Hostname: page-component-857557d7f7-nhjpk Total loading time: 0 Render date: 2025-11-22T12:13:16.388Z Has data issue: false hasContentIssue false

Existence of ground states for free energies on the hyperbolic space

Published online by Cambridge University Press:  07 August 2025

José Antonio Carrillo
Affiliation:
Mathematical Institute, University of Oxford , Oxford OX2 6GG, United Kingdom e-mail: carrillo@maths.ox.ac.uk
Razvan Fetecau*
Affiliation:
Department of Mathematics, Simon Fraser University , Burnaby V5A 1S6, BC, Canada
Hansol Park
Affiliation:
Department of Mathematics, National Tsing Hua University , Hsinchu 30013, Taiwan e-mail: hansolpark@math.nthu.edu.tw
*
Rights & Permissions [Opens in a new window]

Abstract

We investigate a free energy functional that arises in aggregation-diffusion phenomena modelled by nonlocal interactions and local repulsion on the hyperbolic space ${\mathbb H}^n$. The free energy consists of two competing terms: an entropy, corresponding to slow nonlinear diffusion, that favours spreading, and an attractive interaction potential energy that favours aggregation. We establish necessary and sufficient conditions on the interaction potential for ground states to exist on the hyperbolic space ${\mathbb H}^n$. To prove our results, we derived several Hardy–Littlewood–Sobolev (HLS)-type inequalities on general Cartan–Hadamard manifolds of bounded curvature, which have an interest in their own.

Information

Type
Article
Creative Commons
Creative Common License - CCCreative Common License - BY
This is an Open Access article, distributed under the terms of the Creative Commons Attribution licence (https://creativecommons.org/licenses/by/4.0), which permits unrestricted re-use, distribution and reproduction, provided the original article is properly cited.
Copyright
© The Author(s), 2025. Published by Cambridge University Press on behalf of Canadian Mathematical Society

1 Introduction

In this work, we investigate the existence of global minimizers of the following free energy functional on a Cartan–Hadamard manifold M:

(1.1) $$ \begin{align} E[\rho]= \frac{1}{m-1} \int_M \rho(x)^{m} \mathrm{d} x+\frac{1}{2}\iint\limits_{M \times M} W(x, y)\rho(x)\rho(y)\mathrm{d} x \mathrm{d} y, \end{align} $$

where $m>1$ and $W: M\times M \to \mathbb R$ is an attractive interaction potential. This definition holds for densities $\rho $ in $ L^m(M) \cap \mathcal {P}(M)$ , where $\mathcal {P}(M)$ denotes the space of probability measures on M. Otherwise, $E[\rho ]$ is defined by means of sequences $E[\rho _k]$ , with $\rho _k \rightharpoonup \rho $ as $k \to \infty $ (see (2.2) for the precise definition).

The minimizers of the energy functional (2.2) are critical points of the following nonlocal evolution equation:

(1.2) $$ \begin{align} \partial_t\rho(x)- \nabla_M \cdot(\rho(x)\nabla_M W*\rho(x))=\Delta \rho^m(x), \end{align} $$

where

$$\begin{align*}W*\rho(x)=\int_M W(x, y)\rho(y) \mathrm{d} y. \end{align*}$$

Here, $\nabla _M \cdot $ and $\nabla _M $ represent the manifold divergence and gradient, respectively, $\Delta $ is the Laplace–Beltrami operator on M, and $\mathrm {d} x$ denotes the Riemannian volume measure on M. In particular, equation (1.2) can be formally expressed as the gradient flow of the energy E on a suitable Wasserstein space [Reference Ambrosio, Gigli and Savaré2, Example 11.2.7].

The free energy functional (1.1) and its corresponding gradient flow (1.2) (aggregation-diffusion equation) have been extensively studied on the Euclidean space ( $M=\mathbb R^n$ ). The strong interest in this variational problem stems from its many applications in various fields, such as mathematical biology [Reference Keller and Segel38, Reference Patlak42, Reference Topaz, Bertozzi and Lewis45], material science [Reference Carrillo, McCann and Villani13], statistical mechanics [Reference Sire and Chavanis43, Reference Sire and Chavanis44], robotics [Reference Gazi and Passino28, Reference Ji and Egerstedt36], and social sciences [Reference Motsch and Tadmor41]. In such applications, the potential W models social interactions (repulsion and attraction) and the nonlinear diffusion models repulsive/anticrowding effects.

Energies in the form (1.1) have two components: the first term is the entropy and the second is an interaction energy. In terms of dynamical evolution (1.2), the entropy gives rise to nonlinear diffusion and the interaction energy to the aggregation term. In this article, we consider exclusively attractive interaction potentials, in which case, the two terms are competing with each other. Specifically, diffusion promotes spreading and interaction leads to blow-up. Global energy minimizers exist when these two competing effects balance each other.

Depending on the range of m, one can classify the diffusion term $\Delta \rho ^m$ into three cases: (i) slow diffusion ( $m>1$ ), (ii) linear diffusion ( $m=1$ ), and (iii) fast diffusion ( $0<m<1$ ); for linear diffusion, the entropy term in (1.1) is replaced by $\int _M \rho (x) \log \rho (x) \mathrm {d} x$ . In the Euclidean space $\mathbb R^n$ , each of these regimes has been studied extensively – we refer the reader to [Reference Carrillo, Craig, Yao, Bellomo, Degond and Tadmor9] for a comprehensive recent review on this subject. In [Reference Calvez, Carrillo and Hoffmann6, Reference Carrillo, Delgadino and Patacchini10, Reference Carrillo, Hoffmann, Mainini and Volzone12] the authors considered a singular interaction potential of Riesz type, i.e., $W(x, y)=-\frac {1}{\beta \|x-y\|^\beta }$ , where $\|\cdot \|$ denotes the Euclidean norm and $0<\beta <n$ , and established existence or non-existence of global energy minimizers for various ranges of m. Uniqueness and qualitative properties (e.g., monotonicity and radial symmetry) of energy minimizers or steady states of (1.2) have been studied in [Reference Bedrossian3, Reference Calvez, Carrillo and Hoffmann7, Reference Carrillo, Hittmeir, Volzone and Yao11, Reference Delgadino, Yan and Yao16, Reference Kaib37]. Well-posedness, long time behaviour and properties of solutions to the dynamical model (1.2) were addressed in [Reference Carrillo, Craig, Yao, Bellomo, Degond and Tadmor9, Reference Carrillo, Hittmeir, Volzone and Yao11, Reference Fernández-Jiménez, Carrillo and Gómez-Castro19]. Relevant for our own work, we note that the key tools used in some of these works are Hardy–Littlewood–Sobolev (HLS)-type inequalites and Riesz rearrangement inequalities [Reference Blanchet, Carrillo and Laurençot4, Reference Carrillo, Delgadino and Patacchini10Reference Carrillo, Hoffmann, Mainini and Volzone12].

Without the diffusion term, model (1.2) has a large literature on its own, in particular for its setup in Euclidean spaces $M=\mathbb R^n$ (see, for instance, [Reference Carrillo and Vázquez14, Reference Kolokolnikov, Carrillo, Bertozzi, Fetecau and Lewis39] and the references therein for some of the results in the literature). The well-posedness of the model without diffusion on general and specific manifolds was studied in [Reference Fetecau, Park and Patacchini25, Reference Fetecau and Patacchini26, Reference Wu and Slepčev46] and emergent behaviours were studied in [Reference Fetecau and Zhang27, Reference Ha, Hwang, Kim, Kim and Min34] (hyperbolic space), [Reference Fetecau, Ha and Park20] (the special orthogonal group $SO(3)$ ), [Reference Ha, Kang and Kim35] (Stiefel manifolds), and [Reference Fetecau and Park22] (general Riemannian manifolds of bounded curvature). Furthermore, a variational approach was taken in [Reference Fetecau and Park21] to study equilibria and energy minimizers of the aggregation model on the hyperbolic space.

In this article, we investigate the existence of global minimizers of the energy (1.1) via a variational approach rather than the dynamical model (1.2). Throughout the article, we assume that the interaction potential is purely attractive and that it is in the form $W(x, y)=h(d(x, y))$ , where h is a non-decreasing and lower-semicontinuous function, and d denotes the geodesic distance on the manifold. Using the geodesic distance for modeling interactions is very natural since it makes the model completely intrinsic to the given manifold. The class of potentials we consider includes Riesz-type potentials, as studied in [Reference Blanchet, Carrillo and Laurençot4, Reference Carrillo, Delgadino and Patacchini10, Reference Carrillo, Hoffmann, Mainini and Volzone12] for $\mathbb R^n$ . In studying the existence of energy minimizers, we only consider interaction potentials $h(\theta )$ that are less singular than $-\frac {1}{\theta ^{\min (n(m-1),n)}}$ at origin; otherwise, attraction is too singular and the energy functional is not bounded below. As shown in our results, this threshold in the singular behaviour of the interaction potential is sharp for existence of ground states of finite energies on ${\mathbb H}^n$ . A similar threshold property holds also in the Euclidean case [Reference Carrillo, Delgadino and Patacchini10].

We also note that in the general manifold setup, instead of the nonlocal interaction term in equation (1.2), local reaction terms have been considered in [Reference Grillo, Meglioli and Punzo31] (with linear diffusion) and [Reference Grillo, Meglioli and Punzo29, Reference Grillo, Meglioli and Punzo30, Reference Grillo, Muratori and Vázquez32, Reference Grillo, Muratori and Vázquez33] (with nonlinear diffusion). In these papers, the authors also need to deal with the competing effects of a local reaction term that leads to blow-up and the diffusion that results in spreading. Issues such as global existence versus blow-up, as well as the long-time behaviour of solutions have been investigated. The methods used in these papers are specific to local reaction terms however.

To the best of our knowledge, our study is the first to consider the manifold setup for free energies containing both nonlocal intrinsic interactions and nonlinear diffusion. The contributions of our article are the following. First, we derive several HLS-type inequalities on general Cartan–Hadamard manifolds with bounded sectional curvatures. Second, we investigate the existence of ground states of the energy functional (1.1) on the hyperbolic space ${\mathbb H}^n$ , with $m>1$ (slow diffusion regime). Our main results show the existence of global energy minimizers on the hyperbolic space ${\mathbb H}^n$ under two behaviors of h at $\infty $ , namely: $\lim _{\theta \to \infty }h(\theta )=\infty $ (Theorem 5.1) and $\lim _{\theta \to \infty }h(\theta )=0$ (Theorem 6.1). In this regard, the methods developed hereafter can also be used to generalize the results in $\mathbb R^n$ to general interaction potentials not necessarily of Riesz-type (see Remark 6.3).

Some of the crucial ingredients in our study are the new HLS-type inequalities that we established together with Riesz rearrangement inequalities on ${\mathbb H}^n$ [Reference Burchard and Schmuckenschläger5]. In particular, the latter inequalities enable us to restrict the minimizing sequences to be radially symmetric and decreasing about a pole. The setup $M={\mathbb H}^n$ for our existence results is motivated by various specific properties of the hyperbolic space that are used in our proofs: the availability of the Riesz rearrangement inequality and the fact that ${\mathbb H}^n$ is a homogeneous manifold (i.e., for any choice of two points on it, there is an isometry of ${\mathbb H}^n$ that sends one of the two points to the other). Finally, we note that for simplicity, we have fixed the diffusion coefficient to value $1$ in our article. For a generic diffusion coefficient $\nu>0$ , one can simply scale the interaction potential by $1/\nu $ , and work with the scaled potential instead.

Variational approaches to the free energy functional with linear diffusion on Cartan–Hadamard manifolds were recently considered in [Reference Fetecau and Park23, Reference Fetecau and Park24]. The main tool there was also based on generalizations of HLS-type inequalities from the Euclidean space to Cartan–Hadamard manifolds; in the case of linear diffusion one needs to work instead with a logarithmic HLS inequality [Reference Carlen and Loss8]. The approach in [Reference Fetecau and Park23, Reference Fetecau and Park24] is to use the Riemannian exponential map (which for Cartan–Hadamard manifolds is a global diffeomorphism) and comparison theorems in differential geometry, to carry an HLS inequality from the tangent space of the manifold at a fixed pole, to the entire manifold. This approach has to be significantly adjusted to the nonlinear diffusion case using the Vitali covering lemma (see Lemma 3.1).

This article is organized as follows. In Section 2, we present assumptions, definitions, and preliminary background on differential geometry and rearrangement inequalities that is needed for the main results. Several HLS-type inequalities on general Cartan–Hadamard manifolds are derived in Section 3. The rest of the article deals exclusively with the case $M={\mathbb H}^n$ . In Section 4, we establish the weak lower-semicontinuity of the energy functional. The existence of global energy minimizers are proven in Sections 5 and 6, respectively – see Theorems 5.1 and 6.1. Finally, we provided proofs of several results in the Appendix.

2 Assumptions and preliminary background

2.1 Assumptions and definitions

The results in Section 3 apply to general Cartan–Hadamard manifolds, while Sections 46 are for the specific case when the manifold is the hyperbolic space. The general assumptions on the manifold M and the interaction potential W are the following.

(M) M is an n-dimensional Cartan–Hadamard manifold, i.e., M is complete, simply connected, and has everywhere non-positive sectional curvature. We denote its intrinsic distance by d and sectional curvature by $\mathcal {K}$ . In particular, for x a point in M and $\sigma $ a two-dimensional subspace of $T_x M$ , $\mathcal {K}(x;\sigma )$ denotes the sectional curvature of $\sigma $ at x.

(H) The interaction potential $W:M\times M \to \mathbb R$ is assumed to be in the form

$$\begin{align*}W(x, y)=h(d(x, y)), \qquad \text{ for all } x,y \in M, \end{align*}$$

where $h:[0, \infty )\to [-\infty , \infty )$ is non-decreasing and lower-semicontinuous. Note that the interactions are purely attractive, and also, that h can be singular at origin.

Since h is non-decreasing, we have two cases for the behaviour of h at infinity. Denote

$$\begin{align*}h_\infty = \lim_{\theta\to\infty}h(\theta); \end{align*}$$

then we can have either $h_\infty =\infty $ or $h_\infty <\infty $ . For the latter case, without loss of generality, we can assume $h_\infty =0$ . Hence, we will consider only these two cases:

(2.1) $$ \begin{align} h_\infty=\infty, \qquad \text{ or } \qquad h_\infty=0. \end{align} $$

Let us make precise now the definition of the energy functional. We define $E: \mathcal {P}(M) \to [-\infty , \infty ]$ by

(2.2) $$ \begin{align} E[\rho]=\begin{cases} \displaystyle \frac{1}{m-1}\int_M \rho(x)^m\mathrm{d} x +\frac{1}{2}\iint\limits_{M \times M} h(d(x, y))\rho(x)\rho(y) \mathrm{d} x\mathrm{d} y, & \text{ if }\rho\in L^m(M) \cap \mathcal{P}(M),\\[10pt] \displaystyle\inf_{\substack{\rho_k \in L^m(M) \cap \mathcal{P}(M) \\ \rho_k \rightharpoonup \rho}} \liminf_{k\to \infty} E[\rho_k], & \text{ otherwise}. \end{cases} \end{align} $$

First, we note that we make an abuse of notation when we write $\rho \in L^m(M) \cap \mathcal {P}(M)$ . Here, we refer to an absolutely continuous measure directly by its density, where the notation means $\rho \in L^m(M)$ and $\mathrm {d} \rho (x) = \rho (x)\, \mathrm {d} x \in \mathcal {P}(M)$ . We denote by $\mathcal {P}_{ac}(M)\subset \mathcal {P}(M)$ the space of probability measures that are absolutely continuous with respect to the Riemannian volume measure $\mathrm {d} x$ ; note that, with the abuse of notation mentioned above, $L^m(M) \cap \mathcal {P}(M) \subset \mathcal {P}_{ac}(M)$ . Second, the definition assumes that (2.2) is well-defined, which needs to be justified. For the interaction potentials considered in this article, this is done in Proposition 2.2.

Also, $\rho _k \rightharpoonup \rho $ denotes weak- $^{*}$ convergence of measures, i.e.,

$$\begin{align*}\int_M f(x) \rho_k(x) \mathrm{d} x \to \int_M f(x) \rho(x) \mathrm{d} x, \qquad \text{ as } k\to \infty, \end{align*}$$

for all $f \in C_0(M)$ , where $C_0(M)$ denotes the space of continuous functions on M that vanish at infinity. The weak- $^{*}$ topology on $\mathcal {P}(M)$ (regarded as a subset of the space of Radon measures $\mathcal {M}(M)$ ) is induced by the duality with the set $C_0(M)$ . Hence, by Banach–Alaoglu theorem, the following compactness property of the weak- $^{*}$ topology holds [Reference Carrillo, Delgadino and Patacchini10].

Theorem 2.1 (Compactness of weak- $^{*}$ topology)

Let $\{\rho _k\}_{k\in \mathbb {N}}$ be a sequence in $\mathcal {P}(M)$ . Then, there exists a subsequence of $\{\rho _k\}_{k\in \mathbb {N}}$ which converges weakly- $^{*}$ to some $\rho _0\in \mathcal {M}_+(M)$ such that $\int _M\rho _0(x)\mathrm {d} x\leq 1$ , where $\mathcal {M}_+(M)$ denotes the space of nonnegative Radon measures on M.

2.2 Comparison results in differential geometry

In this section, we present some comparison results in Riemannian geometry, which are used in the article. Recall that M is assumed to be a Cartan–Hadamard manifold, and hence, it has no conjugate points, it is diffeomorphic to $\mathbb R^n$ and the exponential map at any point is a global diffeomorphism. Hence, the tangent space $T_p M$ at a generic point $p \in M$ can be identified globally with $\mathbb R^n$ , a property that will used several times in the article.

2.2.1 Rauch comparison theorem

One tool in our proofs is Rauch’s comparison theorem, which enables us to compare lengths of curves on different manifolds. The result is the following [Reference do Carmo17, Chapter 10, Proposition 2.5].

Theorem 2.2 (Rauch comparison theorem)

Let M and $\tilde {M}$ be two n-dimensional Riemannian manifolds and suppose that for all $p\in M$ , $\tilde {p}\in \tilde {M}$ , and $\sigma \subset T_pM$ , $\tilde {\sigma }\subset T_{\tilde {p}}\tilde {M}$ , the sectional curvatures $\mathcal {K}$ and $\tilde {\mathcal {K}}$ of M and $\tilde {M}$ , respectively, satisfy

$$\begin{align*}\tilde{\mathcal{K}}(\tilde{p}; \tilde{\sigma})\geq \mathcal{K}(p;\sigma). \end{align*}$$

Let $p\in M$ , $\tilde {p}\in \tilde {M}$ and fix a linear isometry $i:T_pM\to T_{\tilde {p}}\tilde {M}$ . Let $r>0$ be such that the restriction ${\exp _p}_{|B_r(0)}$ is a diffeomorphism and ${\exp _{\tilde {p}}}_{|\tilde {B}_r(0)}$ is non-singular. Let ${c:[0, a]\to \exp _p(B_r(0))\subset M}$ be a differentiable curve and define a curve $\tilde {c}:[0, a]\to \exp _{\tilde {p}}(\tilde {B}_r(0))\subset \tilde {M}$ by

$$\begin{align*}\tilde{c}(s)=\exp_{\tilde{p}}\circ i\circ\exp^{-1}_p(c(s)),\qquad s\in[0, a]. \end{align*}$$

Then, the length of c is greater or equal than the length of $\tilde {c}$ .

2.2.2 Bounds on the Jacobian of the exponential map and volume comparison theorems

On manifolds with bounded sectional curvatures, the Jacobian $J(\exp _x)$ of the Riemannian exponential map can be bounded below and above as follows [Reference Chavel15].

Theorem 2.3 [Reference Chavel15, Theorems III.4.1 and III.4.3]

Suppose the sectional curvatures of M satisfy

(2.3) $$ \begin{align} -c_m\leq\mathcal{K}(x;\sigma)\leq -c_M<0, \end{align} $$

for all $x \in M$ and all two-dimensional subspaces $\sigma \subset T_x M$ , where $c_m$ and $c_M$ are positive constants. Then, for any $x \in M$ and $u \in T_x M$ , it holds that

$$\begin{align*}\left( \frac{\sinh(\sqrt{c_M}\|u\|)}{\sqrt{c_M}\|u\|} \right)^{n-1}\leq |J(\exp_x)(u)|\leq\left( \frac{\sinh(\sqrt{c_m}\|u\|)}{\sqrt{c_m}\|u\|} \right)^{n-1}. \end{align*}$$

A point on a manifold is called a pole if the exponential map at that point is a global diffeomorphism. On Cartan–Hadamard manifolds, all points are poles. For a fixed pole o on M, we denote

$$\begin{align*}\theta_x := d(o,x), \qquad \text{ for all } x \in M. \end{align*}$$

The bounds in Theorem 2.3 can be used to bound the volume of geodesic balls in M [Reference Chavel15]. Denote by $B_\theta (o)$ the open geodesic ball centred at o of radius $\theta $ , and by $|B_\theta (o)|$ its volume. The following volume comparison result hold.

