewritten as follows: e−1∑ r=0 (( n r ) − ( n r − 1 )) 3F2 ( −ξ,−r, r − n− 1 ℓ− n,−ℓ ∣∣∣∣ 1) E. Bannai et al.: Tight relative t-designs on two shells in hypercubes, . . . 189 = e−1∑ r=0 n!(n− 2r + 1) r!(n− r + 1)! Qr(ξ;α− 1, β,N + 1) = n! ℓ(n− ℓ) e−1∑ r=0 ( (ℓ− n+ r)(r − n− 1)(ℓ− r) r!(n− r + 1)! Qr(ξ − 1;α, β,N) −r(r + ℓ− n− 1)(r − ℓ− 1) r!(n− r + 1)! Qr−1(ξ − 1;α, β,N) ) = n! ℓ(n− ℓ) · (−1)(ℓ− n+ e− 1)(ℓ− e+ 1) (e− 1)!(n− e+ 1)! Qe−1(ξ − 1;α, β,N) = ( n e− 1 ) (n− ℓ− e+ 1)(ℓ− e+ 1) ℓ(n− ℓ) 3 F2 ( 1− ξ, 1− e, e− n− 1 ℓ− n+ 1, 1− ℓ ∣∣∣∣ 1) . Likewise, the first term of the RHS in (5.34) is given by e−1∑ r=0 (( n r ) − ( n r − 1 )) 3F2 ( −ξ,−r, r − n− 1 m− n,−m ∣∣∣∣ 1) = ( n e− 1 ) (n−m− e+ 1)(m− e+ 1) m(n−m) 3 F2 ( 1− ξ, 1− e, e− n− 1 m− n+ 1, 1−m ∣∣∣∣ 1) . 6 Zeros of the Hahn and Hermite polynomials Recall the Hahn polynomials Qr(ξ;α, β,N) from (4.11). Recall also that the zeros of orthogonal polynomials are always real and simple; see, e.g., [42, Theorem 3.3.1]. It is well known that we can obtain the Hermite polynomials as limits of the Hahn polynomials; cf. [30, 31]. In this section, we revisit this limit process and describe the limit behavior of the zeros of the Qr(ξ;α, β,N), in a special case which is suited to our purpose. Assumption 6.1. Throughout this section, we assume that α < −N and β < −N , so that the Qr(ξ;α, β,N) satisfy the orthogonality relation (4.12). We consider the following limit: ϵ := − α+ β√ αβN → +0. We write α = αϵ ϵ2 , β = βϵ ϵ2 , N = Nϵ ϵ2 , and assume further that lim ϵ→+0 Nϵ αϵ + βϵ = 0, lim ϵ→+0 βϵ αϵ + βϵ = ρ ∈ [0, 1]. Remark 6.2. We do not require in Assumption 6.1 that αϵ, βϵ, and Nϵ are uniquely deter- mined by ϵ. In other words, these are multi-valued functions of ϵ in general (for admissible values of ϵ), but their limit behaviors are uniformly governed by ϵ. With reference to Assumption 6.1, observe that lim ϵ→+0 αϵ = lim ϵ→+0 αϵ2 = lim ϵ→+0 αϵ + βϵ βϵ · αϵ + βϵ Nϵ = −∞. 190 Ars Math. Contemp. 22 (2022) #P2.01 / 163–205 Likewise, we have lim ϵ→+0 βϵ = −∞, lim ϵ→+0 Nϵ = 1 ρ(1− ρ) ∈ [4,∞]. We will work with the normalized (or monic) Hahn polynomials: qr(ξ) = qr(ξ; ϵ) = (α+ 1)r(−N)r (r + α+ β + 1)r Qr(ξ;α, β,N). (6.1) Their recurrence relation is given by (cf. [31, Section 1.5]) ξqr(ξ) = qr+1(ξ) + (ar + br)qr(ξ) + ar−1brqr−1(ξ), (6.2) where q−1(ξ) := 0, and ar = (r + α+ β + 1)(r + α+ 1)(N − r) (2r + α+ β + 1)(2r + α+ β + 2) , br = r(r + α+ β +N + 1)(r + β) (2r + α+ β)(2r + α+ β + 1) . For convenience, let λϵ = √ 2(αϵ + βϵ +Nϵ) αϵ + βϵ . Note that lim ϵ→+0 λϵ = √ 2. (6.3) Consider the polynomial q̃r(η; ϵ) in the new indeterminate η defined by q̃r(η) = q̃r(η; ϵ) = qr ( λϵη ϵ + αϵNϵ (αϵ + βϵ)ϵ2 ) · ϵ r (λϵ)r ∈ R[η]. Note that q̃r(η) is also monic with degree r in η. Then (6.2) becomes ηq̃r(η) = q̃r+1(η) + 1 λϵ ( (ar + br)ϵ− αϵNϵ (αϵ + βϵ)ϵ ) q̃r(η) + ar−1brϵ 2 (λϵ)2 q̃r−1(η). (6.4) It is a straightforward matter to show that 1 λϵ ( (ar + br)ϵ− αϵNϵ (αϵ + βϵ)ϵ ) = −(µϵ + rσϵ)ϵ+O(ϵ3), (6.5) ar−1brϵ 2 (λϵ)2 = r 2 +O(ϵ2), (6.6) where µϵ := (αϵ − βϵ)Nϵ λϵ(αϵ + βϵ)2 , σϵ := (αϵ − βϵ)(αϵ + βϵ + 2Nϵ) λϵ(αϵ + βϵ)2 are convergent: lim ϵ→+0 µϵ = 0, lim ϵ→+0 σϵ = 1− 2ρ√ 2 . (6.7) E. Bannai et al.: Tight relative t-designs on two shells in hypercubes, . . . 