arXiv:1907.00260v1 [math.RA] 29 Jun 2019 Introduction 1 atclrcsso hs arcsi re odrv xlctexpres explicit derive others). to among order therein). [20] in references [19], matrices [17], the these [16], and eng [14] of in or a cases applications varied [12] particular also a [4], but only (see not [2]) theory have instance, control matrices for of (see, type computations these mial fact, of matter a As fteecag arx,dsiemlil ymtisla osm re some to lead matrices. symmetries of multiple classes two despite kno these ), to Hank exchange allow persymmetric the not corresponding of the does of matrix th eigenvectors Toeplitz general, and symmetric values in a that, of emphasize eigenvectors We Toe and symmetric matrices. ar band Hankel of for persymmetric state developed the in matrices. techniques for Toeplitz contribution that small show symmetric a matrices, band being expens for pr besides the eighties here the at nounced the in obtained in ingredient matrices, settled essential of originally sort The these per available. for anti-heptadiagonal made decomposition any is of nonsingularity) inverse its the Particula calculate matrices. comp to referred the the parameter for for formulae diagonalization announce explic a purpose an thereunto, our getting is it and Additionally, eigenvalues its tors. locating namely matrices, Hankel 1 pcrlpoete fanti-heptadiagonal of properties Spectral e words: Key 00MteaisSbetClassification: Subject Mathematics 2010 h motneo aklmtie ncmuainlmteaisa mathematics computational in matrices Hankel of importance The h i fti ae st xlr pcrlpoete fgnrla general of properties spectral explore to is paper this of aim The -aladdress: E-mail esmercHne arcsi provided. f parameter is unknown matrices any compu Hankel on to persymmetric depending not parameters formula localizable explicit represent on an a also depending giving expression polynomials an explicit of zeros the as nti ae eepesteegnauso anti-heptadiago of eigenvalues the express we paper this In nihpaignlmti,Hne arx ievle ei eigenvalue, matrix, , Anti-heptadiagonal esmercHne matrices Hankel persymmetric [email protected] unad or,22-1 aaia Portugal Caparica, 2829-516 Torre, da Quinta eateto ahmtc n GeoBioTec and Mathematics of Department ; aut fSine n Technology and Sciences of Faculty [email protected] OAUiest fLisbon of University NOVA oa iad Silva da Jo˜ao Lita Abstract 51,15B05 15A18, 1 to fisegnetr.W rsn also present We eigenvectors. its of ation eisitgrpwr.I particular, In powers. integer its te 1 eety eea uhr aestudied have authors several Recently, rteivreo anti-heptadiagonal of inverse the or l,a xrsinfe fayunknown any of free expression an rly, a esmercHne matrices Hankel persymmetric nal lmti otie ymultiplication by (obtained matrix el fteso-called the of e ymti aklmti (assuming matrix Hankel symmetric nseta etrso anti-banded of features spectral on t lt arcscnb loemployed also be can matrices plitz dnmru eain ihpolyno- with relations numerous nd evco,diagonalization genvector, tto fisitgrpwr using, powers integer its of utation in o t oes(e 7,[11], [7], (see powers its for sions cuitneo h eigenvalues the of acquaintance e neigpolm fsse and system of problems ineering t-etdaoa persymmetric nti-heptadiagonal trpeetto fiseigenvec- its of representation it ainhpbtenegnasof eigenpars between lationship hc r xcl h eigen- the exactly are which w degneigi well-known. is engineering nd hrb,tesaeet an- statements the Thereby, sn prahi suitable a is approach esent odrn technique bordering An n × n matrix Bn is said to be banded when all its nonzero elements are confined within a band formed by diagonals parallel to the main diagonal, i.e. [Bn]k,ℓ = 0 when |k − ℓ| >h, and [Bn]k,k−h 6=0 or [Bn]k,k+h 6= 0 for at least one value of k, where h is the half-bandwidth and 2h + 1 is the bandwidth (see [15], page 13). An n × n matrix An is said to be anti-banded if JnAn is a banded matrix where Jn is the n × n (that is, the matrix such that [Jn]k,n−k+1 = 1 for each 1 6 k 6 n with the remaining entries zero). Particularly, an n × n matrix which has the form

0 ...... 0 αn βn γn dn . ..  . . αn−1 βn−1 γn−1 dn−1 cn−1  . . .  . .. .. β γ d c b   n−2 n−2 n−2 n−2 n−2   . . . .   ......   . γn−3 dn−3 cn−3 bn−3 an−3   . . . . .   ......   . dn−4 cn−4 bn−4 an−4 0   . .   ......   ......     ......   0 α5 β5 γ5 d5 . . . . .     ......   α4 β4 γ4 d4 c4 . . . .     .. .. .   β3 γ3 d3 c3 b3 . . .     .. .   γ2 d2 c2 b2 a2 . .     d1 c1 b1 a1 0 ...... 0      is called anti-heptadiagonal. Throughout, we shall consider the following n × n anti-heptadiagonal persymmetric Hankel matrix

0 ...... 0 a b cd . .  . .. a b c dc  . . .  . .. .. b c d cb     . . . .   ......   . c d c ba   . . . . .   ......   . d c b a 0   ......  Hn =  ......  . (1.1)  ......     ......   0 a b c d . . . . .     ......   ab c d c . . . .     .. .. .   bc d c b . . .     . .   cd c b a .. .     dc b a 0 ...... 0      2 Auxiliary tools

Consider the class of complex matrices defined by

A M C n := An ∈ n( ):[An]k−1,ℓ + [An]k+1,ℓ = [An]k,ℓ−1 + [An]k,ℓ+1 , k,ℓ =1, 2,...,n n o where it is assumed [An]n+1,ℓ = [An]k,n+1 = [An]0,ℓ = [An]k,0 = 0 for all k,ℓ. We begin by retrieve a pair of results stated in [1] (see Propositions 2.1 and 2.2), announced here for complex matrices. The proofs do not suffer any changes from the original ones, whence we will omit the details.

2 Lemma 1 (a) The class An is the algebra generated over C by the n × n matrix

0 1 0 ...... 0 .. .  1 0 1 . .  .. .. .  0 1 0 . . .     ......  Ωn =  ......  . (2.1)    . . .   . .. .. 0 1 0     . .   . .. 1 0 1     0 ...... 0 1 0      ⊤ (b) If An ∈ An and a is its first row then

n−1 k An = ωk+1Ωn k X=0 ⊤ where Ωn is the n × n matrix (2.1) and ω = (ω1,ω2,...,ωn) is the solution of the upper triangular ω system Tn = a, with [Tn]0,0 := 1, [Tn]0,ℓ := 0 for all 1 6 ℓ 6 n,

[T ] + [T ] , 1 6 k 6 ℓ 6 n [T ] := n k−1,ℓ−1 n k+1,ℓ−1 . n k,ℓ 0, otherwise 

Given a positive integer n, Sn+2 will denote the (n + 2) × (n + 2) symmetric, involutory and defined by

2 kℓπ [S ] := sin . (2.2) n+2 k,ℓ n +3 n +3 r   Our second auxiliary result makes useful Lemma 1 above and provides us an orthogonal diagonalization for the following (n + 2) × (n + 2) complex matrix

0 ...... 0 a b c − a d − b . .  . .. abc dc − a  . . .  . .. .. bcd c b     . . . .   ......   . cdc b a   . . . . .   ......   . dcb a 0   ......  Hn =  ......  . (2.3) +2  ......     ......  b  0 a bcd . . . . .     ......   a b cdc . . . .     .. .. .   b c dcb . . .     . .   c − ad cba .. .     d − b c − a b a 0 ...... 0      Lemma 2 Let n ∈ N, a,b,c,d ∈ C and

nkπ (n+1)kπ (n+2)kπ λk = −2a cos n+3 − 2b cos n+3 − 2c cos n+3 − d cos(kπ) , k =1, 2,...,n +2. (2.4)   h i h i 3 If Hn+2 is the (n + 2) × (n + 2) matrix (2.3) then

b Hn+2 = Sn+2Λn+2Sn+2 where Λn+2 = diag (λ1, λ2,...,λn+2) andb Sn+2 is the (n + 2) × (n + 2) matrix (2.2).

