1 Introduction

Let \((Z,\omega )\) be a connected Kähler manifold with a holomorphic action of a complex reductive group \(U^\mathbb {C},\) where \(U^\mathbb {C}\) is the complexification of a compact connected Lie group U with Lie algebra \({\mathfrak {u}}.\) We also assume \(\omega \) is U-invariant and that there is a U-equivariant momentum map \(\mu : Z \rightarrow {\mathfrak {u}}^*.\) By definition, for any \(\xi \in {\mathfrak {u}}\) and \(z\in Z,\) \(d\mu ^\xi = i_{\xi _Z}\omega ,\) where \(\mu ^\xi (z) := \mu (z)( \xi )\) and \(\xi _Z\) denotes the fundamental vector field induced on Z by the action of U,  i.e.,

$$\begin{aligned} \xi _Z(z) := \frac{d}{dt}\bigg \vert _{t=0} \exp (t\xi )z \end{aligned}$$

(see, for example, Kirwan 1984 for more details on the momentum map). Since U is compact we may identify \({\mathfrak {u}} \cong {\mathfrak {u}}^*\) by means of a \({\textrm{Ad}}(U)\)-invariant scalar product \(\langle \cdot , \cdot \rangle \) on \({\mathfrak {u}}.\) Hence, we consider a momentum map as a \({\mathfrak {u}}\)-valued map, i.e., \(\mu : Z \rightarrow {\mathfrak {u}}.\) Recently, the momentum map has been generalized to the following settings (Heinzner and Schwarz 2007; Heinzner et al. 2008).

We say that a subgroup G of \(U^\mathbb {C}\) is compatible if G is closed and the Cartan decomposition \(U^\mathbb {C}=U\exp (\text {i} {\mathfrak {u}})\) induces a Cartan decomposition of G. This means that the map \(K\times {\mathfrak {p}} \rightarrow G,\) \((k,\beta ) \mapsto k\text {exp}(\beta )\) is a diffeomorphism where \(K := G\cap U\) and \({\mathfrak {p}} := {\mathfrak {g}}\cap \text {i}{\mathfrak {u}};\) \({\mathfrak {g}}\) is the Lie algebra of G. In particular K is a maximal compact subgroup of G with Lie algebra \({\mathfrak {k}}\) and \({\mathfrak {g}} = {\mathfrak {k}}\oplus {\mathfrak {p}}.\)

Using \(\langle \cdot , \cdot \rangle ,\) we define an \({\textrm{Ad}}(U)\)-invariant scalar product on \(\text {i}{\mathfrak {u}}\) requiring multiplication by \(\text {i}\) to be an isometry between \({\mathfrak {u}}\) and \(\text {i}{\mathfrak {u}}.\) The G-gradient map \(\mu _{\mathfrak {p}}:Z \longrightarrow {\mathfrak {p}}\) associated with \(\mu \) is the orthogonal projection of \(\text {i}\mu \) onto \({\mathfrak {p}}.\) If \(\beta \in {\mathfrak {p}}\) then

$$\begin{aligned} \mu _{\mathfrak {p}}^\beta (z) := \langle \mu _{\mathfrak {p}}(z), \beta \rangle = \langle \text {i}\mu (z), \beta \rangle = \langle \mu (z), -\text {i}\beta \rangle = \mu ^{-\text {i}\beta }(z), \end{aligned}$$

for any \(\beta \in {\mathfrak {p}}\) and \( z\in Z.\) In this paper, a G-invariant compact connected locally closed real submanifold X of Z is fixed and the restriction of \(\mu _{\mathfrak {p}}\) to X is also denoted by \(\mu _{\mathfrak {p}}.\) Then \(\mu _{\mathfrak {p}}:X \longrightarrow {\mathfrak {p}}\) is a K-equivariant map such that \( \text {grad}\mu _{\mathfrak {p}}^\beta = \beta _X, \) where the gradient is computed with respect to the induced Riemannian metric on X denoted by \((\cdot ,\cdot ).\) By the linearization Theorem (Heinzner et al. 2008; Sjamaar 1998), \(\mu _{\mathfrak {p}}^\beta \) is a Morse–Bott function (Biliotti et al. 2013; Heinzner et al. 2008) and the limit