Theorem 2.4 [Reference Chavel15, Theorems III.4.2 and III.4.4]

Suppose the sectional curvatures of M satisfy (2.3). Then,

$$\begin{align*}n w(n)\int_0^\theta \left(\frac{\sinh(\sqrt{c_M}t)}{\sqrt{c_M}}\right)^{n-1} \mathrm{d} t \leq |B_{\theta}(o)|\leq n w(n) \int_0 ^\theta \left(\frac{\sinh(\sqrt{c_m} t)}{\sqrt{c_m}}\right)^{n-1} \mathrm{d} t, \end{align*}$$

where $w(n)$ denotes the volume of the unit ball in $\mathbb R^n$ .

We note that the two bounds in Theorem 2.4 represent the volume of the ball of radius $\theta $ in the hyperbolic space ${\mathbb H}^n$ of constant curvatures $-c_M$ , and $-c_m$ , respectively.

Remark 2.1 A limiting case of the curvature upper bound (2.3) is $c_M=0$ , i.e., when the sectional curvatures of M satisfy

(2.4) $$ \begin{align} \mathcal{K}(x;\sigma) \leq 0, \end{align} $$

for all $x \in M$ and all two-dimensional subspaces $\sigma \subset T_x M$ . In such case, the volume of geodesic balls in M is bounded below by the volume of balls of the same radius in the Euclidean space $\mathbb R^n$ :

(2.5) $$ \begin{align} |B_\theta(o)|\geq w(n)\theta^n. \end{align} $$

2.3 General variational considerations

It is not difficult to infer that the attractive potential h cannot be too singular at origin, or else blow-up occurs. The formal result is the following, with proof presented in the Appendix.

Proposition 2.1 (Attraction too singular at origin)

Let $m>1$ and assume the manifold and the interaction potential satisfy assumptions (M) and (H), respectively. If h satisfies either

(2.6) $$ \begin{align} \lim_{\theta\to 0+}\left(h(\theta)+\left(\frac{2^{n(m-1)+1}}{(m-1)w(n)^{m-1}}\right)\frac{1}{\theta^{n(m-1)}}\right)=-\infty \end{align} $$

or

(2.7) $$ \begin{align} \liminf_{\theta\to0+}\theta^n h(\theta)<0, \end{align} $$

then the energy functional (2.2) is not bounded from below.

Proof See Appendix A.

Remark 2.2 Let the function h be given by

$$\begin{align*}h(\theta)=-\frac{\alpha}{\theta^{\beta}}, \end{align*}$$

for some $\alpha>0$ and $\beta>0$ . According to Proposition 2.1, the energy is not bounded from below if $\beta> \min (n, n(m-1))$ , and also for $1<m<2$ , $\beta =n(m-1)$ and ${\alpha>\frac {2^{n(m-1)-1}}{(m-1)w(n)^{m-1}}}$ .

This result is consistent with the non-existence of minimizers in the Euclidean case $M=\mathbb R^n$ , as established in [Reference Carrillo, Delgadino and Patacchini10, Theorem 4.1]. We also note that in [Reference Carrillo, Delgadino and Patacchini10], it is assumed that $\beta < n$ , as otherwise the interaction potential is non-integrable, and hence, the interaction energy is not bounded below.

Due to Proposition 2.1, the study in our article is restricted to potentials $h(\theta )$ that are less singular than $-\frac {1}{\theta ^{\min (n(m-1),n)}}$ at origin. For such potentials, let us show first that the energy from (2.2) is properly defined when the sectional curvatures of M are bounded below. For this purpose, we will use the following lemma and an HLS inequality on Cartan–Hadamard manifolds presented in Section 3.

Lemma 2.1 Let the interaction potential satisfy (H), for a function h that also satisfies

(2.8) $$ \begin{align} \lim_{\theta\to0+}\theta^\alpha h(\theta)=0, \qquad\text{and}\qquad \lim_{\theta\to\infty}h(\theta)\geq0, \end{align} $$

for some $0<\alpha <\min (n(m-1),n)$ . Then, there exist non-negative constants $\gamma _1$ and $\gamma _2$ such that

(2.9) $$ \begin{align} h(\theta)\geq -\gamma_1 \theta^{-\alpha}-\gamma_2,\qquad\text{ for all } \theta>0. \end{align} $$

Proof Fix a positive constant $\epsilon $ . If $h(\theta )\geq -\epsilon $ for all $\theta>0$ , then (2.9) holds for all $\theta>0$ with $\gamma _1=0$ and $\gamma _2=\epsilon $ .

If $h(\theta )<-\epsilon $ for some $\theta>0$ , then there exists $\theta _2>0$ such that

$$\begin{align*}h(\theta)\geq -\epsilon, \qquad\forall \theta\geq \theta_2, \end{align*}$$

as $\lim _{\theta \to \infty }h(\theta )\geq 0$ . From $\lim _{\theta \to 0+}\theta ^\alpha h(\theta )=0$ , we can also choose $\gamma>0$ and $\theta _1<\theta _2$ such that

$$\begin{align*}h(\theta)\geq -\gamma \theta^{-\alpha},\qquad\forall 0<\theta\leq\theta_1. \end{align*}$$

Since h is a lower-semicontinuous function, we can find

$$\begin{align*}C:=\inf_{\theta_1\leq \theta\leq \theta_2}\left(h(\theta)+\gamma\theta^{-\alpha}\right). \end{align*}$$

If $C\geq 0$ , then (2.9) holds with $\gamma _1=\gamma $ and $\gamma _2=\epsilon $ . If $C<0$ , then

$$\begin{align*}h(\theta)+(\gamma-C\theta_2^\alpha)\theta^{-\alpha}\geq C+(-C)\theta_2^\alpha \theta^{-\alpha}\geq C+(-C)=0,\qquad\forall~\theta_1\leq \theta\leq \theta_2. \end{align*}$$

For this case, (2.9) holds for all $\theta>0$ with $\gamma _1=\gamma -C\theta _2^\alpha $ and $\gamma _2=\epsilon $ .

Note that to prove Lemma 2.1, assuming the lower bound $\alpha>0$ is sufficient. We included however, the upper bound $\alpha <\min (n(m-1),n)$ as well, as this is the range of $\alpha $ needed to prove subsequent results, such as Proposition 2.2.

Consider first the case $\rho \in L^m(M) \cap \mathcal {P}(M)$ and show that the energy (2.2) is well-defined.

Proposition 2.2 Let $m>1$ and assume M satisfies assumption (M) and that its sectional curvatures are bounded below by a constant $-c_m<0$ – see (3.5). Let the interaction potential satisfy (H), for a function h that also satisfies (2.8). If $\rho \in L^m(M) \cap \mathcal {P}(M)$ , then $E[\rho ]$ given by (2.2) is well-defined.

Proof See Appendix B.

Now, consider the case $\rho \in \mathcal {P}(M)\backslash L^m(M)$ . If h satisfies (2.8), we can show that the energy is in fact infinite in this case, so such densities are not candidates for energy minimizers.

Proposition 2.3 Make the same assumptions as in Proposition 2.2. Then, if ${\rho \in \mathcal {P}(M)\backslash L^m(M)}$ , we have $E[\rho ]=\infty $ .

Proof See Appendix C.

We conclude the considerations above with the following remark.

Remark 2.3 Assume the manifold and the interaction potential satisfy the assumptions in Propositions 2.2 or 2.3. Then,

$$\begin{align*}\inf_{\rho\in\mathcal{P}(M)}E[\rho]=\inf_{\rho\in L^m(M)\cap\mathcal{P}(M)}E[\rho]. \end{align*}$$

2.4 Riesz rearrangement inequality on ${\mathbb H}^n$

We first present a rearrangement inequality on the hyperbolic space $\mathbb {H}^n$ [Reference Burchard and Schmuckenschläger5]. Rearrangement inequalities in $\mathbb R^n$ are discussed in detail in the classical monograph by Lieb and Loss [Reference Lieb and Loss40].

Fix a pole o on the hyperbolic space and take $A\subset {\mathbb H}^n$ a generic Borel set of finite Lebesgue measure. We define $A^{*}$ , the symmetric rearrangement (about the pole o) of the set A, as the open geodesic ball centred at o with the same volume as A, i.e.,

$$\begin{align*}A^{*}= B_r(o), \quad\text{with}\quad |B_r(o)|=|A|. \end{align*}$$

The symmetric-decreasing rearrangement of the characteristic function $\chi _A$ of the set A is defined as

$$\begin{align*}\chi_A^{*}=\chi_{A^{*}}. \end{align*}$$

For a Borel measurable function $f: {\mathbb H}^n \to \mathbb R$ , its symmetric-decreasing rearrangement (about the pole o) is the function $f^{*}:{\mathbb H}^n \to \mathbb R$ defined by

$$\begin{align*}f^{*}(x)=\int_0^\infty \chi^{*}_{|f|>t}(x)\mathrm{d} t. \end{align*}$$

The following theorem in [Reference Burchard and Schmuckenschläger5] generalizes the Riesz rearrangement inequality to ${\mathbb H}^n$ (the result is more general in fact, as it also applies to the sphere and it can work with a general number of functions – see [Reference Burchard and Schmuckenschläger5, Theorem 2]).

Theorem 2.5 (Riesz rearrangement inequality on ${\mathbb H}^n$ [Reference Burchard and Schmuckenschläger5])

Let $f_1$ and $f_2$ be two nonnegative measurable functions on ${\mathbb H}^n$ , which vanish at infinity, and $g:[0,\infty ) \to \mathbb R$ be a non-increasing function. Then, we have

$$\begin{align*}\iint\limits_{{\mathbb H}^n\times {\mathbb H}^n}f_1(x)g(d(x, y))f_2(y)\mathrm{d} x\mathrm{d} y\leq\iint\limits_{{\mathbb H}^n\times {\mathbb H}^n}f_1^{*}(x)g(d(x, y))f_2^{*}(y)\mathrm{d} x\mathrm{d} y, \end{align*}$$

where $f_1^{*}$ and $f_2^{*}$ are the symmetric-decreasing rearrangements of $f_1$ and $f_2$ about a fixed pole $o \in {\mathbb H}^n$ .

Our interest is to apply Theorem 2.5 for the interaction component of the energy in (2.2). Indeed, for an interaction potential that satisfies assumption (H) (note that h is non-decreasing, so $-h$ is non-increasing) and a density $\rho \in \mathcal {P}_{ac}({\mathbb H}^n)$ , we have

$$\begin{align*}\frac{1}{2}\iint\limits_{{\mathbb H}^n \times {\mathbb H}^n} h(d(x, y))\rho(x)\rho(y)\mathrm{d} x \mathrm{d} y \geq \frac{1}{2}\iint\limits_{{\mathbb H}^n \times {\mathbb H}^n} h(d(x, y))\rho^{*}(x)\rho^{*}(y)\mathrm{d} x \mathrm{d} y, \end{align*}$$

where $\rho ^{*}$ is the symmetric-decreasing rearrangement of $\rho $ .

On the other hand, decreasing rearrangements preserve the $L^p$ norm, for any $p \geq 1$ [Reference Lieb and Loss40, Chapter 3.3], which implies

$$\begin{align*}\int_{{\mathbb H}^n} \rho(x)^m\mathrm{d} x = \int_{{\mathbb H}^n} \rho^{*}(x)^m\mathrm{d} x. \end{align*}$$

Hence, by symmetric-decreasing rearrangements of the density, the entropy component of the energy remains the same, while the interaction component decreases. Combining the two observations, we then find that for an interaction potential that satisfies (H), the energy (2.2) decreases by symmetric-decreasing rearrangements of the density, i.e.,

$$\begin{align*}E[\rho] \geq E[\rho^{*}], \qquad \text{ for any } \rho \in\mathcal{P}_{ac}({\mathbb H}^n). \end{align*}$$

From these considerations, to find global minimizers of $E[\rho ]$ on $\mathcal {P}({\mathbb H}^n)$ , we can only look for densities that are radially symmetric with respect to a fixed point o, and decreasing along the radial direction. We will use the following notation:

$$ \begin{align*} \mathcal{P}_o^{sd}({\mathbb H}^n)=\{\rho\in\mathcal{P}_{ac}({\mathbb H}^n): \rho \text{ is radially symmetric and decreasing w.r.t pole } o\}. \end{align*} $$

3 HLS inequalities on Cartan–Hadamard manifolds

In this section, we derive several HLS-type inequalities on general Cartan–Hadamard manifolds. In particular, the inequalities hold for the hyperbolic space ${\mathbb H}^n$ , and they will be a key tool used in subsequent sections to show existence of global minimizers of the energy (2.2).

First, we list the following variation of HLS-type inequality on the Euclidean space $\mathbb R^n$ (see [Reference Carrillo, Delgadino and Patacchini10, Theorem 2.6] and [Reference Calvez, Carrillo and Hoffmann6, Theorem 3.1]).

Theorem 3.1 (Variation of HLS inequality on $\mathbb R^n$ [Reference Carrillo, Delgadino and Patacchini10])

Let $0<\lambda <n$ and denote ${m_c:=1+\frac {\lambda }{n}}$ . Then, for any $\rho \in L^{m}(\mathbb R^n) \cap \mathcal {P}(\mathbb R^n)$ , with $m \geq m_c$ , we have

(3.1) $$ \begin{align} \iint\limits_{\mathbb R^n \times \mathbb R^n} \frac{\mathrm{d}\rho(x)\mathrm{d}\rho(y)}{|x-y|^\lambda}\leq C(\lambda, n)\|\rho\|_{L^{m}}^{(1-\theta)m_c}, \end{align} $$

for a positive constant $C(\lambda , n)$ that depends on $\lambda $ and n, and $\theta \in [0,1)$ given by

$$\begin{align*}\theta = \frac{m- m_c}{m_c(m-1)}. \end{align*}$$

Theorem 3.1 is the statement of [Reference Carrillo, Delgadino and Patacchini10, Theorem 2.6]. The case $m=m_c$ ( $\theta =0$ ) was derived in [Reference Calvez, Carrillo and Hoffmann6, Theorem 3.1]. In fact, $\theta $ is an interpolation parameter, that satisfies

$$\begin{align*}\frac{1}{m_c} = \theta + \frac{1-\theta}{m}. \end{align*}$$

By using the expression for $\theta $ , inequality (3.1) can also be written as

$$ \begin{align*} \iint\limits_{\mathbb R^n \times \mathbb R^n} \frac{\mathrm{d}\rho(x)\mathrm{d}\rho(y)}{|x-y|^\lambda}\leq C(\lambda, n) \left( \int_{\mathbb R^n}\rho(x)^m \mathrm{d} x \right)^{\frac{\lambda}{n(m-1)}}, \end{align*} $$

which for $m=m_c$ it simply reduces to

(3.2) $$ \begin{align} \iint\limits_{\mathbb R^n \times \mathbb R^n} \frac{\mathrm{d}\rho(x)\mathrm{d}\rho(y)}{|x-y|^\lambda}\leq C(\lambda, n) \int_{\mathbb R^n}\rho(x)^{m_c} \mathrm{d} x. \end{align} $$

We will use inequality (3.2) to derive an HLS-type inequality on Cartan–Hadamard manifolds.

Let M be an n-dimensional Cartan–Hadamard manifold. Take a density $\rho \in \mathcal {P}_{ac}(M)$ and $\lambda \in (0,n)$ . Fix a pole $o\in M$ and denote by $f:M\to T_o M$ the Riemannian logarithm map at o, i.e.,

(3.3) $$ \begin{align} f(x)=\log_o x,\qquad \text{ for all }x\in M. \end{align} $$

Since M has non-positive curvature, consider the Rauch Comparison Theorem (see Theorem 2.2) in the following setup. Fix any two points $x,y \in M$ , take $\tilde {M} = T_o M$ , i to be the identity map, and c to be the geodesic curve joining x and y. The assumption on the curvatures $\mathcal {K}$ and $\tilde {\mathcal {K}}$ holds, as $\mathcal {K} \leq 0 = \tilde {\mathcal {K}}$ . The curve $\tilde {c}$ constructed in Theorem 2.2 joins $\log _o x$ and $\log _o y$ , and hence, its length is greater than or equal to the length of the straight segment joining $\log _o x$ and $\log _o y$ . Hence, by Theorem 2.2, we infer:

$$\begin{align*}|\log_o x - \log_o y| \leq l(\tilde{c}) \leq l(c) = d(x,y). \end{align*}$$

Using the notation (3.3), we write

(3.4) $$ \begin{align} |f(x)-f(y)| \leq d(x, y), \qquad \text{ for all } x,y \in M. \end{align} $$

Note that inequality (3.4) holds with equal sign for $M=\mathbb R^n$ .

The following lemma is a generalization of the HLS inequality (3.2) to Cartan–Hadamard manifolds with curvature bounded from below.

Proposition 3.1 (HLS-type inequality on Cartan–Hadamard manifolds - variant I)

Let M be an n-dimensional Cartan–Hadamard manifold with sectional curvatures that satisfy, for all $x\in M$ and all sections $\sigma \subset T_x M$ ,

(3.5) $$ \begin{align} -c_m\leq \mathcal{K}(x;\sigma)\leq 0, \end{align} $$

where $c_m$ is a positive constant. Let o be a fixed (but arbitrary) pole in M, let ${0<\lambda <n}$ and denote $m_c:=1+\frac {\lambda }{n}$ . Then, for any compactly supported density $\rho \in L^{m_c}(M) \cap \mathcal {P}(M)$ , we have

(3.6) $$ \begin{align} \iint\limits_{M\times M}\frac{\mathrm{d} \rho(x) \mathrm{d} \rho(y)}{d(x, y)^\lambda}\leq C(\lambda, n)\int_M\left(\frac{\sinh(\sqrt{c_m} \theta_x)}{\sqrt{c_m} \theta_x}\right)^{(n-1)(m_c-1)}\rho(x)^{m_c} \mathrm{d} x. \end{align} $$

for a positive constant $C(\lambda , n)$ that depends on $\lambda $ and n.

Proof By (3.4) and a change of variable, we find

(3.7) $$ \begin{align} \iint\limits_{M\times M}\frac{\mathrm{d}\rho(x) \mathrm{d}\rho(y)}{d(x, y)^\lambda} & \leq \iint\limits_{M\times M}\frac{\mathrm{d}\rho(x) \mathrm{d}\rho(y)}{|f(x)-f(y)|^\lambda} \nonumber \\ & =\iint\limits_{T_o M\times T_o M}\frac{\mathrm{d} f_\#\rho(u)\, \mathrm{d} f_\#\rho(v)}{|u-v|^\lambda}, \end{align} $$

where $f_\#\rho \in \mathcal {P}_{ac}(T_o M)$ denotes the pushforward (as measures) of $\rho $ by f. Note that

(3.8) $$ \begin{align} f_\#\rho(u)=\rho(f^{-1}(u))|J(f^{-1})(u)|, \qquad \text{ for all } u\in T_o M, \end{align} $$

where $J(f^{-1})$ denotes the Jacobian of the map $f^{-1}$ . The inverse map $f^{-1}:T_o M\to M$ is the Riemannian exponential map

$$ \begin{align*} f^{-1}(u)=\exp_o u, \qquad \text{ for all } u\in T_o M, \end{align*} $$

whose Jacobian was discussed in Section 2.2.

Since $T_o M$ can be identified with $\mathbb R^n $ , we can apply (3.2) to get

(3.9) $$ \begin{align} \iint\limits_{T_o M\times T_o M}\frac{\mathrm{d} f_\#\rho(u)\, \mathrm{d} f_\#\rho(v)}{|u-v|^\lambda}&\leq C(\lambda, n) \int_{T_o M}(f_\#\rho(u))^{m_c}\mathrm{d} u \nonumber \\ &=C(\lambda, n)\int_M |J(f^{-1})(f(x))|^{m_c-1}\rho(x)^{m_c} \mathrm{d} x, \end{align} $$

where and for the second line we used (3.8) and the change of variable $u = f(x)$ . We also note here since $\rho $ is compactly supported and in $L^{m_c}(M)$ , $f_\#\rho $ is in $L^{m_c}(T_o M)$ . This follows from the equality in (3.9), given that $|J(f^{-1})|$ is bounded on compact sets.

In (3.9), we use the upper bound on the Jacobian of the exponential map in Theorem 2.3 (note that $\|\log _o x\| = \theta _x$ ) to get

(3.10) $$ \begin{align} \iint\limits_{T_o M\times T_o M}\frac{\mathrm{d} f_\#\rho(u)\, \mathrm{d} f_\#\rho(v)}{|u-v|^\lambda}\leq C(\lambda, n)\int_M\left(\frac{\sinh(\sqrt{c_m} \theta_x)}{\sqrt{c_m} \theta_x}\right)^{(n-1)(m_c-1)} \rho(x)^{m_c} \mathrm{d} x. \end{align} $$

Finally, combine (3.7) and (3.10) to obtain (3.6).

Remark 3.1 Note that (3.6) holds for an arbitrary pole o, with $\theta _x$ denoting the distance from x to the pole. Inequality (3.6) can be regarded as a generalization to Cartan–Hadamard manifolds of (3.2), with an extra factor in the integrand, that corresponds to the bound on the Jacobian of the exponential map on M at o.

In the sequel, we will use the following classical result from analysis [Reference Evans and Gariepy18].

