Signatures of hermitian forms, positivity, and an answer to a question of Procesi and Schacher

Signatures of hermitian forms, positivity, and an answer to a question of Procesi and Schacher

Journal of Algebra 508 (2018) 339–363 Contents lists available at ScienceDirect Journal of Algebra www.elsevier.com/locate/jalgebra Signatures of h...

506KB Sizes 0 Downloads 18 Views

Journal of Algebra 508 (2018) 339–363

Contents lists available at ScienceDirect

Journal of Algebra www.elsevier.com/locate/jalgebra

Signatures of hermitian forms, positivity, and an answer to a question of Procesi and Schacher Vincent Astier, Thomas Unger ∗ School of Mathematics and Statistics, University College Dublin, Belfield, Dublin 4, Ireland

a r t i c l e

i n f o

Article history: Received 5 January 2017 Available online 7 May 2018 Communicated by Louis Rowen MSC: 16K20 11E39 13J30 Keywords: Central simple algebra Involution Formally real field Hermitian form Signature Positivity Sum of hermitian squares

a b s t r a c t Using the theory of signatures of hermitian forms over algebras with involution, developed by us in earlier work, we introduce a notion of positivity for symmetric elements and prove a noncommutative analogue of Artin’s solution to Hilbert’s 17th problem, characterizing totally positive elements in terms of weighted sums of hermitian squares. As a consequence we obtain an earlier result of Procesi and Schacher and give a complete answer to their question about representation of elements as sums of hermitian squares. © 2018 Elsevier Inc. All rights reserved.

1. Introduction We use the theory of signatures of hermitian forms, a tool we developed and studied in [4], [5] and [3], to introduce a natural notion of positivity for symmetric elements * Corresponding author. E-mail addresses: [email protected] (V. Astier), [email protected] (T. Unger). https://doi.org/10.1016/j.jalgebra.2018.05.004 0021-8693/© 2018 Elsevier Inc. All rights reserved.

340

V. Astier, T. Unger / Journal of Algebra 508 (2018) 339–363

in an algebra with involution, inspired by the theory of quadratic forms; signatures of one-dimensional hermitian forms over algebras with an involution can take values outside of {−1, 1} and it is therefore natural to single out those symmetric elements whose associated hermitian form has maximal signature at a given ordering. We call such elements maximal at the ordering and characterize the elements that are maximal at all orderings in terms of weighted sums of hermitian squares, thus obtaining an analogue of Artin’s solution to Hilbert’s 17th problem for algebras with involution, cf. Section 3. The proof is obtained via signatures, allowing us to use the hermitian version of Pfister’s local-global principle. This provides a short and conceptual argument, based on torsion in the Witt group. Procesi and Schacher [19] already considered such a noncommutative version of Artin’s theorem in this context, using a notion of positivity based on involution trace forms that seems to go back to Albert (e.g. [1]) and Weil [23]. They showed that every totally positive element (in their sense) in an algebra with involution is a sum of squares of symmetric elements, and thus of hermitian squares, with weights, cf. [19, Theorem 5.4]. They also asked if these weights could be removed [19, p. 404]. The answer to this question is in general no, as shown in [12]. Our approach via signatures makes it possible to obtain the sum of hermitian squares version of their theorem as a consequence of Theorem 3.6. It also allows us to single out the set of orderings relevant to their question (the non-nil orderings) and to rephrase it in a natural way, which can then be fully answered (Theorem 4.19). 2. Algebras with involution and signatures of hermitian forms We present the notation and main tools used in this paper and refer to the standard references [13], [14], [15] and [22] as well as [4] and [5] for the details. 2.1. Algebras with involution, hermitian forms For a ring A, an involution σ on A and ε ∈ {−1, 1}, we denote the set of ε-symmetric elements of A with respect to σ by Symε (A, σ) = {a ∈ A | σ(a) = εa}. We also denote the set of invertible elements of A by A× and let Symε (A, σ)× := Symε (A, σ) ∩ A× . Let F be a field of characteristic different from 2. We denote by W (F ) the Witt ring of F , by XF the space of orderings of F , and by FP a real closure of F at an ordering P ∈ XF . We allow for the possibility that F is not formally real, i.e. that XF = ∅. By an F -algebra with involution we mean a pair (A, σ) where A is a finite-dimensional simple F -algebra with centre a field K, equipped with an involution σ : A → A, such that F = K ∩ Sym(A, σ). Observe that dimF K  2. We say that σ is of the first kind if K = F and of the second kind (or of unitary type) otherwise. Involutions of the first kind can be further subdivided into those of orthogonal type and those of symplectic type, depending on the dimension of Sym(A, σ). We let ι = σ|K and note that ι = idF if σ

V. Astier, T. Unger / Journal of Algebra 508 (2018) 339–363

341

is of the first kind. If A is a division algebra, we call (A, σ) an F -division algebra with involution. Let (A, σ) be an F -algebra with involution. It follows from the structure theory of F -algebras with involution that A is isomorphic to a full matrix algebra M (D) for a unique  ∈ N and an F -division algebra D (unique up to isomorphism) which is equipped with an involution ϑ of the same kind as σ, cf. [14, Thm. 3.1]. For B = (bij ) ∈ M (D) we let ϑt (B) = (ϑ(bji )). We denote Brauer equivalence by ∼, isomorphism by ∼ = and isometry of forms by . For ε ∈ {−1, 1} we write Wε (A, σ) for the Witt group of Witt equivalence classes of nonsingular ε-hermitian forms, defined on finitely generated right A-modules. Note that Wε (A, σ) is a W (F )-module. For a nonsingular ε-hermitian form h over (A, σ) the notation h ∈ Wε (A, σ) signifies that h is identified with its Witt class in Wε (A, σ). For a1 , . . . , ak ∈ F the notation a1 , . . . , ak  stands for the quadratic form (x1 , . . . , xk ) k ∈ F k → i=1 ai x2i ∈ F , as usual, whereas for a1 , . . . , ak in Symε (A, σ) the notation a1 , . . . , ak σ stands for the diagonal ε-hermitian form k    (x1 , . . . , xk ), (y1 , . . . , yk ) ∈ Ak × Ak → σ(xi )ai yi ∈ A. i=1

In each case, we call k the dimension of the form. In this paper, we are mostly interested in hermitian forms (ε = 1) and only occasionally in skew-hermitian forms (ε = −1). When ε = 1, we write Sym(A, σ) and W (A, σ) instead of Sym1 (A, σ) and W1 (A, σ), respectively. Let h : M × M → A be a hermitian form over (A, σ). We sometimes write (M, h) instead of h. The rank of h, rk(h), is the rank of the A-module M . Observe that if h is diagonal then rk(h) =  dim(h). The set of elements represented by h is denoted by D(A,σ) (h) := {u ∈ Sym(A, σ) | ∃x ∈ M such that h(x, x) = u}. We denote by Int(u) the inner automorphism determined by u ∈ A× , where Int(u)(x) := uxu−1 for x ∈ A. Remark 2.1. If F is not formally real, many results in this paper are trivially true since W (A, σ) is torsion in this case (see [17, Theorem 4.1] and note that this theorem, being a reformulation of [17, Theorem 3.2], is actually valid for any field of characteristic not 2). Remark 2.2. The class of F -algebras with involution is often enlarged to include algebras with unitary involution with centre F × F , cf. [14, 2.B]. With this convention, it becomes stable under scalar extension. We do not consider this case in the main body of the paper, since the properties we are interested in become trivial for them. For the details we refer to Appendix B.

342

V. Astier, T. Unger / Journal of Algebra 508 (2018) 339–363

2.2. Morita theory For the remainder of the paper we fix some field F of characteristic not 2 and some F -algebra with involution (A, σ), where dimK A = m = n2 and A ∼ = M (D) for some F -division algebra D which is equipped with an involution ϑ of the same kind as σ. Recall that the integer n is called the degree of A, deg A. By [14, 4.A], there exists ε ∈ {−1, 1} and an invertible matrix Φ ∈ M (D) such that ϑ(Φ)t = εΦ and (A, σ) ∼ = (M (D), adΦ ), where adΦ = Int(Φ−1 ) ◦ ϑt . (In fact, Φ is the Gram matrix of an ε-hermitian form over (D, ϑ).) Note that Φ is only defined up to a factor in F × , with adΦ = adλΦ for all λ ∈ F × , and that ε = 1 when σ and ϑ are of the same type. We fix an isomorphism of F -algebras with involution f : (A, σ) → (M (D), adΦ ). Lemma 2.3. We may choose ϑ above such that ε = 1, except when A ∼ = M (F ) with  even and σ symplectic, in which case (D, ϑ, ε) = (F, idF , −1). Proof. We consider all possible cases, with reference to [14, Corollary 2.8] for involutions of the first kind. Case 1: σ, and thus ϑ, of the second kind. In this case, if ε = −1, let u ∈ K × be such that ϑ(u) = −u and scale by u, i.e. replace Φ by uΦ and note that Int(u) ◦ ϑ = ϑ. Case 2: σ, and thus ϑ, of the first kind and deg D even. Then D can be equipped with both orthogonal and symplectic involutions and so we may choose ϑ to be of the same type as σ so that ϑ(Φ)t = Φ. Case 3: σ, and thus ϑ, of the first kind, deg D odd and deg A also odd. In this case, D = F , ϑ = idF , A is split (i.e. A ∼ F ) and σ must be orthogonal. Thus ε = 1 since ϑ and σ are both orthogonal. Case 4: σ, and thus ϑ, of the first kind, deg D odd and deg A even. In this case, D = F , ϑ = idF and A is split. If σ is orthogonal, then ε = 1 since ϑ and σ are both orthogonal. If σ is symplectic, then ε = −1. 2 Given an F -algebra with involution (B, τ ) we denote by Hermε (B, τ ) the category of ε-hermitian forms over (B, τ ) (possibly singular), cf. [13, p. 12]. The isomorphism f trivially induces an equivalence of categories f∗ : Herm(A, σ) −→ Herm(M (D), adΦ ). Furthermore, the F -algebras with involution (A, σ) and (D, ϑ) are Morita equivalent, cf. [13, Chapter I, Theorem 9.3.5]. In this paper we make repeated use of a particular Morita equivalence between (A, σ) and (D, ϑ), following the approach in [18] (see also [4, §2.4] for the case of nonsingular forms and [4, Proposition 3.4] for a justification of why using this equivalence is as good as using any other equivalence), namely:

Herm(A, σ)

f∗

Herm(M (D), adΦ )

s

Hermε (M (D), ϑt )

g

Hermε (D, ϑ), (2.1)

V. Astier, T. Unger / Journal of Algebra 508 (2018) 339–363

343

where s is the scaling by Φ Morita equivalence, given by (M, h) → (M, Φh) and g is the collapsing Morita equivalence, given by (M, h) → (Dk , b), where k is the rank of M as M (D)-module. Under the isomorphism M ∼ = (D )k , h can be identified with the form (Mk, (D), Bϑt ) for some matrix B ∈ Mk (D) that satisfies ϑt (B) = εB and we take for b the ε-hermitian form whose Gram matrix is B. Note that Bϑt (X, Y ) := ϑ(X)t BY for all X, Y ∈ Mk, (D). 2.3. Signatures of hermitian forms We defined signatures of nonsingular hermitian forms over (A, σ) in [4], inspired by [8], and gave a more concise presentation in [5, §2], which we will follow in this section and to which we refer for the details. (We called them H-signatures in [4] and [5] to differentiate them from the signatures in [8].) Let P ∈ XF and consider the sequence of group morphisms (cf. [5, Diagram (1)])

W (A, σ)

rP

W (A ⊗F FP , σ ⊗ id)

μP ∼ =

WεP (DP , ϑP )

signP

Z,

(2.2)

where • rP is induced by the canonical extension of scalars map,  √ • (DP , ϑP ) is one of (FP , idFP ), (FP ( −1), ), ((−1, −1)FP , ), (−1, −1)FP ×  (−1, −1)FP ,  , (FP ×FP , ), where  denotes the exchange involution and denotes conjugation (in the first three cases we often simply write instead of ϑP , even when DP = FP ), • A ⊗F FP is a matrix algebra over DP , • μP is any isomorphism induced by Morita equivalence (for example, the isomorphism induced by (2.1) with (A ⊗F FP , σ ⊗ id) in the role of (A, σ)), • If σ and ϑP are unitary, we are free to take εP = 1 or εP = −1, cf. [4, Lemma 2.1]. If σ and ϑP are of the first kind, we have εP = 1 if σ and ϑP are of the same type and εP = −1 otherwise.   • signP is zero if (DP , ϑP ) is one of (−1, −1)FP × (−1, −1)FP ,  , (FP × FP , ) or if εP = −1 and (DP , ϑP ) is one of (FP , idFP ), ((−1, −1)FP , ), • signP is the Sylvester signature at the unique ordering of FP if (DP , ϑP ) = √ (FP ( −1), ) or if εP = 1 and (DP , ϑP ) is one of (FP , idFP ), ((−1, −1)FP , ). Diagram (2.2) defines a morphism of groups sμP : W (A, σ) → Z. The map μP is not canonical and a different choice may at most result in multiplying sμP by −1. We define the set of nil-orderings of (A, σ) as follows: Nil[A, σ] := {P ∈ XF | sμP = 0}

344

V. Astier, T. Unger / Journal of Algebra 508 (2018) 339–363

and note that it does not depend on the choice of μP , but only on the Brauer class of A and the type of σ. For convenience we also introduce F := XF \ Nil[A, σ], X which does not indicate the dependence on (A, σ) in order to avoid cumbersome notation. Given P ∈ XF , we define signηP , the signature at P of nonsingular hermitian forms over (A, σ), as follows (see also [4] and [5]): (i) if P ∈ Nil[A, σ], we let signηP = 0; F , signη will be either sμ or −sμ . In [4, Theorem 6.4] we proved that (ii) if P ∈ X P P P there exists a finite tuple η = (η1 , . . . , ηt ) of nonsingular hermitian forms (which F , can all be chosen to be diagonal of dimension 1) such that for every Q ∈ X sμQ (η) = (0, . . . , 0). Using η as provided by this theorem, let i be the least integer such that sμP (ηi ) = 0. We choose signηP ∈ {−sμP , sμP } such that signηP ηi > 0. In [5, Proposition 3.2] we showed that the tuple η (called a tuple of reference forms for (A, σ)) can be replaced by a single diagonal hermitian form (called a reference form for (A, σ)) which may have dimension greater than one. Remark 2.4. If η = (η1 , . . . , ηt ) is a tuple of reference forms for (A, σ), then η  = (1σ , η1 , . . . , ηt ) is also a tuple of reference forms, with the property that if sμP 1σ = 0,  then signηP 1σ > 0. More generally, for every hermitian form η0 over (A, σ), the tuple (η0 , η1 , . . . , ηt ) will also be a tuple of reference forms. Remark 2.5. Let (A, σ) and (B, τ ) be Morita equivalent F -algebras with involution of the same type. Denoting this equivalence by μ and letting η = (η1 , . . . , ηt ) be a tuple of reference forms for (A, σ), it follows from [5, Theorem 4.2] that (μ(η1 ), . . . , μ(ηt )) is a tuple of reference forms for (B, τ ). F = ∅. Lemma 2.6. If (D, ϑ, ε) = (F, idF , −1), then X Proof. It follows from (2.1) that W (A, σ) ∼ = W−1 (D, ϑ) = W−1 (F, idF ), which is zero.  Therefore, sμP = 0 for every P ∈ XF , i.e. XF = ∅. 2 F =  ∅. Remark 2.7. Following Lemmas 2.3 and 2.6 we take ε = 1 whenever X Use of the notation signηP h assumes that η is some tuple of reference forms for (A, σ) and that h is a nonsingular hermitian form over (A, σ). Also, if F has only one ordering P , we write signη instead of signηP .

V. Astier, T. Unger / Journal of Algebra 508 (2018) 339–363

345

2.4. The nonsingular part of a hermitian form Let u be an element in Sym(A, σ), not necessarily invertible. In the next sections we examine the “positivity” of u and its relation to sums of hermitian squares in terms of the associated hermitian form uσ over (A, σ), which may be singular. The properties that we are interested in only depend on the nonsingular part of uσ , which motivates the remainder of this section. Lemmas 2.8 and 2.9 below correspond to [13, Chapter I, Lemma 6.2.3] and [13, Chapter I, Proposition 6.2.4], but are stated for possibly singular ε-hermitian forms. Lemma 2.8. Let (D, ϑ) be an F -division algebra with involution and let (M, h) be an ε-hermitian form over (D, ϑ), where ε ∈ {−1, 1}. Assume that h(x, x) = 0 for all x ∈ M . Then h=0

or

(D, ϑ, ε) = (F, idF , −1).

Proof. Assume h = 0 and let x, z ∈ M be such that h(x, z) = α = 0. Let d ∈ D× and let y = zα−1 d. Then h(x, y) = d, and the proof proceeds as in the proof of [13, Chapter I, Lemma 6.2.3]: if ϑ is nontrivial, we reach a contradiction; if ϑ is trivial we must have D = F and therefore ε = −1. 2 Recall from Proposition A.3 in Appendix A that if h is a hermitian form over (A, σ), then h  h0 ⊥ hns where h0 is the zero form of appropriate rank and hns is nonsingular, and that h0 and hns are unique up to isometry. Lemma 2.9. Let (D, ϑ) be an F -division algebra with involution and let (M, h) be an ε-hermitian form over (D, ϑ), where ε ∈ {−1, 1}. Assume that the Gram matrix of h is H. Then there exists an invertible matrix G ∈ M (D) such that ϑ(G)t HG = diag(u1 , . . . , uk , 0, . . . , 0), where u1 , . . . , uk ∈ Sym(D, ϑ)× , except when (D, ϑ, ε) = (F, idF , −1), in which case u1 , . . . , uk ∈ Sym−1 (M2 (F ), t )× . Proof. Assume first that (D, ϑ, ε) = (F, idF , −1). If h = 0, there is nothing to prove. Otherwise, there exists x ∈ M such that h(x, x) = 0, by Lemma 2.8. Then M = xD ⊕ (xD)⊥h and the result follows by induction. Finally, if (D, ϑ, ε) = (F, idF , −1), the result is well-known. 2 Let (A, σ) be an F -algebra with involution and fix an isomorphism f : (A, σ) → (M (D), Int(Φ−1 ) ◦ ϑt ) as at the start of Section 2.2. Let u ∈ Sym(A, σ). Since Φf (u) ∈ Symε (M (D), ϑt ), it is the Gram matrix of an ε-hermitian form over (D, ϑ) and thus, by Lemma 2.9, there exists an invertible matrix G ∈ M (D) such that

V. Astier, T. Unger / Journal of Algebra 508 (2018) 339–363

346

ϑ(G)t (Φf (u))G = diag(u1 , . . . , uk , 0, . . . , 0),

(2.3)

where u1 , . . . , uk are as in Lemma 2.9. For i = 1, . . . , k, let ϕi denote the ε-hermitian form over (D, ϑ) with Gram matrix ui . The F -algebras with involution (A, σ) and (D, ϑ) are Morita equivalent, cf. [13, Chapter I, Theorem 9.3.5]. Consider the hermitian form uσ over (A, σ). Under the equivalences depicted in (2.1), uσ corresponds to the scaled ε-hermitian form Φf (u)ϑt over (M (D), ϑt ), which then corresponds to the collapsed -dimensional ε-hermitian form ϕ with Gram matrix diag(u1 , . . . , uk , 0, . . . , 0). Note that ϕ = ϕ1 ⊥ . . . ⊥ ϕk ⊥ 0 ⊥ . . . ⊥ 0. For i ∈ {1, . . . , k}, the preimage of ϕi under these equivalences is a nonsingular hermitian form over (A, σ) which we denote by hi . Consequently we obtain the orthogonal decomposition uσ  h1 ⊥ . . . ⊥ hk ⊥ 0 ⊥ . . . ⊥ 0, where 0 denotes the zero form of rank 1 over (A, σ). The form h1 ⊥ . . . ⊥ hk is the nonsingular part uns σ of uσ . The following result characterizes the representation of not necessarily invertible elements in Sym(A, σ) in terms of hermitian forms. Proposition 2.10. Let h be a hermitian form over (A, σ) and let u ∈ Sym(A, σ). The following statements are equivalent: (i) u ∈ D(A,σ) (r × h) for some r ∈ N.   (ii) The form uns σ is a subform of r × h for some r ∈ N. Proof. We use the notation from the beginning of this section and denote being a subform by . Assume first that (D, ϑ, ε) = (F, idF , −1). With reference to the equivalences in (2.1), we have the following equivalent statements (with justifications below): ∃r ∈ N

u ∈ D(A,σ) (r × h)

⇔ ∃r ∈ N

Φf (u) ∈ D(M (D),ϑt ) (r × Φf∗ (h))

⇔ ∃r ∈ N

ϑ(G)t (Φf (u))G = diag(u1 , . . . , uk , 0, . . . , 0) ∈ D(M (D),ϑt ) (r × Φf∗ (h))

(2.4)

(2.5)

⇔ ∃s ∈ N ∀i = 1, . . . , k

diag(ui , . . . , ui ) ∈ D(M (D),ϑt ) (s × Φf∗ (h))

(2.6)

⇔ ∃s ∈ N ∀i = 1, . . . , k

diag(ui , . . . , ui )ϑt  s × Φf∗ (h)

(2.7)

⇔ ∃s ∈ N  × u1 ϑ , . . . ,  × uk ϑ  s × g(Φf∗ (h))

(2.8)

V. Astier, T. Unger / Journal of Algebra 508 (2018) 339–363

347

⇔ ∃s1 ∈ N u1 ϑ , . . . , uk ϑ  s1 × g(Φf∗ (h)) ⇔ ∃s2 ∈ N u1 ϑ ⊥ . . . ⊥ uk ϑ  s2 × g(Φf∗ (h)) ⇔ ∃r ∈ N

 uns σ = h1 ⊥ . . . ⊥ hk  r × h.

(2.9)

The justifications are as follows: (2.4) follows by scaling, (2.7) is a standard argument (recalled in Lemma A.1 in Appendix A), (2.8) follows by collapsing and (2.9) follows by the full sequence of equivalences in (2.1) (between (D, ϑ) and (A, σ)) and the observations preceding the proposition. Both directions of (2.6) follow by applying sufficiently many transformations of the form X → ϑ(Q)t XQ to diag(u1 , . . . , uk , 0, . . . , 0) or u1 I , . . . , uk I , where Q is diag(0, . . . , 0, 1, 0, . . . , 0)

(where 1 can be in any position)

or a permutation matrix, and summing the results. Finally, if (D, ϑ, ε) = (F, idF , −1), the same argument works mutatis mutandis, using ui ∈ Sym−1 (M2 (D), ϑt )× , noting that the step from (2.5) to (2.6) works since  is even (indeed, Φ is an invertible skew-symmetric matrix over F in the case under consideration, and is thus of even dimension). 2 3. Maximal elements and sums of hermitian squares In contrast to quadratic forms, the signature of nonsingular hermitian forms of dimension one can take more than just two values. It is therefore natural to single out those elements u in Sym(A, σ) whose associated hermitian form uσ has maximal possible signature, leading to a natural notion of positivity, which we call η-maximality (where η is a tuple of reference forms for (A, σ)), cf. Definition 3.1. Our main result, Theorem 3.6, shows that, as in the quadratic forms case, Pfister’s local-global principle can be used to characterize “totally positive” elements in terms of (weighted) sums of hermitian squares, providing an extension of Artin’s result to algebras with involution. We treat the case of invertible elements first in Theorem 3.3 since its proof is more streamlined and the arguments appear more clearly. Definition 3.1. Let P ∈ XF and let η be a tuple of reference forms for (A, σ). (i) Let mP := max{signηP aσ | a ∈ Sym(A, σ)× }. We call u ∈ Sym(A, σ)× η-maximal at P if signηP uσ = mP . (ii) We call a nonsingular hermitian form h of rank k over (A, σ) η-maximal at P if for every nonsingular form h of rank k over (A, σ) we have signηP h  signηP h .

348

V. Astier, T. Unger / Journal of Algebra 508 (2018) 339–363

(iii) We call a hermitian form h over (A, σ) (resp. an element u ∈ Sym(A, σ)) η-maximal at P if hns (resp. uns σ ) is η-maximal at P . Observe that mP does not depend on the choice of η. Proposition 3.2. Let P ∈ XF and let MP := max{signηP h | h is a rank 1 nonsingular hermitian form over (A, σ)}. Then (i) max{signηP h | h is a rank t nonsingular hermitian form over (A, σ)} = tMP ; (ii) mP = MP . F . Proof. If P ∈ Nil[A, σ], then mP = MP = 0, so we may assume that P ∈ X (i) Let h be a nonsingular form of rank t. Since h is an orthogonal sum of forms of rank 1, signηP h  tMP . The equality follows by taking a form h0 of rank 1 such that signηP h0 = MP and considering t × h0 . (ii) The inequality mP  MP follows from the fact that a form of dimension 1 has rank  and thus is an orthogonal sum of  hermitian forms of rank 1. For the other inequality, we now construct a form of dimension 1 and signature MP . Using the notation introduced in Section 2.2, the tuple η of reference forms for (A, σ) obviously behaves as follows under the equivalences in (2.1): η

(s ◦ f∗ )(η)

f∗ (η)

(g ◦ s ◦ f∗ )(η),

F , cf. Remark 2.7. Since signature and rank are preserved under where ε = 1 since P ∈ X Morita equivalence (cf. [5, Theorem 4.2] and [7, §2.2]), there exists a form dϑ of rank (g◦s◦f∗ )(η) 1 over (D, ϑ) such that signP dϑ = MP . Let w = diag(d, . . . , d) ∈ M (D) and consider the form wϑt . Then (2.1) yields forms f −1 (Φ−1 w)σ and Φ−1 wadΦ such that f −1 (Φ−1 w)σ

Φ−1 wadΦ

wϑt

 × dϑ

(note that s(uadΦ ) := ΦuadΦ = Φuϑt for u ∈ M (D), which is easy to check). Then, by [5, Theorem 4.2], signηP f −1 (Φ−1 w)σ =  signP

(g◦s◦f∗ )(η)

dϑ = MP .

