<<

arXiv:1703.09577v4 [math.DG] 13 Jul 2020 cur emti ao yipsn itnergdt ewe leave [6]). between Ghys rigidity distance [25, called imposing Thurston by is and flavor [9] geometric Haefliger a e.g. acquire (see subjects topological deep olwn ojcue omnycle GoesCnetr”(e a sup (see is Conjecture” question “Grove’s [18]): this totally Tapp called to nat with answer commonly is affirmative conjecture, The it Riemannian following only groups. [20], such the Ranjan on are following leaves foliations Therefore, coset foliations. whether coset [8, are Grove leaves by asked was metrics submers bi-invariant Riemannian with of groups classification Lie The compact [7]). Gromoll–Walschap or [19] flsaecalled are ifolds ml faReana oito:asubmersion a : Riemannian a of ample restriction aia umnflso nivltv subbundle involutive an of submanifolds maximal eaiescinlcurvature. sectional negative ubr2017/10892-7. number cosets, ol s t leri tutr ocntutapten n i s gr Lie any Lie a pattern: with a construct starting to instance, t structure For how algebraic fabric? [23]: its Thurston use stripped following could of Or, out pattern? manifold geometric a a with space ric nue eopsto of decomposition a induces ned nsc rusalkonReana oitoswt oal geo totally with foliations Riemannian known all groups such in Indeed, h eopsto notefieso imninsbeso sami e main a is submersion Riemannian a of fibers the into decomposition The ngnrl a general, In hswr a atal upre yCP rn ubr40426 number grant CNPq by supported partially was work This phrases. and words Key 1 2010 h rsn oki eiae otesml usin o ofilagive a fill to how question: simple the to dedicated is work present The nynnsnua oitoswt once evsaecon are leaves connected with foliations non-singular Only OAL EDSCREANA OITOSON FOLIATIONS RIEMANNIAN GEODESIC TOTALLY F ahmtc ujc Classification. Subject na pnsbe of submersion subset the open when an even geodesi in affirmatively, totally coset question with a this Riemannian is answer is metric submersion bi-invariant the with provided Lie simple compact Abstract. H Riemannian − dπ = p { | Hg (ker foliation leaves dπ n8 ajnqetoe hte submersion a whether questioned Ranjan 86 In | h p ) xsec,osrcin n lsictoso oitosare foliations of classifications and obstructions Existence, . ∈ ⊥ fislae r oal qiitn seeg oio[7 or [17] Molino e.g. (see equidistant locally are leaves its if i rus imninflain,smercsae,holo spaces, symmetric foliations, Riemannian groups, Lie G sa smtyto isometry an is 1 OPC I GROUPS LIE COMPACT G } F uhdcmoiin r aldas called are decompositions Such . G asmn utbecmateshypothesis). compactness suitable (assuming LHN .SPERANC¸ D. A LLOHANN on ybt ih cosets, right both by 1. M Introduction stedcmoiinof decomposition the is S 33,MC5C0adMC53C12. MSC and 53C20 MSC 53C35, MSC 1 T π ( p ) B π o every for V : M F = iee nti paper. this in sidered H + T → = ⊆ F /069adFPS grant FAPESP and 6/2016-9 B { p π br.Hr we Here fibers. c gH M : oito rnot, or foliation ∈ is sdfie only defined is G M T M oe foliations coset noteintegrable the into Riemannian → | 4 3)adthey and 23]) 24, h ubgroup seeg O’Neill e.g. (see B uhsubman- Such . s Munteanu– lso ∈ rma from otdb the by ported rbe 5.4]. Problem :afoliation a s: G construct o rlt ask to ural oup } geomet- n oy non- nomy, osfrom ions n left and , geodesic G < H G fthe if desic we , . x- 2 LLOHANN D. SPERANC¸A

Conjecture 1. Let G be a compact simple Lie group with a bi-invariant metric. A Riemannian submersion π : G → B with connected totally geodesic fibers is induced either by left or right cosets. Here Ranjan’s question together with Conjecture 1 are proved affirmatively, with- out the simplicity assumption. Theorem 1.1. Let π : G → B be a Riemannian submersion with totally geodesic connected fibers on G, a compact connected Lie group with bi-invariant metric. Then π is isometric to a coset foliation. Actually, our proof has only one non-local instance, which can be circumvented by suitable compactness hypothesis. In particular, the proof works well for foliations on compact groups and reduce the general problem to foliations whose leaves are totally geodesic flats (as the foliation defined by the fibers of a vector bundle): Theorem 1.2. Let F be a Riemannian foliation with connected totally geodesic leaves on a connected open subset U of a compact Lie group with bi-invariant met- ric G. Then V = T F splits as V = ∆0 ⊕ ∆1, where ∆0 defines a totally geodesic Riemannian foliation by Euclidean spaces and ∆1 is isometric to a coset foliation. Moreover, F is a coset foliation if it satisfies one of the following additional hy- pothesis:

(a) L˜e, the universal cover of a fixed leaf, has no Euclidean factors; (b) Le is complete and the integrability tensor of F is bounded along Le; (c) the closure of Le is a compact subset of G. The proof is essentially algebraic, with two main geometrical instances: Theo- rems 1.3 and 1.4. Both are interesting on their own: Theorem 1.3 is a refinement of the celebrated Ambrose–Singer Theorem for foliations with totally geodesic fibers on spaces with non-negative sectional curvature; Theorem 1.4 could be readily used in an attempt to generalize Theorem 1.1 to symmetric spaces. The author is tempted to believe that the foliation defined by ∆0 is (locally) of a metric product G = G′ × Rk, therefore it is a coset foliation. The thesis of Theorem 1.1 follows once we show that g decomposes in ideals g = g+ ⊕ g− such that, for every vectors X, Y orthogonal to leaves, v v (1) [X, Y ] = [X−, Y−] − [X+, Y+] ,   v where denotes the orthogonal projection to V = T F and X±, Y± are the g±- components of X, Y . It follows from [18, Corollary 4.2] that π is given by cosets (see section 1.1 for details). The decomposition g+ ⊕ g− is obtained by refining [18, Theorem 1.5] and ideas in [20]. We observe that the hypothesis on Theorem 1.1 can not be relaxed: Kerin– Shankar [12] presented infinite families of Riemannian submersions from compact Lie groups with bi-invariant metrics that can not be realized as principal bundles (for instance, the composition h ◦ pr : SO(16) → S8 of the orthonormal frame bundle pr : SO(16) → S15 with the Hopf map S15 → S8 is one such submersion.) Moreover, the simple group SO(8) admits a foliation, FSO(8), by totally geodesic round 7-spheres (obtained by trivializing the orthonormal frame bundle SO(8) → 7 S ). Kerin–Shankar examples does not have totally geodesic fibers and FSO(8) is not Riemannian. RIEMANNIAN FOLIATIONS ON COMPACT LIE GROUPS 3

The general classification of Riemannian foliations is wide open. For instance, classifications neither for totally geodesic Riemannian foliations on symmetric spaces, nor for generic Riemannian foliations on Lie groups are known (we refer to Lytchak [15], Lytchak–Wilking [16] and Wilking [26] for important developments in other cases). The author believes that the proof here can be partially replicated for the symmetric space case, giving important first steps.

1.1. Preliminaries and description of each step. Given a Riemannian foliation F on M, we might think of F locally as an stripped fabric (or a Riemannian submersion) with leaves vertically placed. At each point x ∈ M, we decompose ⊥ TxM as the tangent to the leaf Vx and its orthogonal complement Hx = (Vx) . We call Vx as the vertical space and Hx as the horizontal space at x. Given X ∈ TM, we denote Xh,Xv the horizontal and vertical components of X, respectively. A vector field X is said to be basic horizontal if it is H-valued and, for every vertical field V ,[X, V ] is vertical. Equivalently, if F is induced by a submersion π, X is basic if dπ(X) is fiberwise constant. The flow of a basic horizontal vector field X induces local diffeomorphisms between leaves (as a standard computation shows – see e.g. Hirsch [11, Proposition 17.6]). These (local) diffeomorphisms are called (local) transformations. It is known that holonomy transformations are (local) isometries if and only if leaves are totally geodesic (see e.g. Gromoll– Walschap [7, Lemma 1.4.3]), which is the case at hand. Given a Riemannian foliation F, the Gray–O’Neill integrability tensor A: H× H→V is defined by 1 ¯ ¯ v AX Y = 2 [X, Y ] , where X,¯ Y¯ are horizontal extensions of X, Y . We follow Ranjan [20] and denote AξX as the opposite dual of A defined by: ξ A X, Y = − hAX Y, ξi .

Let φt be the flow of a basic horizontal field X and c an integral curve of φt. For any given ξ ∈Vc(0), we define its holonomy field along c by

ξ(t)= dφt(ξ). Alternatively, ξ(t) is the only vector field along c that satisfies ξ ∇X ξ(t)= A X, ξ(0) = ξ.

# The dual leaf at p ∈ M, Lp , is the subset of points in M that can be joined to p by horizontal curves (compare Wilking [27] or Gromoll–Walschap [7, section 1.8]). When the leaves of F are the fibers of a principal G-bundle π : P → B, the integrability tensor, infinitesimal holonomy fields and dual leaves replace classi- cal objects: given a connection 1-form ω : T P → g, the curvature 2-form satis- fies Ω(X, Y ) = −2ω(AX Y ); for any holonomy field along a horizontal curve c, ωc(t)(ξ(t)) = ωc(0)(ξ(0)). That is, ξ(t) is the restriction to c of the action field # defined by ξ; Lp is the holonomy bundle through p (see e.g. [14, secnomition II] for a definition of the last). The celebrated Ambrose–Singer Theorem [1, Theorem # 2] identifies the Lie algebra of the holonomy group of π with ω(TpP (p)) = ω(Lp ). Theorem 1.3 refines this result in the case of Riemannian foliations/submersions on non-negatively curved ambient spaces: 4 LLOHANN D. SPERANC¸A