Proof. Suppose n ∈ N and a,b,c,d ∈ C. Since Hn+2 ∈ An+2 and its first row is

⊤ a = (0,...,b0,a,b,c − a, d − b) we have, from Lemma 1,

2 3 Hn+2 = (d − 2b)In+2 + (c − 3a)Ωn+2 + b Ωn+2 + a Ωn+2 Jn+2   where Jn+2 is the (bn + 2) × (n + 2) exchange matrix (that is, the matrix such that [Jn+2]k,n−k+3 =1 for every 1 6 k 6 n + 2 with the remaining entries zero). Using the spectral decomposition

n+2 ℓπ Ω = 2cos s s⊤, n+2 n +3 ℓ ℓ ℓ X=1   where 2 ℓπ n+3 sin n+3  q 2  2ℓπ   n+3 sin n+3 sℓ =  q .     .     2 (n+1)ℓπ   n sin n   +3 +3   q h i  (i.e. the ℓth column of Sn+2 in (2.2)) and the decomposition

2 3 n+3 Jn+2 = Sn+2 diag (−1) , (−1) ,..., (−1) Sn+2   (see Lemma 1 of [8]) with Sn+2 given by (2.2), it follows

n+2 ℓπ 2 ℓπ 3 ℓπ ⊤ Hn+2 = (d − 2b)+2(c − 3a)cos +4b cos +8a cos sℓ sℓ Jn+2 ( n +3 n +3 n +3 ) ℓ=1        . X 2 3 n+3 2 3 n+3 b = Sn+2 diag (−1) λ1, (−1) λ2,..., (−1) λn+2 diag (−1) , (−1) ,..., (−1) Sn+2 = Sn+2 diag (λ1, λ2,...,λn+2) Sn+2    The proof is completed. 

The decomposition below plays a central role in the study of the spectral properties of anti- heptadiagonal persymmetric Hankel matrices. The statement is split in two cases of n ∈ N: n is an even number and n is an odd number. For the case in which n is even, we shall take the same steps as in [1] (see pages 106 and 107); the case in which n is odd, this procedure must be improved and we shall follow closely the bordering technique of Fasino (see [5], pages 303 and 304).

Lemma 3 Let n ∈ N, a,b,c,d ∈ C, λk, k =1, 2,...,n +2 be given by (2.4) and Hn the n × n matrix (1.1). (a) If n is even then

⊤ ⊤ diag (λ3, λ5,...,λn+1)+ λ1uu O Hn = RnPn ⊤ PnRn (2.5a) O diag (λ , λ ,...,λn)+ λn vv  2 4 +2 

4 where 3π 2π sin( n+3 ) sin( n+3 ) π π sin( n+3 ) sin( n+3 )

 5π   4π  sin( n+3 ) sin( n+3 ) π π  sin( n )   sin( n )  u =  +3  , v =  +3  , (2.5b)  .   .   .   .   (n+1)π   nπ   sin[ n+3 ]   sin( n+3 )   π   π   sin( n )   sin( n )   +3   +3      Rn is the n × n matrix given by

2 (k + 1)(ℓ + 1)π [R ] = sin (2.5c) n k,ℓ n +3 n +3 r   and Pn is the n × n defined by

1 if ℓ =2k − n − 1 or ℓ =2k [Pn]k,ℓ = . (2.5d) ( 0 otherwise

(b) If n is odd then

⊤ diag (λ3, λ5,...,λn, λn+2)+ λ1uu O ⊤ ⊤ Hn = RnPn ⊤ Pn Rn O diag (λ , λ ,...,λn− )+ λn vv  2 4 1 +1  (2.6a) where 3π sin n ( +3 ) 2π π sin( n+3 ) sin n ( +3 ) (n+1)π sin[ n+3 ]  5π  sin n ( +3 )  4π  π sin( n+3 ) sin n  ( +3 )  (n+1)π    sin[ n ]  u =  .  , v =  +3  , (2.6b)  .   .   (n+1)π   .  sin n  [ +3 ]   (n−1)π   π   sin[ n+3 ]  sin n  ( +3 )   (n+1)π     sin[ n ]   1   +3        Rn is the n × n matrix given by

2 (k+1)(ℓ+1)π n+3 sin n+3 if ℓ

[Pn]k,ℓ =  1 if k = n and ℓ = (n + 1)/2 . (2.6d)  0 otherwise  Proof. Consider n ∈ N, a,b,c,d ∈ C and λk, k =1, 2,...,n + 2 given by (2.4).

(a) Supposing n even and Sn+2 given by (2.2), we have

θ x⊤ θ Sn+2 =  x Rn y  θ y⊤ −θ     5 2 π where θ := n+3 sin n+3 , q   2 2π 2 2π n+3 sin n+3 (−1) n+3 sin n+3  q     q    2 sin 3π (−1)2 2 sin 3π x =  n+3 n+3  and y =  n+3 n+3  . (2.7)  .   .   q .     q .     .   .       2 (n+1)π   n 2 (n+1)π   n+3 sin n+3   (−1) n+3 sin n+3       q h i   q h i  From Lemma 2, we obtain Hn+2 = Sn+2 diag (λ1, λ2,...,λn+2) Sn+2 and since Hn is obtained by deleting the first and last row and column of the matrix Hn+2 in (2.3), it follows b −1 −1 ⊤ −1 −1 ⊤ Hn = Rn diag(λ2, λ3,...,λn+1)+ λ1Rn x Rnb x + λn+2Rn y Rn y Rn h   i with x + y R−1x = − , (2.8) n 2θ x − y R−1y = − . (2.9) n 2θ The odd components of (2.8) and the even components of (2.9) are zero; moreover, the kth even component of (2.8) and the kth odd component of (2.9) is

(k+1)π sin n+3 − . h π i sin n+3   −1 −1 ⊤ −1 −1 ⊤ Permuting rows and columns of diag(λ2, λ3,...,λn+1)+ λ1Rn x Rn x + λn+2Rn y Rn y ac- cording to (2.5d) yields (2.5a).   (b) Consider n odd and the (n + 2) × (n + 2) orthogonal matrix Sn+2 := Sn+2Kn+2, where Sn+2 is the matrix (2.2) and Kn+2 is the (n + 2) × (n + 2) permutation matrix b In 0 0 ⊤ Kn+2 :=  0 0 1  .  0⊤ 1 0      ⊤ Using Lemma 2, we have Hn+2 = Sn+2 diag (λ1, λ2,...,λn, λn+2, λn+1) Sn+2. Putting

b b θ z⊤ η b Sn+2 =  x Rn y  θ w⊤ −η   b   2 π 2 (n+1)π where θ := n+3 sin n+3 , η := n+3 sin n+3 , q   q h i

6 2 2π 2 4π n+3 sin n+3 (−1) n+3 sin n+3  q     q    2 sin 3π (−1)2 2 sin 6π x =  n+3 n+3  , y =  n+3 n+3   .   .   q .     q .     .   .       2 (n+1)π   n 2 2(n+1)π   n+3 sin n+3   (−1) n+3 sin n+3       q 2 h 2π i   q 2 h 2π i  n+3 sin n+3 (−1) n+3 sin n+3 (2.10)  q     q    2 sin 3π (−1)2 2 sin 3π  n+3 n+3   n+3 n+3   .   .  z =  q .    , w =  q .     .   .       2 sin nπ   (−1)n−1 2 sin nπ   n+3 n+3   n+3 n+3       q     q     2 (n+2)π   2 (n+2)π   n+3 sin n+3   n+3 sin n+3       q h i   q h i  and Rn given by (2.6c), we get

−1 −1 ⊤ −1 −1 ⊤ ⊤ Hn = Rn diag(λ2, λ3,...,λn, λn+2)+ λ1Rn x Rn x + λn+1Rn y Rn y Rn h i with   z + w R−1x = − , (2.11) n 2θ z − w R−1y = − (2.12) n 2η since Hn is obtained by deleting the first and last row and column of the matrix Hn+2. The first n − 1 odd components of (2.11) are zero; additionally, the even components of (2.12) are zero as well as its last component. In the first n − 1 components of (2.11), the kth even componentb is

(k+1)π sin n+3 − , h π i sin n+3   and the kth odd component of (2.12) is

(k+1)π sin n+3 − . h (n+1)π i sin n+3 h i Permuting rows and columns of

−1 −1 ⊤ −1 −1 ⊤ diag(λ2, λ3,...,λn, λn+2)+ λ1Rn x Rn x + λn+1Rn y Rn y according to (2.6d) we obtain (2.6a) establishing the thesis.  

ℓ+1 Remark Let us observe that for n odd the condition Sn+2 = (−1) Sn+2 , ℓ =1, 2,...,n+2 n+2,ℓ 1,ℓ is not satisfied (unlike the case in which n is evenh andi we have [S h] i= (−1)ℓ+1 [S ] , b n+2 nb+2,ℓ n+2 1,ℓ ℓ = 1, 2,...,n + 2). This foreshadowed that the bordering technique can be founded taking account only the four corners of the orthogonal matrix (in order to get det(Rn) 6= 0) and discarding any relation between the remaining entries of its first and last rows.