2

3.1. The case of invertible elements Let b1 , . . . , bt ∈ F × . We use the notation b1 , . . . , bt  := 1, b1  ⊗· · ·⊗1, bt  for Pfister forms and also write

V. Astier, T. Unger / Journal of Algebra 508 (2018) 339–363

349

H(b1 , . . . , bt ) := {P ∈ XF | b1 , . . . , bt ∈ P } for the corresponding Harrison set. Note that such Harrison sets form a basis of the Harrison topology on XF . Theorem 3.3. Let b1 , . . . , bt ∈ F × , π = b1 , . . . , bt , Y = H(b1 , . . . , bt ) and η be a tuple of reference forms for (A, σ). Assume that a ∈ Sym(A, σ)× is η-maximal at all P ∈ Y . Let u ∈ Sym(A, σ)× . The following statements are equivalent: (i) u is η-maximal at all P ∈ Y . (ii) u ∈ D(A,σ) (s × π ⊗ aσ ) for some s ∈ N. Proof. Assume (i). It follows from the assumptions that signηP a, −uσ = 0 for all P ∈ Y . Hence signηP (π⊗a, −uσ ) = signP π·signηP a, −uσ = 0 for all P ∈ XF . Thus π⊗a, −uσ is torsion in W (A, σ) by [17, Theorem 4.1]. In other words, there exists s ∈ N such that  2s × π ⊗ a, −uσ = 0 in W (A, σ) by [21, Theorem 5.1], from which (ii) follows. Assume (ii), i.e. assume that u ∈ D(A,σ) (h), where h = s × π ⊗ aσ . Then u = h(x, x) for some x ∈ M = Ar , where r = s2t . Since u is invertible, a standard argument shows that M = xA ⊕ (xA)⊥h . Thus h  uσ ⊥ h , for some hermitian form h over (A, σ) of rank (s2t − 1) (since A ∼ = M (D), for some F -division algebra D). By assumption we have for every P ∈ Y that signηP h = s2t mP = signηP uσ + signηP h .

(3.1)

Since signηP uσ  mP and signηP h  mP rk(h ) = mP (s2t − 1) (by Proposition 3.2), these inequalities are in fact equalities by (3.1), and (i) follows. 2 Remark 3.4. If P ∈ Nil[A, σ], then the statement “u is η-maximal at P ” is trivially true. F . Thus Theorem 3.3(i) only needs to be checked for P ∈ Y ∩ X 3.2. The general case The following result is the equivalent of Theorem 3.3 when u is not necessarily invertible. Proposition 3.5. Let b1 , . . . , bt ∈ F × , π = b1 , . . . , bt , Y = H(b1 , . . . , bt ) and η be a tuple of reference forms for (A, σ). Assume that a ∈ Sym(A, σ)× is η-maximal at all P ∈ Y . Let h be a hermitian form over (A, σ). The following statements are equivalent:

350

V. Astier, T. Unger / Journal of Algebra 508 (2018) 339–363

(i) hns is η-maximal at all P ∈ Y . (ii) hns is a subform of k × π ⊗ aσ for some k ∈ N. Proof. (i) ⇒ (ii): We write h  hns ⊥ 0 and let r := rk(hns ). Let P ∈ Y . By Proposition 3.2 it follows that signηP hns = rmP /. Note that signηP aσ = mP and that rk(aσ ) = . It follows that signηP (r × aσ −  × hns ) = 0 for every P ∈ Y . Therefore, by Pfister’s local-global principle ([17, Theorem 4.1], [21, Theorem 5.1]), there exists k ∈ N   such that 2k  × π ⊗ hns  2k r × π ⊗ aσ and the result follows. (ii) ⇒ (i): Let P ∈ Y . By the assumption on a and Proposition 3.2, k × π ⊗ aσ is η-maximal. The conclusion follows by the additivity of signηP . 2 It follows from Proposition 2.10 and Proposition 3.5 that Theorem 3.6. Let b1 , . . . , bt ∈ F × , π = b1 , . . . , bt , Y = H(b1 , . . . , bt ) and η be a tuple of reference forms for (A, σ). Assume that a ∈ Sym(A, σ)× is η-maximal at all P ∈ Y . Let u ∈ Sym(A, σ). The following statements are equivalent: (i) u is η-maximal at all P ∈ Y . (ii) u ∈ D(A,σ) (k × π ⊗ aσ ) for some k ∈ N. Concerning the existence of the element a in Theorem 3.6: in a forthcoming paper [2, F , there exists a ∈ Sym(A, σ)× that Propositions 6.6, 6.7] we show that for every P ∈ X is η-maximal at P (and we also determine the value of mP ). By continuity of the map signη aσ : XF → Z, cf. [4, Theorem 7.2], there is an open set Y around P such that a is η-maximal at all P ∈ Y . To conclude this section we consider (A, σ) = (Mn (F ), t), where t denotes transposition, and obtain a result similar to a classical theorem of Gondard and Ribenboim [11, Théorème 1]: Corollary 3.7. A symmetric matrix over F is positive semidefinite at all P ∈ XF if and only if it is a sum of hermitian squares in (Mn (F ), t). Proof. We may take η = (1t ) as a tuple of reference forms for (A, σ) since signηP 1t = n F = XF . Let U ∈ Sym(Mn (F ), t). Then U is positive for every P ∈ XF . Note that X semidefinite at all P ∈ XF if and only if all nonzero eigenvalues of U are positive at all P ∈ XF if and only if U ns t is η-maximal at all P ∈ XF . Finally, by Theorem 3.6 with a = 1 and Y = H(1) = XF , this happens if and only if U is a sum of hermitian squares in (Mn (F ), t). 2 4. A theorem and a question of Procesi and Schacher Procesi and Schacher already considered a notion of positivity of elements in an algebra with involution and proved a result characterizing totally positive elements (in their

V. Astier, T. Unger / Journal of Algebra 508 (2018) 339–363

351

sense) in terms of weighted sums of squares of symmetric elements, cf. [19, Theorem 5.4]. They also raised the question of whether positive elements are always sums of hermitian squares (and not necessarily squares of symmetric elements), cf. [19, p. 404]. In this spirit, after showing how their notion of positivity relates to ours, we prove a sums of hermitian squares version of [19, Theorem 5.4], using Theorem 3.6, and use our techniques to fully answer the question raised in [19, p. 404] of whether positive elements are always sums of hermitian squares. Let (A, σ) be an F -algebra with involution, let u ∈ Sym(A, σ). In [19], Procesi and Schacher define the positivity of u in terms of the corresponding scaled involution trace form T(A,σ,u) . Consider T(A,σ) : A × A → K, (x, y) → TrdA (σ(x)y)

for x, y ∈ A

and T(A,σ,u) : A × A → K, (x, y) → TrdA (σ(x)uy)

for x, y ∈ A,

where u ∈ Sym(A, σ). These forms are both symmetric bilinear over F if σ is of the first kind and hermitian over (K, ι) if σ is of the second kind. The first form is always nonsingular, whereas the second form is nonsingular if and only if u is invertible, cf. [14, §11]. Recall the following definitions from [19, Definitions 1.1 and 5.1]: Definition 4.1. Let P ∈ XF . (i) The involution σ is called positive at P if the form T(A,σ) is positive semidefinite at P . We also introduce the notation Xσ := {P ∈ XF | σ is positive at P }. (ii) Assume that σ is positive at P . An element u ∈ Sym(A, σ) is called positive at P if the form T(A,σ,u) is positive semidefinite at P . Remark 4.2. Recall that a nonsingular symmetric bilinear form over F or a hermitian form over (K, ι) is positive semidefinite at a given ordering P on F if and only if it is positive definite at P . Another way of looking at the Procesi–Schacher notion of positivity is from the point of view of signatures of involutions and signatures of hermitian forms, and specifically the signature of the form uσ . Propositions 4.8 and 4.10 give the precise connections between these approaches, whereas Remark 4.9 describes positivity of u at P in terms of a different trace form, T(A,σu ) , under a weaker hypothesis.

