In our specific setting, the measures are always nonnegative, and we only use the equivalence between CD ( K , N ) $\mathsf {CD}(K,N)$ and the smooth Ricci lower bound Ric N , V ⩾ K $\mathrm{Ric}_{N,V} \geqslant K$ . Once the gaps in the RCD $\mathsf {RCD}$ theory for negative effective dimension will have been filled, one should be able to replace CD ( K , N ) $\mathsf {CD}(K, N)$ with RCD ( K , N ) $\mathsf {RCD}(K, N)$ when N < 0 $N < 0$ in this paper, for example, in (iii) of Theorem 1.2
Pontryagin's Maximum PrincipleIf γ $\gamma$ is length minimiser parameterized by constant speed, then there exists a Lipschitz curve λ ( t ) ∈ T γ ( t ) ∗ M $\lambda (t) \in T^*_{\gamma (t)}M$ such that one and only one of the following is satisfied:
(i) λ ̇ = H ⃗ ( λ ) $\dot{\lambda } = \overrightarrow{H}(\lambda)$ , where H ⃗ $\overrightarrow{H}$ is the unique vector field in T ∗ M $T^*M$ such that σ ( · , H ⃗ ( λ ) ) = d λ H $\sigma (\cdot, \overrightarrow{H}(\lambda)) = \mathrm{d}_\lambda H$ for all λ ∈ T ∗ M $\lambda \in T^*M$ ;
(ii) ⟨ λ ( t ) , X i ( γ ( t ) ) ⟩ = 0 $\langle \lambda (t), X_i(\gamma (t)) \rangle = 0$ for all i = 1 , ⋯ , m $i = 1, \dots, m$ , and λ ( t ) ≠ 0 $\lambda (t) \ne 0$ for all t ∈ [ 0 , 1 ] $t\in [0,1]$ .
A curve λ : 0 , 1 → T ∗ M $\lambda: \left[0, 1\right] \rightarrow T^*M$ satisfying (i) (resp., (ii)) in the theorem above is called a normal (resp., abnormal) extremal. Theorem 2.5 is thus stating that a (constant speed) minimizing geodesic has a cotangent lift that is a normal or an abnormal extremal. Note that an extremal in a two-dimensional almost-Riemannian manifold is abnormal if and only if its projection is a constant curve that lies on the singular set (see [1, Theorem 9.2]).
For the remainder of this work, we will adopt the notation ( x ) 2 α : = ( x 2 ) α ∈ R ⩾ 0 $(x)^{2 \alpha }:= (x^2)^\alpha \in \mathbb {R}_{\geqslant 0}$ for every α ⩾ 0 $\alpha \geqslant 0$ and x ∈ R $x\in \mathbb {R}$ .
GEOMETRY OF α $\alpha$ -GRUSHIN HALF-SPACES
The α $\alpha$ -Grushin sphere and hemisphere
On the two-dimensional Riemannian sphere ( S 2 , g S 2 ) $(\mathbb {S}^2, g_{\mathbb {S}^2})$ , we introduce the following coordinate chart, the validity of which can be found in [11]. Fix a large circle γ : R / 2 π Z → S 2 $\gamma:\mathbb {R}/2\pi \mathbb {Z}\rightarrow \mathbb {S}^2$ , and let N $N$ and S $S$ be the north pole and south pole of S 2 $\mathbb {S}^2$ , respectively, with respect to γ $\gamma$ . For p ∈ S 2 ∖ { N , S } $p\in \mathbb {S}^2\setminus \lbrace N,S\rbrace$ , we define x : = x ( p ) ∈ ( − π / 2 , π / 2 ) $x:=x(p)\in (-\pi/2,\pi/2)$ as the signed (spherical) distance d S 2 ( p , Im ( γ ) ) $\mathop {}\!\mathrm{d}_{\mathbb {S}^2}(p,\mathrm{Im}(\gamma))$ , where the sign is positive (resp., negative) if p $p$ belongs to the hemisphere containing N $N$ (resp., S $S$ ). Furthermore, define the number y = y ( p ) ∈ R / 2 π Z $y=y(p)\in \mathbb {R}/2\pi \mathbb {Z}$ so that γ ( y ) $\gamma (y)$ is the perpendicular foot from p $p$ to Im ( γ ) $\mathrm{Im}(\gamma)$ . The map S 2 → ( − π / 2 , π / 2 ) × R / 2 π Z : p ↦ ( x ( p ) , y ( p ) ) ${\mathbb{S}}^{2}\to (-\pi/2,\pi/2)\ensuremath{\times{}}\mathbb{R}/2 \pi \mathbb{Z}:p\mapsto (x(p),y(p))$ is a well-defined coordinate chart which can be naturally extended to N $N$ and S $S$ under the identification ( π 2 , y 1 ) ∼ ( π 2 , y 2 ) $(\frac{\pi }{2},y_1) \sim (\frac{\pi }{2},y_2)$ and ( − π 2 , y 1 ) ∼ ( − π 2 , y 2 ) $(-\frac{\pi }{2},y_1) \sim (-\frac{\pi }{2},y_2)$ for any y 1 , y 2 ∈ R / 2 π Z $y_1,y_2\in \mathbb {R}/2\pi \mathbb {Z}$ . Note that N $N$ (resp., S $S$ ) corresponds to the equivalence class of ( π 2 , y ) $(\frac{\pi }{2},y)$ (resp., ( − π 2 , y ) $(-\frac{\pi }{2},y)$ ). It is easily shown that in these coordinates, the spherical Riemannian metric tensor g S 2 $g_{\mathbb {S}^2}$ possesses the warped product structureg S 2 = d x ⊗ d x + cos 2 ( x ) d y ⊗ d y . $$\begin{equation*} g_{\mathbb {S}^2} = \mathop {}\!\mathrm{d}x \otimes \mathop {}\!\mathrm{d}x + \cos ^2(x) \mathop {}\!\mathrm{d}y \otimes \mathop {}\!\mathrm{d}y. \end{equation*}$$ 3.1
Remark
The coordinate system ( x , y ) $(x, y)$ is, up to rotations, the same as the spherical coordinates ( φ , θ ) ∈ ( − π / 2 , π / 2 ) × R / 2 π Z $(\varphi,\theta) \in (-\pi /2, \pi /2)\times \mathbb {R}/ 2\pi \mathbb {Z}$ which parameterizes the standard sphere S 2 ∖ { N , S } ⊆ R 3 $\mathbb {S}^2 \setminus \lbrace N, S\rbrace \subseteq \mathbb {R}^3$ by( θ , φ ) ↦ ( cos ( θ ) cos ( φ ) , sin ( θ ) cos ( φ ) , sin ( φ ) ) . $$\begin{equation*} (\theta,\varphi)\mapsto (\cos (\theta)\cos (\varphi), \sin (\theta)\cos (\varphi),\sin (\varphi)). \end{equation*}$$
With this in mind, we can introduce the α $\alpha$ -Grushin sphere and hemisphere.
3.2
Definition
For α ⩾ 0 $\alpha \geqslant 0$ , the α $\alpha$ -Grushin sphere S α $\mathbb {S}_\alpha$ is the sub-Riemannian structure on S 2 $\mathbb {S}^2$ induced from the vector field X $X$ and Y α $Y^\alpha$ given byX : = ∂ x , Y α : = | sin ( x ) | α cos ( x ) ∂ y . $$\begin{equation*} X:={\partial}_{x},\hspace*{0.202em}\hspace*{0.202em}{Y}^{\alpha}:=\frac{|\sin (x){|}^{\alpha}}{\cos (x)}{\partial}_{y}. \end{equation*}$$ The α $\alpha$ -Grushin hemisphere S ¯ α + $\overline{\mathbb {S}}^+_\alpha$ (resp., open hemisphere S α + $\mathbb {S}^+_\alpha$ ) is the subset of S α $\mathbb {S}_\alpha$ defined byS ¯ α + : = { p ∈ S 2 ∣ x ∈ [ 0 , π / 2 ] } (resp., S α + : = { p ∈ S 2 ∣ x ∈ ( 0 , π / 2 ] } ) . $$\begin{equation*} \overline{\mathbb {S}}^+_\alpha:= \lbrace p \in \mathbb {S}^2 \mid x \in [0,\pi /2] \rbrace \text{ (resp., $\mathbb {S}^+_\alpha:= \lbrace p \in \mathbb {S}^2 \mid x \in (0,\pi /2] \rbrace)$}. \end{equation*}$$
3.3
Remark
When α $\alpha$ is noninteger, the vector fields are not smooth and they are not bracket-generating. However, any pair of points can still be joined with a horizontal curve and Pontryagin's Maximum Principle stated in Theorem 2.5 can be applied. Therefore, the Carnot–Carathéodory metric Equation (4) can be constructed as in smooth sub-Riemannian geometry. This remark remains valid for the other model spaces introduced in this work.
3.4
Remark
Note that, strictly speaking, the structure introduced in Definition 3.2 does not fall into the definition of sub-Riemannian structure laid out in Subsection 2.2. Indeed, the vector fields X $X$ and Y α $Y^\alpha$ are not global vector fields: they are not defined at the poles, that is, at x = ± π / 2 $x=\pm \pi /2$ . To be completely rigorous, one should therefore check that it is still a sub-Riemannian manifold but according to the general definition of [1, Definition 3.2]. There, a sub-Riemannian manifold is given by the pair ( E , f ) $(E, f)$ where E $E$ is a Euclidean vector bundle and f : E → T M $f: E \rightarrow T M$ is a morphism of vector bundles. In our setting, E = T S 2 $E = T\mathbb {S}^2$ and f $f$ is a morphism that we now construct explicitly.
