This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Appendix: Adequate subgroups

Robert Guralnick Department of Mathematics, University of Southern California, Los Angeles, CA 90089-2532, USA guralnic@usc.edu Florian Herzig Institute for Advanced Study, Einstein Drive, Princeton, NJ 08540, USA herzig@math.ias.edu Richard Taylor Department of Mathematics, Harvard University, 1 Oxford Street, Cambridge, MA 02138, USA rtaylor@math.harvard.edu  and  Jack Thorne Department of Mathematics, Harvard University, 1 Oxford Street, Cambridge, MA 02138, USA thorne@math.harvard.edu
The first author was partially supported by NSF grants DMS-0653873 and DMS-1001962. The second author was partially supported by NSF grant DMS-0902044 and agreement DMS-0635607. The third author was partially supported by NSF grant DMS-0600716 and by the IAS Oswald Veblen and Simonyi Funds. The second and third authors are grateful to the IAS for its hospitality during some of the work on this appendix.

Let ll be a prime, and let Γ\Gamma be a finite subgroup of GLn(𝔽¯l)=GL(V)\operatorname{GL}_{n}(\overline{{\mathbb{F}}}_{l})=\operatorname{GL}(V). With these assumptions we say that Condition (C) holds if for every irreducible Γ\Gamma-submodule Wad0VW\subset\operatorname{ad}^{0}V there exists an element gΓg\in\Gamma with an eigenvalue α\alpha such that treg,αW0\operatorname{tr}e_{g,\alpha}W\neq 0. Here, eg,αe_{g,\alpha} denotes the projection to the generalised α\alpha-eigenspace of gg. This condition arises in the definition of adequacy in section 2.

Let Γss\Gamma^{\mathrm{ss}} denote the subset of Γ\Gamma consisting of the elements that are semisimple (i.e. of order prime to ll).

Lemma 1.

Suppose that Γ\Gamma acts irreducibly on VV. The following are equivalent.

  1. (i)

    Condition (C).

  2. (ii)

    For every irreducible submodule Wad0VW\subset\operatorname{ad}^{0}V there exists gΓssg\in\Gamma^{\mathrm{ss}} and α𝔽¯l\alpha\in\overline{{\mathbb{F}}}_{l} such that treg,αW0\operatorname{tr}e_{g,\alpha}W\neq 0.

  3. (iii)

    The set Γss\Gamma^{\mathrm{ss}} spans adV\operatorname{ad}V as an 𝔽¯l\overline{{\mathbb{F}}}_{l}-vector space.

Proof.

Note that for any gΓg\in\Gamma, Γ\Gamma contains both its semisimple and unipotent parts gsg_{s} and gug_{u}, respectively. (They are powers of gg, as we work over 𝔽¯l\overline{{\mathbb{F}}}_{l}.) Since eg,α=egs,αe_{g,\alpha}=e_{g_{s},\alpha} for all gΓg\in\Gamma, the first two conditions are equivalent.

To show that the last two conditions are equivalent, let ZadVZ\subset\operatorname{ad}V be the span of the semisimple elements in Γ\Gamma. Let UU denote the annihilator of ZZ under the (non-degenerate, Γ\Gamma-invariant) trace pairing:

(1) U\displaystyle U ={wadV:tr(gw)=0gΓss}\displaystyle=\{w\in\operatorname{ad}V:\operatorname{tr}(gw)=0\quad\forall g\in\Gamma^{\mathrm{ss}}\}
(2) ={wadV:tr(eg,αw)=0gΓss,α𝔽¯l},\displaystyle=\{w\in\operatorname{ad}V:\operatorname{tr}(e_{g,\alpha}w)=0\quad\forall g\in\Gamma^{\mathrm{ss}},\ \alpha\in\overline{{\mathbb{F}}}_{l}\},

where we used that eg,αe_{g,\alpha} is a polynomial in gg and that g=αeg,αg=\sum\alpha e_{g,\alpha} for gg semisimple.

Note that Uad0VU\subset\operatorname{ad}^{0}V by taking g=1g=1 in (1). From (2) it thus follows that the second condition is equivalent to U=0U=0. Equivalently, Z=adVZ=\operatorname{ad}V, which is the third condition. ∎

Lemma 2.
  1. (i)

    Suppose that Γ\Gamma acts irreducibly on VV. Condition (C) holds whenever Γ\Gamma has order prime to ll.

  2. (ii)

    Suppose that VV, VV^{\prime} are finite-dimensional vector spaces over 𝔽¯l\overline{{\mathbb{F}}}_{l} and that ΓGL(V)\Gamma\subset\operatorname{GL}(V), ΓGL(V)\Gamma^{\prime}\subset\operatorname{GL}(V^{\prime}) are finite subgroups that act irreducibly. If they both satisfy (C), then the image of Γ×Γ\Gamma\times\Gamma^{\prime} in GL(VV)\operatorname{GL}(V\otimes V^{\prime}) also satisfies (C).

Proof.

By Burnside’s theorem, Γ\Gamma spans adV\operatorname{ad}V. If Γ\Gamma has order prime to ll, then every element is semisimple, so the lemma above applies.

The second part of the proposition follows on noting that if gg, hh are semisimple elements then ghg\otimes h is semisimple, and appealing to the third characterization of condition (C) in the lemma above. ∎

Next we establish some preliminary results to prepare for our main theorem.

Lemma 3.

Suppose that TT is a torus over 𝔽l{\mathbb{F}}_{l}. Let X=X(T/𝔽¯l)X^{*}=X^{*}(T_{/\overline{{\mathbb{F}}}_{l}}) and X=X(T/𝔽¯l)X_{*}=X_{*}(T_{/\overline{{\mathbb{F}}}_{l}}). There is a natural action of Frobenius Fr\operatorname{Fr} as an automorphism of XX^{*} and XX_{*}. Suppose that ΔX\Delta_{*}\subset X_{*} is a finite subset that is stable under the action of Fr\operatorname{Fr} and spans XX_{*}\otimes{\mathbb{Q}}.

  1. (i)

    If μX\mu\in X^{*} with |μ,δ|<l1|\langle\mu,\delta\rangle|<l-1 for all δΔ\delta\in\Delta_{*} then μ(T(𝔽l))\mu(T({\mathbb{F}}_{l})) is trivial iff μ=0\mu=0.

  2. (ii)

    If VV is a T/𝔽¯lT_{/\overline{{\mathbb{F}}}_{l}}-module and all the weights μ\mu of T/𝔽¯lT_{/\overline{{\mathbb{F}}}_{l}} on VV satisfy |μ,δ|<(l1)/2|\langle\mu,\delta\rangle|<(l-1)/2 for all δΔ\delta\in\Delta_{*} then the 𝔽¯l\overline{{\mathbb{F}}}_{l}-span of T(𝔽l)T({\mathbb{F}}_{l}) in adV\operatorname{ad}V equals the 𝔽¯l\overline{{\mathbb{F}}}_{l}-span of T(𝔽¯l)T(\overline{{\mathbb{F}}}_{l}).

Proof.

We can identify Hom(T(𝔽l),𝔽¯l×)\operatorname{Hom}(T({\mathbb{F}}_{l}),\overline{{\mathbb{F}}}_{l}^{\times}) with X/(lFr)XX^{*}/(l-\operatorname{Fr})X^{*}. To prove the first part, suppose that |μ,δ|<l1|\langle\mu,\delta\rangle|<l-1 for δΔ\delta\in\Delta_{*} and that μ(T(𝔽l))\mu(T({\mathbb{F}}_{l})) is trivial, so μ=(lFr)λ\mu=(l-\operatorname{Fr})\lambda. Choose δ1\delta_{1} in Δ\Delta_{*} with |λ,δ1||\langle\lambda,\delta_{1}\rangle| maximal. If λ,δ10\langle\lambda,\delta_{1}\rangle\neq 0 then

l1>|μ,δ1|l|λ,δ1||λ,Fr1δ1|(l1)|λ,δ1|l1,l-1>|\langle\mu,\delta_{1}\rangle|\geq l|\langle\lambda,\delta_{1}\rangle|-|\langle\lambda,\operatorname{Fr}^{-1}\delta_{1}\rangle|\geq(l-1)|\langle\lambda,\delta_{1}\rangle|\geq l-1,

a contradiction. Therefore λ,δ1=0\langle\lambda,\delta_{1}\rangle=0, so λ=0\lambda=0 and μ=0\mu=0. In particular we see that if μ1\mu_{1} and μ2\mu_{2} are two elements of XX^{*} with |μi,δ|<(l1)/2|\langle\mu_{i},\delta\rangle|<(l-1)/2 for δΔ\delta\in\Delta_{*} and i=1i=1, 22 then μ1|T(𝔽l)=μ2|T(𝔽l)\mu_{1}|_{T({\mathbb{F}}_{l})}=\mu_{2}|_{T({\mathbb{F}}_{l})} iff μ1=μ2\mu_{1}=\mu_{2}. The second part now follows since both subspaces of adV\operatorname{ad}V equal the 𝔽¯l\overline{{\mathbb{F}}}_{l}-linear span of the T/𝔽¯lT_{/\overline{{\mathbb{F}}}_{l}}-equivariant projectors onto the weight spaces of T/𝔽¯lT_{/\overline{{\mathbb{F}}}_{l}} in VV. ∎

Lemma 4.