$$\begin{aligned} \varphi _\infty ^\beta (x):=\lim _{t\rightarrow +\infty } \exp (t\beta )x, \end{aligned}$$

exists and belongs to \(X^\beta :=\{z\in X:\, \beta _X (z)=0\}\) for any \(x\in X.\) The linearization theorem (Heinzner et al. 2008; Sjamaar 1998) also proves that any connected component of \(X^\beta \) is an embedded submanifold, see for instance (Biliotti et al. 2013; Heinzner et al. 2008).

Let \(C_1,\ldots , C_k\) be the connected components of \(X^\beta .\) Let \(W_i:=\{x\in X:\, \lim _{t\mapsto +\infty } \exp (t\beta )x \in C_i\}.\) Then \(\mu _{\mathfrak {p}}^\beta (C_i)=c_i\) and applying again the linearization theorem (Heinzner et al. 2008; Sjamaar 1998), the submanifold \(C_i\) is a connected component of \((\mu _{\mathfrak {p}}^\beta )^{-1} (c_i).\) One of the most important Theorem of Morse theory proves that \(W_i\) is an embedded submanifold, which is called unstable manifold of the critical submanifold \(C_i,\) and \(\varphi _\infty ^\beta : W_i \longrightarrow C_i\) is smooth (Bott 1954).

Let T be a torus of U. This means that T is a connected compact Abelian subgroup of U (Adams 1969). By a Theorem of Koszul, (Duistermaat and Kolk 2000), the connected components of \(Z^T:=\{x\in Z:\, T\cdot x=x\}\) are embedded Kähler submanifolds of Z. Let \({\mathfrak {t}}\) be the Lie algebra of T. It is well-known that the set

$$\begin{aligned} \left\{ \beta \in {\mathfrak {t}}:\, \overline{\exp (\mathbb {R}\beta )}=T \right\} , \end{aligned}$$

is dense in \({\mathfrak {t}},\) see for instance (Adams 1969). Hence,

$$\begin{aligned} Z^T=Z^{T^\mathbb {C}}=\left\{ p\in Z:\, \beta _Z(p)=0\right\} , \end{aligned}$$
(1)

for some \(\beta \in {\mathfrak {t}}.\) This means \(Z^T\) is the set of the singularities of the vector field \(\beta _Z,\) i,e., the zero of the vector field \(\beta _Z.\) Moreover, \(Z^T\) is the image of the gradient flow \(\varphi _\infty ^\beta \) defined by \(\mu ^\beta .\)

In this paper, we investigate the fixed point set of the action of an Abelian compatible subgroup of \(U^\mathbb {C}\) acting on a real submanifold of Z.

Let \({\mathfrak {a}}\subset {\mathfrak {p}}\) be an Abelian subalgebra and \(A=\exp ({\mathfrak {a}}).\) Notice that A is automatically closed in G and hence compatible, since \(\{e\} \times {\mathfrak {a}}\) is closed in \(K\times {\mathfrak {p}}.\) Then the A-gradient map on X is given by \(\mu _{{\mathfrak {a}}}=\pi _{\mathfrak {a}}\circ \mu _{\mathfrak {p}},\) where \(\pi _{\mathfrak {a}}:{\mathfrak {p}}\longrightarrow {\mathfrak {a}}\) denotes the orthogonal projection of \({\mathfrak {p}}\) onto \({\mathfrak {a}}.\) Since A is Abelian, then by Lemma 2.2 below for any \(p\in X\) the stabilizer \(A_p=\{a\in A:\, ap=p\}=\exp ({\mathfrak {a}}_p),\) where \({\mathfrak {a}}_p\) is the Lie algebra of \(A_p.\) Therefore \(X^A=\{p\in X:\, A\cdot p=p \}=\{ p\in X:\, \beta _X (p)=0,\, \forall \beta \in {\mathfrak {a}}\}.\) Hence, if \(\alpha _1,\ldots ,\alpha _n\) is a basis of \({\mathfrak {a}}\) then \(X^A\) is the set of the common singularities of the commuting vector fields \((\alpha _1)_X, \ldots , (\alpha _n)_X.\) Our first main result is the following