Denote by p : T S 2 → S 2 $p:T\mathbb {S}^2\rightarrow \mathbb {S}^2$ the bundle projection. Letting U 1 : = S 2 ∖ { N , S } $\mathcal {U}_1:= \mathbb {S}^2 \setminus \lbrace N, S\rbrace$ , the coordinate chart φ 1 : = ( x , y ) : U 1 → R 2 $\varphi _1:= (x, y): \mathcal {U}_1 \rightarrow \mathbb {R}^2$ induces a chart on T S 2 $T\mathbb {S}^2$ , and we define f 1 : p − 1 ( U 1 ) → T S 2 $f_{1}: p^{-1}(\mathcal {U}_1) \rightarrow T \mathbb {S}^2$ as the map, linear on fibers, that satisfiesf 1 ( ∂ x ) = ∂ x , f 1 ( ∂ y ) = | sin ( x ) | α ∂ y . $$\begin{equation*} {f}_{1}({\partial}_{x})={\partial}_{x},\hspace*{2em}{f}_{1}({\partial}_{y})=|\sin (x){|}^{\alpha}{\partial}_{y}. \end{equation*}$$ Letting U 2 : = { ( s , t , z ) ∈ S 2 ⊂ R 3 ∣ z > 0 } $\mathcal {U}_2:= \lbrace (s, t, z) \in \mathbb {S}^2\subset \mathbb {R}^3 \mid z > 0 \rbrace$ , the mapφ 2 : U 2 → R 2 : ( s , t , 1 − s 2 − t 2 ) ↦ ( s , t ) $$\begin{equation*} \varphi _2: \mathcal {U}_2 \rightarrow \mathbb {R}^2: (s, t, \sqrt {1 - s^2 - t^2}) \mapsto (s, t) \end{equation*}$$ is another coordinate chart on S 2 $\mathbb {S}^2$ , which also induces a chart on T S 2 $T \mathbb {S}^2$ . We set f 2 : p − 1 ( U 2 ) → T S 2 $f_{2}: p^{-1}(\mathcal {U}_2) \rightarrow T \mathbb {S}^2$ the map, linear on fibers, satisfyingf 2 ( ∂ s ) = 1 s 2 + t 2 s 2 + ( 1 − s 2 − t 2 ) α / 2 t 2 ∂ s + s t ( 1 − ( 1 − s 2 − t 2 ) α / 2 ) ∂ t $$\begin{equation*} f_2(\partial _s) = \frac{1}{s^2+t^2} {\left[ {\left(s^2+(1-s^2-t^2)^{\alpha /2}t^2\right)} \partial _s + st(1-(1-s^2-t^2)^{\alpha /2}) \partial _t \right]} \end{equation*}$$ andf 2 ( ∂ t ) = 1 s 2 + t 2 s t ( 1 − ( 1 − s 2 − t 2 ) α / 2 ) ∂ s + ( 1 − s 2 − t 2 ) α / 2 s 2 + t 2 ∂ t . $$\begin{equation*} f_2(\partial _t) = \frac{1}{s^2+t^2} {\left[ st(1-(1-s^2-t^2)^{\alpha /2}) \partial _s + {\left((1-s^2-t^2)^{\alpha /2}s^2 + t^2 \right)} \partial _t \right]}. \end{equation*}$$ This ensures f 1 = f 2 $f_1=f_2$ on p − 1 ( U 1 ∩ U 2 ) $p^{-1}(\mathcal {U}_1\cap \mathcal {U}_2)$ via the coordinate transformation s = cos ( x ) cos ( y ) $s=\cos (x)\cos (y)$ , t = cos ( x ) sin ( y ) $t=\cos (x)\sin (y)$ . The apparent singularity at ( s , t ) = ( 0 , 0 ) $(s,t)=(0,0)$ is removable becausef 2 ( ∂ s ) → ∂ s , f 2 ( ∂ t ) → ∂ t as ( s , t ) → ( 0 , 0 ) , $$\begin{equation*} f_2(\partial _s)\rightarrow \partial _s, \quad f_2(\partial _t)\rightarrow \partial _t \quad \text{as }(s,t)\rightarrow (0,0), \end{equation*}$$ so f 2 $f_2$ extends smoothly over the origin, giving the identity on the fiber at N $N$ . The maps f 1 $f_1$ and f 2 $f_2$ are patched together to define a smooth, globally defined bundle morphism f : E → T S 2 $f\colon E\rightarrow T\mathbb {S}^2$ , and the pair ( E , f ) $(E,f)$ defines a sub-Riemannian manifold in the sense of [1, Definition 3.2].
The 0-Grushin sphere is simply the two-dimensional Riemannian sphere S 2 $\mathbb {S}^2$ . When α > 0 $\alpha > 0$ , the α $\alpha$ -Grushin sphere is a two-dimensional almost-Riemannian structure with { x = 0 } $\lbrace x = 0\rbrace$ being its set of singular points. The Grushin sphere studied in [11, 34] corresponds to the 1-Grushin sphere. At nonsingular points, this sub-Riemannian structure admits the Riemannian metric
5
g S α = d x ⊗ d x + cos 2 ( x ) sin 2 α ( x ) d y ⊗ d y . $$\begin{equation} g_{\mathbb {S}_{\alpha }} = \mathop {}\!\mathrm{d}x \otimes \mathop {}\!\mathrm{d}x + \frac{\cos ^2(x)}{\sin ^{2\alpha }(x)} \mathop {}\!\mathrm{d}y \otimes \mathop {}\!\mathrm{d}y. \end{equation}$$ A simple computation shows that the Riemannian volume induced from Equation 5, is given by
dvol S α = cos ( x ) | sin ( x ) | − α d x d y . $$\begin{equation*} \text{dvol}_{{\mathbb{S}}_{\alpha}}=\cos (x)|\sin (x){|}^{-\alpha} \mathrm{d}x \mathrm{d}y. \end{equation*}$$ We introduce the following weighted measure.
3.5
Definition
For β ⩾ α $\beta \geqslant \alpha$ , we consider the weighted measure given bym S α β : = | sin ( x ) | β dvol S α = cos ( x ) | sin ( x ) | β − α d x d y = e − V S α dvol S α , $$\begin{equation*} {\mathfrak{m}}_{{\mathbb{S}}_{\alpha}}^{\beta}:=|\sin (x){|}^{\beta}\text{dvol}_{{\mathbb{S}}_{\alpha}}=\cos (x)|\sin (x){|}^{\beta -\alpha} \mathrm{d}x \mathrm{d}y={e}^{-{V}_{{\mathbb{S}}_{\alpha}}}\text{dvol}_{{\mathbb{S}}_{\alpha}}, \end{equation*}$$ where V S α ( x , y ) : = − β log | sin ( x ) | ${V}_{{\mathbb{S}}_{\alpha}}(x,y):=-\beta \log |\sin (x)|$ .
The α $\alpha$ -Grushin hemisphere is a geodesically convex subset of S α $\mathbb {S}_\alpha$ and a geodesic space when seen as a length subspace of S α $\mathbb {S}_\alpha$ . This is made clear by the next result.
3.6
Proposition
There is a minimizing geodesic contained within S ¯ α + $\overline{\mathbb {S}}^+_\alpha$ that joins any two given points in S ¯ α + $\overline{\mathbb {S}}^+_\alpha$ . Furthermore, the α $\alpha$ -Grushin open hemisphere S α + $\mathbb {S}_\alpha ^+$ is a geodesically convex subset of S ¯ α + $\overline{\mathbb {S}}^+_\alpha$ and has the structure of an incomplete weighted Riemannian manifold when equipped with the restriction of the measure m S α β $\mathfrak {m}^\beta _{\mathbb {S}_\alpha }$ .
Proof
We start by noting that between every two points on the α $\alpha$ -Grushin sphere, there is indeed a horizontal path controlled by the vector fields X $X$ and Y α $Y^{\alpha }$ joining them. This means that the induced sub-Riemannian distance d S α $\mathop {}\!\mathrm{d}_{\mathbb {S}_\alpha }$ is well-defined and that ( S α , d S α ) $(\mathbb {S}_\alpha, \mathop {}\!\mathrm{d}_{\mathbb {S}_\alpha })$ is a locally compact metric space. Even though the bracket generating condition is not verified when α ∉ N $\alpha \notin \mathbb {N}$ , it is easy to see that the metric topology still coincides with the original topology of S 2 $\mathbb {S}^2$ by the monotonicity property of d S α $\mathrm{d}_{\mathbb {S}_\alpha }$ with respect to α ⩾ 0 $\alpha \geqslant 0$ . Furthermore, any metric ball is compact and thus the metric space ( S α , d S α ) $(\mathbb {S}_\alpha, \mathop {}\!\mathrm{d}_{\mathbb {S}_\alpha })$ is complete.
The sub-Riemannian Hamiltonian H : T ∗ ( S α ) → R $H: T^*(\mathbb {S}_\alpha) \rightarrow \mathbb {R}$ can be written in the canonical coordinates ( x , y , u , v ) $(x, y, u,v)$ induced from ( x , y ) $(x, y)$ asH ( λ ) : = 1 2 ⟨ λ , X ⟩ 2 + ⟨ λ , Y α ⟩ 2 = 1 2 u 2 + sin 2 α ( x ) cos 2 ( x ) v 2 . $$\begin{equation*} H(\lambda):=\frac{1}{2}{\left[\langle \lambda,X\rangle ^2+\langle \lambda,Y^\alpha \rangle ^2\right]}=\frac{1}{2}{\left[u^2+\frac{\sin ^{2\alpha }(x)}{\cos ^2(x)}v^2\right]}. \end{equation*}$$ A normal extremal λ : 0 , T → T ∗ ( S α ) : t ↦ ( x ( t ) , y ( t ) , u ( t ) , v ( t ) ) $\lambda: \left[0, T\right] \rightarrow T^*(\mathbb {S}_\alpha) : t \mapsto (x(t), y(t), u(t), v(t))$ satisfies the following Hamiltonian system of equations6
x ̇ = u , y ̇ = sin 2 α ( x ) cos 2 ( x ) v , u ̇ = − v 2 sin 2 ( α − 1 ) ( x ) tan ( x ) ( α + tan 2 ( x ) ) , v ̇ = 0 . $$\begin{equation} {\begin{cases} \dot{x}=u,\\ \displaystyle \dot{y}=\frac{\sin ^{2\alpha }(x)}{\cos ^2(x)}v,\\ \displaystyle \dot{u}= - v^2 \ \sin ^{2(\alpha - 1)}(x) \tan (x) (\alpha + \tan ^2(x)),\\ \dot{v}=0. \end{cases}} \end{equation}$$
Here note that the extremal reaches to the undefined point x = ± π / 2 $x=\pm \pi /2$ only if v ≡ 0 $v\equiv 0$ (this follows from the nonintegrability of tan ( x ) $\tan (x)$ near x = ± π / 2 $x=\pm \pi /2$ ). In this case, a (Euclidean) large circle passing through the north pole becomes a length minimizing geodesic. By completeness, there is a sub-Riemannian geodesic between every two points of S α $\mathbb {S}_\alpha$ by [12, Theorem 2.5.23]. These are obtained from Hamilton's equation Equation 6 because there are no nontrivial abnormal geodesics.