Suppose that GG is a connected simply connected semisimple algebraic group over 𝔽¯l\overline{{\mathbb{F}}}_{l} and ϕ:GGL(V)\phi:G\to\operatorname{GL}(V) a finite-dimensional representation. Let GBTG\supset B\supset T denote a Borel and maximal torus, and suppose that |μ1μ2,α|<l|\langle\mu_{1}-\mu_{2},\alpha^{\vee}\rangle|<l for all weights μ1\mu_{1}, μ2\mu_{2} of TT on VV and all simple roots α\alpha. Then there exist connected simply connected semisimple algebraic subgroups II and JJ of GG such that G=I×JG=I\times J, ϕ(J)=1\phi(J)=1, and ϕ\phi induces a central isogeny of II onto its image I¯\overline{I}, which is a semisimple algebraic group.

Proof.

Let JJ denote the connected component of the kernel of ϕ\phi with its reduced scheme structure. Then JJ is smooth ([Mil], Proposition I.5.18). By Theorem 8.1.5 of [Spr09] and its proof, JJ is semisimple and there is a second semisimple algebraic group IGI\subset G which commutes with JJ and such that I×JGI\times J\rightarrow G is a central isogeny. It follows from the simply-connectedness of GG that it is an isomorphism of I×JI\times J onto GG. In particular, II and JJ are simply connected. Note that T=TI×TJT=T_{I}\times T_{J} and that B=BI×BJB=B_{I}\times B_{J} where (BI,TI)(B_{I},T_{I}) (resp. (BJ,TJ)(B_{J},T_{J})) is a Borel and maximal torus in II (resp. JJ). (This follows from the fact that any smooth connected soluble subgroup of (resp. torus in) GG is conjugate to a subgroup of BB (resp. TT).) Moreover U=UI×UJU=U_{I}\times U_{J}, where UU denotes the unipotent radical of BB. Let I¯\overline{I} denote the image of II under ϕ\phi. Then I¯\overline{I} is again reduced and connected and hence also smooth. In fact it is semisimple. (See Proposition 14.10(1)(c) of [Bor91].) The map ϕ\phi factors through an isogeny II¯GL(V)I\rightarrow\overline{I}\subset\operatorname{GL}(V). Let B¯\overline{B}, T¯\overline{T}, U¯\overline{U} denote the images of BIB_{I}, TIT_{I}, UIU_{I} in I¯\overline{I}. Then these are all reduced and hence smooth. Moreover T¯\overline{T} is a torus, B¯\overline{B} is connected and soluble, U¯\overline{U} is connected unipotent and B¯=T¯U¯\overline{B}=\overline{T}\overline{U}. As dimI¯=dimI=dimTI+2dimUI=dimT¯+2dimU¯\dim\overline{I}=\dim I=\dim T_{I}+2\dim U_{I}=\dim\overline{T}+2\dim\overline{U} we see that B¯\overline{B} must be a Borel subgroup of I¯\overline{I} with unipotent radical U¯\overline{U} and that T¯\overline{T} is a maximal torus in I¯\overline{I}. The isogeny II¯I\rightarrow\overline{I} induces an ll-morphism from the root datum of I¯\overline{I} to the root datum of II. (See section 9.6.3 of [Spr09].) Then II¯I\rightarrow\overline{I} is a central isogeny, as otherwise TT would have a weight occurring in LieI¯adV\operatorname{Lie}\overline{I}\subset\operatorname{ad}V of the form lμl\mu with μ\mu non-zero and this would contradict our assumption on the weights of TT on VV. ∎

Suppose that we are given 𝔽¯l\overline{{\mathbb{F}}}_{l}-vector spaces WiW_{i} with dimWil\dim W_{i}\leq l for i=1i=1, …, rr. Then the maps

exp:X\displaystyle\exp:X 1+X+X22!++Xl1(l1)!\displaystyle\mapsto 1+X+\frac{X^{2}}{2!}+\cdots+\frac{X^{l-1}}{(l-1)!}
log:1+u\displaystyle\log:1+u uu22+u33±ul1l1\displaystyle\mapsto u-\frac{u^{2}}{2}+\frac{u^{3}}{3}\pm\cdots-\frac{u^{l-1}}{l-1}

define inverse bijections between the set of nilpotent elements in End(Wi)\prod\operatorname{End}(W_{i}) and the set of unipotent elements in GL(Wi)\prod\operatorname{GL}(W_{i}).

Lemma 5.

Suppose that GGL(Wi)G\subset\prod\operatorname{GL}(W_{i}) is a connected reductive group over 𝔽¯l\overline{{\mathbb{F}}}_{l} with dimWil\dim W_{i}\leq l for all ii. Let TT be a maximal torus and UU be the unipotent radical of a Borel subgroup of GG that contains TT. Suppose that |μ1μ2,α|<l|\langle\mu_{1}-\mu_{2},\alpha^{\vee}\rangle|<l for all weights μ1\mu_{1}, μ2\mu_{2} of TT on V=WiV=\bigoplus W_{i} and all simple roots α\alpha.

  1. (i)

    The maps exp\exp and log\log induce inverse isomorphisms of varieties between LieUEnd(V)\operatorname{Lie}U\subset\operatorname{End}(V) and UGL(V)U\subset\operatorname{GL}(V).

  2. (ii)

    For any positive root α\alpha we have exp(LieUα)=Uα\exp(\operatorname{Lie}U_{\alpha})=U_{\alpha}.

  3. (iii)

    The map exp:LieUU\exp:\operatorname{Lie}U\to U depends only on GG and UU, but not on VV, WiW_{i}, or the representation GGL(V)G\hookrightarrow\operatorname{GL}(V).

  4. (iv)

    If θ\theta is an automorphism of GG that preserves TT and UU, then we have a commutative diagram:

    LieU\textstyle{\operatorname{Lie}U\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}dθ\scriptstyle{d\theta}exp\scriptstyle{\exp}LieU\textstyle{\operatorname{Lie}U\ignorespaces\ignorespaces\ignorespaces\ignorespaces}exp\scriptstyle{\exp}U\textstyle{U\ignorespaces\ignorespaces\ignorespaces\ignorespaces}θ\scriptstyle{\theta}U\textstyle{U}
Proof.

By the Lie–Kolchin theorem we may suppose UU is contained in the group U=UiU^{\prime}=\prod U_{i}^{\prime}, where UiU^{\prime}_{i} denotes the unipotent radical of a Borel subgroup of GL(Wi)\operatorname{GL}(W_{i}). The maps exp\exp and log\log provide mutually inverse isomorphisms of varieties between UU^{\prime} and LieU\operatorname{Lie}U^{\prime}. It remains to show that expLieU=U\exp\operatorname{Lie}U=U. Note that the product of any ll elements of LieU\operatorname{Lie}U^{\prime} is zero. Thus the Zassenhaus formula (see [Mag54], section IV) tells us that to check that expLieUU\exp\operatorname{Lie}U\subset U it suffices to check that for any root α\alpha we have exp(LieUα)U\exp(\operatorname{Lie}U_{\alpha})\subset U. Let xα:𝔾aUαx_{\alpha}:{\mathbb{G}}_{a}\to U_{\alpha} be the root homomorphism corresponding to α\alpha and let Xα=dxα(1)LieUαX_{\alpha}=dx_{\alpha}(1)\in\operatorname{Lie}U_{\alpha}. Then formula II.1.19(6) of [Jan03] shows that for a𝔽¯la\in\overline{{\mathbb{F}}}_{l},

(3) xα(a)=n=0l1anXαnn!=exp(aXα)x_{\alpha}(a)=\sum_{n=0}^{l-1}a^{n}\frac{X_{\alpha}^{n}}{n!}=\exp(aX_{\alpha})

in GL(V)\operatorname{GL}(V), on noting that for n<ln<l we have Xα,n=Xαn/n!X_{\alpha,n}=X_{\alpha}^{n}/n! while Xα,nX_{\alpha,n} acts trivially on VV for nln\geq l. (This latter assertion follows from formula II.1.19(5) of [Jan03] because VλV_{\lambda} and Vλ+nαV_{\lambda+n\alpha} cannot both be non-zero.) Now by the Baker–Campbell–Hausdorff formula (see section IV.8 in part I of [Ser92]) and the fact that the product of any ll elements of LieU\operatorname{Lie}U^{\prime} is zero we see that expLieU\exp\operatorname{Lie}U is a subgroup of UU. As UU is connected and smooth and dimLieUdimU\dim\operatorname{Lie}U\geq\dim U we deduce that expLieU=U\exp\operatorname{Lie}U=U. This proves the first two parts.

The third part follows inductively from equation (3) and the Zassenhaus formula: fix a total order << on the set of positive roots such that if α\alpha, β\beta, α+β\alpha+\beta are positive roots, then max(α,β)<α+β\max(\alpha,\beta)<\alpha+\beta. We induct on the positive root γ\gamma. Suppose that we know that exp\exp depends only on GG and UU on the subspace α>γLieUα\bigoplus_{\alpha>\gamma}\operatorname{Lie}U_{\alpha}. Then the same is true for exp(X+Y)\exp(X+Y) for any XLieUγX\in\operatorname{Lie}U_{\gamma} and Yα>γLieUαY\in\bigoplus_{\alpha>\gamma}\operatorname{Lie}U_{\alpha} by the Zassenhaus formula. (Note that [LieUα,LieUβ]LieUα+β[\operatorname{Lie}U_{\alpha},\operatorname{Lie}U_{\beta}]\subset\operatorname{Lie}U_{\alpha+\beta} whenever α\alpha, β\beta are positive roots.) This completes the proof of the third part.

The last part follows from the third part, by considering the representation G𝜃GGL(V)G\xrightarrow{\theta}G\hookrightarrow\operatorname{GL}(V). ∎

Lemma 6.