Theorem 1.1

The set \(\left\{ \beta \in {\mathfrak {a}}:\, X^\beta =X^A \right\} \) is dense in \({\mathfrak {a}}.\)

Hence \(X^A\) is the set of the singularities of a vector field \(\beta _X\) for some \(\beta \in {\mathfrak {a}}\) and so the critical points of the Morse–Bott function \(\mu _{\mathfrak {p}}^\beta .\)

We point out that \(X^A\) contains a lot of information of the geometry of both the A gradient map and the G gradient map. Indeed, for any \(x\in X,\) \(\mu _{\mathfrak {a}}(A\cdot x)\) is an open convex subset of \(\mu _{\mathfrak {a}}(x)+{\mathfrak {a}}_x\) and \(\overline{ \mu _{\mathfrak {a}}(A\cdot x) }={{\textrm{conv}}} (\mu _{\mathfrak {a}}(X^A \cap \overline{A\cdot x)}),\) see Atiyah (1982), Biliotti and Ghigi (2018) and Heinzner and Schützdeller (2010), where \({\textrm{conv}}(\cdot ) \) denotes the convex hull of \((\cdot ).\) In particular \(\mu _{\mathfrak {a}}(X^A)\) is a finite set and \({\textrm{conv}} (\mu _{\mathfrak {a}}(X))={\textrm{conv}} (\mu _{\mathfrak {a}}(X^A))\) and so a polytope. Moreover, if \({\mathfrak {a}}\subset {\mathfrak {p}}\) is a maximal Abelian subalgebra, then \({\textrm{conv}} (\mu _{\mathfrak {p}}(X))\) is given by \(K{\textrm{conv}} ( \mu _{\mathfrak {a}}(X))\) (Biliotti et al. 2016).

The second main result proves the existence of \(\beta \in {\mathfrak {a}}\) such that the limit map associated with the gradient flow of \(\mu _{\mathfrak {p}}^\beta \) defines a map from X onto \(X^A.\) Hence, the set \(X^A\) is the image of the gradient flow of the Morse–Bott function \(\mu _{\mathfrak {p}}^\beta \) for some \(\beta \in {\mathfrak {a}}.\)

Let \(\alpha _1,\ldots ,\alpha _n\in {\mathfrak {a}}\) be a basis of \({\mathfrak {a}}.\) Then \(\varphi _\infty ^{\alpha _n}\circ \cdots \circ \varphi _\infty ^{\alpha _1}\) defines a map from the manifold X onto \(X^{\alpha _1} \cap \cdots \cap X^{\alpha _n}=X^A.\)

Theorem 1.2

Let \(\alpha _1,\ldots ,\alpha _n\in {\mathfrak {a}}\) be a basis of \({\mathfrak {a}}.\) There exists \(\delta >0\) such that for any \(0<\epsilon _2,\ldots ,\epsilon _n < \delta \) we have \( \varphi _\infty ^{\alpha _1+\epsilon _2 \alpha _2 +\cdots + \epsilon _n \alpha _n}=\varphi _\infty ^{\alpha _n}\circ \cdots \circ \varphi _\infty ^{\alpha _1}. \)

2 Proof of the Main Results

Suppose \(X\subset Z\) is G-invariant compact connected real submanifold of Z with the gradient map \(\mu _{\mathfrak {p}} : X\rightarrow {\mathfrak {p}}.\) If \(x\in X\) then \(G_x=\{g\in G:\, gx=x\}\) denotes the stabilizer of G at x. If \(G_x\) acts on a manifold S,  then \(G \times ^{G_x}S\) denotes the associated bundle with principal bundle \(G \rightarrow G/G_x\) defined as the quotient of \(G \times S\) by the \(G_x\)-action \(h (g, s) = (gh^{-1}, h s).\) We recall the Slice Theorem, see Heinzner et al. (2008) for details.