Lemma 3.1 (Vitali Covering Lemma)

Let $\mathcal {S}=\{B_{r_\alpha }(x_\alpha )\}_{\alpha \in \Lambda }$ be a family of balls in a separable metric space X such that

$$\begin{align*}\sup\{r_\alpha:\alpha\in \Lambda\}<\infty. \end{align*}$$

Then, there exists a countable subset $\Lambda '$ of $\Lambda $ such that $\{B_{r_\alpha }(x_\alpha )\}_{\alpha \in \Lambda '}$ are pairwise disjoint and satisfy

$$\begin{align*}\bigcup_{\alpha\in\Lambda}B_{r_\alpha}(x_\alpha)\subset \bigcup_{\alpha\in\Lambda'}B_{5r_\alpha}(x_\alpha). \end{align*}$$

Moreover, for any $\alpha \in \Lambda $ , there exists $\beta \in \Lambda '$ such that

$$\begin{align*}B_{r_\alpha}(x_\alpha)\subset B_{5r_\beta}(x_\beta). \end{align*}$$

For fixed $r>0$ , we consider the following setup of Lemma 3.1. Consider the family of balls in M given by

(3.11) $$ \begin{align} \mathcal{S}=\{B_r(x): x\in M\}. \end{align} $$

From Lemma 3.1, we know that there exists a countable subcollection

$$\begin{align*}\mathcal{S}'=\{B_r(x):x\in \mathcal{C}\} \end{align*}$$

for some countable subset $\mathcal {C}$ of M, such that

$$\begin{align*}M=\bigcup_{x\in M}B_r(x)\subset \bigcup_{x\in \mathcal{C}}B_{5r}(x). \end{align*}$$

In particular,

(3.12) $$ \begin{align} M=\bigcup_{x\in \mathcal{C}}B_{5r}(x), \end{align} $$

and any two balls $B_r(x_1)$ and $B_r(x_2)$ , with $x_1, x_2\in \mathcal {C}$ and $x_1\neq x_2$ , are disjoint.

Lemma 3.2 Let M be an n-dimensional Cartan–Hadamard manifold with sectional curvatures that satisfy (3.5), with $c_m>0$ . Let $r>0$ fixed and consider the countable subset $\mathcal {C}$ of M constructed by Vitali Covering Lemma from the covering $\mathcal {S}$ given by (3.11) (see (3.12)). For fixed $x\in M$ , denote

$$\begin{align*}\mathcal{S}_x:=\{y\in \mathcal{C}: x\in B_{6r}(y)\}. \end{align*}$$

Then,

(3.13) $$ \begin{align} \#(\mathcal{S}_x)\leq N(n,c_m, r), \end{align} $$

where $\#(\mathcal {S}_x)$ denotes the cardinal number of $\mathcal {S}_x$ , and

(3.14) $$ \begin{align} N(n,c_m, r):= \frac{n}{r^{n}}\int_0^{7r}\left(\frac{\sinh(\sqrt{c_m}s)}{\sqrt{c_m}}\right)^{n-1}\mathrm{d} s. \end{align} $$

Proof If $y\in \mathcal {S}_x$ , then we have $d(x, y)<6r$ , which implies $y\in B_{6r}(x)$ . Hence,

$$\begin{align*}\mathcal{S}_x\subset B_{6r}(x). \end{align*}$$

We also infer from here that for any $y \in \mathcal {S}_x\subset B_{6r}(x)$ , $B_r(y)$ is included in $B_{7r}(x)$ . Also note that by construction, $\{B_r(y)\}_{y\in \mathcal {S}_x}$ are disjoint. Hence, conclude that $\{B_r(y)\}_{y\in \mathcal {S}_x}$ are disjoint sets in $B_{7r}(x)$ . From this observation, we get

$$\begin{align*}\#(\mathcal{S}_x)\inf_{z\in M}|B_r(z)|\leq\sum_{y\in\mathcal{S}_x}|B_r(y)|\leq |B_{7r}(x)|\leq \sup_{z\in M}|B_{7r}(z)|. \end{align*}$$

We then find

(3.15) $$ \begin{align} \#(\mathcal{S}_x)\leq \frac{\sup_{z\in M}|B_{7r}(z)|}{\inf_{z\in M}|B_r(z)|}. \end{align} $$

From the comparison with volumes of balls in $\mathbb R^n$ (see (2.5)), we have

$$\begin{align*}\inf_{z\in M}|B_r(z)|\geq w(n)r^n, \end{align*}$$

while from the lower bound of curvatures and Theorem 2.4, we get

$$\begin{align*}\sup_{z\in M}|B_{7r}(z)|\leq n w(n)\int_0^{7r}\left(\frac{\sinh(\sqrt{c_m}s)}{\sqrt{c_m}}\right)^{n-1}\mathrm{d} s. \end{align*}$$

Using these bounds in (3.15), we find (3.13).

The main result of this section is the following theorem.

Theorem 3.2 (HLS-type inequality on Cartan–Hadamard manifolds - variant II)

Let M be an n-dimensional Cartan–Hadamard manifold whose sectional curvatures satisfy (3.5), for a positive constant $c_m$ . Let $\lambda \in (0, n)$ , $m\geq m_c$ be given (recall $m_c:=1+\frac {\lambda }{n}$ ), and $r>0$ fixed. Then, for any $\rho \in L^m(M) \cap \mathcal {P}(M)$ , we have

$$\begin{align*}\iint\limits_{M\times M}\frac{\mathrm{d}\rho(x) \mathrm{d}\rho(y)}{d(x, y)^\lambda}\leq r^{-\lambda}+\tilde{C}(\lambda, n, r, c_m)\left(\int_M \rho(x)^m\mathrm{d} x\right)^{\frac{\lambda}{n(m-1)}}, \end{align*}$$

where the constant $\tilde {C}$ is given by

(3.16) $$ \begin{align} \tilde{C}(\lambda, n, r, c_m)=C(\lambda, n)\left(\frac{\sinh(6\sqrt{c_m}r)}{6\sqrt{c_m}r}\right)^{(n-1)(m_c-1)}N(n,c_m, r). \end{align} $$

In (3.16), $C(\lambda ,n)$ is the constant introduced in Theorem 3.1, and $N(n,c_m, r)$ is given by (3.14).

Proof Take $x, y$ to be two points in M which satisfy $d(x, y)<r$ . Since $x\in M$ and $M=\bigcup _{z\in \mathcal {C}}B_{5r}(z)$ , there exists $z\in \mathcal {C}$ such that $x\in B_{5r}(z)$ . Then,

$$\begin{align*}d(z, y)\leq d(z, x)+d(x, y)<5r+r=6r, \end{align*}$$

which implies $y\in B_{6r}(z)$ . Since $x\in B_{5r}(z)\subset B_{6r}(z)$ , we have

$$\begin{align*}(x, y)\in B_{6r}(z)\times B_{6r}(z). \end{align*}$$

This implies

(3.17) $$ \begin{align} \{(x, y)\in M\times M: d(x, y)<r\}\subset \bigcup_{z\in \mathcal{C}}B_{6r}(z)\times B_{6r}(z). \end{align} $$

Use (3.17) to estimate

(3.18) $$ \begin{align} \iint\limits_{M\times M}\frac{\mathrm{d}\rho(x)\mathrm{d}\rho(y)}{d(x, y)^\lambda}& = \iint\limits_{d(x, y)\geq r}\frac{\mathrm{d}\rho(x)\mathrm{d}\rho(y)}{d(x, y)^\lambda}+\iint\limits_{d(x, y)<r}\frac{\mathrm{d}\rho(x)\mathrm{d}\rho(y)}{d(x, y)^\lambda} \nonumber \\ &\leq r^{-\lambda}+\sum_{z\in\mathcal{C}}\iint\limits_{B_{6r}(z)\times B_{6r}(z)}\frac{\mathrm{d}\rho(x)\mathrm{d}\rho(y)}{d(x, y)^\lambda}. \end{align} $$

We remove all elements $z\in \mathcal {C}$ which satisfy

$$\begin{align*}\int_{B_{6r}(z)}\mathrm{d} \rho(x)=0, \end{align*}$$

and define $\mathcal {C}' \subset \mathcal {C}$ by

$$\begin{align*}\mathcal{C}'=\mathcal{C}\cap\left\{z\in M:\int_{B_{6r}(z)}\mathrm{d} \rho(x)>0\right\}. \end{align*}$$

Then, by (3.18) we have

(3.19) $$ \begin{align} \iint\limits_{M\times M}\frac{\mathrm{d}\rho(x)\mathrm{d}\rho(y)}{d(x, y)^\lambda}&\leq r^{-\lambda}+\sum_{z\in\mathcal{C}'}\iint\limits_{B_{6r}(z)\times B_{6r}(z)}\frac{\mathrm{d}\rho(x)\mathrm{d}\rho(y)}{d(x, y)^\lambda}. \end{align} $$

Now, for all $z\in \mathcal {C}'$ we consider the normalized restrictions of measure $\rho $ on $B_{6r}(z)$ , given by

$$\begin{align*}\rho_z(x)=\begin{cases} \displaystyle\frac{\rho(x)}{\|\rho\|_{L^1(B_{6r}(z))}}\qquad&\text{ when }x\in B_{6r}(z),\\[10pt] \displaystyle 0&\text{ otherwise}. \end{cases} \end{align*}$$

Note that $\rho _z$ is compactly supported, for all z. Fix $z \in \mathcal {C}'$ and substitute $\rho _z$ into (3.6) with z as pole (use $\rho _z$ in place of $\rho $ , and z in place of o in Proposition 3.1 – see also Remark 3.1), to get

(3.20) $$ \begin{align} & \frac{1}{\|\rho\|_{L^1(B_{6r}(z))}^2}\iint\limits_{B_{6r}(z)\times B_{6r}(z)}\frac{\mathrm{d}\rho(x) \mathrm{d}\rho(y)}{d(x, y)^\lambda}= \iint\limits_{M\times M}\frac{\mathrm{d}\rho_z(x) \mathrm{d}\rho_z(y)}{d(x, y)^\lambda} \nonumber \\ &\hspace{2cm} \leq C(\lambda, n)\int_M \left(\frac{\sinh(\sqrt{c_m} d(x, z))}{\sqrt{c_m} d(x, z)}\right)^{(n-1)(m_c-1)}\rho_z(x)^{m_c} \mathrm{d} x \nonumber \\ &\hspace{2cm} =\frac{C(\lambda, n)}{\|\rho\|_{L^1(B_{6r}(z))}^{m_c}}\int_{B_{6r}(z)}\left(\frac{\sinh(\sqrt{c_m}d(x, z))}{\sqrt{c_m}d(x, z)}\right)^{(n-1)(m_c-1)}\rho(x)^{m_c}\mathrm{d} x. \end{align} $$

Since

$$\begin{align*}\frac{\mathrm{d}}{\mathrm{d}\theta}\left(\frac{\sinh\theta}{\theta}\right)= \frac{\cosh\theta}{\theta^2}(\theta-\tanh\theta), \end{align*}$$

and $\theta \geq \tanh \theta $ for all $\theta \geq 0$ , we know that $\frac {\sinh \theta }{\theta }$ is non-decreasing for all $\theta \geq 0$ . From this property, we get

$$\begin{align*}\frac{\sinh(\sqrt{c_m}d(x, z))}{\sqrt{c_m}d(x, z)}\leq \frac{\sinh(6\sqrt{c_m}r)}{6\sqrt{c_m}r}, \qquad \text{ for all } x\in B_{6r}(z). \end{align*}$$

Then, multiply (3.20) by $\|\rho \|_{L^1(B_{6r}(z))}^2$ and use the inequality above to find

$$\begin{align*}\iint\limits_{B_{6r}(z)\times B_{6r}(z)}\frac{\mathrm{d}\rho(x) \mathrm{d}\rho(y)}{d(x, y)^\lambda}\leq \frac{C(\lambda, n)}{\|\rho\|_{L^1(B_{6r}(z))}^{m_c-2}}\left(\frac{\sinh(6\sqrt{c_m}r)}{6\sqrt{c_m}r}\right)^{(n-1)(m_c-1)}\int_{B_{6r}(z)}\rho(x)^{m_c}\mathrm{d} x. \end{align*}$$

Furthermore, use $\|\rho \|_{L^1(B_{6r}(z))}\leq \|\rho \|_{L^1(M)}=1$ and $m_c-2=\frac {\lambda }{n}-1<0$ , to get

(3.21) $$ \begin{align} \iint\limits_{B_{6r}(z)\times B_{6r}(z)}\frac{\mathrm{d}\rho(x) \mathrm{d}\rho(y)}{d(x, y)^\lambda}\leq C(\lambda, n)\left(\frac{\sinh(6\sqrt{c_m}r)}{6\sqrt{c_m}r}\right)^{(n-1)(m_c-1)}\int_{B_{6r}(z)}\rho(x)^{m_c}\mathrm{d} x. \end{align} $$

Finally, use (3.21) in (3.19) to get

(3.22) $$ \begin{align} \iint\limits_{M\times M}\frac{\mathrm{d}\rho(x)\mathrm{d}\rho(y)}{d(x, y)^\lambda} \leq r^{-\lambda}+C(\lambda, n)\left(\frac{\sinh(6\sqrt{c_m}r)}{6\sqrt{c_m}r}\right)^{(n-1)(m_c-1)}\sum_{z\in \mathcal{C}'}\int_{B_{6r}(z)}\rho(x)^{m_c}\mathrm{d} x. \end{align} $$

By Lemma 3.2, for any $x\in M$ there are at most $N(n,c_m, r)$ balls $\{B_{6r}(z)\}_{z\in \mathcal {C}'}$ which contain x. This implies

$$\begin{align*}\sum_{z\in\mathcal{C}'}\chi_{B_{6r}(z)}\leq N(n,c_m, r)\chi_{M}, \end{align*}$$

where $\chi _A$ is the characteristic function of the set A, defined as

$$\begin{align*}\chi_A(x)=\begin{cases} 1&\quad\text{when }x\in A,\\ 0&\quad\text{when }x\not\in A. \end{cases} \end{align*}$$

Then, we have

(3.23) $$ \begin{align} \sum_{z\in\mathcal{C}'}\int_{B_{6r}(z)}\rho(x)^{m_c}\mathrm{d} x&=\sum_{z\in\mathcal{C}'}\int_{M}\chi_{B_{6r}(z)}(x)\rho(x)^{m_c}\mathrm{d} x=\int_M\left(\sum_{z\in\mathcal{C}'}\chi_{B_{6r}(z)}(x)\right)\rho(x)^{m_c}\mathrm{d} x \nonumber \\ &\leq \int_M N(n,c_m, r)\rho(x)^{m_c}\mathrm{d} x=N(n,c_m, r)\int_M \rho(x)^{m_c}\mathrm{d} x. \end{align} $$

Finally, using (3.23) in (3.22), we arrive at

(3.24) $$ \begin{align} \iint\limits_{M\times M}\frac{\mathrm{d}\rho(x)\mathrm{d}\rho(y)}{d(x, y)^\lambda}\leq r^{-\lambda}+C(\lambda, n)\left(\frac{\sinh(6\sqrt{c_m}r)}{6\sqrt{c_m}r}\right)^{(n-1)(m_c-1)}N(n,c_m, r)\int_M \rho(x)^{m_c}\mathrm{d} x. \end{align} $$

For $m>m_c$ , the Hölder inequality with $p=\frac {m-1}{m-m_c}$ and $q=\frac {m-1}{m_c-1}$ , yields

$$\begin{align*}\int_M\rho(x)^{m_c}\mathrm{d} x\leq \left(\int_M\rho(x)\mathrm{d} x\right)^{\frac{m-m_c}{m-1}} \left(\int_M\rho(x)^{m}\mathrm{d} x\right)^{\frac{m_c-1}{m-1}}. \end{align*}$$

Since $\rho $ integrates to $1$ , this inequality together with (3.24) lead to the conclusion of the theorem.

Remark 3.2 The relationship between m, n, and $\lambda $ is $m \geq 1+\frac {\lambda }{n}$ and $0<\lambda <n$ . Hence, the range of $\lambda $ is

$$\begin{align*}\begin{cases} 0<\lambda\leq n(m-1),&\qquad\text{when }1<m<2,\\[2pt] 0<\lambda<n,&\qquad\text{when }m\geq 2. \end{cases} \end{align*}$$

For both cases, $0<\lambda <\min (n, n(m-1))$ .

4 Lower-semicontinuity of the energy functional on ${\mathbb H}^n$

From here on in this article, we consider the manifold M to be the hyperbolic space ${\mathbb H}^n$ of constant sectional curvature $-$ 1. Note that ${\mathbb H}^n$ is a Cartan–Hadamard manifold which is also a homogeneous manifold. In this section, we study the lower semicontinuity of the entropy and the interaction energy, under suitable conditions on the interaction potential.

We note that in working with a minimizing sequence $\{\rho _k\}_{k\in \mathbb {N}}$ of the energy functional, one can assume without loss of generality that the energy values $\{E[\rho _k]\}_{k\in \mathbb {N}}$ are non-increasing. Indeed, this property can be achieved by passing to subsequence of the minimizing sequence.

Proposition 4.1 Let $m>1$ and consider the energy functional (2.2) on ${\mathbb H}^n$ , with an interaction potential that satisfies assumption (H), and in addition, that

(4.1) $$ \begin{align} \lim_{\theta\to0+}\theta^\alpha h(\theta)=0, \quad \qquad \text{ for some } 0<\alpha<\min(n(m-1), n). \end{align} $$

Let $\{\rho _k\}_{k\in \mathbb {N}}\subset L^m({\mathbb H}^n) \cap \mathcal {P}({\mathbb H}^n)$ be a minimizing sequence of $E[\rho ]$ . Then, for any ${0\leq \beta <\min (n(m-1), n)}$ ,

(4.2) $$ \begin{align} \int_{{\mathbb H}^n}\rho_k(x)^m\mathrm{d} x,\quad \iint\limits_{ {\mathbb H}^n \times {\mathbb H}^n}\frac{\rho_k(x)\rho_k(y)\mathrm{d} x\mathrm{d} y}{d(x, y)^{\beta}}, \quad \text{ and} \quad\! \iint\limits_{{\mathbb H}^n\times {\mathbb H}^n}h(d(x, y))\rho_k(x)\rho_k(y)\mathrm{d} x\mathrm{d} y \end{align} $$

are uniformly bounded (from above and below) in k. Also,

$$\begin{align*}\lim_{\delta\to0+}\iint\limits_{d(x, y)<\delta}h(d(x, y))\rho_k(x)\rho_k(y)\mathrm{d} x\mathrm{d} y=0, \end{align*}$$

uniformly in k.

Proof Fix $0\leq \beta <\min (n(m-1), n)$ , and define

$$\begin{align*}\tilde{\alpha}=\max\left(\frac{1}{2}(\alpha+\min(n(m-1), n)), \beta\right), \end{align*}$$

where $\alpha $ is the constant from (4.1). Then, $0<\alpha <\tilde {\alpha }<\min (n(m-1), n)$ , and hence

$$\begin{align*}\lim_{\theta\to0+}h(\theta)\theta^{\tilde{\alpha}}=\lim_{\theta\to0+}\left(h(\theta)\theta^\alpha\times \theta^{\tilde{\alpha}-\alpha}\right)=0. \end{align*}$$

From the assumptions on h (see (H), (2.1) and (4.1)) and Lemma 2.1, there exist ${\gamma _1,\gamma _2\geq 0}$ such that

$$\begin{align*}h(\theta)\geq -\gamma_1\theta^{-\tilde{\alpha}}-\gamma_2,\qquad\forall \theta>0. \end{align*}$$

Then, given that $\{\rho _k\}_{k\in \mathbb {N}}$ is a minimizing sequence, we have

(4.3) $$ \begin{align} \begin{aligned} E[\rho_1] &\geq E[\rho_k] \\ & = \frac{1}{m-1}\int_{{\mathbb H}^n}\rho_k(x)^m\mathrm{d} x+\frac{1}{2}\iint\limits_{{\mathbb H}^n \times {\mathbb H}^n} h(d(x, y))\rho_k(x)\rho_k(y)\mathrm{d} x\mathrm{d} y \\ & \geq\frac{1}{m-1}\int_{{\mathbb H}^n}\rho_k(x)^m\mathrm{d} x-\frac{\gamma_1}{2}\iint\limits_{{\mathbb H}^n \times {\mathbb H}^n}\frac{\rho_k(x)\rho_k(y)\mathrm{d} x\mathrm{d} y}{d(x, y)^{\tilde{\alpha}}} - \frac{\gamma_2}{2}. \end{aligned} \end{align} $$

Fix $r>0$ arbitrary. By using the HLS inequality on Cartan–Hadamard manifolds (Theorem 3.2) for $M={\mathbb H}^n$ , $c_m=1$ and $\lambda =\tilde {\alpha ,}$ we find

(4.4) $$ \begin{align} \iint\limits_{{\mathbb H}^n\times {\mathbb H}^n}\frac{\mathrm{d} \rho_k(x)\mathrm{d} \rho_k(y)}{d(x, y)^{\tilde{\alpha}}}\leq r^{-\tilde{\alpha}}+\tilde{C}(\tilde{\alpha}, n, r, 1)\left(\int_{{\mathbb H}^n} \rho_k(x)^m\mathrm{d} x\right)^{\frac{\tilde{\alpha}}{n(m-1)}}. \end{align} $$

Combine (4.3) and (4.4) to get

$$\begin{align*}E[\rho_1]\geq \frac{1}{m-1}\int_{{\mathbb H}^n}\rho_k(x)^m\mathrm{d} x-\frac{\gamma_1\tilde{C}(\tilde{\alpha}, n, r, 1)}{2}\left(\int_{{\mathbb H}^n} \rho_k(x)^m\mathrm{d} x\right)^{\frac{\tilde{\alpha}}{n(m-1)}}-\frac{1}{2}(\gamma_1 r^{-\tilde{\alpha}}+\gamma_2). \end{align*}$$

Since $0<\frac {\tilde {\alpha }}{n(m-1)}<1$ , if $\int _{{\mathbb H}^n}\rho _k(x)^m\mathrm {d} x$ tends to $\infty $ as $k \to \infty $ , then the right hand side of the above inequality also tends to $\infty $ . Therefore, we conclude that $ \int _{{\mathbb H}^n}\rho _k(x)^m\mathrm {d} x $ is bounded from above uniformly in $k\in \mathbb {N}$ , which shows the statement for the first term in (4.2); note that a uniform lower bound is immediate, as the expressions are non-negative.