If a horizontal path of S α $\mathbb {S}_\alpha$ is contained in both S ¯ α + $\overline{\mathbb {S}}^+_\alpha$ and S ¯ α − : = { p ∈ S 2 ∣ x ∈ [ − π / 2 , 0 ] } $\overline{\mathbb {S}}^-_\alpha:= \lbrace p \in \mathbb {S}^2 \mid x \in [-\pi /2,0] \rbrace$ , then a reflection ( x , y ) ↦ ( − x , y ) $(x, y) \mapsto (-x,y)$ of the part of path that is in S ¯ α − $\overline{\mathbb {S}}^-_\alpha$ produces a curve contained in S ¯ α + $\overline{\mathbb {S}}^+_\alpha$ with the same length. This shows that a geodesic between points in the α $\alpha$ -Grushin hemisphere S ¯ α + $\overline{\mathbb {S}}^+_\alpha$ is contained within S ¯ α + $\overline{\mathbb {S}}^+_\alpha$ . Length-minimizers are smooth because they satisfy Hamilton's equation Equation 6. Thus, a constant-speed minimizing geodesic γ ( t ) = ( x ( t ) , y ( t ) ) $\gamma (t) = (x(t), y(t))$ that touches the singular equator at a point other than its endpoints must do so tangentially, and Equation 6 implies that x ( t ) $x(t)$ vanishes for all t $t$ . Consequently, y ( t ) $y(t)$ also vanishes, and γ $\gamma$ becomes a constant curve. In particular, a minimizing geodesic between points of S α + $\mathbb {S}_\alpha ^+$ is also contained within S α + $\mathbb {S}_\alpha ^+$ .
The fact that S α + $\mathbb {S}_\alpha ^+$ is also an incomplete Riemannian manifold follows easily because it does not contain any singular points of S α $\mathbb {S}_\alpha$ .□ $\Box$
The α $\alpha$ -Grushin hyperbolic plane and half-plane
On the two-dimensional hyperbolic plane ( H 2 , g H 2 ) $(\mathbb {H}^2, g_{\mathbb {H}^2})$ , we consider the following coordinate chart, called Lobachevsky's coordinates (see [29, section 33.1], for example). We fix an infinite minimizing geodesic ray γ : R → H 2 $\gamma: \mathbb {R}\rightarrow \mathbb {H}^2$ , and for p ∈ H 2 $p \in \mathbb {H}^2$ , we let x : = x ( p ) ∈ R $x:= x(p) \in \mathbb {R}$ be signed hyperbolic distance d H 2 ( p , Im ( γ ) ) $\mathop {}\!\mathrm{d}_{\mathbb {H}^2}(p, \mathrm{Im}(\gamma))$ , where the signature is positive (resp., negative) if p $p$ belongs to the left-hand side (resp., right-hand side) of γ $\gamma$ . Furthermore, let y = y ( p ) ∈ R $y=y(p)\in \mathbb {R}$ be the unique number such that γ ( y ) $\gamma (y)$ is the perpendicular foot from p $p$ to Im ( γ ) $\mathrm{Im}(\gamma)$ . The map H 2 → R × R : p ↦ ( x ( p ) , y ( p ) ) $\mathbb {H}^2 \rightarrow \mathbb {R}\times \mathbb {R}: p \mapsto (x(p), y(p))$ defines a global coordinate chart, and a short computation shows that the hyperbolic Riemannian metric g H 2 $g_{\mathbb {H}^2}$ has the warped product structure
g H 2 = d x ⊗ d x + cosh 2 ( x ) d y ⊗ d y . $$\begin{equation*} g_{\mathbb {H}^2} = \mathop {}\!\mathrm{d}x\otimes \mathop {}\!\mathrm{d}x+\cosh ^2(x)\mathop {}\!\mathrm{d}y\otimes \mathop {}\!\mathrm{d}y. \end{equation*}$$
3.7
Definition
For α ⩾ 0 $\alpha \geqslant 0$ , the α $\alpha$ -Grushin hyperbolic plane H α $\mathbb {H}_\alpha$ is the sub-Riemannian structure on H 2 $\mathbb {H}^2$ induced from the vector field X $X$ and Y α $Y^\alpha$ given byX : = ∂ x , Y α : = | sinh ( x ) | α cosh ( x ) ∂ y . $$\begin{equation*} X:={\partial}_{x},\hspace*{0.202em}\hspace*{0.202em}{Y}^{\alpha}:=\frac{|\mathrm{\sinh}(x){|}^{\alpha}}{\cosh (x)}{\partial}_{y}. \end{equation*}$$ The α $\alpha$ -Grushin hyperbolic half-plane H ¯ α + $\overline{\mathbb {H}}^+_\alpha$ (resp., open half-plane H α + $\mathbb {H}^+_\alpha$ ) is the subset of H α $\mathbb {H}_\alpha$ defined byH ¯ α + : = { p ∈ H 2 ∣ x ⩾ 0 } (resp., H α + : = { p ∈ H 2 ∣ x > 0 } ) . $$\begin{equation*} \overline{\mathbb {H}}^+_\alpha:= \lbrace p \in \mathbb {H}^2 \mid x \geqslant 0 \rbrace \text{ (resp., $\mathbb {H}^+_\alpha:= \lbrace p \in \mathbb {H}^2 \mid x > 0 \rbrace $)}. \end{equation*}$$
The 0-Grushin hyperbolic plane is simply the two-dimensional Riemannian hyperbolic plane H 2 $\mathbb {H}^2$ . When α > 0 $\alpha > 0$ , the α $\alpha$ -Grushin hyperbolic plane is a two-dimensional almost-Riemannian structure with { x = 0 } $\lbrace x = 0\rbrace$ being its set of singular points. To the best of our knowledge, this definition, although very natural, is new. At nonsingular points, this sub-Riemannian structure admits the Riemannian metric
7
g H α = d x ⊗ d x + cosh 2 ( x ) sinh 2 α ( x ) d y ⊗ d y . $$\begin{equation} g_{\mathbb {H}_{\alpha }} = \mathop {}\!\mathrm{d}x \otimes \mathop {}\!\mathrm{d}x + \frac{\cosh ^2(x)}{\sinh ^{2\alpha }(x)} \mathop {}\!\mathrm{d}y \otimes \mathop {}\!\mathrm{d}y. \end{equation}$$ A simple computation shows that the Riemannian volume induced from Equation 7 is given by
dvol H α = cosh ( x ) | sinh ( x ) | − α d x d y . $$\begin{equation*} \text{dvol}_{{\mathbb{H}}_{\alpha}}=\cosh (x)|\mathrm{\sinh}(x){|}^{-\alpha} \mathrm{d}x \mathrm{d}y. \end{equation*}$$ We introduce the following weighted measure.
3.8
Definition
For β ⩾ α $\beta \geqslant \alpha$ , we consider the weighted measure given by
m H α β : = | sinh ( x ) | β dvol H α = cosh ( x ) | sinh ( x ) | β − α d x d y = e − V H α dvol H α , $$\begin{equation*} {\mathfrak{m}}_{{\mathbb{H}}_{\alpha}}^{\beta}:=|\mathrm{\sinh}(x){|}^{\beta}\text{dvol}_{{\mathbb{H}}_{\alpha}}=\cosh (x)|\mathrm{\sinh}(x){|}^{\beta -\alpha} \mathrm{d}x \mathrm{d}y={e}^{-{V}_{{\mathbb{H}}_{\alpha}}}\text{dvol}_{{\mathbb{H}}_{\alpha}}, \end{equation*}$$ where V H α ( x , y ) : = − β log | sinh ( x ) | ${V}_{{\mathbb{H}}_{\alpha}}(x,y):=-\beta \log |\mathrm{\sinh}(x)|$ .
As for the previous section, the α $\alpha$ -Grushin hyperbolic half-plane is a geodesically convex subset of H α $\mathbb {H}_\alpha$ and is a geodesic space when seen as a length subspace of H α $\mathbb {H}_\alpha$ . The open half-plane H α + $\mathbb {H}_\alpha ^+$ is also a geodesically convex subset and it is an incomplete Riemannian manifold because it does not contain any singular points of H α $\mathbb {H}_\alpha$ .
3.9
Proposition
There is a minimizing geodesic contained within H ¯ α + $\overline{\mathbb {H}}^+_\alpha$ that joins any two given points in H ¯ α + $\overline{\mathbb {H}}^+_\alpha$ . Furthermore, The α $\alpha$ -Grushin hyperbolic open half-plane H α + $\mathbb {H}_\alpha ^+$ is a geodesically convex subset of H ¯ α + $\overline{\mathbb {H}}_\alpha ^+$ and has the structure of an incomplete weighted Riemannian manifold when equipped with the restriction of the measure m H α β $\mathfrak {m}^\beta _{\mathbb {H}_\alpha }$ .
Proof
Here, the Hamiltonian and the corresponding Hamilton's equation are given in Lobachevsky's coordinates, byH ( λ ) = 1 2 u 2 + sinh 2 α ( x ) cosh 2 ( x ) v 2 , and x ̇ = u , y ̇ = sinh 2 α ( x ) cosh 2 ( x ) v , u ̇ = − 2 v 2 sinh 2 ( α − 1 ) ( x ) tanh ( x ) ( α − tanh 2 ( x ) ) , v ̇ = 0 . $$\begin{equation*} H(\lambda)=\frac{1}{2}{\left[u^2+\frac{\sinh ^{2\alpha }(x)}{\cosh ^2(x)}v^2\right]}, \text{ and } {\begin{cases} \dot{x}=u,\\ \displaystyle \dot{y}=\frac{\sinh ^{2\alpha }(x)}{\cosh ^2(x)}v,\\ \displaystyle \dot{u}= - 2 v^2 \ \sinh ^{2(\alpha - 1)}(x) \tanh (x) (\alpha - \tanh ^2(x)),\\ \dot{v}=0. \end{cases}} \end{equation*}$$ The rest of the proof follows exactly the arguments of Proposition 3.6.□ $\Box$
The ∞ $\infty$ -Grushin plane and half-plane
The geometry of the so-called α $\alpha$ -Grushin plane, where α ⩾ 0 $\alpha \geqslant 0$ , has been studied in [6, 13], and [7]. The α $\alpha$ -Grushin half-plane and the validity of the CD $\mathsf {CD}$ condition in this space is studied in [36], and we recalled some details in Section 1. Instead, we introduce a model of a Grushin plane with infinite Hausdorff dimension. The global chart ( x , y ) $(x, y)$ simply denotes the cartesian coordinates in this section.