Theorem 2.1

(Slice Theorem (Heinzner et al. 2008, Thm. 3.1; Sjamaar 1998) If \(x \in X\) and \(\mu _{\mathfrak {p}}(x) = 0,\) there are a \(G_x\)-invariant decomposition \(T_x X = {\mathfrak {g}}\cdot x \oplus W,\) open \(G_x\)-invariant neighborhood S of \(0 \in W,\) a G-stable open neighborhood \(\Omega \) of \(x\in X\) and a G-equivariant diffeomorphism \(\Psi : G \times ^{G_x}S \rightarrow \Omega \) where \(\Psi ([e, 0]) =x.\)

Corollary 2.1.1

If \(x \in X\) and \(\mu _{\mathfrak {p}}(x) = \beta ,\) there are a \((G^\beta )_x\)-invariant decomposition \(T_x X = {\mathfrak {g}}^\beta \cdot x \, \oplus W,\) open \((G^\beta )_x\)-invariant neighborhood S of \(0\subset W,\) a \(G^\beta \)-stable open neighborhood \(\Omega \) of \(x\in X\) and a \(G^\beta \)-equivariant diffeomorphism \(\Psi : G^\beta \times ^{{(G^\beta )}_x} S \rightarrow \Omega \) where \(\Psi ([e, 0]) =x.\)

This follows applying the previous theorem to the action of \(G^\beta \) on X. Indeed, it is well known that \(G^\beta =K^\beta \exp ({\mathfrak {p}}^\beta )\) is compatible (Biliotti et al. 2013, Lemma 2.7, p.584) and the orthogonal projection of \(\text {i} \mu \) onto \({\mathfrak {p}}^\beta \) is the \(G^\beta \)-gradient map \(\mu _{{\mathfrak {p}}^\beta }\) associated with \(\mu \) (Heinzner et al. 2008). The group \(G^\beta \) is also compatible with the Cartan decomposition of \((U^\mathbb {C})^{\beta }=(U^\mathbb {C})^{\text {i} \beta }=(U^{\text {i}\beta } )^\mathbb {C}\) and \(\text {i} \beta \) is fixed by the \(U^{\text {i}\beta }\)-action on \({\mathfrak {u}}^{\text {i}\beta }.\) A momentum map of the \((U^\mathbb {C})^{\text {i} \beta }\)-action on Z is given by \(\widehat{\mu _{{\mathfrak {u}}^{\text {i}\beta }}(z)}=\pi _{{\mathfrak {u}}^{\text {i}\beta }} \circ \mu +\text {i}\beta ,\) where \(\pi _{{\mathfrak {u}}^{\text {i}\beta }}\) is the orthogonal projection of \({\mathfrak {u}}\) onto \({\mathfrak {u}}^{\text {i}\beta },\) i.e., \(U^{\text {i}\beta }\)-shifted momentum map by an element of the center of \({\mathfrak {u}}^{\text {i}\beta }.\) Then, the associated \(G^{\beta }\)-gradient map with \(\widehat{\mu _{{\mathfrak {u}}^{\text {i}\beta }}(z)}\) is given by \(\widehat{\mu _{{\mathfrak {p}}^\beta }} := \mu _{{\mathfrak {p}}^\beta } - \beta \) and so \(\widehat{\mu _{{\mathfrak {p}}^\beta }} (x)=0.\) Now, the result follows by Theorem 2.1. In particular, if G is commutative, then we have a Slice Theorem for G at every point of X,  see Heinzner et al. (2008, p.169) and Sjamaar (1998) for more details.