Given that $ \int _{{\mathbb H}^n}\rho _k(x)^m\mathrm {d} x $ are uniformly bounded from above, from (4.4), we then get the uniform boundedness in $k \in \mathbb {N}$ of $\iint\limits _{{\mathbb H}^n \times {\mathbb H}^n}\frac {\rho _k(x)\rho _k(y)\mathrm {d} x\mathrm {d} y}{d(x, y)^{\tilde {\alpha }}}$ . This allows us to define the uniform upper bound $0<U<\infty $ as

$$\begin{align*}U:=\sup_{k\in\mathbb{N}}\iint\limits_{{\mathbb H}^n\times {\mathbb H}^n}\frac{\rho_k(x)\rho_k(y)\mathrm{d} x\mathrm{d} y}{d(x, y)^{\tilde{\alpha}}}. \end{align*}$$

Since $\tilde {\alpha }\geq \beta $ , we can apply the Hölder inequality to find

$$\begin{align*}&\iint\limits_{{\mathbb H}^n \times {\mathbb H}^n}\frac{\rho_k(x)\rho_k(y)\mathrm{d} x\mathrm{d} y}{d(x, y)^{\beta}}\\ &\quad\leq \left(\iint\limits_{{\ \mathbb H}^n \times {\mathbb H}^n}\frac{\rho_k(x)\rho_k(y)\mathrm{d} x\mathrm{d} y}{d(x, y)^{\tilde{\alpha}}}\right)^{\frac{\beta}{\tilde{\alpha}}}\left(\iint\limits_{\ {\mathbb H}^n \times {\mathbb H}^n}\rho_k(x)\rho_k(y)\mathrm{d} x\mathrm{d} y\right)^{\frac{\tilde{\alpha}-\beta}{\tilde{\alpha}}}. \end{align*}$$

In the above, use $\rho _k\in \mathcal {P}({\mathbb H}^n)$ and the definition of U to further get

$$\begin{align*}&\iint\limits_{{\mathbb H}^n \times {\mathbb H}^n}\frac{\rho_k(x)\rho_k(y)\mathrm{d} x\mathrm{d} y}{d(x, y)^{\beta}}\leq U^{\frac{\beta}{\tilde{\alpha}}}. \end{align*}$$

Conclude then that $\iint\limits _{{\mathbb H}^n \times {\mathbb H}^n}\frac {\rho _k(x)\rho _k(y)\mathrm {d} x\mathrm {d} y}{d(x, y)^{\beta }}$ is uniformly bounded from above in $k\in \mathbb {N}$ , for arbitrary $0\leq \beta <\min (n(m-1), n)$ . This shows that the second term in (4.2) is uniformly bounded from above in $k \in \mathbb {N}$ , and since the expressions are non-negative, it is also bounded uniformly from below.

It remains to argue about $ \iint\limits _{{\mathbb H}^n \times {\mathbb H}^n} h(d(x, y))\rho _k(x)\rho _k(y)\mathrm {d} x\mathrm {d} y$ , the third term in (4.2). Its uniform boundedness from above follows immediately (see (4.3)) as

$$ \begin{align*} E[\rho_1]&\geq E[\rho_k]\geq \frac{1}{2}\iint\limits_{{\mathbb H}^n \times {\mathbb H}^n } h(d(x, y))\rho_k(x)\rho_k(y)\mathrm{d} x\mathrm{d} y. \end{align*} $$

We will now show that it is also uniformly bounded from below.

From assumption (4.1) on h and Lemma 2.1, there exist $\gamma _1', \gamma _2'\geq 0$ such that

$$\begin{align*}h(\theta)\geq -\gamma_1'\theta^{-\alpha}-\gamma_2', \qquad \forall \theta>0. \end{align*}$$

If h is singular at $0+$ , then there exists $\delta _0>0$ such that $h(\theta )<0$ for any $0<\theta <\delta _0$ . Hence, for any $0<\delta <\delta _0$ , we have

$$ \begin{align*} &\Biggl|\ \iint\limits_{d(x, y)<\delta}h(d(x, y))\rho_k(x)\rho_k(y) \mathrm{d} x\mathrm{d} y\Biggr|=\iint\limits_{d(x, y)<\delta}|h(d(x, y))|\mathrm{d}\rho_k(x)\mathrm{d}\rho_k(y)\\ &\quad\leq \iint\limits_{d(x, y)<\delta}(\gamma_1'd(x, y)^{-\alpha}+\gamma_2')\rho_k(x)\rho_k(y)\mathrm{d} x\mathrm{d} y\\ &\quad\leq \iint\limits_{d(x, y)<\delta}(\gamma_1'+\gamma_2'\delta^\alpha )d(x, y)^{-\alpha}\rho_k(x)\rho_k(y)\mathrm{d} x\mathrm{d} y\\ &\quad\leq (\gamma_1'+\gamma_2'\delta^\alpha )\iint\limits_{d(x, y)<\delta}d(x, y)^{-\tilde{\alpha}}\delta^{\tilde{\alpha}-\alpha}\rho_k(x)\rho_k(y)\mathrm{d} x\mathrm{d} y\\ &\quad\leq (\gamma_1'+\gamma_2'\delta^\alpha ) U\delta^{\tilde{\alpha}-\alpha}. \end{align*} $$

We conclude from the above (note that $\alpha <\tilde {\alpha }$ ) that

(4.5) $$ \begin{align} \lim_{\delta\to0+}\iint\limits_{d(x, y)<\delta}h(d(x, y))\rho_k(x)\rho_k(y)\mathrm{d} x\mathrm{d} y=0, \end{align} $$

uniformly in $k \in \mathbb {N}$ . On the other hand, if h is not singular at $0$ , showing (4.5) is immediate.

Now fix $\epsilon>0$ . By (4.5), there exists $\delta _1>0$ such that

$$ \begin{align*} \Biggl|\iint\limits_{d(x, y)<\delta_1}h(d(x, y))\rho_k(x)\rho_k(y) \mathrm{d} x \mathrm{d} y\Biggr|<\epsilon, \qquad\forall k\in\mathbb{N}. \end{align*} $$

Then, compute and estimate using the monotonicity of h:

$$ \begin{align*} & \iint\limits_{{\mathbb H}^n \times {\mathbb H}^n} h(d(x, y))\rho_k(x)\rho_k(y)\mathrm{d} x\mathrm{d} y \\ &\qquad =\iint\limits_{d(x, y)<\delta_1} h(d(x, y))\rho_k(x)\rho_k(y)\mathrm{d} x\mathrm{d} y+\iint\limits_{d(x, y)\geq\delta_1} h(d(x, y))\rho_k(x)\rho_k(y)\mathrm{d} x\mathrm{d} y\\ &\qquad \geq-\epsilon+\iint\limits_{d(x, y)\geq \delta_1}h(\delta_1)\rho_k(x)\rho_k(y)\mathrm{d} x\mathrm{d} y\\ &\qquad \geq-\epsilon+\min(0, h(\delta_1)), \end{align*} $$

which yields that $ \iint\limits _{{\mathbb H}^n \times {\mathbb H}^n} h(d(x, y))\rho _k(x)\rho _k(y)\mathrm {d} x\mathrm {d} y$ is uniformly bounded from below in $k\in \mathbb {N}$ .

Remark 4.1 Using the uniform bound from above of $\int _{{\mathbb H}^n}\rho _k^m(x)\mathrm {d} x$ , one can show that $\{\rho _k\}_{k \geq 1}$ is uniformly integrable. Indeed, from

$$\begin{align*}\int_{\{x:\rho_k(x)\geq Q\}}\rho_k(x)\mathrm{d} x\leq \frac{1}{Q^{m-1}}\int_{{\mathbb H}^n}\rho_k^m(x)\mathrm{d} x, \end{align*}$$

we get

$$\begin{align*}\lim_{Q\to\infty}\sup_{k}\int_{\{x:\rho_k(x)\geq Q\}}\rho_k(x)\mathrm{d} x=0, \end{align*}$$

which yields the uniform integrability.

The main result of this section is given by the following proposition.

Proposition 4.2 (Weak lower semicontinuity of the energy)

Let $m>1$ and consider an interaction potential that satisfies the assumptions in Proposition 4.1. Let $\{\rho _k\}_{k\in \mathbb {N}}\subset L^m({\mathbb H}^n) \cap \mathcal {P}_o^{sd}({\mathbb H}^n)$ be a minimizing sequence of the energy functional (2.2). Then, if $\{\rho _k\}$ converges weakly- $^{*}$ to $\rho _0$ , then

$$\begin{align*}\liminf_{k\to\infty}E[\rho_k]\geq E[\rho_0]. \end{align*}$$

Proof We will investigate the entropy and the interaction energy separately.

Part 1: Entropy. In this part, we study the lower semicontinuity of the entropy functional

$$\begin{align*}\frac{1}{m-1}\int_{{\mathbb H}^n}\rho(x)^m\mathrm{d} x. \end{align*}$$

We use again $f(x)=\log _o x$ (see (3.3) and (3.8)) with o a fixed pole in ${\mathbb H}^n$ , to transform

$$\begin{align*}\int_{{\mathbb H}^n}\rho(x)^m\mathrm{d} x=\int_{T_o {\mathbb H}^n}\left(\frac{f_\#\rho(u)}{J(f^{-1}(u))}\right)^m J(f^{-1}(u))\mathrm{d} u. \end{align*}$$

Here, $\mathrm {d} u$ denotes the Lebegue measure on ${T_o {\mathbb H}^n}$ , which is identified to $\mathbb R^n $ . In the notations of [Reference Ambrosio, Fusco and Pallara1, Section 2.6], the r.h.s. above can be written as

$$\begin{align*}G (\nu, \mu) = \int_{T_o {\mathbb H}^n}U \left(\frac{\nu }{\mu}\right) \mathrm{d} \mu(u), \end{align*}$$

where

$$\begin{align*}U(s)=s^m,\quad \nu=f_\#\rho(u)\mathrm{d} u,\quad \text{ and } \mu=J(f^{-1}(u))\mathrm{d} u. \end{align*}$$

Hence, for the sequence $\rho _k$ we can write

$$\begin{align*}\int_{{\mathbb H}^n}\rho_k(x)^m\mathrm{d} x = G (\nu_k,\mu), \end{align*}$$

where $\nu _k=f_\#\rho _k(u)\mathrm {d} u$ . Note that since $\rho _k$ converges weakly- $^{*}$ to $\rho _0$ , $\nu _k$ converges weakly- $^{*}$ to $\nu _0:=f_\#\rho _0(u)\mathrm {d} u$ . By [Reference Ambrosio, Fusco and Pallara1, Theorem 2.34], identifying $T_o {\mathbb H}^n$ with $\mathbb R^n$ , and using that U is convex and lower semicontinuous, we have

$$\begin{align*}\liminf_{k\to \infty} G (\nu_k,\mu) \geq G (\nu_0,\mu), \end{align*}$$

which then implies the lower-semicontinuity of the entropy component:

(4.6) $$ \begin{align} \liminf_{k\to\infty}\frac{1}{m-1}\int_{{\mathbb H}^n}\rho_k(x)^m\mathrm{d} x\geq \frac{1}{m-1}\int_{{\mathbb H}^n}\rho_0(x)^m\mathrm{d} x. \end{align} $$

Part 2: Interaction energy. We now study the lower semicontinuity of the interaction energy functional

$$\begin{align*}\iint\limits_{{\mathbb H}^n \times {\mathbb H}^n} h(d(x, y))\rho_k(x)\rho_k(y)\mathrm{d} x\mathrm{d} y. \end{align*}$$

Fix $\epsilon>0$ . For $0<\delta <1$ (to be specified later), and $k\geq 1$ fixed, we divide the domain of the integration as follows:

$$ \begin{align*} &\iint\limits_{{\mathbb H}^n\times {\mathbb H}^n} h(d(x, y))\rho_k(x)\rho_k(y)\mathrm{d} x\mathrm{d} y\\ & \quad =\iint\limits_{d(x, y)<\delta} h(d(x, y))\rho_k(x)\rho_k(y)\mathrm{d} x\mathrm{d} y +\iint\limits_{d(x, y)>\delta^{-1}} h(d(x, y))\rho_k(x)\rho_k(y)\mathrm{d} x\mathrm{d} y\\ & \qquad +\iint\limits_{\delta\leq d(x, y)\leq \delta^{-1}} h(d(x, y))\rho_k(x)\rho_k(y)\mathrm{d} x\mathrm{d} y \\ & \quad =:\mathcal{I}_1+\mathcal{I}_2+\mathcal{I}_3. \end{align*} $$

We will investigate the three terms $\mathcal {I}_1$ , $\mathcal {I}_2,$ and $\mathcal {I}_3$ separately.

Estimate of $\mathcal {I}_1$ . We have already investigated $\mathcal {I}_1$ in Proposition 4.1. By the uniform (in k) convergence in (4.5), we can choose $\delta _1>0$ such that

(4.7) $$ \begin{align} -\epsilon<\mathcal{I}_1<\epsilon, \qquad \text{ for any } \delta\in(0, \delta_1). \end{align} $$

Estimate of $\mathcal {I}_2$ . Recall that we only have two cases: $h_\infty =0$ and $h_\infty =\infty $ . Since h is increasing and converges to $h_\infty $ , there exists $\delta _2>0$ such that $-\epsilon <h(\theta )$ for any $\theta>\delta _2^{-1}$ . Hence, for any $\delta \in (0, \delta _2)$ , we can estimate $\mathcal {I}_2$ as

(4.8) $$ \begin{align} \mathcal{I}_2\geq \iint\limits_{d(x, y)>\delta^{-1}}(-\epsilon)\rho_k(x)\rho_k(y)\mathrm{d} x \mathrm{d} y\geq -\epsilon. \end{align} $$

Note that both (4.7) and (4.8) hold for arbitrary $k \geq 1$ fixed, while $\delta _1$ and $\delta _2$ depend on $\epsilon $ , but do not depend on k.

Estimate of $\mathcal {I}_3$ . Since $\rho _k$ is radially symmetric and decreasing, for any $x\in {\mathbb H}^n,$ we have

$$\begin{align*}1\geq \int_{B_{\theta_x}(o)}\rho_k(y)\mathrm{d} y\geq \int_{B_{\theta_x}(o)}\rho_k(x)\mathrm{d} y=\rho_k(x)|B_{\theta_x}(o)|. \end{align*}$$

This implies

$$\begin{align*}\rho_k(x)\leq |B_{\theta_x}(o)|^{-1},\qquad \text{ for all } x\in {\mathbb H}^n, \text{ and } k \geq 1, \end{align*}$$

which yields

(4.9) $$ \begin{align} \|\rho_k\|_{L^\infty(B_R(o)^c)}\leq |B_R(o)|^{-1},\qquad \text{ for all } R>0, \text{ and } k \geq 1. \end{align} $$

Note that in these considerations we used implicitly the Riesz rearrangement property on ${\mathbb H}^n$ (Theorem 2.5), which allows us to consider only radially symmetric and decreasing minimizing sequences $\rho _k$ ; this is in fact the only place in the article where we use Theorem 2.5.

Fix $R>0$ . We divide $\mathcal {I}_3$ as follows:

$$ \begin{align*} \mathcal{I}_3&=\iint\limits_{\substack{\delta\leq d(x, y)\leq \delta^{-1}\\\theta_x,\theta_y\leq R}} h(d(x, y))\rho_k(x)\rho_k(y)\mathrm{d} x\mathrm{d} y+\iint\limits_{\substack{\delta\leq d(x, y)\leq \delta^{-1}\\\theta_x> R}} h(d(x, y))\rho_k(x)\rho_k(y)\mathrm{d} x\mathrm{d} y\\ &\iint\limits_{\substack{\delta\leq d(x, y)\leq \delta^{-1}\\\theta_y> R}} h(d(x, y))\rho_k(x)\rho_k(y)\mathrm{d} x\mathrm{d} y-\iint\limits_{\substack{\delta\leq d(x, y)\leq \delta^{-1}\\\theta_x,\theta_y> R}} h(d(x, y))\rho_k(x)\rho_k(y)\mathrm{d} x\mathrm{d} y\\ &=\iint\limits_{\substack{\delta\leq d(x, y)\leq \delta^{-1}\\\theta_x,\theta_y\leq R}} h(d(x, y))\rho_k(x)\rho_k(y)\mathrm{d} x\mathrm{d} y+2\!\iint\limits_{\substack{\delta\leq d(x, y)\leq \delta^{-1}\\\theta_x> R}} h(d(x, y))\rho_k(x)\rho_k(y)\mathrm{d} x\mathrm{d} y\\ &-\iint\limits_{\substack{\delta\leq d(x, y)\leq \delta^{-1}\\\theta_x,\theta_y> R}} h(d(x, y))\rho_k(x)\rho_k(y)\mathrm{d} x\mathrm{d} y\\ &=:\mathcal{I}_{31}+2\mathcal{I}_{32}-\mathcal{I}_{33}. \end{align*} $$

To estimate $\mathcal {I}_{32}$ , note that

(4.10) $$ \begin{align} \iint\limits_{\substack{\delta\leq d(x, y)\leq \delta^{-1}\\\theta_x> R}} \rho_k(y)\mathrm{d} x\mathrm{d} y&\leq \iint\limits_{d(x, y)\leq \delta^{-1}}\rho_k(y)\mathrm{d} x\mathrm{d} y =\int_{{\mathbb H}^n} \rho_k(y)|B_{\delta^{-1}}(y)|\mathrm{d} y \nonumber \\ &=\int_{{\mathbb H}^n} \rho_k(y)|B_{\delta^{-1}}(o)|\mathrm{d} y= |B_{\delta^{-1}}(o)|, \end{align} $$

where we integrated in x for the first equal sign, and for the second equal sign we used that on ${\mathbb H}^n$ all geodesic balls of a certain radius have the same volume. Then, by the monotonicity of h, along with (4.9) and (4.10), we get

(4.11) $$ \begin{align} |\mathcal{I}_{32}|&\leq \|\rho_k\|_{L^\infty(B_R(o)^c)}\max \left(|h(\delta)|, |h(\delta^{-1})|\right)\iint\limits_{\substack{\delta\leq d(x, y)\leq \delta^{-1}\\\theta_x> R}} \rho_k(y)\mathrm{d} x\mathrm{d} y \nonumber \\ &\leq|B_R(o)|^{-1}\max\left(|h(\delta)|, |h(\delta^{-1})|\right)|B_{\delta^{-1}}(o)|. \end{align} $$

For $\mathcal {I}_{33}$ , one can proceed exactly as in (4.10) and write

$$ \begin{align*} \iint\limits_{\substack{\delta\leq d(x, y)\leq \delta^{-1}\\\theta_x,\theta_y> R}} \rho_k(y)\mathrm{d} x\mathrm{d} y&\leq \iint\limits_{d(x, y)\leq \delta^{-1}}\rho_k(y)\mathrm{d} x\mathrm{d} y =\int_M \rho_k(y)|B_{\delta^{-1}}(y)|\mathrm{d} y \\ &=\int_M \rho_k(y)|B_{\delta^{-1}}(o)|\mathrm{d} y= |B_{\delta^{-1}}(o)|, \end{align*} $$

This estimate, together with the monotonicity of h and (4.9), lead to

(4.12) $$ \begin{align} |\mathcal{I}_{33}|&\leq \|\rho_k\|_{L^\infty(B_R(o)^c)}\max \left(|h(\delta)|, |h(\delta^{-1})|\right)\iint\limits_{\substack{\delta\leq d(x, y)\leq \delta^{-1}\\\theta_x,\theta_y> R}} \rho_k(y)\mathrm{d} x\mathrm{d} y \nonumber \\ &\leq|B_R(o)|^{-1}\max \left(|h(\delta)|, |h(\delta^{-1})|\right)|B_{\delta^{-1}}(o)|. \end{align} $$

Then, combining (4.11) and (4.12), we find

(4.13) $$ \begin{align} \begin{aligned} \mathcal{I}_3&\geq\mathcal{I}_{31}-2|\mathcal{I}_{32}|-|\mathcal{I}_{33}| \\ &\geq\mathcal{I}_{31}-3|B_R(o)|^{-1}\max \left(|h(\delta)|, |h(\delta^{-1})|\right)|B_{\delta^{-1}}(o)|. \end{aligned} \end{align} $$

Note that (4.13) holds for an arbitrary $k\geq 1$ , with $R>0$ and $\delta>0$ fixed, independent of k.