7 3 Spectral properties of Hn 3.1 Eigenvalue localization

The decomposition presented in Lemma 3 allows us to express the eigenvalues of Hn as the zeros of explicit polynomials.

Theorem 1 Let n ∈ N, a,b,c,d ∈ C, λk, k =1, 2,...,n+2 be given by (2.4) and Hn the n×n matrix (1.1).

(a) If n is even then the eigenvalues of Hn are precisely the zeros of

n n 2 (2j+1)π n 2 2 2 sin n+3 f(t)= (t − λ2j+1) + (t − λ1) (t − λ2m+1) (3.1a) 2h π i j=1 j=1 sin n+3 m=1 Y X mY=6 j   and n n 2 2jπ n 2 2 2 sin n+3 g(t)= (t − λ2j ) + (t − λn+2) (t − λ2m). (3.1b) 2  π  j=1 j=1 sin n+3 m=1 Y X mY=6 j   (b) If n is odd then the eigenvalues of Hn are precisely the zeros of

n+1 n+1 2 (2j+1)π n+1 2 2 2 sin n+3 f(t)= (t − λ2j+1) + (t − λ1) (t − λ2m+1) (3.2a) 2h π i j=1 j=1 sin n+3 m=1 Y X mY=6 j   and n−1 n−1 2 2jπ n−1 2 2 2 sin n+3 g(t)= (t − λ2j ) + (t − λn+1) (t − λ2m). (3.2b) 2 (n+1)π j=1 j=1 sin n+3 m=1 Y X mY=6 j h i Proof. Let n ∈ N, a,b,c,d ∈ C and λk, k =1, 2,...,n + 2 be given by (2.4).

(a) Consider n even and Rn the matrix defined by (2.5c). We have

2 ⊤ ⊤ Rn = In − xx − yy

2 16 4 π where x, y are given by (2.7) and det(Rn)= (n+3)2 sin n+3 6= 0 (see [1], page 106), so that   −2 ⊤ ⊤ −1 Rn = In − xx − yy . Setting 

Σ n := diag (λ , λ ,...,λ )+ λ uu⊤ 2 3 5 n+1 1 Γ n := diag (λ , λ ,...,λ )+ λ vv⊤ 2 2 4 n n+2 with u, v given by (2.5b), we obtain

det(tIn − Hn)=

n 2 −2 ⊤ Σ O = [det(Rn)] det tR − P 2 Pn n n O Γ n   2   n 2 −2 ⊤ Σ O = det R det tPnR P − 2 n n n O Γ n   2   8 n 2 ⊤ ⊤ ⊤ −1 Σ O = det R det t Pn In − xx − yy P − 2 n n O Γ n   2    2 π   −1 4 sin ( n ) I n − +3 uu⊤ O Σ n O 2 2 n+3 2 = det Rn det t 2 π −  4 sin n n   n ( +3 ) ⊤  O Γ O I − n vv  2   2 +3      n 1 ⊤  I − ⊤ n+3 uu O 2 u u− 2 π n 4 sin ( n ) Σ O R2 +3 2 = det n det t n 1 ⊤ − n   O I − ⊤ n vv  O Γ  2 v v− +3 2 2 π   4 sin ( n )    +3       2 t ⊤ = det R det tI n − diag (λ , λ ,...,λn ) − λ + uu · n 2 3 5 +1 1 ⊤ n+3   u u − 2 π   4 sin ( n )   +3     t  ⊤ det tI n − diag (λ , λ ,...,λn) − λn + vv 2 2 4 +2 ⊤ n+3   v v − π   4 sin2  ( n+3 )    (see [9], page 88 and [10], page 19) where Pn the permutation matrix (2.5d). Thus, supposing t 6= λ2k+1 n u⊤u n+3 for each k =1,..., 2 and noting that − 2 π = −1, we get 4 sin ( n+3 )

t ⊤ det tI n − diag (λ , λ ,...,λn ) − λ + uu = 2 3 5 +1 1 ⊤ n+3   u u − π   4 sin2  ( n+3 )  n  2 (2j+1)π n  2 2  sin n+3  = 1 − (λ1 − t) (t − λ2j+1)  h 2 i π  (3.3) j=1 (t − λ2j+1) sin j=1  X n+3  Y n n 2 (2j+1)π n 2 2 2  sin n+3  = (t − λ2j+1) + (t − λ1) (t − λ2m+1). 2h π i j=1 j=1 sin n+3 m=1 Y X mY=6 j   Since the determinant is a continuous function, relation (3.3) holds for all t. Analogously, we have ⊤ n+3 v v − 2 π = −1 and 4 sin ( n+3 )

t ⊤ det tI n − diag (λ , λ ,...,λn) − λn + vv = 2 2 4 +2 ⊤ n+3   v v − π   4 sin2  ( n+3 )  n  2 2jπ n  2 2  sin n+3  = 1 − (λn+2 − t) (t − λ2j )   2  π  j=1 (t − λ2j ) sin j=1 X n+3 Y n n  2jπ  n 2 2 2 2 sin n+3 = (t − λ2j ) + (t − λn+2) (t − λ2m) 2  π  j=1 j=1 sin n+3 m=1 Y X mY=6 j   for all t, so that the characteristic polynomial of Hn is

4 n n 2 (2j+1)π n π 2 2 2 16 sin n+3 sin n+3 det(tIn − Hn)= (t − λ2j+1) + (t − λ1) · (t − λ2m+1) · (n + 3)2   2h π i  j=1 j=1 sin n+3 m=1  Y X mY=6 j     

9 n n 2 2jπ n 2 2 2 sin n+3 (t − λ2j ) + (t − λn+2) (t − λ2m) .  2  π   j=1 j=1 sin n+3 m=1 Y X mY=6 j     

(b) Supposing n odd and Rn the matrix defined by (2.6c), it follows

⊤ ⊤ ⊤ Rn Rn = In − zz − ww where z and w are given in (2.10). Hence,

⊤ ⊤ ⊤ det(Rn Rn) = det(In − zz − ww ) (z⊤w)2 = (1 − z⊤z) 1 − w⊤w − 1 − z⊤z   16 π (n + 1)π = sin2 sin2 6=0 (n + 3)2 n +3 n +3     (see [13], page 70) and ⊤ −1 ⊤ ⊤ −1 Rn Rn = In − zz − ww . Putting  

⊤ Σ n+1 := diag (λ3, λ5,...,λn, λn+2)+ λ1uu 2 ⊤ Γ n−1 := diag (λ2, λ4,...,λn−1)+ λn+1vv 2 with u, v given by (2.6b), we obtain det(tIn − Hn)=

Σ n+1 O −1 ⊤ −1 2 ⊤ ⊤ = det(Rn) det tRn (Rn ) − Pn Pn det(Rn ) O Γ n−1 " 2 # !

Σ n+1 O ⊤ ⊤ ⊤ −1 2 = det Rn Rn det tPn (Rn Rn) Pn − O Γ n−1 " 2 #!  −1 Σ n+1 O ⊤ ⊤ ⊤ ⊤ 2 = det Rn Rn det t Pn In − zz − ww Pn − O Γ n−1 " 2 #!