3.10
Definition
The ∞ $\infty$ -Grushin plane G ∞ $\mathbb {G}_\infty$ is the sub-Riemannian structure on R 2 $\mathbb {R}^2$ induced from the vector field X $X$ and Y $Y$ given byX : = ∂ x , Y : = e − 1 / | x | ∂ y . $$\begin{equation*} X:={\partial}_{x},\hspace*{0.202em}\hspace*{0.202em}Y:={e}^{-1/|x|}{\partial}_{y}. \end{equation*}$$ The ∞ $\infty$ -Grushin half-plane G ¯ ∞ + $\overline{\mathbb {G}}^+_\infty$ (resp., open half-plane G ∞ + $\mathbb {G}^+_\infty$ ) is the subset of G ∞ $\mathbb {G}_\infty$ defined byG ¯ ∞ + : = { p ∈ R 2 ∣ x ⩾ 0 } (resp., G ∞ + : = { p ∈ R 2 ∣ x > 0 } ) . $$\begin{equation*} \overline{\mathbb {G}}^+_\infty:= \lbrace p \in \mathbb {R}^2 \mid x \geqslant 0 \rbrace \text{ (resp., $\mathbb {G}^+_\infty:= \lbrace p \in \mathbb {R}^2 \mid x > 0 \rbrace $)}. \end{equation*}$$
3.11
Lemma
The ∞ $\infty$ -Grushin plane and half-plane have infinite Hausdorff dimension.
Proof
We will show that the Hausdorff dimension of S : = { ( 0 , y ) ∣ y ∈ R } ⊆ G ∞ $\mathcal {S}:=\lbrace (0,y)\mid y\in \mathbb {R}\rbrace \subseteq \mathbb {G}_\infty$ is + ∞ $+\infty$ . Let us denote by d α $\mathsf {d}_\alpha$ (resp., d ∞ $\mathsf {d}_\infty$ ) the induced distance on G α $\mathbb {G}_\alpha$ (resp., G ∞ $\mathbb {G}_\infty$ ). It is well-known that the Hausdorff dimension of S ⊆ G α $\mathcal {S}\subseteq \mathbb {G}_\alpha$ is α + 1 $\alpha +1$ , see, for example, [17]. As the inequality | x | α ⩾ e − 1 / | x | $|x|^\alpha \geqslant e^{-1/|x|}$ holds for sufficiently small | x | $|x|$ , we have the inequality d α ⩾ d ∞ $\mathsf {d}_\alpha \geqslant \mathsf {d}_\infty$ in a small neighbourhood of an arbitrary point in S $\mathcal {S}$ . This implies that dim H ( S , d α ) ⩽ dim H ( S , d ∞ ) $\dim _H(\mathcal {S},\mathsf {d}_\alpha)\leqslant \dim _H(\mathcal {S},\mathsf {d}_\infty)$ and concludes the lemma.□ $\Box$
The ∞ $\infty$ -Grushin plane is a two-dimensional almost-Riemannian structure with { x = 0 } $\lbrace x = 0\rbrace$ being its set of singular points. To the best of our knowledge, this definition of ∞ $\infty$ -Grushin plane is also new. At nonsingular points, this sub-Riemannian structure admits the Riemannian metric
8
g G ∞ = d x ⊗ d x + e 2 / | x | d y ⊗ d y . $$\begin{equation} {g}_{{\mathbb{G}}_{\infty}}= \mathrm{d}x\otimes \mathrm{d}x+{e}^{2/|x|} \mathrm{d}y\otimes \mathrm{d}y. \end{equation}$$ A simple computation shows that the Riemannian volume induced from Equation 8 is given by
dvol G ∞ = e 1 / | x | d x d y . $$\begin{equation*} \text{dvol}_{{\mathbb{G}}_{\infty}}={e}^{1/|x|} \mathrm{d}x \mathrm{d}y. \end{equation*}$$ We introduce the following weighted measure.
3.12
Definition
For β ⩾ 0 $\beta \geqslant 0$ and γ > 0 $\gamma >0$ , we consider the weighted (Radon) measure given bym G ∞ β , γ : = | x | β e − γ / x 2 dvol G ∞ = | x | β e − γ / x 2 + 1 / | x | d x d y = e − V dvol G ∞ , $$\begin{equation*} {\mathfrak{m}}_{{\mathbb{G}}_{\infty}}^{\beta ,\gamma}:=|x{|}^{\beta}{e}^{-\gamma /{x}^{2}}\text{dvol}_{{\mathbb{G}}_{\infty}}=|x{|}^{\beta}{e}^{-\gamma /{x}^{2}+1/|x|} \mathrm{d}x \mathrm{d}y={e}^{-V}\text{dvol}_{{\mathbb{G}}_{\infty}}, \end{equation*}$$ where V G ∞ ( x , y ) : = γ x 2 − β log | x | ${V}_{{\mathbb{G}}_{\infty}}(x,y):=\frac{\gamma}{{x}^{2}}-\beta \log |x|$ .
The next result is analogous to the corresponding one in the previous two sections. The ∞ $\infty$ -Grushin half-plane is a geodesically convex subset of G ∞ $\mathbb {G}_\infty$ and a geodesic space when seen as a length subspace of G ∞ $\mathbb {G}_\infty$ . Similarly, the open half-plane G ∞ + $\mathbb {G}_\infty ^+$ is also a geodesically convex subset and it is an incomplete Riemannian manifold because it does not contain any singular points of G ∞ $\mathbb {G}_\infty$ .
3.13
Proposition
There is a minimizing geodesic contained within G ¯ ∞ + $\overline{\mathbb {G}}^+_\infty$ that joins any two given points in G ¯ ∞ + $\overline{\mathbb {G}}^+_\infty$ . Furthermore, the ∞ $\infty$ -Grushin open half-plane G ∞ + $\mathbb {G}_\infty ^+$ is a geodesically convex subset of G ¯ ∞ + $\overline{\mathbb {G}}_\infty ^+$ and has the structure of an incomplete weighted Riemannian manifold when equipped with the restriction of the measure m G ∞ β , γ $\mathfrak {m}^{\beta, \gamma }_{\mathbb {G}_\infty }$ .
Proof
The argument is again analogous to the proofs of Proposition 3.6 and Proposition 3.9, with the Hamiltonian and Hamilton's equation given in cartesian coordinates by
H ( λ ) = 1 2 u + e − 2 / | x | v 2 , and x ̇ = u , y ̇ = e − 2 / | x | v , u ̇ = − v 2 e − 2 / | x | | x | 2 x , v ̇ = 0 . $$\begin{equation*} H(\lambda)=\frac{1}{2}{\left[u+e^{-2/\vert x \vert}v^2\right]}, \text{ and } {\begin{cases} \dot{x}=u,\\ \displaystyle \dot{y}=e^{-2/\vert x \vert }v,\\ \displaystyle \dot{u}= - v^2 \frac{e^{-2/\vert x \vert }}{{\vert x \vert}^2} x,\\ \dot{v}=0. \end{cases}} \end{equation*}$$ □ $\Box$
Hereafter, we will collectively refer to the α $\alpha$ -Grushin plane (resp., half-plane), the ∞ $\infty$ -Grushin plane (resp., half-plane), the α $\alpha$ -Grushin sphere (resp., hemisphere), and the α $\alpha$ -Grushin hyperbolic plane (resp., half-plane) as the α $\alpha$ -Grushin spaces (resp., half-spaces).
EQUIVALENCE BETWEEN CD ( K , N ) $\mathsf {CD}(K,N)$ AND Ric N ⩾ K $\mathrm{Ric}_N\geqslant K$ FOR SOME ALMOST RIEMANNIAN MANIFOLDS
The α $\alpha$ -Grushin spaces and half-spaces introduced in Section 3 are metric measure spaces ( X , d , m ) $(\mathsf {X}, \mathsf {d}, \mathfrak {m})$ with a weighted Riemannian manifold ( M , d g , e − V dvol g ) $(M, \mathop {}\!\mathrm{d}_g, e^{-V}\mathrm{dvol}_g)$ as their interior. For N ∈ ( − ∞ , 0 ) ∪ [ n , + ∞ ] $N \in (-\infty,0)\cup [n,+\infty]$ , we recall that the N $N$ -Ricci tensor of an n $n$ -dimensional weighted Riemannian ( M , d g , e − V dvol g ) $(M, \mathop {}\!\mathrm{d}_g, e^{-V} \mathrm{dvol}_g)$ is defined by
9
Ric N , V : = Ric if N = n , Ric + Hess ( V ) if N = + ∞ , Ric + Hess ( V ) − d V ⊗ d V N − n otherwise , $$\begin{equation} \mathrm{Ric}_{N,V}:= {\begin{cases} \mathrm{Ric}& \text{ if } N=n, \\ \mathrm{Ric}+ \mathrm{Hess}(V) & \text{ if } N=+\infty,\\ \mathrm{Ric}+\mathrm{Hess}(V)-\dfrac{\mathop {}\!\mathrm{d}V\otimes \mathop {}\!\mathrm{d}V}{N-n} & \text{ otherwise}, \end{cases}} \end{equation}$$ with the convention that V $V$ must be constant when N = n $N=n$ .
The following theorem provides sufficient conditions under which the differential condition Ric N , V ⩾ K $\mathrm{Ric}_{N,V} \geqslant K$ on M $M$ is equivalent to ( X , d , m ) $(\mathsf {X}, \mathsf {d}, \mathfrak {m})$ satisfying the CD ( K , N ) $\mathsf {CD}(K, N)$ condition. The proof generalizes the sketch found in [36, section 3.5], which is specific to the α $\alpha$ -Grushin half-plane. This theorem is, to some extent, related to the conjecture stated in [20].
4.1
Theorem
Let K ∈ R $K \in \mathbb {R}$ , N ∈ ( − ∞ , 0 ) ∪ [ n , + ∞ ] $N\in (-\infty ,0)\cup [n,+\infty ]$ , ( X , d , m ) $(\mathsf {X}, \mathsf {d}, \mathfrak {m})$ be a metric measure space and M $M$ be an open subset of X $\mathsf {X}$ such that
(i) M $M$ is a geodesically convex subset of ( X , d ) $(\mathsf {X}, \mathsf {d})$ , that is, for every x , y ∈ M $x,y\in M$ , there is a geodesic joining x $x$ and y $y$ and any such curve is contained in M $M$ ,
(ii) ( M , d | M , m | M ) $(M, \mathsf {d}|_M, \mathfrak {m}|_M)$ possesses a weighted n $n$ -dimensional Riemannian manifold structure,
(iii) m ( X ∖ M ) = 0 $\mathfrak {m}(\mathsf {X}\setminus M) = 0$ . Then, the metric measure space ( X , d , m ) $(\mathsf {X}, \mathsf {d}, \mathfrak {m})$ satisfies the CD ( K , N ) $\mathsf {CD}(K,N)$ condition if and only if Ric N , V ⩾ K $\mathrm{Ric}_{N,V} \geqslant K$ on M $M$ .