Theorem 1.3. Let F be a totally geodesic Riemannian foliation on a manifold M # of non-negative sectional curvature. Let Lp be the dual leaf of F through p. Then # TLp ∩Vp = span{AX Y | X, Y ∈ Hp}. Theorem 1.3 is used twice: to obtain a local version of the splitting theorem [15, Corollary 3.3] and as one of the last steps in the paper. When we are in the scope of Theorem 1.2 (i.e., F is a totally geodesic Riemannian foliation on a neighborhood of a Lie group with bi-invariant metric), Ranjan [20] makes a key observation: let e be the identity of G. For every ξ ∈Ve, X ∈ He, the Grey–O’Neill’s formulas imply: ξ 2 1 2 (2) (A ) = ( 2 adξ) . By using it, [20] proves Conjecture 1 for simple groups that have a maximal torus inside a leaf. Such torus provides a decomposition of the basic horizontal fields, producing candidates for g±. Then the simplicity of the group is used to prove that either g+ or g− is trivial. Without the maximal torus assumption, we introduce a new root system based on the integrability tensor of F. Theorem 1.4. Let F be a Riemannian foliation with totally geodesic leaves on a manifold M. Let Lp be the leaf through p ∈ M and assume that a neighborhood of v 0 in a subspace t ⊆ Vp exponentiates to a totally geodesic flat. Suppose that one of the hypothesis hold: (a) exp(tv) is a complete totally geodesic flat and A is bounded; (b) Le is the open neighborhood of a compact symmetric space without Euclidean factors. Then, R(η, ξ)h = AηAξ − AξAη for all ξ, η ∈ tv. Equation (2) together with Theorem 1.4 readily gives a decomposition X = v v X+ + X−, producing spaces H+(t )+ H−(t ) ⊇ He. The Lie algebras g+, g− are v v the ideals generated by H+(t ), H−(t ). The bulk of the paper is to prove that these ideals commute with each other. To this aid, we build upon Munteanu–Tapp [18, Theorem 1.5] (Theorem 1.5 below), which provides Lie algebraic relations between the original root system and the one in Theorem 1.4: A triple {X,V, A} ⊆ TpM is called a good triple if expp(tV (s)) = expp(sX(t)) for all s,t ∈ R, where V (s),X(t) denote the Jacobi fields along exp(sX) and exp(tV ) that satisfy V (0) = V , X(0) = X and V ′(0) = A = X′(0), respectively. Such conditions are achieved in totally geodesic Riemannian foliations by X ∈ H, V ∈V and A = AV X. [18, Theorem 1.5] provides a key identity that is used in section 4: Theorem 1.5 (Theorem 1.5, [18]). Let G be a compact Lie group with a bi-invariant metric and denote its Lie algebra by g. The triple {X,V,A}⊆ g is good if and only if, for all integers n,m ≥ 0, n m ¯ [adX B, adV B]=0, 1 ¯ 1 where B = 2 adV X − A and B = − 2 adV X − A.

Once proved that g+ commutes with g−, it follows from [18, Proposition 4.1, Corollary 4.2] that the foliation is given by cosets: on the one hand

Proposition 1.6 (Munteanu–Tapp [18], Proposition 4.1). Let F1, F2 be Riemann- ian foliations with totally geodesic leaves on M. Suppose that their vertical spaces, together with their integrability tensors coincide at a point e. Then F1 = F2. RIEMANNIAN FOLIATIONS ON COMPACT LIE GROUPS 5

On the other hand, assume that M is a (open) subset of the product Lie group G = G1 × G2, equipped with a bi-invariant metric. Consider a H < G1 × G2 and define FH as: −1 (3) L(g1,g2) = {(h1g1,g2h2 ) | (h1,h2) ∈ H}. Then,

(ξ1,ξ2) 1 1 (4) A (X1,X2) = ( 2 adξ1 X1, − 2 adξ2 X2).

Observe that inverting the second coordinate on G1 × G2 is an isometry and inter- ξ 1 changes F to a Riemannian foliation whose A-tensor satisfy A X = 2 adξ X, for all ξ ∈Ve,X ∈ He. Corollary 1.7 (Munteanu–Tapp [18], Corollary 4.2). Let F be a Riemannian fo- liation with totally geodesic leaves on a connected Lie group G with bi-invariant ξ 1 metric. If A X = 2 adξ X, for all ξ ∈ Ve,X ∈ He, then Ve is a subalgebra and F is the foliation defined by the left cosets of the subgroup whose subalgebra is Ve. Although [18, Proposition 4.1] assumes completeness of M, its proof can be carried out as far as every two points of M can be joined by the concatenation of vertical and horizontal geodesics. Theorem 1.2 follows by putting together (1), (4) and Corollary (1.7). The paper is divided as follow: in section 2 we prove Theorem 1.3 and use it to reduce Theorem 1.2 to the case of an irreducible foliation. Section 3 deals with the proof of Theorem 1.4 and presents the splitting V = ∆0 ⊕ ∆1. The proof of Theorems 1.1, 1.2 are completed in section 4.

The author would like to thank C. Dur´an, K. Shankar and K. Tapp for sugges- tions and insightful conversations. Specially K. Shankar for pointing out [2] (which was a crucial reference for an earlier version of the paper). The author also would like to thank Miguel Dom´ıngues V´azquez and the anonymous referee for many sug- gestions, and Universidade Federal do Paran´afor hosting the author for most part of this work.

2. An Ambrose-Singer theorem for non-negatively curved foliations Let π : M → B be a Riemannian submersion. For simplicity we assume that all submersions and foliations here have totally geodesic fibers/leaves. Define # Lp = {c(1) ∈ M | c: [0, 1] → M horizontal, c(0) = p}. We recall that, given a curvec ˜: [0, 1] → B, its horizontal lifts define the holonomy −1 −1 −1 diffeomorphism φc˜ : π (˜c(0)) → π (˜c(1)) (by sending a point q ∈ π (˜c(0)) to the endpoint of the horizontal lift ofc ˜ starting at q). When F is a Riemannian foliation, one can still define local diffeomorphisms, since F is locally given by submersions, therefore their differentials define holonomy fields. Thus, given a horizontal curve c: [0, 1] → M and ξ ∈ Vc(0), we denote dφc(ξ) = ξ(1), where ξ(t) is the holonomy field defined by ξ, i.e., it satisfies: ξ (5) ∇c˙ξ(t)= A c,˙ ξ(0) = ξ. # −1 When π is a principal H-bundle Lp ∩ π (b) coincides with an orbit of the holo- nomy group of π at b. In this case, the Ambrose–Singer Theorem [1] characterizes 6 LLOHANN D. SPERANC¸A the Lie algebra of the holonomy group through the connection 2-form Ω. The result naturally extends to the case of a (not necessarily principal) Riemannian foliation (as one can see from the proof in [3, section 3.4.2]): Theorem 2.1 (Ambrose–Singer [1], Theorem 2). Let F be a Riemannian foliation # and Lp the dual leaf at p ∈ M. Then, # −1 TpLp ∩Vp = span{dφc (AX Y ) | c horizontal, c(0) = p, X,Y ∈ H}. # Although the Theorem gives a semi-local characterization for TLp , one must understand the behavior of the holonomy fields and of the A-tensor, which might be quite arbitrary objects. The situation can be greatly improved when the ambient space has non-negative sectional curvature. Theorem 1.3. Let F be a totally geodesic Riemannian foliation on a manifold M # of non-negative sectional curvature. Let Lp be the dual leaf of F at p. Then # TLp ∩Vp = span{AX Y | X, Y ∈ Hp}.

Under such hypothesis, a vector in the cokernel of AX never leaves it (see Lemma 2.3). Observe that the result is absolutely local since there is no hypothesis on the completeness of M. Moreover, punctual information of A spreads out through M: for instance, by combining Theorems 2.1 and 1.3, one concludes that A = # # 0 at p if and only if Lp is a polar section for F (i.e., Lp intersects every leaf perpendicularly). In this case, one can show that the universal cover M˜ metrically ˜ ˜# ˜ splits as M = Lp × Lp (see [7, Theorem 1.4.1]). Now we proceed to the proof of Theorem 1.3. As in the introduction, denote ξ A : Hq → Hq as ξ A X,Z = − hAX Z, ξi . v ξ ξ ∇X ξ ξ h Denote (∇X A) Z = ∇X (A Z) − A Z − A ∇X Z. By extending ξ as a holonomy field, one sees that: ξ h(∇X A)X Z, ξi = − (∇X A) X,Z . In the remaining of the section, we assume F with totally geodesic leaves and M with non-negative sectional curvature. Theorem 1.3 is based on the next inequality. Lemma 2.2. For each p ∈ M there is a neighborhood U and a constant a> 0 such that ξ ξ (6) akXkkZkkA Xk ≥ | (∇X A) X,Z | for all X,Z ∈ Hq and ξ ∈Vq, q ∈ U. Proof. Given X,Z ∈ H and ξ ∈ V, Gray–O’Neill equations ([7, page 44]) states that the unreduced sectional curvature K(X, ξ + tZ) = R(X, ξ + tZ,ξ + tZ,X) satisfies 2 ξ 2 (7) K(X, ξ + tZ)= t K(X,Z)+2t h(∇X A)X Z, ξi + kA Xk . Since K(X, ξ + tZ) ≥ 0, the discriminant of the polynomial (7) satisfies ξ 2 2 0 ≤ K(X,Z)kA Xk − h(∇X A)X Z, ξi . On small neighborhoods, continuity of K guarantees the existence of some a > 0 such that K(X,Z) ≤ akXk2kZk2.  RIEMANNIAN FOLIATIONS ON COMPACT LIE GROUPS 7

Lemma 2.3. Let ξ(t) be a holonomy field along γ(t) = exp(tX), X ∈ H. If Aξ(0)X =0 then Aξ(t)γ˙ (t)=0 for all t.

ξ v Proof. Take kXk = 1 and Z = A γ˙ in (6). Recalling that ∇γ˙ ξ = 0, we get 1 d (8) akAξγ˙ k2 ≥ (∇ A)ξγ,˙ Aξγ˙ = ∇ (Aξγ˙ ), Aξγ˙ = kAξγ˙ k2. γ˙ γ˙ 2 dt

Equation (8) is Gronwall’s inequality for u(t)= kAξ(t)c˙(t)k2, implying that kAξ(t)γ ˙ (t)k2 ≤kAξ(0)γ˙ (0)k2e2at for all t> 0. In particular, if Aξ(0)γ˙ (0) = 0, Aξ(t)γ˙ (t) = 0. Analogously, Aξ(t)γ˙ (t)= 0 for t< 0 by replacing X by −X in the argument. 

ξ(t) The main ingredient in the proof is the constancy of the rank of ker(A : Hc(t) → ξ Hc(t)) along exp(ker A ) (Proposition 2.6). First we present two algebraic lemmas. Lemma 2.4. Let X, Y ∈ H be orthonormal and AξX =0. Then, ξ 2 ξ ξ ξ 2akA Y k ≥ (∇X A) Y + (∇Y A) X, A Y .

Proof. For any unitary Z ∈ H, (6) gives: ξ ξ ξ ξ ξ 2akZkkA (Y + X)k≥ (∇X A) X + (∇X A) Y + (∇Y A) X + (∇Y A) Y,Z , ξ ξ ξ ξ ξ 2akZkkA (Y − X)k≥− (∇X A) X − (∇X A) Y − (∇Y A) X + (∇Y A) Y,Z .

Observe that Aξ(X + Y )= Aξ(Y − X)= AξY and take Z = AξY . The result now follows by summing up both inequalities. 