Suppose that GG is a connected simply connected semisimple algebraic group over 𝔽¯l\overline{{\mathbb{F}}}_{l}. Suppose that l>3l>3 and that GG has no simple factor isomorphic to SLn\operatorname{SL}_{n} with l|nl|n. Let 𝔤{\mathfrak{g}} denote the Lie algebra of GG. Then 𝔤{\mathfrak{g}} contains no non-trivial abelian ideal, and the natural map Aut(G)Aut(𝔤)\operatorname{Aut}(G)\to\operatorname{Aut}({\mathfrak{g}}) is a bijection. Moreover, a connected normal subgroup of GG is preserved by an automorphism θAut(G)\theta\in\operatorname{Aut}(G) if and only if its Lie algebra is preserved by dθAut(𝔤)d\theta\in\operatorname{Aut}({\mathfrak{g}}).

Here, Aut(G)\operatorname{Aut}(G) (resp., Aut(𝔤)\operatorname{Aut}({\mathfrak{g}})) denotes the abstract group of automorphisms of the algebraic group GG (resp., its Lie algebra 𝔤{\mathfrak{g}}). In the proof we use Chevalley groups in the sense of Steinberg’s Yale notes [Ste68b].

Proof.

The universal Chevalley group over 𝔽¯l\overline{{\mathbb{F}}}_{l} constructed using the complex semisimple Lie algebra {\mathcal{L}} of the same root system as GG is an algebraic group isomorphic to GG (see [Ste68b], §5). (In the notation of [Ste68b], we can let VV be any representation whose weights span the weight lattice, so that {\mathcal{L}}_{\mathbb{Z}}\subset{\mathcal{L}} is the {\mathbb{Z}}-lattice spanned by the fixed Chevalley basis HiH_{i}, XαX_{\alpha}; see Cor. 2 on p. 18 of [Ste68b].) In particular, 𝔤𝔽¯l{\mathfrak{g}}\cong{\mathcal{L}}_{\mathbb{Z}}\otimes\overline{{\mathbb{F}}}_{l} (by the remark on p. 64 of [Ste68b]). Write G=GiG=\prod G_{i} as a product of almost simple simply connected algebraic groups and correspondingly 𝔤=𝔤i{\mathfrak{g}}=\bigoplus{\mathfrak{g}}_{i}. Then Z(𝔤i)=0Z({\mathfrak{g}}_{i})=0 by our assumption on ll and GG (see Theorem 2.3 in [Hur82]) and hence all 𝔤i{\mathfrak{g}}_{i} are simple ([Ste61], 2.6(5)). Moreover 𝔤i𝔤j{\mathfrak{g}}_{i}\cong{\mathfrak{g}}_{j} implies GiGjG_{i}\cong G_{j} ([Ste61], 8.1). The GiG_{i} (resp., 𝔤i{\mathfrak{g}}_{i}) are uniquely characterised as the minimal non-trivial connected normal subgroups of GG (resp., minimal non-trivial ideals of 𝔤{\mathfrak{g}}), so they are permuted by automorphisms. Therefore if Aut(Gi)Aut(𝔤i)\operatorname{Aut}(G_{i})\to\operatorname{Aut}({\mathfrak{g}}_{i}) is a bijection for all ii, then so is Aut(G)Aut(𝔤)\operatorname{Aut}(G)\to\operatorname{Aut}({\mathfrak{g}}), and also the final claim of the proposition follows. (Note that any connected normal subgroup is a product of some of the GiG_{i}.) We can thus assume, without loss of generality, that GG is almost simple.

Let GadG^{\operatorname{ad}} denote the adjoint form of GG. As GG is the universal cover of GadG^{\operatorname{ad}} and as Gad=G/Z(G)G^{\operatorname{ad}}=G/Z(G), we have Aut(G)=Aut(Gad)\operatorname{Aut}(G)=\operatorname{Aut}(G^{\operatorname{ad}}). As Z(𝔤)=0Z({\mathfrak{g}})=0 we see that the natural map 𝔤LieGad{\mathfrak{g}}\to\operatorname{Lie}G^{\operatorname{ad}} is an isomorphism. Thus it suffices to show that Aut(G)=Aut(𝔤)\operatorname{Aut}(G)=\operatorname{Aut}({\mathfrak{g}}) whenever GG is simple of adjoint type and 𝔤=LieG{\mathfrak{g}}=\operatorname{Lie}G. Thus we write GG for GadG^{\operatorname{ad}} from now on.

As an algebraic group GG is isomorphic to the adjoint Chevalley group over 𝔽¯l\overline{{\mathbb{F}}}_{l} (again by [Ste68b], §5). (In the notation of [Ste68b], we take VV to be the adjoint representation 𝔤{\mathfrak{g}}.) Thus we can identify G(𝔽¯l)G(\overline{{\mathbb{F}}}_{l}) with the subgroup of GL(𝔤)\operatorname{GL}({\mathfrak{g}}) generated by the elements xα(t):=exp(ad(tXα))x_{\alpha}(t):=\exp(\operatorname{ad}(tX_{\alpha})), where t𝔽¯lt\in\overline{{\mathbb{F}}}_{l} and α\alpha is any root. As each ad(tXα)\operatorname{ad}(tX_{\alpha}) is a derivation of 𝔤{\mathfrak{g}}, the group G(𝔽¯l)G(\overline{{\mathbb{F}}}_{l}) is actually contained in Aut(𝔤)\operatorname{Aut}({\mathfrak{g}}). For any ηAut(𝔤)\eta\in\operatorname{Aut}({\mathfrak{g}}), we have ηadXη1=ad(ηX)\eta\circ\operatorname{ad}X\circ\eta^{-1}=\operatorname{ad}(\eta X) in GL(𝔤)\operatorname{GL}({\mathfrak{g}}). It follows that the natural action of G(𝔽¯l)GL(𝔤)G(\overline{{\mathbb{F}}}_{l})\subset\operatorname{GL}({\mathfrak{g}}) on 𝔤{\mathfrak{g}} agrees with the adjoint action of G(𝔽¯l)G(\overline{{\mathbb{F}}}_{l}) on 𝔤End(𝔤){\mathfrak{g}}\subset\operatorname{End}({\mathfrak{g}}).

The choice of Chevalley basis gives rise to a maximal torus TT and a Borel BB that contains it ([Ste68b], §5). From Theorem 9.6.2 in [Spr09] we deduce the following, using that GG is adjoint. For each symmetry π\pi of the Dynkin diagram 𝒟{\mathcal{D}} there is a unique πAut(G)\pi^{\prime}\in\operatorname{Aut}(G) that preserves (B,T)(B,T) and that permutes the xαi(1)Bx_{\alpha_{i}}(1)\in B according to π\pi (where αi\alpha_{i} are the simple roots). Moreover, Aut(G)\operatorname{Aut}(G) is the semidirect product of GG (acting by inner automorphisms) and Aut(𝒟)\operatorname{Aut}({\mathcal{D}}). Also, the elements of Aut(𝒟)\operatorname{Aut}({\mathcal{D}}) biject with the “graph automorphisms” of 𝔤{\mathfrak{g}} ([Ste61], §3).

The result now follows from ([Ste61], 4.2 and 4.5), as the group \mathfrak{H} in [Ste61] is actually contained in G(𝔽¯l)G(\overline{{\mathbb{F}}}_{l}) since 𝔽¯l\overline{{\mathbb{F}}}_{l} is algebraically closed (see Lemma 19 on p. 27 of [Ste68b]). (Note that the uniqueness statement in ([Ste61], 4.2) is incorrect and seems to be a typo.) ∎

The following proposition may be of independent interest. The proof uses the classification of finite simple groups. Without it, the proof still goes through for ll sufficiently large (depending on dd and ineffective) by appealing to [LP] instead of [Gur99].

Proposition 7.

Suppose that VV is a finite-dimensional 𝔽¯l\overline{{\mathbb{F}}}_{l}-vector space and that ΓGL(V)\Gamma\subset\operatorname{GL}(V) is a finite subgroup that acts semisimply on VV. Let Γ0Γ\Gamma^{0}\subset\Gamma be the subgroup generated by elements of ll-power order. Then VV is a semisimple Γ0\Gamma^{0}-module. Let d1d\geq 1 be the maximal dimension of an irreducible Γ0\Gamma^{0}-submodule of VV. Suppose that l2(d+1)l\geq 2(d+1). Then there exists an algebraic group GG over 𝔽l{\mathbb{F}}_{l} and a semisimple representation r:G/𝔽¯lGL(V)r:G_{/\overline{{\mathbb{F}}}_{l}}\to\operatorname{GL}(V) with the following properties:

  1. (i)

    The connected component G0G^{0} is semisimple, simply connected.

  2. (ii)

    GG0HG\cong G^{0}\rtimes H, where HH is a finite group of order prime to ll.

  3. (iii)

    r(G(𝔽l))=Γr(G({\mathbb{F}}_{l}))=\Gamma.

Moreover, if TG0T\subset G^{0} is a maximal torus and if μ\mu is a weight of T/𝔽¯lT_{/\overline{{\mathbb{F}}}_{l}} on VV then |μ,α|<2d\sum|\langle\mu,\alpha^{\vee}\rangle|<2d, where α\alpha ranges over the roots of G/𝔽¯l0G^{0}_{/\overline{{\mathbb{F}}}_{l}}. Also, Γ\Gamma does not have any composition factor of order ll.

Proof.

Write V=iWiV=\bigoplus_{i}{W_{i}} as a direct sum of irreducible Γ0\Gamma^{0}-modules. Since dimWil\dim W_{i}\leq l for all ii, we see that every element of ll-power order in the image of Γ0GL(Wi)\Gamma^{0}\rightarrow\operatorname{GL}(W_{i}) actually has order dividing ll. Since Γ0GL(Wi)\Gamma^{0}\hookrightarrow\prod\operatorname{GL}(W_{i}), we deduce that every element of Γ0\Gamma^{0} of ll-power order actually has order dividing ll. Note that Γ/Γ0\Gamma/\Gamma^{0} has order prime to ll.

