- 1Laboratoire de Physique Théorique, CNRS, UPS, Université de Toulouse, Toulouse, France
- 2Kavli Institute for Theoretical Physics (KITP), University of California Santa Barbara, Santa Barbara, CA, United States
We review basic results on the Bose–Einstein condensate dark matter (BECDM) model. Self-gravitating BECs experience a collisionless process of gravitational cooling and violent relaxation, leading to BECDM halos with a “core-envelope” structure. The quantum core (soliton), which is the ground state of the Gross–Pitaevskii–Poisson (GPP) equations, may solve the core–cusp problem of the cold dark matter (CDM) model. The approximately isothermal envelope, resulting from the quantum interferences of the excited states, is similar to the Navarro–Frenk–White (NFW) profile of CDM halos and accounts for the flat rotation curves of the galaxies. We derive the core mass–radius relation, the halo mass–radius relation, and the core mass–halo mass relation of BECDM halos. We show that the core mass increases with the halo mass and we discuss the possibility that it collapses above a maximum mass arising from general relativity or from the attractive self-interaction of the bosons. We discuss the secular evolution of BECDM halos induced by the formation of granules (or quasiparticles) in the envelope, and we mention the analogy with the evolution of globular clusters. We also discuss the basic elements of BECDM cosmology. Throughout this review, we emphasize the importance of the maximum mass of dilute axion stars with an attractive self-interaction [P.H. Chavanis, Phys. Rev. D 84, 043531 (2011)] and its consequences.
1 Introduction
Even after 100 years of research, the nature of dark matter (DM) remains elusive. The cold dark matter (ΛCDM) model works extremely well at large (cosmological) scales and can account for precise measurements of the cosmic microwave background (CMB) from the Wilkinson Microwave Anisotropy Probe (WMAP) (Hinshaw et al., 2009) and Planck missions (Ade et al., 2014; Ade et al., 2016). However, in addition to the lack of evidence for any CDM particle such as a weakly interacting massive particle (WIMP) with a mass in the GeV–TeV range, the CDM model faces serious problems at small (galactic) scales.
One of these problems is the “core-cusp problem” (Moore et al., 1999). Classical
Another problem is the “missing satellites problem” (Kauffmann et al., 1993; Klypin et al., 1999; Kamionkowski and Liddle, 2000). For a cold classical gas, the Jeans length vanishes or is extremely small
for its characteristic mass and characteristic radius (this corresponds to a typical density
The “small-scale crisis of CDM” (Bullock and Boylan-Kolchin, 2017) is somehow related to the assumption that DM is pressureless. A first possibility to solve the CDM crisis is to take into account the feedback of the baryons that can transform cusps into cores (Romano-Díaz et al., 2008; Pontzen and Governato, 2014; Oñorbe et al., 2015). Another possibility is to consider self-interacting dark matter (SIDM) (Spergel and Steinhardt, 2000) or warm dark matter (WDM) (Bode et al., 2001), where a pressure arises as a result of particle dispersion. In that case, we have to deal with a gas at nonzero temperature
Some authors have considered the case where the DM particle is a fermion, like a massive sterile neutrino. In this model, gravitational collapse is prevented by the quantum pressure arising from the Pauli exclusion principle. Other authors have considered the case where the DM particle is a boson like the QCD axion1 or like an ultralight axion (ULA) or an axion-like particle (ALP) predicted by string theory. In this model, gravitational collapse is prevented by the quantum pressure arising from the Heisenberg uncertainty principle or by the scattering of the bosons when their self-interaction is repulsive. In these quantum (fermionic and bosonic) models, the minimum halo is interpreted as the ground state of the self-gravitating quantum gas at
We shall not review the literature on fermionic and bosonic DM in detail here because this literature is extensive, and this review has already been done in our previous contributions where we cite numerous references. A short history of fermionic DM can be found in the introductions of Chavanis (2006), Chavanis et al. (2015a), Chavanis et al. (2015b), Chavanis (2020d), Alberti and Chavanis (2020b), Chavanis (2023d), Chavanis (2022g) and a short history of bosonic DM can be found in the introductions of Chavanis (2011c), Chavanis (2016a), Chavanis (2018b), Chavanis (2019d), Chavanis (2020c), Chavanis (2023c). It is fascinating to note that the basic equations governing self-gravitating Fermi and Bose gases have been used to describe not only astrophysical objects (white dwarfs, neutron stars, fermion stars, boson stars, axion stars, and DM halos) but also extended elementary particles with an effective gravity accounting, for example, for the strong interaction (see the discussion in the introduction of Chavanis (2023c)). To some extent, the two communities working on these similar equations do not know each other.
In this paper, we focus on bosonic DM. At
The evolution of a self-gravitating BEC presents three main periods:
(i) Initially, the universe (assumed to be made of bosons in the form of BECs) is spatially homogeneous. If slightly perturbed,3 it experiences a quantum Jeans instability, leading to the growth of the perturbations and the formation of condensations (clumps) (Jeans, 1902; Khlopov et al., 1985; Bianchi et al., 1990; Hu et al., 2000; Sikivie and Yang, 2009; Chavanis, 2011c; Chavanis, 2012b; Suárez and Matos, 2011; Suárez and Chavanis, 2015a; Suárez and Chavanis, 2018; Chavanis, 2020b; Chavanis, 2021a). This corresponds to the linear regime of structure formation describing the initiation of the large-scale structures of the universe.
(ii) When the density contrast reaches a sufficiently large value, the condensations (overdensities) decouple from the Hubble flow, experience free fall and virialization, and grow through collisions, merging, and accretion before being diluted by the expansion of the universe. This corresponds to the nonlinear regime of structure formation, leading to the DM halos that are observed today.4 BECDM halos are formed by a dissipationless (“collisionless”) relaxation process called gravitational cooling (Seidel and Suen, 1994) or violent relaxation (Lynden-Bell, 1967). By this process, BECDM halos acquire a “core-envelope” structure with a quantum core (soliton) surrounded by an extended quasi-isothermal envelope that results from the quantum interferences of the excited states. This core–envelope structure has been evidenced in numerical simulations of the SP, GPP, and KGE equations (Seidel and Suen, 1994; Schive et al., 2014a; Schive et al., 2014b; Schwabe et al., 2016; Mocz et al., 2017; Mocz et al., 2018; Veltmaat et al., 2018; Mocz et al., 2019; Mocz et al., 2020; Veltmaat et al., 2020; Glennon and Prescod-Weinstein, 2021; Mocz et al., 2023; Liu et al., 2023; Nori et al., 2023; Painter et al., 2024). The quantum core (soliton) results from the balance between the gravitational attraction and the quantum pressure arising from the Heisenberg uncertainty principle or the self-interaction of the bosons. This is the ground state of the SP, GPP, or KGE equations. The envelope results from the balance between the gravitational attraction and an effective thermal pressure. The quantum core may solve the core–cusp problem of the CDM model and the approximately isothermal envelope, which is similar to the classical NFW profile, accounts for the flat rotation curves of the galaxies at large distances (Chavanis, 2019d).5 The mass of the quantum core
(iii) The envelope of BECDM halos has a granular structure as a result of quantum fluctuations (Schive et al., 2014a; Schive et al., 2014b; Hui et al., 2017). These granules (or quasiparticles) induce the secular evolution of the envelope (Levkov et al., 2018; Bar-Or et al., 2019; Marsh and Niemeyer, 2019; El-Zant et al., 2020; Bar-Or et al., 2021; Chavanis, 2021b). During this “collisional” process, some particles of the envelope condense and the mass of the soliton increases, being fed by the envelope (Levkov et al., 2018; Eggemeier and Niemeyer, 2019; Chen et al., 2021; Chen et al., 2022). This scenario was originally introduced to describe the Bose–Einstein condensation and the formation of axion stars in virialized DM halos and axion miniclusters (Levkov et al., 2018). It was then applied to the formation of inflaton stars in inflaton clusters (Musoke et al., 2020; Niemeyer and Easther, 2020; Eggemeier et al., 2021; Eggemeier et al., 2022).
There have been several recent reviews on the scalar field, fuzzy, ultralight, wave, or BEC dark matter model and on axion stars (Suárez et al., 2014; Rindler-Daller et al., 2014; Chavanis, 2015d; Marsh, 2016; Lee, 2018; Braaten and Zhang, 2019; Niemeyer, 2020; Chavanis, 2023b; Ferreira, 2021; Hui, 2021; Visinelli, 2021; Khoury, 2022; Matos et al., 2024; Padilla et al., 2024). In this paper, we review the basic aspects of self-gravitating BECs. We discuss results (mainly centered on our own contributions) that have not been treated in detail in the previous reviews. We only quote the literature closely related to these topics and refer to other reviews and to the introductions in our papers (Chavanis, 2011c; Chavanis, 2016a; Chavanis, 2018b; Chavanis, 2020c; Chavanis, 2023c) for additional references on a broader context.
This review is organized as follows. Section 2 discusses the KGE and GPP equations governing self-gravitating BECs in the relativistic and nonrelativistic regimes, respectively. In the following sections, we focus on the nonrelativistic limit. Section 3 introduces the Madelung (1927) hydrodynamic representation of the GPP equations (Böhmer and Harko, 2007; Sikivie and Yang, 2009; Chavanis, 2011c; Chavanis and Delfini, 2011; Chavanis, 2011a; Chavanis, 2012b; Rindler-Daller and Shapiro, 2012). Section 4 discusses the ground state of the GPP equations, called a soliton. This ground-state solution describes either the minimum halo or the quantum core of large DM halos. The core mass–radius relation can be obtained exactly by solving the GPP equations numerically, or approximately (analytically) by using a variational approach based on a Gaussian ansatz (Chavanis, 2011c; Chavanis and Delfini, 2011). In the case of self-gravitating BECs with no self-interaction or with a repulsive self-interaction, we recall the existence of a maximum mass set by general relativity (Kaup, 1968; Ruffini and Bonazzola, 1969; Colpi et al., 1986; Chavanis and Harko, 2012). In the case of self-gravitating BECs with a strongly attractive self-interaction, we show that equilibrium states exist only below a maximum mass obtained in (Chavanis, 2011c; Chavanis and Delfini, 2011) that has a nonrelativistic origin. It corresponds to the maximum mass of dilute axion stars (Chavanis, 2023b). We introduce an interpolation formula that connects these different limiting masses (Chavanis, 2023c). We also discuss the outcomes of the collapse above the maximum mass, leading to a dense axion star, a bosenova, a black hole, axion drops, fast radio bursts, etc (Braaten et al., 2016; Chavanis, 2016a; Chavanis, 2018b; Cotner, 2016; Davidson and Schwetz, 2016; Eby et al., 2016; Helfer et al., 2017; Levkov et al., 2017; Michel and Moss, 2018; Visinelli et al., 2018). Section 5 describes the core–envelope structure of BECDM halos resulting from gravitational cooling (Seidel and Suen, 1994) and violent relaxation (Lynden-Bell, 1967). We introduce a coarse-grained Wigner equation and derive hydrodynamic equations as well as a generalized GPP equation relaxing toward an equilibrium state with a core–envelope structure (Chavanis, 2022a). In Section 6, we develop a predictive model of BECDM halos (Chavanis, 2019d). In this model, the quantum core (soliton), corresponding to the ground state of the GPP equation, can solve the core–cusp problem. It is surrounded by an isothermal envelope that accounts for the flat rotation curves of the galaxies. By using the observational fact that the surface density of DM halos is universal (Kormendy et al., 2004; Spano et al., 2008; Donato et al., 2009) we obtain the halo mass–radius relation. Then, by using a “velocity dispersion tracing” relation justified by thermodynamical arguments based on a maximum entropy principle (most probable state), we obtain a general core mass–halo mass relation valid for bosons with repulsive or attractive self-interactions (Chavanis, 2019d; Chavanis, 2019a; Chavanis, 2020a; Chavanis, 2021a; Chavanis, 2023c; Padilla et al., 2021). Previously known results (Schive et al., 2014b) are recovered in particular limits of this relation. In the case of a repulsive self-interaction, we show by a general argument that the quantum core of realistic BECDM halos can never overcome the maximum mass set by general relativity and collapse toward a black hole by this process. In the case of an attractive self-interaction, we determine the halo mass above which the soliton (“axion” star) overcomes the maximum mass of Chavanis (2011c) and collapses. Section 7 discusses the secular evolution of BECDM halos due to the granular structure of the envelope (quasiparticles) (Schive et al., 2014a; Schive et al., 2014b; Hui et al., 2017) and the resulting increase in mass of the soliton by Bose–Einstein gravitational condensation (Levkov et al., 2018; Eggemeier and Niemeyer, 2019; Chen et al., 2021; Chen et al., 2022). We show that these results can be understood in terms of a kinetic theory based on the bosonic Landau equation (Bar-Or et al., 2019; Bar-Or et al., 2021; Marsh and Niemeyer, 2019; El-Zant et al., 2020; Chavanis, 2021b) and that BECDM halos behave in a sense like globular clusters in astrophysics. We thus suggest the possibility that they experience a gravothermal catastrophe stopped by quantum mechanics (Heisenberg’s uncertainty principle) or by repulsive scattering. This provides another means to form a solitonic core (Chavanis, 2019d). This quantum core becomes unstable if the halo mass is sufficiently large, leading to a supermassive black hole (SMBH). Section 8 discusses basic results of BECDM cosmology. We determine the Jeans length and the Jeans mass of a spatially homogeneous self-gravitating BEC (Khlopov et al., 1985; Bianchi et al., 1990; Chavanis, 2011c). Then, we take into account the expansion of the universe and discuss the different stages taking place during the cosmological evolution of a spatially homogeneous BECDM universe (Li et al., 2014; Suárez and Chavanis, 2017). Finally, we consider the growth of perturbations in an expanding universe filled with BECDM and again identify different regimes of evolution (Chavanis, 2012b; Suárez and Chavanis, 2015a).
2 From the KGE equations to the GPP equations
We assume that DM is made of spinless bosons in the form of BECs at
where
is the energy-momentum (stress) tensor of the SF. These equations have been introduced in the context of boson stars, first in the noninteracting case (Kaup, 1968; Ruffini and Bonazzola, 1969) and then for a self-interaction of the form Colpi et al. (1986):
where
This review especially focuses on the nonrelativistic limit of the KGE equations, which is suitable to describe BECDM in most cases of interest. By making the Klein transformation6
where
Then, by taking the nonrelativistic limit
These equations describe a nonrelativistic self-gravitating BEC at
For the standard BEC, we have
where
Remark: For a real SF, like the one describing axions, the counterpart of Equation 6 is
This transformation allows us to separate the fast oscillations of the SF with proper pulsation
where
where
3 Quantum hydrodynamics
In astrophysics, one is used to describe stars, DM, and even the universe as a whole in terms of fluid equations. At first sight, the BECDM model relies on a completely different formalism since we now have to work in terms of a complex wavefunction
where
The velocity field is irrotational
where
is the quantum potential taking into account the Heisenberg uncertainty principle and
is the pressure arising from the self-interaction of the bosons.8 For the standard BEC, using Equations 10, 21, the equation of state is
corresponding to a polytrope of index
Equations 16–19 have a clear physical interpretation. Equation 16 is the continuity equation, Equation 17 is the quantum Hamilton–Jacobi (or Bernoulli) equation, Equation 18 is the quantum Euler equation, and Equation 19 is the Poisson equation. We will call them the quantum Euler–Poisson equations. For
In the context of BECDM, the hydrodynamic representation of the GPP equations through the Madelung transformation was first considered by Böhmer and Harko (2007), Sikivie and Yang (2009), Chavanis (2011c), Chavanis and Delfini (2011), Chavanis (2011a), Chavanis (2012b), Rindler-Daller and Shapiro (2012). It is also possible to write the KGE equations in the form of quantum hydrodynamic equations by using the de Broglie transformation (de Broglie, 1927b; de Broglie, 1927c; de Broglie, 1927a). In the context of BECDM, the hydrodynamic representation of the KGE equations was first considered by Suárez and Matos (2011), Suárez and Matos (2014), Suárez and Chavanis (2015a), Suárez and Chavanis (2015b) in the weak gravity limit and by Chavanis and Harko (2012), Chavanis (2015c), Chavanis and Matos (2017), Matos et al. (2019), Chavanis (2022c), Chavanis (2023c) in a curved spacetime with an arbitrary metric (possibly including the electromagnetic field).9 In this review, we will essentially use the hydrodynamic representation of the GPP equations (see Chavanis (2011c), Chavanis (2017b) for the expression of the following results using the wave representation). We stress, however, that the fluid approach cannot resolve interference patterns at very small scales, in regions where the quantum pressure becomes ill-defined due to the vanishing of the wavefunction and the density. This is in general accompanied by the formation of singularities and topological defects like vortex lines.
Remark: We note that the quantum force in Equation 18 can be written as (Takabayasi, 1952)
where
is an anisotropic quantum pressure tensor. Therefore, the quantum Euler Equation 18 may be rewritten as
This shows that the quantum force is equivalent to an anisotropic pressure force. This remark will make more sense in Section 5.
4 Soliton
In this section, we consider the ground state of the GPP equations, which is usually referred to as a soliton. As explained in the introduction, this ground state solution can either describe the minimum halo of typical mass
4.1 Equilibrium state
In the hydrodynamic representation of the GPP equations, the equilibrium state of a BECDM halo is determined by the condition of quantum hydrostatic equilibrium
This equation describes the balance between the pressure due to the self-interaction of the bosons, the gravitational force, and the quantum force arising from the Heisenberg uncertainty principle. The quantum force always tends to stabilize a BECDM halo with respect to gravitational collapse. A repulsive self-interaction also stabilizes a BECDM halo similarly to the Pauli exclusion principle for fermionic DM. By contrast, an attractive self-interaction adds its effect to the gravitational attraction and destabilizes a BECDM halo above a maximum mass (see below). By combining Equation 26 with the Poisson Equation 19, we obtain the fundamental differential equation of quantum hydrostatic equilibrium (Chavanis, 2011c; Chavanis, 2017b)
For the standard BEC, using Equation 22, it becomes
The solution of this equation without node corresponds to the ground state of the GPP equations (soliton). This differential equation has been solved numerically in Chavanis and Delfini (2011) in the general case of a repulsive or an attractive self-interaction (or no self-interaction as in Membrado et al. (1989b)).10
For noninteracting self-gravitating BECs, a Gaussian density profile of the form (Chavanis, 2011c)
provides a relatively good fit of the soliton up to a few halo radii. A more accurate fit of the form
was later proposed by Schive et al. (2014a), Schive et al. (2014b). The radius

Figure 1. Normalized density profile of the soliton from Chavanis and Delfini (2011) (here
In the Thomas–Fermi (TF) limit where the quantum potential can be neglected (Tkachev, 1986; Membrado et al., 1989a; Lee and Koh, 1996; Goodman, 2000; Arbey et al., 2003; Böhmer and Harko, 2007; Chavanis, 2011c), the density profile has a compact support. It is analytically expressed as
This is the well-known density profile of a polytrope of index
Remark: In Appendix E of Chavanis (2019a), we have shown that, in the noninteracting case, the soliton resulting from the equilibrium between the gravitational attraction and the quantum repulsion (Heisenberg’s uncertainty principle) is similar to a polytrope of index
depending on the gravitational constant
4.2 Mass–radius relation
The mass–radius relation of BECDM halos at
Based on these works, the mass–radius relation of self-gravitating BECs can be parametrized by a function of the form
The value of the coefficients
In the noninteracting case, the mass–radius relation is given by (Membrado et al., 1989b; Chavanis and Delfini 2011)
where
In the TF limit where we can neglect the quantum potential, the equilibrium states have a unique radius (independent of the halo mass
where
The mass–radius relation with

Figure 2. Mass–radius relation of self-gravitating BECs with repulsive self-interaction
If we apply Equation 34 to the minimum halo of typical mass
If we apply Equation 35 to the minimum halo of typical mass
More generally, by fixing the values of

Figure 3. Mass
We can also use the Gaussian ansatz to obtain the pulsation of the soliton as a function of the self-interaction parameter
which is in fair agreement with the numerical value
4.3 Maximum mass of dilute axion stars
Self-gravitating BECs with an attractive self-interaction
The corresponding radius is
We note that

Table 1. Maximum mass
The mass–radius relation with

Figure 4. Mass–radius relation of self-gravitating BECs with attractive self-interaction
When
This relation shows that the equilibrium state is stable when

Figure 5. Squared pulsation (obtained from the Gaussian ansatz) as a function of the radius for self-gravitating BECs with
There is no equilibrium state with
where
We also note that dilute axion stars are metastable (local but not global minima of energy at fixed mass). Because of quantum fluctuations, they can penetrate the barrier of energy by tunnel effect and collapse. The calculation of their lifetime (Chavanis, 2020c), which is given by the WKB formula, is an interesting problem in physics. It can be performed by using the instanton theory, leading to an expression of the form (Chavanis, 2020c)
where the coefficients are determined by the Gaussian ansatz (see Appendix 2). We note that the detailed expression of the lifetime of dilute axion stars is unnecessary because it scales as
It is instructive to write the maximum mass and the minimum radius of dilute axion stars in different forms (Chavanis, 2011c; Chavanis, 2016a; Chavanis, 2018b)
where
For QCD axions with
For ULAs, we can have a much larger maximum mass, of the order of galactic masses (or, more precisely, of their DM quantum cores). Its precise value depends on the values of
In these two examples, the nonrelativistic approximation is justified because
Remark: For the minimum halo of typical mass
4.4 Outcome of the collapse
When
(i) The first possibility is to form a dense axion star (Braaten et al., 2016). When the star becomes overdense as a consequence of the collapse, one needs to take into account higher-order terms in the expansion of the self-interaction potential. The next-order term is a
(ii) The second possibility is a bosenova phenomenon (Levkov et al., 2017) if special relativity is taken into account in the wave equation. The collapse of the axion star may be accompanied by a burst of relativistic axions when the density reaches high values. In that case, the implosion of the core is followed by an explosion of the halo, like in supernovae (Chavanis et al., 2019). This may lead to an emission of radio-photons via parametric resonance (Hertzberg et al., 2020).
(iii) The third possibility, in the case where general relativity is taken into account, is the formation of a black hole if the mass of the axion star is sufficiently large or if the self-interaction is sufficiently weak
(iv) The axion star may also fragment in several stable pieces (axion drops) of mass
(v) Finally, it has been proposed that fast radio bursts, whose origin is one of the major mysteries of high energy astrophysics, could be caused by axion stars that have reached the maximum mass

Figure 6. Mass–radius relation of self-gravitating BECs with an attractive
4.5 Maximum mass of boson stars in general relativity
We have seen that the radius of Newtonian self-gravitating BECs with a repulsive self-interaction (or no self-interaction) decreases as their mass increases. General relativity must be taken into account when the radius of the object
The maximum mass and the minimum radius of a noninteracting boson star at
The minimum radius of the boson star is of the order of the Compton wavelength
The maximum mass and the minimum radius of a self-interacting boson star or BEC star at
The minimum radius of a boson star in the TF limit is of the same order as its radius in the nonrelativistic limit (see Equation 35). On the other hand, for
One can show [see Appendix C of Chavanis (2020a) and Section 6.4 below] that the mass of the quantum core of a DM halo is always much smaller than the maximum mass set by general relativity. Therefore, the soliton can be treated in the nonrelativistic limit and does not collapse toward a black hole by this process.
Remark: The maximum mass (Equation 47) of self-interacting boson stars can be obtained either by solving the KGE equations (Colpi et al., 1986) or from a fluid approach (Chavanis and Harko, 2012). Indeed, in the TF regime, one can show [see Suárez and Chavanis (2017), Chavanis (2022c), Chavanis (2023b) and Appendix B.8 of Chavanis (2023c) for details] that the SF behaves as a relativistic fluid with an equation of state (Colpi et al., 1986; Chavanis and Harko, 2012)
where

Figure 7. Mass–radius relation of general relativistic BEC stars with a repulsive
4.6 Interpolation formula for the maximum mass of boson stars with repulsive or attractive self-interactions
In the case of boson stars with a repulsive self-interaction
with corresponding radius
As expected, the maximum mass and the minimum radius increase as the self-interaction becomes more and more repulsive since the pressure due to the self-interaction opposes itself to the gravitational attraction and stabilizes the boson star.
In the case of dilute axion stars with an attractive self-interaction
with corresponding radius
As expected, the maximum mass decreases while the minimum radius increases as the self-interaction becomes more and more attractive since the negative pressure due to the self-interaction adds itself to the gravitational attraction and destabilizes the axion star.
We see that the transition between these two regimes (weak self-interaction and strong self-interaction) occurs for a scattering length of the order of
The noninteracting limit corresponds to
The noninteracting limit corresponds to
The results of this section together with those of Section 4.4 allow us to construct the phase diagram of boson stars with a

Figure 8. Phase diagram of boson stars with a
5 Core–envelope structure of BECDM halos
5.1 Violent relaxation and gravitational cooling
A self-gravitating BEC that is initially out-of-equilibrium undergoes a rapid relaxation toward a virialized state with a core–envelope structure. The quantum core (soliton) corresponds to the ground state of the KGE or GPP equations discussed in the previous section. It is surrounded by an extended envelope arising from the quantum interferences of excited states. This relaxation process was first identified by Seidel and Suen (1994), who numerically solved the KGE equations with spherical symmetry and showed the formation of a soliton (boson star) accompanied by an outward traveling scalar wave radiation. They called it “gravitational cooling.” They mentioned the similarity of this process with the “violent relaxation” of collisionless stellar systems discussed by Lynden-Bell (1967). In this sense, the core–envelope structure of the equilibrium (virialized) state can be interpreted as being the “most probable state” reached by the system as a result of a violent relaxation. It can be understood from a maximum entropy principle based on the Lynden–Bell entropy. The relaxation of BECDM halos governed by the SP equations was illustrated more recently by Schive et al. (2014a), Schive et al. (2014b) with use of spectacular colorful numerical simulations. They clearly showed the formation of a solitonic core surrounded by an NFW-like envelope. They also discovered that the envelope possesses a fluctuating granular structure.
5.2 Wigner–Kramers equation
In order to theoretically understand this core–envelope structure and make the link with previous works on the theory of violent relaxation, we have to introduce a phase-space description and use a set of equations similar to the Vlasov–Poisson equations governing the evolution of collisionless stellar systems. This can be accomplished in the framework of the Wigner–Poisson equations. It can be shown [see Appendix A of Chavanis (2022a)] that the Schrödinger equation is equivalent to the Wigner equation
where the Wigner distribution function is defined by
One can check that
Like in the case of collisionless stellar systems (Lynden-Bell, 1967), we need to introduce a coarse-grained description in order to describe the relaxation of the system toward a quasiequilibrium state. By analogy with our former paper (Chavanis et al., 1996) on the kinetic theory of the violent relaxation of stellar systems, we introduced in Chavanis (2022a) a coarse-grained Wigner equation of the form
The effective collision term on the right hand side of Equation 59 parameterizes the process of violent relaxation in terms of a Fokker–Planck (Kramers) operator. Therefore, Equation 59 can be called the Wigner–Kramers equation.24 The friction coefficient and the diffusion coefficient are linked by a form of Einstein relation
which corresponds to the dilute (nondegenerate) limit of the Lynden–Bell distribution (here
The Wigner–Kramers Equation 59 well accounts for the core–envelope structure of BECDM halos. The advection (Wigner) term on the left-hand side characterizes the quantum core (soliton) studied in Section 4, while the collision (Kramers) term on the right-hand side characterizes the envelope. If we take into account tidal effects from other galaxies and use the fact that quantum effects are negligible in the envelope in an average sense (i.e., apart from the presence of granules considered in Section 7), we can show that the distribution function governed by the Wigner–Kramers Equation 59 relaxes, in the envelope, toward the King distribution
which is a truncated Boltzmann distribution (King, 1965). A more elaborate model would be the fermionic King model (Chavanis, 1998) taking into account the effects of degeneracy in the sense of Lynden–Bell. These degeneracy effects arising from the coarse-grained Vlasov and Wigner equations are similar to those arising from the Pauli exclusion principle in quantum mechanics, but we will not consider them here (see Chavanis (2022a)).
Remark: The connection between the wave description (in terms of
5.3 Hydrodynamic equations
We can derive a hierarchy of hydrodynamic equations by taking the successive moments of the Wigner–Kramers Equation 59 (Chavanis et al., 1996; Mocz et al., 2018; Chavanis, 2022a), also called the quantum damped Jeans equations. With introduction of the density
where
This hierarchy of equations can be closed by making a local thermodynamic equilibrium (LTE) approximation. In Mocz et al. (2018), Chavanis (2022a), we argued heuristically that the pressure tensor
Combining the previous results, and adding the pressure due to a possible self-interaction between the bosons, we deduce that the quantum hydrodynamic equations parameterizing the complex dynamics of BECDM halos in our model are
In the quantum damped Euler Equation 66, the friction term forces the system to relax toward an equilibrium state in which the pressure balances the gravitational attraction.25 The resulting DM halo has a core–envelope structure. The quantum pressure and the self-interaction pressure (in the repulsive case) stabilize the system against gravitational collapse and lead to a central density core (soliton) instead of a cusp. Quantum effects can thus solve the core–cusp problem of CDM. On the other hand, quantum effects are negligible as a first approximation in the envelope. An isothermal envelope leads to flat rotation curves like in the CDM model.26
Some comments are as follows:
(i) We note that the anisotropic quantum tensor from Equation 24, which is equivalent to the quantum potential from Equation 20, can be interpreted as a stress tensor of the form
(ii) To obtain the simple isothermal equation of state (Equation 64), we have assumed that the distribution function in the envelope can be represented by a Boltzmann-like distribution (see Equation 60). We recall, however, that an isothermal distribution cannot be valid at infinitely large distances since it yields DM halos of infinite mass (Chandrasekhar, 1957; Binney and Tremaine, 1987). In principle, we should use the King distribution function (see Equations 61, 62), or the fermionic King distribution function (Chavanis, 1998), for which the density drops to 0 at a finite radius. This would give a more complicated (nonlinear) equation of state, generalizing Equation 64, which would account for the fact that the envelope is not exactly isothermal. Actually, depending on the degree of central concentration, the King distribution evolves from a polytrope of index
(iii) We have closed the hierarchy of hydrodynamic equations at the level of the momentum equation (second moment) by assuming a uniform temperature

Figure 9. Normalized density profiles of the classical King model in logarithmic scales for different values of the central concentration (adapted from Chavanis et al., 2015a). Dotted line: modified Hubble profile. Dashed line: Burkert profile. Dashed–dotted line: Hénon’s isochrone profile. The critical King profile is relatively close to the Burkert profile up to
Remark: In the strong friction limit
The friction term in the damped quantum Euler equation, therefore, allows us to establish a link between quantum mechanics and Brownian theory. However, the strong friction limit is usually not relevant in astrophysics. Rather, the low friction limit is found relevant. Therefore, in the present context, the relation to Brownian theory is just formal (see Chavanis (2023a) for more details about the analogy between generalized equations in quantum mechanics and Brownian theory and for a short history of these topics).
5.4 Generalized wave equation
Based on the aforementioned results derived from the coarse-grained Wigner (or Wigner–Kramers) equation, we have introduced a generalized wave equation (Chavanis, 2017b; Chavanis, 2019d; Chavanis, 2022a)
which is equivalent through the Madelung transformation to the damped Euler Equations 65, 66. This wave equation can be viewed as a generalized GP equation that includes a thermal term and a friction term resulting from the processes of gravitational cooling and violent relaxation. The thermal term, which is associated to the isothermal equation of state (Equation 64), is equivalent to a logarithmic potential
The friction term also leads to a logarithmic nonlinearity but of a different nature.
Interestingly, we can derive this generalized wave equation (in the noninteracting case) from a completely different formalism based on Nottale’s theory of scale relativity (Nottale, 2011; Chavanis, 2017a; Chavanis, 2018a). We know that the Schrödinger equation can be derived from Newton’s equation by using a principle of scale covariance (Nottale, 2011). Similarly, the generalized wave Equation 68 can be derived from Newton’s equation with a linear friction by writing it under the form (Chavanis, 2017a; Chavanis, 2018a)
where
between the cosmological constant (vacuum energy) and the mass of the electron
6 Predictive model of BECDM halos
In this section, following Chavanis (2019d), we develop a predictive model of BECDM halos with a core–envelope structure.
6.1 Equilibrium state
An equilibrium state of the generalized GPP Equations 9, 68 (equivalent to Equations 19, 65, 66) is determined, in the hydrodynamic representation, by the condition of quantum hydrostatic equilibrium from Equation 26, where the pressure
It has a linear part
It can be expressed under the form of a generalized Lane–Emden equation (Chavanis, 2019d; Chavanis, 2022a). The quadratic equation of state and the quantum potential dominate in the core where the density is high, and the isothermal (linear) equation of state dominates in the envelope where the density is low. As a result, the BECDM halo presents a quantum core (soliton) surrounded by an isothermal envelope. This composite model of BECDM halos with a core–envelope structure has been studied in detail in Chavanis (2019d).29
In the quantum core, we can ignore the thermal pressure, and we recover the equations of Section 4 describing the soliton. The equilibrium of the core is due to the balance between the self-interaction pressure, the quantum pressure, and the gravitational attraction. The quantum core (soliton), which is the ground state of the GPP equations, may solve the core–cusp problem of CDM. The core mass–radius relation is given by Equation 33.
In the envelope, we can ignore quantum effects as well as the self-interaction of the bosons, and Equation 73 reduces to
This equation is equivalent to the Boltzmann–Poisson (or Emden) equation, which describes a classical self-gravitating isothermal gas (Chandrasekhar, 1957). The equilibrium of the envelope is due to the balance between the effective thermal pressure and the gravitational attraction. The Boltzmann–Poisson (or Emden) equation has no simple analytical solution and must be solved numerically. However, its asymptotic behavior is known analytically (Chandrasekhar, 1957). The density of a self-gravitating isothermal halo decreases as
The equilibrium state of a BECDM halo can also be obtained by maximizing the Lynden–Bell entropy at fixed mass and energy (Chavanis, 2022a).30 This variational principle determines the most probable state of the system. It directly leads to Equation 73 and provides, in addition, a condition of thermodynamical stability: Only BECDM halos that are entropy maxima at fixed mass and energy are thermodynamically stable in the sense of Lynden–Bell. Actually, the statistical mechanics of self-gravitating systems is complicated because of incomplete relaxation, evaporation, and the gravothermal catastrophe (see Chavanis (2019d), Chavanis (2022a) and the reviews Padmanabhan (1990), Katz (2003), Chavanis (2006) for more details of these issues).31
The differential Equation 73 has been solved numerically in Chavanis (2019d); Chavanis (2018a). A density profile with a core–envelope structure is represented in Figure 10.

Figure 10. Density profile of a DM halo of mass
6.2 Halo mass–radius relation
To determine the halo mass–radius relation, we use the observational evidence that the surface density of DM halos has the universal value (Kormendy et al., 2004; Spano et al., 2008; Donato et al., 2009)
where
In the case of BECDM halos, we must be careful to correctly define what we call the “central” density. If we were to apply the relation from Equation 75 to the quantum core (soliton) of BECDM halos with
Assuming that the envelope is isothermal, we obtain the halo mass–radius relation (Chavanis, 2019d; Chavanis, 2022g)
We also find that
Remark: In the isothermal envelope model, the temperature must change from halo to halo according to the law
where
where
6.3 Core mass–halo mass relation
An important challenge of the BECDM model is to predict the relation between the core mass
Combining the velocity dispersion tracing relation from Equation 79 with the core mass–radius relation
Writing
This equation determines the mass
For noninteracting bosons, the core mass–halo mass relation is
For
For bosons with a repulsive self-interaction in the TF limit, the core mass–halo mass relation is
Note that the core radius
In these two examples, the quantum core represents a bulge or a nucleus. It cannot mimic an SMBH because it is too much extended
The general core mass–halo mass relation from Equation 80 is plotted in Figure 11 adapted from Chavanis (2019a) for an arbitrary (repulsive or attractive) self-interaction.

Figure 11. Core mass
Remark: Our results (Chavanis, 2019d; Chavanis, 2019a; Chavanis, 2020a) have been rederived and confirmed by Padilla et al. (2021) repeating the same arguments (compare Figure 2 in Padilla et al. (2021) to Fig. 19 in Chavanis (2019a)). Their scalings appear different from ours at first sight because their relations are expressed in terms of the virial mass instead of the halo mass (see Section 6.6), but the results turn out to be fully equivalent.
6.4 Collapse toward a black hole?
In the case of BECDM halos made of bosons with a repulsive self-interaction (or no self-interaction), the mass of the quantum core (soliton) increases monotonically with the halo mass (see Figure 11). Therefore, for sufficiently large halos, the core mass
According to Equations 45, 47, 82, 83, the mass of the soliton would be equal to the general relativistic maximum mass
Strikingly, this expression is independent of the characteristics of the DM particle (Chavanis, 2020a). Such a large halo mass is clearly unrealistic (the biggest DM halos observed in the universe have a mass
Remark: When these arguments are applied to the context of inflaton clusters that could form in the very early universe (Musoke et al., 2020; Niemeyer and Easther, 2020; Eggemeier et al., 2021; Eggemeier et al., 2022) it is possible that the solitonic cores (inflaton stars) of sufficiently massive inflaton clusters may overcome the critical mass
6.5 Collapse toward a dense axion star?
In the case of BECDM halos made of bosons with an attractive self-interaction (axions), the core mass–halo mass relation (see Figure 11) presents a maximum when the quantum core mass (dilute axion star) reaches the critical value from Equation 37, at which it becomes unstable and collapses (Chavanis, 2011c; Chavanis, 2016a).37 According to Equations 37, 80, the collapse of the core, leading to the formation of a dense axion star, a bosenova, or a black hole (Braaten et al., 2016; Davidson and Schwetz, 2016; Cotner, 2016; Chavanis, 2016a; Eby et al., 2016; Levkov et al., 2017; Helfer et al., 2017; Chavanis, 2018b; Visinelli et al., 2018; Michel and Moss, 2018) occurs in a DM halo of mass
We note that the maximum halo mass
For this phase transition (collapse) to take place in practice, the critical halo mass
Remark: For
6.6 Virial mass
In the previous sections, we have defined the halo mass
Numerically,
6.7 Fermionic DM halos
Although this review is devoted to the case of bosonic DM, we briefly mention in this section what the previous results become in the case of fermionic DM.
For degenerate fermions, the pressure is due to the Pauli exclusion principle. The equation of state of a nonrelativistic Fermi gas at
This is a polytropic equation of state of index
which is equivalent to the Lane–Emden equation of index
The radius decreases monotonically as the mass increases. An equilibrium state exists for any mass
The maximum mass and the minimum radius of a fermion star at
These results were first obtained in the case of neutron stars (Oppenheimer and Volkoff, 1939). The scalings of Equations 90, 91 can be obtained qualitatively by equating the mass–radius relation from Equation 89 with the Schwarzschild relation
The core mass–halo mass relation is (Chavanis, 2021a)
It terms of the virial mass we have
Like in Section 6.4, we can show that the quantum core (fermion ball) present at the center of a realistic fermionic DM halo is always below the general relativistic maximum mass from Equation 90, so it can never collapse toward a black hole by this process [see Appendix C of Chavanis (2020a)]. Since
We have proposed in Chavanis (2022a); Chavanis (2022g) a generalized wave equation of the form of Equation 68 for fermionic DM with an effective potential
accounting for the Pauli exclusion principle (see Equations 21, 87).
6.8 Can the quantum core mimic an SMBH?
A very massive object (Sagittarius A*) resides at the center of our galaxy of mass
(i) Let us first assume that this object is a noninteracting boson star (soliton).
Using the expression of the maximum mass from Equation 45 with
Using the nonrelativistic mass–radius relation from Equation 34 with
Using the core mass–halo mass relation from Equation 82 with
(ii) Let us now assume that the supermassive object at the center of the galaxy is a self-interacting boson star, which can be described with the TF approximation.
Using the expression of the maximum mass from Equation 47 with
Using the nonrelativistic relation from Equation 35 with
Using the core mass–halo mass relation from Equation 83 with
(iii) Finally, let us assume that the supermassive object at the center of the galaxy is a fermion ball.
Using the expression of the maximum mass from Equation 90 with
Using the nonrelativistic mass–radius relation from Equation 89 with
Using the core mass–halo mass relation from Equation 92 with
In conclusion, if the core mass–halo mass relations from Equations 82, 83, 92 are valid in the present context, it is not possible to construct a quantum core which mimics an SMBH at the center of our galaxy (see Figure 12). It would be too extended

Figure 12. Core–envelope density profile of a BECDM halo with a repulsive self-interaction (TF limit) in the scenario where the quantum core mimics an SMBH of mass
7 Secular evolution of BECDM halos
7.1 Classical relaxation time
The results of the previous sections have been obtained in the “collisionless” regime where the evolution of the system is due to a process of gravitational cooling and violent relaxation. This process takes place on a few dynamical times
where
7.2 Granules and quasiparticles
As shown by Schive et al. (2014a), Schive et al. (2014b), quantum wave interferences produce time-dependent small-scale density granules of the size
which depends on the local halo density
7.3 Heating and cooling
The evolution of test particles of mass
7.4 Kinetic theory based on the bosonic Landau equation
The granules (quasiparticles) also induce a secular evolution of the BECDM halo itself. This evolution can be described by the bosonic Landau equation Chavanis (2021b):
where
It corresponds to the relaxation time of a classical collisional stellar system (e.g., a globular cluster) from Equation 94 with an effective particle mass given by Equation 95. As discussed above, everything happens as if the envelope of the BECDM halo contained
7.5 Growth of the soliton
This kinetic description was originally introduced by Levkov et al. (2018) who used it to investigate the condensation phenomenon and the growth of the soliton due to collisions between quasiparticles. The relaxation time
We stress that this mechanism of condensation, which operates on a secular timescale, is physically different from the processes of gravitational cooling and collisionless violent relaxation discussed in Section 5.1 which operate on a much shorter timescale (a few dynamical times). In the present case, the formation and growth of the soliton is due to the condensation caused by gravitational interactions (collisions). There are therefore two distinct mechanisms (collisionless vs. collisional) to form the soliton (see the introduction of Chavanis (2021b) for more details). The distinction between the phase of violent relaxation (formation of the soliton) and the phase of slow collisional relaxation (formation and/or growth of the soliton) is illustrated numerically in Seidel and Suen (1994), Schive et al. (2014a), Schive et al. (2014b), Schwabe et al. (2016), Mocz et al. (2017), Mocz et al. (2018), Veltmaat et al. (2018), Mocz et al. (2019), Mocz et al. (2020), Veltmaat et al. (2020), Glennon and Prescod-Weinstein (2021), Mocz et al. (2023), Liu et al. (2023), Nori et al. (2023), Painter et al. (2024) and in Levkov et al. (2018), Eggemeier and Niemeyer (2019), Chen et al. (2021), Chen et al. (2022), respectively.
7.6 Self-interacting dark matter
For completeness, we mention another possible scenario of evolution of BECDM halos (Chavanis, 2019d). If the bosons have a strongly repulsive self-interaction, the collisional evolution of the envelope may be driven by self-interactions instead of gravitational encounters.39 This is usually referred to as self-interacting dark matter (SIDM) (Spergel and Steinhardt, 2000). The evolution of SIDM halos can be described by the ordinary Boltzmann kinetic equation. If the particles are self-interacting, with a scattering length
Since the Bullet Cluster constraint determines
Remark: If we account for Bose enhancement, the relaxation time scales generically as
7.7 Gravothermal catastrophe
Gravitational encounters between quasiparticles or collisions between bosons due to their self-interaction induce the secular evolution of a BECDM halo on a relevant timescale and establish an isothermal distribution, especially in the core of the halo where the density is high and the relaxation time is short. If we account for evaporation and tidal effects, this collisional evolution can provide another justification of the King distribution (see Equations 61, 62 without the bar on

Figure 13. Series of equilibria (caloric curve) of the classical King model (Chavanis et al., 2015a; Chavanis et al., 2015b) giving the inverse temperature
Most DM halos are close to the point of marginal stability. The marginal King density profile, which is flat at the center and decreases approximately as
If the BECDM halo is not too massive, the gravothermal catastrophe is stopped by quantum mechanics (Heisenberg’s uncertainty principle) or by the repulsive self-interaction of the bosons. This leads to BECDM halos with a core–envelope structure made of a bosonic condensate (soliton) surrounded by an approximately isothermal envelope or, more realistically, a King envelope (Chavanis, 2019d). Therefore, the soliton may result from the gravothermal catastrophe stopped by quantum mechanics or by repulsive scattering. This scenario provides another mechanism for the formation of a solitonic core (Chavanis, 2019d). Depending on the characteristics of the DM particle (mass, scattering length…), the condensed object (soliton) may mimic an SMBH at the center of a galaxy or represent a large DM bulge (see Section 6.8).41
If the BECDM halo is sufficiently massive, the gravothermal catastrophe may be followed by an instability of general relativistic origin (Balberg et al., 2002), leading to the formation of an SMBH instead of a quantum core (soliton). This soliton/SMBH transition (Chavanis, 2019d) occurs when the mass of the solitonic core overcomes the maximum mass
Note that during the gravothermal catastrophe and the gravitational collapse, the envelope remains unaltered. Therefore, the resulting structure is a marginal King profile (with a flat core and a
Similar results are obtained for fermionic DM where the soliton is replaced by a fermion ball (Bilic et al., 2003; Destri et al., 2013; Chavanis, 2006; Chavanis et al., 2015a; Chavanis et al., 2015b; Ruffini et al., 2015; Chavanis and Alberti, 2020; Alberti and Chavanis, 2020b; Chavanis, 2020d; Chavanis, 2020e; Argüelles et al., 2021; Chavanis, 2023d; Chavanis, 2022g; Argüelles et al., 2023).42 We refer to Chavanis (2019d), Chavanis (2022g), Chavanis (2023d), Argüelles et al. (2023) and references therein for more details about this scenario valid for fermionic and bosonic DM. At this stage, it is not possible to favor fermionic DM over bosonic DM or the converse. Note that fermionic and bosonic DM halos involve particles with a very different mass differing by 30 orders of magnitude (see Section 4, Section 6.7 and Chavanis, 2024b). It is also possible that DM is made of several types of fermions and bosons, as suggested in Sec. XII.E of Chavanis (2022g).
8 Basics of BECDM cosmology
The large-scale structures of the universe such as galaxies and DM halos are formed in the homogeneous background by Jeans instability (Jeans, 1902).43 For a cold classical gas, the Jeans length vanishes or is extremely small
The Jeans instability of a self-gravitating BEC (or a complex SF) with repulsive or attractive self-interactions was first considered by Khlopov et al. (1985) and Bianchi et al. (1990) in a general relativistic framework based on the KGE equations. The Jeans instability of a noninteracting self-gravitating BEC in Newtonian gravity described by the SP equations was studied by Hu et al. (2000) and Sikivie and Yang (2009). The Jeans instability of a Newtonian self-gravitating BEC with repulsive or attractive self-interactions described by the GPP equations was studied by Chavanis (2011c). These results were extended in general relativity by Suárez and Chavanis (2015a), and Suárez and Chavanis (2018). In these different studies, the authors determined the Jeans length and the Jeans mass of the unstable structures and used them to obtain an estimate of the minimum size and minimum mass of BECDM halos. These studies were performed in a static universe. The Jeans instability of an infinite homogeneous self-gravitating BEC in an expanding universe was studied by Bianchi et al. (1990), Suárez and Matos (2011) and Suárez and Chavanis (2015a) in general relativity and by Sikivie and Yang (2009) (without self-interaction) and Chavanis (2012b) (with self-interaction) in Newtonian gravity. These studies are valid for a complex SF describing the wavefunction of a BEC. They rely on a hydrodynamical representation of the wave equation. We refer to Chavanis (2020b) for a review about the Jeans instability of self-gravitating BECs. The Jeans instability of a real SF in general relativity has been studied by numerous authors, and a detailed list of references is given in Chavanis (2021a).
The Jeans instability is only valid in the linear regime of structure formation. It describes the initiation of the large-scale structures of the universe. The Jeans instability leads to the increase in the perturbations and the formation of condensations (clumps). When the density contrast reaches a sufficiently large value, the condensations (overdensities) decouple from the Hubble flow and experience a free fall, followed by a complicated process of gravitational cooling (Seidel and Suen, 1994) and violent relaxation (virialization) (Lynden-Bell, 1967). They can also grow through collisions, merging, and accretion before being diluted by the expansion of the universe. This corresponds to the nonlinear regime of structure formation, leading to the DM halos that we observe today. BECDM halos result from the balance between the gravitational attraction, the quantum pressure arising from the Heisenberg uncertainty principle, the pressure due to the self-interaction of the bosons, and the effective thermal pressure. They have a core–envelope structure with a quantum core (soliton) surrounded by an approximately isothermal envelope due to quantum interferences, as discussed in the previous sections.
8.1 Gravitational instability of an infinite homogeneous BEC in a static universe
Let us apply the GPP Equations 8, 9 to the universe as a whole in order to study the initiation of structure formation. We use a hydrodynamic approach. Specifically, following Chavanis (2011c), we study the linear dynamical stability of an infinite homogeneous self-gravitating BEC with density
Considering a small perturbation about an infinite homogeneous self-gravitating BEC, linearizing the quantum Euler–Poisson Equations 16–19, and decomposing the perturbation into plane waves, we obtain the generalized dispersion relation (Chavanis, 2011c)
where

Figure 14. Dispersion relation of a self-gravitating BEC with repulsive or attractive self-interaction. The pulsation is normalized by
The generalized Jeans wavenumber
A perturbation with wavelength
In the noninteracting limit
For a repulsive self-interaction
The maximum growth rate of the instability
From the Jeans length
Remark: The study of the Jeans instability for a self-gravitating BEC has been extended in general relativity in Suárez and Chavanis (2018) by using a hydrodynamic representation of the KGE equations in the weak gravity limit. The case of a dissipative self-gravitating BEC has been considered in Chavanis (2017b), Chavanis (2020b), and Ourabah (2020). The effect of a rotation and a magnetic field is discussed in Appendix 5.
8.2 Cosmological evolution of a BECDM universe
The cosmological evolution of a spatially homogeneous complex SF with a self-interaction potential
where
where

Figure 15. Phase diagram of a BEC universe showing the different eras of the SF during the evolution of the universe as a function of the scattering length of the bosons in the case of a repulsive self-interaction.
Remark: The study of the cosmological evolution of a relativistic SF with a
For
8.3 Growth of perturbations in an expanding universe filled with BECDM
The growth of perturbations in an expanding universe filled with BECDM has been studied in Chavanis (2012b). The expansion of the universe avoids the Jeans swindle (Peebles, 1980). In Chavanis (2012b), we focused on the matter era where the evolution of the universe is described by the Einstein-de Sitter (EdS) solution, and we used a nonrelativistic approach.48 We generalized the classical Bonnor (1957) study to a BEC universe by taking into account the quantum potential arising from the Heisenberg uncertainty principle and the pressure due to the self-interaction of the bosons.
The GPP equations in an expanding universe read (Chavanis, 2012b; Suárez and Chavanis, 2015a; Chavanis, 2020b; Chavanis, 2021a)
where
with the quantum potential
The evolution of the homogeneous background (
In the linear regime of structure formation, we showed that the evolution of the Fourier transform of the density contrast
This equation extends the classical Bonnor (1957) equation to a quantum (BEC) gas. It has been solved analytically in Chavanis (2012b) by considering just the effect of the quantum potential (noninteracting SF) or just the effect of a polytropic equation of state corresponding to a power-law self-interaction potential (TF limit).49 In the noninteracting regime, the solutions of Equation 115 representing the damped mode (upper sign) and the growing mode (lower sign) are
with
with

Figure 16. Evolution of the density contrast
Remark: The differential Equation 115 for the density contrast of a nonrelativistic BEC in an expanding universe originally derived in Chavanis (2012b) has been used in Marsh (2015), Marsh (2016), Fan (2016), Desjacques et al. (2018), Brax et al. (2019), Brax et al. (2020c), Mocz et al. (2023), Hwang and Noh (2022), Proukakis et al. (2024), Arakawa et al. (2024), and Guth et al. (2015). It has been extended to the context of general relativity in Suárez and Chavanis (2015a) by using the hydrodynamic representation of the KGE equations in the weak gravity regime.
9 Conclusion
In this paper, we have reviewed basic results on the BECDM model. We have discussed the core–envelope structure of BECDM halos and determined the core mass–radius relation, the halo mass–radius relation, and the core mass–halo mass relation. We have distinguished the violent (rapid) collisionless relaxation of BECDM halos through gravitational cooling and their secular (slow) collisional evolution induced by the formation of granules. We have explained how quantum mechanics may solve certain problems of the CDM model such as the core–cusp problem and the missing satellite problem. We have emphasized the importance of the maximum mass of dilute axion stars with attractive self-interactions (Chavanis, 2011c; Chavanis and Delfini, 2011) and its consequences.50 We have also discussed basic results of BECDM cosmology.
Many interesting studies have been developed in relation to the BECDM model. Although we cannot discuss them in detail, we briefly list some of them (we apologize for the many other valuable works that we cannot mention). They include, for example, the effect of temperature on the formation of compact objects in self-gravitating BECs (Slepian and Goodman, 2012; Robles and Matos, 2013; Verma et al., 2021; Proukakis et al., 2023; Shukla et al., 2024b), models of pulsar glitches interpreted as rotating self-gravitating BECs with a crust (Verma et al., 2022; Shukla et al., 2024a), the infall of an SF onto a central black hole (Brax et al., 2020a; Brax et al., 2020b), angular momentum and vortex formation in BECDM halos (Rindler-Daller and Shapiro, 2012; Harko, 2019), the formation of scalar clumps (solitons) in the early universe (Amin and Mocz, 2019; Brax et al., 2020c), self-similar solutions for FDM (Galazo-García et al., 2022), dynamical friction in FDM (Lora et al., 2012; Berezhiani et al., 2019; Lancaster et al., 2020; Buehler and Desjacques, 2023; Glennon et al., 2024; Bar-Or et al., 2019; Chavanis, 2021b; Hui et al., 2017; Boudon et al., 2022; Boudon et al., 2023), and gravitational waves from binary black holes in a self-interacting scalar dark matter cloud (Boudon et al., 2024). Some works more specifically related to axion stars and their maximum mass (Chavanis, 2011c) include, for example, the detectability of accretion-induced bosenovae in the Milky Way (Arakawa et al., 2024; Eby et al., 2022; Maseizik et al., 2024), the detectability of axion stars via photon emission (Bai and Hamada, 2018; Hertzberg et al., 2020; Patkós, 2023) or by gravitational microlensing events (Fujikura et al., 2021; Chang et al., 2024; Yin and Visinelli, 2024), the signature in the Lyman-
In this review, we have mainly focused on nonrelativistic self-gravitating BECs. There is of course a huge literature on boson stars in general relativity that we have just briefly mentioned. We refer to the introduction of our paper Chavanis (2023c) and to specific reviews on boson stars (Jetzer, 1992; Liddle and Madsen, 1992; Schunck and Mielke, 2003; Liebling and Palenzuela, 2017; Visinelli, 2021) for more details and references. It could also be interesting to further develop the analogies between BEC stars, superfluid stars, quark stars, and quark matter in neutron stars (Chavanis and Harko, 2012; Haghani and Harko, 2022; Sharma et al., 2022; Tangphati et al., 2024; Banerjee et al., 2025).
Author contributions
P-HC: writing – original draft and writing – review and editing.
Funding
The author(s) declare that no financial support was received for the research and/or publication of this article.
Conflict of interest
The author declares that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.
The handling editor TM declared a past co-authorship with the author.
Generative AI statement
The authors declare that no Generative AI was used in the creation of this manuscript.
Publisher’s note
All claims expressed in this article are solely those of the authors and do not necessarily represent those of their affiliated organizations, or those of the publisher, the editors and the reviewers. Any product that may be evaluated in this article, or claim that may be made by its manufacturer, is not guaranteed or endorsed by the publisher.
Footnotes
1The QCD axion is a pseudo-Nambu–Goldstone boson of the Peccei–Quinn phase transition associated with a
2The condensation occurs when the particles in the gas become correlated in a quantum mechanical manner, i.e., when the de Broglie thermal wavelength of a particle
3These perturbations may come from quantum fluctuations produced during the inflation. In this sense, the universe is never exactly homogeneous. It is spatially homogeneous only in an average sense. We speak of the homogeneous background.
4Spherical DM halos are connected by filaments. This network pattern with a filamentary structure is similar to the initiation of a vasculature in biology (see Gamba et al. (2003), Chavanis and Sire (2007), Chavanis and Sire (2008) for the development of the analogy between gravity and chemotaxis).
5Fermionic DM halos have a similar core–envelope structure, except that the soliton is replaced by a “fermion ball” similar to a white dwarf or a neutron star (Chavanis, 2022g). This core–envelope structure, involving a fermion ball surrounded by an isothermal envelope, directly results from the Fermi–Dirac distribution function coupled to gravity. Because of its isothermal distribution function, fermionic DM is an example of WDM.
6A short history of the early development of quantum mechanics (Schrödinger, KG and Dirac wave equations) is given in the introductions of Chavanis and Matos (2017), Chavanis, In preparation.
7Actually, the Madelung hydrodynamic equations are not fully equivalent to the Schrödinger equation (Bialynicki-Birula and Bialynicka-Birula, 1971; Wallstrom, 1994). To achieve perfect equivalence, we must assume that
8This relation, which can be rewritten as
9The Madelung and the de Broglie hydrodynamic equations (which aim at a causal interpretation of quantum mechanics) were much criticized by the founders of quantum mechanics, notably by Pauli, because their interpretation is not clear in the case of a single particle like an electron. However, they take more sense in the case of BECs with many bosons
10The density profile of the soliton can also be obtained by directly solving the GPP equations for the wavefunction (Guzmán and Ureña-López, 2006).
11In a recent paper, Indjin et al. (2024) have refined this variational approach by using an ansatz for the wavefunction that depends on the self-interaction parameter
12We stress the very small value of
13In a very interesting paper, Delgado and Muñoz Mateo (2023) determined the core mass and core radius of several DM-dominated halos by fitting their rotation curves with the velocity profile from Equation A8 and located them on the
14Dilute axion stars can be described by the nonrelativistic GPP equations with an attractive
15The noninteracting limit corresponds to
16A detailed exposition of the Poincaré theory of linear series of equilibria is given in Appendix C.2 of Alberti and Chavanis (2020b).
17The detailed expression of the lifetime of dilute axion stars could be useful in order to interpret future laboratory experiments of self-gravitating BECs with an effective gravity. In that case, the number
18Recall that the KGE and GPP equations have sometimes been used in the context of particle physics (independently from the context of boson stars) in order to describe classically extended particles consisting, e.g., of confined quarks with an effective gravity accounting for strong interactions (see the introduction of Chavanis (2023c)).
19The maximum mass can be smaller if
20An axion star may grow overcritical
21Visinelli et al. (2018) and Eby et al. (2019b), Eby et al. (2019a) argue that relativistic effects are crucial on the branch of dense axion stars, while self-gravity is negligible. As a result, dense axion stars correspond to “pseudobreathers” or “oscillons,” which are described by the sine-Gordon equation. For a real SF, these objects are known to be unstable and to decay via emission of relativistic axions on a timescale much shorter than any cosmological timescale. This conclusion is, however, contested by Braaten and Zhang (2019). On the other hand, dense axion stars may be stable in the relativistic regime if they are made of a complex SF. In that case, they are called “axion boson stars” (Guerra et al., 2019).
22These results were first obtained qualitatively in Tkachev (1986).
23It is also similar to the caloric curve of self-gravitating isothermal spheres (see Figure 3 in Chavanis (2023d)).
24We have neglected degenerate effects in the sense of Lynden–Bell. More general kinetic equations are the fermionic Wigner–Kramers equation and the fermionic Wigner–Landau equation (see Chavanis (2022a) for details).
25The Kramers equation and the damped Euler equations satisfy a form of
26The density profile of the self-gravitating isothermal sphere decreases at large distances as
27The recent results of Gaia DR3 (Jiao et al., 2023) indicate that the rotation curve of the Milky Way decreases more rapidly than previously thought. In the framework of the King model (Chavanis et al., 2015a), this suggests that the halo is more “polytropic” than “isothermal” (see Figs. 21 and 27 in Chavanis et al. (2015a)) in the sense that it is strongly concentrated (see Krut et al. (2025) for fermionic DM halos). It has been suggested (Chavanis et al., 2015a; Chavanis et al., 2015b; Simon-Petit et al., 2019) that the virialized state resulting from violent relaxation could correspond to Hénon’s isochrone profile with a
28We have written here the refined Eddington relation obtained in Chavanis (2019c), Chavanis (2022f), Chavanis (2024b), and Chavanis (2024c). This relation provides a remarkable agreement with the observations and solves the cosmological constant problem.
29The equation of state (Equation 72) and the corresponding generalized Lane–Emden equation introduced in Chavanis (2019d), Chavanis (2022a) have been used in Dawoodbhoy et al. (2021), Shapiro et al. (2022), Pils and Rindler-Daller (2022), and Foidl et al. (2023) to approximate their accurately simulated halo profiles (see also Korshynska et al. (2023)). Their studies confirm that an isothermal equation of state correctly describes the envelope of BECDM halos.
30We call it the Lynden–Bell entropy (for the coarse-grained distribution function
31A short history of the statistical mechanics of self-gravitating systems is given in the introductions of Alberti and Chavanis (2020a), Alberti and Chavanis (2020b), and Chavanis (2020e).
32Equation 75 with
33Similarly, for fermionic DM halos, using the mass–radius relation from Equation 89, we find
34This relation was empirically introduced in Mocz et al. (2017), Bar et al. (2018) without rigorous justification. In Chavanis (2019d), Chavanis (2019a), we provided a thermodynamical justification. We determined the actual core mass
35The numerical values in this expression rely on the Gaussian ansatz. As a result, the values of the parameters obtained below slightly differ from the “exact” values given in the previous sections.
36Of course, there is always the possibility that the quantum core grows by accretion and reaches the maximum mass (see the conclusion for references on this topic).
37Note that the maximum of the core mass–halo mass relation
38Conversely, the values of the DM particle mass computed in the present section lead to a minimum halo of mass
39These comments also apply to fermionic DM (Chavanis, 2022g) for which gravitational encounters are completely negligible (Bose enhancement which increases the effective mass of the particles is replaced by Pauli blocking, which reduces the effective mass of the particles and increases the gravitational relaxation time (Chavanis, 2021b)).
40The Burkert profile is similar to the NFW profile (Navarro et al., 1996) at large distances, but it presents a central core instead of a
41In the second case, we need to assume the presence of a primordial central SMBH to account for the observations.
42This scenario assumes that fermionic DM is self-interacting so as to justify its secular evolution. These ideas also apply to other models of SIDM and WDM. Indeed, fermionic DM, SIDM, and WDM halos present an isothermal core and may experience a gravothermal catastrophe. For not too massive DM halos, the collapse of the core is stopped by quantum mechanics (Pauli’s exclusion principle) or by the repulsive self-interaction of the particles. For massive DM halos, core collapse leads to the formation of an SMBH as proposed in Balberg et al. (2002) and further discussed in Chavanis et al. (2015a), Chavanis et al. (2015b), Chavanis (2020e), Chavanis and Alberti (2020), Alberti and Chavanis (2020b), Chavanis (2023d), and Chavanis (2022g) (see in particular Sec. XIII.D of Chavanis (2022g)).
43A short history of the Jeans instability is given in the introductions of Chavanis (2012a), Suárez and Chavanis (2018).
44WDM also implies a nonzero Jeans length due to the velocity dispersion of the particles.
45This is essentially a toy model, useful in a first approach, because it turns out that one can never neglect the expansion of the universe when studying the growth of structures in cosmology (see Suárez and Chavanis (2018) for details). Therefore, the study of Section 8.3 is more relevant.
46For a repulsive self-interaction
47A cosmology with a stiff matter era has been developed in Chavanis (2015a) at a general level (not necessarily connected to BECs). New analytical solutions of the Friedmann equations have been obtained for a Universe undergoing a stiff matter era, a DM era, and a DE era due to the cosmological constant.
48See the introductions of (Chavanis, 2013; Chavanis, 2014) for a short history of the early development of modern cosmology.
49These results can be applied to the standard BEC described by the quadratic equation of state from Equation 22 in the TF limit. They can also be applied to fermions described by the polytropic equation of state from Equation 87.
50In a sense, the maximum mass of dilute axion stars (Chavanis, 2023b) is the counterpart of the Chandrasekhar mass of white dwarfs (Chandrasekhar, 1931) or the Oppenheimer–Volkoff mass of neutron stars (Oppenheimer and Volkoff, 1939), but it has a nonrelativistic origin.
51For a Gaussian density profile, the relation between the radius
References
Ade, P. A. R., Aghanim, N., Armitage-Caplan, C., Arnaud, M., Ashdown, M., Atrio-Barandela, F., et al. (2014). Planck 2013 results. XVI. Cosmological parameters. Astron. Astrophys. 571, A16. doi:10.1051/0004-6361/201321591
Ade, P. A. R., Aghanim, N., Arnaud, M., Ashdown, M., Aumont, J., Baccigalupi, C., et al. (2016). Planck 2015 results. XIII. Cosmological parameters. Astron. Astrophys. 594, A13. doi:10.1051/0004-6361/201525830
Alberti, G., and Chavanis, P.-H. (2020a). Caloric curves of classical self-gravitating systems in general relativity. Phys. Rev. E 101, 052105. doi:10.1103/PhysRevE.101.052105
Alberti, G., and Chavanis, P.-H. (2020b). Caloric curves of self-gravitating fermions in general relativity. Eur. Phys. J. B 93, 208. doi:10.1140/epjb/e2020-100557-6
Alcubierre, M., Becerril, R., Guzmán, F. S., Matos, T., Núñez, D., and Ureña-López, L. A. (2003). Numerical studies of Φ2-oscillatons. Class. Quantum Gravity 20, 2883–2903. doi:10.1088/0264-9381/20/13/332
Álvarez-Rios, I., Bernal, T., Chavanis, P.-H., and Guzmán, F. S. (2024). Galactic rotation curves of low surface brightness galaxies using core-halo fuzzy dark matter configurations. Phys. Rev. D. 110, 063502. doi:10.1103/PhysRevD.110.063502
Amin, M. A., and Mocz, P. (2019). Formation, gravitational clustering, and interactions of nonrelativistic solitons in an expanding universe. Phys. Rev. D. 100, 063507. doi:10.1103/PhysRevD.100.063507
Arakawa, J., Eby, J., Safronova, M. S., Takhistov, V., and Zaheer, M. H. (2024). Detection of bosenovae with quantum sensors on Earth and in space. Phys. Rev. D. 110, 075007. doi:10.1103/PhysRevD.110.075007
Arbey, A., Lesgourgues, J., and Salati, P. (2001). Quintessential halos around galaxies. Phys. Rev. D. 64, 123528. doi:10.1103/PhysRevD.64.123528
Arbey, A., Lesgourgues, J., and Salati, P. (2003). Galactic halos of fluid dark matter. Phys. Rev. D. 68, 023511. doi:10.1103/PhysRevD.68.023511
Argüelles, C. R., Díaz, M. I., Krut, A., and Yunis, R. (2021). On the formation and stability of fermionic dark matter haloes in a cosmological framework. Mon. Not. R. Astron. Soc. 502, 4227–4246. doi:10.1093/mnras/staa3986
Argüelles, C. R., Becerra-Vergara, E. A., Rueda, J. A., and Ruffini, R. (2023). Fermionic dark matter: physics, astrophysics, and cosmology. Universe 9, 197. doi:10.3390/universe9040197
Bai, Y., and Hamada, Y. (2018). Detecting axion stars with radio telescopes. Phys. Lett. B 781, 187–194. doi:10.1016/j.physletb.2018.03.070
Balberg, S., Shapiro, S. L., and Inagaki, S. (2002). Self-interacting dark matter halos and the gravothermal catastrophe. Astrophys. J. 568, 475–487. doi:10.1086/339038
Baldeschi, M. R., Gelmini, G. B., and Ruffini, R. (1983). On massive fermions and bosons in galactic halos. Phys. Lett. B 122, 221–224. doi:10.1016/0370-2693(83)90688-3
Bañados, M., and Ferreira, P. G. (2010). Eddington’s theory of gravity and its progeny. Phys. Rev. Lett. 105, 011101. doi:10.1103/PhysRevLett.105.011101
Banerjee, A., Pradhan, A., Sakallı, İ., and Dixit, A. (2025). Properties of interacting quark star in light of Rastall gravity. Class. Quantum Gravity 42, 025008. doi:10.1088/1361-6382/ad9c0f
Bar, N., Blas, D., Blum, K., and Sibiryakov, S. (2018). Galactic rotation curves versus ultralight dark matter: implications of the soliton-host halo relation. Phys. Rev. D. 98, 083027. doi:10.1103/PhysRevD.98.083027
Bar-Or, B., Fouvry, J.-B., and Tremaine, S. (2019). Relaxation in a fuzzy dark matter halo. Astrophys. J. 871, 28. doi:10.3847/1538-4357/aaf28c
Bar-Or, B., Fouvry, J.-B., and Tremaine, S. (2021). Relaxation in a fuzzy dark matter halo. II. Self-Consistent kinetic equations. Astrophys. J. 915, 27. doi:10.3847/1538-4357/abfb66
Berezhiani, L., Elder, B., and Khoury, J. (2019). Dynamical friction in superfluids. J. Cosmol. Astropart. Phys. 2019, 074. doi:10.1088/1475-7516/2019/10/074
Berezhiani, L., Cintia, G., and Khoury, J. (2023). Thermalization, fragmentation, and tidal disruption: the complex galactic dynamics of dark matter superfluidity. Phys. Rev. D. 107, 123010. doi:10.1103/PhysRevD.107.123010
Bialynicki-Birula, I., and Bialynicka-Birula, Z. (1971). Magnetic monopoles in the hydrodynamic formulation of quantum mechanics. Phys. Rev. D. 3, 2410–2412. doi:10.1103/PhysRevD.3.2410
Bianchi, M., Grasso, D., and Ruffini, R. (1990). Jeans mass of a cosmological coherent scalar field. Astron. Astrophys. 231, 301–308.
Bilic, N., Tupper, G. B., and Viollier, R. D. (2003). “Dark matter in the galaxy,” in Particle physics in the new millennium. Editors J. Trampeti, and J. Wess, 616, 24–38. doi:10.1007/3-540-36539-7_2
Bode, P., Ostriker, J. P., and Turok, N. (2001). Halo Formation in warm dark matter models. Astrophys. J. 556, 93–107. doi:10.1086/321541
Böhmer, C. G., and Harko, T. (2007). Can dark matter be a Bose Einstein condensate? J. Cosmol. Astropart. Phys. 2007, 025. doi:10.1088/1475-7516/2007/06/025
Bonnor, W. B. (1957). Jeans’ formula for gravitational instability. Mon. Not. R. Astron. Soc. 117, 104–117. doi:10.1093/mnras/117.1.104
Boudon, A., Brax, P., and Valageas, P. (2022). Subsonic accretion and dynamical friction for a black hole moving through a self-interacting scalar dark matter cloud. Phys. Rev. D. 106, 043507. doi:10.1103/PhysRevD.106.043507
Boudon, A., Brax, P., and Valageas, P. (2023). Supersonic friction of a black hole traversing a self-interacting scalar dark matter cloud. Phys. Rev. D. 108, 103517. doi:10.1103/PhysRevD.108.103517
Boudon, A., Brax, P., Valageas, P., and Wong, L. K. (2024). Gravitational waves from binary black holes in a self-interacting scalar dark matter cloud. Phys. Rev. D. 109, 043504. doi:10.1103/PhysRevD.109.043504
Boyle, L. A., Caldwell, R. R., and Kamionkowski, M. (2002). Spintessence! New models for dark matter and dark energy. Phys. Lett. B 545, 17–22. doi:10.1016/S0370-2693(02)02590-X
Braaten, E., and Zhang, H. (2019). Colloquium: the physics of axion stars. Rev. Mod. Phys. 91, 041002. doi:10.1103/RevModPhys.91.041002
Braaten, E., Mohapatra, A., and Zhang, H. (2016). Dense axion stars. Phys. Rev. Lett. 117, 121801. doi:10.1103/PhysRevLett.117.121801
Brax, P., Valageas, P., and Cembranos, J. A. R. (2019). Impact of kinetic and potential self-interactions on scalar dark matter. Phys. Rev. D. 100, 023526. doi:10.1103/PhysRevD.100.023526
Brax, P., Valageas, P., and Cembranos, J. A. R. (2020a). Fate of scalar dark matter solitons around supermassive galactic black holes. Phys. Rev. D. 101, 023521. doi:10.1103/PhysRevD.101.023521
Brax, P., Valageas, P., and Cembranos, J. A. R. (2020b). K-essence scalar dark matter solitons around supermassive black holes. Phys. Rev. D. 101, 063510. doi:10.1103/PhysRevD.101.063510
Brax, P., Valageas, P., and Cembranos, J. A. R. (2020c). Nonrelativistic formation of scalar clumps as a candidate for dark matter. Phys. Rev. D. 102, 083012. doi:10.1103/PhysRevD.102.083012
Buehler, R., and Desjacques, V. (2023). Dynamical friction in fuzzy dark matter: circular orbits. Phys. Rev. D. 107, 023516. doi:10.1103/PhysRevD.107.023516
Bullock, J. S., and Boylan-Kolchin, M. (2017). Small-scale challenges to the ΛCDM paradigm. Annu. Rev. Astron. Astrophys. 55, 343–387. doi:10.1146/annurev-astro-091916-055313
Burkert, A. (1995). The structure of dark matter halos in dwarf galaxies. Astrophys. J. 447, L25–L28. doi:10.1086/309560
Burkert, A. (2020). Fuzzy dark matter and dark matter halo cores. Astrophys. J. 904, 161. doi:10.3847/1538-4357/abb242
Calogero, F. (1997). Cosmic origin of quantization. Phys. Lett. A 228, 335–346. doi:10.1016/S0375-9601(97)00107-2
Carvente, B., Jaramillo, V., Escamilla-Rivera, C., and Núñez, D. (2021). Observational constraints on complex quintessence with attractive self-interaction. Mon. Not. R. Astron. Soc. 503, 4008–4015. doi:10.1093/mnras/stab650
Chandrasekhar, S., and Fermi, E. (1953). Problems of gravitational stability in the presence of a magnetic field. Astrophys. J. 118, 116. doi:10.1086/145732
Chandrasekhar, S. (1931). The maximum mass of ideal white dwarfs. Astrophys. J. 74, 81. doi:10.1086/143324
Chandrasekhar, S. (1954). The gravitational instability of an infinite homogeneous medium when Coriolis force is acting and a magnetic field is present. Astrophys. J. 119, 7. doi:10.1086/145790
Chandrasekhar, S. (1955). The gravitational instability of an infinite homogeneous medium when a Coriolis acceleration is acting. Vistas Astronomy 1, 344–347. doi:10.1016/0083-6656(55)90045-3
Chang, J. H., Fox, P. J., and Xiao, H. (2024). Axion stars: mass functions and constraints. J. Cosmol. Astropart. Phys. 2024, 023. doi:10.1088/1475-7516/2024/08/023
Chavanis, P.-H., and Alberti, G. (2020). Gravitational phase transitions and instabilities of self-gravitating fermions in general relativity. Phys. Lett. B 801, 135155. doi:10.1016/j.physletb.2019.135155
Chavanis, P.-H., and Delfini, L. (2011). Mass-radius relation of Newtonian self-gravitating Bose-Einstein condensates with short-range interactions. II. Numerical results. Phys. Rev. D. 84, 043532. doi:10.1103/PhysRevD.84.043532
Chavanis, P.-H., and Harko, T. (2012). Bose-Einstein condensate general relativistic stars. Phys. Rev. D. 86, 064011. doi:10.1103/PhysRevD.86.064011
Chavanis, P.-H., and Kumar, S. (2017). Comparison between the Logotropic and ΛCDM models at the cosmological scale. J. Cosmol. Astropart. Phys. 2017, 018. doi:10.1088/1475-7516/2017/05/018
Chavanis, P.-H., and Matos, T. (2017). Covariant theory of Bose-Einstein condensates in curved spacetimes with electromagnetic interactions: the hydrodynamic approach. Eur. Phys. J. Plus 132, 30. doi:10.1140/epjp/i2017-11292-4
Chavanis, P.-H., and Sire, C. (2004). Anomalous diffusion and collapse of self-gravitating Langevin particles in D dimensions. Phys. Rev. E 69, 016116. doi:10.1103/PhysRevE.69.016116
Chavanis, P.-H., and Sire, C. (2007). Kinetic and hydrodynamic models of chemotactic aggregation. Phys. A Stat. Mech. its Appl. 384, 199–222. doi:10.1016/j.physa.2007.05.069
Chavanis, P.-H., and Sire, C. (2008). Jeans type analysis of chemotactic collapse. Phys. A Stat. Mech. its Appl. 387, 4033–4052. doi:10.1016/j.physa.2008.02.025
Chavanis, P. H., Sommeria, J., and Robert, R. (1996). Statistical mechanics of two-dimensional vortices and collisionless stellar systems. Astrophys. J. 471, 385–399. doi:10.1086/177977
Chavanis, P.-H., Lemou, M., and Méhats, F. (2015a). Models of dark matter halos based on statistical mechanics: the classical King model. Phys. Rev. D. 91, 063531. doi:10.1103/PhysRevD.91.063531
Chavanis, P.-H., Lemou, M., and Méhats, F. (2015b). Models of dark matter halos based on statistical mechanics: the fermionic King model. Phys. Rev. D. 92, 123527. doi:10.1103/PhysRevD.92.123527
Chavanis, P.-H., Denet, B., Le Berre, M., and Pomeau, Y. (2019). Supernova implosion-explosion in the light of catastrophe theory. Eur. Phys. J. B 92, 271. doi:10.1140/epjb/e2019-100435-6
Chavanis, P.-H. (1998). On the “coarse-grained” evolution of collisionless stellar systems. Mon. Not. R. Astron. Soc. 300, 981–991. doi:10.1046/j.1365-8711.1998.01867.x
Chavanis, P. H. (2002). Gravitational instability of finite isothermal spheres in general relativity. Analogy with neutron stars. Astron. Astrophys. 381, 709–730. doi:10.1051/0004-6361:20011424
Chavanis, P. H. (2005). On the lifetime of metastable states in self-gravitating systems. Astron. Astrophys. 432, 117–138. doi:10.1051/0004-6361:20041114
Chavanis, P. H. (2006). Phase transitions in self-gravitating systems. Int. J. Mod. Phys. B 20, 3113–3198. doi:10.1142/S0217979206035400
Chavanis, P.-H. (2007). White dwarf stars in D dimensions. Phys. Rev. D. 76, 023004. doi:10.1103/PhysRevD.76.023004
Chavanis, P. H. (2008). Relativistic stars with a linear equation of state: analogy with classical isothermal spheres and black holes. Astron. Astrophys. 483, 673–698. doi:10.1051/0004-6361:20078287
Chavanis, P.-H. (2011a). BEC dark matter, Zeldovich approximation, and generalized Burgers equation. Phys. Rev. D. 84, 063518. doi:10.1103/PhysRevD.84.063518
Chavanis, P.-H. (2011b). Instability of a uniformly collapsing cloud of classical and quantum self-gravitating Brownian particles. Phys. Rev. E 84, 031101. doi:10.1103/PhysRevE.84.031101
Chavanis, P.-H. (2011c). Mass-radius relation of Newtonian self-gravitating Bose-Einstein condensates with short-range interactions. I. Analytical results. Phys. Rev. D. 84, 043531. doi:10.1103/PhysRevD.84.043531
Chavanis, P. H. (2012a). Dynamical stability of infinite homogeneous self-gravitating systems and plasmas: application of the Nyquist method. Eur. Phys. J. B 85, 229. doi:10.1140/epjb/e2012-21012-9
Chavanis, P. H. (2012b). Growth of perturbations in an expanding universe with Bose-Einstein condensate dark matter. Astron. Astrophys. 537, A127. doi:10.1051/0004-6361/201116905
Chavanis, P.-H. (2013). “A simple model of universe describing the early inflation and the late accelerated expansion in a symmetric manner,” in IX Mexican school on gravitation and mathematical physics: cosmology for the XXIst century: gravitation and mathematical physics division of the Mexican physical society DGFM-SMF. Editors L. A. Ureña-López, R. Becerril-Bárcenas, and R. