Skip to content
Publicly Available Published by De Gruyter March 20, 2020

Injective stability for odd unitary K1

  • Egor Voronetsky EMAIL logo
From the journal Journal of Group Theory

Abstract

We give a new purely algebraic approach to odd unitary groups using odd form rings. Using these objects, we give a self-contained proof of injective stability for the odd unitary K1-functor under the stable rank condition.

1 Introduction

The modern definition of odd unitary groups was given in [9] by Victor Petrov. His definition generalizes Anthony Bak’s unitary groups and split odd orthogonal groups, hence all classical Chevalley groups over arbitrary commutative rings. In [19], we introduced quadratic structures that may be used to construct these groups in a more geometric way.

The natural problem in this context is to prove stabilization for the K1-functor. For general linear groups, this was done in the groundbreaking paper [6] by Hyman Bass and also may be found in his book [7]. Injective stability for the linear K1-functor was reproved in [14] by Suslin and Tulenbaev using simple factorization of the elementary group. For the usual classical groups and Bak’s unitary groups, stability was proved in [2, 4, 15, 16] and in the unpublished paper [10], which can be found on Max-Albert Knus’s homepage. Finally, for Petrov’s unitary groups, the surjective part was already proved in [9], and the injective part in the main result of Yu Weibo’s paper [20]. All injectivity proofs for various unitary groups used the corresponding result for linear groups.

For unitary groups, the condition for injective stability is formulated in terms of the Λ-stable rank condition starting from [2]. This condition is weaker than all the previous ones; in particular, Λ-stable rank may be estimated by the absolute stable rank. For algebras R over a commutative ring K, the absolute stable rank may be bounded by the Bass–Serre dimension of K (more precisely, asr(R)δ(K)+1). On the other hand, in some cases, the injective stability was proved using only the stable rank of R. For Chevalley groups, it was done in Stein’s paper [13], and for even unitary groups, it is the main result of S. Sinchuk’s work [11].

In his paper [1], Bak proved the nilpotence of linear K1-groups using the localization-completion method. This result is much stronger than surjective stability and may be proved independently. The nilpotence of Bak’s unitary K1-groups was proved in the absolute case by Hazrat in [8] and in the relative case by Bak, Hazrat, and Vavilov in [3]. For Petrov’s unitary groups, it was done in a recent paper [22] of Weibo and Tang.

When the authors finished this work, it turned out that Weibo also proved injective stability under the stable rank condition; see [21]. In contrast to his paper, our approach is based on a more refined factorization than the usual Dennis–Vaserstein factorization of the elementary group (see [18] for details on various parabolic factorizations). Ultimately, this approach makes the proof self-contained without even using linear stability. Also, we use a more general definition of odd unitary groups. With this definition, the classical result for linear groups is a corollary of our main theorem, and in this case, our proof reduces to that of Suslin and Tulenbaev.

In the paper [19], we already implicitly used odd form rings in the definition of levels. A (unital special) odd form ring is a pair (R,Δ), where R is a ring with involution and Δ is Petrov’s odd form parameter on the right R-module R with the canonical hermitian form. As was shown in that paper, unitary groups of arbitrary regular quadratic bimodules may be considered as unitary groups of appropriate odd form rings. We show that the regularity condition is redundant.

In order to prove relative results for various groups, the common approach is Stein’s relativization technique from [12]. However, even for Bak’s unitary groups, it becomes complicated because the definition of Bak’s form rings is not truly algebraic (in the sense of universal algebra). This difficulty leads to the relativization with two parameters; see [5] for Bak’s unitary groups and [9] for Petrov’s unitary groups. Our odd form rings are defined using operations and axioms; hence the original Stein variant may be applied.

In Section 2, we review odd quadratic modules from [19], and in Sections 35, we develop the general theory of odd form rings. The main result of Section 6 is the following.

Theorem.

Let (R,Δ) be an odd form ring with a free orthogonal family of rank n, and suppose that sr(R1)n-2. Then the map

KU1(n-1;R,Δ)KU1(n;R,Δ)

is injective. If (I,Γ)(R,Δ) is an odd form ideal, then the map

KU1(n-1;R,Δ;I,Γ)KU1(n;R,Δ;I,Γ)

is also injective.

2 Quadratic and hermitian forms

Every ring in this paper is associative but not necessarily with 1. All commutative rings have identity elements and only trivial involutions; homomorphisms between them are unital. When we work with algebras over a fixed commutative ring K, we always consider only those bimodules that have the same K-module structure from the left and from the right. Recall that if R is a non-unital K-algebra, then RK is a unital K-algebra with an ideal R; it is the K-module RK with the multiplication (r,k)(r,k)=(rr+rk+rk,kk).

If R is an arbitrary ring, then R is the multiplicative semigroup of R (it is a monoid if R is unital), i.e. the set R with the multiplication operation. By C(G) and C(R), we denote the center of a group G or a ring R. The subgroups CG(H) and NG(H) are the centralizer and the normalizer of a subgroup H in a group G. We use the notation hg=ghg-1, hg=g-1hg and [g,h]=ghg-1h-1 for elements g and h of an arbitrary group.

If R is a ring, then Rop={roprR} is the opposite ring, rop(r)op=(rr)op. The same notation is used for the opposite modules (note that the opposite of a left module is right and vice versa). Clearly, (Rop)opR, and similarly for modules. An involution on a ring R is an additive map (-)¯:RR such that rr¯=r¯r¯, and r¯¯=r (if R is unital it follows that 1¯=1). For example, if R=Sop×S for a ring S, then (aop,b)¯=(bop,a) is an involution.

Until the end of this section, all rings, ring homomorphisms and modules are unital.

Recall the definitions from [19]. Let R be an arbitrary ring and λR*. A map (-)¯:RR,rr¯ is called a λ-involution (or pseudo-involution) if it is additive,

1¯=1,rr¯=r¯r¯,r¯¯=λrλ-1andλ¯=λ-1.

For example, if R is commutative, then the identity map on R is a 1-involution (i.e. an involution) and a (-1)-involution simultaneously. Let MR be a right module. A map B:M×MR is called a hermitian form if it is biadditive,

B(m,mr)=B(m,m)randB(m,m)=B(m,m)¯λ.