Step 1. We show that there exists a connected simply connected semisimple algebraic group G0G^{0} over 𝔽l{\mathbb{F}}_{l} and a finite central subgroup Z0G0(𝔽l)Z_{0}\subset G^{0}({\mathbb{F}}_{l}) with G0(𝔽l)/Z0Γ0G^{0}({\mathbb{F}}_{l})/Z_{0}\cong\Gamma^{0}. Let Γi\Gamma_{i} denote the image of Γ0\Gamma^{0} in GL(Wi)\operatorname{GL}(W_{i}). Note that Γi\Gamma_{i} has no non-trivial normal subgroup of ll-power order (since Γi\Gamma_{i} acts faithfully on WiW_{i}, and an ll-group acting on a non-zero 𝔽¯l\overline{{\mathbb{F}}}_{l}-vector space has non-zero fixed points). So by Theorem B of [Gur99], Γi\Gamma_{i} is a central product of quasisimple Chevalley groups. (Note that if l=11l=11 then dimWi<7\dim W_{i}<7.) Now Γ0\Gamma^{0} is a subgroup of Γi\prod\Gamma_{i} that surjects onto each factor, so Z(Γ0)=Γ0Z(Γi)Z(\Gamma^{0})=\Gamma^{0}\cap\prod Z(\Gamma_{i}). Thus Γ0/Z(Γ0)\Gamma^{0}/Z(\Gamma^{0}) is a subgroup of Γi/Z(Γi)\prod\Gamma_{i}/Z(\Gamma_{i}), a product of simple Chevalley groups, that surjects onto each factor. By a theorem of Hall (Lemma 3.5 in [Kup]), Γ0/Z(Γ0)\Gamma^{0}/Z(\Gamma^{0}) is itself isomorphic to a direct product of simple Chevalley groups. It follows that Γ0=[Γ0,Γ0]Z(Γ0)\Gamma^{0}=[\Gamma^{0},\Gamma^{0}]Z(\Gamma^{0}). Since Γ0\Gamma^{0} is generated by elements of order ll and Z(Γ0)Z(\Gamma^{0}) is of order prime to ll, it follows moreover that Γ0\Gamma^{0} is perfect. Therefore Γ0\Gamma^{0} is a perfect central extension of a product Hj\prod H_{j} of simple Chevalley groups HjH_{j}, so there exists a surjective homomorphism π:H~jΓ0\pi:\prod\widetilde{H}_{j}\to\Gamma^{0} with central kernel, where H~j\widetilde{H}_{j} is the universal perfect central extension of HjH_{j}.

As l>3l>3 (to rule out Suzuki and Ree groups) there exist connected simply connected algebraic groups Gj{G_{j}} over 𝔽l{\mathbb{F}}_{l} such that HjGj(𝔽l)/Z(Gj(𝔽l))H_{j}\cong G_{j}({\mathbb{F}}_{l})/Z(G_{j}({\mathbb{F}}_{l})). (Note that GjG_{j} is the restriction of scalars of an absolutely almost simple algebraic group over a finite extension of 𝔽l{\mathbb{F}}_{l}.) Since l>3l>3 it is known that H~jGj(𝔽l)\widetilde{H}_{j}\cong G_{j}({\mathbb{F}}_{l}) (see section 6.1 in [GLS98], particularly table 6.1.3). So we can take G0=GjG^{0}=\prod G_{j} and Z0=kerπZ_{0}=\ker\pi.

Since Γ0/Z(Γ0)\Gamma^{0}/Z(\Gamma^{0}) is a product of nonabelian simple groups and since Z(Γ0)Z(\Gamma^{0}) and Γ/Γ0\Gamma/\Gamma^{0} are of order prime to ll, it follows that Γ\Gamma does not have any composition factor of order ll.

Let G0BTG^{0}\supset B\supset T denote a Borel and maximal torus defined over 𝔽l{\mathbb{F}}_{l}.

Step 2. We lift VV to a G/𝔽¯l0G^{0}_{/\overline{{\mathbb{F}}}_{l}}-module and compare the actions of T(𝔽l)T({\mathbb{F}}_{l}) and T(𝔽¯l)T(\overline{{\mathbb{F}}}_{l}) on VV. Let UU denote the unipotent radical of BB and set N=NG0(T)N=N_{G^{0}}(T). Let BopB^{\mathrm{op}} denote the opposite Borel subgroup to BB containing TT and let UopU^{\mathrm{op}} denote its unipotent radical. (See Theorem 14.1 of [Bor91]. By uniqueness we see it is defined over 𝔽l{\mathbb{F}}_{l}.) Let X=X(T/𝔽¯l)X=X^{*}(T_{/\overline{{\mathbb{F}}}_{l}}) with its subset Φ\Phi of roots and Φ+\Phi^{+} (resp. Δ\Delta) the set of positive (resp. simple) roots corresponding to BB. Let X+XX^{+}\subset X be the subset of dominant weights. There is a semisimple algebraic action of G/𝔽¯l0G^{0}_{/\overline{{\mathbb{F}}}_{l}} on VV, say ϕ:G/𝔽¯l0GL(V)\phi:G^{0}_{/\overline{{\mathbb{F}}}_{l}}\rightarrow\operatorname{GL}(V), such that:

  1. (i)

    the highest weight λ\lambda of a simple submodule is restricted (i.e. 0λ,α<l0\leq\langle\lambda,\alpha^{\vee}\rangle<l for all αΔ\alpha\in\Delta),

  2. (ii)

    the action of G0(𝔽l)G^{0}({\mathbb{F}}_{l}) is the one induced by the map G0(𝔽l)Γ0G^{0}({\mathbb{F}}_{l})\rightarrow\Gamma^{0},

  3. (iii)

    the subspaces WiW_{i} are G/𝔽¯l0G^{0}_{/\overline{{\mathbb{F}}}_{l}}-stable.

(This follows from a result of Steinberg: see Theorem 2.11 in [Hum06]. Note that [Hum06] works with an algebraic group 𝐆\mathbf{G} that is simple, but the proof given does not depend on that assumption.) By Proposition 3 of [Ser94] we see that if λ\lambda in X+X^{+} is a weight of T/𝔽¯lT_{/\overline{{\mathbb{F}}}_{l}} on VV then αΦ+λ,α<d\sum_{\alpha\in\Phi^{+}}\langle\lambda,\alpha^{\vee}\rangle<d; in particular, λ,α<(l1)/2\langle\lambda,\alpha^{\vee}\rangle<(l-1)/2 for all αΦ+\alpha\in\Phi^{+}. (Note that dimWi(l1)/2\dim W_{i}\leq(l-1)/2 and that the proof of that proposition does not require that G/𝔽¯l0G^{0}_{/\overline{{\mathbb{F}}}_{l}} be almost simple.) If μ\mu is a weight of T/𝔽¯lT_{/\overline{{\mathbb{F}}}_{l}} on VV then we see that there is ww in the Weyl group with wμX+w\mu\in X^{+} and 0wμ,α<(l1)/20\leq\langle w\mu,\alpha^{\vee}\rangle<(l-1)/2 for all αΦ+\alpha\in\Phi^{+}, and we deduce that |μ,α|<(l1)/2|\langle\mu,\alpha^{\vee}\rangle|<(l-1)/2 for all αΦ\alpha\in\Phi. We also deduce that if μ\mu is a weight of T/𝔽¯lT_{/\overline{{\mathbb{F}}}_{l}} on adV\operatorname{ad}V then |μ,α|<l1|\langle\mu,\alpha^{\vee}\rangle|<l-1 for all αΔ\alpha\in\Delta.

Step 3. The semisimple group I¯GL(V)\overline{I}\subset\operatorname{GL}(V) and its simply connected cover IG/𝔽¯l0I\subset G^{0}_{/\overline{{\mathbb{F}}}_{l}}. Since |μ,α|<l/2|\langle\mu,\alpha^{\vee}\rangle|<l/2 for all weights μ\mu of T/𝔽¯lT_{/\overline{{\mathbb{F}}}_{l}} on VV and all αΔ\alpha\in\Delta we may apply Lemma 4 to ϕ:G/𝔽¯l0GL(V)\phi:G^{0}_{/\overline{{\mathbb{F}}}_{l}}\rightarrow\operatorname{GL}(V). We obtain connected simply connected semisimple algebraic subgroups II, JJ of G/𝔽¯l0G^{0}_{/\overline{{\mathbb{F}}}_{l}} such that G/𝔽¯l0=I×JG^{0}_{/\overline{{\mathbb{F}}}_{l}}=I\times J, ϕ(J)=1\phi(J)=1, and ϕ\phi induces a central isogeny of II onto its image I¯\overline{I}, which is a semisimple algebraic group. Note that T/𝔽¯l=TI×TJT_{/\overline{{\mathbb{F}}}_{l}}=T_{I}\times T_{J} and that B/𝔽¯l=BI×BJB_{/\overline{{\mathbb{F}}}_{l}}=B_{I}\times B_{J} where (BI,TI)(B_{I},T_{I}) (resp. (BJ,TJ)(B_{J},T_{J})) is a Borel and maximal torus in II (resp. JJ). Moreover U/𝔽¯l=UI×UJU_{/\overline{{\mathbb{F}}}_{l}}=U_{I}\times U_{J}. Let B¯\overline{B}, T¯\overline{T}, U¯\overline{U}, B¯op\overline{B}^{\mathrm{op}}, U¯op\overline{U}^{\mathrm{op}} denote the images of BIB_{I}, TIT_{I}, UIU_{I}, BIopB_{I}^{\mathrm{op}}, UIopU_{I}^{\mathrm{op}} in I¯\overline{I}. Then T¯\overline{T} is a maximal torus of I¯\overline{I}, and B¯\overline{B}, B¯op\overline{B}^{\mathrm{op}} are opposite Borel subgroups containing it. Also U¯\overline{U}, U¯op\overline{U}^{\mathrm{op}} are the unipotent radicals of B¯\overline{B}, B¯op\overline{B}^{\mathrm{op}}. Since II¯I\to\overline{I} is a central isogeny, UIU¯U_{I}\rightarrow\overline{U} and UIopU¯opU_{I}^{\mathrm{op}}\rightarrow\overline{U}^{\mathrm{op}} are isomorphisms.