Overall, from (4.7), (4.8), and (4.13), we can conclude that

(4.14) $$ \begin{align} \begin{aligned} &\iint\limits_{{\mathbb H}^n \times {\mathbb H}^n} h(d(x, y))\rho_k(x)\rho_k(y)\mathrm{d} x\mathrm{d} y\\ &\quad \geq \iint\limits_{\substack{\delta\leq d(x, y)\leq \delta^{-1}\\\theta_x,\theta_y\leq R}} h(d(x, y))\rho_k(x)\rho_k(y)\mathrm{d} x\mathrm{d} y\\ &\qquad-2\epsilon-3|B_R(o)|^{-1}\max \left(|h(\delta)|, |h(\delta^{-1})|\right)|B_{\delta^{-1}}(o)|. \end{aligned} \end{align} $$

In (4.14), $\epsilon>0$ and $R>0$ are fixed, independent of k, and $0<\delta \leq \min (\delta _1,\delta _2)$ with $\delta _1$ and $\delta _2$ depending on $\epsilon $ but not on k.

Since $\{(x, y)\in {\mathbb H}^n \times {\mathbb H}^n: \delta \leq d(x, y)\leq \delta ^{-1}, \theta _x\leq R, \theta _y\leq R\}$ is a compact set in ${\mathbb H}^n \times {\mathbb H}^n$ , we can send $k\to \infty $ in (4.14) to get

$$ \begin{align*} &\liminf_{k\to\infty}\iint\limits_{{\mathbb H}^n \times {\mathbb H}^n} h(d(x, y))\rho_k(x)\rho_k(y)\mathrm{d} x\mathrm{d} y\\ & \geq \liminf_{k\to\infty}\iint\limits_{\substack{\delta\leq d(x, y)\leq \delta^{-1}\\\theta_x,\theta_y\leq R}} h(d(x, y))\rho_k(x)\rho_k(y)\mathrm{d} x\mathrm{d} y\\ &\quad-2\epsilon-3|B_R(o)|^{-1}\max(|h(\delta)|, |h(\delta^{-1})|)|B_{\delta^{-1}}(o)|\\ &=\iint\limits_{\substack{\delta\leq d(x, y)\leq \delta^{-1}\\\theta_x,\theta_y\leq R}} h(d(x, y))\rho_0(x)\rho_0(y)\mathrm{d} x\mathrm{d} y\\ &\quad-2\epsilon-3|B_R(o)|^{-1}\max(|h(\delta)|, |h(\delta^{-1})|)|B_{\delta^{-1}}(o)|. \end{align*} $$

By taking the limits $R\to \infty $ first and then $\delta \to 0+$ in the above, we get

$$\begin{align*}\liminf_{k\to\infty}\iint\limits_{{\mathbb H}^n \times {\mathbb H}^n} h(d(x, y))\rho_k(x)\rho_k(y)\mathrm{d} x\mathrm{d} y\geq \iint\limits_{{\mathbb H}^n\times {\mathbb H}^n} h(d(x, y))\rho_0(x)\rho_0(y)\mathrm{d} x\mathrm{d} y-2\epsilon. \end{align*}$$

Since $\epsilon>0$ was arbitrary, we get the lower-semicontinuity of the interaction energy:

(4.15) $$ \begin{align} \liminf_{k\to\infty}\iint\limits_{{\mathbb H}^n\times {\mathbb H}^n} h(d(x, y))\rho_k(x)\rho_k(y)\mathrm{d} x\mathrm{d} y\geq \iint\limits_{{\mathbb H}^n\times {\mathbb H}^n} h(d(x, y))\rho_0(x)\rho_0(y)\mathrm{d} x\mathrm{d} y. \end{align} $$

The conclusion of the theorem now follows from (4.6) and (4.15).

5 Existence of global minimizers on ${\mathbb H}^n$ : Case $h_\infty = \infty $

In this section, we will show the existence of global energy minimizers when the interaction potential is not too singular at origin (see (4.1)) and it grows to infinity ( $h_\infty = \infty $ ). In addition, we show in this case that the global minimizers are compactly supported.

Theorem 5.1 (Existence of global minimizers on ${\mathbb H}^n$ : case $h_\infty = \infty $ )

Consider the energy functional (2.2) on ${\mathbb H}^n$ , with $m>1$ . Assume the interaction potential satisfies assumption (H), for a function h that satisfies (4.1) and $\lim _{\theta \to \infty } h(\theta )=\infty $ . Then, the energy functional $E[\cdot ]$ has a global minimizer in $\mathcal {P}({\mathbb H}^n)$ .

Proof Let o be an arbitrary pole in ${\mathbb H}^n$ and $\{\rho _k\}_{k \geq 1}\subset L^m({\mathbb H}^n) \cap \mathcal {P}_o^{sd}({\mathbb H}^n)$ be a minimizing sequence of $E[\rho ]$ . From Proposition 4.1, we know that $\iint\limits _{{\mathbb H}^n\times {\mathbb H}^n}h(d(x, y))\rho _k(x)\rho _k(y)\mathrm {d} x\mathrm {d} y$ is uniformly bounded (above and below) and also, that (4.5) holds uniformly in $k\geq 1$ . Therefore, we can fix $\delta _1>0$ such that

$$\begin{align*}\Biggl|\iint\limits_{d(x, y)<\delta_1}h(d(x, y))\rho_k(x)\rho_k(y)\mathrm{d} x\mathrm{d} y\Biggr|<1, \qquad \text{ for all } k\geq 1. \end{align*}$$

This implies that for any $\delta>\delta _1$ ,

$$ \begin{align*} &\iint\limits_{d(x, y)<\delta}h(d(x, y))\rho_k(x)\rho_k(y)\mathrm{d} x\mathrm{d} y\\ &\quad = \iint\limits_{d(x, y)<\delta_1}h(d(x, y))\rho_k(x)\rho_k(y)\mathrm{d} x\mathrm{d} y+\iint\limits_{\delta_1\leq d(x, y)<\delta}h(d(x, y))\rho_k(x)\rho_k(y)\mathrm{d} x\mathrm{d} y\\ &\quad \geq -1+\min(h(\delta_1), 0), \end{align*} $$

for all $k \geq 1$ . Furthermore, for any $\delta>\delta _1$ , we then have

(5.1) $$ \begin{align} \begin{aligned} &\iint\limits_{d(x, y) \geq \delta}h(d(x, y))\rho_k(x)\rho_k(y)\mathrm{d} x\mathrm{d} y \\ &\quad =\iint\limits_{{\mathbb H}^n\times {\mathbb H}^n}h(d(x, y))\rho_k(x)\rho_k(y)\mathrm{d} x\mathrm{d} y\kern1.6pt{-}\iint\limits_{d(x, y)<\delta}h(d(x, y))\rho_k(x)\rho_k(y)\mathrm{d} x\mathrm{d} y\\ &\quad \leq \iint\limits_{{\mathbb H}^n\times {\mathbb H}^n}h(d(x, y))\rho_k(x)\rho_k(y)\mathrm{d} x\mathrm{d} y+1-\min(h(\delta_1), 0), \end{aligned} \end{align} $$

for all $k\geq 1$ .

Define

$$\begin{align*}V:=\sup_{k\in\mathbb{N}}\Biggl(\, \iint\limits_{{\mathbb H}^n\times {\mathbb H}^n}h(d(x, y))\rho_k(x)\rho_k(y)\mathrm{d} x\mathrm{d} y \Biggr) +1-\min(h(\delta_1), 0). \end{align*}$$

Recall that $\delta _1$ is a fixed number, so V is constant. Since $h_\infty =\infty $ , there exists $\delta _2>0$ such that $h(\delta )>0$ for all $\delta>\delta _2$ . Then, for any $\delta>\delta _2$ , we have

(5.2) $$ \begin{align} \iint\limits_{d(x, y)\geq \delta}h(d(x, y))\rho_k(x)\rho_k(y)\mathrm{d} x\mathrm{d} y\kern1.6pt{\geq} \iint\limits_{\substack{d(x, y)\geq \delta\\ \angle (xo y)\geq \pi/2}}h(d(x, y))\rho_k(x)\rho_k(y)\mathrm{d} x\mathrm{d} y, \end{align} $$

for all $k \geq 1$ . By combining (5.1) and (5.2), we infer that for any $\delta>\max (\delta _1, \delta _2)$ , we have

(5.3) $$ \begin{align} V\geq \iint\limits_{\substack{d(x, y)\geq \delta\\ \angle (xo y)\geq \pi/2}}h(d(x, y))\rho_k(x)\rho_k(y)\mathrm{d} x\mathrm{d} y, \qquad \text{ for all } k\geq 1. \end{align} $$

We will use the Rauch comparison theorem – see setup in Section 3 leading to (3.4). First, by the cosine law in the Euclidean space $T_o {\mathbb H}^n,$ we have

$$\begin{align*}|\log_o x - \log_o y|^2 = \theta_x^2+\theta_y^2-2\theta_x\theta_y\cos\angle(xo y), \qquad \text{ for all } x,y \in {\mathbb H}^n. \end{align*}$$

By combining the above with (3.4), we then get

$$ \begin{align*} d(x, y)^2 &\geq \theta_x^2+\theta_y^2-2\theta_x\theta_y\cos\angle(xo y) \\[2pt] &\geq \theta_x^2+\theta_y^2, \end{align*} $$

for all points $x,y$ such that $\angle (xo y)\geq \pi /2$ . Furthermore, if x and y satisfy $\theta _x, \theta _y\geq \delta $ and $\angle (xo y)\geq \pi /2$ , for some $\delta>0$ , then

$$\begin{align*}d(x, y)\geq\sqrt{2}\delta>\delta. \end{align*}$$

It implies

$$ \begin{align*} &\{(x, y)\in {\mathbb H}^n\times {\mathbb H}^n: d(x, y)\geq \delta,\angle(xo y)\geq \pi/2\}\supset \\ &\quad\{(x, y)\in {\mathbb H}^n\times {\mathbb H}^n: \theta_x\geq \delta, \theta_y\geq \delta,\angle(xo y)\geq \pi/2\}. \end{align*} $$

Using these considerations in (5.3), we then find for any $\delta>\max (\delta _1, \delta _2)$ ,

(5.4) $$ \begin{align} \begin{aligned} V &\geq \iint\limits_{\substack{\theta_x, \theta_y\geq \delta\\ \angle (xo y)\geq \pi/2}}h(d(x, y))\rho_k(x)\rho_k(y)\mathrm{d} x\mathrm{d} y \\ &\geq \iint\limits_{\substack{\theta_x, \theta_y\geq \delta\\\angle (xo y)\geq \pi/2}}h(\delta)\rho_k(x)\rho_k(y)\mathrm{d} x\mathrm{d} y, \end{aligned} \end{align} $$

for all $k \geq 1$ . Now, from the radial symmetry of $\rho _k$ with respect to o, for any fixed $x\neq o$ , we have

$$\begin{align*}\int_{\substack{\theta_y\geq \delta\\ \angle(xo y)\geq \pi/2}}\rho_k(y)\mathrm{d} y=\frac{1}{2}\int_{\theta_y\geq\delta}\rho_k(y)\mathrm{d} y. \end{align*}$$

Hence, we get for any $\delta>\max (\delta _1, \delta _2)$ ,

$$ \begin{align*} \iint\limits_{\substack{\theta_x, \theta_y\geq \delta\\ \angle (xo y)\geq \pi/2}}h(\delta)\rho_k(x)\rho_k(y)\mathrm{d} x\mathrm{d} y&=\frac{h(\delta)}{2}\iint\limits_{\theta_x,\theta_y\geq\delta}\rho_k(x)\rho_k(y)\mathrm{d} x\mathrm{d} y \\ &=\frac{h(\delta)}{2}\left(\int_{\theta_x\geq \delta}\rho_k(x)\mathrm{d} x\right)^2, \end{align*} $$

which used in (5.4) yields

(5.5) $$ \begin{align} V\geq\frac{h(\delta)}{2}\left(\int_{\theta_x\geq \delta}\rho_k(x)\mathrm{d} x\right)^2, \qquad \text{ for all } k \geq 1. \end{align} $$

Write (5.5) as

$$\begin{align*}\int_{\theta_x\geq \delta}\rho_k(x)\mathrm{d} x\leq\left(\frac{2V}{h(\delta)}\right)^{1/2},\qquad\text{ for all } \delta>\max(\delta_1, \delta_2) \text{ and } k\geq 1. \end{align*}$$

Finally, use $\lim _{\delta \to \infty }h(\delta )=\infty $ to get

$$\begin{align*}\lim_{\delta\to\infty}\int_{\theta_x\geq \delta}\rho_k(x)\mathrm{d} x=0, \qquad \text{ uniformly in } k \geq 1. \end{align*}$$

We infer from here that $\rho _k$ is tight. Then, from Prokhorov’s theorem, there exists a subsequence $\{ \rho _{k_l}\}$ of $\{\rho _k\}$ such that $\{ \rho _{k_l}\}$ converges weakly- $^{*}$ to $\rho _0\in \mathcal {P}({\mathbb H}^n)$ as ${l\to \infty }$ . Define

$$\begin{align*}E_{\inf}:=\inf_{\rho\in L^m({\mathbb H}^n) \cap \mathcal{P}({\mathbb H}^n)}E[\rho]. \end{align*}$$

From the definition of the minimizing sequence and the weak lower semicontinuity of the energy functional (see Proposition 4.2), we have

(5.6) $$ \begin{align} E[\rho_0] \leq \liminf_{l\to\infty}E[\rho_{k_l}] = E_{\inf}. \end{align} $$

If $\rho _0\not \in L^m({\mathbb H}^n)$ , then by Proposition 2.3 $E[\rho _0]=\infty $ , and the above inequality cannot hold. Hence, $\rho _0\in L^m({\mathbb H}^n) \cap \mathcal {P}({\mathbb H}^n)$ , which yields

(5.7) $$ \begin{align} E_{\inf}\leq E[\rho_0]. \end{align} $$

From (5.6) and (5.7), we conclude that $E[\rho _0]=E_{\inf }$ and hence $\rho _0$ is a global energy minimizer over $\mathcal {P}({\mathbb H}^n)$ .

Remark 5.1 Theorem 5.1 considers interaction potentials that are less singular than $-\frac {1}{\theta ^{\min (n(m-1),n)}}$ at origin. This is consistent with Propositions 2.2 and 2.3, where the energy is shown to be well-defined for such interaction potentials. Theorem 5.1 also complements the result from Proposition 2.1, where the attraction is too singular and leads to energies unbounded from below.

We will now investigate the support of the global minimizers. By the considerations in Section 2.4, we can choose a radially symmetric and decreasing global minimizer $\rho \in \mathcal {P}_o^{sd}({\mathbb H}^n)$ , for a pole $o \in {\mathbb H}^n$ fixed. Then, the support of $\rho $ should be either a closed ball with radius $R<\infty $ or the entire ${\mathbb H}^n$ . We will show that the latter case is not possible.

Proposition 5.1 Assume the interaction potential satisfies the assumptions in Theorem 5.1 and consider a global energy minimizer $\rho \in \mathcal {P}_o^{sd}({\mathbb H}^n)$ , for a pole $o \in {\mathbb H}^n$ fixed. Then, the support of $\rho $ is a compact set.

Proof Since $\rho $ is a global minimizer, in particular, it is a critical point of the energy. The derivation of the Euler-Lagrange equation for the Euclidean space (see, for instance, [Reference Carrillo, Delgadino and Patacchini10, Section 2.3] or [Reference Kaib37]) extends immediately to the manifold setup. For the energy (2.2), the Euler-Lagrange equation is given by

(5.8) $$ \begin{align} \frac{m}{m-1}\rho(x)^{m-1}+\int_{{\mathbb H}^n} h(d(x, y))\rho(y)\mathrm{d} y=C,\qquad \forall x\in\mathrm{supp}(\rho). \end{align} $$

Assume that the support of $\rho $ is the entire ${\mathbb H}^n$ , so (5.8) holds for all $x \in {\mathbb H}^n$ . Since h is non-decreasing and $h_\infty =\infty $ , there exists $s\geq 0$ such that

$$\begin{align*}s:=\inf\{\theta:h(\theta)\geq 0\}. \end{align*}$$

If $s=0$ , then $h(\theta )\geq 0$ for all $\theta>0$ . If $s>0$ then $h(\theta )<0$ for $0<\theta < s$ , and $h(\theta )\geq 0$ for $\theta \geq s$ . Also, take $R>0$ fixed such that

(5.9) $$ \begin{align} \int_{\theta_y\geq R}\rho(y) \mathrm{d} y \geq \frac{1}{2}. \end{align} $$

For $x\in \mathrm {supp}(\rho ) = {\mathbb H}^n$ fixed, but arbitrary, that satisfies $\theta _x\geq s+R$ , write

(5.10) $$ \begin{align} \int_{{\mathbb H}^n} h(d(x, y))\rho(y)\mathrm{d} y\kern1.6pt{=}\int_{d(x, y) < s} h(d(x, y))\rho(y)\mathrm{d} y\kern1.6pt{+}\int_{d(x, y)\geq s} h(d(x, y))\rho(y)\mathrm{d} y. \end{align} $$

We will inspect separately the two integrals in the r.h.s. of (5.10). For the first integral, note that for $y \in \{y: d(x,y)<s \}$ , we have

$$\begin{align*}\theta_y\geq \theta_x-d(x, y)\geq s+R-s=R, \end{align*}$$

which implies

$$ \begin{align*} \{y: d(x,y) < s\}\subset B_R(o)^c. \end{align*} $$

We, then infer

(5.11) $$ \begin{align} \|\rho\|_{L^\infty(\{y:d(x, y)\leq s\})}&\leq \|\rho\|_{L^\infty(B_R(o)^c)} \nonumber \\[2pt] & \leq |B_R(o)|^{-1}, \end{align} $$

where for the second inequality sign we used that $\rho $ is radially symmetric and decreasing (see (4.9) and the argument leading to it).

From Lemma 2.1, there exist $\gamma _1, \gamma _2\geq 0$ such that

$$ \begin{align*} h(\theta)\geq -\gamma_1\theta^{-\alpha}-\gamma_2,\qquad\forall \theta>0. \end{align*} $$

If we set $\beta =\gamma _1+\gamma _2s^\alpha $ , then we have

(5.12) $$ \begin{align} h(\theta)\geq -\beta\theta^{-\alpha},\qquad\forall \theta\in(0, s]. \end{align} $$

Using (5.12) followed by (5.11), we then find

(5.13) $$ \begin{align} \int_{d(x, y) < s} h(d(x, y))\rho(y)\mathrm{d} y &\geq-\beta\int_{d(x, y) < s}\frac{\rho(y)\mathrm{d} y}{d(x, y)^\alpha} \nonumber \\ &\geq-\beta|B_R(o)|^{-1}\int_{d(x, y)\leq s}\frac{\mathrm{d} y}{d(x, y)^\alpha} \nonumber \\ & = - \beta|B_R(o)|^{-1} \int_0^s\frac{| \partial B_\theta(o)|}{\theta^\alpha}\mathrm{d}\theta. \end{align} $$

Since $\lim _{\theta \to 0+}\frac {| \partial B_\theta (o)|}{\theta ^{n-1}} = n w(n)$ and $\alpha <n$ , the integral in the r.h.s. above is convergent, and we can define $U>0$ as

$$ \begin{align*} U:=\beta|B_R(o)|^{-1} \int_0^s\frac{|\partial B_\theta(o)|}{\theta^\alpha} \mathrm{d}\theta. \end{align*} $$

For the second integral in the r.h.s. of (5.10), note that if y is such that $\theta _y\leq R$ , then $d(x, y)\geq \theta _x-\theta _y\geq s+R-R=s$ . This implies

$$\begin{align*}\{y: \theta_y\leq R\}\subset \{y:d(x, y)\geq s\}. \end{align*}$$

and hence,

(5.14) $$ \begin{align} \int_{d(x, y)\geq s} h(d(x, y))\rho(y)\mathrm{d} y\geq \int_{B_{R}(o)}h(d(x, y))\rho(y)\mathrm{d} y. \end{align} $$

Finally, by using (5.10) in (5.8), along with (5.13) and (5.14), we get

$$ \begin{align*} C&=\frac{m}{m-1}\rho(x)^{m-1}+\int_{d(x, y)< s} h(d(x, y))\rho(y)\mathrm{d} y+\int_{d(x, y)\geq s} h(d(x, y))\rho(y)\mathrm{d} y\\ &\geq0-U+\int_{B_{R}(o)}h(d(x, y))\rho(y)\mathrm{d} y, \end{align*} $$

for all $x\in {\mathbb H}^n$ , with $\theta _x\geq s+R$ . Hence, for all such $x,$ we get

$$ \begin{align*} C+U &\geq \int_{B_{R}(o)}h(d(x, y))\rho(y)\mathrm{d} y \\ & \geq \int_{B_{R}(o)}h(\theta_x-R)\rho(y)\mathrm{d} y \\ & =h(\theta_x-R) \int_{B_{R}(o)}\rho(y)\mathrm{d} y, \end{align*} $$

where for the second inequality sign we used that $d(x, y)\geq \theta _x-\theta _y\geq \theta _x-R$ for $\theta _y \leq R$ , and that by the monotonicity of h and the definition of s, we have $h(\theta _x-R)\geq h(s)\geq 0$ . Now, use how R was defined (see (5.9)) to conclude

$$\begin{align*}C+U \geq \frac{1}{2} h(\theta_x-R), \end{align*}$$

for any $x \in {\mathbb H}^n$ which satisfies $\theta _x\geq R+s$ . Since x is arbitrary, we then find

$$\begin{align*}C+U\geq \frac{1}{2}\lim_{\theta_x\to\infty}h(\theta_x-R)=\infty \end{align*}$$

which yields a contradiction. The only remaining possibility is that the support of $\rho $ is a closed ball centred of o with radius $R<\infty $ .

6 Existence of global minimizers on ${\mathbb H}^n$ : Case $h_\infty =0$

In this section, we show the existence of global energy minimizers on ${\mathbb H}^n$ for the case $h_\infty =0$ . The methods used to prove Theorem 5.1 do not work in this case, as the attraction is not strong enough at infinity to ensure tightness of the minimizing sequences. Consequently, the arguments in this section and more delicate and involved.

Proposition 6.1 Consider the energy functional (2.2) on ${\mathbb H}^n$ , with $m>2$ . Assume the interaction potential satisfies assumption (H), for a function h such that $\lim _{\theta \to \infty } h(\theta ) =0$ . Then, there exists $\rho \in L^m({\mathbb H}^n) \cap \mathcal {P}({\mathbb H}^n)$ such that

$$\begin{align*}E[\rho]<0. \end{align*}$$

Proof Let o be a fixed pole in ${\mathbb H}^n$ . Given that h is non-decreasing and $h_\infty = 0$ , there exists $\tilde {\theta }>0$ such that

$$\begin{align*}|B_{\tilde{\theta}}(o)|\leq1\qquad\text{and}\qquad h(2\tilde{\theta})<0. \end{align*}$$

Then, for any $x, y\in B_{\tilde {\theta }}(z)$ for some z, we have

(6.1) $$ \begin{align} h(d(x, y))\leq h(2\tilde{\theta}) <0. \end{align} $$

For any $\alpha>0$ , define $\rho _\alpha =\alpha \chi _{B_{\tilde {\theta }}(o)}$ . By a simple calculation,

$$\begin{align*}E[\rho_\alpha]=A \alpha^m+B \alpha^2, \end{align*}$$

where

$$\begin{align*}A:=\frac{1}{m-1}\int_{B_{\tilde{\theta}}(o)}1\mathrm{d} x>0,\quad \text{ and } \quad B:=\frac{1}{2}\iint\limits_{B_{\tilde{\theta}}(o)\times B_{\tilde{\theta}}(o)}h(d(x, y))\mathrm{d} x\mathrm{d} y<0. \end{align*}$$

Since $m>2$ , by comparing the two terms of orders $\alpha ^m$ and $\alpha ^2$ , we have

$$\begin{align*}\lim_{\alpha\to0+}E[\rho_\alpha]<0. \end{align*}$$

This implies that there exists $\alpha _0>0$ such that

$$\begin{align*}E[\rho_\alpha]<0, \qquad\forall\alpha\in(0,\alpha_0]. \end{align*}$$

Note however that $\rho _\alpha $ is not a probability measure, so it cannot be used to conclude the theorem.

Let N be an integer such that $ N \geq \frac {1}{\alpha _0|B_{\tilde {\theta }}(o)|}$ . Then,

$$\begin{align*}\alpha_*:=\frac{1}{N|B_{\tilde{\theta}}(o)|}\leq \alpha_0, \end{align*}$$

and hence,

(6.2) $$ \begin{align} E[\rho_{\alpha_*}]<0. \end{align} $$

Since N is finite, we can choose N points $x_1, x_2, \cdots , x_N \in {\mathbb H}^n$ which satisfy

$$\begin{align*}B_{\tilde{\theta}}(x_i)\cap B_{\tilde{\theta}}(x_j)=\emptyset,\qquad \text{ for all } 1\leq i<j\leq N. \end{align*}$$

Using these points, construct the measure $\rho $ by

$$\begin{align*}\rho=\frac{1}{N|B_{\tilde{\theta}}(o)|}\sum_{i=1}^N\chi_{B_{\tilde{\theta}}(x_i)}. \end{align*}$$

Note that $\rho $ is a probability measure, as

$$\begin{align*}\int_{{\mathbb H}^n}\rho(x)\mathrm{d} x=\alpha_*\sum_{i=1}^N|B_{\tilde{\theta}}(x_i)| =\alpha_* N|B_{\tilde{\theta}}(o)|=1, \end{align*}$$

where we used that all balls with radius $\tilde {\theta }$ have the same volume since ${\mathbb H}^n$ has constant sectional curvature.

The conclusion can then be inferred from the following calculation:

$$ \begin{align*} E[\rho]& =\alpha_*^m\sum_{i=1}^N\left(\frac{1}{m-1}\int_{B_{\tilde{\theta}}(x_i)}1\mathrm{d} x\right)+\frac{\alpha_*^2}{2}\sum_{i, j=1}^N\left(\iint\limits_{B_{\tilde{\theta}}(x_i)\times B_{\tilde{\theta}}(x_j)}h(d(x, y))\mathrm{d} x \mathrm{d} y\right)\\ &\leq\alpha_*^m\sum_{i=1}^N\left(\frac{1}{m-1}\int_{B_{\tilde{\theta}}(x_i)}1\mathrm{d} x\right)+\frac{\alpha_*^2}{2}\sum_{i=1}^N\left(\iint\limits_{B_{\tilde{\theta}}(x_i)\times B_{\tilde{\theta}}(x_i)}h(d(x, y))\mathrm{d} x \mathrm{d} y\right)\\ &=\alpha_*^m\sum_{i=1}^N\left(\frac{1}{m-1}\int_{B_{\tilde{\theta}}(o)}1\mathrm{d} x\right)+\frac{\alpha_*^2}{2}\sum_{i=1}^N\left(\iint\limits_{B_{\tilde{\theta}}(o)\times B_{\tilde{\theta}}(o)}h(d(x, y))\mathrm{d} x \mathrm{d} y\right)\\ &=NE[\rho_{\alpha_*}]<0, \end{align*} $$

where for the second line we only retained the diagonal part ( $1\leq i=j\leq N$ ) of the double sum ( $1\leq i, j\leq N$ ) and used (6.1). Also, for the third line, we used that ${\mathbb H}^n$ is a homogeneous manifold, and translated by isometries the balls $B_{\tilde {\theta }}(x_i)$ to $B_{\tilde {\theta }}(o)$ . For the last line, we used (6.2).

Proposition 6.1 covers only the range $m>2$ . The analogous result for $1<m\leq 2$ , given by the following proposition, assumes an additional assumption on h.

Proposition 6.2 Consider the energy functional (2.2) on ${\mathbb H}^n$ , with $1<m\leq 2$ , and let o be a fixed pole in ${\mathbb H}^n$ . Assume the interaction potential satisfies assumption (H), with a function h that satisfies $\lim _{\theta \to \infty } h(\theta ) =0$ and

(6.3) $$ \begin{align} h(\theta_0)<-\frac{2}{(m-1)|B_{\theta_0/2}(o)|^{m-1}}, \end{align} $$

for some $\theta _0>0$ . Then, there exists $\rho \in L^m({\mathbb H}^n) \cap \mathcal {P}({\mathbb H}^n)$ such that

$$\begin{align*}E[\rho]<0. \end{align*}$$

Proof Consider again densities in the form (A.1), for $R>0$ . Then, by the monotonicity of h, we have

$$ \begin{align*} E[\rho_R] &= \frac{1}{(m-1)|B_R(o)|^{m-1}} + \frac{1}{2} \frac{1}{|B_R(o)|^2} \iint\limits_{B_R(o) \times B_R(o)} h(d(x,y)) \mathrm{d} x \mathrm{d} y \\ &\leq\frac{1}{(m-1)|B_R(o)|^{m-1}}+\frac{h(2R)}{2}. \end{align*} $$

By the extra assumption (6.3) on h, by choosing $R = \frac {\theta _0}{2}$ , we find

$$\begin{align*}E[\rho_R]<0, \end{align*}$$

and this concludes the proof.

Remark 6.1 The volumes of geodesic balls on the hyperbolic space grow exponentially fast as their radius increases to infinity – see Theorem 2.4. From this property we infer that a Riesz potential of the form $h(\theta ) = -\frac {1}{\beta \, \theta ^\beta }$ with $\beta>0$ , satisfies (6.3) for a large enough $\theta _0$ . Hence, while the extra condition in (6.3) can be interpreted as h not being allowed to increase too fast, it is sufficiently mild to include large classes of interaction potentials considered in the literature.

Theorem 6.1 (Existence of global minimizers on ${\mathbb H}^n$ : case $h_\infty = 0$ )

Consider the energy functional (2.2) on ${\mathbb H}^n$ , with $m>1$ , and let o be a fixed pole on ${\mathbb H}^n$ . Assume the interaction potential satisfies assumption (H), for a function h that satisfies (4.1) and $\lim _{\theta \to \infty } h(\theta )=0$ . In addition, if $1<m\leq 2$ , assume that h satisfies the extra assumption (6.3), for some $\theta _0>0$ . Then, the energy functional $E[\cdot ]$ has a global minimizer in $\mathcal {P}({\mathbb H}^n)$ .

Proof Let $\{\rho _k\}_{k\geq 1}\subset L^m({\mathbb H}^n) \cap \mathcal {P}_o^{sd}({\mathbb H}^n)$ be a minimizing sequence of the energy functional. From Theorem 2.1, there exists a subsequence $\{\rho _{k_l}\}_{l\geq 1 }$ that converges weakly- $^{*}$ to $\rho _0\in \mathcal {M}_+({\mathbb H}^n)$ , which satisfies $\int _{{\mathbb H}^n}\rho _0(x)\mathrm {d} x\leq 1$ .

From Propositions 6.1 and 6.2, we know that there exists $\rho \in L^m({\mathbb H}^n) \cap \mathcal {P}({\mathbb H}^n)$ such that $E[\rho ]<0$ . Then, since $\rho _{k_l}$ is a minimizing sequence we infer that $\liminf _{l \to \infty } E[\rho _{k_l}]<0$ , and hence, by the weak lower semicontinuity of the energy (Proposition 4.2), we get

(6.4) $$ \begin{align} E[\rho_0]\leq \liminf_{l \to \infty} E[\rho_{k_l}]<0. \end{align} $$

We want to show that $\rho _0$ is the global energy minimizer, and for this purpose we need to show that it is a probability measure (i.e., $\int _{{\mathbb H}^n}\rho _0(x)\mathrm {d} x=1$ ). Since $E[0]=0$ , $\rho _0$ cannot be the zero measure, and if $\|\rho _0\|_{L^1({\mathbb H}^n)}=1$ , then the proof is done. Hence, we assume $0<\|\rho _0\|_{L^1({\mathbb H}^n)}<1$ , and follow a proof by contradiction argument.

We consider the two cases $1<m\leq 2$ and $m>2$ separately.

Case 1: $1<m\leq 2$ . For this case, the proof is simple. Define $\tilde {\rho }:=\frac {1}{\|\rho _0\|_{L^1({\mathbb H}^n)}}\rho _0\in \mathcal {P}({\mathbb H}^n)$ . Then, we have

$$ \begin{align*} E[\tilde{\rho}]&=\frac{1}{\|\rho_0\|_{L^1({\mathbb H}^n)}^m} \times \frac{1}{m-1}\int_{{\mathbb H}^n} \rho_0^m(x)\mathrm{d} x+\frac{1}{\|\rho_0\|_{L^1({\mathbb H}^n)}^2}\\&\quad\times \frac{1}{2}\iint\limits_{{\mathbb H}^n \times {\mathbb H}^n}h(d(x, y))\rho_0(x)\rho_0(y)\mathrm{d} x\mathrm{d} y\\ &\leq \frac{1}{\|\rho_0\|_{L^1({\mathbb H}^n)}^2} \times \frac{1}{m-1}\int_{{\mathbb H}^n} \rho_0^m(x)\mathrm{d} x+\frac{1}{\|\rho_0\|_{L^1({\mathbb H}^n)}^2}\\&\quad\times \frac{1}{2}\iint\limits_{{\mathbb H}^n \times {\mathbb H}^n}h(d(x, y))\rho_0(x)\rho_0(y)\mathrm{d} x\mathrm{d} y\\ &=\frac{1}{\|\rho_0\|_{L^1({\mathbb H}^n)}^2} E[\rho_0], \end{align*} $$

where for the second line we used that $\|\rho _0\|_{L^1({\mathbb H}^n)}^2 \leq \|\rho _0\|_{L^1({\mathbb H}^n)}^m$ (as $\|\rho _0\|_{L^1({\mathbb H}^n)}<1$ ).

Furthermore, since $E[\rho _0]<0$ and $\|\rho _0\|_{L^1({\mathbb H}^n)}<1$ , we infer from above that

$$\begin{align*}E[\tilde{\rho}]<E[\rho_0], \end{align*}$$

which combined with (6.4) leads to the following contradiction:

$$\begin{align*}E_{\inf} \leq E[\tilde{\rho}] < \liminf_{l\to\infty}E[\rho_{k_l}] = E_{\inf}. \end{align*}$$

Hence, necessarily $\rho _0 \in \mathcal {P}({\mathbb H}^n)$ .

Case 2: $m>2$ . From Proposition 6.1, there exists $\rho \in L^m({\mathbb H}^n) \cap \mathcal {P}({\mathbb H}^n)$ such that $E[\rho ]<0$ . We can assume that both $\rho _0$ and $\rho $ are radially symmetric and decreasing about the pole o.

As ${\mathbb H}^n$ is a homogeneous manifold, for any arbitrary point $o'\in {\mathbb H}^n$ , there exists an isometry $i_{o'}:{\mathbb H}^n \to {\mathbb H}^n$ which sends o into $o'$ , i.e., $i_{o'}(o)=o'$ . Set

$$\begin{align*}\rho_{o'}:={(i_{o'})}_\# \rho. \end{align*}$$

Since $\rho \in \mathcal {P} ({\mathbb H}^n) $ and $i_{o'}$ is an isometry, we also have $\rho _{o'} \in \mathcal {P} ({\mathbb H}^n) $ . Also, as $i_{o'}$ is an isometry, we have

$$\begin{align*}\int_{{\mathbb H}^n} \rho(x)^m\mathrm{d} x=\int_{{\mathbb H}^n} \rho_{o'}(x)^m\mathrm{d} x, \end{align*}$$

and

(6.5) $$ \begin{align} \quad \iint\limits_{{\mathbb H}^n \times {\mathbb H}^n}h(d(x, y))\rho_{o'}(x)\rho_{o'}(y)\mathrm{d} x\mathrm{d} y=\iint\limits_{{\mathbb H}^n \times {\mathbb H}^n} h(d(x, y))\rho(x)\rho(y)\mathrm{d} x\mathrm{d} y, \end{align} $$

which put together yield

$$\begin{align*}E[\rho_{o'}]=E[\rho]<0. \end{align*}$$

By a similar argument as above, we also have

$$\begin{align*}E[(1-\|\rho_0\|_{L^1({\mathbb H}^n)})\rho_{o'}]=E[(1-\|\rho_0\|_{L^1({\mathbb H}^n)})\rho]. \end{align*}$$

Since $0<1-\|\rho _0\|_{L^1({\mathbb H}^n)}<1$ and $m>2$ , we also get

(6.6) $$ \begin{align} \begin{aligned} &E[(1-\|\rho_0\|_{L^1({\mathbb H}^n)})\rho_{o'}] \\ &=(1-\|\rho_0\|_{L^1({\mathbb H}^n)})^m\times \frac{1}{m-1}\int_{{\mathbb H}^n} \rho_{o'}(x)^m\mathrm{d} x+(1-\|\rho_0\|_{L^1({\mathbb H}^n)})^2\\&\quad\times \frac{1}{2}\iint\limits_{{\mathbb H}^n \times {\mathbb H}^n}h(d(x, y))\rho_{o'}(x)\rho_{o'}(y)\mathrm{d} x\mathrm{d} y \\ &\leq (1-\|\rho_0\|_{L^1({\mathbb H}^n)})^2\times \frac{1}{m-1}\int_{{\mathbb H}^n} \rho_{o'}(x)^m\mathrm{d} x+(1-\|\rho_0\|_{L^1({\mathbb H}^n)})^2\\&\quad\times \frac{1}{2}\iint\limits_{{\mathbb H}^n \times {\mathbb H}^n}h(d(x, y))\rho_{o'}(x)\rho_{o'}(y)\mathrm{d} x\mathrm{d} y \\ &=(1-\|\rho_0\|_{L^1({\mathbb H}^n)})^2E[\rho_{o'}]<0. \end{aligned} \end{align} $$

The goal is to justify the following claim:

$\underline {\textit {Claim}\!:}$ There exists $o'\in {\mathbb H}^n$ such that

(6.7) $$ \begin{align} E[\rho_0+(1-\|\rho_0\|_{L^1({\mathbb H}^n)})\rho_{o'}]<E[\rho_0]. \end{align} $$

Given that $\rho _0+(1-\|\rho _0\|_{L^1({\mathbb H}^n)})\rho _{o'} \in \mathcal {P}({\mathbb H}^n)$ , this will lead to the desired contradiction.

By a direct calculation, we get

$$ \begin{align*} &E[\rho_0+(1-\|\rho_0\|_{L^1({\mathbb H}^n)})\rho_{o'}]\\ &\quad = \frac{1}{m-1}\int_{{\mathbb H}^n} \left(\rho_0(x)+(1-\|\rho_0\|_{L^1({\mathbb H}^n)})\rho_{o'}(x)\right)^m\mathrm{d} x\\ & \qquad + \frac{1}{2}\iint\limits_{{\mathbb H}^n \times {\mathbb H}^n}h(d(x, y))\left(\rho_0(x)+ (1-\|\rho_0\|_{L^1({\mathbb H}^n)})\rho_{o'}(x)\right)\left(\rho_0(y)\right.\\ &\qquad\left.+(1-\|\rho_0\|_{L^1({\mathbb H}^n)})\rho_{o'}(y)\right)\mathrm{d} x\mathrm{d} y\\ & \quad = \frac{1}{m-1}\int_{{\mathbb H}^n} \left(\rho_0(x)+(1-\|\rho_0\|_{L^1({\mathbb H}^n)})\rho_{o'}(x)\right)^m\mathrm{d} x\\ &\qquad + \frac{1}{2}\iint\limits_{{\mathbb H}^n \times {\mathbb H}^n}h(d(x, y))\rho_0(x)\rho_0(y)\mathrm{d} x\mathrm{d} y\\ &\qquad + (1-\|\rho_0\|_{L^1({\mathbb H}^n)}) \iint\limits_{{\mathbb H}^n \times {\mathbb H}^n}h(d(x, y))\rho_0(x)\rho_{o'}(y)\mathrm{d} x\mathrm{d} y\\ &\qquad + \frac{1}{2} (1-\|\rho_0\|_{L^1({\mathbb H}^n)})^2 \iint\limits_{{\mathbb H}^n \times {\mathbb H}^n}h(d(x, y)) \rho_{o'}(x) \rho_{o'}(y)\mathrm{d} x\mathrm{d} y\\ & \quad \leq \frac{1}{m-1}\int_{{\mathbb H}^n} \left(\rho_0(x)+ (1-\|\rho_0\|_{L^1({\mathbb H}^n)} )\rho_{o'}(x) \right)^m\mathrm{d} x\\ &\qquad + \frac{1}{2}\iint\limits_{{\mathbb H}^n \times {\mathbb H}^n}h(d(x, y))\rho_0(x)\rho_0(y)\mathrm{d} x\mathrm{d} y \\ &\qquad + \frac{1}{2} (1-\|\rho_0\|_{L^1({\mathbb H}^n)})^2 \iint\limits_{{\mathbb H}^n \times {\mathbb H}^n}h(d(x, y)) \rho_{o'}(x) \rho_{o'}(y)\mathrm{d} x\mathrm{d} y, \end{align*} $$

where for the last inequality we used $h\leq 0$ and dropped the mixed term. Furthermore, use (6.5) to write the result above as

(6.8) $$ \begin{align} \begin{aligned} &E[\rho_0+(1-\|\rho_0\|_{L^1({\mathbb H}^n)})\rho_{o'}]\\[2pt]& \quad \leq \frac{1}{m-1}\int_{{\mathbb H}^n} \left(\rho_0(x)+(1-\|\rho_0\|_{L^1({\mathbb H}^n)})\rho_{o'}(x)\right)^m\mathrm{d} x\\ & \qquad + \frac{1}{2}\iint\limits_{{\mathbb H}^n \times {\mathbb H}^n}h(d(x, y))\rho_0(x)\rho_0(y)\mathrm{d} x\mathrm{d} y \\& \qquad + \frac{1}{2} (1-\|\rho_0\|_{L^1({\mathbb H}^n)})^2 \iint\limits_{{\mathbb H}^n \times {\mathbb H}^n}h(d(x, y)) \rho(x) \rho(y)\mathrm{d} x\mathrm{d} y. \end{aligned} \end{align} $$

Next, we will focus the entropy part, i.e., the first term in the r.h.s. of (6.8). From the mean value theorem, for any nonnegative numbers a and b, we have

$$\begin{align*}(a+b)^m-a^m= mb(a+b_*)^{m-1}, \end{align*}$$

where $b_*\in [0, b]$ . This implies

$$\begin{align*}(a+b)^m-a^m\leq mb(a+b)^{m-1}. \end{align*}$$

Then, for any non-negative functions $a(x)$ and $b(x)$ on a measurable set $S\subset {\mathbb H}^n$ , we get

(6.9) $$ \begin{align} \int_S \left((a(x)+b(x))^m-a(x)^m \right)\mathrm{d} x\leq\int_S mb(x)(a(x)+b(x))^{m-1}\mathrm{d} x. \end{align} $$

Also, by Hölder inequality, we can further estimate

(6.10) $$ \begin{align} \int_S b(x)(a(x)+b(x))^{m-1}\mathrm{d} x\leq \left(\int_S b(x)^m\mathrm{d} x\right)^{\frac{1}{m}}\left(\int_S(a(x)+b(x))^m \mathrm{d} x\right)^{\frac{m-1}{m}}. \end{align} $$

Given the following elementary inequality:

(6.11) $$ \begin{align} (a(x)+b(x))^m\leq (2\max(a(x), b(x)))^m\leq (2a(x))^m+(2b(x))^m=2^m(a(x)^m+b(x)^m), \end{align} $$

we can combine (6.9), (6.10), and (6.11) to get

(6.12) $$ \begin{align} \int_S \left((a(x)+b(x))^m-a(x)^m \right)\mathrm{d} x\leq m2^{m-1}\left(\int_S b(x)^m\mathrm{d} x\right)^{\frac{1}{m}}\left(\int_S a(x)^m\mathrm{d} x+\int_S b(x)^m \mathrm{d} x\right)^{\frac{m-1}{m}}, \end{align} $$

which holds for all non-negative functions $a(x)$ and $b(x)$ , and measurable subsets $S\subset {\mathbb H}^n$ .