  2 π   −1 4 sin ( n+3 ) ⊤ I n+1 − uu O Σ n+1 O ⊤ 2 n+3 2 = det R Rn det t 2 (n+1)π − n 4 sin n−   [ n+3 ] ⊤  O Γ 1  n−1 " # O I − n+3 vv 2   2  ⊤     = det Rn Rn ·

n 1 ⊤ I +1 − ⊤ n+3 uu O  2 u u− 2 π n+1 4 sin n Σ O ( +3 ) 2 det t 1 −   n− ⊤  n−  O I 1 − v⊤v n+3 vv O Γ 1 2 − (n+1)π " 2 # 4 sin2   [ n+3 ]       ⊤ t ⊤ = det Rn Rn det tI n+1 − diag (λ3, λ5,...,λn, λn+2) − λ1 + uu · 2 ⊤ n+3   u u − 2 π   4 sin ( n )   +3     t  ⊤ det tI n−1 − diag (λ2, λ4,...,λn−1) − λn+1 + vv 2 ⊤ n+3   v v − n π   4 sin2 ( +1)  [ n+3 ]      10 (see [9], page 88 and [10], page 19) where Pn the permutation matrix (2.6d). Thus, admitting t 6= λ2k+1 n+1 for each k =1,..., 2 , we have

t ⊤ det tI n+1 − diag (λ3, λ5,...,λn, λn+2) − λ1 + uu = 2 ⊤ n+3   u u − π   4 sin2  ( n+3 )  n−1 2 (2j+1)π  n−1 2 2  sin n+3 1  = 1 − (λ1 − t) + (t − λn+2) (t − λ2j+1)    2 π t − λn  j=1 (t − λ2j+1) sin +2 j=1  X n+3  Y n+1 2 (2j+1)π   n+1  (3.4) 2 2  sin n+3  = 1 − (λ1 − t) (t − λ2j+1)  h 2 i π  j=1 (t − λ2j+1) sin j=1  X n+3  Y n+1 n+1 2 (2j+1)π n+1 2 2 2  sin n+3  = (t − λ2j+1) + (t − λ1) (t − λ2m+1) 2h π i j=1 j=1 sin n+3 m=1 Y X mY=6 j   ⊤ n+3 provided that u u − 2 π = −1. Since the determinant is a continuous function, relation (3.4) 4 sin ( n+3 ) ⊤ n+3 is valid for every t. Analogously, we have v v − 2 (n+1)π = −1 and 4 sin [ n+3 ]

t ⊤ det tI n−1 − diag (λ2, λ4,...,λn−1) − λn+1 + vv = 2 ⊤ n+3   v v − n π   4 sin2 ( +1)  ( n+3 )  n−1  n−1 2 2jπ n−1 2 2 2  sin n+3 = (t − λ2j ) + (t − λn+1) (t − λ2m) 2 (n+1)π j=1 j=1 sin n+3 m=1 Y X mY=6 j h i for all t, whence the characteristic polynomial of Hn is

n−1 16 π (n + 1)π 2 det(tI − H )= sin2 sin2 (t − λ ) (t − λ )+ n n (n + 3)2 n +3 n +3 n+2 2j+1 ( j=1     Y n−1 2 (2j+1)π n−1 n−1 2 2 2 (λn+2 − t) sin n+3 (t − λ1) (t − λ2j+1)+ (t − λ2j+1) · 2 π  " j=1 sin n+3 m=1 j=1 #) X mY=6 j Y   n−1 n−1 2 2jπ n−1 2 2 2 sin n+3 (t − λ2j ) − (λn+1 − t) (t − λ2m) .  2 (n+1)π  j=1 j=1 sin n+3 m=1   Y X mY=6 j  h i The proof is completed.  

The previous theorem leads us to the following separation result for the eigenvalues of real matrices Hn.

Corollary 1 Let n ∈ N, a,b,c,d ∈ R, λk, k = 1, 2,...,n +2 be given by (2.4) and Hn the n × n matrix (1.1). (a) If n is even and:

11 i. λτ(1) 6 λτ(3) 6 ... 6 λτ(n+1) are arranged in nondecreasing order by some bijection τ then

λτ(2k−1) 6 µk 6 λτ(2k+1) (3.5a)

for any k = 1,..., n , where µ ,...,µ n are the zeros of (3.1a). Moreover, if λ are 2 1 2 τ(2k−1) all distinct then the kth inequalities (3.5a) are strict, and λτ(2k−1) = λτ(2k+1) if and only if λτ(2k−1) = µk = λτ(2k+1).

ii. λσ(2) 6 λσ(4) 6 ... 6 λσ(n+2) are arranged in nondecreasing order by some bijection σ then

λσ(2k) 6 νk 6 λσ(2k+2) (3.5b)

for all k = 1,..., n , where ν ,...,ν n are the zeros of (3.1b). Furthermore, if λ are all 2 1 2 σ(2k) distinct then the kth inequalities (3.5b) are strict, and λσ(2k) = λσ(2k+2) if and only if λσ(2k) = νk = λσ(2k+2).

(b) If n is odd and:

i. λτ(1) 6 λτ(3) 6 ... 6 λτ(n) 6 λτ(n+2) are arranged in nondecreasing order by some bijection τ then λτ(2k−1) 6 µk 6 λτ(2k+1) (3.6a) n+1 for every k = 1,..., , where µ1,...,µ n+1 are the zeros of (3.2a). Moreover, if λτ(2k−1) are 2 2 all distinct then the kth inequalities (3.5a) are strict, and λτ(2k−1) = λτ(2k+1) if and only if λτ(2k−1) = µk = λτ(2k+1).

ii. λσ(2) 6 λσ(4) 6 ... 6 λσ(n+1) are arranged in nondecreasing order by some bijection σ then

λσ(2k) 6 νk 6 λσ(2k+2) (3.6b)

n−1 for each k = 1,..., , where ν1,...,ν n−1 are the zeros of (3.2b). Furthermore, if λσ(2k) 2 2 are all distinct then the kth inequalities (3.6b) are strict, and λσ(2k) = λσ(2k+2) if and only if λσ(2k) = νk = λσ(2k+2).

Proof. Since all assertions can be proven in the same way, we only prove i. of (a). Let n ∈ N be even 6 6 n and consider τ(2ξ − 1)=1 for some 1 ξ 2 +1. According to Theorem 1, we have

n n 2 (2j+1)π n 2 2 2 sin n+3 f(t)= (t − λ2j+1) + (t − λ1) (t − λ2m+1) 2h π i j=1 j=1 sin n+3 m=1 Y X mY=6 j   n n 2 τ(2j−1)π n 2 +1 2 +1 2 +1 sin n+3 = t − λτ(2j−1) + t − λτ(2ξ−1) t − λτ(2m−1) h2 π i j=1 j=1 sin n+3 m=1 jY=6 ξ     Xj=6 ξ mY=6 ξ,j     6 n 6 6 so that f λτ(2k−1) f λτ(2k+1) 0 for all k = 1,..., 2 . Hence, λτ(2k−1) µk λτ(2k+1). If λ are all distinct then λ <µk 6 λ or λ 6 µk < λ . In the first case, if τ(2k−1)    τ(2 k−1) τ(2k+1) τ(2k−1) τ(2k+1) µk = λτ(2k+1) then f λτ(2k+1) = 0, implying λτ(2k+1) = λτ(2ζ−1) for some ζ 6= k +1 which contradicts the hypothesis; the case λτ(2k−1) 6 µk < λτ(2k+1) is analogous. Hence, the kth inequalities (3.5a) are strict.   

Remark Let us observe that, in the assumptions of Corollary 1, if λτ(1) > 0 and λσ(2) > 0 then Hn is positive definite.

12 3.2 Eigenvectors

From Lemma 3 we can get also an explicit representation for the eigenvectors of Hn.

Theorem 2 Let n ∈ N, a,b,c,d ∈ R, λk, k =1, 2,...,n+2 be given by (2.4) and Hn the n×n matrix (1.1).

(a) If n is even, u, v are given in (2.5b), Rn is the n × n matrix (2.5c), Pn is the n × n permutation matrix (2.5d),

i. µ ,...,µ n are the zeros of (3.1a) and λ , λ ,...,λ are all distinct then 1 2 1 3 n+1

3π sin( n+3 ) π k (µ −λ3) sin( n+3 )  5π  sin( n+3 ) π (µk−λ ) sin  5 ( n+3 )   .  n ⊤  .  ⊤ I + uu O   R P 2 (n+1)π n n n ⊤  sin  (3.7a) O I + vv  [ n+3 ]  2 π    (µk−λn ) sin   +1 ( n+3 )     0     .   .       0    n is an eigenvector of Hn associated to µk, k =1,..., 2 . ii. ν ,...,ν n are the zeros of (3.1b) and λ , λ ,...,λ are all distinct then 1 2 2 4 n+2

0 .  .  0    2π  sin( n+3 ) n ⊤  π  ⊤ I + uu O  (νk −λ ) sin  R P 2 2 ( n+3 ) n n n ⊤   (3.7b) O I + vv 4π 2  sin     ( n+3 )  π  (νk −λ ) sin   4 ( n+3 )   .   .   .   sin nπ   ( n+3 )  π  (νk−λn) sin   ( n+3 )    n is an eigenvector of Hn associated to νk, k =1,..., 2 .