If \(\beta \in {\mathfrak {p}},\) then \(\beta _X\) is a vector field on X,  i.e. a section of the bundle TX. For \(x\in X,\) the differential is a map \(T_x X \rightarrow T_{\beta _X (x)}(TX).\) If \(\beta _X (x) =0,\) there is a canonical splitting \(T_{\beta _X (x)}(TX) = T_x X \oplus T_x X.\) Accordingly, the differential of \(\beta _X,\) regarded as a section of TX,  splits into a horizontal and a vertical part. The horizontal part is the identity map. We denote the vertical part by \(\mathrm d \beta _X (x).\) The linear map \(\mathrm d \beta _X (x)\in {\text {End}}(T_x X)\) is indeed the so-called intrinsic differential of \(\beta _X,\) regarded as a section in the tangent bundle TX,  at the vanishing point x. Let \(\{\varphi _t=\exp (t\beta )\} \) be the flow of \(\beta _X.\) There is a corresponding flow on TX. Since \(\varphi _t(x)=x,\) the flow on TX preserves \(T_x X\) and there it is given by \(d\varphi _t(x) \in {\text {Gl}}(T_x X).\) Thus we get a linear \(\mathbb {R}\)-action on \(T_x X\) given by \(\mathbb {R}\times T_x X \longrightarrow T_x X,\, (t,v)\mapsto d\varphi _t(x) (v).\) The flow of the vector field \(\beta _X\) defines an action of \(\mathbb {R}\) on X,  i.e., \(\mathbb {R}\times X \longrightarrow X,\, (t,x)\mapsto \exp (t\beta )x.\)

Corollary 2.1.2

If \(\beta \in {\mathfrak {p}}\) and \(x \in X\) is a critical point of \(\mu _{\mathfrak {p}}^\beta ,\) then there are open \(\mathbb {R}\)-invariant neighborhoods \(S \subset T_x X\) and \(\Omega \subset X\) and an \(\mathbb {R}\)-equivariant diffeomorphism \(\Psi : S \rightarrow \Omega ,\) such that \(0\in S, x\in \Omega ,\) \(\Psi ( 0) =x.\) (Here \(t\in \mathbb {R}\) acts as \(d\varphi _t(x)\) on S and as \(\varphi _t\) on \(\Omega .)\)

Proof

Since \(\exp :{\mathfrak {p}}\longrightarrow G\) is a diffeomorphism onto the image, the subgroup \(H:=\exp (\mathbb {R}\beta )\) is closed and so it is compatible. Hence, it is enough to apply the previous corollary to the H-action on X and the value at x of the corresponding gradient map.\(\square \)

Lemma 2.2

Let \({\mathfrak {a}}\subset {\mathfrak {p}}\) be an Abelian subalgebra and let \(A=\exp ({\mathfrak {a}})\) which is closed and compatible. If \(x\in X,\) then \(A_x\) is compatible,  i.e.,  \(A_x=\exp ({\mathfrak {a}}_x).\)

Proof

If \(a\in A_x,\) then \(a=\exp (\beta )\) for a \(\beta \in {\mathfrak {a}}.\) Let \( f(t)=\langle \mu _{\mathfrak {a}}(\exp (t\beta )x), \beta \rangle .\) Then \(f(1)= \langle \mu _{\mathfrak {a}}(\exp (\beta )x), \beta \rangle = \langle \mu _{\mathfrak {a}}(ax), \beta \rangle = \langle \mu _{\mathfrak {a}}(x), \beta \rangle = f(0)\) and \(f'(t)=\parallel \beta _X (\exp (t\beta )x)\parallel ^2\ge 0 .\) This implies \(\beta _X (x)=0\) and so \(\beta \in {\mathfrak {a}}_x,\) proving \(A_x = \exp ({\mathfrak {a}}_x).\) \(\square \)

Let \(\alpha ,\beta \in {\mathfrak {p}}\) be such that \([\alpha ,\beta ]=0\) and let \({\mathfrak {a}}\) be the vector space in \({\mathfrak {p}}\) generated by \(\alpha \) and \(\beta .\) By the above Lemma, it follows that \( X^A=X^\beta \cap X^\beta ,\) where \(A=\exp ({\mathfrak {a}}),\) which is closed and compatible due to the fact that the exponential map is a diffeomorphism restricted on \({\mathfrak {p}}.\)

