ISSN 2590-9770 The Art of Discrete and Applied Mathematics 3 (2020) #P2.04 https://doi.org/10.26493/2590-9770.1271.e54 (Also available at http://adam-journal.eu) On the Terwilliger algebra of a certain family of bipartite distance-regular graphs with ∆2 = 0 Štefko Miklavič* , Safet Penjić† University of Primorska, Andrej Marušič Institute, Muzejski trg 2, 6000 Koper, Slovenia Received 27 September 2018, accepted 4 January 2019, published online 10 August 2020 Abstract Let Γ denote a bipartite distance-regular graph with diameterD ≥ 4 and valency k ≥ 3. Let X denote the vertex set of Γ, and let Ai (0 ≤ i ≤ D) denote the distance matrices of Γ. We abbreviate A := A1. For x ∈ X and for 0 ≤ i ≤ D, let Γi(x) denote the set of vertices in X that are distance i from vertex x. Fix x ∈ X and let T = T (x) denote the subalgebra of MatX(C) generated by A,E∗0 , E ∗ 1 , . . . , E ∗ D, where for 0 ≤ i ≤ D, E∗i represents the projection onto the ith subconstituent of Γ with respect to x. We refer to T as the Terwilliger algebra of Γ with respect to x. By the endpoint of an irreducible T -module W we mean min{i | E∗iW 6= 0}. In this paper we assume Γ has the property that for 2 ≤ i ≤ D− 1, there exist complex scalars αi, βi such that for all y, z ∈ X with ∂(x, y) = 2, ∂(x, z) = i, ∂(y, z) = i, we have αi + βi|Γ1(x) ∩ Γ1(y) ∩ Γi−1(z)| = |Γi−1(x) ∩ Γi−1(y) ∩ Γ1(z)|. We study the structure of irreducible T -modules of endpoint 2. Let W denote an irre- ducible T -module with endpoint 2, and let v denote a nonzero vector in E∗2W . We show that W = span ( {E∗i Ai−2E∗2v | 2 ≤ i ≤ D} ∪ {E∗i Ai+2E∗2v | 2 ≤ i ≤ D − 2} ) . It turns out that, except for a particular family of bipartite distance-regular graphs with D = 5, this result is already known in the literature. Assume now that Γ is a member of this particular family of graphs. We show that if Γ is not almost 2-homogeneous, then up to isomorphism there exists exactly one irreducible T -module with endpoint 2 and it is not thin. We give a basis for this T -module. Keywords: Distance-regular graphs, Terwilliger algebra, irreducible modules. Math. Subj. Class. (2020): 05E30, 05C50 *The author acknowledge the financial support from the Slovenian Research Agency (research core funding No. P1-0285 and research projects N1-0032, N1-0038, N1-0062, J1-5433, J1-6720, J1-7051, J1-9108, J1-9110). †The author acknowledges the financial support from the Slovenian Research Agency (research core funding No. P1-0285 and Young Researchers Grant). E-mail addresses: stefko.miklavic@upr.si (Štefko Miklavič), safet.penjic@iam.upr.si (Safet Penjić) cb This work is licensed under https://creativecommons.org/licenses/by/4.0/ 2 Art Discrete Appl. Math. 3 (2020) #P2.04 1 Introduction Throughout this introduction let Γ denote a bipartite distance-regular graph with diameter D ≥ 4, valency k ≥ 3 and path-length function ∂. Let X denote the vertex set of Γ. For x ∈ X and 0 ≤ i ≤ D, let Γi(x) denote the set of vertices in X that are distance i from vertex x, and let T = T (x) denote the Terwilliger algebra of Γ with respect to x (see Section 2 for formal definitions). It is known that there exists a unique irreducible T -module with endpoint 0, and this module is thin [8, Proposition 8.4]. Moreover, Curtin showed that up to isomorphism Γ has exactly one irreducible T -module with endpoint 1, and this module is thin [4, Corol- lary 7.7]. We now discuss the irreducible T -modules of endpoint 2. It turns out that the structure of these modules is particularly nice if we assume that Γ has the following combinatorial property: for 2 ≤ i ≤ D − 1, there exist complex scalars αi, βi such that for all y, z ∈ X with ∂(x, y) = 2, ∂(x, z) = i, ∂(y, z) = i, we have αi + βi|Γ1(x) ∩ Γ1(y) ∩ Γi−1(z)| = |Γi−1(x) ∩ Γi−1(y) ∩ Γ1(z)|. Irreducible modules of endpoint 2 of these graphs were studied extensively, see [10, 11, 12, 13, 15]. We are motivated by the fact that the above equation holds if Γ is Q-polynomial. Assume that Γ has the above mentioned combinatorial property. We show that if W is an irreducible T -module