(b) If n is odd, u, v are given in (2.6b), Rn is the n × n matrix (2.6c), Pn is the n × n permutation matrix (2.6d),

i. µ1,...,µ n+1 are the zeros of (3.1a) and λ1, λ3,...,λn, λn+2 are all distinct then 2

13 3π sin( n+3 ) π k (µ −λ3) sin( n+3 )  5π  sin( n+3 ) π (µk−λ ) sin  5 ( n+3 )   .   .   .  ⊤ n π I n+1 + uu O  sin ( +1)  2  [ n+3 ]  RnPn ⊤ π (3.8a) O I n−1 vv  (µk−λn) sin  " + #  ( n+3 )  2    1   µk−λn+2     0     .   .       0  n+1   is an eigenvector of Hn associated to µk, k =1,..., 2 .

ii. ν1,...,ν n−1 are the zeros of (3.1b) and λ2, λ4,...,λn+1 are all distinct then 2

0 .  .  0    sin 2π  ⊤ ( n+3 ) n+1  (n+1)π  I + uu O k 2  (ν −λ2) sin[ n+3 ]  RnPn ⊤   (3.8b) O I n−1 + vv  sin 4π  " 2 # ( n+3 )  (n+1)π   (νk −λ ) sin   4 [ n+3 ]   .   .   .  n− π  sin ( 1)   [ n+3 ]  (n+1)π  (νk −λn− ) sin   1 [ n+3 ]    n−1 is an eigenvector of Hn associated to νk, k =1,..., 2 . Proof. Since both assertions can be proven in the same way, we only prove (a). Let n ∈ N be even.

i. Supposing λ1, λ3,...,λn+1 all distinct, Corollary 1 guarantees that the zeros of (3.1a) are all simple. Setting

Υ n := diag(λ , λ ,...,λ ) 2 3 5 n+1 ∆ n := diag(λ , λ ,...,λ ) 2 2 4 n

we can rewrite the matricial equation (µkIn − Hn)q = 0 as

µ I n − Υ n + (µ − λ )uu⊤ O ⊤ k 2 2 k 1 RnPn PnRnq = 0 (3.9) O µ I n − ∆ n + (µ − λ )vv⊤ " k 2 2 k n+2 #

with u, v given by (2.5b), Rn defined in (2.5c) and Pn the permutation matrix (2.5d). Thus,

µ I n − Υ n + (µ − λ )uu⊤ q = 0, k 2 2 k 1 1 ⊤ µ I n − ∆ n + (µ − λ )vv q = 0, k 2 2 k n+2  2 q  1 = P R q  q n n  2 

14 that is,

− q = α(µ I n − Υ n ) 1u, 1 k 2 2

q2 = 0

for α 6= 0 and −1 α(µkI n − Υ n ) u q = R−1P⊤ 2 2 n n 0   is a nontrivial solution of (3.9), and so an eigenvector of Hn associated to the eigenvalue µk. Since

−1 ⊤ −2 ⊤ Rn Pn = RnRn Pn ⊤ ⊤ −1 ⊤ = Rn In − xx − yy Pn ⊤ ⊤ ⊤ ⊤ −1 (3.10) = RnP n Pn In − xx − yy Pn I n + uu⊤ O ⊤  2   = RnP , n O I n + vv⊤  2  where x, y are given by (2.7), the result follows choosing α = 1.

ii. Considering λ2, λ4,...,λn+2 all distinct, the zeros of (3.1b) are all simple according to Corollary 1. Hence,

ν I n − Υ n + (ν − λ )uu⊤ O ⊤ k 2 2 k 1 RnPn PnRnq = 0 O ν I n − ∆ n + (ν − λ )vv⊤ " k 2 2 k n+2 # leads to −1 ⊤ 0 q = Rn Pn −1 , α(νkI n − ∆ n ) v  2 2  for α 6= 0, which is an eigenvector of Hn associated to the eigenvalue νk. The conclusion follows from (3.10).



Remark Recall that if n is even and some λ2k−1 is a root of (3.1a) (and a fortiori an eigenvalue of Hn) then λ2k−1 = λ2ℓ−1 for a suitable ℓ 6= k according to Corollary 1, so that

n ⊤ ⊤ I + uu O q1 RnP 2 , n O I n + vv⊤ 0  2   where u, v are given in (2.5b), Rn is the n × n matrix (2.5c), Pn is the n × n permutation matrix (2.5d), (2ℓ−1)π (2k−1)π sin n+3 ik − sin n+3 iℓ if k,ℓ 6=1 h i h i q1 =  i ,  ℓ if k =1  ik if ℓ =1 and i denotes the kth column of I n , is an eigenvector of H associated to λ ; if a λ is root of k  2 n 2k−1 2k (3.1b) then λ2k = λ2ℓ for some ℓ 6= k and

I n + uu⊤ O 0 ⊤ 2 RnPn ⊤ , O I n + vv q2  2  

15 with u, v given by (2.5b), Rn the n × n matrix (2.5c), Pn the n × n permutation matrix (2.5d) and

2ℓπ 2kπ n+2 sin n+3 ik − sin n+3 iℓ if k,ℓ 6= 2      q2 = i n+2  ℓ if k = 2   n+2 ik if ℓ = 2   is an eigenvector of Hn associated to λ2k. Analogously, if n is odd and some λ2k−1 is a root of (3.2a) then λ2k−1 = λ2ℓ−1 for a suitable ℓ 6= k and

⊤ I n+1 + uu O q 2 1 RnPn ⊤ , O I n−1 + vv 0 " 2 #   where u, v are given in (2.6b), Rn is the n × n matrix (2.6c), Pn is the n × n permutation matrix (2.6d), (2ℓ−1)π (2k−1)π sin n+3 ik − sin n+3 iℓ if k,ℓ 6=1 h i h i q1 =  i ,  ℓ if k =1  ik if ℓ =1  and ik denotes the kth column of I n+1 , is an eigenvector of Hn associated to λ2k−1; if a λ2k is a root  2 of (3.2b) then λ2k = λ2ℓ for some ℓ 6= k and

⊤ I n+1 + uu O 0 2 RnPn ⊤ , O I n−1 + vv q2 " 2 #   with u, v given by (2.6b), Rn the n × n matrix (2.6c), Pn the n × n permutation matrix (2.6d),

2ℓπ 2kπ n+1 sin n+3 ik − sin n+3 iℓ, if k,ℓ 6= 2 q     2 =  i , if k = n+1  ℓ 2  n+1 ik, if ℓ = 2  and ik denotes the kth column of I n−1 , is an eigenvector of Hn associated to λ2k. 2

4 Integer powers of Hn

The decomposition in Lemma 3 can still be used to compute explicitly the inverse of Hn.

Theorem 3 Let n ∈ N, a,b,c,d ∈ C, λk, k =1, 2,...,n+2 be given by (2.4) and Hn the n×n matrix (1.1). If λk 6=0 for every k =1, 2,...,n +2, Hn is nonsingular and: (a) n is even then n −1 ⊤ U O H = RnP 2 PnRn n n O V n  2  where Rn is the n × n matrix (2.5c), Pn is the n × n permutation matrix (2.5d),

n n ⊤ −1 −1 −1 λ1 U = I + uu diag λ3 , λ5 ,...,λn+1 − u⊤ −1 −1 −1 u · 2 2 1+λ1 diag(λ ,λ ,...,λn )  3 5 +1   (4.1a) − − − ⊤ − − − ⊤ diag λ 1, λ 1,...,λ 1 uu diag λ 1, λ 1,...,λ 1 I n + uu 3 5 n+1 3 5 n+1 2     16 and

n n n ⊤ −1 −1 −1 λ +2 V = I + vv diag λ2 , λ4 ,...,λn − v⊤ −1 −1 −1 v · 2 2 1+λn+2 diag(λ ,λ ,...,λn )  2 4   (4.1b) − − − − diag λ 1, λ 1,...,λ−1 vv⊤diag λ 1, λ 1,...,λ−1 I n + vv⊤ 2 4 n 2 4 n 2     with u, v given by (2.5b). (b) n is odd then

U n+1 O −1 2 ⊤ ⊤ Hn = RnPn Pn Rn O V n−1 " 2 # where Rn is the n × n matrix (2.6c), Pn is the n × n permutation matrix (2.6d),

⊤ − − − − λ n n 1 1 1 1 1 U +1 = I +1 + uu diag λ3 , λ5 ,...,λn , λn+2 − u⊤ −1 −1 −1 −1 u · 2 2 1+λ1 diag(λ ,λ ,...,λn ,λn )  3 5 +2    (4.2a) −1 −1 −1 −1 ⊤ −1 −1 −1 −1 ⊤ diag λ , λ ,...,λn , λn uu diag λ , λ ,...,λn , λn I n+1 + uu 3 5 +2 3 5 +2 2      and

⊤ − − − λn+1 n− n− 1 1 1 V 1 = I 1 + vv diag λ2 , λ4 ,...,λn−1 − v⊤ −1 −1 −1 v · 2 2 1+λn+1 diag(λ ,λ ,...,λn− )  2 4 1    (4.2b) −1 −1 −1 ⊤ −1 −1 −1 ⊤ diag λ , λ ,...,λn− vv diag λ , λ ,...,λn− I n−1 + vv 2 4 1 2 4 1 2      with u, v given by (2.6b).