The hermitian form B is called regular (or non-degenerate) if MR is finitely generated projective and B induces an bijection MHomR(M,R),mB(m,-). The module RR has the canonical regular hermitian form B1(r,r)=r¯r.

A quadratic structure on a ring R with a λ-involution is a right R-module A (i.e. an abelian group with a right action of the monoid R) with additive maps φ:RA and tr:AR such that

  1. φ(r¯rr)=φ(r)r, tr(ar)=r¯tr(a)r,

  2. tr(φ(r))=r+r¯λ,

  3. tr(a)=tr(a)¯λ, φ(r)=φ(r¯λ),

  4. a(r+r)=ar+φ(r¯tr(a)r)+ar.

For example, if ΛR is a form parameter (i.e. Λ is an additive subgroup, {r-r¯λ}Λ{rr+r¯λ=0} and r¯ΛrΛ for all r), then A=R/Λ is a quadratic structure with φ(r)=r+Λ and tr(r+Λ)=r+r¯λ.

Let (MR,B) be a hermitian module and A a quadratic structure on R. A quadratic form on M with values in A is a map q:MA such that

  1. q(mr)=q(m)r,

  2. tr(q(m))=B(m,m),

  3. q(m+m)=q(m)+φ(B(m,m))+q(m).

Such a triple (M,B,q) is called a quadratic module. The unitary group of a quadratic module (M,B,q) is

U(M,B,q)={gAutR(M)B(gm,gm)=B(m,m),q(gm)=q(m)for allm,m}.

In the case of Bak’s quadratic forms, we have A=R/Λ for a form parameter Λ,

B(m,m)=Q(m,m)+Q(m,m)¯λ

for a sesquilinear map Q, and q(m)=φ(Q(m,m)). Quadratic modules with regular hermitian forms are also called quadratic spaces.

As was shown in [19], Petrov’s quadratic forms are almost the same as our quadratic forms if A is generated by the images of φ and q. Indeed, recall that the Heisenberg group of (M,B) is Heis(M,B)=M×R with the operation

(m,r)(m,r)=(m+m,r-B(m,m)+r)

(the identity element 0˙=(0,0) and the inverses -˙(m,r)=(-m,-B(m,m)-r)). The monoid R acts on the group Heis(M,B) from the right via

(m,r)r=(mr,r¯rr),

and there are natural maps

φ:RHeis(M,B),r(0,r),
tr:Heis(M,B)R,(m,r)B(m,m)+r+r¯λ,
q:MHeis(M,B),m(m,0).

A Petrov’s odd form parameter is an R-subgroup Heis(M,B) such that

{(0,r-r¯λ)}Ker(tr).

Then φ, tr and q are well-defined on A=Heis(M,B)/. Conversely, if a quadratic structure A with a fixed quadratic form q is generated by the images of φ and q, then A is isomorphic to Heis(M,B)/ for a unique .

Recall also that if (P,B,q) and (P,B,q) are quadratic modules over a ring R with a λ-involution and a quadratic structure A, then

(P,B,q)(P,B,q)=(PP,BB,qq)

is a quadratic module, where

(BB)(p1p1,p2p2)=B(p1,p2)+B(p1,p2),
(qq)(pp)=q(p)+q(p).

Conversely, if a quadratic module (P,B,q) splits as a direct sum of orthogonal submodules P=P1P2 (i.e. B(P1,P2)=0), then P=P1P2.

Let P be a finitely generated projective module over a ring R with a λ-involution and a quadratic structure A. Then Q=HomR(P,R)op is also an R-module and the map -,=:Q×PR,(xop,p)x(p) is the canonical pairing. Now the module H(P)=QP has a hermitian form

B(xopp,xopp)=xop,p+xop,p

and a quadratic form

q(xopp)=φ(xop,p).

The module H(P) is called a hyperbolic space (clearly, the hermitian form of H(P) is regular), P and Q are the lagrangians of H(P). Conversely, if M is a quadratic space and there is a direct sum decomposition M=PQ such that B|Q×Q=0=B|P×P and q|Q=0=q|P, then MH(P).

Note that if (N,B|N×N,q|N)(M,B,q) is a regular submodule of a quadratic module, then M canonically splits as an orthogonal sum of N and its orthogonal complement N={mMB(m,N)=0}. We usually consider the situation when a given quadratic module (M,B,q) has several pairwise orthogonal hyperbolic submodules Mi=H(Pi), 1in, i.e.

Mi=PiP-i,B|Pi×Pi=0=B|P-i×P-i,qPi=0=qP-i.

The orthogonal complement to all Mi is denoted by M0. If Pi=R, 0<|i|n, then M is exactly an odd hyperbolic space of rank n from [9].

3 Odd form rings

Let (M,B,q) be a quadratic module over a unital ring R with a λ-involution and a quadratic structure A. Consider the ring

T={(xop,y)EndR(M)op×EndR(M)B(xm,m)=B(m,ym)for allm,m};

clearly, T possess an involution (xop,y)¯=(yop,x). If B is regular, then

TEndR(M),(xop,y)y.

Also, there is the odd form parameter

Ξ={((xop,y),(zop,w))q(ym)+φ(B(m,wm))=0for allm,xy+z+w=0}Heis(T,B1).

It is easy to see that U(M,B,q)U(T,B1,qΞ), where

qΞ(a)=(1,0)aΞHeis(T,B1)/Ξ.

Conversely, let T be a unital involution ring, let ΞHeis(T,B1) be an odd form parameter, and let U(T,B1,qΞ) be the corresponding unitary group. We call the pair (T,Ξ) a special unital odd form ring. Clearly, there are natural maps

π:ΞT,(a,b)a,
ρ:ΞT,(a,b)b,
ϕ:TΞ,aφ(a-a¯)=(0,a-a¯).

For any gU(T,B1) (i.e. if gT* and g-1=g¯), let

γ(g)=-˙qΞ(g)qΞ(1)=(g-1,g¯-1);

then U(T,B1,qΞ)={gT*g-1=g¯,γ(g)Ξ}. Note that π(γ(g))=g-1 and ρ(γ(g))=g¯-1. In other words, we may express the unitary group using only Ξ instead of the full Heisenberg group.