191 Recall the Hermite polynomials [31, Section 1.13] Hr(η) = (2η) r 2F0 ( −r/2,−(r − 1)/2 − ∣∣∣∣− 1η2 ) ∈ R[η] (r = 0, 1, 2, . . .). Their normalized recurrence relation is given by ηhr(η) = hr+1(η) + r 2 hr−1(η), (6.8) where hr(η) = Hr(η) 2r , (6.9) and h−1(η) := 0. We also note that dhr dη (η) = rhr−1(η), (6.10) and that hr(−η) = (−1)rhr(η). (6.11) Since q̃0(η) = h0(η) = 1, it follows from (6.4) – (6.8) that lim ϵ→+0 q̃r(η; ϵ) = hr(η) (6.12) in the sense of coefficient-wise convergence. We now set q̃r(η; 0) = hr(η), and discuss partial derivatives of q̃r(η; ϵ) as a bivariate function of η and ϵ. First, it follows from (6.10) and (6.12) that lim ϵ→+0 ∂q̃r ∂η (η; ϵ) = dhr dη (η) = rhr−1(η). (6.13) Concerning the partial differentiability of q̃r(η; ϵ) with respect to ϵ, it follows that Lemma 6.3. The function q̃r(η; ϵ) is partially right differentiable with respect to ϵ at (η, 0), and we have ∂q̃r ∂ϵ (η; 0) = r(1− 2ρ) 3 √ 2 ( (r − 1 + η2)hr−1(η)− ηhr(η) ) . Proof. Throughout the proof, we fix η ∈ R and set ∆r(ϵ) = ∆r(η; ϵ) = q̃r(η; ϵ)− hr(η) ϵ . It follows from (6.4) – (6.8) and (6.12) that η∆r(ϵ) = ∆r+1(ϵ)− (µϵ + rσϵ)q̃r(η; ϵ) + r 2 ∆r−1(ϵ) +O(ϵ) = ∆r+1(ϵ)− rσ0hr(η) + r 2 ∆r−1(ϵ) + o(1), (6.14) 192 Ars Math. Contemp. 22 (2022) #P2.01 / 163–205 where we set σ0 := lim ϵ→+0 σϵ = 1− 2ρ√ 2 for brevity. Since q̃0(η; ϵ) = 1, we have ∆0(ϵ) = 0. Solving the recurrence (6.14) using this initial condition and (6.8), we routinely obtain ∆r(ϵ) = r(r − 1) 2 σ0hr−1(η) + r(r − 1)(r − 2) 12 σ0hr−3(η) + o(1), where h−1(η) = h−2(η) = h−3(η) := 0. It follows that q̃r(η; ϵ) is partially right differen- tiable with respect to ϵ at (η, 0): ∂q̃r ∂ϵ (η; 0) = lim ϵ→+0 ∆r(ϵ) = r(r − 1) 2 σ0hr−1(η) + r(r − 1)(r − 2) 12 σ0hr−3(η). Finally, from (6.8) it follows that ∂q̃r ∂ϵ (η; 0) = r(r − 1) 2 σ0hr−1(η) + r(r − 1) 6 σ0 ( ηhr−2(η)− hr−1(η) ) = r(r − 1) 3 σ0hr−1(η) + r 3 σ0η ( ηhr−1(η)− hr(η) ) = rσ0 3 ( (r − 1 + η2)hr−1(η)− ηhr(η) ) , as desired. Proposition 6.4. Recall Assumption 6.1. Fix a positive integer e, and let ξ−⌊e/2⌋ < · · · < ξ−1 < (ξ0) < ξ1 < · · · < ξ⌊e/2⌋, η−⌊e/2⌋ < · · · < η−1 < (η0) < η1 < · · · < η⌊e/2⌋ be the zeros of qe(ξ; ϵ) and he(η) from (6.1) and (6.9), respectively, where ξ0 and η0 appear only when e is odd. Then ξi satisfies lim ϵ→+0 ( ξi − λϵηi ϵ − αϵNϵ (αϵ + βϵ)ϵ2 ) = 2ρ− 1 3 ( e− 1 + (ηi)2 ) as a function of ϵ, for i = −⌊e/2⌋, . . . ,−1, (0), 1, . . . , ⌊e/2⌋. Proof. Define τi by ξi = λϵ(ηi + τi) ϵ + αϵNϵ (αϵ + βϵ)ϵ2 , so that ηi + τi is a zero of q̃e(η; ϵ). Then, from (6.12) it follows that lim ϵ→+0 τi = 0. (6.15) For the moment, fix i. Then we have 0 = q̃e(ηi + τi; ϵ) = q̃e(ηi; ϵ) + ∂q̃e ∂η (ηi + θτi; ϵ)τi E. Bannai et al.: Tight relative t-designs on two shells in hypercubes, . . . 193 for some θ ∈ (0, 1) depending on ϵ. Hence, from (6.13), (6.15), Lemma 6.3, and since q̃e(ηi; 0) = he(ηi) = 0, it follows that lim ϵ→+0 τi ϵ = − 1 ehe−1(ηi) lim ϵ→+0 q̃e(ηi; ϵ) ϵ = − 1 ehe−1(ηi) ∂q̃e ∂ϵ (ηi; 0) = 2ρ− 1 3 √ 2 ( e− 1 + (ηi)2 ) , where we note that he(η) and he−1(η) have no common zero by the general theory of orthogonal polynomials; see, e.g., [42, Theorem 3.3.2]. By (6.3), we have lim ϵ→+0 ( ξi − λϵηi ϵ − αϵNϵ (αϵ + βϵ)ϵ2 ) = lim ϵ→+0 λϵτi ϵ = 2ρ− 1 3 ( e− 1 + (ηi)2 ) . This completes the proof. The following is part of the estimates on the zeros of he(η) used in [1].3 Proposition 6.5 ([1, Proposition 13]). Fix a positive integer e, and let the ηi be as in Proposition 6.4. Then η−i = −ηi for all i. Moreover, the following hold: 1. If e is odd and e ⩾ 5, then η0 = 0 and (η1)2 < 3/2. 2. If e is even and e ⩾ 8, then (η2)2 − (η1)2 < 3/2. Proof. That η−i = −ηi is immediate from (6.11). We now write ηi = ηei to compare these zeros for different values of e. Then, as an application of Sturm’s method, it follows that √ 2e+ 1 ηei < √ 2e′ + 1 ηe ′ i (i = 1, 2, . . . , ⌊e′/2⌋), whenever e′ < e and e′ ≡ e (mod 2); see the comments preceding (6.31.19) in [42]. Since h3(η) = η 3 − 3 2 η, h4(η) = η 4 − 3η2 + 3 4 , we have η31 = √ 3 2 , η42 = √ 3 + √ 6 2 . Hence, for odd e ⩾ 5 we have (ηe1) 2 < 7 2e+ 1 (η31) 2 = 21 4e+ 2 < 3 2 , and for even e ⩾ 8 we have (ηe2) 2 − (ηe1)2 < (ηe2)2 < 9 2e+ 1 (η42) 2 = 27 + 9 √ 6 4e+ 2 < 3 2 , as desired. 