Lemma 2.3

Let \(\beta , \alpha \in {\mathfrak {p}}\) be such that \([\beta ,\alpha ] = 0.\) If X is compact,  then there exists \(\delta > 0\) such that for any \(\epsilon \in (0, \delta ),\) \(X^{\beta + \epsilon \alpha } = X^\beta \cap X^\alpha .\)

Proof

Let \(\epsilon >0\) and let \(A = \exp ({\mathfrak {a}}),\) where \({\mathfrak {a}}={\textrm{span}}(\alpha ,\beta ).\) Let \(X^A\) denote the fixed point set of A,  i.e., \(X^A=\{z\in X:\, A\cdot x=x\}.\) By Lemma 2.2, \(X^A=X^\beta \cap X^\alpha .\) Corollary 2.1.2 applies for A and \(H=\exp (\mathbb {R}(\alpha +\epsilon \beta )).\) Therefore \(X^\beta \cap X^\alpha \) and \(X^{\alpha +\epsilon \beta }\) are compact submanifolds satisfying \(X^\beta \cap X^\alpha \subseteq X^{\alpha +\epsilon \beta }.\)

Let C be a connected component of \(X^\alpha \cap X^\beta .\) C is a compact connected submanifold of X and so it is arcwise connected. If \(x\in C\) then \(C\subseteq C',\) where \(C'\) is the connected component of \(X^{\alpha +\epsilon \beta }\) containing x. On the other hand, if L is a connected component of \(X^{\alpha +\epsilon \beta }\) then L is A-stable and so there exists a A-gradient map (Heinzner et al. 2008). Since L is compact the norm square A-gradient map has a maximum. By Heinzner et al. (2008, Corollary 6.12) L has a fixed point of A. This implies that L contains a connected component of \(X^{\alpha }\cap X^\beta .\) Summing up, we have proved that the number of the connected components of \(X^\alpha \cap X^\beta \) is greater than or equal to the number of connected components of \(X^{\alpha +\epsilon \beta }\) and any connected component of \(X^{\alpha +\epsilon \beta }\) contains at least a connected component of \(X^\alpha \cap X^\beta .\)

Let \(C_1,\ldots ,C_m\) be the connected components of \(X^\alpha \cap X^\beta .\) Let \(C_i'\) denote the connected component of \(X^{\alpha +\epsilon \beta }\) containing \(C_i,\) for \(i=1,\ldots ,m.\) We point out that \(C_k'\) would coincide with \(C_j'\) for \(k\ne i.\) We shall prove that there exists \(\delta >0\) such that for any \(\epsilon < \delta \) the connected components \(C_1',\ldots ,C_m'\) are pairwise disjoints and \(C_i=C_i'\) for \(i=1,\ldots ,m.\)

Let \(x_i \in C_i.\) Since \(x_i\) is fixed by A,  Corollary 2.1.2 implies there exists A-invariant open subsets \(\Omega \) of \(x_i\in X\) and S of \(0\in T_{x_i} X\) and a A-equivariant diffeomorphism \(\varphi : S \rightarrow \Omega \) such that \(\varphi (0) = x_i,\) \(d\varphi _0 = id_{T_{x_i} X}.\) Since the \(\mathbb {R}\)-action on S is linear and \(\varphi \) is \(\mathbb {R}\)-equivariant, we may assume that \(S=\Omega =\mathbb {R}^n\) by means of \(\varphi ,\) \(\alpha ,\beta \) are symmetric matrices of order n satisfying \([\alpha ,\beta ] = 0.\) Moreover, \(T_{x_i} X^{\alpha +\epsilon \beta }={\textrm{Ker}}\, (\alpha +\epsilon \beta )\) and \(T_{x_i} X^{\alpha }\cap T_{x_i} X^{\beta }={\textrm{Ker}}\, \alpha \cap {\textrm{Ker}}\, \beta .\)