with endpoint 2 and v is a nonzero vector in E∗2W , then W = span ( {E∗i Ai−2E∗2v | 2 ≤ i ≤ D} ∪ {E∗i Ai+2E∗2v | 2 ≤ i ≤ D − 2} ) . Except for a particular family of bipartite distance-regular graphs with D = 5, this result is already known in the literature. To define this particular family we introduce a certain parameter ∆2 in terms of the intersection numbers of Γ by ∆2 = (k− 2)(c3− 1)− (c2 − 1)p222. It turns out that ∆2 ≥ 0 and that ∆2 = 0 implies c2 ∈ {1, 2} or D ≤ 5. The above mentioned family of bipartite distance-regular graphs with D = 5 is exactly the family of such graphs with ∆2 = 0. Assume now that Γ is such a graph. We show that if Γ is not almost 2-homogeneous, then up to isomorphism there exists exactly one irreducible T -module with endpoint 2, and this module is not thin. We give a basis for this T -module. If Γ is almost 2-homogeneous, then the structure of irreducible T -modules with endpoint 2 is described in [7]. 2 Preliminaries In this section we review some definitions and basic results concerning distance-regular graphs. See the book of A. E. Brouwer, A. M. Cohen and A. Neumaier [2] for more background information. Let C denote the complex number field and let X denote a nonempty finite set. Let MatX(C) denote the C-algebra consisting of all matrices whose rows and columns are indexed by X and whose entries are in C. Let V = CX denote the vector space over C consisting of column vectors whose coordinates are indexed by X and whose entries are in C. We observe MatX(C) acts on V by left multiplication. We call V the standard module. We endow V with the Hermitean inner product 〈 , 〉 that satisfies 〈u, v〉 = utv for u, v ∈ V , where t denotes transpose and denotes complex conjugation. Recall that 〈u,Bv〉 = 〈Btu, v〉 (2.1) Š. Miklavič and S. Penjić: On the Terwilliger algebra of BDRG with ∆2 = 0 3 for u, v ∈ V and B ∈ MatX(C). For y ∈ X let ŷ denote the element of V with a 1 in the y coordinate and 0 in all other coordinates. Note that {ŷ | y ∈ X} is an orthonormal basis for V. Let Γ = (X,R) denote a finite, undirected, connected graph, without loops or multiple edges, with vertex set X and edge set R. Let ∂ denote the path-length distance function for Γ, and set D := max{∂(x, y) | x, y ∈ X}. We call D the diameter of Γ. For a vertex x ∈ X and an integer i let Γi(x) denote the set of vertices at distance i from x. For an integer k ≥ 0 we say Γ is regular with valency k whenever |Γ1(x)| = k for all x ∈ X . We say Γ is distance-regular whenever for all integers h, i, j (0 ≤ h, i, j ≤ D) and for all vertices x, y ∈ X with ∂(x, y) = h, the number phij = |Γi(x) ∩ Γj(y)| is independent of x and y. The phij are called the intersection numbers of Γ. For the rest of this paper we assume Γ is distance-regular with diameter D ≥ 4. Note that phij = p h ji for 0 ≤ h, i, j ≤ D. For convenience set ci := pi1,i−1 (1 ≤ i ≤ D), ai := p i 1i (0 ≤ i ≤ D), bi := pi1,i+1 (0 ≤ i ≤ D − 1), ki := p0ii (0 ≤ i ≤ D), and c0 = bD = 0. By the triangle inequality the following hold for 0 ≤ h, i, j ≤ D: (i) phij = 0 if one of h, i, j is greater than the sum of the other two; (ii) phij 6= 0 if one of h, i, j equals the sum of the other two. In particular ci 6= 0 for 1 ≤ i ≤ D and bi 6= 0 for 0 ≤ i ≤ D−1. We observe that Γ is regular with valency k = k1 = b0 and that ci + ai + bi = k (0 ≤ i ≤ D). (2.2) Note that ki = |Γi(x)| for x ∈ X and 0 ≤ i ≤ D. By [2, p. 127], ki = b0b1 · · · bi−1 c1c2 · · · ci (1 ≤ i ≤ D). (2.3) Recall Γ is bipartite whenever ai = 0 for 0 ≤ i ≤ D. Setting ai = 0 in (2.2) we find bi + ci = k (0 ≤ i ≤ D). (2.4) The following formulae for the bipartite case will be useful. Lemma 2.1 ([2, Lemma 4.1.7]). Let Γ denote a bipartite distance-regular graph with diameter D ≥ 4 and valency k ≥ 3. Then pi2i = ci(bi−1 − 1) + bi(ci+1 − 1) c2 (1 ≤ i ≤ D − 1), pD2D = k(bD−1 − 1) c2 . We recall the Bose-Mesner algebra of Γ. For 0 ≤ i ≤ D let Ai denote the matrix in MatX(C) with (x, y)-entry (Ai)xy = { 1 if ∂(x, y) = i, 0 if ∂(x, y) 6= i (x, y ∈ X). (2.5) For notational convenience, we define Ai to be the zero matrix for all integers i < 0 or i > D. We call Ai the ith distance matrix of Γ. We abbreviate A := A1 and call this the adjacency matrix of Γ. We observe (i) A0 = I; (ii) ∑D i=0Ai = J ; (iii) Ai = Ai (0 ≤ i ≤ D); (iv) Ati = Ai (0 ≤ i ≤ D); (v) AiAj = ∑D h=0 p h ijAh (0 ≤ i, j ≤ D), where I (resp. J) denotes the identity matrix (resp. all 1’s matrix) in MatX(C). Using these facts we find A0, A1, . . . , AD is a basis for a commutative subalgebra M of MatX(C). We call M the Bose-Mesner algebra of Γ. It turns out that A generates M [1, p. 190]. 4 Art Discrete Appl. Math. 3 (2020) #P2.04 3 Terwilliger algebra Let Γ denote a distance-regular with diameter D ≥ 4 and valency k ≥ 3. We first recall the dual idempotents of Γ. To do this fix a vertex x ∈ X. We view x as a “base vertex”. For 0 ≤ i ≤ D let E∗i = E∗i (x) denote the diagonal matrix in MatX(C) with (y, y)-entry (E∗i )yy = { 1 if ∂(x, y) = i, 0 if ∂(x, y) 6= i (y ∈ X). We call E∗i the ith dual idempotent of Γ with respect to x [16, p. 378]. We observe (ei)∑D i=0E ∗ i = I; (eii) E ∗ i = E ∗ i (0 ≤ i ≤ D); (eiii) E∗ti = E∗i (0 ≤ i ≤ D); (eiv) E∗i E ∗ j = δijE ∗ i (0 ≤ i, j ≤ D). By these facts E∗0 , E∗1 , . . . , E∗D form a basis for a commutative subalgebra M∗ = M∗(x) of MatX(C). We call M∗ the dual Bose-Mesner algebra of Γ with respect to x [16, p. 378]. For 0 ≤ i ≤ D we have E∗i V = span{ŷ | y ∈ X, ∂(x, y) = i}, so dimE∗i V = ki. We call E ∗ i V the ith subconstituent of Γ with respect to x. Note that V = E∗0V + E ∗ 1V + · · ·+ E∗DV (orthogonal direct sum). (3.1) Moreover E∗i is the projection from V onto E ∗ i V for 0 ≤ i ≤ D. We now recall the Terwilliger algebra of Γ. Let T = T (x) denote the subalgebra of MatX(C) generated by M , M∗. We call T the Terwilliger algebra of Γ with respect to x [16, Definition 3.3]. Recall M is generated by A, so T is generated by A and the dual idempotents. We observe T has finite dimension. By construction T is closed under the conjugate-transpose map so T is semisimple [16, Lemma 3.4(i)]. By a T -module we mean a subspace W of V such that BW ⊆ W for all B ∈ T . Let W denote a T -module. Then W is said to be irreducible whenever W is nonzero and W contains no T -modules other than 0 and W . By [9, Corollary 6.2] any T -module is an orthogonal direct sum of irreducible T - modules. In particular the standard module V is an orthogonal direct sum of irreducible T -modules. Let W , W ′ denote T -modules. By an isomorphism of T -modules from W to W ′ we mean an isomorphism of vector spaces σ : W →W ′ such that (σB −Bσ)W = 0 for all B ∈ T . The T -modules W , W ′ are said to be isomorphic whenever there exists an isomorphism of T -modules from W to W ′. By [4, Lemma 3.3] any two nonisomor- phic irreducible T -modules are orthogonal. Let W denote an irreducible T -module. By [16, Lemma 3.4(iii)] W is an orthogonal direct sum of the nonvanishing spaces among E∗0W,E ∗ 1W, . . . , E ∗ DW . By the endpoint ofW we mean min{i | 0 ≤ i ≤ D, E∗iW 6= 0}. By the diameter of W we mean |{i | 0 ≤ i ≤ D, E∗iW 6= 0}| − 1. We say W is thin whenever the dimension of E∗iW is at most 1 for 0 ≤ i ≤ D. The following matrices of MatX(C) will be useful later in the paper. Definition 3.1. Let Γ denote a distance-regular with diameter D ≥ 4 and valency k ≥ 3. Fix x ∈ X and let E∗i = E∗i (x) (0 ≤ i ≤ D) and T = T (x). We define matrices L = L(x), R = R(x) by L = D∑ h=1 E∗h−1AE ∗ h, R = D−1∑ h=0 E∗h+1AE ∗ h. Š. Miklavič and S. Penjić: On the Terwilliger algebra of BDRG with ∆2 = 0 5 Note thatA = L+R [4, Lemma 4.4] andLt = R. We callL andR the lowering matrix and the raising matrix of Γ with respect to x, respectively. Observe that L and R are contained in T . Definition 3.2 ([7, Definition 3.2]). Let Γ denote a distance-regular with diameter D ≥ 4 and valency k ≥ 3. Fix x ∈ X . For 1 ≤ i ≤ D we define matrices Λi = Λi(x) in MatX(C) by (Λi)zy = { |Γ1(x) ∩ Γ1(y) ∩ Γi−1(z)|, if ∂(x, y) = 2, ∂(x, z) = ∂(y, z) = i, 0, otherwise for z, y ∈ X . 