Now we define odd form rings using axioms. A pair (R,Δ) is called an odd form ring if R is a ring with involution (non-unital in general), Δ is a group with a right R-action, and there are maps ϕ:RΔ, π:ΔR, ρ:ΔR such that

  1. π is a group homomorphism, π(ux)=π(u)x,

  2. ϕ is a group homomorphism, ϕ(x¯yx)=ϕ(y)x, ϕ(z)=0 if z=z¯,

  3. uϕ(x)=ϕ(x)u, uv=vuϕ(-π(u)¯π(v)),

  4. ρ(uv)=ρ(u)-π(u)¯π(v)+ρ(v), ρ(u)¯=ρ(-˙u), ρ(ux)=x¯ρ(u)x,

  5. π(ϕ(x))=0, ρ(ϕ(x))=x-x¯,

  6. u(x+y)=uxϕ(y¯ρ(u)x)uy.

Note that, in every odd form ring, we have

ϕ(x¯)=-˙ϕ(x),ρ(0˙)=0,ρ(-˙u)+ρ(u)+π(u)¯π(u)=0,u0=0˙.

An odd form ring (R,Δ) is called unital if R is unital and u1=u for all uΔ; for such odd form rings, the identity u(-1)u=ϕ(ρ(u)) holds.

Clearly, any special unital odd form ring is an odd form ring. Conversely, if (R,Δ) is an odd form ring and (π,ρ):ΔR×R is injective, then (R,Δ) is called special (and if R is unital, then (R,Δ) is a special unital odd form ring as in the definition above). Our odd form rings are preferable in comparison with the special ones since they behave better under Stein’s relativization and other algebraic constructions like tensor products.

An odd form ideal of an odd form ring (R,Δ) is a pair (I,Γ) such that IR is a two-sided ideal, ΓΔ is a normal subgroup,

I=I¯,ΓRΓandΓminΓΓmax,

where Γmin=ΔIϕ({rr-r¯I}) and Γmax={uΔπ(u),ρ(u)I} (clearly, (I,Γmin) and (I,Γmax) are odd form ideals). If (I,Γ)(R,Δ) is an odd form ideal, then (R,Δ)/(I,Γ)=(R/I,Δ/Γ) is an odd form ring. This does not always hold for special odd form rings since the factor may not be special.

We say that an odd form ring (R,Δ) is an odd form algebra over a commutative ring K if R is an involution K-algebra and (RK,Δ) is a unital odd form ring (i.e. the action of K on Δ is defined and satisfies appropriate identities). Any odd form algebra (R,Δ) naturally becomes an ideal in a unital odd form algebra (RK,Δ). Any odd form ring is an odd form -algebra.

Now let us define the unitary group of an odd form ring (R,Δ). Since the ring may be non-unital, we should use quasi-regular elements of R instead of invertible ones. Also, we need to explicitly use the odd form parameter; hence we define the unitary group as a certain subset of R×Δ. The components of gR×Δ are denoted as β(g)R and γ(g)Δ. To simplify further computations, it is useful to set α(g)=β(g)+1R (we may consider α(g) as an element of RK if (R,Δ) is an odd form K-algebra, or even as an element of R itself if R is unital). Finally,

U(R,Δ)={gR×Δα(g)-1=α(g)¯,π(γ(g))=ρ(γ(g))¯=β(g)}.

The first equation in the definition of U(R,Δ) may be written as

β(g)β(g)¯+β(g)+β(g)¯=β(g)¯β(g)+β(g)+β(g)¯=0.

The group operation is given by α(gh)=α(g)α(h), γ(gh)=γ(g)α(h)γ(h). Note that if (R,Δ) is special unital, then we may identify an element gU(R,Δ) with α(g)R.

Lemma 1.

There are the following identities:

  1. β(1)=0, β(gh)=β(g)β(h)+β(g)+β(h), β(g-1)=β(g)¯,

  2. γ(1)=0˙, γ(gh)=γ(g)β(h)γ(h)γ(g), γ(g-1)=-˙-˙-˙γ(g)γ(g)β(g)¯ϕ(β(g)¯2),

  3. β(hg)=βα(g)(h)=(β(g)β(h)+β(h))α(g)¯,

  4. γ(hg)=(γ(g)β(h)γ(h))α(g)¯,

  5. β([g,h])=(βα(g)(h)-β(h))α(h)¯=(β(g)β(h)-β(h)β(g))α(g)¯α(h)¯,

  6. γ([g,h])=(-˙γ(g)β(h)γ(h)β(g)ϕ(β(g)¯β(h)))α(g)¯α(h)¯.

Proof.

This follows from direct calculations. ∎

For example, if (R,Δ) is obtained from a quadratic module (M,B,q), then U(R,Δ)U(M,B,q). Also, if R is a C*-algebra, then

Δ={(x,y)Heis(R)x¯x+y+y¯=0}

is an odd form parameter, (R,Δ) is a special odd form -algebra (it is not a -algebra by our definition since the involution is anti-linear) and U(R,Δ) is isomorphic to the unitary group of R.

If (R,Δ) is an arbitrary odd form ring, M is a right R-module, then (R,ΔM) is also an odd form ring with π(um)=π(u) and ρ(um)=ρ(u). It is easy to see that U(R,ΔM)MU(R,Δ). In particular, various affine groups like KnGL(n,K) are unitary groups of certain non-special odd form rings.

Now we formulate Stein’s relativization in our context. Let (I,Γ)(R,Δ) be an odd form ideal. The double (R,Δ)×(R/I,Δ/Γ)(R,Δ) is the fiber product of (R,Δ) with itself over (R/I,Δ/Γ). In other words,

R×R/IR={(x,y)R×Rx-yI},
Δ×Δ/ΓΔ={(u,v)Δ×Δ-˙uvΓ}.