The matrices \(\alpha \) and \(\beta \) are simultaneously diagonalizable. Let \(\{e_1,\ldots ,e_n\}\) be a basis of \(\mathbb {R}^n\) such that \(\alpha e_k =a_k e_k\) and \(\beta e_k =b_k e_k\) for \(k=1,\ldots , n.\) Let \(J=\{1\le k \le n:\, a_k b_k \ne 0\}.\) Pick \(\delta _{i} ={\textrm{min}}\{\frac{| a_k |}{| b_k |}:\, k\in J\}.\) Now, \((\alpha +\epsilon \beta )e_k=0\) if and only if \(a_k+\epsilon b_k=0.\) If \(a_k \ne 0,\) then \(b_k \ne 0\) and vice-versa. If \(\epsilon <\delta _i\) then \( (\alpha +\epsilon \beta )e_k=0, \) if and only if \(a_k=b_k=0.\) Therefore, \({\textrm{Ker}}\, (\alpha +\epsilon \beta )={\textrm{Ker}}\, \alpha \cap {\textrm{Ker}}\, \beta .\) This implies \(T_{x_i} C_i = T_{x_i} C_i'.\) Although \(\delta _i\) depends on \(x_i,\) since \(C_i\subseteq C_i'\) and both are compact submanifolds it follows that \(C_i=C_i'.\) Pick \(\delta ={\textrm{min}}(\delta _1,\ldots ,\delta _k).\) Then for any \(\epsilon <\delta \) we have \(C_i=C_i'\) for \(i=1,\ldots ,m.\) In particular \(C_1',\ldots ,C_m'\) are pairwise disjoints. Since the number of the connected components of \(X^{\alpha +\epsilon \beta }\) is less than or equal to the number of connected components of \(X^\alpha \cap X^\beta ,\) it follows that for any \(\epsilon <\delta \) both \(X^\alpha \cap X^\beta \) and \(X^{\alpha +\epsilon \beta }\) have the same connected components and so \(X^\alpha \cap X^\beta =X^{\alpha +\epsilon \beta }\) concluding the proof.\(\square \)

Theorem 2.4

Let \({\mathfrak {a}}\subset {\mathfrak {p}}\) be an Abelian subalgebra and let \(A=\exp ({\mathfrak {a}}).\) Then the set

$$\left\{ \alpha \in {\mathfrak {a}}:\, X^A=X^\alpha \right\} $$

is dense.

Proof

Let \(\alpha _1,\ldots ,\alpha _n\) be a basis of \({\mathfrak {a}}.\) Then

$$\begin{aligned} X^A=X^{\alpha _1} \cap \cdots \cap X^{\alpha _n}. \end{aligned}$$

By the above Lemma, there exists \(\delta >0\) such that for any \(\epsilon _2,\ldots ,\epsilon _n < \delta ,\) we have

$$\begin{aligned} X^A =X^{\alpha _1+ \epsilon _2 \alpha _2+\cdots +\epsilon _n \alpha _n} \end{aligned}$$
(2)

Let \(\alpha \in {\mathfrak {a}}\) different form 0. It is well known that there exists \(\alpha _2,\ldots \alpha _n\in {\mathfrak {a}}\) such that \(\alpha ,\alpha _2,\ldots ,\alpha _n\) is a basis of \({\mathfrak {a}}.\) By (2), for any neighborhood U of \(\alpha ,\) there exists \(\beta \in U\) such that \( X^A =X^\beta , \) concluding the proof.\(\square \)

The following lemma is proved in Biliotti and Windare (2023), see also Bruasse and Teleman (2005, pag. 1036).

Lemma 2.5

Let \(x\in X\) and \(\beta , \alpha \in {\mathfrak {p}}\) be such that \([\beta , \alpha ]= 0.\) Set \(y:= \lim _{t\rightarrow \infty }\exp (t\beta )x\) and \(z:= \lim _{t\rightarrow \infty }\exp (t\alpha )y.\) Let \(\delta \) be as in Lemma 2.3. Then for \(0<\epsilon <\delta ,\)

$$\begin{aligned} \lim _{t\rightarrow \infty }\exp (t(\beta + \epsilon \alpha ))x = z. \end{aligned}$$

As a consequence of the above lemma, we get the following result.