4 The scalars ∆i and γi Let Γ denote a distance-regular graph with diameter D ≥ 4 and valency k ≥ 3. From now on we assume that Γ is bipartite. In this section we introduce certain scalars ∆i and γi (2 ≤ i ≤ D − 1) which we find useful. Definition 4.1. Let Γ denote a distance-regular with diameter D ≥ 4 and valency k ≥ 3. Then for 2 ≤ i ≤ D − 1 we define ∆i = (bi−1 − 1)(ci+1 − 1)− (c2 − 1)pi2i and γi = ci(bi−1 − 1) pi2i (observe that pi2i > 0 by [3, Lemma 11]). By [3, Theorem 12] we have ∆i ≥ 0 for 2 ≤ i ≤ D − 1. Moreover, the scalars ∆i and γi are related as follows. Lemma 4.2 ([3, Theorem 13]). Let Γ denote a distance-regular with diameter D ≥ 4 and valency k ≥ 3 and fix an integer 2 ≤ i ≤ D − 1. Then the following (i),(ii) are equivalent. (i) ∆i = 0. (ii) For all x, y, z ∈ X with ∂(x, y) = 2, ∂(x, z) = i, ∂(y, z) = i, |Γ1(x) ∩ Γ1(y) ∩ Γi−1(z)| = γi. If ∆i = 0 for 2 ≤ i ≤ D − 2, then Γ is called almost 2-homogeneous, see [7]. In this case the structure of irreducible T -modules is well understood, so we will assume that Γ is not almost 2-homogeneous. In the rest of the paper we therefore consider the following situation. Notation 4.3. Let Γ = (X,R) denote a bipartite distance-regular graph with diameter D ≥ 4, valency k ≥ 3 and intersection numbers bi, ci, which is not almost 2-homogeneous. Let Ai (0 ≤ i ≤ D) be the distance matrices of Γ, and let V denote the standard module for Γ. We fix x ∈ X and let E∗i = E∗i (x) (0 ≤ i ≤ D) and T = T (x) denote the dual idempotents and the Terwilliger algebra of Γ with respect to x, respectively. We assume 6 Art Discrete Appl. Math. 3 (2020) #P2.04 that for 2 ≤ i ≤ D − 1, there exist complex scalars αi, βi such that for all y, z ∈ X with ∂(x, y) = 2, ∂(x, z) = i, ∂(y, z) = i, we have αi + βi|Γ1(x) ∩ Γ1(y) ∩ Γi−1(z)| = |Γi−1(x) ∩ Γi−1(y) ∩ Γ1(z)|. Let matrices L = L(x), R = R(x) and Λi = Λi(x) (1 ≤ i ≤ D) be as in Definitions 3.1 and 3.2. Let scalars ∆i, γi (2 ≤ i ≤ D − 1) be as in Definition 4.1. With reference to Notation 4.3, pick 2 ≤ i ≤ D − 1 and assume that ∆i 6= 0. By [12, Theorem 5.4] scalars αi and βi are uniquely determined and given by αi = ci(ci − 1)(bi−1 − c2)− cici−1(bi − 1)(c2 − 1) c2∆i , βi = ci(ci+1 − ci)(bi−1 − 1)− bi(ci+1 − 1)(ci − ci−1) c2∆i . (4.1) If ∆i = 0, then scalars αi and βi are not uniquely determined. For example, if ∆2 = 0, then one of the possible values for α2 and β2 is α2 = 0, β2 = 1. Note however that by Lemma 4.2 this is not the only possible solution. 5 Some products in T With reference to Notation 4.3, in this section we compute some products of matrices of T . We start by recalling the following results. Lemma 5.1 ([14, Lemma 6.1]). With reference to Notation 4.3, for 0 ≤ h, i, j ≤ D and y, z ∈ X the (y, z)-entry of E∗hAiE∗j is 1 if ∂(x, y) = h, ∂(y, z) = i, ∂(x, z) = j, and 0 otherwise. Lemma 5.2 ([14, Lemma 6.5]). With reference to Notation 4.3, for 0 ≤ h, i, j, r, s ≤ D and y, z ∈ X the (y, z)-entry of E∗hArE∗i AsE∗j is |Γi(x) ∩ Γr(y) ∩ Γs(z)| if ∂(x, y) = h, ∂(x, z) = j, and 0 otherwise. Lemma 5.3 ([7, Lemma 3.3]). With reference to Notation 4.3, we have Λ1 = E ∗ 1AE ∗ 2 , Λi = E ∗ i Ai−1E ∗ 1AE ∗ 2 − c2E∗i Ai−2E∗2 (2 ≤ i ≤ D). In particular, Λi ∈ T (1 ≤ i ≤ D). Theorem 5.4. With reference to Notation 4.3 the following holds for 3 ≤ i ≤ D: LE∗i Ai−2E ∗ 2 = bi−1E ∗ i−1Ai−3E ∗ 2 + (ci−1 − αi−1)E∗i−1Ai−1E∗2 − βi−1Λi−1. (5.1) Proof. Pick z, y ∈ X and an integer 3 ≤ i ≤ D. We show that (z, y)-entries of both sides of (5.1) agree. Note that by the property (eiv) of Section 3 and Lemma 5.2, (LE∗i Ai−2E ∗ 2 )zy = { |Γi(x) ∩ Γi−2(y) ∩ Γ1(z)| if ∂(x, y) = 2, ∂(x, z) = i− 1, 0 otherwise. (5.2) It follows from (5.2), Lemma 5.1 and Definition 3.2 that the (z, y)-entries of both sides of (5.1) are 0 if ∂(x, y) 6= 2 or ∂(x, z) 6= i−1. Assume now ∂(x, y) = 2 and ∂(x, z) = i−1. Š. Miklavič and S. Penjić: On the Terwilliger algebra of BDRG with ∆2 = 0 7 Observe that by the triangle inequality we have that ∂(z, y) ∈ {i − 3, i − 1, i + 1}. We consider each of these three cases separately. Case 1: ∂(x, y) = 2, ∂(x, z) = i−1 and ∂(z, y) = i−3. Note that in this case we have (LE∗i Ai−2E ∗ 2 )zy = bi−1 by (5.2). By Lemma 5.1 and Definition 3.2 the (z, y)-entries of both sides of (5.1) agree. Case 2: ∂(x, y) = 2, ∂(x, z) = i − 1 and ∂(z, y) = i − 1. Observe that by (5.2) we have (LE∗i Ai−2E ∗ 2 )zy = ci−1 − |Γ1(z) ∩ Γi−2(x) ∩ Γi−2(y)| = ci−1 − (αi−1 + βi−1|Γi−2(z) ∩ Γ1(x) ∩ Γ1(y)|). By Lemma 5.1 and Definition 3.2 the (z, y)-entries of both sides of (5.1) agree. Case 3: ∂(x, y) = 2, ∂(x, z) = i − 1 and ∂(z, y) = i + 1. By (5.2), Lemma 5.1 and Definition 3.2 the (z, y)-entries of both sides of (5.1) are 0. 