Consider the non-negative symmetric operator D = −AξAξ and recall that ker Aξ = ker D. Given DY = λ2Y , λ > 0, define Y¯ = λ−1AξY . We have kY¯ k = kY k and AξY¯ = −λY . In particular, if kY k = 1, kDY k = λ2 and kAξY k = kAξY¯ k = λ. Lemma 2.5. Let X, Y be unitary horizontals satisfying AξX =0 and DY = λ2Y 6= 0. Then, ξ ξ ξ ξ ¯ ξ ¯ ξ ¯ (∇Y A) X, A Y + (∇Y¯ A) X, A Y = (∇X A) Y , A Y .

Proof. By combining the Bianch identity of Rv(X, Y )Z and Grey–O’Neill’s equa- v v tion R (X, Y )Z = −(∇Z A)X Y we get (see also Lemma 1.5.1 in [7, page 26]): ξ ¯ ξ ¯ ξ (∇Y A) X, Y = − (∇X A) Y , Y − (∇Y¯ A) Y,X .

By replacing AξY = λY¯ and AξY¯ = −λY , we have: ξ ξ ξ ¯ ξ ¯ ξ (∇Y A) X, A Y =λ (∇Y A) X, Y = −λ[ (∇X A) Y , Y + (∇Y¯ A) Y,X ] ξ ¯ ξ ¯ ξ ξ ¯ = (∇X A) Y , A Y − (∇Y¯ A) X, A Y . 

Proposition 2.6. Let ξ(t) be a holonomy field along γ(t) = exp(tX0), X0 ∈ ξ(0) 2 Hp, and suppose that A X0 = 0. If λ(t) is a continuous eigenvalue of D = −Aξ(t)Aξ(t), then λ(t) either vanishes identically or it never vanishes. Proof. We argue by contradiction. Assume that λ vanishes at t = 0 but there is l> 0 such that λ(t) > 0 for all t ∈ (0,l). We further assume (by possibly reducing l) 8 LLOHANN D. SPERANC¸A that D has a smooth frame of eigenvectors along γ((0,l)). The Proposition follows from Gronwall’s inequality once we prove that d (9) a′λ2 ≥ λ2, dt ′ for some a′ > 0. In particular, λ(t)2 ≤ λ(ǫ)2ea t for all ǫ ∈ (0,l), t ∈ (ǫ,l). Thus, λ must vanish on (0,l), a contradiction. Inequality (9) follows from Lemmas 2.4 and 2.5: let Y be a smooth unitary vector field satisfying DY = λ2Y . Applying Lemma 2.4 on both Y and Y¯ gives 2 ξ ξ ξ 2aλ ≥ (∇X A) Y + (∇Y A) X, A Y , 2 ξ ¯ ξ ξ ¯ 2aλ ≥ (∇X A) Y + (∇Y¯ A) X, A Y . Summing up gives: 2 ξ ξ ξ ξ 4aλ ≥ (∇X A) Y, A Y + (∇X A) Y¯ , A Y¯ ξ ξ ξ ξ ¯ + (∇Y A) X, A Y + (∇Y¯ A) X, A Y . ξ ξ ξ ξ ¯ ξ ¯ ξ ¯ Applying Lemma 2.5, we have (∇Y A) X, A Y + (∇Y¯ A) X, A Y = (∇X A) Y , A Y . On the other hand, ξ ξ ξ ξ ξ ξ (∇X A) Y¯ , A Y¯ = ∇X (A Y¯ ), A Y¯ − A (∇X Y¯ ), A Y¯ d d = 1 λ2 − λ2 ∇ Y,¯ Y¯ = 1 λ2. 2 dt X 2 dt

ξ ξ 1 d 2 ′ 8 Analogously, (∇X A) Y, A Y = 2 dt λ . Thus we can take a = 3 a, concluding the proof. 

As the last step, fix p ∈ M and denote aq = span{AX Y | X, Y ∈ Hq}. Observe ⊥ ξ that aq = {ξ ∈ Vq | A = 0} and recall that a horizontal curve can be smoothly approximated by a broken horizontal geodesic. Then, Proposition 2.6 gives: ⊥ ⊥ Corollary 2.7. For any curve c, dφc(ap )= ac(1).

Theorem 1.3 follows directly from Corollary 2.7: since dφc is an isometry, Lemma −1 2.7 implies that dφc (aφc(p)) = ap. Applying Theorem 2.1, one directly gets the #  equality TLp ∩Vp = ap.

⊥ # Once established Theorem 1.3, one can observe that ap coincides with ν(Lp ), # the space normal to Lp . It gives the very important local version of [15, Proposition 3.1] below. It guarantees the same thesis of [15, Proposition 3.1] by exchanging the completeness of M by the assumption of totally geodesic leaves in F. Corollary 2.8. Let F be a totally geodesic Riemannian foliation on M, a non- negatively curved manifold. Then the sectional curvature sec(ξ,X)=0, for every ξ ∈ ν(L#) and X ∈ H and ν(L#) is translated along exp(tX). Proof. Along the proof of Theorem 1.3, we have shown that the distribution p 7→ ⊥ # ap = ν(Lp ) is invariant along holonomy transformation defined by horizontal ⊥ geodesics. Moreover, if ξ(t) is the holonomy field defined by ξ ∈ ap along γ(t) = exp(tX), X ∈ Hp, then kAξ(t)γ˙ (t)k2 sec(ξ(t), γ˙ (t)) = ≡ 0. kγ˙ (t)k2kξ(t)k2 RIEMANNIAN FOLIATIONS ON COMPACT LIE GROUPS 9

⊥ On the one hand, ∇γ˙ (t)ξ(t) = 0, thus ξ(t) is parallel. On the other hand, ξ(t) ∈ aγ(t) for every t, therefore a⊥ is parallel along γ. 

2.1. Reduction to the single-dual-leaf case. Here we reduce the proof of The- orem 1.2 to the case of an irreducible F (i.e., with only one dual leaf). First we observe that it is sufficient to prove Theorem 1.2 locally: if {Ui} is an open cover of U such that F|Ui is the restriction of a coset foliation defined by a connected subgroup, then F|Ui and F|Uj must be the restriction of the same coset foliation whenever Ui ∩ Uj 6= ∅. To reduce to the case of only one dual leaf, we give a local version of the following result due to Lytchak (see also [21]): Theorem 2.9 (Corollary 3.3 [15]). Let F be a regular Riemannian foliation on a simply connected compact symmetric space M. Then there is a metric decomposi- tion M = M1 × M2 and a foliation F1 on M1 such that each slice M1 ×{x2} is a dual leaf for F and F satisfies

F = {L × M2 | L ∈F1}. Again, let F be a Riemannian submersion and consider the decomposition G = G1 × G2 given by Theorem 2.9. If F1 is given by left H-cosets, H < G1, then F is given by the left H × G2-cosets. We now establish our local version of Theorem 2.9: Proposition 2.10. Let F be a totally geodesic Riemannian foliation on a simply connected symmetric space M with non-negative sectional curvature. Then there is a metric decomposition M = M1 × M2 and a foliation F1 on M1 such that each slice M1 ×{x2} is a dual leaf for F and F satisfies

F = {L × M2 | L ∈F1}. Proof. Theorem 2.9 is based on [15, Proposition 3.1] and the completeness of dual leaves of regular Riemannian foliations on compact non-negatively curved manifolds ([27, Theorem 3, item (b)]). Here we argue on how to trade completeness by the assumption of totally geodesic leaves on both points. We first observe that dual leaves of (non-singular) totally geodesic foliations must be complete. In such case, holonomy transformations are local isometries of leaves, and can be lifted as full isometries defined on the whole universal cover o each leaf. Therefore the intersection of the dual leaf to a leaf L, L# ∩ L is (finitely covered by) the orbit of a proper on L˜. It follows that L# is complete since L# is invariant under holonomy transformations. On the other hand, the proof of Proposition 3.1 in [15] is based on simple Lie algebraic computations (by identifying M = G/K and g = TM ⊕ k) and the following two properties: sec(ν(L#), H) = 0 and that ν(L#) is invariant with respect to parallel translations along horizontal geodesics. These two facts are guaranteed by the arguments in [27] for singular Riemannian foliation on a complete ambient space with non-negative sectional curvature. In our case, these two facts were recovered by Corollary 2.8. 

Considering the arguments above, it is sufficient to prove Theorems 1.1, 1.2 assuming that F has only one dual leaf. This assumption is used in section 4.5 in order to apply Theorem 1.3. 10 LLOHANN D. SPERANC¸A

3. The A-root system Both in here and in section 4 we work with the complexification of some related spaces, specially g and He. Given a V , the complexification of V will be denoted by V C. Given an operator A: V → V , its natural complexification is denoted by the same letter A: V C → V C and is defined by A(x+iy)= A(x)+iA(y). We follow Knapp [13] and extend inner products on V to C-bilinear symmetric products on V C (not to Hermitian, positive definite ones). The usual setting for a root system consists of an abelian real Lie algebra t acting on a linear space V through a Lie algebra morphism ρ: t → End(V ). For instance, one may endow V with an inner product and suppose that ρ(t) ⊆ End(V ) is a subspace of commuting skew-adjoint linear endomorphisms of V . In this case, ρ naturally defines an action on V C and the subset ρ(t) ⊆ End(V C) consists of endomorphisms with pure imaginary eigenvalues that can be diagonalized in a single bases. The root decomposition induced by ρ(t) is defined by C V = Vα. αX∈Π where Vα is the weight space of the linear function α : t → iR: C Vα = {X ∈ V | ρ(A)X = α(A)X, ∀A ∈ t}.