Step 4. The maps log\log and exp\exp provide inverse isomorphisms of varieties between U¯GL(V)\overline{U}\subset\operatorname{GL}(V) and LieU¯adV\operatorname{Lie}\overline{U}\subset\operatorname{ad}V. This follows from Lemma 5 applied to I¯GL(V)\overline{I}\subset\operatorname{GL}(V) since dimWil\dim W_{i}\leq l for all ii and |μ,α|<l/2|\langle\mu,\alpha^{\vee}\rangle|<l/2 for all weights μ\mu of T/𝔽¯lT_{/\overline{{\mathbb{F}}}_{l}} on VV and all αΔ\alpha\in\Delta. (Note that TIT¯T_{I}\to\overline{T} induces a bijection on coroots since II¯I\to\overline{I} is a central isogeny; thus TT¯T\to\overline{T} induces a surjection on coroots.)

Step 5. The 𝔽¯l\overline{{\mathbb{F}}}_{l}-span of logU(𝔽l)\log U({\mathbb{F}}_{l}) is LieU¯\operatorname{Lie}\overline{U}. Since dϕ:LieULieU¯d\phi:\operatorname{Lie}U\rightarrow\operatorname{Lie}\overline{U} is surjective, it suffices to show that there is an isomorphism log:ULieU\log:U\to\operatorname{Lie}U defined over 𝔽l{\mathbb{F}}_{l} such that dϕlog=logϕd\phi\circ\log=\log\circ\phi. Pick an 𝔽l{\mathbb{F}}_{l}-structure on VV. The map G/𝔽¯l0GL(V)G^{0}_{/\overline{{\mathbb{F}}}_{l}}\to\operatorname{GL}(V) can be defined over some 𝔽ls{\mathbb{F}}_{l^{s}} and so taking restrictions of scalars from 𝔽ls{\mathbb{F}}_{l^{s}} to 𝔽l{\mathbb{F}}_{l} we get an 𝔽l{\mathbb{F}}_{l}-vector space VV^{\prime} and a map ψ:G0GL(V)\psi:G^{0}\to\operatorname{GL}(V^{\prime}). The map G/𝔽¯l0GL(V)G^{0}_{/\overline{{\mathbb{F}}}_{l}}\to\operatorname{GL}(V) is obtained from ψ\psi by extending scalars to 𝔽¯l\overline{{\mathbb{F}}}_{l} and projecting to a direct summand VV of V𝔽¯lV^{\prime}\otimes\overline{{\mathbb{F}}}_{l}. The dimension of all irreducible factors of V𝔽¯lV^{\prime}\otimes\overline{{\mathbb{F}}}_{l} is at most ll. Moreover for any weight λ\lambda of T/𝔽¯lT_{/\overline{{\mathbb{F}}}_{l}} on V𝔽¯lV^{\prime}\otimes\overline{{\mathbb{F}}}_{l} we have |λ,α|<(l1)/2|\langle\lambda,\alpha^{\vee}\rangle|<(l-1)/2 for all αΦ+\alpha\in\Phi^{+}.

By Lemma 4 we see that ψ:G0GL(V)\psi:G^{0}\to\operatorname{GL}(V^{\prime}) is a central isogeny onto its image. (By construction we have (kerψ)(𝔽l)=Z0(\ker\psi)({\mathbb{F}}_{l})=Z_{0}. Suppose that kerψ\ker\psi is not finite. Then it has to contain one of the 𝔽l{\mathbb{F}}_{l}-almost simple factors of G0=GjG^{0}=\prod G_{j}. But Gj(𝔽l)G_{j}({\mathbb{F}}_{l}) is nonabelian.)

In particular, ψ\psi induces an isomorphism Uψ(U)U\to\psi(U). Then Lemma 5 (applied to the image of ψ/𝔽¯l\psi_{/\overline{{\mathbb{F}}}_{l}}) gives the desired map log:ULieUadV\log:U\to\operatorname{Lie}U\subset\operatorname{ad}V^{\prime}.

Step 6: Some properties of G0(𝔽l)G^{0}({\mathbb{F}}_{l}). The pair (B(𝔽l),N(𝔽l))(B({{\mathbb{F}}_{l}}),N({{\mathbb{F}}_{l}})) is a split BNBN pair in G0(𝔽l)G^{0}({{\mathbb{F}}_{l}}) (see section 1.18 of [Car93]). Also U(𝔽l)U({{\mathbb{F}}_{l}}) is a Sylow ll-subgroup of G0(𝔽l)G^{0}({{\mathbb{F}}_{l}}) and B(𝔽l)=NG0(𝔽l)(U(𝔽l))=NG0(𝔽l)(B(𝔽l))B({{\mathbb{F}}_{l}})=N_{G^{0}({{\mathbb{F}}_{l}})}(U({{\mathbb{F}}_{l}}))=N_{G^{0}({{\mathbb{F}}_{l}})}(B({{\mathbb{F}}_{l}})) (see Proposition 2.5.1 of [Car93]).

Moreover T(𝔽l)T({{\mathbb{F}}_{l}}) is a Sylow ll-complement in B(𝔽l)B({{\mathbb{F}}_{l}}). Note that Uop(𝔽l)U^{\mathrm{op}}({{\mathbb{F}}_{l}}) is N(𝔽l)N({{\mathbb{F}}_{l}})-conjugate to U(𝔽l)U({{\mathbb{F}}_{l}}). (The longest Weyl element w0w_{0} is stable under Frobenius, hence represented by an element n0N(𝔽l)n_{0}\in N({\mathbb{F}}_{l}). Then use that Uop=n0Un01U^{\mathrm{op}}=n_{0}Un_{0}^{-1}.) Moreover the second-last displayed equation on page 74 (section 2.9) of [Car93] shows that Uop(𝔽l)U^{\mathrm{op}}({{\mathbb{F}}_{l}}) is the unique N(𝔽l)N({{\mathbb{F}}_{l}})-conjugate of U(𝔽l)U({{\mathbb{F}}_{l}}) with trivial intersection with U(𝔽l)U({{\mathbb{F}}_{l}}).

Step 7. We have N(𝔽l)=NG0(𝔽l)(T(𝔽l))N({{\mathbb{F}}_{l}})=N_{G^{0}({{\mathbb{F}}_{l}})}(T({{\mathbb{F}}_{l}})) so that NG0(𝔽l)(T(𝔽l))NG0(𝔽l)(B(𝔽l))=T(𝔽l)N_{G^{0}({{\mathbb{F}}_{l}})}(T({{\mathbb{F}}_{l}}))\cap N_{G^{0}({{\mathbb{F}}_{l}})}(B({{\mathbb{F}}_{l}}))=T({{\mathbb{F}}_{l}}) and Z0Z(G0(𝔽l))T(𝔽l)Z_{0}\subset Z(G^{0}({\mathbb{F}}_{l}))\subset T({\mathbb{F}}_{l}).

Suppose that gg is in NG0(𝔽l)(T(𝔽l))N_{G^{0}({{\mathbb{F}}_{l}})}(T({{\mathbb{F}}_{l}})). One can write gg uniquely as unuunu^{\prime} where uU(𝔽l),nN(𝔽l)u\in U({{\mathbb{F}}_{l}}),n\in N({{\mathbb{F}}_{l}}) maps to wnw_{n} in the Weyl group and uUwnu^{\prime}\in U_{w_{n}} in the notation of Theorem 2.5.14 of [Car93]. Then for any hh in T(𝔽l)T({{\mathbb{F}}_{l}}) we can find hh^{\prime} and h′′h^{\prime\prime} in T(𝔽l)T({{\mathbb{F}}_{l}}) such that

hunu=unuh and h′′unu=unuh,hunu^{\prime}=unu^{\prime}h^{\prime}\text{\quad and\quad}h^{\prime\prime}unu^{\prime}=unu^{\prime}h,

i.e.,

(huh1)(hn)u=u(nh)(h1uh)(huh^{-1})(hn)u^{\prime}=u(nh^{\prime})(h^{\prime-1}u^{\prime}h^{\prime})

and

(h′′uh′′1)(h′′n)u=u(nh)(h1uh).(h^{\prime\prime}uh^{\prime\prime-1})(h^{\prime\prime}n)u^{\prime}=u(nh)(h^{-1}u^{\prime}h).