From

$$\begin{align*}{\mathbb H}^n = \{x\in {\mathbb H}^n:d(x, o)\geq d(o, o')/2\}\cup \{x\in M: d(x, o')\geq d(o, o')/2\}, \end{align*}$$

we can separate the integration on ${\mathbb H}^n$ of the entropy part as follows:

(6.13) $$ \begin{align} \begin{aligned} &\int_{{\mathbb H}^n} \left(\rho_0(x)+ (1-\|\rho_0\|_{L^1({\mathbb H}^n)})\rho_{o'}(x)\right)^m\mathrm{d} x \\&\quad \leq \int_{d(x, o)\geq d(o, o')/2} \left(\rho_0(x)+(1-\|\rho_0\|_{L^1({\mathbb H}^n)})\rho_{o'}(x)\right)^m\mathrm{d} x \\ & \qquad+\int_{d(x, o')\geq d(o, o')/2} \left(\rho_0(x)+(1-\|\rho_0\|_{L^1({\mathbb H}^n)})\rho_{o'}(x)\right)^m\mathrm{d} x. \end{aligned} \end{align} $$

Substitute

$$\begin{align*}a(x)=(1-\|\rho_0\|_{L^1({\mathbb H}^n)})\rho_{o'}(x), \quad b(x)=\rho_0(x), \quad \text{ and } \quad S=\{x:d(x, o)\geq d(o, o')/2\} \end{align*}$$

into (6.12), to get

(6.14) $$ \begin{align} \begin{aligned} &\int_{d(x, o)\geq d(o, o')/2} \left(\rho_0(x)+ (1-\|\rho_0\|_{L^1({\mathbb H}^n)})\rho_{o'}(x)\right)^m\mathrm{d} x\\&\quad-\int_{d(x, o)\geq d(o, o')/2} \left((1-\|\rho_0\|_{L^1({\mathbb H}^n)})\rho_{o'}(x) \right)^m\mathrm{d} x\\ & \quad \leq m2^{m-1}\left(\int_{d(x, o)\geq d(o, o')/2} \rho_0(x)^m\mathrm{d} x\right)^{\frac{1}{m}}\\ &\qquad\times\left(\int_{d(x, o)\geq d(o, o')/2} \rho_0(x)^m\mathrm{d} x+\int_{d(x, o)\geq d(o, o')/2} \left((1-\|\rho_0\|_{L^1({\mathbb H}^n)})\rho_{o'}(x)\right)^m\mathrm{d} x\right)^{\frac{m-1}{m}}\\ & \quad \leq m2^{m-1}\left(\int_{d(x, o)\geq d(o, o')/2} \rho_0(x)^m\mathrm{d} x\right)^{\frac{1}{m}} \\&\qquad\left(\int_{{\mathbb H}^n} \rho_0(x)^m\mathrm{d} x+\int_{{\mathbb H}^n} \left( (1-\|\rho_0\|_{L^1({\mathbb H}^n)})\rho_{o'}(x) \right)^m\mathrm{d} x\right)^{\frac{m-1}{m}}. \end{aligned} \end{align} $$

Similarly, substitute

$$\begin{align*}a(x)=\rho_0(x),\quad b(x)=(1-\|\rho_0\|_{L^1({\mathbb H}^n)})\rho_{o'}(x), \quad \text{ and } \quad S=\{x:d(x, o')\geq d(o, o')/2\} \end{align*}$$

into (6.12), to get

(6.15) $$ \begin{align} \begin{aligned} &\int_{d(x, o')\geq d(o, o')/2} \left(\rho_0(x)+(1-\|\rho_0\|_{L^1({\mathbb H}^n)})\rho_{o'}(x)\right)^m\mathrm{d} x-\int_{d(x, o')\geq d(o, o')/2} \rho_0(x)^m\mathrm{d} x\\ &\quad \leq m2^{m-1}\left(\int_{d(x, o')\geq d(o, o')/2} \left((1-\|\rho_0\|_{L^1({\mathbb H}^n)})\rho_{o'}(x)\right)^m\mathrm{d} x\right)^{\frac{1}{m}}\\ &\qquad\times \left(\int_{d(x, o')\geq d(o, o')/2} \rho_0(x)^m\mathrm{d} x+\int_{d(x, o')\geq d(o, o')/2} \left( (1-\|\rho_0\|_{L^1({\mathbb H}^n)})\rho_{o'}(x) \right)^m\mathrm{d} x\right)^{\frac{m-1}{m}}\\ & \quad \leq m2^{m-1}\left(\int_{d(x, o')\geq d(o, o')/2} \left( (1-\|\rho_0\|_{L^1({\mathbb H}^n)})\rho_{o'}(x) \right)^m\mathrm{d} x\right)^{\frac{1}{m}} \\ & \qquad \times \left(\int_{{\mathbb H}^n} \rho_0(x)^m\mathrm{d} x+\int_{{\mathbb H}^n} \left( (1-\|\rho_0\|_{L^1({\mathbb H}^n)})\rho_{o'}(x) \right)^m\mathrm{d} x\right)^{\frac{m-1}{m}}. \end{aligned} \end{align} $$

Now combine (6.13), (6.14), and (6.15) to get

(6.16) $$ \begin{align} \begin{aligned} &\int_{{\mathbb H}^n} \left(\rho_0(x)+(1-\|\rho_0\|_{L^1({\mathbb H}^n)})\rho_{o'}(x)\right)^m\mathrm{d} x\\ &\quad \leq \int_{d(x, o)\geq d(o, o')/2} \left( ( 1-\|\rho_0\|_{L^1({\mathbb H}^n)})\rho_{o'}(x)\right)^m\mathrm{d} x+\int_{d(x, o')\geq d(o, o')/2} \rho_0(x)^m\mathrm{d} x\\ &\qquad + m2^{m-1}\left(\int_{d(x, o)\geq d(o, o')/2} \rho_0(x)^m\mathrm{d} x\right)^{\frac{1}{m}}\\&\qquad\left(\int_{{\mathbb H}^n} \rho_0(x)^m\mathrm{d} x+\int_{{\mathbb H}^n} \left( (1-\|\rho_0\|_{L^1({\mathbb H}^n)})\rho_{o'}(x)\right)^m\mathrm{d} x\right)^{\frac{m-1}{m}}\\ &\quad\quad + m2^{m-1}\left(\int_{d(x, o')\geq d(o, o')/2} \left( (1-\|\rho_0\|_{L^1({\mathbb H}^n)} )\rho_{o'}(x) \right)^m\mathrm{d} x\right)^{\frac{1}{m}} \\ &\qquad \times \left(\int_{{\mathbb H}^n} \rho_0(x)^m\mathrm{d} x+\int_{{\mathbb H}^n} \left( (1-\|\rho_0\|_{L^1({\mathbb H}^n)})\rho_{o'}(x) \right)^m\mathrm{d} x\right)^{\frac{m-1}{m}}. \end{aligned} \end{align} $$

Since $\int _{{\mathbb H}^n}\rho _k(x)^m\mathrm {d} x$ is uniformly bounded from above in $k\geq 1$ (see Proposition 4.1), from the lower semicontinuity of the entropy functional we infer that $\int _{{\mathbb H}^n}\rho _0(x)^m\mathrm {d} x $ is also bounded from above. Also, we know that $\int _{{\mathbb H}^n}\rho (x)^m\mathrm {d} x$ is bounded, as $\rho \in L^m({\mathbb H}^n)$ . Hence, set

(6.17) $$ \begin{align} U:=\max\left(\int_{{\mathbb H}^n}\rho_0(x)^m\mathrm{d} x, \int_{{\mathbb H}^n} \left( (1-\|\rho_0\|_{L^1({\mathbb H}^n)})\rho(x) \right)^m\mathrm{d} x\right) < \infty. \end{align} $$

Then, we have

$$\begin{align*}\left(\int_{{\mathbb H}^n} \rho_0(x)^m\mathrm{d} x+\int_{{\mathbb H}^n} \left((1-\|\rho_0\|_{L^1({\mathbb H}^n)})\rho_{o'}(x)\right)^m\mathrm{d} x\right)^{\frac{m-1}{m}}\leq (2U)^{\frac{m-1}{m}}, \end{align*}$$

which used in (6.16) yields

(6.18) $$ \begin{align} \begin{aligned} &\int_{{\mathbb H}^n} \left(\rho_0(x)+(1-\|\rho_0\|_{L^1({\mathbb H}^n)})\rho_{o'}(x)\right)^m\mathrm{d} x\\ & \quad \leq \int_{d(x, o)\geq d(o, o')/2} \left((1-\|\rho_0\|_{L^1({\mathbb H}^n)})\rho_{o'}(x) \right)^m\mathrm{d} x+\int_{d(x, o')\geq d(o, o')/2} \rho_0(x)^m\mathrm{d} x\\[2pt] &\qquad +m2^{m-1} \left(2U\right)^{\frac{m-1}{m}} \\ & \qquad \quad \times \left( \left(\int_{d(x, o)\geq d(o, o')/2} \rho_0(x)^m\mathrm{d} x\right)^{\frac{1}{m}}\right.\\&\quad \left.+\left(\int_{d(x, o')\geq d(o, o')/2} \left( (1-\|\rho_0\|_{L^1({\mathbb H}^n)})\rho_{o'}(x) \right)^m\mathrm{d} x\right)^{\frac{1}{m}} \right) \\ & \quad \leq \int_{{\mathbb H}^n} \left((1-\|\rho_0\|_{L^1({\mathbb H}^n)})\rho_{o'}(x) \right)^m\mathrm{d} x+\int_{{\mathbb H}^n} \rho_0(x)^m\mathrm{d} x\\[2pt] &\qquad + m2^{m-\frac{1}{m}}U^{\frac{m-1}{m}} \\ &\qquad \quad \times \left( \left(\int_{d(x, o)\geq d(o, o')/2}\rho_0(x)^m\mathrm{d} x\right)^{\frac{1}{m}} \right.\\&\qquad\left.+\left(\int_{d(x, o')\geq d(o, o')/2} \left( (1-\|\rho_0\|_{L^1({\mathbb H}^n)})\rho_{o'}(x) \right)^m\mathrm{d} x\right)^{\frac{1}{m}} \right). \end{aligned} \end{align} $$

By the property of isometries, we have

$$\begin{align*}&\int_{d(x, o')\geq d(o, o')/2} \left((1-\|\rho_0\|_{L^1({\mathbb H}^n)})\rho_{o'}(x) \right)^m\mathrm{d} x\\&\quad=\int_{d(x, o)\geq d(o, o')/2} \left( (1-\|\rho_0\|_{L^1({\mathbb H}^n)})\rho(x) \right)^m\mathrm{d} x, \end{align*}$$

and

$$\begin{align*}\int_{{\mathbb H}^n} \left( (1-\|\rho_0\|_{L^1({\mathbb H}^n)})\rho_{o'}(x) \right)^m\mathrm{d} x=\int_{{\mathbb H}^n} \left( (1-\|\rho_0\|_{L^1({\mathbb H}^n)})\rho(x) \right)^m\mathrm{d} x, \end{align*}$$

and these used in (6.18) yield

(6.19) $$ \begin{align} \begin{aligned} &\int_{{\mathbb H}^n} \left(\rho_0(x)+(1-\|\rho_0\|_{L^1({\mathbb H}^n)})\rho_{o'}(x)\right)^m\mathrm{d} x\\[2pt] & \quad \leq \int_{{\mathbb H}^n} \left((1-\|\rho_0\|_{L^1({\mathbb H}^n)})\rho(x) \right)^m\mathrm{d} x+\int_{{\mathbb H}^n} \rho_0(x)^m\mathrm{d} x\\ & \qquad + m2^{m-\frac{1}{m}}U^{\frac{m-1}{m}} \\ & \qquad \quad \times \left( \left(\int_{d(x, o)\geq d(o, o')/2}\rho_0(x)^m\mathrm{d} x\right)^{\frac{1}{m}}\right.\\&\quad\left.+\left(\int_{d(x, o)\geq d(o, o')/2} ((1-\|\rho_0\|_{L^1({\mathbb H}^n)})\rho(x))^m\mathrm{d} x\right)^{\frac{1}{m}} \right). \end{aligned} \end{align} $$

Now, choose an arbitrary $\epsilon $ that satisfies

(6.20) $$ \begin{align} 0<\epsilon<-\frac{E[(1-\|\rho_0\|_{L^1({\mathbb H}^n)})\rho]}{m2^{m+1-\frac{1}{m}}U^{\frac{m-1}{m}}}. \end{align} $$

Note that this is possible, since $E[(1-\|\rho _0\|_{L^1({\mathbb H}^n)})\rho ]<0$ by (6.6). From the boundedness of integrals (see (6.17)), there exists $L>0$ (depending on $\epsilon $ , $\rho _0$ and $\rho $ ) such that

$$\begin{align*}\int_{d(x, o)\geq L}\rho_0(x)^m\mathrm{d} x<\epsilon^m\qquad\text{and}\qquad \int_{d(x, o)\geq L}((1-\|\rho_0\|_{L^1({\mathbb H}^n)})\rho(x))^m\mathrm{d} x<\epsilon^m. \end{align*}$$

Hence, for any $o'$ which satisfies $d(o,o')\geq 2L$ , we get from (6.19) that

(6.21) $$ \begin{align} \begin{aligned} &\int_{{\mathbb H}^n} \left(\rho_0(x)+(1-\|\rho_0\|_{L^1({\mathbb H}^n)})\rho_{o'}(x)\right)^m\mathrm{d} x\\[2pt] & \quad \leq \int_{{\mathbb H}^n} \left((1-\|\rho_0\|_{L^1({\mathbb H}^n)})\rho(x) \right)^m\mathrm{d} x+\int_{{\mathbb H}^n} \rho_0(x)^m\mathrm{d} x +m2^{m-\frac{1}{m}}U^{\frac{m-1}{m}}\left(\epsilon+\epsilon\right)\\[2pt] & \quad < \int_{{\mathbb H}^n} \left((1-\|\rho_0\|_{L^1({\mathbb H}^n)})\rho(x)\right)^m\mathrm{d} x+\int_{{\mathbb H}^n} \rho_0(x)^m\mathrm{d} x-E[(1-\|\rho_0\|_{L^1({\mathbb H}^n)})\rho], \end{aligned} \end{align} $$

where in the last inequality we used the assumption (6.20) on $\epsilon $ .

Finally, we use (6.21) in (6.8) to get

$$ \begin{align*} &E[\rho_0+(1-\|\rho_0\|_{L^1({\mathbb H}^n)})\rho_{o'}]\\[2pt] &\quad < \int_{{\mathbb H}^n} \left((1-\|\rho_0\|_{L^1({\mathbb H}^n)})\rho(x) \right)^m\mathrm{d} x+\int_{{\mathbb H}^n} \rho_0(x)^m\mathrm{d} x-E[(1-\|\rho_0\|_{L^1({\mathbb H}^n)})\rho]\\[2pt] &\qquad + \frac{1}{2}\iint\limits_{{\mathbb H}^n \times {\mathbb H}^n}h(d(x, y))\rho_0(x)\rho_0(y)\mathrm{d} x\mathrm{d} y \\&\qquad+ \frac{1}{2} (1-\|\rho_0\|_{L^1({\mathbb H}^n)})^2 \iint\limits_{{\mathbb H}^n \times {\mathbb H}^n}h(d(x, y))\rho(x)\rho(y)\mathrm{d} x\mathrm{d} y\\ &\quad = E[\rho_0]. \end{align*} $$

This shows the claim - see (6.7). Similar to Case 1, (6.7) combined with (6.4), together with the fact that $\rho _0+(1-\|\rho _0\|_{L^1({\mathbb H}^n)})\rho _{o'} \in \mathcal {P}({\mathbb H}^n)$ , leads to the following contradiction:

$$\begin{align*}E_{\inf} \leq E[\rho_0+(1-\|\rho_0\|_{L^1({\mathbb H}^n)})\rho_{o'}] < \liminf_{l\to\infty}E[\rho_{k_l}] = E_{\inf}. \end{align*}$$

We conclude that $\rho _0\in \mathcal {P}({\mathbb H}^n)$ and hence, $\rho _0$ is a global minimizer of the energy functional.

Remark 6.2 Note that in the case $h_\infty = \infty $ (Theorem 5.1) we were able to show that minimizing sequences are tight. Hence, Prokhorov’s theorem applies and minimizing sequences converge weak- $^{*}$ (on subsequences) to probability measures. In the case $h_\infty = 0,$ we could not ensure tightness of minimizing sequences, so we had to use Theorem 2.1 instead, which resulted in a more involved argument to prove existence of energy minimizers.

Remark 6.3 By similar arguments, one can prove that Theorem 6.1 holds also in the Euclidean space $\mathbb R^n$ . Such result (not stated separately here) generalizes [Reference Carrillo, Delgadino and Patacchini10, Theorem 5.1], which only considers Riesz potentials.

A Proof of Proposition 2.1

On a Cartan–Hadamard manifold the sectional curvatures satisfy (2.4) and hence (2.5) holds. Fix a pole $o \in M$ and consider the following family of probability density functions $\rho _R$ that depend on $R>0$ :

(A.1) $$ \begin{align} \rho_{R}(x)=\begin{cases} \displaystyle\frac{1}{|B_R(o)|},&\quad\text{when }x\in B_R(o),\\[10pt] 0,&\quad\text{otherwise}. \end{cases} \end{align} $$

Assume first that h satisfies (2.6). We will estimate $E[\rho _R]$ in the limit $R \to 0$ . The entropy term can be calculated from (A.1) and then estimated using (2.5) as

(A.2) $$ \begin{align} \frac{1}{m-1}\int_M \rho_R(x)^{m} \mathrm{d} x&=\frac{1}{(m-1)|B_R(o)|^{m-1}} \nonumber \\ & \leq \frac{1}{(m-1)\left(w(n) R^n \right)^{m-1}}. \end{align} $$

On the other hand, the interaction energy can be estimated as

(A.3) $$ \begin{align} \frac{1}{2}\iint\limits_{M \times M} h(d(x, y))\rho_R(x)\rho_R(y)\mathrm{d} x\mathrm{d} y &\leq \frac{1}{2}\iint\limits_{M \times M} h(2R)\rho_R(x)\rho_R(y)\mathrm{d} x \mathrm{d} y \nonumber \\[2pt] &=\frac{h(2R)}{2}, \end{align} $$

where we used that h is non-decreasing and $\sup _{x, y\in \mathrm {supp}(\rho _R)}d(x, y)= 2R$ . By combining (A.2) and (A.3), we then get

(A.4) $$ \begin{align} E[\rho_R] \leq \frac{1}{(m-1)w(n)^{m-1}R^{n(m-1)}}+\frac{h(2R)}{2}. \end{align} $$

If the r.h.s. of (A.4) satisfies

$$\begin{align*}\lim_{R\to 0+}\left(\frac{1}{(m-1)w(n)^{m-1}R^{n(m-1)}}+\frac{h(2R)}{2}\right)=-\infty, \end{align*}$$

then $\lim _{R\to 0+}E[\rho _R]=-\infty $ , and a global minimizer cannot exist (the attraction is too strong and blow-up occurs). Note that we can write

$$\begin{align*}\lim_{R\to 0+}\left(\frac{1}{(m-1)w(n)^{m-1}R^{n(m-1)}}+\frac{h(2R)}{2}\right)=\lim_{R\to 0+}\left(\frac{2^{n(m-1)}}{(m-1)w(n)^{m-1}R^{n(m-1)}}+\frac{h(R)}{2}\right). \end{align*}$$

These considerations prove the statement if h satisfies (2.6).