Whenever Vα 6= {0}, for α 6= 0, we call α a root. We denote the set of roots by Π(t). v Let F be a Riemannian foliation. Let ı: t # Lp be an immersed totally geodesic ξ flat with ı(0) = p. Here we prove that, under certain conditions, ρA(ξ)= A defines v a representation of t on Hp. We use the following convention for the Riemannian curvature:

R(X, Y )Z = ∇X ∇Y Z − ∇Y ∇X Z − ∇[X,Y ]Z. Theorem 1.4. Let F be a Riemannian foliation with totally geodesic leaves on a manifold M. Let Lp be the leaf through p ∈ M and that a neighborhood of 0 in a v subspace t ⊆ Vp exponentiates to a totally geodesic flat. Suppose that one of the hypothesis hold: (a) exp(tv) is complete and A is bounded; (b) Le is the open neighborhood of a compact symmetric space. Then, R(η, ξ)h = [Aη, Aξ] for all ξ, η ∈ tv. We begin by proving item (a). Proof of item (a). Consider basic horizontal fields X, Y and vertical fields ξ, η such v that ∇ξη = ∇ηξ = 0 along exp(t ). We have, ξ η hR(η, ξ)X, Y i = h∇η∇ξX − ∇ξ∇ηX, Y i = ∇η(A X) − ∇ξ(A X), Y ξ η η ξ = − η hAX Y, ξi + ξ hAX Y, ηi− A X, A Y + A X, A Y η ξ ξ η = − h∇η(AX Y ), ξi + h∇ξ(AX Y ), ηi + (A A − A A )X, Y . h The proof is concluded by observing that R (η, ξ)Vp = 0, since fibers are totally geodesic, and h∇ξ(AX Y ), ηi = − h∇η(AX Y ), ξi = 0. For the last, consider the geodesic γ(s) = exp(tξ) and recall that AX Y is a Jacobi field along γ. We have,

ξξ hAX Y, ηi = h∇ξ∇ξ(AX Y ), ηi = hR(ξ, η)ξ, AX Y i =0. RIEMANNIAN FOLIATIONS ON COMPACT LIE GROUPS 11

Therefore, ϕ(t) = hAX Y, ηi (γ(t)) is a bounded affine function in the real line. In particular, ϕ(t) must be constant and h∇ξ(AX Y ), ηi = ξ hAX Y, ηi vanishes. 

In the proof above, the boundedness of A is used to show that AX Y |exp(tξ) is bounded. Fortunately, there is a natural way to ensure such a bound using only local information. Let F be as in Theorem 1.2 i.e., F is a totally geodesic Riemannian foliation on an open subset U ⊆ G of a Lie group G with bi-invariant metric. We assume that e ∈ U without lost of generality. Recall that, although Le is defined only on U, Le can be isometrically identified with an open neighborhood of a complete symmetric space L˜e ([10, Theorem 5.1]). Since L˜e must have non-negative sectional curvature, it is locally isometric to a product L0 × L1, where L0 is an and L1 is a compact symmetric space. It is easy to conclude that AX Y can only have unbounded components on L0. With this motivation in mind, we prove item (b)

Proof of item (b). Observe that every isometry ϕ : Le → Le can be extended to an isometry of L˜e (for instance, if φ: Le → Le is an isometry, then its graph is a closed totally geodesic submanifold Γ ⊆ Le × Le, thus a symmetric space by itself. Therefore, there is a unique symmetric space Γ˜ containing Γ as an open subset. One clearly sees that Γ˜ can be naturally identified as a submanifold of L˜e × L˜e which is the graph of an isometryϕ ˜). In particular, Killing fields on Le are the restriction of Killing fields in L˜e. Since L˜e is locally the product L0 × L1, a Killing field ζ in Le decomposes as the sum of a component ζ0 in L0 and ζ1 in L1. Since L1 is compact, ζ is unbounded only if ζ0 is unbounded. The result now follows since AX Y |Le is a Killing field, whenever X, Y are basic horizontal.  3.1. Splitting of totally geodesic foliations. At last, we observe that, even if Le has an Euclidean factor, we can still split the foliation and use Theorem 1.4 on the compact factor. The splitting we mean is stated in the next result, which should be either known or expected to hold among specialists. Theorem 3.1. Let F be a totally geodesic Riemannian foliation with simply con- s ˜ nected dual leaves. For a fixed L ∈F, let TL = i=0 ∆i be the de Rham decompo- sition of TL. Then, there are smooth integrableL distributions ∆0, ..., ∆s on M such that, for every i: s (1) ∆i is vertical and V = i=0 ∆i; s ′ ′ ′ ′ (2) for every leaf L ∈F, TLL = i=0 ∆i|L is the de Rham decomposition of TL ; (3) Fi, the foliation defined by ∆Li, is Riemannian and has totally geodesic leaves.

Let F be an irreducible Riemannian foliation with totally geodesic leaves. Let Lp ˜ be the leaf through p ∈ M and denote TLp = i ∆i as the de Rham decomposition of TLp. L Let c: [0, 1] → M be a horizontal curve. Recall that holonomy fields define a linear isometry dφc : Vc(0) → Vc(1), dφc(ξ)(ξ(1)), where ξ(t) is the holonomy field defined by ξ along c. Moreover, by recalling that Riemannian foliations are locally given by Riemannian submersions, one concludes that c defines an isometry between ˜ ˜ universal covers φc : Lc(0) → Lc(1). Fix a leaf L ∈F and TL = ∆˜ i, its de Rham decomposition. Define ∆i(q)= dφc(∆˜ i), where c is a horizontalL curve joining p to q. We claim that ∆i is well defined if dual leaves are simply connected. 12 LLOHANN D. SPERANC¸A

−1 Let c1,c2 : [0, 1] → M be horizontal curves joining L to q. Then (dφc2 ) dφc1 = dφc, where c is the concatenation of c1 with the reverse of c2. In particular, ˜ ˜ dφc1 (∆i)= dφc2 (∆i) for every pair of horizontal curves c1,c2, such that c1(0),c2(0) ∈ L and c1(1) = c2(1), if and only if dφc(∆˜ i)= ∆˜ i for every horizontal curve c, such that c(0),c(1) ∈ L. Denote

HolL = {φc : L˜p → L˜p | c(0),c(1) ∈ L, c horiontal}.

Observe that HolL is a subgroup of isometries of L˜p. Moreover, according to Eschenburg–Heintze [5], it is sufficient to show that HolL does not exchange factors of the de Rham decomposition of L˜. On the one hand, if HolL exchange factors, the action of HolL on L˜ has non- connected isotropy group at some point (if L˜ has two isometric factors M1 × M1, then the points in the diagonal have non-connected isotropy). On the other hand, if φc(p)= p, then c(0) = c(1) = p and dφc is naturally identified with the isotropy representation of φc at p. Therefore, it is sufficient to prove that the group:

Hp = {dφc : Vp →Vp | c(0) = c(1) = p, c horiontal} is connected.

Claim 3.2. The distribution ∆˜ i is Hp-invariant. Proof. Letπ ¯ : O(V) → M be the bundle of orthonormal frames of V, i.e.,

k O(V)= {b: R →Vp | b linear isometry}.

O(k) acts on O(V) by right composition. Observe that F˜ = {π¯−1(L) | L ∈ F} ˜ V defines a foliation on O(V). One can make F Riemannian by observing that ∇X ξ = v (∇X ξ) defines a O(k)-invariant F˜-horizontal distribution H˜ and equipping each fiber with a bi-invariant metric. Let b ∈ O(V) be such thatπ ¯(b) = p. Denote the F˜-dual leaf at b by Eb. One # −1 can observe thatπ ¯|Eb : Eb → Lp is a principal bundle with principal group b (Eb ∩ −1 −1 π¯ (p)) = b Hpb. By definition, every point in Eb is connected to b by a F˜- horizontal curve, in particular, by aπ ¯-horizontal curve (i.e., orthogonal to the

π¯-fibers). Therefore,π ¯|Eb is irreducible as a principal bundle, concluding that H is #  connected, since Lp is simply connected (see [14] for details).

Observe that ∆i ⊆V and ∆i|Lq is a de Rham decomposition in each leaf Lq.

Therefore ∆i|Lq integrates aL Riemannian foliation with totally geodesic leaves on each Lq. Since Lq is totally geodesic on M, the integral submanifolds of ∆i are totally geodesic on M. It is left to prove that the foliation defined by ∆i on M is Riemannian.

Claim 3.3. Let Fi be the foliation defined by ∆i. Then Fi is Riemannian. Proof. We follow Gromoll–Walschap [7, Theorem 1.2.1] and show that the Lie ⊥ ∆i ⊥ derivative LU g = 0 for every U ∈ ∆i. Observe that ∆i = H ⊕ (⊕i=6 j ∆j ) and: LU g(H, V) = 0 and LU g(H, H) = 0 since F is Riemannian and U verti- cal; LU g(∆j , ∆k) = 0, j, k 6= i, since the leaves of F are totally geodesic and the restriction of ∆i to each leaf is Riemannian.  RIEMANNIAN FOLIATIONS ON COMPACT LIE GROUPS 13

4. Good triples and the H±-decomposition From now on, we specialize to the case of a totally geodesic Riemannian foliation F on U ⊆ G, a compact Lie group with bi-invariant metric. Furthermore, we assume that F satisfy the thesis hypothesis in Theorem 1.4 and use it throughout. We call such a foliation as a Ranjan foliation. This section is the technical bulk of the paper and concludes the proof of Theorem 1.1. Here we decompose g in commuting ideals g+, g− satisfying (1) (we actually consider a third ideal g0 for technical reasons, but it can be incorporated in either g+ or g−.) The decomposition g+, g− will be achieved step by step: in section 4.1 we fix a maximal abelian subalgebra inside Ve and decompose the elements of He according to the relation between the root system of g and of Theorem 1.4. v v v The process produces subspaces H+(t ), H−(t ), H0(t ) ⊆ g; section (4.2) proves a v v strong commuting identity for H+(t ), H−(t ) which is used throughout; in section v 4.3 we expand H±(t ) to subspaces H±(F) which are independent of the choice v of t ; in 4.4 we prove that H±(F) commute, providing the very important Lemma 4.20. Using Lemma 4.20 and the irreducibility hypothesis, we put Theorem 1.3 into play in order to prove that adVe preserves the subalgebras generated by H±(F); finally, in section 4.5 we prove that the algebras generated by H±(F) are ideals and that they commute. Together with Lemma 4.20 and Proposition 1.6, it concludes the proof of Theorem 1.1. The foliation defined by (3) gives a picture of H±(F): let H < G = G1 × G2 and consider FH as the foliation defined by the orbits of (h1,h2) · (g1,g2) = −1 (h1g1,g2h2 ). Each vector in H(g1,g2) has a component tangent to G1 ×{g2} and other tangent to {g1}× G1. The subsets spanned by such components are the desired subspaces H+(F) and H−(F), respectively. In this case the ideals generated by H+(F), H−(F) clearly commute. The whole paper is dedicated to show that this is the general situation. All arguments in section 4.1-4.4 follows from Theorem 1.4 and Munteanu–Tapp’s Theorem 1.5. That is, assuming that both theorems hold at e, the results in section 4.1-4.4 holds. Section 4.5 further requires Lytchak’s decomposition Theorem 2.9 (or its local version, Proposition 2.10) to reduce the general case to the case of a single dual leaf, which is required to apply Theorem 1.3. Given a Ranjan foliation, the leaf through e, Le, is a totally geodesic submanifold v of a symmetric space, thus a symmetric space itself. In particular, t = t ∩Ve exponentiates to a maximal totally geodesic flat in Le, as long as t is a totally geodesic abelian subalgebra in g. At last, we recall that the (4, 0) Riemannian curvature tensor of a bi-invariant metric satisfies:

1 (10) R(X,Y,Z,W )= − h[X, Y ], [Z, W ]i . 4

4.1. The horizontal decomposition I. Consider a maximal vertical abelian sub- v v ′ algebra t ⊆Ve completed to a maximal abelian subalgebra t = t ⊕ t ⊆ g = TeG. v 1 ξ t and t act on g through the representations ρad(ξ) = 2 adξ and ρA(ξ) = A , respectively. 14 LLOHANN D. SPERANC¸A

Given linear maps α: tv → iR, α′ : t′ → iR, we consider the spaces:

C ′ 1 ′ ′ ′ v ′ ′ gα,α (t)= {X ∈ g | 2 adξ+ξ X = (α(ξ)+ α (ξ ))X, for all ξ ∈ t , ξ ∈ t }, v C 1 v gα(t )= {X ∈ g | 2 adξ X = α(ξ)X, for all ξ ∈ t }, v C ξ v Hα(t )= {X ∈ He | A X = α(ξ)X, for all ξ ∈ t }.