Proof. Consider n ∈ N even, a,b,c,d ∈ C λk 6= 0, k = 1, 2,...,n + 2 and Hn nonsingular. According to Sherman-Morrison-Woodbury formula (see [10], page 19 or [9], page 424), we have

−1 ⊤ −1 −1 ⊤ −1 −1 ⊤ −1 Υ n + λ1uu = Υ n − λ1 1+ λ1u Υ n u Υ n uu Λ n , 2 2 2 2 2    where Υ n := diag (λ , λ ,...,λ ), u is given in (2.5a) and U n is the matrix (4.1a). Similarly, we 2 3 5 n+1 2 obtain −1 ⊤ −1 −1 ⊤ −1 −1 ⊤ −1 ∆ n + λn+2vv = ∆ n − λn+2 1+ λn+2v ∆ n v ∆ n vv ∆ n , 2 2 2 2 2   where ∆ n := diag (λ , λ ,...,λ ), v is given in (2.5a) and V n is the matrix (4.1b). Hence, the 2 2 4 n 2 decomposition of Lemma 3, 8.5b of [9] (page 88), identity (3.10) and

−1 −2 PnRn = PnRn Rn ⊤ ⊤ −1 = Pn In − xx − yy Rn I n + uu⊤ O 2  = PnRn, O I n + vv⊤  2  where x, y are given by (2.7), establish the thesis in (a). The proof of (b) is analogous and so will be omitted. 

Remark Note that if n is even, λk 6= 0 for all k =2, 3,...,n + 1,

n 2 (2j+1)π n 2 2jπ 2 2 λ sin n+3 λ sin n+3 1 6= −1 and n+2 6= −1 2 π λh j i 2 π λ j  sin j=1 2 +1 sin j=1 2 n+3 X n+3 X     17 then Hn is nonsingular. Similarly, if n is odd, λk 6= 0 for all k =2, 3,...,n,n + 2,

n+1 (2j+1)π n−1 2jπ 2 2 2 2 λ sin n+3 λ sin n+3 1 6= −1 and n+1 6= −1 2 π λh j i 2 (n+1)π λ j  sin j=1 2 +1 sin j=1 2 n+3 X n+3 X   h i then Hn is a nonsingular matrix.

Let us point out that the previous Theorems 1 and 2 lead to an immediate eigenvalue decomposition with orthogonal eigenvector matrix for any real anti-heptadiagonal persymmetric Hankel matrix having all the eigenvalues distinct. Indeed, admitting Hn in (1.1) real with its eigenvalues all different it is well- known that the set of its eigenvectors is orthogonal (see, for instance, [6] page 109); hence, normalizing each of them we can obtain an orthonormal set, and so an orthogonal eigenvector matrix (see [18], page 257). However, it is not guaranteed, in general, that the eigenvalues of Hn are all different even assuming the distinctness of every λ2k−1 and all λ2k, as the following examples show: 01000 1 0 1 0 10000 0 1 0 1 H = or H =  00000  . 4  1 0 1 0  5  00001   0 1 0 1       00010        It is possible to establish an alternative diagonalization for Hn avoiding its eigenvalues. The following auxiliary result is the central key in its accomplishment.

Lemma 4 Let n ∈ N, a,b,c,d ∈ C, λk, k =1, 2,...,n +2 be given by (2.4) and Hn the n × n matrix (1.1). (a) If n is even and:

i. u is given in (2.5b) then the eigenvalues of

2 π 4 sin n+3 diag (λ , λ ,...,λ )+ diag (λ − λ , λ − λ ,...,λ − λ ) uu⊤ (4.3a) 3 5 n+1 n +3  1 3 1 5 1 n+1 are the zeros of the polynomial

n n n 2 4 2 (2j + 1)π 2 F (t)= (t − λ )+ sin2 (λ − λ ) (t − λ ). (4.3b) 2j+1 n +3 n +3 2j+1 1 2m+1 j=1 j=1   m=1 Y X mY=6 j

Moreover, if λ , λ ,...,λ are all real and distinct then the eigenvalues φ , φ ,...,φ n of (4.3b) 1 3 n+1 1 2 2 are all simple and 3π (λ3−λ1) sin( n+3 ) λ3−φk (λ −λ ) sin 5π  5 1 ( n+3 )  f := λ5−φk (4.3c) k  .   .   (n+1)π  λn −λ  ( +1 1) sin[ n+3 ]     λn+1−φk   n  is an eigenvector associated to φk, k =1,..., 2 . ii. v is given in (2.5b) then the eigenvalues of

2 π 4 sin n+3 diag (λ , λ ,...,λ )+ diag (λ − λ , λ − λ ,...,λ − λ ) vv⊤ (4.3d) 2 4 n n +3  n+2 2 n+2 4 n+2 n

18 are the zeros of the polynomial

n n n 2 4 2 2jπ 2 G(t)= (t − λ )+ sin2 (λ − λ ) (t − λ ). (4.3e) 2j n +3 n +3 2j n+2 2m j=1 j=1   m=1 Y X mY=6 j

Furthermore, if λ , λ ,...,λ are all real, nonzero and different then the eigenvalues ψ , ψ ,...,ψ n 2 4 n+2 1 2 2 of (4.3e) are all simple and 2π n (λ2−λ +2) sin( n+3 ) λ2−ψk π (λ −λn ) sin 4  4 +2 ( n+3 )  g := λ4−ψk (4.3f) k  .   .   nπ  λn−λn  ( +2) sin( n+3 )     λn−ψk   n  is an eigenvector associated to ψk, k =1,..., 2 .

(b) If n is odd and:

i. u is given in (2.6b) then the eigenvalues of

2 π 4 sin n+3 diag (λ , λ ,...,λ , λ )+ diag (λ − λ , λ − λ ,...,λ − λ , λ − λ ) uu⊤ 3 5 n n+2 n +3  1 3 1 5 1 n 1 n+2 (4.4a) are the zeros of the polynomial

n+1 n+1 n+1 2 4 2 (2j + 1)π 2 F (t)= (t − λ )+ sin2 (λ − λ ) (t − λ ) (4.4b) 2j+1 n +3 n +3 2j+1 1 2m+1 j=1 j=1   m=1 Y X mY=6 j

Moreover, if λ1, λ3,...,λn, λn+2 are all real and distinct then the eigenvalues φ1, φ2,...,φ n+1 of 2 (4.4b) are all simple and 3π (λ3−λ1) sin( n+3 ) λ3−φk (λ −λ ) sin 5π  5 1 ( n+3 )  λ5−φk  .  fk :=  .  (4.4c) nπ  (λn−λ ) sin   1 ( n+3 )   λn−φk   (n+2)π  λn −λ  ( +2 1) sin[ n+3 ]     λn+2−φk   n  is an eigenvector associated to φk, k =1,..., 2 . ii. v is given in (2.6b) then the eigenvalues of

2 (n+1)π 4 sin n+3 diag (λ , λ ,...,λ )+ diag (λ − λ , λ − λ ,...,λ − λ ) vv⊤ 2 4 n−1 nh+3 i n+1 2 n+1 4 n+1 n−1 (4.4d) are the zeros of the polynomial

n−1 n−1 n−1 2 4 2 2jπ 2 G(t)= (t − λ )+ sin2 (λ − λ ) (t − λ ). (4.4e) 2j n +3 n +3 2j n+1 2m j=1 j=1   m=1 Y X mY=6 j

19 Furthermore, if λ2, λ4,...,λn−1, λn+1 are all real and different then the eigenvalues ψ1, ψ2,...,ψ n−1 2 of (4.4e) are all simple and

2π n (λ2−λ +1) sin( n+3 ) λ2−ψk π (λ −λn ) sin 4  4 +1 ( n+3 )  g := λ4−ψk (4.4f) k  .   .   (n−1)π  λn− −λn  ( 1 +1) sin[ n+3 ]     λn−1−ψk    n−1 is an eigenvector associated to ψk, k =1,..., 2 . Proof. Since both assertions can be proven in the same way, we only prove (a). Consider n ∈ N, a,b,c,d ∈ C and λk, k = 1, 2,...,n + 2 given by (2.4). The proof follows the same steps of the ones displayed in Section 3.