352

V. Astier, T. Unger / Journal of Algebra 508 (2018) 339–363

Recall from [16] and [20] (or [14, §11]) that the signature of σ at P ∈ XF is defined as  signP σ :=

signP T(A,σ) .

(4.1)

Remark 4.3. If follows from (4.1) that σ is positive at P ∈ XF if and only if signP σ = deg A(= n). √ F then A ⊗F FP ∼ DP , where DP is one of FP , FP ( −1) or Recall that if P ∈ X √ (−1, −1)FP . We define λP := deg DP , i.e. λP = 1 if DP = FP or FP ( −1) and λP = 2 if DP = (−1, −1)FP . We also let nP = n/λP , so that A ⊗F FP ∼ = MnP (DP ). Now let h be a hermitian form over (A, σ) with adjoint involution adh . Then for P ∈ XF , signP adh = λP | signηP h|

(4.2)

(if P ∈ Nil[A, σ], both sides of (4.2) are zero), cf. [4, Lemma 4.6]. Note that the correspondence between adh and h is unique only up to multiplication of h by a nonzero element in F and that λP only depends on the Brauer class of A. In the following proposition we collect a few elementary statements about signatures of involutions and one-dimensional forms. For u ∈ Sym(A, σ)× we write σu := Int(u−1 ) ◦ σ. Proposition 4.4. Let u ∈ Sym(A, σ)× and let P ∈ XF . (i) (ii) (iii) (iv)

signP σu = λP | signηP uσ |. signP σu ∈ {0, . . . , n}. signηP uσ ∈ {−nP , . . . , nP }. signP σu = n ⇔ | signηP uσ | = nP .

Proof. (i) follows from (4.2) since the involution σu is adjoint to the form uσ , as can easily be verified. (ii): Since dimK A = m = n2 we have dim T(A,σu ) = m. Using that signP T(A,σu ) is always a square (cf. [16], [20]) we obtain signP T(A,σu ) ∈ {0, 1, 4, . . . , (n − 1)2 , n2 } and thus signP σu ∈ {0, . . . , n} by (4.1). (iii) follows from (i) and (ii), whereas (iv) follows from (i). 2 Remark 4.5. It is clear that P ∈ Xσ if and only if the form T(A,σ) is positive definite at P , cf. (4.1). Furthermore, mP  nP and if P ∈ Xσ , then mP = nP by Proposition 4.4(iv). As an immediate consequence of Proposition 4.4 we obtain:

V. Astier, T. Unger / Journal of Algebra 508 (2018) 339–363

353

Corollary 4.6. The following statements are equivalent: (i) P ∈ Xσ . (ii) | signηP 1σ | = nP for all tuples of reference forms η. (iii) signηP 1σ = nP for all tuples of reference forms η of the form (1σ , . . .). F and so we may take εP = 1 by Remark 4.7. Let P ∈ Xσ . By Corollary 4.6, P ∈ X ∼ definition of signature. Hence (A ⊗F FP , σ ⊗ id) = (MnP (DP ), adΦP ), for some matrix ΦP ∈ Sym(MnP (DP ), t ). It follows from [4, Lemma 3.10] and Corollary 4.6 that sign ΦP = ±nP , where sign denotes the Sylvester signature of hermitian matrices. In other words, ΦP is positive definite or negative definite and, up to replacing ΦP by −ΦP (since adΦP = ad−ΦP ) we may assume that ΦP is positive definite. In the following result we make the link between Procesi and Schacher’s notion of positivity (statement (ii); see also Definition 4.1) and signatures of hermitian forms. Proposition 4.8. Let η be a tuple of reference forms for (A, σ), P ∈ XF and u ∈ Sym(A, σ)× . Assume that σ is positive at P . The following statements are equivalent: (i) The involution σu is positive at P . (ii) The form T(A,σ,u) is positive definite or negative definite at P . (iii) u or −u is η-maximal at P . Proof. By [14, (11.1)] the involution σu ⊗ ι σ corresponds to adT(A,σ,u) under the isomorphism A ⊗K ι A −→ EndK (A), where (ι A, ι σ) is the conjugate algebra with involution of (A, σ). It follows from the definition of ι σ that signP σ = signP ι σ and from [4, Remark 4.2] that signP adT(A,σ,u) = signP σu · signP σ. From [16] and [20] we obtain that | signP T(A,σ,u) | = signP adT(A,σ,u) . These two equalities prove the equivalence (i) ⇔ (ii). The equivalence (i) ⇔ (iii) follows from Proposition 4.4(iv) and the fact that nP = mP , since σ is positive at P . 2 Remark 4.9. If we drop the assumption that σ is positive at P in Proposition 4.8, we obtain (from (4.1) and Proposition 4.4(iv)) a similar sequence of equivalences, but in terms of a different form, namely T(A,σu ) : let η be a tuple of reference forms for (A, σ), P ∈ XF and u ∈ Sym(A, σ)× . The following statements are equivalent:

354

V. Astier, T. Unger / Journal of Algebra 508 (2018) 339–363

(i) The involution σu is positive at P . (ii) The form T(A,σu ) is positive definite at P . (iii) | signηP uσ | = nP . The equivalence between (ii) and (iii) in Proposition 4.8 can be made more precise: Proposition 4.10. Let η be a tuple of reference forms for (A, σ), P ∈ XF and u ∈ Sym(A, σ)× . Assume that σ is positive at P . (i) If 1 is η-maximal at P , then T(A,σ,u) is positive definite at P if and only if u is η-maximal at P . (ii) If −1 is η-maximal at P , then T(A,σ,u) is negative definite at P if and only if u is η-maximal at P . Proof. (ii) follows from (i) upon replacing η by −η and u by −u. Thus, it suffices to prove (i). Observe that by Corollary 4.6 and Remark 4.5, σ positive at P implies that either 1 F . or −1 is η-maximal at P . Also note that the assumption on σ implies that P ∈ X Assume that 1 is η-maximal at P . By Proposition 4.8 and since T(A,σ,−u) = −T(A,σ,u) , we only need to show the sufficient condition in (i). Thus, assume that u (and hence u−1 ) is η-maximal at P . Then signηP 1, −uσ = 0. Consider the involution τ = ad1,−uσ on M2 (A). It follows from (4.1) and (4.2) that signP T(M2 (A),τ ) = 0. Using [10, §22, Theorem 1] it is not difficult to see that T(M2 (A),τ )  T(A,σ) ⊥ T(A,σu−1 ) ⊥ T(A,σ,−u−1 ) ⊥ ι(T(A,σ,−u) ).

(4.3)

Observe that signP T(A,σ) = signP T(A,σu−1 ) = n2 by the assumptions on σ and u, cf. Remark 4.9. Also, T(A,σ,−u−1 )  T(A,σ,−u) = −T(A,σ,u) and signP ι(T(A,σ,−u) ) = signP T(A,σ,−u) = − signP T(A,σ,u) . After taking signatures at P in (4.3) we obtain signP T(A,σ,u) = n2 , i.e. T(A,σ,u) is positive definite at P . 2 We record the next result for future use: Proposition 4.11. Let (A, σ) be an F -algebra with involution such that Xσ = ∅. Then there exists an F -linear involution τ on D, of the same type as σ, such that Xσ ⊆ Xτ . Proof. Write (A, σ) ∼ = (M (D), adΦ ) with ϑ, ε and Φ as in Section 2.2. Since Xσ = ∅,  we have XF = ∅. Recall from Remark 2.7 that we take ε = 1 in this case, and thus ϑ of the same type as σ. Consider the hermitian form 1σ . It corresponds to an -dimensional hermitian form a1 , . . . , a ϑ via the isomorphisms in (2.1). We show that Xσ ⊆ Xτ , where τ is the involution ϑa1 on D.