We call their elements as (α, α′)-weights, vertical α-weights and α-A-weight, respec- tively. Whenever one of such spaces is non-trivial, the corresponding linear map is called a root, vertical root or A-root, respectively. We denote the corresponding set of roots as Π(t), Πv(tv) and ΠA(tv). Observe that a vertical root α can always be completed to a root (α, α′): t → iR ′ ′ ′ v v of g by a linear function α : t → iR: since t commutes with t , gα(t ) can be further decomposed as v gα(t )= gα,α′ (t). Xα′ Conversely, the identification (t ⊕ t′)∗ = t∗ ⊕ (t′)∗ is given by the restrictionα ˜ 7→ ′ (˜α|tv , α˜|t′ ). In particular, if (α, α ) is a root, α is a vertical root. We take advantage of this two-level decomposition:

C v C v C g = (t ) + gα(t )= t + gα,α′ (t), α∈XΠv(tv ) (α,αX′)∈Π(t) and the decomposition based on Theorem 1.4:

C v v He = H0(t )+ Hα(t ), α∈XΠA(tv )

v ξ v v where H0(t ) = ∩ξ∈tv ker A . The gα,β(t)-, gα(t )-, Hα(t )-components of X will α be denoted by Xα,β, Xα, X , respectively. Our first step in this algebraic part is to relate A-weights to vertical weights.

Lemma 4.1. Let tv be a maximal vertical abelian subalgebra. Then,

A v v v C v v v v Π (t )= {α ∈ Π (t ) | He ∩ (gα(t )+ g−α(t )) 6= ∅} ⊆ Π (t ).

C Moreover, if X ∈ He is an α-A-weight, then

v v X = Xα + X−α ∈ gα(t )+ g−α(t ).

Proof. Since we are dealing with Ranjan foliations, Grey–O’Neill’s equations gives (compare Ranjan [20], equation (1.3)): 1 −AξAξX = R(X, ξ)ξ = − ad2 X 4 ξ

1 2 2 for every ξ ∈ Ve. Therefore, if X is either a α-A-weight or ( 2 adξ) X = α(ξ) X, we get 1 (11) (Aξ)2X = α(ξ)2X = ad2 X. 4 ξ RIEMANNIAN FOLIATIONS ON COMPACT LIE GROUPS 15

Recall that two roots α, β ∈ Πv(tv) satisfying α(ξ)2 = β(ξ)2 for all ξ must satisfy α = ±β. In particular,

v v 1 2 2 gα(t )+ g−α(t )= ker ( 2 adξ) − α(ξ) id , ξ\∈tv   v v ξ 2 2 Hα(t )+ H−α(t )= ker (A ) − α(ξ) id . ξ\∈tv   v v v v v v C Thus, Hα(t )+ H−α(t ) ⊆ gα(t )+ g−α(t ). Since Hα(t )+ H−α(t ) ⊆ He , we conclude: v v C v v  Hα(t )+ H−α(t )= He ∩ (gα(t )+ g−α(t )). In particular, all vertical roots appearing on horizontal vectors are the A-roots, i.e., α X = X0 + X = X0 + Xα, α∈ΠXA(tv ) α∈XΠA(tv ) where v ξ C X0 ∈ H0(t )= ∩ξ∈tv ker A = He ∩ (∩ξ∈tv ker adξ). v C C Following Lemma 4.1, we define the projections πǫ(t ): He → g , ǫ =0, +, − by v α v α α v π0(t )(X0 + X )= X0, π±(t )(X ) = (X )±α ∈ g±α(t ). v P v v C C So, X = π0(t )(X)+ π+(t )(X)+ π−(t )(X) for every X ∈ He . Since g is the complexification of g, gC inherits two natural objects: a complex conjugation and the extension of the bi-invariant inner product h, i to a symmetric C-bilinear form, also denoted by h, i. We emphasize that we choose the C-extension of h, i so both ξ A and adξ are skew-symmetric. In particular: α −α hX, Y i = hX0, Y0i + hXα, Y−αi = hX0, Y0i + X , Y . α∈ΠXA(tv ) α∈ΠXA(tv ) v We also observe that Hǫ(t ) are real subspaces, i.e., invariant by the complex conjugation, since they are (a sum of) the image of real operators restricted to the kernel of real operators: v v v ξ 1 v v H±(t ) ∩ (gα(t ) ⊕ g−α(t )) = (A ± 2 adξ)(Hα(t ) ⊕ H−α(t )) ξ 1 ξ 2 2 = (A ± ad )ker (A ) − α(ξ) id C . 2 ξ He   v Analogously, H0(t ) is the intersection of kernels of real operators. In particular, v v v H0(t ) is orthogonal to H+(t )+ H−(t ). We gather the notations/statements in the last paragraphs as a lemma: v v Lemma 4.2. Let t be a vertical abelian subagebra and πǫ(t ) be the projections defined by Lemma 4.1. Then: C v (1) X = X0 + X+ + X− for every X ∈ He , where Xǫ = πǫ(t )(X); v C (2) Hǫ(t ) are real subspaces of g ; v v v (3) H0(t ) ⊥ (H+(t )+ H−(t )), We state another technical lemma to be used in the next section. Lemma 4.3. Let tv be a maximal vertical subalgebra and t ⊇ tv a maximal torus. v ′ ′ v Then, t decomposes orthogonally as t = t ⊕ t with t ⊆ H0(t ). 16 LLOHANN D. SPERANC¸A

Proof. Let t ∈ t, l ∈ tv and decompose t in its vertical and horizontal components, v h 1 t = t + t . On the one hand, R(t,l) = 2 ad[t,l] = 0 . On the other hand, since fibers are totally geodesic, R(H,l,ξ,η) = 0 for all ξ, η ∈Ve. Thus 0= R(t,l,ξ,η)= R(th,l,ξ,η)+ R(tv,l,ξ,η)= R(tv,l,ξ,η). v v 1 v 2 v v In particular, R(t ,l,l,t ) = 4 ||[t ,l]|| = 0. Since l ∈ t is arbitrary and t maximal, tv ∈ tv. Since th = t − tv ∈ t, we conclude that [th,l] = 0 for all l ∈ tv, h v thus t ∈ H0(t ). 

′′ v We claim that, if t is an abelian subalgebra in H0(t ), there is a maximal abelian ′ v ′ ′′ subalgebra t ⊆ H0(t ) such that t ⊇ t . Indeed, recall that any abelian subalgebra in a compact Lie group can always be extended to a maximal abelian subalgebra. Therefore, t⊕t′′ can be extended to a maximal t. Since t, tv are arbitrary in Lemma, we conclude that there is a t′ ⊇ t′′, such that t = tv ⊕ t′.

4.2. The bracket identity. Fix a maximal vertical abelian subalgebra tv and complete it to a maximal abelian subalgebra t = tv ⊕ t′. We have:

C ′ ′ Proposition 4.4. For every X, Y ∈ He and (α, α ), (β,β ) ∈ Π(t),

[(X+)α,α′ , (Y−)β,β′ ]=0. v α We brake the proof into steps, stated as the next lemmas, and write π±(t )(X )= α X±. The first Lemma is a restatement of Theorem 1.5 taking into account the v π±(t )-decomposition. v C Lemma 4.5. Let ξ ∈ t , X ∈ He . Then, for all n,m ≥ 0, m n+1 [adX adξ X−, adξ X+]=0. Proof. Recall from the discussion in section 1 that {X,ξ,AξX} is a good triple for every X ∈ H and ξ ∈ V. We apply Theorem 1.5 to it. Let Xα be a α-A-weight and observe that 1 α α ξ α α α 2 adξ X± = ±α(ξ)X±, A (X+ + X−)= α(ξ)X . Thus: 1 ξ α α α α B = ( 2 adξ −A )X = α(ξ) (X+ − X−) − (X+ + X−) α∈XΠA(tv )  α = −2α(ξ)X− = adξ X−. α∈XΠA(tv )

Analogously, B¯ = − adξ X+. 

α Expanding the sum X± = X±, we get: C P v v v Corollary 4.6. Let X ∈ He , ξ ∈ t . Then, for all m ≥ 0 and β ∈ Π (t ), m α β α(ξ)[adX X−,X+]=0. α∈ΠXA(tv ) Proof. From Lemma 4.5, we have m n+1 n+1 m α β 0 = [adX adξ X−, adξ X+]= α(ξ)β(ξ) [adX X−,X+] Xα,β RIEMANNIAN FOLIATIONS ON COMPACT LIE GROUPS 17 for every n ≥ 0. Suppose ξ is such that α(ξ) 6= β(ξ) for every pair of distinct A-roots α 6= β. In this case, by taking enough values of n we conclude that m α β α(ξ)[adX X−,X+] = 0 (recall that the determinant of the Vandermonde matrix Pof a set of pairwise distinct values is non-zero). However, the set of such ξ’s is dense in tv, concluding the result for every ξ.  C v v Lemma 4.7. Let X ∈ He . Then, for all l ≥ 0 and α, β ∈ Π (t ), α l β [X−, adX0 X+]=0. Proof. We use induction on s in: for all m ≥ 0, m α s β (12) α(ξ)[adX X−, adX0 X+]=0. α∈XΠv(tv ) Observe that (12) holds for s = 0 (Corollary 4.6). As the induction hypothesis, we m α k+1 β assume that (12) holds for s ≤ k and compute [adX X−, adX0 X+] backwards:

m+1 α k β m α k β m α k β [adX X−, adX0 X+] = [[X0, adX X−], adX0 X+] + [[X−, adX X−], adX0 X+] m α k β m α k β m α k+1 β + [[X+, adX X−], adX0 X+]=adX0 [adX X−, adX0 X+] − [adX X−, adX0 X+] m α k β m α k β m α k β +adX− [adX X−, adX0 X+]−[adX X−, [X−, adX0 X+]]+[[X+, adX X−], adX0 X+]. That is,

m α k+1 β m α k β m+1 α k β (13) [adX X−, adX0 X+]=adX0 [adX X−, adX0 X+] − [adX X−, adX0 X+] m α k β m α k β m α k β +adX− [adX X−, adX0 X+]−[adX X−, [X−, adX0 X+]]−[[adX X−,X+], adX0 X+]. In order to apply the induction hypothesis, we multiply both sides by α(ξ) and sum in α. It follows that the first three terms on the right-hand-side vanish. We deal with the last term in a separate claim. m α Claim 4.8. α(ξ)[X+, adX X−]=0. P β m α A v Proof. It is sufficient to prove that α α(ξ)[X+, adX X−] = 0 for every β ∈ Π (t ). We use induction on s in: for everyPr ≥ 0, r β s α (14) α(ξ)[adX X+, adX X−]=0. α∈ΠXA(tv ) The case s = 0 is Corollary 4.6. Assuming that (14) holds for s ≤ k, we have r β k+1 α r β k α α(ξ)[adX X+, adX X−] =adX α(ξ)[adX X+, adX X−] α∈XΠA(tv ) α∈ΠXA(tv ) r+1 β k α  − α(ξ)[adX X+, adX X−]=0. α∈XΠA(tv )