3Bannai [1] worked with the polynomial √ 2ehe(η/ √ 2). We may remark that the upper bounds √ 3 men- tioned in Proposition 13 (i) and (ii) in [1] should both be 3. See also [22, Proposition 2.4]. 194 Ars Math. Contemp. 22 (2022) #P2.01 / 163–205 7 A finiteness result for tight relative 2e-designs on two shells in Qn In this section, we prove that Theorem 7.1. For any δ ∈ (0, 1/2), there exists e0 = e0(δ) > 0 with the property that, for every given integer e ⩾ e0 and each constant c > 0, there are only finitely many tight relative 2e-designs (Y, ω) (up to scalar multiples of ω) supported on two shells Xℓ ⊔Xm in Qn satisfying Assumption 5.2 such that ℓ < c · nδ. (7.1) Our proof is an application of Bannai’s method from [1]. We will use the following result, which is a variation of [40, Satz I]: Proposition 7.2. For any ϑ > 0 and δ ∈ (0, 1/ϑ), there exists k0 = k0(ϑ, δ) > 0 such that the following holds for every given integer k ⩾ k0 and each constant c > 0: for all but finitely many pairs (a, b) of positive integers with b < c · aδ, the product of k consecutive odd integers (2a+ 1)(2a+ 3) · · · (2a+ 2k− 1) has a prime factor which is greater than 2k + 1 and whose exponent in this product is greater than that in (b+ 1)(b+ 2) · · · (b+ ⌊ϑk⌋). The proof of Proposition 7.2 will be deferred to the appendix. We will establish Theorem 7.1 by contradiction: Assumption 7.3. We fix δ ∈ (0, 1/2). Let k0 = k0(2, δ) > 0 be as in Proposition 7.2 (applied to ϑ = 2), and set e0 = e0(δ) = max{2k0, 8}. We also fix a positive integer e ⩾ e0 and a constant c > 0. Throughout the proof, we assume that there exist infinitely many tight relative 2e-designs (Y, ω) in question. Let Θ denote the set of triples (ℓ,m, n) ∈ N3 taken by those (Y, ω) in Assumption 7.3. Recall from Proposition 5.4 that ω is uniquely determined by Y up to a scalar multiple. Moreover, for each (ℓ,m, n) ∈ Θ there are only finitely many choices for Y . Hence we have |Θ| =∞. (7.2) For the moment, we fix (ℓ,m, n) ∈ Θ and consider the polynomial ψℓ,me (ξ) (which also depends on n) from (5.36). We recall that ψℓ,me (ξ) = Qe(ξ;α, β,N), where α = m− n− 1, β = −m− 1, N = ℓ. (7.3) E. Bannai et al.: Tight relative t-designs on two shells in hypercubes, . . . 195 We note that α, β < −N in view of Assumption 5.2. By Theorem 5.3(ii), if we let ξ−⌊e/2⌋ < · · · < ξ−1 < (ξ0) < ξ1 < · · · < ξ⌊e/2⌋ (7.4) denote the zeros of ψℓ,me (ξ) (cf. Proposition 6.4), then we have ξi ∈ {0, 1, . . . , ℓ} for all i. (7.5) We also rewrite ψℓ,me (ξ) as follows: ψℓ,me (ξ) = e∑ i=0 se−i(−1)i(−ξ)i, where se−i = ( e i ) (e− n− 1)i (m− n)i(−ℓ)i (0 ⩽ i ⩽ e). From (7.5) it follows that the polynomial ψℓ,me (ξ)/s0 is monic and integral: ψℓ,me (ξ) s0 = e∑ i=0 se−i s0 (−1)i(−ξ)i = (ξ − ξ−⌊e/2⌋) · · · (ξ − ξ⌊e/2⌋) ∈ Z[ξ], (7.6) where the factor (ξ − ξ0) appears only when e is odd. Since (−1)i(−ξ)i is also monic and integral, and has degree i for 0 ⩽ i ⩽ e, it follows that si s0 = (−1)i ( e i ) (n−m− e+ 1)i(ℓ− e+ 1)i (n− 2e+ 2)i ∈ Z\{0} (0 ⩽ i ⩽ e), (7.7) where that these coefficients are non-zero follows from Assumption 5.2. We now consider the map f : Θ→ [0, 1]2 defined by f(ℓ,m, n) = ( ℓ n , m n ) ∈ [0, 1]2 ((ℓ,m, n) ∈ Θ). Recall (7.2). Moreover, from (7.1) it follows that |f−1(a, b)| <∞ ((a, b) ∈ [0, 1]2). (7.8) Hence it follows that |f(Θ)| =∞, so that f(Θ) has at least one accumulation point in [0, 1]2. Again by (7.1), such an accu- mulation point must be of the form (0, ρ) ∈ [0, 1]2. We next show that the parameters α, β, and N from (7.3) satisfy Assumption 6.1 when f(ℓ,m, n)→ (0, ρ). Claim 7.4. ℓ,m, n−m→∞ as f(ℓ,m, n)→ (0, ρ). 196 Ars Math. Contemp. 