Proof
We provide the proof only for N ∈ n , + ∞ $N \in \left[n, +\infty \right]$ . The details for N ∈ − ∞ , 0 $N \in \left(-\infty, 0\right)$ are exactly the same, but one needs to replace the relevant key results used in the proof by the analogous ones when N $N$ is negative. In particular, the Gromov–Hausdorff convergence which we use below must be replaced by the pointed iKRW $\mathsf {iKRW}$ convergence described in [26].
First, we note that the CD ( K , N ) $\mathsf {CD}(K,N)$ condition on ( X , d , m ) $({\rm X},{\rm d},{\rm m})$ directly implies that Ric N , V ⩾ K $\mathrm{Ric}_{N,V} \geqslant K$ on M $M$ by [39, Theorem 17.36] (see also [38, Part (e) of the proof of Theorem 1.7]). We therefore focus now on the other implication.
For a closed metric subspace A ⊆ X $A\subseteq \mathsf {X}$ , we shall denote by W 2 A $W_2^A$ the Wasserstein distance on A $A$ . Consider μ 0 , μ 1 ∈ P 2 ac ( X , d , m ) $\mu _0, \mu _1 \in \mathcal {P}_2^{\mathrm{ac}}(\mathsf {X}, \mathsf {d}, \mathfrak {m})$ with continuous densities ρ 0 $\rho _0$ and ρ 1 $\rho _1$ , respectively. We need to argue that there exists a W 2 X $W_2^\mathsf {X}$ -geodesic ( μ s ) s ∈ 0 , 1 $(\mu _s)_{s \in \left[0, 1\right]}$ joining μ 0 $\mu _0$ to μ 1 $\mu _1$ with μ s = ( e s ) ♯ ν $\mu _s = (e_s)_\sharp \nu$ for some ν ∈ P ( Geo ( X , d ) ) $\nu \in \mathcal {P}(\mathrm{Geo}(\mathsf {X}, \mathsf {d}))$ such that the inequality in Definition 2.1 is satisfied. For all k ∈ N $k \in \mathbb {N}$ and given an arbitrary x 0 ∈ X $x_0 \in \mathsf {X}$ , we define the setsM k : = cl ( geo ( N k ) ) , and N k : = { x ∈ M ∣ k ⩾ d ( x 0 , x ) , and d ( x , X ∖ M ) ⩾ 1 / k } , $$\begin{equation*} M_k:=\mathrm{cl}(\mathrm{geo}(N_k)), \text{ and } \ \ N_k:= \lbrace x\in M\mid k \geqslant \mathsf {d}(x_0,x), \text{ and }\mathsf {d}(x,\mathsf {X}\setminus M)\geqslant 1/k\rbrace, \end{equation*}$$ where geo ( A ) $\mathrm{geo}(A)$ is the geodesic hull of a subset A ⊆ X $A \subseteq \mathsf {X}$ , that is, the union of all geodesics starting at x ∈ A $x \in A$ and ending at y ∈ A $y \in A$ . The set N k $N_k$ is clearly closed and bounded. As ( X , d , m ) $(\mathsf {X}, \mathop {}\!\mathrm{d}, \mathfrak {m})$ is a complete metric measure space, the Heine–Borel property (see [12, Theorem 2.5.28]) implies that N k $N_k$ is also compact in X $\mathsf {X}$ .
Step 1. The closure of the geodesic hull of a compact subset of M $M$ is in M $M$
Let K $K$ be a compact set of X $\mathsf {X}$ contained in M $M$ . We claim that cl ( geo ( K ) ) $\mathrm{cl}(\mathrm{geo}(K))$ is contained within M $M$ . As M $M$ is assumed to be geodesically convex in X $\mathsf {X}$ , we note that their closure in X $\mathsf {X}$ satisfies cl ( geo ( K ) ) ⊆ cl ( M ) $\mathrm{cl}(\mathrm{geo}(K)) \subseteq \mathrm{cl}(M)$ . Because M $M$ is open in X $\mathsf {X}$ , we also have that d ( x , ∂ M ) ⩾ d ( K , ∂ M ) > 0 $\mathop {}\!\mathrm{d}(x, \partial M) \geqslant \mathop {}\!\mathrm{d}(K, \partial M) > 0$ for all x ∈ K $x \in K$ . Suppose by contradiction that there exists p ∈ cl ( geo ( K ) ) $p \in \mathrm{cl}(\mathrm{geo}(K))$ such that p ∈ ∂ M $p \in \partial M$ . This means that for all n ∈ N $n \in \mathbb {N}$ , there exist x n , y n ∈ K $x_n, y_n \in K$ , a geodesic γ n $\gamma _n$ joining x n $x_n$ to y n $y_n$ contained in M $M$ , and a point p n ∈ M $p_n \in M$ lying on γ n $\gamma _n$ such that p n → p $p_n \rightarrow p$ as n → + ∞ $n \rightarrow +\infty$ . By compactness, we have, after extracting a converging subsequence, that x n → x $x_n \rightarrow x$ and y n → y $y_n \rightarrow y$ for some x , y ∈ K $x, y \in K$ . In particular, the geodesics γ n $\gamma _n$ have uniformly bounded lengths. As ( X , d ) $(\mathsf {X}, \mathop {}\!\mathrm{d})$ is a complete locally compact length space, the Arzela–Ascoli theorem from [12, Theorem 2.5.14] implies that the sequence of curves γ n $\gamma _n$ contains a uniformly converging subsequence. This limit, which we denote by γ $\gamma$ , is a curve of X $\mathsf {X}$ with endpoints x $x$ and y $y$ and passing through p $p$ . Actually, the curve γ $\gamma$ is a minimizing geodesic between x $x$ and y $y$ by [12, Proposition 2.5.17]. By geodesic convexity of M $M$ assumed at (ii), the curve γ $\gamma$ is contained in M $M$ while the point p ∉ M $p \notin M$ , leading to a contradiction.
Step 2. The support of μ 0 $\mu _0$ and μ 1 $\mu _1$ are compact and contained in N k $N_k$
Assume that supp ( μ 0 ) $\operatorname{supp}(\mu _0)$ and supp ( μ 1 ) $\operatorname{supp}(\mu _1)$ are compact and that supp ( μ 0 ) ∪ supp ( μ 1 ) ⊆ N k $\operatorname{supp}(\mu _0) \cup \operatorname{supp}(\mu _1) \subseteq N_k$ for some k ∈ N $k \in \mathbb {N}$ . The previous step applied to the compact N k $N_k$ implies that each pair of points x ∈ supp ( μ 0 ) $x \in \operatorname{supp}(\mu _0)$ and y ∈ supp ( μ 1 ) $y \in \operatorname{supp}(\mu _1)$ are connected by a geodesic contained in M $M$ . By [30, Theorem 13], there exists a unique W 2 M k $W_2^{M_k}$ -geodesic ( μ s ) s ∈ 0 , 1 $(\mu _s)_{s \in \left[0, 1\right]}$ joining μ 0 $\mu _0$ to μ 1 $\mu _1$ . By [24, Proposition 2.10], we also know that there exists ν ∈ P ( Geo ( X , d ) ) $\nu \in \mathcal {P}(\mathrm{Geo}(\mathsf {X}, \mathsf {d}))$ such that μ s = ( e s ) ♯ ν $\mu _s = (e_s)_\sharp \nu$ for all s ∈ 0 , 1 $s \in \left[0, 1\right]$ and supp ( ν ) ⊆ Γ : = ( e 0 × e 1 ) − 1 ( supp ( μ 0 ) × supp ( μ 1 ) ) $\mathrm{supp}(\nu) \subseteq \Gamma:= (e_0 \times e_1)^{-1}(\mathrm{supp}(\mu _0) \times \mathrm{supp}(\mu _1))$ . A geodesic in γ ∈ Γ $\gamma \in \Gamma$ joins points in supp ( μ 0 ) $\mathrm{supp}(\mu _0)$ with points in supp ( μ 1 ) $\mathrm{supp}(\mu _1)$ and is contained in M $M$ thanks to Step 1. As dim ( M ) ⩽ N $\mathrm{dim}(M) \leqslant N$ and Ric V N ⩾ K $\mathrm{Ric}_{V}^N \geqslant K$ on ( M , g ) $(M, g)$ , [39, Theorem 17.36] (or also [38, Parts (a), (b), (c), and (d) of the proof of Theorem 1.7]) implies that the inequality in Definition 2.1 for all N ′ ⩾ N $N^{\prime } \geqslant N$ .
Step 3. The support of μ 0 $\mu _0$ and μ 1 $\mu _1$ are compact and contained in X $\mathsf {X}$
We can assume without loss of generality that X ∖ M ⊆ ∂ M $\mathsf {X}\setminus M \subseteq \partial M$ . Indeed, when it comes to the CD ( K , N ) $\mathsf {CD}(K, N)$ condition or the Gromov–Hausdorff convergence that we are going to discuss in this step, only the support of the measure matters (see [19, Remark 3.1]). The definition of N k $N_k$ and assumption (iii) imply that N k → X $N_k \rightarrow \mathsf {X}$ as k → + ∞ $k \rightarrow +\infty$ in the pointed measured Gromov–Hausdorff convergence (see [19, Definition 3.9]), when taking the inclusion map ι k : N k → X $\iota _k:N_k\rightarrow \mathsf {X}$ as approximation maps. Although N k $N_k$ is not a geodesically convex subspace, we have shown in Step 1 that M k : = cl ( geo ( N k ) ) ⊆ M $M_k:= \mathrm{cl}(\mathrm{geo}(N_k)) \subseteq M$ . Then, an argument similar to the proof of [39, Theorem 28.13] shows that the metric space ( P ( N k ) , W 2 N k ) $(\mathcal P(N_k),W_2^{N_k})$ converges to ( P ( X ) , W 2 X ) $(\mathcal P(\mathsf {X}),W_2^\mathsf {X})$ in the geodesic local Gromov–Hausdorff topology, via the inclusion map ( ι k ) # $(\iota _k)_\#$ . Hence, there are sequences of probability measures ( μ i k ) k ∈ N $(\mu _i^k)_{k\in \mathbb {N}}$ ( i = 0 , 1 ) $(i=0,1)$ , supported on N k $N_k$ , such that their pushforwards ( ι k ) # μ i k $(\iota _k)_\#\mu _i^k$ converges to μ i $\mu _i$ in ( P 2 ( X ) , W 2 X ) $(\mathcal P_2(\mathsf {X}),W_2^\mathsf {X})$ .