α k β The 5th and the proof is completed once we observe that [X−, adX0 X+]=0 for A v any α ∈ Π (t ), provided (12) holds for s ≤ k. However, since X0 commutes with v α k β v t , the term [X−, adX0 X+] lies in gβ−α(t ). Therefore, each term in the induction hypothesis (12): α k β α(ξ)[X−, adX0 X+]=0. α∈ΠXA(tv ) lies in a different weight space. Since ξ is arbitrary, each term must vanish, con- cluding the proof.  18 LLOHANN D. SPERANC¸A

Proof of Proposition 4.4. We first prove the intermediate step X = Y by following along the same lines as in the proof of Corollary 4.6. Since X0 is itself horizontal ′ and does not influence X±, we consider X = t + X+ + X− where t ∈ t can be A v ′ ′ ′ chosen at our will. Given β ∈ Π (t ), denote Πβ = {β : t → iR | (β,β ) ∈ Π(t)}. Lemma 4.7 gives for all l ≥ 0, α l β ′ l α β ′ l α β 0 = [X−, adt X+]= β (t) [X−, (X+)β,β′ ]= β (t) [X−, (X+)β,β′ ]. ′ ′ βX∈Πβ βX∈Πβ ′ ′ Consider t such that the values β (t), β ∈ Πβ, are all distinct and nonzero. Taking enough values of l gives α β α β (15) 0 = [X−, (X+)β,β′ ]= [(X−)−α,α′ , (X+)β,β′ ]. ′ α X∈Π−α

′ α β Since α,β,β are fixed, each term [(X−)−α,α′ , (X+)β,β′ ] lies in a different root space. α β ′ ′ Therefore, (15) implies that [(X−)−α,α′ , (X+)β,β′ ] = 0 for all α, α ,β,β . Noting α α α β that (X±)±α,α′ = (X )±α,α′ , and writing X+ = X+, X− = X−, we get [(X−)−α,α′ , (X+)β,β′ ] = 0. P P To proceed, recall that g is the product of an abelian Lie algebra and a semi- simple Lie algebra. In particular, the root space gα,α′ (t) is 1-dimensional and C ′ ′ the brackets [, ]: gα,α′ (t) × gβ,β′ (t) → g are either zero, when (α + α ,β + β ) ∈/ Π(t) ∪{(0, 0)}, or are non-degenerate, i.e., [x, y] = 0 only if x =0 or y = 0. C ± ′ ′ ′ Let πα,α : g → gα,α (t) be the linear projection onto gα,α (t) and denote πα,α′ = v ′ ′ πα,α′ ◦ π±(t ). Suppose (−α + β, α + β ) is a root (if not, [(X−)−α,α′ , (Y+)β,β′ ] + − C trivially vanishes). Since [π−α,α′ (X), πβ,β′ (X)] = 0 for every X ∈ He , we conclude C + − that He = ker π−α,α′ ∪ker πβ,β′ . This is only possible if one of the kernels coincides C ′ ′ + v − v with He . In particular, for every pair α, α ,β,β ,[π−α,α′ (H+(t )), πβ,β′ (H−(t ))] = {0}. 

+ − C Proposition 4.4 shows, in particular, that ker πα,α′ ∪ ker π−α,−α′ = He . On the v ± other hand, Hǫ(t ) are real spaces therefore πα,α′ 6= {0} if and only if its complex ± ± conjugate, π−α,−α′ , satisfy π−α,−α′ 6= {0}. Putting together these two pieces of + − C information, we conclude that ker πα,α′ ∪ ker πα,α′ = He , i.e., each root (together with its negative) can appear as a component of at most one of th two spaces v v H+(t ), H−(t ). These arguments derive a central property of: v ′ v C (16) Υ±(t )= {(α, α ) ∈ Π(t ) | ∃X ∈ He , (X+)α,α′ 6=0}, which is a main object in the next section. We have shown: v v Corollary 4.9. Υ+(t) ∩ Υ−(t) = ∅. In particular, H+(t ) ∩ H−(t ) = {0} and v v H+(t ) ⊥ H−(t )= {0}. v The orthogonality follows since H±(t ) are real spaces. Remark 4.10. An important step both in here (see Corollary 4.20) and in [20] is to show that: v 1 (17) AX Y = 2 [X−, Y−] − [X+, Y+] .   ξ 1 1 v Such equality implies that A X = 2 adξ X+ − 2 adξ X− for all ξ ∈Ve. If one fix t and the same computations as in [20], one can show that equation (17) holds for RIEMANNIAN FOLIATIONS ON COMPACT LIE GROUPS 19

X = Xα, Y = Y β, when α 6= β. However, the information is lost for α = β. We prove (17) by using a much more refined decomposition. The horizontal decomposition II. We call an immersed subgroup H ֒→ G .4.3 as V-maximal if its adjoint representation leaves Ve invariant and it is transitive in the set of maximal vertical abelian subalgebras. That is, AdH (Ve)= Ve and, fixed v v t , every other maximal vertical abelian subalgebra is of the form Adh t . We recall a few points:

(1) if Le is a subgroup, then H = Fe is V-maximal; (2) if Le is an irreducible symmetric space which is not a Lie group, then a V- maximal H can be chosen as the subgroup whose Lie algebra is h = [Ve, Ve] (see Conlon [4] or Berestovskii-Nikonorov [2, Lemma 7]). It follows that h ⊆ He; ∗ −1 −1 (3) writing h (α, β) = (α ◦ Adh ,β ◦ Adh ), ∗ Π(Adh t)= {h (α, β) | (α, β) ∈ Π(t)};

(4) gh∗(α,β)(Adh t) = Adh(g(α,β)(t)). Given a V-maximal H, define the vector spaces

v H±(F)= H±(Adh t ), hX∈H C ⊥ H0(F)= He ∩ (H+(F)+ H−(F)) . v It is clear that Hǫ(F) is independent of H and t , however both are crucial in our prove of the commutation. We now state a main result: Theorem 4.11. Let F be a Ranjan foliation. Then

(1) H+(F)⊥H−(F); (2) H+(F) ∩ H−(F)= {0}; (3) [H+(F), H−(F)] = {0}. The current section is a preliminary step for Theorem 4.11, where we prove Proposition 4.12. Theorem 4.11 is proved in section 4.4. We fix an arbitrary maximal abelian subalgebras tv ⊆ t throughout.

v v Proposition 4.12. For every h ∈ H, H±(Adh t ) = Adh H±(t ).

The proof of Proposition 4.12 uses Proposition 4.4, to control the set of Adht- roots, and the next three lemmas.

v v Lemma 4.13. For every h ∈ H, H0(Adh t ) = Adh H0(t ). v C Proof. H0(t )= He ∩ξ∈tv ker adξ. Therefore:

v C Adh H0(t ) = (Adh He ) ∩ ∩ Adh ker adξ ξ∈tv  C C  = He ∩ ∩vker adAdhξ = He ∩ ∩ kerv adξ . ξ∈t  ξ∈Adht 

v v ∗ Let Υ(t)=Υ+(t ) ∪ Υ−(t ). Since Adg fixes He, h Υ(t) = Υ(Adh t). Moreover: ∗ Lemma 4.14. For every h ∈ H, Υ±(Adh t)= h (Υ±(t)). 20 LLOHANN D. SPERANC¸A

′ Proof. Given (α, α ) ∈ Υ+(t) ∪ Υ−(t), we prove that ± ∗ ′ Hα,α′ = {h ∈ H | h (α, α ) ∈ Υ±(Adh t)} + − + are open subsets of H. Note that Hα,α′ ∩ Hα,α′ = ∅ (Corollary 4.9) and Hα,α′ ∪ − v ∗ v Hα,α′ = H, since Υ(Adh t ) = h Υ(t ). Since H is connected, it is sufficient to ± prove that Hα,α′ are open. Analogous to the proof of Proposition 4.4, consider the projections ± ∗ ∗ ′ ± v πα,α′ (h)= πh α,h α ◦ π (Adh t ). ′ v ± Now suppose that (α, α ) ∈ Υ+(t ). Then, there is X such that πα,α′ (e) 6= 0. Moreover,

± ± ∗ 1 h α 1 Adh ξ ∗ ∗ ′ πα,α′ (h)(X)= πα,α′ (h)(X )= ∗ (( 2 adAdh ξ ±A )X)h α,h α 2h α(Adh ξ) v ∗ v for any Adh ξ ∈ Adh t such that h α(Adh ξ) = α(ξ) 6= 0. By fixing ξ ∈ t we ± see that πα,α′ (h) is continuous as a family of operators with respect to h. Thus, ± ± ′ ′ ± if πα,α′ (h)(X) 6= 0, πα,α′ (h )(X) 6= 0 for h close to h, concluding that Hα,α′ is open.  Given a set S ⊆ g, denote L(S) as the subalgebra generated by S. Define the auxiliary spaces:

H±(t)= L  g(α,β)(t) ; (α,βM)∈Υ±(t)   ⊥ H0(t) = (H+(t)+ H−(t)) . Proposition 4.4, Corollary 4.9 and invariance by complex conjugation guarantees that [H+(t), H−(t)] = {0}, H+(t)⊥H−(t) and H+(t) ∩ H−(t) = {0}. Moreover, Lemma 4.14 implies that H±(Adh t) = Adh H±(t) for all h ∈ H. C Proof of Proposition 4.12. Let π±(t): He → H±(t) be the projections defined by the decomposition g = H+(t) ⊕ H−(t) ⊕ H0(t). From Corollary 4.9, it follows that C v π±(t)(He )= H±(t ). From Lemma 4.14, π±(Adh t) = Adh ◦π±(t) ◦ Adh−1 , h ∈ H. Therefore, v C C H±(Adh t )= π±(Adh t)(He ) = Adh(π±(t)(Adh−1 He )) C v  = Adh(π±(t)(He )) = Adh(H±(t )).