6 Irreducible T -modules with endpoint 2 With reference to Notation 4.3, let W denote an irreducible T -module with endpoint 2. In this section we find a spanning set for W . Definition 6.1. With reference to Notation 4.3, letW denote an irreducible T -module with endpoint 2 and let v denote a nonzero vector in E∗2W . For 0 ≤ i ≤ D, define v+i = E ∗ i Ai−2E ∗ 2v, v − i = E ∗ i Ai+2E ∗ 2v. Note that v+2 = v, v + i = 0 if i < 2, and v − i = 0 if i < 2 or i > D − 2. Lemma 6.2 ([5, Corollary 9.3(i), Theorem 9.4]). With reference to Definition 6.1, the following (i)–(iv) hold. (i) E∗i AiE ∗ 2v = −(v+i + v − i ) (2 ≤ i ≤ D). (ii) Rv+i = ci−1v + i+1 (2 ≤ i ≤ D − 1) and Rv + D = 0. (iii) Lv−i = bi+1v − i−1 (2 ≤ i ≤ D − 2). (iv) Lv+i+1 −Rv − i−1 = biv + i − civ − i (1 ≤ i ≤ D − 1). Lemma 6.3. With reference to Definition 6.1, the following (i)–(iii) hold. (i) Λiv = −c2v+i (2 ≤ i ≤ D). (ii) Lv+2 = 0 and Lv+i = (bi−1 − ci−1 + αi−1 + c2βi−1)v + i−1 − (ci−1 − αi−1)v − i−1 for 3 ≤ i ≤ D. (iii) Rv−i = (c2βi+1 − ci+1 + αi+1)v + i+1 + αi+1v − i+1 for 2 ≤ i ≤ D − 2. Proof. (i) Immediate from Lemma 5.3 and Definition 6.1. (ii) Note that Lv+2 = 0 as the endpoint of W is 2. To obtain the result for Lv + i (3 ≤ i ≤ D) apply (5.1) to v and use Definition 6.1, Lemma 6.2(i) and (i) above. (iii) Immediately by (ii) above and Lemma 6.2(iv). 8 Art Discrete Appl. Math. 3 (2020) #P2.04 Theorem 6.4. With reference to Definition 6.1, W = span{v+2 , v + 3 , . . . , v + D, v − 2 , v − 3 , . . . , v − D−2}. Proof. Denote W ′ = span{v+2 , v + 3 , . . . , v + D, v − 2 , v − 3 , . . . , v − D−2} and note that W ′ ⊆ W . We now show that W = W ′. Note that E∗i v + j = δijv + j for 2 ≤ j ≤ D and E∗i v − j = δijv − j for 2 ≤ j ≤ D − 2. Therefore, W ′ is invariant under the action of E∗i for 0 ≤ i ≤ D. Ob- serve also that W ′ is invariant under the action of L by Lemma 6.2(iii) and Lemma 6.3(ii), and also invariant under the action of R by Lemma 6.2(ii) and Lemma 6.3(iii). As A = R+L, W ′ is invariant under the action ofA. As T is generated byA andE∗i (0 ≤ i ≤ D), this implies that W ′ is a T -module. Recall that W is irreducible and that W ′ contains a nonzero vector v. It follows that W = W ′. Corollary 6.5. With reference to Definition 6.1, we have dim (E∗D−1W ) ≤ 1, dim (E∗DW ) ≤ 1. Proof. Immediately from Theorem 6.4. As already mentioned, the result from Theorem 6.4 is already known in the literature, except for the case D = 5 and ∆2 = 0, see [11, 12, 15]. In the rest of the paper we study this case in detail. If D = 5 and ∆2 = ∆3 = 0, then Γ is almost 2-homogeneous, contradicting our assumption in Notation 4.3. Therefore, we have that ∆3 6= 0. 7 Case ∆2 = 0 and ∆3 6= 0 With reference to Notation 4.3, in this section we study graphs with ∆2 = 0 and ∆3 6= 0. We first have the following observation. Lemma 7.1. With reference to Definition 6.1, assume that ∆2 = 0 and ∆3 6= 0. Then the following (i), (ii) hold. (i) c3 = (c22 − c2 + 1)k − c2(c2 + 1) k + c22 − 3c2 . (ii) α3 = 0, β3 = c2(k − 2) k + c22 − 3c2 . Proof. (i) Solve ∆2 = 0 for c3. Note that k+ c22 − 3c2 = (c2 − 1)(c2 − 2) + k− 2 > 0 as k ≥ 3. (ii) Use Definition 4.1, (4.1) and (i) above. Lemma 7.2. With reference to Definition 6.1, assume that ∆2 = 0 and ∆3 6= 0. Then E∗2A2E ∗ 2v = − c2(k − 2) k + c22 − 3c2 v. Š. Miklavič and S. Penjić: On the Terwilliger algebra of BDRG with ∆2 = 0 9 Proof. Let Γ22 = Γ 2 2(x) denote the graph with vertex set X̃ = Γ2(x) and edge set R̃ = {yz | y, z ∈ X̃, ∂(y, z) = 2}. The graph Γ22 has exactly k2 vertices and it is regular with valency p222 ([6, Lemma 3.2]). Let à denote the