22 (2022) #P2.01 / 163–205 Proof. Since m,n −m ⩾ ℓ by Assumption 5.2, it suffices to show that ℓ → ∞. Suppose the contrary, i.e., that there is a sequence (ℓk,mk, nk) (k ∈ N) of distinct elements of Θ such that lim k→∞ f(ℓk,mk, nk) = (0, ρ), sup k ℓk <∞. Since the ℓk are bounded, it follows from (7.5) and (7.6) that there are only finitely many choices for ψℓ,me (ξ)/s0 when (ℓ,m, n) ranges over this sequence. In particular, there are only finitely many choices for each of the coefficients s1/s0 and s2/s0, and hence the same is true (cf. (7.7)) for each of n−m− e+ 1 n− 2e+ 2 , n−m− e+ 2 n− 2e+ 3 . However, it is immediate to see that these distinct scalars in turn determine n and m uniquely, from which it follows that the nk are bounded, a contradiction. Claim 7.5. ℓm(n−m)/n2 →∞ as f(ℓ,m, n)→ (0, ρ). Proof. If 0 < ρ < 1 then the result follows from Claim 7.4 and since m(n−m) n2 → ρ(1− ρ) > 0. Suppose next that ρ = 1. Suppose moreover that there is a sequence (ℓk,mk, nk) (k ∈ N) of distinct elements of Θ such that lim k→∞ f(ℓk,mk, nk) = (0, 1), sup k ℓkmk(nk −mk) (nk)2 <∞. Since mk/nk → 1, we then have sup k ℓk(nk −mk) nk <∞. Let rk = (nk −mk − e+ 1)(ℓk − e+ 1) nk − 2e+ 2 , tk = (nk −mk − e+ 2)(ℓk − e+ 2) nk − 2e+ 3 . Then the rk and the tk are bounded since rk ≈ tk ≈ ℓk(nk −mk) nk by Claim 7.4. From (7.7) it follows that s1/s0 and s2/s0 are bounded as well, and hence take only finitely many non-zero integral values when (ℓ,m, n) ranges over this sequence. It follows that the rk and the tk can assume only finitely many values, and then since rk ≈ tk we must have rk = tk for sufficiently large k. However, it is again immediate to see that rk ̸= tk for every k ∈ N, and hence this is absurd. It follows that the result holds when ρ = 1. Finally, suppose that ρ = 0. For every (ℓ,m, n) ∈ Θ we have e (m− e+ 1)(ℓ− e+ 1) n− 2e+ 2 = s1 s0 + e(ℓ− e+ 1), E. Bannai et al.: Tight relative t-designs on two shells in hypercubes, . . . 197 ( e 2 ) (m− e+ 1)2(ℓ− e+ 1)2 (n− 2e+ 2)2 = s2 s0 + (e− 1)(ℓ− e+ 2)s1 s0 + ( e 2 ) (ℓ− e+ 1)2. From (7.7) and Assumption 5.2 it follows that these scalars are non-zero integers. By the same argument as above, but working with these two scalars instead of s1/s0 and s2/s0, we conclude that the result holds in this case as well. By Claims 7.4 and 7.5, it follows that the parameters α, β, and N from (7.3) satisfy Assumption 6.1 when f(ℓ,m, n)→ (0, ρ), since − α+ β√ αβN ≈ n√ ℓm(n−m) , N α+ β ≈ − ℓ n , β α+ β ≈ m n . Note that the scalar ρ in Assumption 6.1 agrees with the one used here in this case. Hence we are now in the position to apply the results of the previous section to ψℓ,me (ξ), which is the Hahn polynomial having these parameters. Claim 7.6. We have ρ = 1/2. In particular, (0, 1/2) is a unique accumulation point of f(Θ). Moreover, we have n = 2m for all but finitely many (ℓ,m, n) ∈ Θ. Proof. Let the ξi be as in (7.4). Then from Propositions 6.4 and 6.5 it follows that ξi + ξ−i − ξj − ξ−j → 4ρ− 2 3 ( (ηi) 2 − (ηj)2 ) for all i, j, (7.9) as f(ℓ,m, n)→ (0, ρ), where the ηi are the zeros of the monic Hermite polynomial he(η) from (6.9) as in Proposition 6.4. Recall that e ⩾ 8 by Assumption 7.3. Set (i, j) = (1, 0) in (7.9) if e is odd, and (i, j) = (2, 1) if e is even. Then, since∣∣∣∣4ρ− 23 ∣∣∣∣ ⩽ 23 , it follows from Proposition 6.5 that the RHS in (7.9) lies in the open interval (−1, 1). However, the LHS in (7.9) is always an integer by (7.5), so that this is possible only when the RHS equals zero, i.e., ρ = 1/2. In particular, we have shown that (0, 1/2) is a unique accumulation point of f(Θ). Again by (7.5) and (7.9), we then have ξi + ξ−i = ξj + ξ−j for all i, j, provided that f(ℓ,m, n) is sufficiently close to (0, 1/2). By the uniqueness of the accu- mulation point and (7.8), this last condition on f(ℓ,m, n) can be rephrased as “for all but finitely many (ℓ,m, n) ∈ Θ.” Now, let ξ̃ be the average of the zeros ξi of ψℓ,me (ξ). Then the above identity means that the ξi are symmetric with respect to ξ̃. Hence, if we write ψℓ,me (ξ) s0 = e∑ i=0 we−i(ξ − ξ̃)i, then we have w2i−1 = 0 (1 ⩽ i ⩽ ⌈e/2⌉) 198 Ars Math. Contemp. 