The projections from the double to (R,Δ) are denoted by p1 and p2, the factor-map (R,Δ)(R/I,Δ/Γ) is denoted by q, and the diagonal map from (R,Δ) into the double is denoted by d (so pid=id). The double is canonically isomorphic to (IR,ΓΔ), where IR=IR as an abelian group, ΓΔ=Γ×Δ as a set (actually, it is the semi-direct product of groups), and the operations are given by

  1. (x,y)(x,y)=(xx+xy+yx,yy),

  2. (u,v)(u,v)=(uuϕ(-π(v)¯π(u)),vv),

  3. (u,v)(x,y)=(uxuyvxϕ(y¯(ρ(u)+ρ(v))x),vy),

  4. π(u,v)=(π(u),π(v)), ϕ(x,y)=(ϕ(x),ϕ(y)), ρ(u,v)=(ρ(u)-π(u)¯π(v),ρ(v)).

The isomorphism (IR,ΓΔ)(R×R/IR,Δ×Δ/ΓΔ) is given by

(x,y)(x+y,y)and(u,v)(uv,v),

and in terms of (IR,ΓΔ), we have p1(x,y)=x+y, p1(u,v)=uv, p2(x,y)=y, p2(u,v)=v, d(x)=(0,x) and d(u)=(0˙,u).

Now if (I,Γ)(R,Δ) is an odd form ideal, then there is a left exact sequence

1U(I,Γ)U(R,Δ)U(R/I,Δ/Γ).

Obviously, U(I,Γ)U(R,Δ). It is easy to see that the sequence

1U(I,Γ)p1-1U(IR,ΓΔ)p2U(R,Δ)1

is split exact with the section d, i.e. is a semi-direct product (where p1-1 takes values in Ker(p2)). It follows from the fact that the functor (R,Δ)U(R,Δ) commutes with fibered products.

4 Idempotents

We start with a construction of the free odd form algebra.

Proposition 1.

Let K be a commutative ring, and let A and B be two sets. Then there is the universal odd form algebra (R,Δ) over K freely generated by elements {xaR}aA and {ubΔ}bB. More explicitly, R is the free K-module with generators r1rm for all m>0, where each rs is equal to one of xa, x¯a, π(ub), π(ub)¯, ρ(ub) or ρ(ub)¯ for some aA or bB, and there are no consecutive factors of type rs=π(ub)¯, rs+1=π(ub). Similarly, Δ has elements ubr1rm for all bB, m0, where r1rm is the K-module generator of R for m>0, such that these elements form a K-module basis of Δ/ϕ(R).

Also, if (R,Δ) is an arbitrary odd form ring and (I,Γ)(R,Δ) is the odd form ideal generated by elements {xaR}aA and {ubΔ}bB, then IR is the ideal generated by all xa, xa¯, π(ub), π(ub)¯, ρ(ub), and

Γ=˙bubRϕ({rr-r¯I}).

Proof.

Let R=r1rmKr1rm, where the direct sum is taken by all products from the statement. It is a K-module with an obvious involution. In order to multiply two generators r1rm and r1rm, we just concatenate them and use the relation π(ub)¯π(ub)=-ρ(ub)-ρ(ub)¯ in order to eliminate bad consecutive pairs of factors. Clearly, this gives a well-defined multiplication on R; hence R is an involution K-algebra.

Now let Δ be the group generated by elements ϕ(r) for rR and ubr for bB, rRK with the relations

  1. ϕ(r)ϕ(s)=ϕ(r+s), ϕ(t)=0˙ if t=t¯,

  2. ϕ(r)ubr=ubrϕ(r),

  3. ubrubr=ubrubrϕ(-r¯π(ub)¯π(ub)r),

  4. ub(r+r)=ubrϕ(r¯ρ(ub)r)ubr.

It is easy to see that Δ/ϕ(R)=bub(RK). The maps π, ρ and the R-action are well-defined by the obvious formulas and satisfy all axioms. By construction, (R,Δ) has the universal property.

The second claim is obvious: the pair (I,Γ) is indeed an odd form ideal containing the generators, and it is the smallest such odd form ideal. ∎

Let (I,Γ)(R,Δ) be an odd form ideal in a unital odd form ring, and let eR be a hermitian idempotent (i.e. e=e¯ and e2=e). It is easy to see that (Ie,Γee)(I,Γ) is an odd form subalgebra, where

Ie=eIe,Γee={uΓeπ(u)Ie}andΓe=Γe.

Usually, we apply this to an arbitrary non-unital odd form K-algebra

(R,Δ)(RK,Δ).

If (R,Δ) is obtained from a quadratic module (M,B,q), then there is a natural bijection between hermitian idempotents of R and orthogonal summands NM; under this bijection, (Re,Δee) are obtained from (N,B|N×N,q|N).

Now let us see what happens when a quadratic module (M,B,q) has a family of orthogonal hyperbolic summands, M=M0H(P1)H(Pn) for some n0. Let EiEnd(M) be the canonical projections on Pi for 0<|i|n and on M0 for i=0; then Ei form a complete system of orthogonal idempotents. Let ei=(E-iop,Ei) for -nin; then

eiT={(xop,y)EndR(M)op×EndR(M)B(xm,m)=B(m,ym)}.

Moreover, these ei also form a complete system of orthogonal idempotents, and ei¯=e-i. Also, (ei,0)Ξ for all i0, where Ξ is the odd form parameter considered above, since q|Pi=0.

We say that η=(e-,e+,q-,q+) is a hyperbolic pair in an odd form ring (R,Δ) if e- and e+ are orthogonal idempotents in R, e+¯=e-, q- and q+ lie in Δ,

π(q-)=e-,ρ(q-)=0,q-e-=q-,
π(q+)=e+,ρ(q+)=0,q+e+=q+

(it follows that q+e-=0˙ and q-e+=0˙). Usually, we work with several hyperbolic pairs η1,,ηn; in this case, we use the notation ηi=(e-i,ei,q-i,qi) and e|i|=e-i+ei. Finally, ea=1-eaR (they may be considered as elements of RK if (R,Δ) is an odd form K-algebra or as elements of R if R is unital), where a is an arbitrary index (integer or of type |i|).