adjacency matrix of Γ 2 2. The matrix à is symmetric with real entries. Therefore à is diagonalizable with all eigenvalues real. Note that eigenvalues for E∗2A2E ∗ 2 and à are the same. Since ∆2 = 0, we know E∗2A2E ∗ 2 has exactly one distinct eigenvalue η on E ∗ 2W by [6, Theorem 4.11, Corollary 4.13, Lemma 5.3]. Thus, every nonzero vector in E∗2W is an eigenvector for E∗2A2E ∗ 2 with eigenvalue η. By [6, Lemmas 5.4, 5.5] we find η = − c2γ2 . The result now follows from Definition 4.1 and Lemma 7.1(i). Corollary 7.3. With reference to Definition 6.1, assume that ∆2 = 0 and ∆3 6= 0. Then v−2 = b2(c2 − 1) k + c22 − 3c2 v+2 . Proof. By Lemma 6.2(i) and Lemma 7.2 we have −v+2 − v − 2 = E ∗ 2A2E ∗ 2v = − c2(k − 2) k + c22 − 3c2 v+2 . The result follows. Corollary 7.4. With reference to Definition 6.1, assume that D = 5, ∆2 = 0 and ∆3 6= 0. Then W = span{v+2 , v + 3 , v + 4 , v + 5 , v − 3 }. (7.1) Proof. Immediately from Theorem 6.4 and Corollary 7.3. Observe that by (3.1) vectors v+2 , v + 3 , v + 4 , v + 5 are linearly independent, provided they are non-zero. 8 Some scalar products With reference to Definition 6.1, assume that D = 5, ∆2 = 0 and ∆3 6= 0. Our goal for the rest of this paper is to find a basis for W . In this section we compute the norms of vectors v+3 , v + 4 , v + 5 , v − 3 in terms of the intersection numbers of Γ and ‖v‖. Note that by [10, Lemma 6.4] we have ∆4 6= 0 as well. The assumptions of [10, Lemma 6.4] are somehow different from assumptions of Notation 4.3. However, the proof of [10, Lemma 6.4] works just fine also under assumptions of Notation 4.3. Lemma 8.1. With reference to Definition 6.1, assume that ∆2 = 0 and ∆3 6= 0. Then ‖v+3 ‖2 = b2(b2 − c2) k + c22 − 3c2 ‖v‖2. In particular, if D ≥ 5 then v+3 6= 0. Proof. By Lemma 6.2(ii), (2.1) and Definition 3.1 we have ‖v+3 ‖2 = 〈v + 3 , v + 3 〉 = 〈Rv + 2 , v + 3 〉 = 〈v + 2 , Lv + 3 〉. 10 Art Discrete Appl. Math. 3 (2020) #P2.04 The result now follows from Lemma 6.3(ii), Corollary 7.3 and since α2 = 0, β2 = 1. Now assume that v+3 = 0. Observe that this implies b2 = c2. If D ≥ 5 then by [2, Proposition 4.1.6](i),(ii) we have c2 ≤ c3 ≤ b2, and so c2 = c3. But then c2 = 1 by Lemma 7.1(i), and so k = b2 + c2 = 2, a contradiction. Lemma 8.2. With reference to Definition 6.1, assume that ∆2 = 0 and ∆3 6= 0. Then 〈v+3 , v − 3 〉 = b2b4(c2 − 1) k + c22 − 3c2 ‖v‖2. Proof. By Lemma 6.2(ii), (2.1) and Definition 3.1 we have 〈v+3 , v − 3 〉 = 〈Rv + 2 , v − 3 〉 = 〈v + 2 , Lv − 3 〉. The result now follows from Lemma 6.2(iii) and Corollary 7.3. Lemma 8.3. With reference to Definition 6.1, assume that D = 5, ∆2 = 0 and ∆3 6= 0. Then ‖v+4 ‖2 = b2((b3 − 1)b2 − c3(c2 − 1)b4) c2(k + c22 − 3c2) ‖v‖2. In particular, v+4 = 0 if and only if c2 6= 1 and b4 = b2(b3 − 1)/(c3(c2 − 1)). Proof. By Lemma 6.2(ii), (2.1) and Definition 3.1 we have 〈v+4 , v + 4 〉 = 1 c2 〈Rv+3 , v + 4 〉 = 1 c2 〈v+3 , Lv + 4 〉. The formula for ‖v+4 ‖2 now follows from Lemma 6.3(ii), Lemma 7.1, Lemma 8.1 and Lemma 8.2. It is clear that v+4 = 0 if c2 6= 1 and b4 = b2(b3 − 1)/(c3(c2 − 1)). Therefore assume now that v+4 = 0. It follows that (b3 − 1)b2 = c3(c2 − 1)b4. If c2 = 1, then also b3 = 1 and c3 = 1 by Lemma 7.1(i). But then k = c3 + b3 = 2, a contradiction. Therefore c2 6= 1 and the result follows. Lemma 8.4. With reference to Definition 6.1, assume that D = 5, ∆2 = 0 and ∆3 6= 0. Then ‖v−3 ‖2 = ( (c2 − 1)(c4 − 1)b2 k + c22 − 3c2 + (k − 1)∆3 b2 − 1 ) b2b4‖v‖2 c2(kc2 − k − c2) + b2 . Proof. By Lemma 6.2(iv), (2.1) and Definition 3.1 we have c3〈v−3 , v − 3 〉 = b3〈v + 3 , v − 3 〉+ 〈Rv − 2 , v − 3 〉 − 〈v + 4 , Rv − 3 〉. The result now follows from Lemmas 6.3(iii), 7.1, 8.2 and 8.3, Corollary 7.3 and (4.1). Corollary 8.5. With reference to Definition 6.1, assume that D = 5, ∆2 = 0 and ∆3 6= 0. Then the following (i), (ii) hold. (i) v−3 6= 0. (ii) v+3 , v − 3 are linearly independent. Š. Miklavič and S. Penjić: On the Terwilliger algebra of BDRG with ∆2 = 0 11 Proof. (i) Note that (c2−1)(c4−1)b2/(k+c22−3c2) ≥ 0 and that (k−1)∆3/(b2−1) > 0 by [3, Theorem 12]. Moreover, it is easy to see