Proposition 4.12 gives a new characterization of H±(F): H±(F) is the smallest v AdH -invariant subset containing H±(t ). 4.4. Proof of Theorem 4.11. In order to prove Theorem 4.11, fix tv and observe from Proposition 4.12 and Jacobi identity that [H+(F), H−(F)] = {0} if and only if v v θ [AdH H+(t ), H−(t )] = {0}. Moreover, every element in H can be written as e for some θ ∈ h, since g is compact and H is connected. Thus, a power series argument k v v guarantees that [H+(F), H−(F)] = {0} if and only if [adθ H+(t ), H−(t )] = {0} for every θ ∈ h and k ≥ 0. k v v Our next aim is to show, by brute force, that [adθ H+(t ), H−(t )] = {0}. We start by studying the elements of h. If Le is a subgroup, we can take h = VI . If Le is an irreducible symmetric space which is not a group, we chose h = [Ve, Ve]. However, Le might be reducible, so RIEMANNIAN FOLIATIONS ON COMPACT LIE GROUPS 21 we write Ve = ∆i, where exp(∆i) are locally irreducible symmetric spaces. An standard argumentL shows that the sum of the aforementioned options produces a V-maximal group.

Lemma 4.15. If i 6= j, then [∆i, ∆j ]=0. In particular, H is V-maximal, for h = hi where hi = ∆i whether ∆i is a subalgebra or hi = [∆i, ∆i] otherwise. P Proof. Since Le is totally geodesic and is locally isometric to a metric product exp(∆0)×· · ·×exp(∆s), the curvature tensor of G at the identity satisfies R(∆i, ∆j )= 1 2 {0}. Therefore, hR(ξ, η)η, ξi = 4 k[ξ, η]k = 0 for all ξ ∈ ∆i, η ∈ ∆j . In particular, h = hi integrates a subgroup which is, up to covering, a product H = H˜0 ×···× HP˜s. To see that H is transitive in the set of maximal vertical abelian subalgebras, v v note that a maximal abelian subalgebra of Ve splits as t = t ∩ ∆i (one can use arguments as in Lemma 4.3, for example). Thus, since eachL H˜i acts transitively on the set of abelian subalgebras of ∆i, H acts transitively on the set of maximal abelian subalgebras of Ve. 

Whenever ∆i is not a subalgebra, Besrestovskii–Nikonorov [2, Lemma 7] guar- antees that exp(∆i ⊕ hi) is the full subgroup of isometries of exp(∆i) and that (∆i ⊕ hi, hi) is a symmetric pair (note that [∆i, hi] ⊆ ∆i, since ∆i is a Lie triple system – see e.g. Helgason [10]). In this case, hi is horizontal: it is orthogonal to ∆j , j 6= i, since h[∆i, ∆i], ∆j i = h∆i, [∆j , ∆i]i = {0}; and orthogonal to ∆i since [∆i, hi] ⊆ ∆i We decompose h = hv ⊕ hh in its vertical and horizontal components, denoting ∨ ∧ by ∆ (respectively, ∆ ) the sum of the ∆i-components which are subalgebras (respectively, which are not subalgebras). Observe that [hh, hv] = {0} and decompose θ ∈ h in its horizontal and vertical components, θ = Z + ζ. We proceed by induction on m to show that: for all m,n ≥ 0, m n v v (18) [adζ adZ H+(t ), H−(t )] = {0}. First we show that (18) holds for m = 0 (Claim 4.16), then, assuming that (18) holds for m ≤ k, we show that it holds for m = k + 1.

n v v n v v Claim 4.16. [adZ H+(t ), H−(t )] = {0} and adZ H+(t ) ⊥ H−(t ) for all n ≥ 0. h v v v ′ Proof. Let Z ∈ h and write Zǫ = πǫ(t )(Z) ∈ Hǫ(t ). We choose t = t ⊕ t such ′ v v v that Z0 ∈ t . Since [H+(t), H−(t )] = {0}, H+(t) ⊥ H−(t ) and H+(t ) ⊆ H+(t), it is sufficient to show that H+(t) is adZ -invariant. But, adZ0 (H+(t)) ⊆ H+(t), since Z0 ∈ t and H+(t) is a sum of weight spaces; adZ+ (H+(t)) ⊆ H+(t), since  Z+ ∈ H+(t); adZ− (H+(t)) = {0}⊆ H+(t), by Proposition 4.4.

From now on, we assume hv 6= {0} and proceed to technical steps.

v ∧ m v C Claim 4.17. X± ∈ ∆ . In particular, adζ (H±(t )) ⊆ He , α α v v Proof. Let X± = (X )+ ∈ H±(t ) be the H±(t ) component of an α-A-weight. ∨ ∨ ∧ Recall that ∆ is a subalgebra and [∆ , ∆ ] = 0. Therefore, ad∆∨ preserves the ∨ ∧ C ∨ v decomposition ∆ ⊕ ∆ ⊕ He . Moreover, if α(ξ) 6= 0 for some ξ ∈ ∆ ∩ t , C 1 α 1 α α α He ∋ (α(ξ) ± 2 adξ)X = (α(ξ) ± 2 adξ)(X+ + X−)=2α(ξ)X±. 22 LLOHANN D. SPERANC¸A

α v ∨ v v C Therefore, (X+) 6= 0 only if α(ξ) = 0 for every ξ ∈ ∆ ∩ t . Since H0(t ) ⊆ He , α v ′ ′ v ∧ we conclude that (X+) 6= 0 only if α(ξ ) =1 for some ξ ∈ t ∩ ∆ . Thus,

α ∨ 1 ′ α ∨ 1 α ′ ∨ X±, ∆ = ± 2 adξ X±, ∆ = − 2 X±, adξ ∆ =0.

The Claim is concluded by observing that ∆∨, ∆∧ are real spaces. 

The next claim is a common induction step in the next proofs.

′ C v v ′ v Claim 4.18. Suppose that X ∈ He ∩ (H0(t )+ H+(t )) and [X , H−(t )] = 0. ′ C v v ′ v Then adζ X ∈ He ∩ (H0(t )+ H+(t )) and [adζ X , H−(t )]=0. ′ C v v Proof. Note that adζ X ∈ He ∩ (H0(t )+ H+(t )): ′ v ′ v (19) hadζ X , H−(t )i = hζ, [X , H−(t )]i =0. ′ v Therefore, since adζ X has no H−(t )-component, ′ ′ ′ ′ [adζ X , Y−] = [(adζ X )0, Y−] = [(adζ X0)0, Y−] + [(adζ X+)0, Y−]. We show that both terms are zero. Since tv ∩ ∆∨ is a maximal torus, we can write v ∨ v ζ = ζ0 + ζα, where ζ0 ∈ t ∩ ∆ and ζα ∈ gα(t ), α 6= 0. On the other hand, ′ v ′ ′ v v adζα X0 ∈Pgα(t ). Thus, (adζ X0)0 = (adζ0 X0)0 = 0, since adt H0(t ) = 0. ′ ′ ′ On its turn, the second term belongs to H−(t), for t such that (adζ X+)0 ∈ t . On ′ ′ ′ the other hand, by replacing X by X+ on equation (19), we have [(adζ (X+))0, Y−]= ′ [adζ X+, Y−]. Thus, ′ ′ ′ [adζ X+, Y−], H−(t) = [adζ Y−,X+], H−(t) = adζ Y−, [X+, H−(t)] =0. ′ ′  Since H−(t) is a real space, we conclude that [adζ X+, Y−] = [adζ X , Y−] = 0.

m C v v In particular, adζ X+ ∈ He ∩ (H0(t )+ H+(t )) for m ≥ 1.

C ′ Claim 4.19. For any X ∈ He , there is a decomposition X+ = X+ + X where C v ′ n C v v Y+ ∈ He ∩ H+(t ) and adζ X =0. Moreover, adZ X+ ∈ He ∩ (H0(t )+ H+(t )). α α Proof. We prove the Claim for each component X+ = (X )+ of X+, considering v ∨ α ′ separate cases: if α(t ∩ ∆ ) 6= {0}, X+ = X+ and X = 0 satisfies the desired conditions. On the other hand, since [∆∨, ∆∧]= {0}: α α ′ α adξ′ adζ X+ = adζ adξ′ X+ =2α(ξ ) adζ X+ ′ v ∧ α v for every ξ ∈ t ∩ ∆ . Note that adζ X+ ∈ H+(t ) (Claim 4.18 plus the fact that α ′ v adξ′ adζ X+ 6= 0 for some ξ ∈ t ), thus, adζ preserves the eigenspaces Vλ of adξ′ . α ′ ′ Thus, X+ can be decomposed as X+ + X , where X+ ∈ adζ (Vλ) and X ∈ ker adζ . C The second statement follows since adZ preserves He and Claim 4.16: n v n−1 v  (20) adZ X+, H−(t ) = Z, [adZ X+, H−(t )] =0.

Proof of Theorem 4.11. The main item in Theorem 4.11 is the third item, from where we start the proof. To this aim, we proceed by induction on m ≥ 1 on: m n C v v (21) adζ adZ X+ ∈ He ∩ (H0(t )+ H+(t )), m n v (22) [adζ adZ X+, H−(t )]=0, RIEMANNIAN FOLIATIONS ON COMPACT LIE GROUPS 23 for every n ≥ 0. From Claim (4.19), induction on (21), (22) is equivalent to induction on m n C v v adζ adZ X+ ∈ He ∩ (H0(t )+ H+(t )), m n v [adζ adZ X+, H−(t )]=0, ¯ C v for X+ ∈ He ∩ H+(t ). The last induction follows from Claim 4.18, concluding item (3) in Theorem 4.11. Since H±(F) are real spaces, item (2) follows from item (1). Moreover, item (1) m n v holds once it is proved that adζ adZ X+ ⊥ H−(t ) for every m,n ≥ 0. The case m = 0 is in Claim 4.16 and m ≥ 1 is (21).  As an application of Theorem 4.11, we characterize the A-tensor.