Hyperbolic pairs η1 and η2 are called orthogonal if e|1| and e|2| are orthogonal idempotents. These pairs are called isomorphic if there are e12e1Re2 and e21e2Re1 such that e1=e12e21 and e2=e21e12. Finally, η1 and η2 are called Morita equivalent if e|1|Re|2|Re|1|=e|1|Re|1|, e|2|Re|1|Re|2|=e|2|Re|2|. Morita equivalence means exactly that the unital rings e|1|Re|1| and e|2|Re|2| are Morita equivalent with respect to the bimodules e|2|Re|1| and e|1|Re|2|. See [7, Chapter II] for general Morita theory.

An odd form ring (R,Δ) has an orthogonal hyperbolic family of rank n0 if there is a family η1,,ηn of pairwise orthogonal and Morita equivalent hyperbolic pairs. In this case, we set

e0=1-(e-n++e-1+e1++en)R.

If (R,Δ) is special unital, then we may consider R as a quadratic R-module with the hermitian form B=B1 (i.e. B(r,r)=r¯r) and the quadratic form given by the odd form parameter. In this case, an orthogonal hyperbolic family is the same as a decomposition R=i=-nneiR into direct summands such that e0R is orthogonal to all e|i|R for 0<in, these e|i|R are pairwise orthogonal and hyperbolic with the lagrangians e-iR and eiR, and also e|i|R are isomorphic as right R-modules to direct summands in (e|j|R)N for N large enough and for all 1i,jn.

Let (R,Δ) be an odd form ring with an orthogonal hyperbolic family. If XR is a subgroup closed under multiplications on all ei from the left and from the right, then we use the notation Xij=eiXej and Xi=eiXei. Similarly, if ΥΔ is a subgroup closed under right multiplications on all ei, then we use the notation Υi=Υei. Let e+=e1++en and e-=e+¯=e-n++e-1. The expressions Xi, X|i|, Υ+, and so on have the obvious meaning. Also, Xi* and X*i mean eiX and Xei. Sometimes, we use the notation

αij(g)=eiα(g)ej,βij(g)=eiβ(g)ejandγi(g)=γ(g)ei

for gU(R,Δ). Finally,

Υ0={uΥπ(u)R0*}andΥ|i|={uΥπ(u)R|i|,*}.

An orthogonal hyperbolic family η1,,ηn is called free if ηi are pairwise isomorphic and these isomorphisms are coherent. In other words, there are eijR for 1i,jn such that eii=ei and eijejk=eik. We set e-i,-j=eji¯ and e|i|,|j|=eij+e-i,-j for 1i,jn. In this case, there are canonical isomorphisms R0M(n,R|1|), R+M(n,R1) and R-M(n,R-1).

The following lemma is needed in the next section. It shows, in particular, that the free odd form algebra is special.

Lemma 2.

Let K be a commutative ring, let n0 be an integer, and let Aij, Bk, Ci, D be abstract sets for -ni,j,kn and k0. Let (R,Δ) be the universal odd form K-algebra with a family of n pairwise orthogonal hyperbolic pairs {(e-i,ei,q-i,qi)}i=1n and families of elements

{xijaRij}aAij,{ykbRk*}bBk,{uicΓi0}cCi,{gdU(R0,Δ00)}dD.

Then (R,Δ) is special.

Proof.

The odd form algebra (R,Δ) may be constructed by Proposition 1. The basis of R over K consists of ek for k0 and various products r1rm for m>0, where every rs is one of xija, xija¯, yk, yk-1, yk¯, yk¯-1, π(uic), π(uic)¯, ρ(uuc), ρ(uic)¯, β(gd) or β(gd)¯, such that all consecutive pairs are compatible (i.e. they lie in R*i and Ri*) and there are no such pairs of type ykyk-1, yk-1yk, yk¯yk¯-1, yk¯-1yk¯, π(uic)¯π(uic), β(gd)¯β(gd) or β(gd)β(gd)¯. The basis of Δ/ϕ(R) over K consists of qkr1rm, uicr1rm, γ(gd)r1rm for -ni,kn, k0 and m0, where r1rm lies in the basis of R and the products are compatible (i.e. of type ΓiRi*) for m>0.

Now it is easy to see that π:Δ/ϕ(R)R is injective. So if π(u)=ρ(u)=0 for some uΔ, then u=ϕ(r) for some rR and r-r¯=ρ(u)=0; hence u=0˙. ∎

5 Elementary transvections

In this section, (R,Δ) is an odd form ring with an orthogonal hyperbolic family of rank n. There are elements in the unitary group U(R,Δ) of a particularly simple structure. An elementary transvection of a short root type is an element Tij(x)U(R,Δ) such that

β(Tij(x))=x-x¯,γ(Tij(x))=-˙-˙qixq-jx¯ϕ(x)

for any i0, j0, i±j and xRij. An elementary transvection of an ultrashort root type is an element Ti(u)U(R,Δ) such that

β(Ti(u))=ρ(u)+π(u)-π(u)¯,
γ(Ti(u))=-˙uϕ(ρ(u)+π(u))q-i(ρ(u)-π(u)¯)

for any i0 and uΔi0=Δ0ei It easy to see that all these elements are indeed in the unitary group. The elementary unitary group is

EU(R,Δ)=Tij(x),Tk(u)i±j;i,j,k0;xRij;uΔk0.

An elementary dilation is an element Di(a)U(R,Δ) such that

β(Di(a))=a+a¯-1-e|i|,
γ(Di(a))=q-i(a¯-1-e-i)-˙qi(a-ei)ϕ(a-ei)

for any i0 and aRi*. Also, D0(g)=g for gU(R0,Δ00) (as an element of U(R,Δ)).

Lemma 3.