Let ( μ s k ) s ∈ 0 , 1 $(\mu _s^k)_{s \in \left[0, 1\right]}$ be the Wasserstein geodesic joining μ 0 k $\mu _0^k$ and μ 1 k $\mu _1^k$ and ν k $\nu _k$ be the optimal transport plan joining μ 0 k $\mu _0^k$ to μ 1 k $\mu _1^k$ in M k $M_k$ . By Step 3, the inequality in Definition 2.1 holds along ( μ s k ) s ∈ 0 , 1 $(\mu _s^k)_{s \in \left[0, 1\right]}$ for all N ′ ⩾ N $N^{\prime } \geqslant N$ . Furthermore, by [39, Theorem 28.9 and Exercise 28.15], ν k $\nu _k$ weakly converges to an optimal transport plan ν $\nu$ in X $\mathsf {X}$ joining μ 0 $\mu _0$ to μ 1 $\mu _1$ , and the W 2 M k $W_2^{M_k}$ -geodesics ( μ s k ) s ∈ 0 , 1 $(\mu _s^k)_{s \in \left[0, 1\right]}$ uniformly converges to a W 2 X $W_2^{\mathsf {X}}$ -geodesic ( μ s ) s ∈ 0 , 1 $(\mu _s)_{s \in \left[0, 1\right]}$ in X $\mathsf {X}$ joining μ 0 $\mu _0$ to μ 1 $\mu _1$ . By the same argument as in [39, Theorem 29.24 and 29.21], we conclude that the inequality in Definition 2.1 remains valid along ( μ s ) s ∈ 0 , 1 $(\mu _s)_{s \in \left[0, 1\right]}$ for all N ′ ⩾ N $N^{\prime } \geqslant N$ , by lower semicontinuity.
Step 4. The general case
The case where the support of μ 0 $\mu _0$ and μ 1 $\mu _1$ are not necessarily compact is obtained by the previous step and by using an exhaustion by compact sets, see [24, appendix E].□ $\Box$
As for the RCD ( K , N ) $\mathsf {RCD}(K, N)$ condition (for K ∈ R $K \in \mathbb {R}$ and N ∈ 1 , + ∞ $N \in \left[1, +\infty \right]$ ), it is enough for our purposes to use the result of [23, Theorem 1.2], which shows that sub-Riemannian manifolds equipped with a nonnegative Radon measure are infinitesimally Hilbertian. Although the results in [23] assume the bracket-generating condition and our α $\alpha$ -Grushin half-spaces do not when α ∉ N $\alpha \notin \mathbb {N}$ , their Finsler approximation techniques can still be applied. It is sufficient to regard the α $\alpha$ -Grushin half-spaces as α $\alpha$ -Grushin (full) spaces equipped with a measure supported on the half-space, for the following theorem to follow directly.
4.2
Theorem
For K ∈ R $K \in \mathbb {R}$ and N ∈ [ 1 , + ∞ ] $N\in [1,+\infty ]$ , the CD ( K , N ) $\mathsf {CD}(K,N)$ condition is equivalent to the RCD ( K , N ) $\mathsf {RCD}(K,N)$ condition for the α $\alpha$ -Grushin half-spaces.
GENERALIZED RICCI CURVATURE OF α $\alpha$ -GRUSHIN HALF-SPACES
The previous section shows that establishing the validity of the CD ( K , N ) $\mathsf {CD}(K, N)$ condition in the α $\alpha$ -Grushin half-spaces introduced in Section 3 is equivalent to a computation of the Bakry–Émery Ricci curvature. The following simple computation will be handy.
5.1
Lemma
Let ( M , g ) $(M, g)$ be a two-dimensional Riemannian manifold, U ⊆ M $\mathcal {U} \subseteq M$ an open set, and N ∈ ( − ∞ , 0 ) ∪ ( 2 , + ∞ ] $N \in (-\infty, 0) \cup (2, +\infty]$ . Assume that ( x , y ) : U → R 2 $(x, y): \mathcal {U} \rightarrow \mathbb {R}^2$ is a chart and that we are given two smooth functions f : U → R ∖ { 0 } $f: \mathcal {U} \rightarrow \mathbb {R}\setminus \lbrace 0\rbrace$ and V : U → R $V: \mathcal {U} \rightarrow \mathbb {R}$ that only depend on x $x$ . If, in this coordinate,g = d x ⊗ d x + 1 f ( x ) 2 d y ⊗ d y , $$\begin{equation*} g = \mathop {}\!\mathrm{d}x \otimes \mathop {}\!\mathrm{d}x + \frac{1}{f(x)^2} \mathop {}\!\mathrm{d}y \otimes \mathop {}\!\mathrm{d}y, \end{equation*}$$ then it holds that
Ric N , V = f ′ f ′ − f ′ f 2 + V ′ ′ − ( V ′ ) 2 N − 2 d x ⊗ d x + 1 f 2 f ′ f ′ − f ′ f 2 − f ′ f V ′ d y ⊗ d y , $$\begin{equation*} \begin{aligned} \mathrm{Ric}_{N,V} \,{=}\, {\left[ {\left(\frac{f^{\prime }}{f}\right)}^{\prime } \,{-}\, {\left(\frac{f^{\prime }}{f}\right)}^2 \,{+}\, V^{\prime \prime } \,{-}\, \frac{(V^{\prime })^2}{N \,{-}\, 2}\right]} \mathop {}\!\mathrm{d}x \otimes \mathop {}\!\mathrm{d}x \,{+}\, \frac{1}{f^2}{\left[{\left(\frac{f^{\prime }}{f}\right)}^{\prime } \,{-}\, {\left(\frac{f^{\prime }}{f}\right)}^2 \,{-}\, \frac{f^{\prime }}{f} V^{\prime }\right]} \mathop {}\!\mathrm{d}y \otimes \mathop {}\!\mathrm{d}y, \end{aligned} \end{equation*}$$ where all the derivatives are understood with respect to the variable x $x$ .
Proof
The two vector fields X = ∂ x $X = \partial _x$ and Y = f ( x ) ∂ y $Y = f(x) \partial _y$ defined on U $\mathcal {U}$ form a family of g $g$ -orthonormal fields. The only nonzero bracket relation is[ X , Y ] = f ′ f Y . $$\begin{equation*} [X, Y] = \frac{f^{\prime }}{f} Y. \end{equation*}$$ Using the Koszul formula, we easily obtain∇ X X = ∇ X Y = 0 , ∇ Y X = − f ′ f Y , ∇ Y Y = f ′ f X . $$\begin{equation*} \nabla _X X = \nabla _X Y = 0, \qquad \nabla _Y X = - \frac{f^{\prime }}{f} Y, \qquad \nabla _Y Y = \frac{f^{\prime }}{f} X. \end{equation*}$$ It follows that the only nonzero entry of the Riemann curvature tensor isR ( X , Y , X , Y ) = f ′ f ′ − f ′ f 2 . $$\begin{equation*} \mathrm{R}(X, Y, X, Y) = {\left(\frac{f^{\prime }}{f}\right)}^{\prime } - {\left(\frac{f^{\prime }}{f}\right)}^2. \end{equation*}$$ Using the symmetries, we obtain
Ric = f ′ f ′ − f ′ f 2 d x ⊗ d x + 1 f 2 f ′ f ′ − f ′ f 2 d y ⊗ d y . $$\begin{equation*} \mathrm{Ric} = {\left[{\left(\frac{f^{\prime }}{f}\right)}^{\prime } - {\left(\frac{f^{\prime }}{f}\right)}^2\right]} \mathop {}\!\mathrm{d}x \otimes \mathop {}\!\mathrm{d}x + \frac{1}{f^2}{\left[{\left(\frac{f^{\prime }}{f}\right)}^{\prime } - {\left(\frac{f^{\prime }}{f}\right)}^2\right]} \mathop {}\!\mathrm{d}y \otimes \mathop {}\!\mathrm{d}y. \end{equation*}$$ We also find that ∇ V = X ( V ) X $\nabla V =X(V)X$ , and that the only nonzero entries of the Hessian are
Hess ( V ) ( X , X ) = V ′ ′ , Hess ( V ) ( Y , Y ) = − f ′ f V ′ . $$\begin{equation*} \mathrm{Hess}(V)(X,X) = V^{\prime \prime },\nobreakspace \nobreakspace \mathrm{Hess}(V)(Y,Y)=-\frac{f^{\prime }}{f}V^{\prime }. \end{equation*}$$ Finally, one has the nonzero entry
d V ⊗ d V ( X , X ) = ( V ′ ) 2 , $$\begin{equation*} \mathop {}\!\mathrm{d}V\otimes \mathop {}\!\mathrm{d}V(X,X)=(V^{\prime })^2, \end{equation*}$$ and the proof is complete with Equation 9.□ $\Box$
The α $\alpha$ -Grushin hemisphere
We make use of the coordinate chart and the notation described in Subsection 3.1.