As T(𝔽l)T({{\mathbb{F}}_{l}}) normalizes U(𝔽l)U({{\mathbb{F}}_{l}}) and UwnU_{w_{n}} and as wnh=wn=whnw_{nh}=w_{n}=w_{hn} the uniqueness assertion of Theorem 2.5.14 of [Car93] tells us that huh1=uhuh^{-1}=u and u=h1uhu^{\prime}=h^{-1}u^{\prime}h. Thus uZU(𝔽l)(T(𝔽l))u\in Z_{U({{\mathbb{F}}_{l}})}(T({{\mathbb{F}}_{l}})) and uZUwn(T(𝔽l))ZU(𝔽l)(T(𝔽l))u^{\prime}\in Z_{U_{w_{n}}}(T({{\mathbb{F}}_{l}}))\subset Z_{U({{\mathbb{F}}_{l}})}(T({{\mathbb{F}}_{l}})). So it suffices to prove that ZU(𝔽¯l)(T(𝔽l))=1Z_{U(\overline{{\mathbb{F}}}_{l})}(T({{\mathbb{F}}_{l}}))={1}. By Proposition 8.2.1 in [Spr09]

it suffices to show that ZUα(𝔽¯l)(T(𝔽l))=1Z_{U_{\alpha}(\overline{{\mathbb{F}}}_{l})}(T({{\mathbb{F}}_{l}}))={1} for all αΦ+\alpha\in\Phi^{+}. By Proposition 8.1.1(i) in [Spr09]

it suffices that α\alpha is non-trivial on T(𝔽l)T({{\mathbb{F}}_{l}}) for all αΦ+\alpha\in\Phi^{+}. As l5l\geq 5, this follows from Lemma 3(i) (applied with Δ\Delta_{*} the set of simple coroots).

Step 8. We find a subgroup HH of order prime to ll such that Γ=Γ0H\Gamma=\Gamma^{0}H. Let HH denote the subgroup of hΓh\in\Gamma which normalize both the image of B(𝔽l)B({{\mathbb{F}}_{l}}) and the image of T(𝔽l)T({{\mathbb{F}}_{l}}) in Γ0\Gamma^{0}. Then by the previous paragraph we see that HΓ0H\cap\Gamma^{0} is T(𝔽l)/Z0T({{\mathbb{F}}_{l}})/Z_{0}. Thus HH has order prime to ll.

Moreover if γΓ\gamma\in\Gamma we see that γ(B(𝔽l)/Z0)γ1\gamma(B({{\mathbb{F}}_{l}})/Z_{0})\gamma^{-1} is the normalizer of a Sylow ll-subgroup of G0(𝔽l)/Z0G^{0}({{\mathbb{F}}_{l}})/Z_{0} and hence G0(𝔽l)G^{0}({{\mathbb{F}}_{l}})-conjugate to B(𝔽l)/Z0B({{\mathbb{F}}_{l}})/Z_{0}, say γ(B(𝔽l)/Z0)γ1=k(B(𝔽l)/Z0)k1\gamma(B({{\mathbb{F}}_{l}})/Z_{0})\gamma^{-1}=k(B({{\mathbb{F}}_{l}})/Z_{0})k^{-1} with kG0(𝔽l)k\in G^{0}({{\mathbb{F}}_{l}}). Then k1γ(T(𝔽l)/Z0)γ1kk^{-1}\gamma(T({{\mathbb{F}}_{l}})/Z_{0})\gamma^{-1}k is a Sylow ll-complement in B(𝔽l)/Z0B({{\mathbb{F}}_{l}})/Z_{0} and hence (by Hall’s theorem) B(𝔽l)/Z0B({{\mathbb{F}}_{l}})/Z_{0}-conjugate to T(𝔽l)/Z0T({{\mathbb{F}}_{l}})/Z_{0}, say

k1γ(T(𝔽l)/Z0)γ1k=k(T(𝔽l)/Z0)k1k^{-1}\gamma(T({{\mathbb{F}}_{l}})/Z_{0})\gamma^{-1}k=k^{\prime}(T({{\mathbb{F}}_{l}})/Z_{0})k^{\prime-1}

for some kB(𝔽l)k^{\prime}\in B({{\mathbb{F}}_{l}}). Then (kk)1γ(kk^{\prime})^{-1}\gamma lies in HH and we deduce that Γ\Gamma is generated by HH and G0(𝔽l)/Z0=Γ0G^{0}({{\mathbb{F}}_{l}})/Z_{0}=\Gamma^{0}.

Step 9. Lifting the conjugation action of HH on Γ0\Gamma^{0} to G0G^{0}. We first show that G/𝔽¯l0G^{0}_{/\overline{{\mathbb{F}}}_{l}} has no simple factor SLn\operatorname{SL}_{n} with l|nl|n by showing that any such factor would act trivially on V=WiV=\bigoplus W_{i}, contradicting that G0(𝔽l)/Z0G^{0}({\mathbb{F}}_{l})/Z_{0} acts faithfully. So suppose that SLn/𝔽¯l{\operatorname{SL}_{n}}_{/\overline{{\mathbb{F}}}_{l}} has an irreducible module of dimension less than l1l-1. Then by Proposition 3 in [Ser94] its highest weight λ\lambda would satisfy λ,α<l1\sum\langle\lambda,\alpha^{\vee}\rangle<l-1, where α\alpha runs through the set of positive roots. A calculation shows that the left-hand side is at least n1n-1 if λ\lambda is non-zero. So if nln\geq l, then λ=0\lambda=0.

Next we claim that dϕ:(LieG0)(𝔽¯l)adVd\phi:(\operatorname{Lie}G^{0})(\overline{{\mathbb{F}}}_{l})\to\operatorname{ad}V is injective on the subspace (LieG0)(𝔽l)(\operatorname{Lie}G^{0})({\mathbb{F}}_{l}). Note first that it is injective on (LieU)(𝔽l)(\operatorname{Lie}U)({\mathbb{F}}_{l}) as ϕ\phi is injective on U(𝔽l)U({\mathbb{F}}_{l}). (Consider the isomorphism log:U(𝔽l)(LieU)(𝔽l)\log:U({\mathbb{F}}_{l})\to(\operatorname{Lie}U)({\mathbb{F}}_{l}) constructed in Step 5.) Similarly dϕd\phi is injective on (LieUop)(𝔽l)(\operatorname{Lie}U^{\mathrm{op}})({\mathbb{F}}_{l}). Since ϕ\phi maps UU to U¯\overline{U}, TT to T¯\overline{T}, UopU^{\mathrm{op}} to U¯op\overline{U}^{\mathrm{op}}, and since LieG0=LieULieTLieUop\operatorname{Lie}G^{0}=\operatorname{Lie}U\oplus\operatorname{Lie}T\oplus\operatorname{Lie}U^{\mathrm{op}}, LieI¯=LieU¯LieT¯LieU¯op\operatorname{Lie}\overline{I}=\operatorname{Lie}\overline{U}\oplus\operatorname{Lie}\overline{T}\oplus\operatorname{Lie}\overline{U}^{\mathrm{op}} it follows that the kernel of dϕd\phi on (LieG0)(𝔽l)(\operatorname{Lie}G^{0})({\mathbb{F}}_{l}) is contained in (LieT)(𝔽l)(\operatorname{Lie}T)({\mathbb{F}}_{l}). But (LieG0)(𝔽¯l)(\operatorname{Lie}G^{0})(\overline{{\mathbb{F}}}_{l}) contains no non-trivial abelian ideal by Lemma 6. This proves the claim.

Note that HH acts by conjugation on GL(V)\operatorname{GL}(V) and adV\operatorname{ad}V, in particular it preserves the Lie algebra structure of adV\operatorname{ad}V. By definition HH stabilises the image of U(𝔽l)U({\mathbb{F}}_{l}) in GL(V)\operatorname{GL}(V) and hence by Step 5 it also stabilises logU(𝔽l)=dϕ((LieU)(𝔽l))\log U({\mathbb{F}}_{l})=d\phi((\operatorname{Lie}U)({\mathbb{F}}_{l})). Because Uop(𝔽l)U^{\mathrm{op}}({{\mathbb{F}}_{l}}) is the unique NG0(𝔽l)(T(𝔽l))N_{G^{0}({{\mathbb{F}}_{l}})}(T({{\mathbb{F}}_{l}}))-conjugate of U(𝔽l)U({{\mathbb{F}}_{l}}) that has trivial intersection with U(𝔽l)U({{\mathbb{F}}_{l}}), it is also stabilised by HH. The previous argument then shows that HH stabilises dϕ((LieUop)(𝔽l))d\phi((\operatorname{Lie}U^{\mathrm{op}})({\mathbb{F}}_{l})). Since [LieU,LieUop]=LieG0[\operatorname{Lie}U,\operatorname{Lie}U^{\mathrm{op}}]=\operatorname{Lie}G^{0} (as we may check over 𝔽¯l\overline{{\mathbb{F}}}_{l}), it follows that HH stabilises the image of (LieG0)(𝔽l)(\operatorname{Lie}G^{0})({\mathbb{F}}_{l}) in adV\operatorname{ad}V. By extending scalars, we get a natural action of HH on (LieG0)(𝔽¯l)(\operatorname{Lie}G^{0})(\overline{{\mathbb{F}}}_{l}). This action lifts uniquely to an action on G/𝔽¯l0G^{0}_{/\overline{{\mathbb{F}}}_{l}} by Lemma 6.