Elementary transvections and dilations have the following properties:

  1. Tij:RijU(R,Δ), Ti:Δi0U(R,Δ), Di:Ri*U(R,Δ) for i0 and D0:U(R0,Δ00)U(R,Δ) are group homomorphisms,

  2. Tij(x)=T-j,-i(-x¯), Di(a)=D-i(a¯-1) for i0,

  3. [Di(a),Dj(b)]=1 for i±j,

  4. [Di(a),Tjk(x)]=1 for j±ik,

  5. [Di(a),Tj(u)]=1 for 0i±j,

  6. TijDi(a)(x)=Tij(ax),

  7. Ti(u)Di(a)=Ti(ua),

  8. TiD0(g)(u)=Ti(γ(g)π(u)u),

  9. [Tij(x),Tkl(y)]=1 for il-j-ki,

  10. [Tij(x),Tjk(y)]=Tik(xy) for i±k,

  11. [T-i,j(x),Tji(y)]=Ti(ϕ(xy)),

  12. [Ti(u),Tj(v)]=T-i,j(-π(u)¯π(v)) for i±j;

  13. [Ti(u),Tjk(x)]=1 for ji-k,

  14. [Ti(u),Tij(x)]=T-i,j(ρ(u)x)Tj(-˙u(-x)).

Proof.

We prove these identities without the assumption that the hyperbolic pairs are Morita equivalent. By Lemma 2, we may assume that the odd form ring is special; hence it suffices to check that values of β on both sides of each identity coincide. This is easy, especially when using the expressions for β(hg) and β([g,h]) from Lemma 1. ∎

Let (I,Γ)(R,Δ) be an odd form ideal. Note that both (R/I,Δ/Γ) and (IR,ΓΔ) have canonical orthogonal hyperbolic families. The relative elementary unitary group is

EU(R,Δ;I,Γ)=Tij(x),Tk(u)xIij,uΓk0EU(R,Δ)U(I,Γ).

Lemma 4.

Let (R,Δ) be an odd form ring with an orthogonal hyperbolic family, (I,Γ)(R,Δ). Then the sequence

1EU(R,Δ;I,Γ)p1-1EU(IR,ΓΔ)p2EU(R,Δ)1

is split exact, i.e. a semi-direct product, where p1-1 takes values in Ker(p2). The section is given by d.

Proof.

Let N=p1-1(EU(R,Δ;I,Γ)) and G=d(EU(R,Δ)); then NG=1, [G,N]N (since EU(R,Δ;I,Γ) is normalized by EU(R,Δ)) and every generator of EU(IR,ΓΔ) lies in NG. Hence EU(IR,ΓΔ)=NG. ∎

Let Tij(*), Ti(*) and Di(*) be the groups of corresponding elementary transvections and dilations. Note that the diagonal group iDi(*) is the direct product of Di(*) for i0. Also, Tij(*)=T-j,-i(*) and Di(*)=D-i(*). The following two lemmas are easy consequences of definitions and Lemma 3, so we state them without proofs.

Lemma 5.

Let (R,Δ) be an odd form ring with an orthogonal hyperbolic family. Then the groups of elementary transvections and dilations have the following equations:

iDi(*)={gU(R,Δ)βij(g)=0𝑓𝑜𝑟ij,γi(g)=qiβi(g)𝑓𝑜𝑟i0},
D0(*)={gU(R,Δ)β(g)=β0(g),γ(g)=γ0(g)},
Ti(*)={gU(R,Δ)β(g)=β0i(g)+β-i,i(g)+β-i,0(g),γi(g)=q-iβ-i,0(g)}.

Lemma 6.

Let (R,Δ) be an odd form ring with an orthogonal hyperbolic family η1,,ηn. If Tij(*), Ti(*), Di(*) are the groups of elementary translations and dilations corresponding to the orthogonal hyperbolic family η2,,ηn, then

  1. Tij(*)=Tij(*) and Di(*)=Di(*) for 2|i|,|j|n,

  2. Ti(*)=Ti(*)(T1i(*)×T-1,i(*)) for 2|i|n,

  3. D0(*)T1(*),T-1(*),D1(*),D0(*).

Similarly, if Tij′′(*), Ti′′(*), Di′′(*) are the groups corresponding to the family η1,,ηn-2,ηn-1+ηn, where

ηn-1+ηn=(e1-n+e-n,en-1+en,q1-nq-n,qn-1qn),

then

  1. Tij′′(*)=Tij(*), Ti′′(*)=Ti(*) and Di′′(*)=Di(*) for |i|,|j|<n-1,

  2. Ti,n-1′′(*)=Ti,n-1(*)×Tin(*) for |i|<n-1,

  3. Tn-1′′(*)=T-n,n-1(*)(Tn-1(*)×Tn(*)) and similarly for T1-n′′(*),

  4. Dn-1′′(*)Tn,n-1(*),Tn-1,n(*),Dn(*),Dn-1(*).

6 Injective stability

Until the end of this section, (R,Δ) is an odd form ring with a free orthogonal hyperbolic family of rank n2. In the unitary group U(n;R,Δ)=U(R,Δ), there is a subgroup

U(n-1;R,Δ)={gU(n;R,Δ)β*,|n|(g)=β|n|,*(g)=0,γ|n|(g)=0˙}.

Clearly, U(n-1;R,Δ)=U(R|n|,Δ|n||n|). The elementary subgroup

EU(n-1;R,Δ)U(n-1;R,Δ)

is generated by all transvections Tij(x),Tj(u)EU(R,Δ) with i,j±n. For large n, we may define similarly U(n-2;R,Δ) and so on.

Conversely, if the orthogonal hyperbolic family is free, then we may define U(n+1;R,Δ). Indeed, in this case, (R,Δ) may be embedded into an odd form algebra (R~,Δ~) with a free orthogonal hyperbolic family of rank n+1 (unique up to the canonical isomorphism) such that R=R~|n+1| and Δ=Δ~|n+1||n+1|. In this case, we set U(n+1;R,Δ)=U(n+1;R~,Δ~)=U(R~,Δ~).

We need the notion of stable rank. Recall that the condition sr(R1)k-1 means that, for any right unimodular sequence x1,,xkR1 of length k (i.e. if there exist yiR1 with iyixi=1), there are a1,,ak-1R1 such that x1+a1xk,,xk-1+ak-1xk is right unimodular of length k-1. See Bass’s book [7] for more details. Vaserstein proved in [17] that the stable ranks of R1 and I1R1=(IR)1 coincide.