Lemma 4.20. Let F be a Ranjan foliation and write X = X0 + X+ + X−, where ξ 1 Xǫ ∈ Hǫ(F). Then A X = 2 adξ(X+ − X−). In particular, 1 v (23) AX Y = 2 ([X−, Y−] − [X+, Y+]) . v Proof. Fix X ∈ He and ξ ∈ Ve. Let t be a maximal abelian vertical subalgebra containing ξ. Using the linearity of Aξ, we divide the proof into two cases: (i) X v v has no H0(t )-component; (ii) X ∈ H0(t ). Denote by ǫ v v v π : H+(t )+ H−(t )+ H0(t ) → Hǫ(F) ǫ v C v π (t ): He → Hǫ(t ) the respective orthogonal projections. Supposing that π0(tv)(X) = 0, we have ξ ξ + v − v 1 + v 1 − v A X = A (π (t )(X)+ π (t )(X)) = 2 adξ(π (t )(X)) − 2 adξ(π (t )(X)). v ± ± v ± v ∓ ± v On the other hand, since H±(t ) ⊆ H±(F), π ◦π (t )= π (t ) and π ◦π (t )= 0. v ± v v Now, assume that X ∈ H0(t ). We claim that π (H0(t )) ⊆ H0(t ). By follow- ing Claim (4.16) and (21), we conclude that an element X± is composed by com- v ponents lying either in H0(t ) (components with m ≥ 1 in (21)) or in t⊇tv H±(t) (Claim 4.16). I.e., H+(F) decomposes as P

v (24) H±(F) = (H±(F) ∩ H0(t )) ⊕ H±(F) ∩ H±(t) . tX⊇tv   v 0 v Since the second space space is orthogonal to H0(t ), π (H0(t )) do not have com- ± v v ponents on it, thus concluding that π (H0(t )) ⊆ H0(t ). ξ 1 With A X = 2 adξ(X+ − X−) at hand, equation (23) is straightforward: ξ −2 hAX Y, ξi =2 A X, Y = hadξ(X+ − X−), Y i = hadξ(X+ − X−), Y+ + Y−i

= hξ, [X+, Y+ + Y−] − [X−, Y+ + Y−]i = hξ, [X+, Y+] − [X−, Y−]i .  4.5. Proof of Theorem 1.1. We now have all elements to prove Theorem 1.1. To simplify notation, we denote Hǫ(F)= Hǫ. As pointed out in section 2.1, it is sufficient to prove Theorem 1.1 for irreducible foliations. In this case, Theorem 1.3 guarantees that Ve is spanned by the image of the A-tensor. Thus Lemma 4.20 gives: C Ve ⊆ He + [H+, H+] + [H−, H−]. 24 LLOHANN D. SPERANC¸A

In particular: C (25) g ⊆ H0 + L(H+)+ L(H−), where L(S) is the Lie algebra generated by S ⊆ gC. Moreover:

Claim 4.21. [H0, Hǫ] ⊆ Hǫ. Proof. Observe that C ⊥ H0(F)= He ∩ (H+(F)+ H−(F)) C v v ⊥ v = He ∩ (H+(Adh t )+ H−(Adh t )) = H0(t ) h\∈H h\∈H C = {X ∈ He | adξ X =0, ∀ξ ∈VI }.

In particular, H0 is a subalgebra: let X, Y ∈ H0, ξ ∈Ve, then

adξ[X, Y ] = [adξ X, Y ] + [X, adξ Y ]=0. ⊥ C ⊥ C Moreover, [H0, Ve]= {0}, thus H0 , He and therefore H0 ∩ He are invariant under C adH0 . In particular, [H0, H+ + H−] ⊆ He ∩ (H+ + H−). On the other hand,

h[H0, H±], H∓i = − hH0, [H±, H∓]i =0. 

Corollary 4.22. L(H±) is an ideal. C Proof. Let Z ∈ g . Then Z = Z0 + Z+ + Z−, where Zǫ ∈ L(Hǫ) (equation (25)). But [Z0, L(H±)] ⊆ L(H±) by Claim 4.21; [Z±, L(H±)] ⊆ L(H±) by the definition of L(H±); and [Z∓, L(H±)] = {0} by Theorem 4.11 and Jacobi identity. 

Since L(H±) are real subspaces, their real parts are ideals of g, which we denote by L(H±) as well. In particular G decomposes as G = G+×G−×G0, where Gǫ is the subgroup whose Lie algebra is ⊥ g± = L(H±) ∩ (L(H+) ∩ L(H−)) , ⊥ g0 = (g+ + g−) + L(H+) ∩ L(H−).

By observing that H± ⊥ L(H∓), we conclude that H+ + H− ⊥ g0. 1 v ǫ We claim that AX Y = 2 ([X−, Y−] − [X+, Y+]) , where Xǫ ∈ g±. Denote π (Z) C 0 the Hǫ-component of Z ∈ He . Since [π (Z), ξ] = 0 for all ξ ∈Ve, we conclude that 0 0 [π (Z)ǫ, ξ] = [π (Z)ǫ, ξǫ]=0 for ǫ =0, +, −. Thus, − 0 − 0 h[X−, Y−], ξi = [π (X)+ π (X)−, π (Y )+ π (Y )−], ξ − − 0 − 0 0 = [π (X), π (Y )+ π (Y )−], ξ + π (Y )+ π (Y )−, [ξ, π (X)−] − − − 0 0 − 0 = [π (X), π (Y )], ξ + π (Y )+ π (Y )−, [ξ, π (X)−] + π (X), [π (Y )−, ξ] = [π−(X), π−(Y )], ξ .

Analogously, h[X+, Y+], ξi = h[π+(X), π(Y )+], ξi. The claim now follows from Lemma 4.20. The proof is almost finished and follows from Munteanu–Tapp Corollary 1.7. It is only left to produce an isometry of G whose resulting foliation satisfy AξX = 1 2 adξ X. Let Φ: G → G be the isometric involution defined by −1 Φ(g+,g−,g0) = (g+,g− ,g0), and consider F˜ = {Φ(L) | L ∈ F}. RIEMANNIAN FOLIATIONS ON COMPACT LIE GROUPS 25

Denote Z˜ = dΦ(Z), V˜e = dΦ(Ve) and by A˜ the A-tensor of F˜. Observe that ˜ ˜ dΦe(Z±)= ±Z± and recall that AX˜ Y = dφ(AX Y ). We have: ˜ ˜ ˜ 1 Ve AX˜ Y = dΦ(AX Y )= 2 (dΦ([X−, Y−] − [X+, Y+])) ˜ ˜ Ve 1 Ve 1 ˜ ˜ ˜ ˜ = 2 (−[X−, Y−] − [X+, Y+]) = − 2 [X−, Y−] + [X+, Y+] ,   where the third equality follows since [X±, Y±] ∈ g± and the last since

[X˜±, Y˜±] = [±X±, ±Y±] = [X±, Y±]. In particular, the respective dual tensor is given by ˜ξ˜ ˜ 1 ˜ A X = 2 adξ˜ X.

Corollary 1.7 guarantees that V˜I is a subalgebra and that F˜ is the coset fibration defined by the subgroup integrated by V˜e, completing the proof. 

References

1. W. Ambrose and I. M. Singer, A theorem on holonomy, Trans. Amer. Math. Soc. 75 (1953), 428–443. MR 0063739 2. V. Berestovskii and Y. Nikonorov, Clifford-wolf homogeneous riemannian manifolds, J. Diff. Geo. 82 (2009), 467–500. 3. A. Clarke and B. Santoro, Holonomy groups in riemannian geometry, arXiv preprint arXiv:1206.3170 (2012). 4. L. Conlon, A class of variationally complete representations, Journal of Differential Geometry 7 (1972), no. 1-2, 149–160. 5. J.-H. Eschenburg and E. Heintze, Unique decomposition of Riemannian manifolds, Proc. Amer. Math. Soc. 126 (1998), no. 10, 3075–3078. MR 1473665 6. E.´ Ghys, Feuilletages riemanniens sur les vari´et´es simplement connexes, Annales de l’institut Fourier, vol. 34, 1984, pp. 203–223. 7. D. Gromoll and G. Walshap, Metric foliations and curvature, Birkhuser Verlag, Basel, 2009. 8. K. Grove, Geometry of, and via, symmetries, University Lecture Series - Amer. Math. Soc. 27 (2002), 31–51. 9. A. Haefliger, Feuilletages sur les vari´et´es ouvertes, Topology 9 (1970), no. 2, 183–194. 10. S. Helgason, Differential geometry, lie groups and symmetric spaces (1978), 1978. 11. M. W. Hirsch, S. Smale, and R. L. Devaney, Differential equations, dynamical systems, and an introduction to chaos, second edition, Academic press, 2004. 12. M. Kerin and K. Shankar, Riemannian submersions from simple compact lie groups, Mnster Journal of Math. 5 (2012), 25?40. 13. A. W. Knapp, Lie groups beyond an introduction, vol. 140, Springer Science & Business Media, 2013. 14. S. Kobayashi and K. Nomizu, Foundations of differential geometry, vol. I, Interscience Pub- lishers, 1963. 15. A. Lytchak, Polar foliations of symmetric spaces, Geometric and Functional Analysis 24 (2014), no. 4, 1298–1315. 16. A. Lytchak and B. Wilking, Riemannian foliations of spheres, Geometry & Topology 20 (2016), no. 3, 1257–1274. 17. P. Molino and G. Cairns, Riemannian foliations, Birkhauser Boston Inc., 1988. 18. M. Munteanu and K. Tapp, Totally geodesic foliations and doubly ruled surfaces in a compact lie group, Proc. Amer. Math. Soc. 139 (2011), 4121–4135. 19. B. O’Neill, The fundamental equations of a submersion, Michigan Math. J. 13 (1966) 4 (1966), 459–469. 20. A. Ranjan, Riemannian subersions of compact simple Lie groups with connected totally geo- desic fibres, Math. Z 191 (1986), 239–246. 26 LLOHANN D. SPERANC¸A

21. R. J. M. Silva and L. D. Speran¸ca, On the completeness of dual foliations on nonnegatively curved symmetric spaces, arXiv preprint arXiv:2006.13809 (2020). 22. L. Speran¸ca, On Riemannian foliations over positively curved manifolds, J. Geo. Anal., 1–19. 23. W. Thurston, On the construction and classification of foliations, Proceedings of the ICM, Vancouver, vol. 1, 1974, pp. 547–549. 24. , The theory of foliations of codimension greater than one, Commentarii Mathematici Helvetici 49 (1974), no. 1, 214–231. 25. , Existence of codimension-one foliations, Annals of Mathematics (1976), 249–268. 26. B. Wilking, Index of closed geodesics and rigidity of hopf fibrations, Inventiones math- ematicae 144 (2001), no. 2, 281–295. 27. , A duality theorem for riemannian foliations in nonnegative sectional curvature, Geom. Func. Anal. 17 (2007), 1297–1320.

Universidade Federal de Sao˜ Paulo, ICT, Av. Cesare Monsueto Giulio Lattes, 1211 - Jardim Santa Ines I, CEP 12231-280, Sao˜ Jose´ dos Campos, SP, Brazil E-mail address: [email protected]