We claim that with respect to the HH-action on G/𝔽¯l0G^{0}_{/\overline{{\mathbb{F}}}_{l}} just constructed, ϕ:G/𝔽¯l0GL(V)\phi:G^{0}_{/\overline{{\mathbb{F}}}_{l}}\to\operatorname{GL}(V) is HH-equivariant. We first show that the conjugation action of HH on GL(V)\operatorname{GL}(V) stabilises I¯\overline{I}. If hHh\in H then hh sends U(𝔽l)U({{\mathbb{F}}_{l}}) to itself and hence logU(𝔽l)\log U({{\mathbb{F}}_{l}}) to itself and hence LieU¯\operatorname{Lie}\overline{U} to itself and hence U¯\overline{U} to itself. Similarly hh stabilises U¯op\overline{U}^{\mathrm{op}}. As the root subgroups generate I¯\overline{I} (by Theorem 8.1.5 in [Spr09]), we see that hh indeed stabilises I¯\overline{I}. This action of HH on I¯\overline{I} lifts uniquely to an action on the simply connected cover II of I¯\overline{I}. (For existence use Theorem 9.6.5 of [Spr09] and the conjugation action of TIT_{I}. For uniqueness use the semisimplicity of II.) On the other hand, Lemma 6 shows that the HH-action on G/𝔽¯l0G^{0}_{/\overline{{\mathbb{F}}}_{l}} respects the decomposition G/𝔽¯l0=I×JG^{0}_{/\overline{{\mathbb{F}}}_{l}}=I\times J. Since JJ is killed by ϕ\phi it suffices to show that the two HH-actions on II (one coming from I¯\overline{I} and one from G/𝔽¯l0G^{0}_{/\overline{{\mathbb{F}}}_{l}}) agree. By Lemma 6 we can check this on the Lie algebra. The same lemma shows that dϕ:LieILieI¯d\phi:\operatorname{Lie}I\to\operatorname{Lie}\overline{I} is an isomorphism, since LieI\operatorname{Lie}I contains no non-trivial abelian ideal. By construction both HH-actions on LieI\operatorname{Lie}I are compatible with the HH-action on LieI¯\operatorname{Lie}\overline{I}, so the two HH-actions on II indeed agree. Therefore ϕ\phi is HH-equivariant. A fortiori, it extends to a homomorphism G/𝔽¯l0HGL(V)G^{0}_{/\overline{{\mathbb{F}}}_{l}}\rtimes H\to\operatorname{GL}(V).

Finally we show that the HH-action on G/𝔽¯l0G^{0}_{/\overline{{\mathbb{F}}}_{l}} descends to G0G^{0}. Suppose that hHh\in H and σGal(𝔽¯l/𝔽l)\sigma\in\operatorname{Gal}(\overline{{\mathbb{F}}}_{l}/{\mathbb{F}}_{l}). The automorphism σhσ1h1\sigma h\sigma^{-1}h^{-1} is trivial on (LieG0)(𝔽l)(\operatorname{Lie}G^{0})({\mathbb{F}}_{l}), hence trivial on (LieG0)(𝔽¯l)(\operatorname{Lie}G^{0})(\overline{{\mathbb{F}}}_{l}), hence trivial on G/𝔽¯l0G^{0}_{/\overline{{\mathbb{F}}}_{l}} by Lemma 6. Therefore the HH-action indeed descends to G0G^{0}.

By construction, the image of G0(𝔽l)HG^{0}({\mathbb{F}}_{l})\rtimes H is Γ\Gamma. Let G=G0HG=G^{0}\rtimes H and r:G/𝔽¯lGL(V)r:G_{/\overline{{\mathbb{F}}}_{l}}\to\operatorname{GL}(V) the homomorphism we just obtained. It remains to show that rr is semisimple. But this follows from Lemma 5(b) in [Ser94] since the restriction of rr to G/𝔽¯l0G^{0}_{/\overline{{\mathbb{F}}}_{l}} is semisimple and (G:G0)(G:G^{0}) is prime to ll. ∎

We remark that for the purpose of proving Theorem 9 we do not need an HH-action on G0G^{0}, we only need an HH-action on G/𝔽¯l0G^{0}_{/\overline{{\mathbb{F}}}_{l}} that is compatible with the HH-action on GL(V)\operatorname{GL}(V). Since G/𝔽¯l0=I×JG^{0}_{/\overline{{\mathbb{F}}}_{l}}=I\times J, we can lift the HH-action on I¯\overline{I} to II as above and let HH act arbitrarily on JJ; for this it is not necessary to appeal to Lemma 6.

Lemma 8.

Suppose that GG is a linear algebraic group over 𝔽¯l\overline{{\mathbb{F}}}_{l} such that the connected component G0G^{0} is semi-simple and simply connected and such that ll does not divide (G:G0)(G:G^{0}). Let G0BTG^{0}\supset B\supset T denote a Borel subgroup and a maximal torus and let 𝒯{\mathcal{T}} denote the normalizer of the pair (B,T)(B,T) in GG. Then the G0(𝔽¯l)G^{0}(\overline{{\mathbb{F}}}_{l})-conjugates of 𝒯(𝔽¯l){\mathcal{T}}(\overline{{\mathbb{F}}}_{l}) equal the semisimple elements of G(𝔽¯l)G(\overline{{\mathbb{F}}}_{l}) and they are Zariski dense in GG. In particular, if VV is an irreducible representation of GG then the G0(𝔽¯l)G^{0}(\overline{{\mathbb{F}}}_{l})-conjugates of 𝒯(𝔽¯l){\mathcal{T}}(\overline{{\mathbb{F}}}_{l}) span adV\operatorname{ad}V over 𝔽¯l\overline{{\mathbb{F}}}_{l}.

Proof.

By Theorem 7.5 in [Ste68a] every semisimple element of G(𝔽¯l)G(\overline{{\mathbb{F}}}_{l}) is G0(𝔽¯l)G^{0}(\overline{{\mathbb{F}}}_{l})-conjugate to an element of 𝒯(𝔽¯l){\mathcal{T}}(\overline{{\mathbb{F}}}_{l}). The converse is clear as 𝒯G0=T{\mathcal{T}}\cap G^{0}=T, an element gG(𝔽¯l)g\in G(\overline{{\mathbb{F}}}_{l}) is semisimple iff gg is of order prime to ll, and ll does not divide (G:G0)(G:G^{0}). Next we have G=G0𝒯G=G^{0}{\mathcal{T}} since Borel subgroups in G0G^{0} are conjugate and maximal tori in BB are conjugate. Consider a fixed coset G0hG^{0}h with h𝒯(𝔽¯l)h\in{\mathcal{T}}(\overline{{\mathbb{F}}}_{l}). By Lemma 4 of [Spr06] the elements g(th)g1=[gt(hgh1)1]hg(th)g^{-1}=[gt(hgh^{-1})^{-1}]h of G0hG^{0}h, where tt runs over T(𝔽¯l)T(\overline{{\mathbb{F}}}_{l}) and gg runs over G0(𝔽¯l)G^{0}(\overline{{\mathbb{F}}}_{l}), are Zariski dense in G0hG^{0}h. (Lemma 4 of [Spr06] does not immediately apply to hh as hh is not a diagram automorphism. However for some sT(𝔽¯l)s\in T(\overline{{\mathbb{F}}}_{l}) the automorphism gshgh1s1g\mapsto shgh^{-1}s^{-1} is a diagram automorphism and hence the elements gt(hgh1)1=gts1(shgh1s1)1sgt(hgh^{-1})^{-1}=gts^{-1}(shgh^{-1}s^{-1})^{-1}s as tt runs over T(𝔽¯l)T(\overline{{\mathbb{F}}}_{l}) and gg runs over G0(𝔽¯l)G^{0}(\overline{{\mathbb{F}}}_{l}) are Zariski dense in G0G^{0}.) Thus the G0(𝔽¯l)G^{0}(\overline{{\mathbb{F}}}_{l})-conjugates of 𝒯(𝔽¯l){\mathcal{T}}(\overline{{\mathbb{F}}}_{l}) are Zariski dense in G(𝔽¯l)G(\overline{{\mathbb{F}}}_{l}). For the last claim note that if tr(gw)=0\operatorname{tr}(gw)=0 for some wadVw\in\operatorname{ad}V and some Zariski dense subset of gG(𝔽¯l)g\in G(\overline{{\mathbb{F}}}_{l}), then w=0w=0. ∎

The proof of our main theorem relies on Proposition 7 and thus on the classification of finite simple groups. (It still holds without it for ll sufficiently large, depending on dd and ineffective, due to the results of Larsen and Pink [LP].)

Theorem 9.

Suppose that VV is a finite-dimensional 𝔽¯l\overline{{\mathbb{F}}}_{l}-vector space and that ΓGL(V)\Gamma\subset\operatorname{GL}(V) is a finite subgroup that acts irreducibly on VV. Let Γ0Γ\Gamma^{0}\subset\Gamma be the subgroup generated by elements of ll-power order. Then VV is a semisimple Γ0\Gamma^{0}-module. Let d1d\geq 1 be the maximal dimension of an irreducible Γ0\Gamma^{0}-submodule of VV. Suppose that l2(d+1)l\geq 2(d+1). Then:

  1. (i)

    H0(Γ,ad0V)=H1(Γ,ad0V)=H1(Γ,𝔽¯l)=0.H^{0}(\Gamma,\operatorname{ad}^{0}V)=H^{1}(\Gamma,\operatorname{ad}^{0}V)=H^{1}(\Gamma,\overline{{\mathbb{F}}}_{l})=0.

  2. (ii)

    The set Γss\Gamma^{\mathrm{ss}} spans adV\operatorname{ad}V as an 𝔽¯l\overline{{\mathbb{F}}}_{l}-vector space.

In particular, for any finite subfield kk of 𝔽¯l\overline{{\mathbb{F}}}_{l} containing the eigenvalues of all elements of Γ\Gamma and such that ΓGLn(k)\Gamma\subset\operatorname{GL}_{n}(k), Γ\Gamma is adequate.

Proof.

Write V=iWiV=\bigoplus_{i}{W_{i}} as a direct sum of irreducible Γ0\Gamma^{0}-modules. Note that Γ/Γ0\Gamma/\Gamma^{0} has order prime to ll.