As usual, let

KU1(n;R,Δ)=U(n;R,Δ)/EU(n;R,Δ),
KU1(n;R,Δ;I,Γ)=U(n;I,Γ)/EU(n;R,Δ;I,Γ)

for an odd form ideal (I,Γ)(R,Δ). In general, these objects are just pointed sets. Note that there is a sequence

1KU1(n;R,Δ;I,Γ)(p1-1)*KU1(n;IR,ΓΔ)(p2)*KU1(n;R,Δ)1;

see Lemma 4. It follows from a simple diagram chase that this sequence is short exact in the following sense: (p1-1)* is injective, (p2)* is surjective, the image of (p1-1)* is exactly the preimage of the distinguished point under (p2)*. Moreover, there is a U(n;R,Δ)-equivariant section

d*:KU1(n;R,Δ)KU1(n;IR,ΓΔ)

of (p2)* and the action of U(n;R,Δ) on KU1(n;IR,ΓΔ) induces a transitive action on the fibers of (p2)*; hence there is a bijection

KU1(n;IR,ΓΔ)KU1(n;R,Δ;I,Γ)×KU1(n;R,Δ).

The bijection is not canonical in general; though if all elementary groups are normal in the corresponding unitary groups, then the sequence is split short exact with the splitting d*.

In order to prove injective stability, we show that the elementary group is a product of certain subgroups. Let T+[2,n](u), T-[2,n](u), D+[2,n](a), T-n(u) and T+n(u) be the elementary transvections and dilations with respect to the orthogonal hyperbolic families η2++ηn and ηn of rank 1. By Lemma 6, we have

  1. T+[2,n](*)=Tj(*),Tij(*)i1 and 2j,

  2. T-[2,n](*)=Tj(*),Tij(*)-1i and j-2,

  3. T-n(*)=T-n(*),Ti,-n(*)0<|i|n-1,

  4. T+n(*)=Tn(*),Ti,n(*)0<|i|n-1,

  5. D+[2,n](*)D+,el[2,n](*)=Tij(*)2i,jn.

We are going to prove that EU(n;R,Δ)=PLQ=PLD+,el[2,n](*)S, where

  1. P=EU(n-1;R,Δ)T+n(*) (an elementary maximal parabolic subgroup),

  2. L=Ti,-n(*),T-n(*)i-1T-n(*),

  3. Q=D+,el[2,n](*)T+[2,n](*) (part of a submaximal parabolic subgroup),

  4. S=Tj(*),Tij(*)i1 and 3jT+[2,n](*).

Note that T-[2,n](*)PL.

Lemma 7.

Let (R,Δ) be an odd form ring with a free orthogonal hyperbolic family of rank n, and suppose that sr(R1)n-2. Then PLQPLD+,el[2,n](*)S.

Proof.

Suppose that g=plq, where pP, lL and

q=D+[2,n](a)T+[2,n](u)Qfor someuanda.

By the stable rank condition, there is wT2,n(*),,Tn-1,n(*) such that

wD+[2,n](a)=D+[2,n](a)

with the sequence a2,2,,an-1,2 being unimodular. Note that pw-1P and lwT-[2,n](*). Hence

g=(pw-1)(lw)D+[2,n](a)T+[2,n](u)=plD+[2,n](a)T+[2,n](u)

for some pP and lL.

By unimodularity there is xe2R(e2++en-1) such that xae2=e2. Let v=ux; then ven=0˙ (so T+[2,n](v)EU(n-1;R,Δ)) and

g=(pT+[2,n](v))(lT+[2,n](v))D+[2,n](a)T+[2,n](-˙uxau),

Here

pT+[2,n](v)P,lT+[2,n](v)T-n(*)LD+,el[2,n](*),
D+[2,n](a)D+,el[2,n](*)andT+[2,n](-˙uxau)S.

Lemma 8.

Let (R,Δ) be an odd form ring with a free orthogonal hyperbolic family of rank n, and suppose that sr(R1)n-2. Then EU(n;R,Δ)=PLQ.

Proof.

Note that EU(n;R,Δ) is generated by Q and

T~=T-2(*),T-1,-2(*),T1,-2(*)

(by Lemma 3). Clearly, 1PLQ, and the set PLQ is closed under multiplications by elements of Q from the right. By Lemma 7, it remains to prove that PLD+,el[2,n](*)StPLQ for any tT~. Note that

StTj(*),Tij(*)i2andj3Q,LD+,el[2,n](*)tT-[2,n](*)D+,el[2,n](*).

Hence

PLD+,el[2,n](*)St=P(LD+,el[2,n](*)t)(St)P(PLD+,el[2,n](*))Q=PLQ.

Finally, let us prove injective stability.

Theorem 1.

Let (R,Δ) be an odd form ring with a free orthogonal family of rank n, and suppose that sr(R1)n-2. Then the map

KU1(n-1;R,Δ)KU1(n;R,Δ)

is injective. If (I,Γ)(R,Δ) is an odd form ideal, then the map

KU1(n-1;R,Δ;I,Γ)KU1(n;R,Δ;I,Γ)

is also injective.

Proof.

In the absolute case, we need to prove that every

gEU(n;R,Δ)U(n-1;R,Δ)

lies in EU(n-1;R,Δ). Lemma 8 implies that g=plq for pP, lL and qQ. But then plqU(n-1;R,Δ) implies that l=1; hence we may assume that pEU(n-1;R,Δ) (the factor from T+n(*) may be pushed into q), and therefore qU(n-1;R,Δ). We have to prove that qEU(n-1;R,Δ). By definition, q=dq for dD+,el[2,n](*) and qT+[2,n](*), and it follows that q,dU(n-1;R,Δ). Write d=sts, where ts=Tisjs(xs) for 2is,jsn. Now let ts=ts if is,jsn, ts=Tis1(xsen1) if js=n, and ts=T1js(e1nxs) if is=n. It is easy to see that d=stsEU(n-1;R,Δ). Obviously, then also qEU(n-1;R,Δ).