V. Astier, T. Unger / Journal of Algebra 508 (2018) 339–363

355

Let P ∈ Xσ . Let η be a tuple of reference forms for (A, σ) of the form (1σ , . . .), cf. Remark 2.4. The assumption signP σ = n = deg A is equivalent with signηP 1σ = nP by Corollary 4.6. Since the form 1σ corresponds to a1 , . . . , a ϑ , we have (g◦s◦f∗ )(η) signP a1 , . . . , a ϑ = nP by [5, Theorem 4.2]. Since deg D = n/, the signature of a one-dimensional hermitian form over (D, ϑ) is bounded by nP / (since such a form gives rise to a matrix in MnP / (DP ) during the signature computation). It follows (g◦s◦f∗ )(η) that signP ai ϑ = nP / for all i ∈ {1, . . . , }. By Corollary 4.6, the involution ϑai on D is positive at P for all i ∈ {1, . . . , }. In particular, P ∈ Xτ . Observe that since a1 ∈ Sym(D, ϑ)× , the involution τ is of the same type as σ. 2 4.1. A theorem of Procesi and Schacher By Proposition 4.11 we may choose the involution ϑ on D such that Xσ ⊆ Xϑ . If √ F , and thus (DP , ) is one of (FP , id), (FP ( −1), ), P ∈ Xσ , we have in particular P ∈ X or ((−1, −1)FP , ), cf. Section 2.3. Recall that we have an isomorphism f : (A, σ) → (M (D), Int(Φ−1 ) ◦ ϑt ). It induces an isomorphism of FP -algebras with involution f ⊗ id : (A ⊗F FP , σ ⊗ id) → (M (D) ⊗F FP , (Int(Φ−1 ) ◦ ϑt ) ⊗ id). Lemma 4.12. Let P ∈ Xϑ . The FP -algebras with involution (M (D) ⊗F FP , ϑt ⊗ id) and (MnP (DP ), t ) are isomorphic. Proof. Since M (D) ⊗F FP ∼ = MnP (DP ), there exists a hermitian form h over (DP , ) such that (M (D) ⊗F FP , ϑt ⊗ id) ∼ = (MnP (DP ), adh ). Since ϑ is positive at P , the involution adh is positive at the unique ordering of FP . It follows from Remark 4.3 and (4.2) that the signature of h equals ±nP . Therefore, h  ±nP × 1− and so (M (D) ⊗F FP , ϑt ⊗ id) ∼ = (MnP (DP ), t ). 2 For P ∈ Xσ , and thus P ∈ Xϑ , we fix one of the isomorphisms whose existence is asserted by Lemma 4.12 and denote it by αP . We define ΨP := αP (Φ−1 ⊗1) and note that t ΨP is invertible and satisfies ΨP = ΨP . A straightforward computation shows that αP is also an isomorphism between the algebras with involution (M (D) ⊗F FP , (Int(Φ−1 ) ◦ ϑt ) ⊗ id) and (MnP (DP ), Int(ΨP ) ◦ t ). In addition we define fP = αP ◦ (f ⊗ id). Lemma 4.13. Let P ∈ Xσ and u ∈ Sym(A, σ). Then T(A,σ,u) is positive semidefinite at P if and only if there exists ρP ∈ {−1, 1} such that T(Mn (DP ), t ,ρP Ψ−1 fP (u⊗1)) is positive P P semidefinite at the unique ordering on FP . t

Proof. Note that ΨP = ΨP . Since σ is positive at P , we have by Remark 4.7 that ρP ΨP is a positive definite matrix over DP for some ρP ∈ {−1, 1}. Thus ρP ΨP has a t square root in MnP (DP ) and we write ρP ΨP = Ω2P with ΩP = ΩP . The form T(A,σ,u) is positive semidefinite at P if and only if it remains so over FP . We have, for x ∈ A ⊗F FP ,

356

V. Astier, T. Unger / Journal of Algebra 508 (2018) 339–363

(T(A,σ,u) ⊗ FP )(x, x) = T(A⊗FP ,σ⊗id,u⊗1) (x, x)   = TrdA⊗FP (σ ⊗ id)(x)(u ⊗ 1)x t

= TrdMnP (DP ) (ρP ΨP fP (x) ρP Ψ−1 P fP (u ⊗ 1)fP (x)) t

= TrdMnP (DP ) (Ω2P fP (x) ρP Ψ−1 P fP (u ⊗ 1)fP (x)) t

= TrdMnP (DP ) (ΩP fP (x) ρP Ψ−1 P fP (u ⊗ 1)fP (x)ΩP ) = TrdMnP (DP ) (y t ρP Ψ−1 P fP (u ⊗ 1)y) = T(Mn

P

(DP ),

t ,ρ Ψ−1 f (u⊗1)) P P P

(y, y),

where y = fP (x)ΩP . The statement follows. 2 Lemma 4.14. Let P ∈ Xσ and u ∈ Sym(A, σ). Then T(A,σ,u) is positive semidefinite at P if and only if there exists ρP ∈ {−1, 1} such that T(M (D),ϑt ,ρP Φf (u)) is positive semidefinite at P .  Proof. Let P ∈ Xσ . From the definition of αP and ΨP we have αP ◦ (Int(Φ−1 ) ◦ ϑt ) ⊗  id = (Int(ΨP ) ◦ t ) ◦ αP , which implies −1 ϑt ⊗ id = Int(Φ ⊗ 1) ◦ αP ◦ Int(ΨP ) ◦

t

◦ αP ,

(4.4)

and αP ◦ Int(Φ ⊗ 1) = Int(Ψ−1 P ) ◦ αP (by definition of ΨP ), which implies −1 αP ◦ Int(Φ ⊗ 1) ◦ αP ◦ Int(ΨP ) = id.

(4.5)

Furthermore, using (4.4) followed by (4.5) we obtain αP ◦ (ϑt ⊗ id) =

t

◦ αP .

(4.6)

Clearly, T(M (D),ϑt ,ρP Φf (u)) is positive semidefinite at the ordering P if and only if T(M (D),ϑt ,ρP Φf (u)) ⊗ FP is positive semidefinite at the unique ordering on FP . We have, for x ∈ M (D) ⊗ FP ,   T(M (D),ϑt ,ρP Φf (u)) ⊗ FP (x, x) = T(M (D)⊗FP ,ϑt ⊗id,ρP Φf (u)⊗1) (x, x)

= TrdM (D)⊗FP (ϑt ⊗ id)(x)(ρP (Φf (u)) ⊗ 1)x

 = TrdMnP (DP ) αP (ϑt ⊗ id)(x)(ρP (Φf (u)) ⊗ 1)x

= TrdMnP (DP )



[since αP is an isomorphism of FP -algebras]   αP ◦ (ϑt ⊗ id) (x)ρP αP (Φ ⊗ 1)αP (f (u) ⊗ 1)αP (x)

V. Astier, T. Unger / Journal of Algebra 508 (2018) 339–363

357

 t = TrdMnP (DP ) αP (x) ρP Ψ−1 f (u ⊗ 1)α (x) [by (4.6)] P P P = T(Mn

P

(DP ),

t ,ρ Ψ−1 f (u⊗1)) P P P

(y, y),

where y = αP (x). The statement follows by Lemma 4.13. 2 Lemma 4.15. With notation as in (2.3) we have T(M (D),ϑt ,Φf (u))   × (T(D,ϑ,u1 ) ⊥ · · · ⊥ T(D,ϑ,uk ) ⊥ 0 · · · ⊥ 0) when (D, ϑ, ε) = (F, idF , −1). Proof. It follows from (2.3) that T(M (D),ϑt ,Φf (u))  T(M (D),ϑt ,diag(u1 ,...,uk ,0,...,0)) . The statement follows from a direct matrix computation starting from the canonical decom· · ⊕ D . 2 position of M (D) into simple M (D)-modules: M (D) ∼ = D ⊕ ·   copies