that c2(kc2 − k − c2) + b2 > 0. The result follows. (ii) Assume on the contrary that v+3 , v − 3 are linearly dependent. Let B = ( 〈v+3 , v + 3 〉 〈v + 3 , v − 3 〉 〈v−3 , v + 3 〉 〈v − 3 , v − 3 〉 ) and note that det(B) = 0. Using Lemmas 8.1, 8.2 and 8.4 one could easily see that the only factor of det(B) which could be zero is c4k − c32k + 2c22k − 2c2k + c32c4 − 2c22c4 − c2c4 + 2c22. Solving this for c4 and then computing ∆3 using Definition 4.1, we obtain ∆3 = 0, a contradiction. This shows that v+3 , v − 3 are linearly independent. Lemma 8.6. With reference to Definition 6.1, assume that D = 5, ∆2 = 0 and ∆3 6= 0. Then ‖v+5 ‖2 = b4 − c4 + α4 + c2β4 c3 ‖v+4 ‖2. In particular, v+5 = 0 if and only if v + 4 = 0 or b4 − c4 + α4 + c2β4 = 0. Proof. By Lemma 6.2(ii), (2.1) and Definition 3.1 we have 〈v+5 , v + 5 〉 = 1 c3 〈Rv+4 , v + 5 〉 = 1 c3 〈v+4 , Lv + 5 〉. The result now follows from Lemma 6.3(ii). 9 A basis With reference to Definition 6.1, assume that D = 5, ∆2 = 0 and ∆3 6= 0. In this section we display a basis for W . We will also show that, up to isomorphism, Γ has a unique irreducible T -module with endpoint 2. Theorem 9.1. With reference to Definition 6.1, assume that D = 5, ∆2 = 0 and ∆3 6= 0. Then the following (i)–(iii) hold. (i) If v+5 6= 0, then the following is a basis for W : v+i (2 ≤ i ≤ 5), v − 3 . (9.1) (ii) If v+4 6= 0 and v + 5 = 0, then the following is a basis for W : v+i (2 ≤ i ≤ 4), v − 3 . (9.2) (iii) If v+4 = 0, then the following is a basis for W : v+i (2 ≤ i ≤ 3), v − 3 . (9.3) In particular, W is not thin. 12 Art Discrete Appl. Math. 3 (2020) #P2.04 Proof. Note that by (7.1), W is spanned by vectors v+i (2 ≤ i ≤ 5) and v − 3 . Vector v+2 = v is nonzero by definition. Vectors v + 3 and v − 3 are nonzero by Lemma 8.1 and Corollary 8.5(i), respectively. We prove part (i) of the theorem. Proofs of parts (ii) and (iii) are similar. If v+5 6= 0, then v + 4 6= 0 by Lemma 8.6. Vectors v + i (2 ≤ i ≤ 5) and v − 3 are linearly independent by (3.1) and Corollary 8.5(ii). This shows that (9.1) is a basis for W . As dim (E∗2 (W )) = 2, W is not thin. The result follows. Theorem 9.2. With reference to Definition 6.1, assume that D = 5, ∆2 = 0 and ∆3 6= 0. Then Γ has, up to isomorphism, exactly one irreducible T -module with endpoint 2. Proof. Let U denote an irreducible T -module with endpoint 2, different from W . Fix nonzero u ∈ E∗2U , and for 2 ≤ i ≤ 5 define u+i = E ∗ i Ai−2E ∗ 2u and let u−3 = E ∗ 3A5E ∗ 2u. It follows from the results of Section 8 and Theorem 9.1 that u+2 , u + 3 , u − 3 are nonzero and that nonzero vectors in the set {u + i | 2 ≤ i ≤ 5} ∪ {u − 3 } form a basis for U . Furthermore, it follows from Lemma 8.3 and Lemma 8.6 that u+4 (u + 5 , respectively) is nonzero if and only if v+4 (v + 5 , respectively) is nonzero. Let σ : W → U be defined by σ(v+i ) = u + i (2 ≤ i ≤ 5) and σ(v − 3 ) = u − 3 . It follows from the comments above that σ is a vector space isomorphism from W to U . We show that σ is a T -module isomorphism. Since A generates M and E∗0 , E ∗ 1 , . . . , E ∗ 5 is a basis for M∗, it suffices to show that σ commutes with each of A,E∗0 , E ∗ 1 , . . . , E ∗ 5 . Using the fact that E∗i E ∗ j = δijE ∗ i and the definition of σ we immediately find that σ commutes with each of E∗0 , E ∗ 1 , . . . , E ∗ 5 . Recall that A = R + L. It follows from Lemma 6.2, Lemma 6.3 and Corollary 7.3 that σ commutes with A. The result follows. We would like to emphasize that together with the results in [10, 12, 15], Theorems 9.1 and 9.2 imply the following characterization. Theorem 9.3. Let Γ = (X,R) denote a bipartite distance-regular graph with diameter D ≥ 4 and valency k ≥ 3. Assume Γ is not almost 2-homogeneous. We fix x ∈ X and let E∗i = E ∗ i (x) (0 ≤ i ≤ D) and T = T (x) denote the dual idempotents and the Terwilliger algebra of Γ with respect to x, respectively. Then the following (i), (ii) are equivalent. (i) Γ has, up to isomorphism, exactly one irreducible T -module W with endpoint 2, and W is non-thin with dim(E∗2W ) = 1, dim(E ∗ D−1W ) ≤ 1 and dim(E∗iW ) ≤ 2 for 