Now, assume h satisfies (2.7). Then, there exist $\theta _0>0$ and $\alpha>0$ such that

(A.5) $$ \begin{align} \theta^n h(\theta)\leq -\alpha,\quad\text{for all}~0<\theta<2\theta_0. \end{align} $$

Consider $\rho _{\theta _0}$ as defined in (A.1) (use $R=\theta _0$ ). Then, its energy can be written as

(A.6) $$ \begin{align} E[\rho_{\theta_0}]=\frac{1}{(m-1)|B_{\theta_0}(o)|^{m-1}}+\frac{1}{2}\iint\limits_{M\times M}h(d(x, y))\rho_{\theta_0}(x)\rho_{\theta_0}(y)\mathrm{d} x\mathrm{d} y. \end{align} $$

As h is singular at $0$ , the above interaction energy part can be expressed as

$$\begin{align*}\iint\limits_{M\times M}h(d(x, y))\rho_{\theta_0}(x)\rho_{\theta_0}(y)\mathrm{d} x\mathrm{d} y=\lim_{\delta\to0+}\iint\limits_{d(x, y)\geq \delta}h(d(x, y))\rho_{\theta_0}(x)\rho_{\theta_0}(y)\mathrm{d} x\mathrm{d} y. \end{align*}$$

From the definition of $\rho _{\theta _0}$ , we get

$$\begin{align*}\iint\limits_{d(x, y)\geq \delta}h(d(x, y))\rho_{\theta_0}(x)\rho_{\theta_0}(y)\mathrm{d} x\mathrm{d} y=\frac{1}{|B_{\theta_0}(o)|^2}\iint\limits_{\substack{\delta\leq d(x, y)\\ \theta_x, \theta_y\leq \theta_0}}h(d(x, y))\mathrm{d} x\mathrm{d} y. \end{align*}$$

Since

$$\begin{align*} \{(x, y)\in M\times M: \delta\leq d(x, y) &\text{ and } \theta_x, \theta_y\leq \theta_0\} \supset \\& \{(x, y)\in M\times M: \delta\leq d(x, y)\leq \theta_0/2 \text{ and } \theta_x\leq \theta_0/2\}, \end{align*}$$

and $h(d(x, y))\leq 0$ for all $(x, y)\in \{(x, y)\in M\times M: \delta \leq d(x, y) \text { and } \theta _x, \theta _y\leq \theta _0\}$ , we also have

$$\begin{align*}\iint\limits_{\substack{\delta\leq d(x, y) \\ \theta_x, \theta_y\leq \theta_0}}h(d(x, y))\mathrm{d} x\mathrm{d} y\leq \int_{\theta_x\leq \theta_0/2}\int_{\delta\leq d(x, y)\leq \theta_0/2}h(d(x, y))\mathrm{d} y\mathrm{d} x. \end{align*}$$

Now, use (A.5) to get

$$ \begin{align*} \int_{\theta_x\leq \theta_0/2}\int_{\delta\leq d(x, y)\leq \theta_0/2}h(d(x, y))\mathrm{d} y\mathrm{d} x&\leq -\alpha\int_{\theta_x\leq \theta_0/2}\int_{\delta\leq d(x, y)\leq \theta_0/2}d(x, y)^{-n}\mathrm{d} y\mathrm{d} x\\ &= -\alpha\int_{\theta_x\leq \theta_0/2}\int_\delta^{\theta_0/2}r^{-n}|\partial B_r(o)|\mathrm{d} r \mathrm{d} x\\ &=-\alpha|B_{\theta_0/2}(o)|\int_\delta^{\theta_0/2}r^{-n}|\partial B_r(o)|\mathrm{d} r. \end{align*} $$

Using $|\partial B_r(o)|\geq n w(n) r^{n-1}$ in the above, we then find

$$ \begin{align*} \int_{\theta_x\leq \theta_0/2}\int_{\delta\leq d(x, y)\leq 2\theta_0/2}h(d(x, y))\mathrm{d} y\mathrm{d} x&\leq-\alpha|B_{\theta_0/2}(o)|n w(n)\int_\delta^{\theta_0/2}r^{-1}\mathrm{d} r\\ &=-\alpha|B_{\theta_0/2}(o)|n w(n)\ln\left(\frac{\theta_0}{2\delta}\right). \end{align*} $$

Finally, using these estimates in (A.6) we get

$$\begin{align*}E[\rho_{\theta_0}]\leq \frac{1}{(m-1)|B_{\theta_0}(o)|^{m-1}}-\frac{\alpha|B_{\theta_0/2}(o)|n w(n)}{2|B_{\theta_0}(o)|^2}\lim_{\delta\to0+}\ln\left(\frac{\theta_0}{2\delta}\right)=-\infty. \end{align*}$$

This implies $E[\rho _{\theta _0}]=-\infty $ , which completes the argument.

B Proof of Proposition 2.2

Since $\rho \in L^m(M)$ , the entropy part is well-defined. We need to show that the interaction energy is also well-defined. The interaction energy can be decomposed as

$$ \begin{align*} &\iint\limits_{M\times M}h(d(x, y))\rho(x)\rho(y)\mathrm{d} x\mathrm{d} y\\ &\quad =\iint\limits_{h(d(x, y))\leq0}h(d(x, y))\rho(x)\rho(y)\mathrm{d} x\mathrm{d} y+\iint\limits_{h(d(x, y))>0}h(d(x, y))\rho(x)\rho(y)\mathrm{d} x\mathrm{d} y. \end{align*} $$

If h is singular at zero and $\lim _{\theta \to \infty }h(\theta )=\infty $ , then we may have

$$\begin{align*}&\iint\limits_{h(d(x, y))\leq0}h(d(x, y))\rho(x)\rho(y)\mathrm{d} x\mathrm{d} y=-\infty,\\&\quad \iint\limits_{h(d(x, y))>0}h(d(x, y))\rho(x)\rho(y)\mathrm{d} x\mathrm{d} y=\infty, \end{align*}$$

and in such situation, the interaction energy may not be well-defined. However, if h satisfies (2.8), then we can prove $\iint\limits _{h(d(x, y))\leq 0}h(d(x, y))\rho (x)\rho (y)\mathrm {d} x\mathrm {d} y>-\infty $ . To show this, we start from Lemma 2.1, and estimate

(B.1) $$ \begin{align} \begin{aligned} &\iint\limits_{h(d(x, y))\leq 0}h(d(x, y))\rho(x)\rho(y)\mathrm{d} x\mathrm{d} y\\&\quad\geq \iint\limits_{h(d(x, y))\leq 0}(-\gamma_1d(x, y)^{-\alpha}-\gamma_2)\rho(x)\rho(y)\mathrm{d} x\mathrm{d} y\\ &\quad\geq -\gamma_1\iint\limits_{M\times M}d(x, y)^{-\alpha}\rho(x)\rho(y)\mathrm{d} x\mathrm{d} y-\gamma_2. \end{aligned} \end{align} $$

Then, by the HLS inequality derived in Section 3 (Theorem 3.2), we get

(B.2) $$ \begin{align} \iint\limits_{M\times M}d(x, y)^{-\alpha}\rho(x)\rho(y)\mathrm{d} x\mathrm{d} y\leq r^{-\alpha}+\tilde{C}(\alpha, n, r, c_m)\left(\int_M\rho(x)^m\mathrm{d} x\right)^{\frac{\alpha}{n(m-1)}}, \end{align} $$

for a fixed $r>0$ and a constant $\tilde {C}$ . Now, by combining (B.1) and (B.2), we find

$$\begin{align*}&\iint\limits_{h(d(x, y))\leq 0}h(d(x, y))\rho(x)\rho(y)\mathrm{d} x\mathrm{d} y\\&\quad\geq -\gamma_1\tilde{C}(\alpha, n, r, c_m)\left(\int_M\rho(x)^m\mathrm{d} x\right)^{\frac{\alpha}{n(m-1)}}-(\gamma_1r^{-\alpha}+\gamma_2), \end{align*}$$

and since $\rho \in L^m(M)$ , we conclude $\iint\limits _{h(d(x, y))\leq 0}h(d(x, y))\rho (x)\rho (y)\mathrm {d} x\mathrm {d} y>-\infty $ .

C Proof of Proposition 2.3

From Lemma 2.1, for any $\rho _k\in \mathcal {P}_{ac}(M)$ , we get

$$\begin{align*}E[\rho_k]\geq \frac{1}{m-1}\int_M \rho_k(x)^m\mathrm{d} x -\frac{\gamma_1}{2}\iint\limits_{M\times M}\frac{\rho_k(x)\rho_k(y)\mathrm{d} x\mathrm{d} y}{d(x, y)^\alpha}-\frac{\gamma_2}{2}, \end{align*}$$

for two non-negative constants $\gamma _1$ and $\gamma _2$ . Also, by the HLS inequality from Theorem 3.2 we estimate

$$\begin{align*}\iint\limits_{M \times M}\frac{\rho_k(x)\rho_k(y)\mathrm{d} x\mathrm{d} y}{d(x, y)^\alpha}\leq r^{-\alpha} +\tilde{C}(\alpha, n, r, c_m)\left(\int_M\rho_k(x)^m\mathrm{d} x\right)^{\frac{\alpha}{n(m-1)}}, \end{align*}$$

for some $r>0$ . By combining the two inequalities above, we then get

(C.1) $$ \begin{align} E[\rho_k]\geq \frac{1}{m-1}\int_M \rho_k(x)^m\mathrm{d} x-\frac{\gamma_1\tilde{C}(\lambda, n, r, c_m)}{2}\left(\int_M\rho_k(x)^m\mathrm{d} x\right)^{\frac{\alpha}{n(m-1)}}-\frac{1}{2}(\gamma_1 r^{-\alpha}+\gamma_2). \end{align} $$

If $\rho $ contains a singular part, then for any sequence $\{\rho _k\}_{k\geq 1}\subset L^m(M) \cap \mathcal {P}(M)$ such that $\rho _k\rightharpoonup \rho $ , one has

$$\begin{align*}\lim_{k\to\infty}\int_M\rho_k(x)^m\mathrm{d} x=\infty. \end{align*}$$

Since $0<\alpha <n(m-1)$ , this implies that the right hand side of (C.1) tends to infinity, and hence, $E[\rho ]=\infty $ .

Acknowledgements

J.A.C. acknowledge support by the EPSRC grant EP/V051121/1. R.F. acknowledges an NSERC Alliance International Catalyst grant, which supported H. Park’s visit to University of Oxford, where the research presented in this article was initiated.

Footnotes

J.A.C. was partially supported by the Advanced Grant Nonlocal-CPD (Nonlocal PDEs for Complex Particle Dynamics: Phase Transitions, Patterns, and Synchronization) of the European Research Council Executive Agency (ERC) under the European Union Horizon 2020 research and innovation programme (Grant agreement No. 883363). J.A.C. was also partially supported by the “Maria de Maeztu” Excellence Unit IMAG, reference CEX2020-001105-M, funded by MCIN/AEI/10.13039/501100011033/. R.F. was supported by NSERC Discovery Grant PIN-341834 during this research.

References

Ambrosio, L., Fusco, N., and Pallara, D., Functions of bounded variation and free discontinuity problems, Oxford Mathematical Monographs, Oxford University Press, Oxford, 2000.10.1093/oso/9780198502456.001.0001CrossRefGoogle Scholar
Ambrosio, L., Gigli, N., and Savaré, G., Gradient flows in metric spaces and in the space of probability measures, Lectures in Mathematics ETH Zürich, Birkhäuser Verlag, Basel, 2005.Google Scholar
Bedrossian, J., Global minimizers for free energies of subcritical aggregation equations with degenerate diffusion . Appl. Math. Lett. 24(2011), no. 11, 19271932.CrossRefGoogle Scholar
Blanchet, A., Carrillo, J. A., and Laurençot, P., Critical mass for a Patlak-Keller-Segel model with degenerate diffusion in higher dimensions . Calc. Var. Partial Differ. Equ. 35(2009), no. 2, 133168.10.1007/s00526-008-0200-7CrossRefGoogle Scholar
Burchard, A. and Schmuckenschläger, M., Comparison theorem for exit times . Geom. Funct. Anal. 11(2001), 651692.CrossRefGoogle Scholar
Calvez, V., Carrillo, J. A., and Hoffmann, F., Equilibria of homogeneous functionals in the fair-competition regime . Nonlinear Anal. 159(2017), 85128.10.1016/j.na.2017.03.008CrossRefGoogle Scholar
Calvez, V., Carrillo, J. A., and Hoffmann, F., Uniqueness of stationary states for singular Keller-Segel type models . Nonlinear Anal. 205(2021), Paper No. 112222. 24.CrossRefGoogle Scholar
Carlen, E. and Loss, M., Competing symmetries, the logarithmic HLS inequality and Onofri’s inequality on ${S}^n$ . Geom. Funct. Anal. 2(1992), 90104.10.1007/BF01895706CrossRefGoogle Scholar
Carrillo, J. A., Craig, K., and Yao, Y., Aggregation-diffusion equations: Dynamics, asymptotics, and singular limits . In: Bellomo, N., Degond, P., and Tadmor, E. (eds.), Active particles, Volume 2: Advances in Theory, Models, and Applications, Modeling and Simulation in Science, Engineering and Technology, Birkhäuser, Basel, 2019, pp. 65108.CrossRefGoogle Scholar
Carrillo, J. A., Delgadino, M., and Patacchini, F. S., Existence of ground states for aggregation-diffusion equations . Anal. Appl. 17(2019), no. 3, 393423.CrossRefGoogle Scholar
Carrillo, J. A., Hittmeir, S., Volzone, B., and Yao, Y., Nonlinear aggregation-diffusion equations: radial symmetry and long time asymptotics . Invent. Math. 218(2019), 889977.10.1007/s00222-019-00898-xCrossRefGoogle Scholar
Carrillo, J. A., Hoffmann, F., Mainini, E., and Volzone, B., Ground states in the diffusion-dominated regime . Calc. Var. Partial Differ. Equ. 57(2018), 127.CrossRefGoogle ScholarPubMed
Carrillo, J. A., McCann, R. J., and Villani, C., Contractions in the 2-Wasserstein length space and thermalization of granular media . Arch. Ration. Mech. Anal. 179(2006), no. 2, 217263.10.1007/s00205-005-0386-1CrossRefGoogle Scholar
Carrillo, J. A. and Vázquez, J. L., Some free boundary problems involving non-local diffusion and aggregation . Philos. Trans. Roy. Soc. A 373(2015), no. 2050, 20140275. 16.CrossRefGoogle ScholarPubMed
Chavel, I., Riemannian geometry: A modern introduction. 2nd ed., Cambridge Studies in Advanced Mathematics, Cambridge University Press, New York, 2006.10.1017/CBO9780511616822CrossRefGoogle Scholar
Delgadino, M. G., Yan, X., and Yao, Y., Uniqueness and nonuniqueness of steady states of aggregation-diffusion equations . Commun. Pure Appl. Math. 75(2022), no. 1, 359.CrossRefGoogle Scholar
do Carmo, M. P., Riemannian geometry. 2nd ed., Mathematics: Theory and Applications, Birkhäuser, Boston, MA, 1992.CrossRefGoogle Scholar
Evans, L. C. and Gariepy, R. F., Measure theory and fine properties of functions, Textbooks in Mathematics, CRC Press, Boca Raton, FL, revised edition, 2015.10.1201/b18333CrossRefGoogle Scholar
Fernández-Jiménez, A., Carrillo, J. A., and Gómez-Castro, D., Partial mass concentration for fast-diffusions with non-local aggregation terms. Available at SSRN 4452239, 2023.10.2139/ssrn.4452239CrossRefGoogle Scholar
Fetecau, R. C., Ha, S.-Y., and Park, H., An intrinsic aggregation model on the special orthogonal group SO(3): Well-posedness and collective behaviours . J. Nonlinear Sci. 31(2021), no. 5, 74.10.1007/s00332-021-09732-2CrossRefGoogle Scholar
Fetecau, R. C. and Park, H., Equilibria and energy minimizers for an interaction model on the hyperbolic space . Physica D 446(2023), 133670.10.1016/j.physd.2023.133670CrossRefGoogle Scholar
Fetecau, R. C. and Park, H., Long-time behaviour of interaction models on Riemannian manifolds with bounded curvature . J. Geom. Anal. 33(2023), no. 7, 218.CrossRefGoogle Scholar
Fetecau, R. C. and Park, H., Aggregation-diffusion energies on Cartan-Hadamard manifolds of unbounded curvature . J. Geom. Anal. 34(2024), no. 12, article 356.Google Scholar
Fetecau, R. C. and Park, H., Ground states for aggregation-diffusion models on Cartan-Hadamard manifolds. J. Lond. Math. Soc. 111(2025), no. 2, e70079.Google Scholar
Fetecau, R. C., Park, H., and Patacchini, F. S., Well-posedness and asymptotic behaviour of an aggregation model with intrinsic interactions on sphere and other manifolds . Anal. Appl. 19(2021), no. 6, 9651017.CrossRefGoogle Scholar
Fetecau, R. C. and Patacchini, F. S., Well-posedness of an interaction model on Riemannian manifolds . Commun. Pure Appl. Anal. 21(2021), no. 11, 35593585.10.3934/cpaa.2022114CrossRefGoogle Scholar
Fetecau, R. C. and Zhang, B., Self-organization on Riemannian manifolds . J. Geom. Mech. 11(2019), no. 3, 397426.10.3934/jgm.2019020CrossRefGoogle Scholar
Gazi, V. and Passino, K. M., Stability analysis of swarms . In: Proceedings of the American control conference, American Automatic Control Council, Anchorage, AK, 2002, pp. 810.Google Scholar
Grillo, G., Meglioli, G., and Punzo, F., Global existence of solutions and smoothing effects for classes of reaction–diffusion equations on manifolds . J. Evol. Equ. 21(2021), no. 2, 23392375.10.1007/s00028-021-00685-3CrossRefGoogle Scholar
Grillo, G., Meglioli, G., and Punzo, F., Smoothing effects and infinite time blowup for reaction-diffusion equations: An approach via Sobolev and Poincaré inequalities . J. Math Pures Appl. 151(2021), 99131.10.1016/j.matpur.2021.04.011CrossRefGoogle Scholar
Grillo, G., Meglioli, G., and Punzo, F., Blow-up versus global existence of solutions for reaction–diffusion equations on classes of Riemannian manifolds . Annali di Matematica Pura ed Applicata 202(2023), no. 3, 12551270.CrossRefGoogle Scholar
Grillo, G., Muratori, M., and Vázquez, J. L., The porous medium equation on Riemannian manifolds with negative curvature. The large-time behaviour . Adv. Math. 314(2017), 328377.CrossRefGoogle Scholar
Grillo, G., Muratori, M., and Vázquez, J. L., The porous medium equation on Riemannian manifolds with negative curvature: the superquadratic case . Mathematische Annalen 373(2019), 119153.10.1007/s00208-018-1680-1CrossRefGoogle Scholar
Ha, S.-Y., Hwang, S., Kim, D., Kim, S.-C., and Min, C., Emergent behaviors of a first-order particle swarm model on the hyperboloid . J. Math. Phys. 61(2020), no. 4, 42701.CrossRefGoogle Scholar
Ha, S.-Y., Kang, M., and Kim, D., Emergent behaviors of high-dimensional Kuramoto models on Stiefel manifolds . Automatica 136(2022), 110072.10.1016/j.automatica.2021.110072CrossRefGoogle Scholar
Ji, M. and Egerstedt, M., Distributed coordination control of multi-agent systems while preserving connectedness . IEEE Trans. Robot. 23(2007), no. 4, 693703.CrossRefGoogle Scholar
Kaib, G., Stationary states of an aggregation equation with degenerate diffusion and bounded attractive potential . SIAM J. Math. Anal. 49(2017), no. 1, 272296.CrossRefGoogle Scholar
Keller, E. F. and Segel, L. A., Initiation of slime mold aggregation viewed as an instability . J. Theor. Biol. 26(1970), no. 3, 399415.CrossRefGoogle Scholar
Kolokolnikov, T., Carrillo, J. A., Bertozzi, A., Fetecau, R., and Lewis, M., Emergent behaviour in multi-particle systems with non-local interactions [Editorial] . Phys. D 260(2013), 14.10.1016/j.physd.2013.06.011CrossRefGoogle Scholar
Lieb, E. and Loss, M., Analysis. 2nd ed., volume 14 of Graduate Studies in Mathematics, American Mathematical Society, 2001.Google Scholar
Motsch, S. and Tadmor, E., Heterophilious dynamics enhances consensus . SIAM Rev. 56(2014), 577621.CrossRefGoogle Scholar
Patlak, C. S., Random walk with persistence and external bias . Bull. Math. Biophys. 15(1953), 311338.10.1007/BF02476407CrossRefGoogle Scholar
Sire, C. and Chavanis, P.-H., Thermodynamics and collapse of self-gravitating Brownian particles in d dimensions . Phys. Rev. E 66(2002), no. 4, 46133.10.1103/PhysRevE.66.046133CrossRefGoogle ScholarPubMed
Sire, C. and Chavanis, P.-H., Critical dynamics of self-gravitating Langevin particles and bacterial populations . Phys. Rev. E 78(2008), no. 6, 61111.CrossRefGoogle ScholarPubMed
Topaz, C. M., Bertozzi, A. L., and Lewis, M. A., A nonlocal continuum model for biological aggregation . Bull. Math. Biol. 68(2006), 16011623.10.1007/s11538-006-9088-6CrossRefGoogle ScholarPubMed
Wu, L. and Slepčev, D., Nonlocal interaction equations in environments with heterogeneities and boundaries . Commun. Partial Differ, Equ. 40(2015), no. 7, 12411281.10.1080/03605302.2015.1015033CrossRefGoogle Scholar