Lemma 4.16. Assume that T(A,σ)  b1 , . . . , bm ι with all bi ∈ F × . Then Xσ = H(b1 , . . . , bm ). Proof. It follows from Definition 4.1(i) and (4.1) that P ∈ Xσ if and only if bi ∈ P for all i = 1, . . . , m. 2 We have now laid the ground work for proving our sums of hermitian squares version of [19, Theorem 5.4]: Theorem 4.17. Let u ∈ Sym(A, σ) and let T(A,σ)  b1 , . . . , bm ι with all bi ∈ F × . The following statements are equivalent: (i) uns σ is η-maximal at all P ∈ Xσ , where η is any tuple of reference forms for (A, σ) of the form (1σ , . . .). (ii) The form T(A,σ,u) is positive semidefinite at all P ∈ Xσ . (iii) u ∈ D(A,σ) (r × b1 , . . . , bm  ⊗ 1σ ) for some r ∈ N. Proof. The equivalence between (i) and (iii) follows from Theorem 3.6. (iii) ⇒ (ii): Assume that u=

 e∈{0,1}m

be



σ(xi,e )xi,e ,

i

where be = be11 · · · bemm and xi,e ∈ A. Let x ∈ A \ {0}. Then

358

V. Astier, T. Unger / Journal of Algebra 508 (2018) 339–363

TrdA (σ(x)ux) =

 e∈{0,1}m

be



TrdA (σ(xi,e x)xi,e x)

i

is nonnegative at all P ∈ Xσ by definition of Xσ , (4.1), and Lemma 4.16. (ii) ⇒ (i): The implication is trivially true if Xσ = ∅. Thus we assume Xσ = ∅. By Proposition 4.11 the involution ϑ on D can be chosen so that Xσ ⊆ Xϑ and of the same type as σ (which is in line with our convention that ε = 1 in this case, cf. Remark 2.7). Let ξ be the tuple of reference forms for (D, ϑ), obtained from η via the Morita equivalences in (2.1). Let P ∈ Xσ . We have the following equivalences (with PD meaning positive definite and PSD meaning positive semidefinite, as usual): T(A,σ,u) is PSD at P ⇔ T(M (D),ϑt ,ρP Φf (u)) is PSD at P [by Lemma 4.14] ⇔ T(D,ϑ,ρP ui ) is PSD at P for i = 1, . . . , k [by Lemma 4.15 since ε = 1] ⇔ T(D,ϑ,ρP ui ) is PD at P for i = 1, . . . , k [since all ui are invertible] ⇔ ∃δ ∈ {−1, 1} such that δρP ui is ξ-maximal at P for i = 1, . . . , k [by Proposition 4.10 since P ∈ Xϑ ] ⇔ ∃δ ∈ {−1, 1} such that

δρP uns σ

is η-maximal at P .

Assume for the sake of contradiction that δρP = −1. Thus P ∈ {Q ∈ Xσ | −uns σ is η-maximal at Q}, which is open in XF since the map signη uns σ : XF → Z is continuous [4, Theorem 7.2]. Therefore, there exist c1 , . . . , ct ∈ F × such that P ∈ H(c1 , . . . , ct ) ⊆ {Q ∈ Xσ | −uns σ is η-maximal at Q}. Applying Theorem 3.6 with Y = H(c1 , . . . , ct ) and a = 1 then gives −u ∈ D(A,σ) (s × c1 , . . . , ct  ⊗ 1σ ) for some s ∈ N. A trace computation as in the proof of (iii) ⇒ (ii) above then shows that the form T(A,σ,u) is negative semidefinite at P , contradiction. 2 4.2. A question of Procesi and Schacher Consider the following property: (PS) for every u ∈ Sym(A, σ), the form T(A,σ,u) is positive semidefinite at all P ∈ Xσ if and only if u ∈ D(A,σ) (s × 1σ ) for some s ∈ N. In [19, p. 404], Procesi and Schacher, motivated by [19, Theorem 5.4], ask if property (PS) holds for all F -algebras with involution (A, σ) and give a positive answer for quaternion algebras [19, Corollary 5.5] and in the case where Xσ = XF [19, Proposition 5.3]. In [12] an elementary counterexample is produced to (PS) in general and some cases

V. Astier, T. Unger / Journal of Algebra 508 (2018) 339–363

359

are studied where (PS) holds. Our previous results yield a slight improvement on [19, Proposition 5.3]: F , then property (PS) holds. Corollary 4.18. If Xσ = X Proof. Let u ∈ Sym(A, σ) and let η be a tuple of reference forms for (A, σ) of the form F if and only if uns is (1σ , . . .). Then T(A,σ,u) is positive semidefinite on Xσ = X σ F (and, trivially, on XF ) by Theorem 4.17, which in turn is η-maximal at all P ∈ X equivalent to u ∈ D(A,σ) (s × 1σ ) for some s ∈ N by Theorem 3.6 with a = 1 and Y = H(1) and because 1 is η-maximal on XF . 2 Consider the following variation on property (PS), where we enlarge the set of orderF : ings on which positivity is verified from Xσ to X F (PS’) for every u ∈ Sym(A, σ), the form T(A,σ,u) is positive semidefinite at all P ∈ X if and only if u ∈ D(A,σ) (s × 1σ ) for some s ∈ N. We can use property (PS’) to reformulate the question of Procesi and Schacher and obtain a full characterization of those F -algebras with involution for which (PS’) holds: F = Xσ . Theorem 4.19. Property (PS’) holds if and only if X F = Xσ . Then (PS) equals (PS’) and the conclusion follows from Proof. Assume that X Corollary 4.18. Conversely, assume that (PS’) holds. Since 1 ∈ D(A,σ) (1σ ), the form F by (PS’) and, since T(A,σ,1) is nonsingular, it is T(A,σ,1) is positive semidefinite on X  F , i.e. in fact positive definite on XF . It follows from (4.1) that σ = σ1 is positive on X F = Xσ . 2 X Acknowledgments We are very grateful to the Referee for many pertinent comments that helped improve the original version of this paper. Appendix A. The nonsingular part of a hermitian form: general case Let (A, σ) be an F -algebra with involution, where we allow Z(A) = F × F . In case Z(A) = F × F , the involution σ is unitary and (A, σ) ∼ = (B × B op , ) for some simple F -algebra B and where  is the exchange involution (x, y) → (y, x), cf. [14, Proposition 2.14]. A standard argument shows: Lemma A.1. Let (M, h) be a hermitian form over (A, σ) and let x ∈ A be such that h(x, x) ∈ A× . Then we have M = xA ⊥ (xA)⊥h .

360

V. Astier, T. Unger / Journal of Algebra 508 (2018) 339–363

Lemma A.2. Let (E, τ ) be an F -division algebra with involution or let (E, τ ) = (D × Dop , ), where D is an F -division algebra. Let (M, h) be a hermitian form over (A, τ ). Then h  h0 ⊥ a1 , . . . , ak σ , where h0 is the zero form of the appropriate rank, and a1 , . . . , ak ∈ Sym(E, τ ) \ {0} = Sym(E, τ )× . The integer k and the rank of h0 are uniquely determined by h. The form a1 , . . . , ak σ is nonsingular and is uniquely determined up to isometry by h. Proof. Since D(E,τ ) (h) \ {0} ⊆ E × , using Lemma A.1 and induction on dimF M , we obtain h  h0 ⊥ a1 , . . . , ak σ , as in the statement, with h0 defined on M0 such that h0 (x, x) = 0 for every x ∈ M0 . The fact that h0 is the zero form follows from Lemma 2.8 if E is division and from the following argument if (E, τ ) = (D × Dop , ): if h0 = 0, there are linearly independent x, y ∈ M0 such that h0 (x, y) = 0. The computation of h0 (z, z) for z ∈ xE ⊕yE show that there exists z0 ∈ M0 such that h0 (z0 , z0 ) = 0, a contradiction. The fact that rk h0 , and hence k, is uniquely determined follows from Witt cancellation, cf. [6, Theorem 6.1]. The form a1 , . . . , ak σ is non-singular since all ai are invertible, and its uniqueness up to isometry follows again from Witt cancellation. 2 Proposition A.3. Let h be a hermitian form over (A, σ). Then h  h0 ⊥ hns where h0 is the zero form of appropriate rank and hns is nonsingular. Furthermore, h0 and hns are unique up to isometry. Proof. Assume first that (A, σ) is split symplectic. Then h is Morita equivalent to a skew-symmetric form over F , for which the result is well-known, cf. [9, p. 80, Corollaire 1]. The statement follows since Morita equivalence preserves orthogonal sums, zero forms, and nonsingularity. Assume now that (A, σ) is not split symplectic. Recall that (A, σ) is Morita equivalent to either (D, ϑ) or (D × Dop , ) where D is a division algebra, ϑ is an F -linear involution on D, and  is the exchange involution on D × Dop . Since (A, σ) is not split symplectic, we can assume that h corresponds to a hermitian form over (D, ϑ) or (D × Dop , ) under this equivalence. By Morita theory, the result now follows from Lemma A.2. 2 Appendix B. F -algebras with unitary involution and centre F × F The main objective of this appendix is to justify why the case Z(A) = F × F is of no interest from the point of view of this paper. Without loss of generality we assume that (A, σ) = (B × B op , ) with B = Mn (D) for some F -division algebra D. Our previous treatment of signature avoided this case, since any meaningful signature in our sense will be zero: If h is a hermitian form over (A, σ), by Proposition A.3 we have h = h0 ⊥ hns . Clearly, the zero form h0 should have signature 0. Since we want the signature to be