22 (2022) #P2.01 / 163–205 for all but finitely many (ℓ,m, n) ∈ Θ. On the other hand, using (7.6) and (7.7), we routinely obtain w3 = ( e 3 ) (n− 2ℓ)(n− 2m) × (ℓ− e+ 1)(m− e+ 1)(n− ℓ− e+ 1)(n−m− e+ 1) (n− 2e+ 2)3(n− 2e+ 3)(n− 2e+ 4) . Hence, by Assumption 5.2, that w3 = 0 forces n = 2m. The claim is proved. By virtue of Claim 7.6, we may now assume without loss of generality that n = 2m ((ℓ,m, n) ∈ Θ), by discarding a finite number of exceptions. Set k = ⌊e 2 ⌋ , and let c′ be a constant such that c′ > 2δc. Note that k ⩾ k0 = k0(2, δ) by Assumption 7.3. Let (ℓ,m, 2m) ∈ Θ. We have c · (2m)δ < c′ · (m− e+ 1)δ provided that m is large. Hence it follows from Proposition 7.2 (applied to ϑ = 2) and (7.1) that if m is sufficiently large then there is a prime p > 2k+ 1 such that νp((2m− 2e+ 3)(2m− 2e+ 5) · · · (2m− 2e+ 2k+ 1)) > νp((ℓ− e+ 1)2k), where νp(n) denotes the exponent of p in n. Assuming that this is the case, let i (1 ⩽ i ⩽ k) be such that νp(2m− 2e+ 2i+ 1) > 0. Observe that i is unique since p > 2k+ 1, so that we have νp(2m− 2e+ 2i+ 1) > νp((ℓ− e+ 1)2k). Moreover, we have gcd(2m− 2e+ 2i+ 1,m− e+ i+ j) = gcd(2j − 1,m− e+ i+ j) < p for 1 ⩽ j ⩽ i, from which it follows that νp((m− e+ i+ 1)i) = 0. By these comments and since 2i ⩽ e < p, it follows from (7.7) (with n = 2m) that νp ( s2i s0 ) = νp ( (m− e+ 1)2i(ℓ− e+ 1)2i (2m− 2e+ 2)2i ) E. Bannai et al.: Tight relative t-designs on two shells in hypercubes, . . . 199 = νp ( (m− e+ i+ 1)i(ℓ− e+ 1)2i 2i(2m− 2e+ 3)(2m− 2e+ 5) · · · (2m− 2e+ 2i+ 1) ) < 0. However, this contradicts the fact that s2i/s0 is a non-zero integer. Hence we now conclude that Θ must be finite. The proof of Theorem 7.1 is complete. ORCID iDs Hajime Tanaka https://orcid.org/0000-0002-5958-0375 Yan Zhu https://orcid.org/0000-0001-7164-9198 References [1] E. Bannai, On tight designs, Q. J. Math. Oxford Ser. (2) 28 (1977), 433–448, doi:10.1093/ qmath/28.4.433. [2] E. Bannai and E. Bannai, Euclidean designs and coherent configurations, in: Combinatorics and Graphs, Amer. Math. Soc., Providence, RI, volume 531 of Contemp. Math., pp. 59–93, 2010, doi:10.1090/conm/531/10457. [3] E. Bannai and E. Bannai, Remarks on the concepts of t-designs, J. Appl. Math. Comput. 40 (2012), 195–207, doi:10.1007/s12190-012-0544-1. [4] E. Bannai and E. Bannai, Tight t-designs on two concentric spheres, Mosc. J. Comb. Number Theory 4 (2014), 52–77, http://mjcnt.phystech.edu/en/article.php?id=78. [5] E. Bannai, E. Bannai and H. Bannai, On the existence of tight relative 2-designs on binary Hamming association schemes, Discrete Math. 314 (2014), 17–37, doi:10.1016/j.disc.2013.09. 013. [6] E. Bannai, E. Bannai, S. Suda and H. Tanaka, On relative t-designs in polynomial association schemes, Electron. J. Comb. 22 (2015), Paper 4.47, 17, doi:10.37236/4889. [7] E. Bannai, E. Bannai, H. Tanaka and Y. Zhu, Design theory from the viewpoint of algebraic combinatorics, Graphs Comb. 33 (2017), 1–41, doi:10.1007/s00373-016-1739-2. [8] E. Bannai, E. Bannai and Y. Zhu, A survey on tight Euclidean t-designs and tight relative t-designs in certain association schemes, Proc. Steklov Inst. Math. 288 (2015), 189–202, doi: 10.1134/s0081543815010149. [9] E. Bannai, E. Bannai and Y. Zhu, Relative t-designs in binary Hamming association scheme H(n, 2), Des. Codes Cryptogr. 