i. Setting

4 sin2 π n+3 ⊤ Φ n := diag (λ3, λ5,...,λn+1)+ diag (λ1 − λ3, λ1 − λ5,...,λ1 − λn+1) uu 2 n +3  with u given in (2.5b), we have

det(tI n − Φ n ) = det [diag (t − λ ,t − λ ,...,t − λ )] · 2 2 3 5 n+1 2 π 4 sin ( n ) λ −λn 1 − +3 u⊤diag λ1−λ3 , λ1−λ5 ,..., 1 +1 u n+3 t−λ3 t−λ5 t−λn+1 n  n  n   2 2 2 4 2 (2j+1)π = (t − λ2j+1)+ n+3 sin n+3 (λ2j+1 − λ1) (t − λ2m+1) =: F (t). j=1 j=1 m=1 Y X h i mY=6 j

Supposing λ1, λ3,...,λn+1 all real, distinct and arranged in increasing order by some bijection ϕ, i.e. λϕ(1) < λϕ(3) <...<λϕ(n+1), we get

n n n 2 +1 2 +1 2 +1 4 2 ϕ(2j−1)π F (t)= t − λϕ(2j−1) + n+3 sin n+3 λϕ(2j−1) − λϕ(2ϑ−1) t − λϕ(2m−1) j=1 j=1 m=1 jY=6 ϑ   jX=6 ϑ h i   mY=6 ϑ,j  

6 6 n for some ϑ satisfying ϕ(2ϑ − 1)=1,1 ϑ 2 + 1 which yields F λϕ(2k−1) F λϕ(2k+1) < 0 for n all k = 1,..., . Hence, the zeros φ1, φ2,...,φ n of F (t) are all simple. From Theorem 5 of [3] 2 2     (see [3], page 41), it follows that fk in (4.3c) is an eigenvector of Φ n associated to the eigenvalue n 2 φk, k =1,..., 2 . ii. Analogously, putting

4 sin2 π n+3 ⊤ Ψ n := diag (λ2, λ4,...,λn)+ diag (λn+2 − λ2, λn+2 − λ4,...,λn+2 − λn) vv 2 n +3  with v given in (2.5b), we obtain

det(tI n − Ψ n ) = det [diag (t − λ ,t − λ ,...,t − λ )] · 2 2 2 4 n 2 π 4 sin ( n ) λn −λ λn −λ λn −λn 1 − +3 v⊤diag +2 2 , +2 4 ,..., +2 v n+3 t−λ2 t−λ4 t−λn     20 n n n 2 2 2 4 2 2jπ = (t − λ2j )+ n+3 sin n+3 (λ2j − λn+2) (t − λ2m) =: G(t). j=1 j=1 m=1 Y X   mY=6 j

Assuming λ2, λ4,...,λn all real, distinct and arranged in increasing order by some bijection ρ, that is λρ(2) < λρ(4) <...<λρ(n), we have

n n n 2 +1 2 +1 2 +1 4 2 ρ(2j)π G(t)= t − λρ(2j) + n+3 sin n+3 λρ(2j) − λρ(2υ) t − λρ(2m) j=1 j=1 m=1 jY=6 υ   jX=6 υ h i   mY=6 υ,j   6 6 n for some υ such that ρ(2υ)= n +2, 1 υ 2 + 1 which implies G λρ(2k) G λρ(2k+2) < 0 for n all k = 1,..., . Thus, the zeros ψ1, ψ2,...,ψ n of G(t) are all simple. Hence, Theorem 5 of [3] 2 2     (see [3], page 41), ensures that gk in (4.3f) is an eigenvector of Ψ n associated to the eigenvalue n 2 ψk, k =1,..., 2 . 

Next, we announce a diagonalization for Hn in (1.1) discarding its eigenvalues.

Theorem 4 Let n ∈ N, a,b,c,d ∈ R, λk, k = 1, 2,...,n +2 be given by (2.4) and Hn the n × n anti-heptadiagonal persymmetric Hankel matrix (1.1).

(a) If n is even, λ1, λ3,...,λn+1 are all distinct, λ2, λ4,...,λn+2 are all distinct then

−1 F n O F n O ⊤ 2 n n 2 −1 Hn = RnPn diag φ1, φ2,...,φ , ψ1, ψ2,...,ψ −1 PnRn (4.5a) O G n 2 2 O G n  2  " 2 #  where φ , φ ,...,φ n are the zeros of (4.3b), ψ , ψ ,...,ψ n are the zeros of (4.3e), 1 2 2 1 2 2

(2k+1)π (λ2k+1 − λ1) sin n+3 F n := (4.5b) 2  λ2k+1 − φhℓ i  k,ℓ  2kπ  (λ2k − λn+2) sin n+3 G n := (4.5c) 2  λ2k − ψℓ   k,ℓ   Rn is the n × n matrix (2.5c) and Pn is the n × n permutation matrix (2.5d).

(b) If n is odd, λ1, λ3,...,λn, λn+2 are all distinct, λ2, λ4,...,λn−1, λn+1 are all distinct then

−1 F n+1 O F n+1 O 2 2 ⊤ −1 Hn = RnPn diag φ1, φ2,...,φ n+1 , ψ1, ψ2,...,ψ n−1 −1 Pn Rn O G n−1 2 2 O G n− " # " 1 # 2   2 (4.6a) where φ1, φ2,...,φ n+1 are the zeros of (4.4b), ψ1, ψ2,...,ψ n−1 are the zeros of (4.4e), 2 2

(2k+1)π (λ2k+1 − λ1) sin n+3 F n+1 := (4.6b) 2  λ2k+1 − φhℓ i k,ℓ  2kπ  (λ2k − λn+1) sin n+3 G n−1 := (4.6c) 2  λ2k − ψℓ   k,ℓ   Rn is the n × n matrix (2.6c) and Pn is the n × n permutation matrix (2.6d).

21 Proof. Consider n ∈ N, a,b,c,d ∈ R and λk, k =1, 2,...,n + 2 given by (2.4). (a) According to Lemma 3 and

2 π 4 sin ( n ) I n − +3 uu⊤ O 2 ⊤ 2 n+3 PnRnPn = 2 π 4 sin ( n )  O I n − +3 vv⊤  2 n+3   we get

⊤ ⊤ diag (λ3, λ5,...,λn+1)+ λ1uu O 2 ⊤ −1 Hn = RnPn ⊤ PnRnPn PnRn O diag (λ , λ ,...,λn)+ λn vv  2 4 +2  2 π 4 sin ( n ) diag (λ , λ ,...,λ )+ λ uu⊤ I n − +3 uu⊤ ⊤ 3 5 n+1 1 2 n+3 = RnP n     O   O 2 π −1 4 sin ( n ) PnR diag (λ , λ ,...,λ )+ λ vv⊤ I n − +3 vv⊤ n 2 4 n n+2 2 n+3     2 π   4 sin ( n+3 ) ⊤ ⊤ diag (λ3, λ5,...,λn+1)+ diag (λ1 − λ3, λ1 − λ5,...,λ1 − λn+1) uu = RnPn n+3 " O O −1 2 π PnR . 4 sin ( n+3 ) ⊤ n diag (λ2, λ4,...,λn)+ n+3 diag (λn+2 − λ2, λn+2 − λ4,...,λn+2 − λn) vv # From Lemma 4, − F n diag φ , φ ,...,φ n F 1 2 1 2 2 n and  − G n diag ψ , ψ ,...,ψ n G 1 2 1 2 2 n are eigenvalue decompositions for (4.3a) and (4.3d), respectively (see [6], page 85). Thus,

−1 F n O F n O ⊤ 2 n n 2 −1 Hn = RnPn diag φ1, φ2,...,φ , ψ1, ψ2,...,ψ −1 PnRn . O G n 2 2 O G n  2  " 2 #  The proof of (b) is analogous being so omitted. 