5.2
Proposition
Given K ∈ R $K \in \mathbb {R}$ , N ∈ − ∞ , 0 ∪ 2 , + ∞ $N \in \left(-\infty, 0\right) \cup \left[2, +\infty \right)$ , α ⩾ 0 $\alpha \geqslant 0$ and β ⩾ α $\beta \geqslant \alpha$ , the N $N$ -Ricci curvature of the α $\alpha$ -Grushin open hemisphere ( S α + , d S α , m S α β ) $(\mathbb {S}_\alpha ^+, \mathop {}\!\mathrm{d}_{\mathbb {S}_\alpha }, \mathfrak {m}^\beta _{\mathbb {S}_\alpha })$ is
Ric N , V = − 1 sin 2 ( x ) 3 α − 1 + ( α − 1 ) 2 cos 2 ( x ) − β + β 2 N − 2 cos 2 ( x ) d x ⊗ d x − cos 2 ( x ) sin 2 α + 2 ( x ) 3 α − 1 + ( α − 1 ) 2 cos 2 ( x ) − β 1 + ( α − 1 ) cos 2 ( x ) d y ⊗ d y . $$\begin{align*} \mathrm{Ric}_{N,V}=&-\frac{1}{\sin ^2(x)}{\left[3\alpha -1+(\alpha -1)^2\cos ^2(x) - \beta + \frac{\beta ^2}{N-2}\cos ^2(x) \right]}\mathop {}\!\mathrm{d}x\otimes \mathop {}\!\mathrm{d}x \\ &-\frac{\cos ^2(x)}{\sin ^{2\alpha +2}(x)}{\left[3\alpha -1+(\alpha -1)^2\cos ^2(x) - \beta {\left(1+(\alpha -1)\cos ^2(x)\right)}\right]}\mathop {}\!\mathrm{d}y\otimes \mathop {}\!\mathrm{d}y. \end{align*}$$ Furthermore, it holds Ric N , V ⩾ K $\mathrm{Ric}_{N,V} \geqslant K$ if and only if
β − α 2 − α + min − K + ( α − 1 ) 2 , − β 2 N − 2 , β ( α − 1 ) ⩾ 0 . $$\begin{equation*} \beta -\alpha ^2-\alpha +\min {\left(-K+(\alpha -1)^2,\nobreakspace -\frac{\beta ^2}{N-2},\nobreakspace \beta (\alpha -1)\right)}\geqslant 0. \end{equation*}$$
Proof
The first part is done by Lemma 5.1 with f ( x ) = | sin ( x ) | α cos ( x ) $f(x)=\frac{|\sin (x){|}^{\alpha}}{\cos (x)}$ and V ( x ) = − β log | sin ( x ) | $V(x)=-\beta \log |\sin (x)|$ . For the second part, we see by comparing the coefficients of the tensors d x ⊗ d x $\mathop {}\!\mathrm{d}x\otimes \mathop {}\!\mathrm{d}x$ and d y ⊗ d y $\mathop {}\!\mathrm{d}y\otimes \mathop {}\!\mathrm{d}y$ , respectively, that Ric N , V ⩾ K $\mathrm{Ric}_{N,V}\geqslant K$ holds if and only if
− 3 α + 1 + β − K + K − ( α − 1 ) 2 − β 2 N − 2 cos 2 ( x ) ⩾ 0 $$\begin{equation*} -3\alpha +1+\beta -K+{\left[K-(\alpha -1)^2-\frac{\beta ^2}{N-2}\right]}\cos ^2(x)\geqslant 0 \end{equation*}$$ and
− 3 α + 1 + β − K + K − ( α − 1 ) 2 + β ( α − 1 ) cos 2 ( x ) ⩾ 0 . $$\begin{equation*} -3\alpha +1+\beta -K+{\left[K-(\alpha -1)^2+\beta (\alpha -1)\right]}\cos ^2(x)\geqslant 0. \end{equation*}$$ As cos ( x ) $\cos (x)$ takes its value in [0,1) on S α + $\mathbb {S}_\alpha ^+$ , the above two inequalities are equivalent to the following inequality:
− 3 α + 1 + β − K + min 0 , K − ( α − 1 ) 2 − β 2 N − 2 , K − ( α − 1 ) 2 + β ( α − 1 ) ⩾ 0 , $$\begin{equation*} -3\alpha +1+\beta -K+\min {\left(0,K-(\alpha -1)^2-\frac{\beta ^2}{N-2},K-(\alpha -1)^2+\beta (\alpha -1)\right)}\geqslant 0, \end{equation*}$$ which is trivially equivalent to the inequality in the statement.□ $\Box$
The α $\alpha$ -Grushin hyperbolic half-plane
We make use of the coordinate chart and the notation described in Subsection 3.2.
5.3
Proposition
Given K ∈ R $K \in \mathbb {R}$ , N ∈ ( − ∞ , 0 ) ∪ [ 2 , + ∞ ) $N\in (-\infty ,0)\cup [2,+\infty )$ , α ⩾ 0 $\alpha \geqslant 0$ and β ⩾ α $\beta \geqslant \alpha$ , the N $N$ -Ricci curvature of the α $\alpha$ -Grushin open hyperbolic half-plane ( H α + , d H α , m H α β ) $(\mathbb {H}_\alpha ^+, \mathop {}\!\mathrm{d}_{\mathbb {H}_\alpha }, \mathfrak {m}^\beta _{\mathbb {H}_\alpha })$ isRic N , V = − 1 sinh 2 ( x ) 3 α − 1 + ( α − 1 ) 2 cosh 2 ( x ) − β + β 2 N − 2 cosh 2 ( x ) d x ⊗ d x − cosh 2 ( x ) sin 2 α + 2 ( x ) 3 α − 1 + ( α − 1 ) 2 cosh 2 ( x ) − β 1 + ( α − 1 ) cosh 2 ( x ) d y ⊗ d y . $$\begin{align*} \mathrm{Ric}_{N,V}=&-\frac{1}{\sinh ^2(x)}{\left[3\alpha -1+(\alpha -1)^2\cosh ^2(x) - \beta + \frac{\beta ^2}{N-2}\cosh ^2(x) \right]}\mathop {}\!\mathrm{d}x\otimes \mathop {}\!\mathrm{d}x \\ &-\frac{\cosh ^2(x)}{\sin ^{2\alpha +2}(x)}{\left[3\alpha -1+(\alpha -1)^2\cosh ^2(x) - \beta {\left(1+(\alpha -1)\cosh ^2(x)\right)}\right]}\mathop {}\!\mathrm{d}y\otimes \mathop {}\!\mathrm{d}y. \end{align*}$$ Furthermore, it holds Ric N , V ⩾ K $\mathrm{Ric}_{N,V} \geqslant K$ if and only ifmin − K − ( α − 1 ) 2 , β − α 2 − α + min − β 2 N − 2 , β ( α − 1 ) ⩾ 0 . $$\begin{equation*} \min {\left(-K-(\alpha -1)^2,\beta -\alpha ^2-\alpha \right)}+\min {\left(-\frac{\beta ^2}{N-2},\nobreakspace \beta (\alpha -1)\right)}\geqslant 0. \end{equation*}$$
Proof
The first part is done by Lemma 5.1 with f ( x ) = | sinh ( x ) | α cosh ( x ) $f(x)=\frac{|\mathrm{\sinh}(x){|}^{\alpha}}{\cosh (x)}$ and V ( x ) = − β log | sinh ( x ) | $V(x)=-\beta \log |\mathrm{\sinh}(x)|$ . For the second part, we see by comparing the coefficients of the tensors d x ⊗ d x $\mathop {}\!\mathrm{d}x\otimes \mathop {}\!\mathrm{d}x$ and d y ⊗ d y $\mathop {}\!\mathrm{d}y\otimes \mathop {}\!\mathrm{d}y$ , respectively, that Ric N , V ⩾ K $\mathrm{Ric}_{N,V}\geqslant K$ holds if and only if− 3 α + 1 + β + K + − K − ( α − 1 ) 2 − β 2 N − 2 cosh 2 ( x ) ⩾ 0 $$\begin{equation*} -3\alpha +1+\beta +K+{\left[-K-(\alpha -1)^2-\frac{\beta ^2}{N-2}\right]}\cosh ^2(x)\geqslant 0 \end{equation*}$$ and− 3 α + 1 + β + K + − K − ( α − 1 ) 2 + β ( α − 1 ) cosh 2 ( x ) ⩾ 0 . $$\begin{equation*} -3\alpha +1+\beta +K+{\left[-K-(\alpha -1)^2+\beta (\alpha -1)\right]}\cosh ^2(x)\geqslant 0. \end{equation*}$$ As cosh ( x ) ∈ [ 1 , + ∞ ) $\cosh (x)\in [1,+\infty)$ , the two inequalities hold for all x ⩾ 0 $x\geqslant 0$ if and only ifmin 0 , − 3 α + 1 + β + K + min − K − ( α − 1 ) 2 − β 2 N − 2 , − K − ( α − 1 ) 2 + β ( α − 1 ) ⩾ 0 $$\begin{equation*} \min {\left(0,-3\alpha +1+\beta +K\right)}+\min {\left(-K-(\alpha -1)^2-\frac{\beta ^2}{N-2},-K-(\alpha -1)^2+\beta (\alpha -1)\right)}\geqslant 0 \end{equation*}$$ holds, which is equivalent to the inequality in the statement.□ $\Box$
The ∞ $\infty$ -Grushin half-plane
We make use of the coordinate chart and the notation described in Subsection 3.3.
5.4
Proposition
Given K ∈ R $K \in \mathbb {R}$ , N ∈ ( − ∞ , 0 ) ∪ [ 2 , + ∞ ) $N\in (-\infty ,0)\cup [2,+\infty )$ and β , γ ⩾ 0 $\beta, \gamma \geqslant 0$ , the N $N$ -Ricci curvature of the ∞ $\infty$ -Grushin open half-plane ( G ∞ + , d G ∞ , m G ∞ β , γ ) $(\mathbb {G}_\infty ^+, \mathop {}\!\mathrm{d}_{\mathbb {G}_\infty }, \mathfrak {m}^{\beta,\gamma }_{\mathbb {G}_\infty })$ isRic N , V = 6 γ − 1 | x | 4 − 2 | x | 3 + β | x | 2 − 1 ( N − 2 ) x 6 2 γ + β x 2 2 d x ⊗ d x + 2 γ | x | 5 − 1 | x | 4 + β − 2 | x | 3 e 2 | x | d y ⊗ d y . $$\begin{align*} \mathrm{Ric}_{N,V}=&{\left[\frac{6\gamma -1}{|x|^4}-\frac{2}{|x|^3}+\frac{\beta }{|x|^2}-\frac{1}{(N-2)x^6}{\left(2\gamma +\beta x^2\right)}^2\right]}\mathop {}\!\mathrm{d}x\otimes \mathop {}\!\mathrm{d}x\\ &+{\left[\frac{2\gamma }{|x|^5}-\frac{1}{|x|^4}+\frac{\beta -2}{|x|^3}\right]}e^{\frac{2}{|x|}}\mathop {}\!\mathrm{d}y\otimes \mathop {}\!\mathrm{d}y. \end{align*}$$ It holds Ric N , V ⩾ K $\mathrm{Ric}_{N,V} \geqslant K$ if and only if the following inequalities hold for any x ⩾ 0 $x\geqslant 0$ :( 6 γ − 1 ) x 2 − 2 x 3 + β x 4 − 1 N − 2 ( 2 γ + β x 2 ) 2 ⩾ K x 6 and 2 γ − x + ( β − 2 ) x 2 ⩾ K x 5 . $$\begin{equation*} (6\gamma -1)x^2-2x^3+\beta x^4-\frac{1}{N-2}(2\gamma +\beta x^2)^2\geqslant Kx^6\nobreakspace \nobreakspace \text{and}\nobreakspace \nobreakspace 2\gamma -x+(\beta -2)x^2\geqslant Kx^5. \end{equation*}$$ In particular, Ric ∞ , V ⩾ 0 $\mathrm{Ric}_{\infty,V}\geqslant 0$ holds if and only ifmin { β ( 6 γ − 1 ) , 8 ( β − 2 ) γ } ⩾ 1 . $$\begin{equation*} \min \lbrace \beta (6\gamma -1),8(\beta -2)\gamma \rbrace \geqslant 1. \end{equation*}$$
Proof
The proof is reached by using Lemma 5.1 with f ( x ) = e − 1 / | x | $f(x)=e^{-1/|x|}$ and V ( x ) = γ | x | 2 − β log | x | $V(x)=\frac{\gamma }{|x|^2}-\beta \log |x|$ , and by comparing the coefficients of d x ⊗ d x $\mathop {}\!\mathrm{d}x\otimes \mathop {}\!\mathrm{d}x$ and d y ⊗ d y $\mathop {}\!\mathrm{d}y\otimes \mathop {}\!\mathrm{d}y$ , respectively.□ $\Box$
Theorems 1.1, 1.2, and 1.4 follow from the Ricci computations this section, noting that the validity of the assumptions in Theorem 4.1 are verified because of the results of Section 3.