V. Astier, T. Unger / Journal of Algebra 508 (2018) 339–363

361

additive, the signature of h must be the signature of hns . Finally, the signature must be a morphism from W (A, σ) to Z, so the signature of h must be 0 since W (A, σ) = 0, cf. [4, Lemma 2.1(iv)]. If follows in particular that F = ∅. Lemma B.1. X Lemma B.2. Let Δ be a division ring and let V be a vector space over Δ. Let f ∈ EndΔ (V ). Then there are invertible f1 , f2 ∈ EndΔ (V ) such that f = f1 + f2 . Proof. This is a direct linear algebra argument, considering ker f . 2 Lemma B.3. Let a ∈ Sym(A, σ). Then there are a1 , a2 ∈ Sym(A, σ)× such that a = a1 + a2 . Proof. To simplify notation, we assume that A = Mn (D) × Mn (Dop ). Let a = (x, y) ∈ Sym(A, σ). Since σ(x, y) = (y, x) we have a = (x, x), with x ∈ Mn (D). By Lemma B.2 there are y, z ∈ Mn (D)× such that x = y + z, so (x, x) = (y, y) + (z, z) and (y, y), (z, z) ∈ Sym(A, σ)× . 2 If M is a left A-module, we denote by M the right A-module where the multiplication is twisted by σ as follows: ma := σ(a)m for m ∈ M and a ∈ A. Let P be a right A-module. We define the twisted dual module of P by P ∗ := HomA (P, A). Consider the hyperbolic space associated to P , i.e. the hermitian module (H(P ), hP ) where H(P ) := P ⊕ P ∗ and hP (x1 + f1 , x2 + f2 ) := f1 (x2 ) + σ(f2 (x1 )) for xi ∈ P and fi ∈ P ∗ , cf. [13, Chapter I, (3.5)]. Any hermitian space isometric to some H(P ) is called hyperbolic. Proposition B.4. (i) HomD×Dop (D, D × Dop ) ∼ = Dop as right D × Dop -modules. In particular H(D) ∼ = op op D ⊕ D as right D × D -modules, with the natural action of D × Dop , (d1 , d2 ) · (e1 , e2 ) := (d1 e2 , d2 ×op e2 ). (ii) HomD×Dop (Dop , D × Dop ) ∼ = D as right D × Dop -modules. In particular H(Dop ) ∼ = op op D ⊕ D as right D × D -modules, with the natural action of D × Dop , (d1 , d2 ) · (e1 , e2 ) := (d1 ×op e2 , d2 e1 ). Proof. The structure of right D × Dop -modules of D and Dop comes from the fact that they are minimal right ideals of D × Dop . The proof of both statements is an elementary, but tedious, computation of the various isomorphisms. 2

362

V. Astier, T. Unger / Journal of Algebra 508 (2018) 339–363

Corollary B.5. (i) H(D) ∼ = H(Dop ) as D × Dop -modules. (ii) Let P and Q be right D ×Dop -modules of the same finite rank. Then H(P ) ∼ = H(Q). Proof. (i) follows from Proposition B.4 by a direct computation. (ii) Write P = P1 ⊕ · · · ⊕ Pk and Q = Q1 ⊕ · · · ⊕ Qk with the Pi and Qi simple right D ×Dop -modules. Then H(P ) ∼ = H(P1 ) ⊕· · ·⊕H(Pk ) and H(Q) ∼ = H(Q1 ) ⊕· · ·⊕H(Qk ). op By (i) and since D and D are the only simple right D ×Dop -modules, we have H(Pi ) ∼ = H(Qi ) and the result follows. 2 Proposition B.6. Any two hyperbolic spaces of the same finite rank over (A, σ) are isometric. Proof. By Morita theory, it suffices to consider the case (D × Dop , ). Let P and Q be right D × Dop -modules such that H(P ) and H(Q) are of the same finite rank. Then P and Q have the same rank and H(P ) ∼ = H(Q) by Corollary B.5(ii) and therefore (H(P ), hP )  (H(Q), hQ ) by [13, p. 15]. 2 Lemma B.7. Let h be a nonsingular hermitian form over (A, σ). Then there exists k ∈ N such that Sym(A, σ)× ⊆ D(A,σ) (k × h). Proof. Let a ∈ Sym(A, σ)× . The forms rk(h) × aσ and rk(aσ ) × h are hyperbolic (since W (A, σ) = 0) and of the same rank. By Proposition B.6 they are isometric and so a ∈ D(A,σ) (rk(aσ ) × h). The result follows since rk(aσ ) does not depend on a. 2 Proposition B.8. Let h be a nonsingular hermitian form over (A, σ). Then there exists k ∈ N such that Sym(A, σ) ⊆ D(A,σ) (k × h). Proof. Let k be the integer obtained in Lemma B.7. Then Sym(A, σ)× + Sym(A, σ)× ⊆ D(A,σ) (2k × h) and the result follows by Lemma B.3. 2 Observe now that in Theorems 3.3 and 3.6, property (i) is always trivially satisfied by Lemma B.1 and property (ii) is always true by Proposition B.8. References [1] A.A. Albert, On involutorial algebras, Proc. Natl. Acad. Sci. USA 41 (1955) 480–482. [2] V. Astier, T. Unger, Positive cones on algebras with involution, http://arxiv.org/abs/1609.06601, 2017. [3] V. Astier, T. Unger, Stability index for algebras with involution, Contemp. Math. 697 (2017) 41–50.

V. Astier, T. Unger / Journal of Algebra 508 (2018) 339–363

363

[4] V. Astier, T. Unger, Signatures of hermitian forms and the Knebusch trace formula, Math. Ann. 358 (3–4) (2014) 925–947. [5] V. Astier, T. Unger, Signatures of hermitian forms and “prime ideals” of Witt groups, Adv. Math. 285 (2015) 497–514. [6] E. Bayer-Fluckiger, D.A. Moldovan, Sesquilinear forms over rings with involution, J. Pure Appl. Algebra 218 (3) (2014) 417–423. [7] E. Bayer-Fluckiger, R. Parimala, Galois cohomology of the classical groups over fields of cohomological dimension  2, Invent. Math. 122 (2) (1995) 195–229. [8] E. Bayer-Fluckiger, R. Parimala, Classical groups and the Hasse principle, Ann. of Math. (2) 147 (3) (1998) 651–693. [9] N. Bourbaki, Éléments de mathématique. Algèbre. Chapitre 9, Reprint of the 1959 original SpringerVerlag, Berlin, 2007. [10] P.K. Draxl, Skew Fields, London Math. Soc. Lecture Note Ser., vol. 81, Cambridge University Press, Cambridge, 1983. [11] D. Gondard, P. Ribenboim, Le 17e problème de Hilbert pour les matrices, Bull. Sci. Math. (2) 98 (1) (1974) 49–56. [12] I. Klep, T. Unger, The Procesi–Schacher conjecture and Hilbert’s 17th problem for algebras with involution, J. Algebra 324 (2) (2010) 256–268. [13] M.-A. Knus, Quadratic and Hermitian Forms over Rings, Grundlehren Math. Wiss., vol. 294, Springer-Verlag, Berlin, 1991. [14] M.-A. Knus, A. Merkurjev, M. Rost, J.-P. Tignol, The Book of Involutions, Colloq. Publ., vol. 44, American Mathematical Society, Providence, RI, 1998. [15] T.Y. Lam, Introduction to Quadratic Forms over Fields, Grad. Stud. Math., vol. 67, American Mathematical Society, Providence, RI, 2005. [16] D.W. Lewis, J.-P. Tignol, On the signature of an involution, Arch. Math. (Basel) 60 (2) (1993) 128–135. [17] D.W. Lewis, T. Unger, A local-global principle for algebras with involution and Hermitian forms, Math. Z. 244 (3) (2003) 469–477. [18] D.W. Lewis, T. Unger, Hermitian Morita theory: a matrix approach, Irish Math. Soc. Bull. (62) (2008) 37–41. [19] C. Procesi, M. Schacher, A non-commutative real Nullstellensatz and Hilbert’s 17th problem, Ann. of Math. (2) 104 (3) (1976) 395–406. [20] A. Quéguiner, Signature des involutions de deuxième espèce, Arch. Math. (Basel) 65 (5) (1995) 408–412. [21] W. Scharlau, Induction theorems and the structure of the Witt group, Invent. Math. 11 (1970) 37–44. [22] W. Scharlau, Quadratic and Hermitian Forms, Grundlehren Math. Wiss., vol. 270, Springer-Verlag, Berlin, 1985. [23] A. Weil, Algebras with involutions and the classical groups, J. Indian Math. Soc. (N.S.) 24 (1961) 589–623, 1960.