The relative case follows from a simple diagram chase, using the absolute case for (IR,ΓΔ) and the split short exact sequence for KU1. ∎

Finally, let us show how the main theorem generalizes the linear case. Let S be a unital ring and n2 an integer. The ring R=M(n,S)op×M(n,S) has the involution (xop,y)¯=(yop,x) and the odd form parameter Δ=Δmax; hence (R,Δ) is a special unital odd form ring. This odd form ring has an orthogonal hyperbolic family of rank n, where ei are the standard matrix idempotents eiiM(n,S) for i>0 and e-i,-iopM(n,S)op for i<0. It is easy to see that U(R,Δ)GL(n,S). Under this identification, the stabilization map is the inclusion GL(n-1,S)GL(n,S), the elementary subgroup is E(n,S), and the condition sr(R1)n-2 is precisely the condition sr(S)n-2. Hence our theorem reduces to the classical result from linear K-theory. Note that here the orthogonal hyperbolic family is free, but R is not a matrix ring (because Rij=0 for ij<0), so this case is not covered by the known results from hermitian K-theory.


Communicated by Radha Kessar


Award Identifier / Grant number: 19-71-30002

Funding statement: Research is supported by the Russian Science Foundation grant 19-71-30002.

Acknowledgements

The author wants to express his gratitude to Nikolai Vavilov for helpful discussions and to Sergey Sinchuk for the idea to use the ordinary stable rank condition.

References

[1] A. Bak, Nonabelian K-theory: The nilpotent class of K1 and general stability, K-Theory 4 (1991), no. 4, 363–397. 10.1007/BF00533991Search in Google Scholar

[2] A. Bak and T. Guoping, Stability for Hermitian K1, J. Pure Appl. Algebra 150 (2000), no. 2, 109–121. 10.1016/S0022-4049(99)00035-3Search in Google Scholar

[3] A. Bak, R. Hazrat and N. Vavilov, Localization-completion strikes again: Relative K1 is nilpotent by abelian, J. Pure Appl. Algebra 213 (2009), no. 6, 1075–1085. 10.1016/j.jpaa.2008.11.014Search in Google Scholar

[4] A. Bak, V. Petrov and G. Tang, Stability for quadratic K1, K-Theory, 30 (2003), 1–11. 10.1023/B:KTHE.0000015340.00470.a9Search in Google Scholar

[5] A. Bak and N. Vavilov, Structure of hyperbolic unitary groups. I. Elementary subgroups, Algebra Colloq. 7 (2000), no. 2, 159–196. 10.1007/s10011-000-0159-1Search in Google Scholar

[6] H. Bass, K-theory and stable algebra, Inst. Hautes Études Sci. Publ. Math. (1964), no. 22, 5–60. 10.1007/BF02684689Search in Google Scholar

[7] H. Bass, Algebraic K-theory, W. A. Benjamin, New York, 1968. Search in Google Scholar

[8] R. Hazrat, Dimension theory and nonstable K1 of quadratic modules, K-Theory 27 (2002), no. 4, 293–328. 10.1023/A:1022623004336Search in Google Scholar

[9] V. A. Petrov, Odd unitary groups, J. Math. Sci. 130 (2005), no. 3, 4752–4766. 10.1007/s10958-005-0372-zSearch in Google Scholar

[10] M. Saliani, On the stability of the unitary group, unpublished. Search in Google Scholar

[11] S. Sinchuk, Injective stability for unitary K1, revisited, J. K-Theory 11 (2013), no. 2, 233–242. 10.1017/is013001028jkt211Search in Google Scholar

[12] M. R. Stein, Relativizing functors on rings and algebraic K-theory, J. Algebra 19 (1971), 140–152. 10.1016/0021-8693(71)90123-2Search in Google Scholar

[13] M. R. Stein, Stability theorems for K1, K2 and related functors modeled on Chevalley groups, Japan. J. Math. (N. S.) 4 (1978), no. 1, 77–108. 10.4099/math1924.4.77Search in Google Scholar

[14] A. A. Suslin and M. S. Tulenbaev, Stabilization theorem for Milnor K2-functor, J. Soviet Math. 17 (1981), 1804–1819. 10.1007/BF01091768Search in Google Scholar

[15] L. N. Vaserstein, Stabilization of unitary and orthogonal groups over a ring with involution, Math. USSR Sb. 10 (1970), no. 3, 307–326. 10.1070/SM1970v010n03ABEH001673Search in Google Scholar

[16] L. N. Vaserstein, Stabilization for the classical groups over rings, Math. USSR Sb. 22 (1974), no. 2, 271–303. 10.1070/SM1974v022n02ABEH001693Search in Google Scholar

[17] L. N. Vaserstein, The stable rank of pullbacks, J. Algebra 410 (2014), 239–243. 10.1016/j.jalgebra.2014.04.008Search in Google Scholar

[18] N. A. Vavilov and S. S. Sinchuk, Parabolic factorizations of split classical groups, Algebra i Analiz 23 (2011), no. 4, 1–30. 10.1090/S1061-0022-2012-01211-2Search in Google Scholar

[19] E. Y. Voronetskiĭ, Groups normalized by an odd unitary group, Algebra i Analiz 31 (2019), no. 6, 38–78. 10.1090/spmj/1630Search in Google Scholar

[20] W. Yu, Stability for odd unitary K1 under the Λ-stable range condition, J. Pure Appl. Algebra 217 (2013), no. 5, 886–891. 10.1016/j.jpaa.2012.09.003Search in Google Scholar

[21] W. Yu, Injective stability for odd-dimensional unitary K1, J. Group Theory 23 (2020), no. 2, 313–325. 10.1515/jgth-2019-0099Search in Google Scholar

[22] W. Yu and G. Tang, Nilpotency of odd unitary K1-functor, Comm. Algebra 44 (2016), no. 8, 3422–3453. 10.1080/00927872.2015.1085543Search in Google Scholar

Received: 2018-10-26
Revised: 2020-02-27
Published Online: 2020-03-20
Published in Print: 2020-09-01

© 2020 Walter de Gruyter GmbH, Berlin/Boston

Downloaded on 19.4.2024 from https://www.degruyter.com/document/doi/10.1515/jgth-2020-0013/html
Scroll to top button