Setting

3π 5π (n + 1)π Ξ n := diag (λ3 − λ1) sin , (λ5 − λ1) sin ,..., (λn+1 − λ1) sin 2 n +3 n +3 n +3        and 1 Q n := 2 λ − φ  2k+1 ℓ k,ℓ we have F n = Ξ n Q n . Straight computations permit us to verify that the inverse of the alternant 2 2 2 matrix Q n is 2 n n 2 2 (φk − λ2j+1) (φj − λ2ℓ+1)  j=1 j=1  n j=6 k −1 2 Q n = (−1) Qn n Q . 2  2 2     (φk − φj ) (λ2ℓ+1 − λ2j+1)  j=1 j=1     jQ=6 k jQ=6 ℓ k,ℓ  

22 Hence,

n n 2 2 (φk − λ2j+1) (φj − λ2ℓ+1)  j=1 j=1  − n j=6 k −1 −1 1 2 Q Q F n = Q n Ξ n = (−1) n n (4.7) 2 2 2  2 2   (2ℓ+1)π  (λ2ℓ+1 − λ1) sin n+3 (φk − φj ) (λ2ℓ+1 − λ2j+1)  j=1 j=1     h i jQ=6 k jQ=6 ℓ k,ℓ   Similarly,

n n 2 2 (ψk − λ2j ) (ψj − λ2ℓ)  j=1 j=1  n j=6 k −1 2 G n = (−1) Q n Q n (4.8) 2  2 2   2ℓπ  (λ2ℓ − λn+2) sin n+3 (ψk − ψj ) (λ2ℓ − λ2j )  j=1 j=1       jQ=6 k jQ=6 ℓ k,ℓ  n+1 n+1  2 2 (φk − λ2j+1) (φj − λ2ℓ+1) j=1 j=1 n   −1 +1 j=6 k n 2 Q Q F +1 = (−1) n+1 n+1 (4.9) 2    (2ℓ+1)π 2 2  (λ2ℓ+1 − λ1) sin (φk − φj ) (λ2ℓ+1 − λ2j+1)  n+3   j=1 j=1   h i jQ=6 k jQ=6 ℓ k,ℓ  n−1 n−1  2 2 (ψk − λ2j ) (ψj − λ2ℓ) j=1 j=1 n−   −1 1 j=6 k n− 2 Q Q G 1 = (−1) n−1 n−1 . (4.10) 2  2 2   2ℓπ  (λ2ℓ − λn+1) sin (ψk − ψj ) (λ2ℓ − λ2j )   n+3   j=1 j=1     jQ=6 k jQ=6 ℓ k,ℓ   The ultimate statement below gives us a formula to compute integer powers of Hn in (1.1) only at the expense of the zeros of F (t) and G(t) presented in Lemma 4.

Corollary 2 Let n,m ∈ N, a,b,c,d ∈ R, λk, k = 1, 2,...,n +2 be given by (2.4) and Hn the n × n anti-heptadiagonal persymmetric Hankel matrix (1.1).

(a) If n is even, λ1, λ3,...,λn+1 are all distinct, λ2, λ4,...,λn+2 are all distinct then

F n O m ⊤ 2 m m m m m m Hn = RnPn diag φ1 , φ2 ,...,φ n , ψ1 , ψ2 ,...,ψ n O G n 2 2  2   −1  (4.11) F n O I n + uu⊤ O 2 2 −1 ⊤ PnRn O G n O I n + vv " 2 #  2  where φ , φ ,...,φ n are the zeros of (4.3a), ψ , ψ ,...,ψ n are the zeros of (4.3e), F n is the n × n 1 2 2 1 2 2 2 2 2 n n −1 n n −1 n n matrix (4.5b), G n is the × matrix (4.5c), F n is the × matrix (4.7), G n is the × matrix 2 2 2 2 2 2 2 2 2 (4.8), Rn is the n × n matrix (2.5c), Pn is the n × n permutation matrix (2.5d) and u, v are given by (2.5b). Moreover, if Hn is nonsingular then (4.11) holds for any integer m.

23 (b) If n is odd, λ1, λ3,...,λn, λn+2 are all distinct, λ2, λ4,...,λn−1, λn+1 are all distinct then

F n+1 O m 2 m m m m m m Hn = RnPn diag φ1 , φ2 ,...,φ n+1 , ψ1 , ψ2 ,...,ψ n−1 O G n−1 2 2 " 2 #   (4.12) −1 ⊤ F n+1 O I n+1 + uu O 2 2 P⊤R⊤ −1 n− ⊤ n n O G n−1 O I 1 + vv " 2 # " 2 # where φ1, φ2,...,φ n+1 are the zeros of (4.3a), ψ1, ψ2,...,ψ n−1 are the zeros of (4.3e), F n+1 is the 2 2 2 − n+1 n+1 n− n−1 n−1 1 n+1 n+1 2 × 2 matrix (4.6b), G 1 is the 2 × 2 matrix (4.6c), F n+1 is the 2 × 2 matrix (4.9), 2 2 −1 n−1 n−1 G n−1 is the 2 × 2 matrix (4.10), Rn is the n × n matrix (2.6c), Pn is the n × n permutation 2 matrix (2.6d) and u, v are defined in (2.6b). Furthermore, if Hn is nonsingular then (4.12) is valid for every integer m.

Proof. According to Theorem 4, it suffices to observe that for n even

n ⊤ −1 I + uu O PnR = 2 PnRn, n O I n + vv⊤  2  and for n odd, ⊤ I n+1 + uu O ⊤ −1 2 ⊤ ⊤ Pn Rn = ⊤ Pn Rn . O I n−1 + vv " 2 # 

Remark It is worth pointing out that all results presented here for matrices Hn having the form (1.1) are still valid for anti-pendiagonal persymmetric Hankel matrices or anti-tridiagonal persymmetric Hankel matrices, that is to say, for matrices Hn in (1.1) such that a =0 or a = b = 0.

Acknowledgements

This work is a contribution to the Project UID/GEO/04035/2013, funded by FCT - Funda¸c˜ao para a Ciˆencia e a Tecnologia, Portugal.

References

[1] D. Bini and M. Capovani, Spectral and computational properties of band symmetric Toeplitz matrices, Appl. 52/53 (1983) 99–126. [2] A. Bultheel and M. Van Barel, Linear Algebra: Rational Approximation and Orthogonal Polyno- mials, Studies in Computational Mathematics 6, North-Holland, 1997. [3] J.R. Bunch, C.P. Nielsen, D.C. Sorensen, Rank-one modification of the symmetric eigenproblem, Numer. Math. 31 (1978) 31–48. [4] B.N. Datta, C.R. Johnson, M.A. Kaashoek, R.J. Plemmons, E.D. Sontag, Linear Algebra in Signals, Systems, and Control, SIAM, 1988. [5] D. Fasino, Spectral and structural properties of some pentadiagonal symmetric matrices, Calcolo 25(4) (1988) 301–310. [6] W. Ford, Numerical Linear Algebra with Applications Using MATLAB, Academic Press, 2015.

24 [7] J. Guti´errez-Guti´errez, Powers of complex persymmetric or skew-persymmetric anti-tridiagonal matrices with constant anti-diagonals, Appl. Math. Comput. 217 (2011) 6125–6132. [8] J. Guti´errez-Guti´errez, Eigenvalue decomposition for persymmetric Hankel matrices with at most three non-zero anti-diagonals, Appl. Math. Comput. 234 (2014) 333–338. [9] D.A. Harville, Matrix Algebra From a Statistician’s Perspective, Springer-Verlag New York, 1997.

[10] R.A. Horn and C.R. Johnson, Matrix Analysis, 2nd ed., Cambridge University Press, 2013. [11] J. Lita da Silva, On anti-pentadiagonal persymmetric Hankel matrices with perturbed corners, Comput. Math. Appl. 72 (2016), 415–426. [12] T. Kailath, Linear Systems, Prentice-Hall, 1980. [13] K.S. Miller, On the inverse of the sum of matrices, Math. Mag. 54(2) (1981) 67–72. [14] V. Olshevsky, M. Stewart, Stable factorization for Hankel and Hankel-like matrices, Numer. Linear Algebra Appl. 8 (2001) 401–434. [15] Pissanetsky S. Technology. London: Academic Press, 1984. [16] J. Rimas, Integer powers of real odd order skew-persymmetric anti-tridiagonal matrices with constant anti-diagonals (antitridiagn(a, c, −a), a ∈ R \{0}, c ∈ R), Appl. Math. Comput. 219 (2013) 7075–7088. [17] J. Rimas, Integer powers of real even order anti-tridiagonal Hankel matrices of the form antitridiagn(a, c, −a), Appl. Math. Comput. 225 (2013) 204–215. [18] G. Williams, Linear Algebra with Applications, 8th ed., Jones & Bartlett Learning, Burlington, MA, 2014. [19] Honglin Wu, On computing of arbitrary positive powers for one type of anti-tridiagonal matrices of even order, Appl. Math. Comput. 217 (2010) 2750–2756. [20] Qingxiang Yin, On computing of arbitrary positive powers for anti-tridiagonal matrices of even order, Appl. Math. Comput. 203 (2008) 252–257.

25