ACKNOWLEDGMENTS
This project has received funding from the European Research Council (ERC) under the European Union's Horizon 2020 Research and Innovation Program (Grant Agreement No. 945655). This research was funded in part by the Austrian Science Fund (FWF) [Grant DOI 10.55776/EFP6]. The authors thank Shouhei Honda for leading the authors’ attention to these problems, and telling us about the consequences in Remark 1.5.
JOURNAL INFORMATION
The Transactions of the London Mathematical Society is wholly owned and managed by the London Mathematical Society, a not-for-profit Charity registered with the UK Charity Commission. All surplus income from its publishing programme is used to support mathematicians and mathematics research in the form of research grants, conference grants, prizes, initiatives for early career researchers and the promotion of mathematics.
References
A. Agrachev, D. Barilari, and U. Boscain, A comprehensive introduction to sub‐Riemannian geometry, Cambridge Studies in Advanced Mathematics, vol. 181, Cambridge University Press, Cambridge, 2020, pp. xviii+745. (From the Hamiltonian viewpoint. With an appendix by Igor Zelenko.) DOI 10.1017/9781108677325.
L. Ambrosio, Calculus, heat flow and curvature‐dimension bounds in metric measure spaces, Proceedings of the International Congress of Mathematicians, vol. I (Rio de Janeiro, 2018), Plenary lectures, World Sci. Publ., Hackensack, NJ, pp. 301–340, DOI 10.1142/9789813272880_0015.
L. Ambrosio, N. Gigli, and G. Savaré, Density of Lipschitz functions and equivalence of weak gradients in metric measure spaces, Revista Matemática Iberoamericana 29 (2013), no. 3, 969–996, DOI 10.4171/RMI/746
L. Ambrosio, N. Gigli, and G. Savaré, Bakry–Émery curvature‐dimension condition and Riemannian Ricci curvature bounds, Ann. Probab. 43 (2015), no. 1, 339–404, DOI 10.1214/14‐AOP907
L. Ambrosio and G. Stefani, Heat and entropy flows in Carnot groups, Rev. Mat. Iberoam. 36 (2020), no. 1, 257–290, DOI 10.4171/rmi/1129.
S. Borza, Distortion coefficients of the α$\alpha$‐Grushin plane, J. Geom. Anal. 32 (2022), no. 3, Paper No. 78, 28, DOI 10.1007/s12220‐021‐00736‐8.
S. Borza, Normal forms for the sub‐Riemannian exponential map of Gα$\mathbb {G}_\alpha$, SU(2)$\mathrm{SU}(2)$, and SL(2)$\mathrm{SL}(2)$, New trends in sub‐Riemannian geometry, Contemporary Mathematics, vol. 809, Amer. Math. Soc., 2025, DOI 10.1090/conm/809.
S. Borza and K. Tashiro, Measure contraction property, curvature exponent and geodesic dimension of sub‐Finsler ℓp$\ell ^p$‐Heisenberg groups, math. MG (2023), arXiv: 2305.16722.
S. Borza, M. Magnabosco, T. Rossi, and K. Tashiro, Measure contraction property and curvature‐dimension condition on sub‐Finsler Heisenberg groups, 2024.
S. Borza, M. Magnabosco, T. Rossi and K. Tashiro, Curvature exponent of sub‐Finsler Heisenberg groups, SIAM J. Math. Anal. 57 (2025), no. 4, 3561–3586, DOI 10.1137/24M1690692.
U. Boscain, D. Prandi, and M. Seri, Spectral analysis and the Aharonov–Bohm effect on certain almost‐Riemannian manifolds, Comm. Partial Differential Equations 41 (2016), no. 1, 32–50, DOI 10.1080/03605302.2015.1095766.
D. Burago, Y. Burago, and S. Ivanov, A course in metric geometry, Graduate Studies in Mathematics, vol. 33, Amer. Math. Soc., Providence, RI, 2001, pp. xiv+415, DOI 10.1090/gsm/033.
D.‐C. Chang and Y. Li, SubRiemannian geodesics in the Grushin plane, J. Geom. Anal. 22 (2012), no. 3, 800–826, DOI 10.1007/s12220‐011‐9215‐y.
Y. Chitour, D. Prandi, and L. Rizzi, Weyl's law for singular Riemannian manifolds, J. Math. Pures Appl. (9) 181 (2024), 113–151, DOI 10.1016/j.matpur.2023.10.004.
X. Dai, S. Honda, J. Pan, and G. Wei, Singular Weyl's law with Ricci curvature bounded below, Trans. Amer. Math. Soc. Ser. B 10 (2023), 1212–1253, DOI 10.1090/btran/160.
M. Erbar, K. Kuwada, and K.‐T. Sturm, On the equivalence of the entropic curvature‐dimension condition and Bochner's inequality on metric measure spaces, Invent. Math. 201 (2015), no. 3, 993–1071, DOI 10.1007/s00222‐014‐0563‐7.
B. Franchi and E. Lanconelli, Une métrique associée à une classe d'opérateurs elliptiques dégénérés, Conference on linear partial and pseudodifferential operators (Torino, 1982), 1983, pp. 105–114.
N. Gigli, On the differential structure of metric measure spaces and applications, Mem. Amer. Math. Soc. 236 (2015), no. 1113, vi+91, DOI 10.1090/memo/1113.
N. Gigli, A. Mondino, and G. Savaré, Convergence of pointed non‐compact metric measure spaces and stability of Ricci curvature bounds and heat flows, Proc. Lond. Math. Soc. (3) 111 (2015), no. 5, 1071–1129, DOI 10.1112/plms/pdv047.
B.‐X. Han, Measure rigidity of synthetic lower Ricci curvature bound on Riemannian manifolds, Adv. Math. 373 (2020), Art. 107327, 31 pp., DOI 10.1016/j.aim.2020.107327.
N. Juillet, Geometric inequalities and generalized Ricci bounds in the Heisenberg group, Int. Math. Res. Not. IMRN (2009), no. 13, 2347–2373, DOI 10.1093/imrn/rnp019.
N. Juillet, Sub‐Riemannian structures do not satisfy Riemannian Brunn–Minkowski inequality, Rev. Mat. Iberoam. 37 (2021), no. 1, 177–188, DOI 10.4171/rmi/1205.
E. Le Donne, D. Lucic, and E. Pasqualetto, Universal infinitesimal Hilbertianity of sub‐Riemannian manifolds, Potential Anal. 59 (2023), no. 1, 349–374, DOI 10.1007/s11118‐021‐09971‐8.
J. Lott and C. Villani, Ricci curvature for metric‐measure spaces via optimal transport, Ann. of Math. (2) 169 (2009), no. 3, 903–991, DOI 10.4007/annals.2009.169.903.
M. Magnabosco and C. Rigoni, Optimal maps and local‐to‐global property in negative dimensional spaces with Ricci curvature bounded from below, Tohoku Math. J. 75 (2023), no. 4, 483–507, DOI 10.2748/tmj.20220420.
M. Magnabosco, C. Rigoni, and G. Sosa, Convergence of metric measure spaces satisfying the CD condition for negative values of the dimension parameter, Nonlinear Anal. 237 (2023), Paper No. 113366, 48, DOI 10.1016/j.na.2023.113366.
M. Magnabosco and T. Rossi, Almost‐Riemannian manifolds do not satisfy the curvature‐dimension condition, Calc. Var. Partial Differential Equations 62 (2023), no. 4, Paper No. 123, 27, DOI 10.1007/s00526‐023‐02466‐x.
M. Magnabosco and T. Rossi, Failure of the curvature‐dimension condition in sub‐Finsler manifolds, math. MG (2022), arXiv: 2307.01820.
G. E. Martin, The foundations of geometry and the non‐Euclidean plane, Intext Series in Mathematics, Intext Educational Publishers, New York, 1975, pp. xvi+509.
R. J. McCann, Polar factorization of maps on Riemannian manifolds, Geom. Funct. Anal. 11 (2001), no. 3, 589–608, DOI 10.1007/PL00001679.
E. Milman, Beyond traditional curvature‐dimension I: new model spaces for isoperimetric and concentration inequalities in negative dimension, Trans. Amer. Math. Soc. 369 (2017), no. 5, 3605–3637, DOI 10.1090/tran/6796
E. Milman, Harmonic measures on the sphere via curvature‐dimension, Ann. Fac. Sci. Toulouse Math. 26 (2017), no. 2, 437–449, DOI 10.5802/afst.1540
S.‐i. Ohta, (K,N)$(K,N)$‐convexity and the curvature‐dimension condition for negative N$N$, J. Geom. Anal. 26 (2016), no. 3, 2067–2096, DOI 10.1007/s12220‐015‐9619‐1.
J. Pan, The Grushin hemisphere as a Ricci limit space with curvature ⩾1$\geqslant 1$, Proc. Amer. Math. Soc. Ser. B 10 (2023), 71–75, DOI 10.1090/bproc/160.
J. Pan and G. Wei, Examples of Ricci limit spaces with non‐integer Hausdorff dimension, Geom. Funct. Anal. 32 (2022), no. 3, 676–685, DOI 10.1007/s00039‐022‐00598‐4.
L. Rizzi and G. Stefani, Failure of curvature‐dimension conditions on sub‐Riemannian manifolds via tangent isometries, J. Funct. Anal. 285 (2023), no. 9, Paper No. 110099, 31, DOI 10.1016/j.jfa.2023.110099.
L. D. Schiavo, M. Magnabosco, and C. Rigoni, Gradient flows of (K,N)$(K,N)$‐convex functions with negative N$N$, math. FA (2024), arXiv: 2412.04574.
K.‐T. Sturm, On the geometry of metric measure spaces. II, Acta Math. 196 (2006), no. 1, 133–177, DOI 10.1007/s11511‐006‐0003‐7.
C. Villani, Optimal transport: old and new, vol. 338, Springer, Berlin, 2009, pp. xxii+973, DOI 10.1007/978‐3‐540‐71050‐9.