Theorem 2.6

Let \(\alpha _1,\ldots ,\alpha _n\) be a basis of \({\mathfrak {a}}.\) Let \(x\in X.\) Set \(x_1:= \lim _{t\rightarrow \infty }\exp (t\alpha _1) x\) and \(x_{i}= \lim _{t\rightarrow \infty }\exp (t\alpha _i) x_{i-1}\) for \(i=2,\ldots ,n.\) Then there exists \(\delta >0\) such that for \(0<\epsilon _2,\ldots ,\epsilon _n<\delta ,\) we have

$$\begin{aligned} \lim _{t\rightarrow \infty }\exp (t(\alpha _1 + \epsilon _2 \alpha _2 +\cdots + \epsilon _n \alpha _n))x = x_n, \end{aligned}$$

for any \(x\in X.\) In particular,  \( \varphi _\infty ^{\alpha _1+\epsilon _2 \alpha _2 +\cdots + \epsilon _n \alpha _n}=\varphi _\infty ^{\alpha _n}\circ \cdots \circ \varphi _\infty ^{\alpha _1}. \)

Proof

By Theorem 2.4, there exists \(\delta >0\) such that for any \(0<\epsilon _2,\ldots ,\epsilon _n<\delta ,\) we have

$$\begin{aligned} X^A=X^{\alpha _1+\epsilon _2 \alpha _2+\cdots +\epsilon \alpha _n}. \end{aligned}$$

Let \(A=\exp ({\mathfrak {a}}).\) Let \(z\in X^A.\) By Corollary 2.1.2, there exists A-invariant open subsets \(\Omega \subset X\) and \(S\subset T_z X\) and a A-equivariant diffeomorphism \(\varphi : S \rightarrow \Omega \) such that \(0\in S,\) \(z\in \Omega ,\) \(\varphi (0) = z,\) \(d\varphi _0 = id_{T_z X}.\) Let \(x\in X.\) Set \(x_1:= \lim _{t\rightarrow \infty }\exp (t\alpha _1) x\) and \(x_{i}= \lim _{t\rightarrow \infty }\exp (t\alpha ) x_{i-1}\) for \(i=2,\ldots ,n.\) If \(x_n \in \Omega ,\) keeping in mind that \(\Omega \) is A-invariant, it follows that \(x_1,\ldots ,x_n\in \Omega .\) If one reads carefully the proof of Lemma 2.3, then \(\delta >0\) works whenever that \(y\in \Omega .\) The same argument applies in this case. Hence there exists \(\delta \) such that for any \(0<\epsilon _2,\ldots ,\epsilon _n <\delta ,\) we have

$$\begin{aligned} \lim _{t\mapsto } \exp (t(\alpha _1 + \epsilon _2\alpha _2+\cdots +\epsilon \alpha _n))x=x_n \end{aligned}$$

whenever \(x_1,\) and so \(x_1,\ldots ,x_n,\) belongs to \(\Omega .\) By compactness of \(X^A\) there exist open subsets \(\Omega _1,\ldots ,\Omega _k\) satisfying the above property and such that

$$\begin{aligned} X^A \subseteq \Omega _1 \cup \cdots \cup \Omega _k. \end{aligned}$$

Let \(\delta _1,\ldots .\delta _k\) as before. Pick \(\delta ={\textrm{min}} (\delta _1,\ldots ,\delta _k).\) Let \(x\in X.\) Set \(x_1:= \lim _{t\rightarrow \infty }\exp (t\alpha _1) x\) and \(x_{i}= \lim _{t\rightarrow \infty }\exp (t\alpha _i ) x_{i-1}\) for \(i=2,\ldots ,n.\) Since \(x_1\in \Omega _j\) for some \(j=1,\ldots ,k,\) it follows that for any \(0<\epsilon _2,\ldots ,\epsilon _n<\delta \) we have

$$\begin{aligned} \lim _{t\mapsto +\infty } \exp (t(\alpha _1 +\epsilon _2 \alpha _2 +\cdots +\epsilon _n \alpha _n))x=x_n. \end{aligned}$$

This holds for any \(x\in X,\) concluding the proof.\(\square \)