84 (2017), 23–53, doi:10.1007/s10623-016-0200-0. [10] E. Bannai and T. Ito, Algebraic Combinatorics. I: Association Schemes, The Ben- jamin/Cummings Publishing Co., Inc., Menlo Park, CA, 1984. [11] E. Bannai and Y. Zhu, Tight t-designs on one shell of Johnson association schemes, Eur. J. Comb. 80 (2019), 23–36, doi:10.1016/j.ejc.2018.02.024. [12] A. E. Brouwer, A. M. Cohen and A. Neumaier, Distance-Regular Graphs, volume 18 of Ergeb- nisse der Mathematik und ihrer Grenzgebiete (3) [Results in Mathematics and Related Areas (3)], Springer-Verlag, Berlin, 1989, doi:10.1007/978-3-642-74341-2. [13] C. J. Colbourn and J. H. Dinitz (eds.), Handbook of Combinatorial Designs, Discrete Mathe- matics and its Applications (Boca Raton), Chapman & Hall/CRC, Boca Raton, FL, 2nd edition, 2007, doi:10.1201/9781420010541. 200 Ars Math. Contemp. 22 (2022) #P2.01 / 163–205 [14] C. W. Curtis and I. Reiner, Methods of Representation Theory. Vol. I, Wiley Classics Library, John Wiley & Sons, Inc., New York, 1990, with applications to finite groups and orders, Reprint of the 1981 original, A Wiley-Interscience Publication, https://books.google.si/ books?id=HWm\_QgAACAAJ. [15] E. R. van Dam, J. H. Koolen and H. Tanaka, Distance-regular graphs, Electron. J. Comb. DS22 (2016), Paper No. DS22, 156, doi:10.37236/4925. [16] P. Delsarte, An algebraic approach to the association schemes of coding theory, Philips Res. Rep. Suppl. (1973), vi+97, https://users.wpi.edu/˜martin/RESEARCH/. [17] P. Delsarte, Association schemes and t-designs in regular semilattices, J. Comb. Theory Ser. A 20 (1976), 230–243, doi:10.1016/0097-3165(76)90017-0. [18] P. Delsarte, Pairs of vectors in the space of an association scheme, Philips Res. Rep. 32 (1977), 373–411. [19] P. Delsarte, Hahn polynomials, discrete harmonics, and t-designs, SIAM J. Appl. Math. 34 (1978), 157–166, doi:10.1137/0134012. [20] P. Delsarte, J. M. Goethals and J. J. Seidel, Spherical codes and designs, Geom. Dedicata 6 (1977), 363–388, doi:10.1007/bf03187604. [21] P. Delsarte and J. J. Seidel, Fisher type inequalities for Euclidean t-designs, Linear Algebra Its Appl. 114/115 (1989), 213–230, doi:10.1016/0024-3795(89)90462-x. [22] P. Dukes and J. Short-Gershman, Nonexistence results for tight block designs, J. Algebraic Comb. 38 (2013), 103–119, doi:10.1007/s10801-012-0395-8. [23] A. L. Gavrilyuk, J. Vidali and J. S. Williford, On few-class Q-polynomial association schemes: feasible parameters and nonexistence results, Ars Math. Contemp. 20 (2021), 103–127, doi: 10.26493/1855-3974.2101.b76. [24] D. Gijswijt, Matrix algebras and semidefinite programming techniques for codes, 2005, arXiv:1007.0906 [math.CO]. [25] D. Gijswijt, A. Schrijver and H. Tanaka, New upper bounds for nonbinary codes based on the Terwilliger algebra and semidefinite programming, J. Comb. Theory Ser. A 113 (2006), 1719– 1731, doi:10.1016/j.jcta.2006.03.010. [26] J. T. Go, The Terwilliger algebra of the hypercube, Eur. J. Comb. 23 (2002), 399–429, doi: 10.1006/eujc.2000.0514. [27] C. D. Godsil, Algebraic Combinatorics, Chapman and Hall Mathematics Series, Chapman & Hall, New York, 1993, https://books.google.si/books?id=eADtlNCkkIMC. [28] S. Kageyama, A property of T -wise balanced designs, Ars Comb. 31 (1991), 237–238. [29] B. G. Kodalen, Cometric association schemes, 2019, arXiv:1905.06959 [math.CO]. [30] R. Koekoek, P. A. Lesky and R. F. Swarttouw, Hypergeometric Orthogonal Polynomials and Their q-Analogues, Springer Monographs in Mathematics, Springer-Verlag, Berlin, 2010, doi: 10.1007/978-3-642-05014-5. [31] R. Koekoek and R. F. Swarttouw, The Askey scheme of hypergeometric orthogonal polynomials and its q-analog, Delft University of Technology, 1998, {http://aw.twi.tudelft.nl/ ˜koekoek/askey.html}. [32] Z. Li, E. Bannai and E. Bannai, Tight relative 2- and 4-designs on binary Hamming association schemes, Graphs Comb. 30 (2014), 203–227, doi:10.1007/s00373-012-1252-1. [33] W. J. Martin and H. Tanaka, Commutative association schemes, Eur. J. Comb. 30 (2009), 1497– 1525, doi:10.1016/j.ejc.2008.11.001. E. Bannai et al.: Tight relative t-designs on two shells in hypercubes, . . . 