3 ≤ i ≤ D. (ii) ∆2 = 0, and there exist complex scalars αi, βi (2 ≤ i ≤ D − 1) such that |Γi−1(x) ∩ Γi−1(y) ∩ Γ1(z)| = αi + βi|Γ1(x) ∩ Γ1(y) ∩ Γi−1(z)| (9.4) for all y ∈ Γ2(x) and z ∈ Γi(x) ∩ Γi(y). With reference to Definition 6.1, assume that ∆2 = 0 and ∆3 6= 0. It is known that this implies c2 ∈ {1, 2}, or D ≤ 5, see [12, Theorem 4.4]. If c2 ∈ {1, 2}, then the structure of irreducible T -modules with endpoint 2 was studied in detail in [12, 15]. Therefore, we are mainly interested in the case c2 ≥ 3. We have to mention however that we are not aware of any of such a graph. Using a computer program we found intersection arrays Š. Miklavič and S. Penjić: On the Terwilliger algebra of BDRG with ∆2 = 0 13 {b0, b1, b2, b3, b4; c1, c2, c3, c4, c5} up to valency k = 20000, which satisfy the following conditions: c2 ≥ 3, ∆2 = 0, ∆3 > 0, ∆4 > 0, γ2 ∈ N, p222 ∈ N. None of them passed the feasibility condition p1ij ∈ N ∪ {0}, see the table below. intersection arrays feasibility condition (58, 57, 49, 21, 1; 1, 9, 37, 57, 58) p123 = 1102/3 /∈ N (112, 111, 100, 45, 4; 1, 12, 67, 108, 112) p134 = 103600/67 /∈ N (186, 185, 161, 35, 1; 1, 25, 151, 185, 186) p123 = 6882/5 /∈ N (274, 273, 256, 120, 10; 1, 18, 154, 264, 274) p123 = 12467/3 /∈ N (274, 273, 256, 120, 1; 1, 18, 154, 273, 274) p123 = 12467/3 /∈ N (1192, 1191, 1156, 561, 28; 1, 36, 631, 1164, 1192) p123 = 118306/3 /∈ N (3236, 3235, 3136, 760, 1; 1, 100, 2476, 3235, 3236) p123 = 523423/5 /∈ N ORCID iDs Štefko Miklavič https://orcid.org/0000-0002-2878-0745 Safet Penjić https://orcid.org/0000-0001-6664-4130 References [1] E. Bannai and T. Ito, Algebraic combinatorics. I, The Benjamin/Cummings Publishing Co., Inc., Menlo Park, CA, 1984, association schemes. [2] 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. [3] B. Curtin, 2-homogeneous bipartite distance-regular graphs, Discrete Math. 187 (1998), 39–70, doi:10.1016/S0012-365X(97)00226-4. [4] B. Curtin, Bipartite distance-regular graphs. I, Graphs Combin. 15 (1999), 143–158, doi:10. 1007/s003730050049. [5] B. Curtin, Bipartite distance-regular graphs. II, Graphs Combin. 15 (1999), 377–391, doi:10. 1007/s003730050072. [6] B. Curtin, The local structure of a bipartite distance-regular graph, European J. Combin. 20 (1999), 739–758, doi:10.1006/eujc.1999.0307. [7] B. Curtin, Almost 2-homogeneous bipartite distance-regular graphs, European J. Combin. 21 (2000), 865–876, doi:10.1006/eujc.2000.0399. [8] E. S. Egge, A generalization of the Terwilliger algebra, J. Algebra 233 (2000), 213–252, doi: 10.1006/jabr.2000.8420. [9] J. T. Go, The Terwilliger algebra of the hypercube, European J. Combin. 23 (2002), 399–429, doi:10.1006/eujc.2000.0514. [10] M. S. MacLean and Š. Miklavič, On bipartite distance-regular graphs with exactly one non- thin T -module with endpoint two, European J. Combin. 64 (2017), 125–137, doi:10.1016/j. ejc.2017.04.004. [11] M. S. MacLean and Š. Miklavič, On bipartite distance-regular graphs with exactly two ir- reducible T-modules with endpoint two, Linear Algebra Appl. 515 (2017), 275–297, doi: 10.1016/j.laa.2016.11.021. 14 Art Discrete Appl. Math. 3 (2020) #P2.04 [12] M. S. MacLean, Š. Miklavič and S. Penjić, On the Terwilliger algebra of bipartite distance- regular graphs with ∆2 = 0 and c2 = 1, Linear Algebra Appl. 496 (2016), 307–330, doi: 10.1016/j.laa.2016.01.040. [13] M. S. MacLean, Š. Miklavič and S. Penjić, An A-invariant subspace for bipartite distance- regular graphs with exactly two irreducible T -modules with endpoint 2, both thin, J. Algebraic Combin. 48 (2018), 511–548, doi:10.1007/s10801-017-0798-7. [14] Š. Miklavič, The Terwilliger algebra of a distance-regular graph of negative type, Linear Alge- bra Appl. 430 (2009), 251–270, doi:10.1016/j.laa.2008.07.013. [15] S. Penjić, On the Terwilliger algebra of bipartite distance-regular graphs with ∆2 = 0 and c2 = 2, Discrete Math. 340 (2017), 452–466, doi:10.1016/j.disc.2016.09.001. [16] P. Terwilliger, The subconstituent algebra of an association scheme. I, J. Algebraic Combin. 1 (1992), 363–388, doi:10.1023/A:1022494701663.