We claim that dimV\dim V is prime to ll. Let UU be an irreducible constituent of VV as a Γ0\Gamma^{0}-module and let VV^{\prime} be the UU-isotypic direct summand of VV. Since Γ\Gamma acts transitively on the set of isotypic components and as (Γ:Γ0)(\Gamma:\Gamma^{0}) is prime to ll, it suffices to show that dimV\dim V^{\prime} is prime to ll. Let ΓΓ0\Gamma^{\prime}\supset\Gamma^{0} be the stabiliser of VV^{\prime}. Then VV^{\prime} is an irreducible Γ\Gamma^{\prime}-module. By Theorem 51.7 in [CR62], UU extends to a projective representation of Γ\Gamma^{\prime} and there is an irreducible projective representation UU^{\prime} of Γ/Γ0\Gamma^{\prime}/\Gamma^{0} such that VUUV^{\prime}\cong U\otimes U^{\prime} (as projective Γ\Gamma^{\prime}-representation). The claim follows as dimU<l\dim U<l and Γ/Γ0\Gamma^{\prime}/\Gamma^{0} is of order prime to ll.

By Proposition 7 there exists an algebraic group G=G0HG=G^{0}\rtimes H over 𝔽l{\mathbb{F}}_{l} and a semisimple representation r:G/𝔽¯lGL(V)r:G_{/\overline{{\mathbb{F}}}_{l}}\to\operatorname{GL}(V), where G0G^{0} is connected simply connected semisimple, HH is a finite group of order prime to ll, and r(G(𝔽l))=Γr(G({\mathbb{F}}_{l}))=\Gamma. Moreover Γ\Gamma has no composition factor of order ll, which implies that no quotient of Γ0\Gamma^{0} contains a non-trivial normal ll-subgroup.

We have

H1(Γ,adV)=i,jH1(Γ0,Hom(Wi,Wj))ΓH^{1}(\Gamma,\operatorname{ad}V)=\bigoplus_{i,j}H^{1}(\Gamma^{0},\operatorname{Hom}(W_{i},W_{j}))^{\Gamma}

and

H1(Γ0,Hom(Wi,Wj))=ExtΓ01(Wi,Wj),H^{1}(\Gamma^{0},\operatorname{Hom}(W_{i},W_{j}))=\operatorname{Ext}^{1}_{\Gamma^{0}}(W_{i},W_{j}),

which vanishes by [Gur99], Theorem A, since dimWi+dimWjl2\dim W_{i}+\dim W_{j}\leq l-2. (We apply that theorem to the quotient of Γ0\Gamma^{0} that acts faithfully. Note that we saw above that this quotient does not have a non-trivial normal ll-subgroup.) Similarly, 2l22\leq l-2 implies that H1(Γ,𝔽¯l)=0H^{1}(\Gamma,\overline{{\mathbb{F}}}_{l})=0. Since dimV\dim V is prime to ll it follows that H0(Γ,ad0V)=0H^{0}(\Gamma,\operatorname{ad}^{0}V)=0 and that ad0V\operatorname{ad}^{0}V is a direct summand of adV\operatorname{ad}V, so H1(Γ,ad0V)=0H^{1}(\Gamma,\operatorname{ad}^{0}V)=0. This proves the first part above.

Let G0BTG^{0}\supset B\supset T denote a Borel and maximal torus defined over 𝔽l{\mathbb{F}}_{l}. Proposition 7 also shows that |μ,α|<(l1)/2|\langle\mu,\alpha^{\vee}\rangle|<(l-1)/2 for all weights μ\mu of T/𝔽¯lT_{/\overline{{\mathbb{F}}}_{l}} on VV and all αΔ\alpha\in\Delta. In particular, all dominant weights of T/𝔽¯lT_{/\overline{{\mathbb{F}}}_{l}} on VV and adV\operatorname{ad}V are restricted. Note that if WW is a semisimple G/𝔽¯l0G^{0}_{/\overline{{\mathbb{F}}}_{l}}-module such that all dominant weights of T/𝔽¯lT_{/\overline{{\mathbb{F}}}_{l}} on WW are restricted, then every G0(𝔽l)G^{0}({{\mathbb{F}}_{l}})-submodule of WW is also a G/𝔽¯l0G^{0}_{/\overline{{\mathbb{F}}}_{l}}-submodule. We apply this first to VV (which is semisimple as G/𝔽¯l0G^{0}_{/\overline{{\mathbb{F}}}_{l}}-module, since rr is semisimple), so the WiW_{i} are G/𝔽¯l0G^{0}_{/\overline{{\mathbb{F}}}_{l}}-submodules. By Proposition 8 of [Ser94] we see that adV=i,jHom(Wi,Wj)\operatorname{ad}V=\bigoplus_{i,j}\operatorname{Hom}(W_{i},W_{j}) is a semisimple G/𝔽¯l0G^{0}_{/\overline{{\mathbb{F}}}_{l}}-module. (Note that dimWi+dimWj<l+2\dim W_{i}+\dim W_{j}<l+2.) Thus every G0(𝔽l)G^{0}({{\mathbb{F}}_{l}})-submodule of adV\operatorname{ad}V is also a G/𝔽¯l0G^{0}_{/\overline{{\mathbb{F}}}_{l}}-submodule.

By Lemma 3 (applied with Δ\Delta_{*} the set of simple coroots), the 𝔽¯l\overline{{\mathbb{F}}}_{l}-linear span of the image of T(𝔽l)T({\mathbb{F}}_{l}) in adV\operatorname{ad}V equals the 𝔽¯l\overline{{\mathbb{F}}}_{l}-linear span of the image of T(𝔽¯l)T(\overline{{\mathbb{F}}}_{l}). Thus the G0(𝔽l)G^{0}({{\mathbb{F}}_{l}})-submodule of adV\operatorname{ad}V generated by the 𝔽¯l\overline{{\mathbb{F}}}_{l}-linear span of r(H)r(H) equals the G0(𝔽¯l)G^{0}(\overline{{\mathbb{F}}}_{l})-submodule generated by r(T(𝔽¯l)H)r(T(\overline{{\mathbb{F}}}_{l})H). By Lemma 8 (noting that 𝒯(𝔽¯l)=T(𝔽¯l)H{\mathcal{T}}(\overline{{\mathbb{F}}}_{l})=T(\overline{{\mathbb{F}}}_{l})H) it follows that r(H)r(H) spans adV\operatorname{ad}V. As r(H)Γssr(H)\subset\Gamma^{\mathrm{ss}}, this completes the proof. ∎

References

  • [Bor91] Armand Borel. Linear algebraic groups, volume 126 of Graduate Texts in Mathematics. Springer-Verlag, New York, second edition, 1991.
  • [Car93] Roger W. Carter. Finite groups of Lie type. Wiley Classics Library. John Wiley & Sons Ltd., Chichester, 1993. Conjugacy classes and complex characters, Reprint of the 1985 original, A Wiley-Interscience Publication.
  • [CR62] Charles W. Curtis and Irving Reiner. Representation theory of finite groups and associative algebras. Pure and Applied Mathematics, Vol. XI. Interscience Publishers, a division of John Wiley & Sons, New York-London, 1962.
  • [GLS98] Daniel Gorenstein, Richard Lyons, and Ronald Solomon. The classification of the finite simple groups. Number 3. Part I. Chapter A, volume 40 of Mathematical Surveys and Monographs. American Mathematical Society, Providence, RI, 1998. Almost simple KK-groups.
  • [Gur99] Robert M. Guralnick. Small representations are completely reducible. J. Algebra, 220(2):531–541, 1999.
  • [Hum06] James E. Humphreys. Modular representations of finite groups of Lie type, volume 326 of London Mathematical Society Lecture Note Series. Cambridge University Press, Cambridge, 2006.
  • [Hur82] James F. Hurley. A note on the centers of Lie algebras of classical type. In Lie algebras and related topics (New Brunswick, N.J., 1981), volume 933 of Lecture Notes in Math., pages 111–116. Springer, Berlin, 1982.
  • [Jan03] Jens Carsten Jantzen. Representations of algebraic groups, volume 107 of Mathematical Surveys and Monographs. American Mathematical Society, Providence, RI, second edition, 2003.
  • [Kup] G. Kuperberg. Denseness and Zariski denseness of Jones braid representations. Preprint.
  • [LP] Michael Larsen and Richard Pink. Finite subgroups of algebraic groups. Preprint.
  • [Mag54] Wilhelm Magnus. On the exponential solution of differential equations for a linear operator. Comm. Pure Appl. Math., 7:649–673, 1954.
  • [Mil] James S. Milne. Algebraic groups, Lie groups, and their arithmetic subgroups. Accessible from www.jmilne.org.
  • [Ser92] Jean-Pierre Serre. Lie algebras and Lie groups, volume 1500 of Lecture Notes in Mathematics. Springer-Verlag, Berlin, second edition, 1992. 1964 lectures given at Harvard University.
  • [Ser94] Jean-Pierre Serre. Sur la semi-simplicité des produits tensoriels de représentations de groupes. Invent. Math., 116(1-3):513–530, 1994.
  • [Spr06] T. A. Springer. Twisted conjugacy in simply connected groups. Transform. Groups, 11(3):539–545, 2006.
  • [Spr09] T. A. Springer. Linear algebraic groups. Modern Birkhäuser Classics. Birkhäuser Boston Inc., Boston, MA, second edition, 2009.
  • [Ste61] Robert Steinberg. Automorphisms of classical Lie algebras. Pacific J. Math., 11:1119–1129, 1961.
  • [Ste68a] Robert Steinberg. Endomorphisms of linear algebraic groups. Memoirs of the American Mathematical Society, No. 80. American Mathematical Society, Providence, R.I., 1968.
  • [Ste68b] Robert Steinberg. Lectures on Chevalley groups. Yale University, New Haven, Conn., 1968. Notes prepared by John Faulkner and Robert Wilson.