201 [34] A. Munemasa, An analogue of t-designs in the association schemes of alternating bilinear forms, Graphs Comb. 2 (1986), 259–267, doi:10.1007/bf01788100. [35] A. Neumaier and J. J. Seidel, Discrete measures for spherical designs, eutactic stars and lat- tices, Nederl. Akad. Wetensch. Indag. Math. 50 (1988), 321–334, doi:10.1016/s1385-7258(88) 80011-8. [36] C. Peterson, On tight 6-designs, Osaka Math. J. 14 (1977), 417–435, doi:10.18910/3990. [37] J. B. Rosser and L. Schoenfeld, Approximate formulas for some functions of prime numbers, Illinois J. Math. 6 (1962), 64–94, doi:10.1215/ijm/1255631807. [38] M. Sawa, M. Hirao and S. Kageyama, Euclidean Design Theory, SpringerBriefs in Statistics, Springer, Singapore, 2019, doi:10.1007/978-981-13-8075-4. [39] A. Schrijver, New code upper bounds from the Terwilliger algebra and semidefinite program- ming, IEEE Trans. Inform. Theory 51 (2005), 2859–2866, doi:10.1109/tit.2005.851748. [40] I. Schur, Einige Sätze über Primzahlen: mit Anwendungen auf Irreduzibilitätsfragen, Sitzungs- berichte der Preussischen Akademie der Wissenschaften. Physikalisch-mathematische Klasse, Akademie der Wissenschaften in Kommission bei Walter de Gruyter, 1929, https:// books.google.si/books?id=wE3kAAAAMAAJ. [41] D. Stanton, t-designs in classical association schemes, Graphs Comb. 2 (1986), 283–286, doi: 10.1007/bf01788103. [42] G. Szegő, Orthogonal Polynomials, American Mathematical Society Colloquium Publications, Vol. XXIII, American Mathematical Society, Providence, R.I., 4th edition, 1975, https: //people.math.osu.edu/nevai.1/SZEGO/. [43] Y.-Y. Tan, Y.-Z. Fan, T. Ito and X. Liang, The Terwilliger algebra of the Johnson scheme J(N,D) revisited from the viewpoint of group representations, Eur. J. Comb. 80 (2019), 157– 171, doi:10.1016/j.ejc.2018.02.029. [44] H. Tanaka, New proofs of the Assmus-Mattson theorem based on the Terwilliger algebra, Eur. J. Comb. 30 (2009), 736–746, doi:10.1016/j.ejc.2008.07.018. [45] H. Tanaka, A note on the span of hadamard products of vectors, Linear Algebra Its Appl. 430 (2009), 865–867, doi:10.1016/j.laa.2008.09.010. [46] P. Terwilliger, The subconstituent algebra of an association scheme. I, J. Algebr. Comb. 1 (1992), 363–388, doi:10.1023/a:1022494701663. [47] P. Terwilliger, The subconstituent algebra of an association scheme. II, J. Algebr. Comb. 2 (1993), 73–103, doi:10.1023/a:1022480715311. [48] P. Terwilliger, The subconstituent algebra of an association scheme. III, J. Algebr. Comb. 2 (1993), 177–210, doi:10.1023/a:1022415825656. [49] F. Vallentin, Symmetry in semidefinite programs, Linear Algebra Appl. 430 (2009), 360–369, doi:10.1016/j.laa.2008.07.025. [50] D. R. Woodall, Square λ-linked designs, Proc. London Math. Soc. (3) 20 (1970), 669–687, doi:10.1112/plms/s3-20.4.669. [51] Z. Xiang, A Fisher type inequality for weighted regular t-wise balanced designs, J. Comb. Theory Ser. A 119 (2012), 1523–1527, doi:10.1016/j.jcta.2012.04.008. [52] Z. Xiang, Nonexistence of nontrivial tight 8-designs, J. Algebr. Comb. 47 (2018), 301–318, doi:10.1007/s10801-017-0776-0. [53] H. Yue, B. Hou and S. Gao, Note on the tight relative 2-designs on H(n, 2), Discrete Math. 338 (2015), 196–208, doi:10.1016/j.disc.2014.09.002. 202 Ars Math. Contemp. 22 (2022) #P2.01 / 163–205 [54] Y. Zhu, E. Bannai and E. Bannai, Tight relative 2-designs on two shells in Johnson association schemes, Discrete Math. 339 (2016), 957–973, doi:10.1016/j.disc.2015.10.024. E. Bannai et al.: Tight relative t-designs on two shells in hypercubes, . . . 203 Appendix A Proof of Proposition 7.2 Our proof of Proposition 7.2 is a slight modification of (the first part of) that of [40, Satz I]. For a positive integer n, let χn = 1 · 3 · 5 · · · (2n− 1) = (2n)! 2nn! . Observe that the exponent νp(χn) of an odd prime p in χn is given by νp(χn) = ⌊logp(2n)⌋∑ i=1 (⌊ 2n pi ⌋ − ⌊ n pi ⌋) = ⌊logp(2n)⌋∑ i=1 ⌊ n pi + 1 2 ⌋ , (A.1) where we have used ⌊ξ⌋+ ⌊ ξ + 1 2 ⌋ = ⌊2ξ⌋ (ξ ∈ R). Now, let (a, b) be a pair of positive integers with b < c · aδ, (A.2) which does not satisfy the desired property about a prime factor; in other words, νp(χa+k)− νp(χa) ⩽ νp((b+ ⌊ϑk⌋)!)− νp(b!) if p > 2k+ 1. (A.3) Our aim is to show that a is bounded in terms of ϑ, δ, c, and k, and hence so is b by (A.2), from which it follows that there are only finitely many such pairs. (We will specify k0 = k0(ϑ, δ) at the end of the proof.) To this end, we may assume for example that a > k, c · aδ > k+ 1. (A.4) Without loss of generality, we may also assume that b > k, (A.5) for otherwise the pair (a, k+ 1) would also satisfy (A.2) and (A.3). Let s = χa+k χkχa · b! (b+ ⌊ϑk⌋)! . Then from (A.3) it follows that s ⩽ ∏ 3⩽p⩽2k+1 pνp(s), (A.6) where the product in the RHS is over the odd primes p ⩽ 2k+ 1, and where νp(s) = νp(χa+k)− νp(χk)− νp(χa) + νp(b!)− νp((b+ ⌊ϑk⌋)!). By (A.1), for every odd prime p we have νp(s) ⩽ νp(χa+k)− νp(χk)− νp(χa) 204 Ars Math. Contemp. 22 (2022) #P2.01 / 163–205 = ⌊logp(2a+2k)⌋∑ i=1 (⌊ a+ k pi + 1 2 ⌋ − ⌊ k pi + 1 2 ⌋ − ⌊ a pi + 1 2 ⌋) . Note that ⌊ ξ + η + 1 2 ⌋ − ⌊ ξ + 1 2 ⌋ − ⌊ η + 1 2 ⌋ ∈ {−1, 0, 1} (ξ, η ∈ R). Hence it follows that νp(s) ⩽ logp(2a+ 2k) = ln(2a+ 2k) ln p (A.7) for every odd prime p. From (A.6) and (A.7) it follows that ln s ⩽ (π(2k+ 1)− 1) ln(2a+ 2k), (A.8) where π(n) denotes the number of primes at most n. On the other hand, we have s = (2a+ 2k)!k!a! (a+ k)!(2k)!(2a)! · b! (b+ ⌊ϑk⌋)! . Using Stirling’s formula ln(n!) = ( n+ 1 2 ) ln n− n+ ln 2π 2 + rn, where 0 < rn < 1 12n , we obtain ln s > (a+ k) ln(a+ k)− k ln k− a ln a+ ϑk− 2 + ( b+ 1 2 ) ln b− ( b+ ϑk+ 1 2 ) ln(b+ ϑk). (A.9) Let ã = a k , b̃ = b k . Note that ã, b̃ > 1, (A.10) in view of (A.4) and (A.5). With this notation, we have ln s > k ln ã+ (ã+ 1)k ln ( 1 + 1 ã ) − ϑk ln k+ ϑk− 2 − ϑk ln b̃− ( (b̃+ ϑ)k+ 1 2 ) ln ( 1 + ϑ b̃ ) > k ln ã− ϑk ln k− 2− ϑk ln b̃− ( ϑk+ 1 2 ) ln ( 1 + ϑ b̃ ) E. Bannai et al.: Tight relative t-designs on two shells in hypercubes, . . . 205 > (1− ϑδ)k ln ã− ϑk ln c− ϑδk ln k− 2− ( ϑk+ 1 2 ) ln(1 + ϑ), (A.11) where the first inequality is a restatement of (A.9), the second follows from 0 < ln(1 + ξ) < ξ (ξ > 0), and the last one follows from (A.2) and (A.10). Concerning the prime-counting function π(n), it is known that [37, (3.6)] π(n) < 1.25506 n ln n (n > 1). By this, (A.8), and (A.10), we have ln s ⩽ (π(2k+ 1)− 1) ( ln ã+ ln 2k ( 1 + 1 ã )) < ( 1.25506 2k+ 1 ln(2k+ 1) − 1 ) (ln ã+ ln 4k). (A.12) Combining (A.11) and (A.12), it follows that( (1− ϑδ)k− 1.25506 2k+ 1 ln(2k+ 1) + 1 ) ln ã < ϑk ln c+ ϑδk ln k+ 2 + ( ϑk+ 1 2 ) ln(1 + ϑ) + ( 1.25506 2k+ 1 ln(2k+ 1) − 1 ) ln 4k. (A.13) If we set k0 = k0(ϑ, δ) = 1 2 ( exp ( 2.51012 1− ϑδ ) − 1 ) > 0 for example, then we have (1− ϑδ)k− 1.25506 2k+ 1 ln(2k+ 1) + 1 ⩾ 1 + ϑδ 2 > 0 (k ⩾ k0). Hence, whenever k ⩾ k0, it follows from (A.13) that ln a = ln ã+ ln k is bounded in terms of ϑ, δ, c, and k, from which and (A.2) it follows that there are only finitely many choices for the pairs (a, b). This completes the proof of Proposition 7.2.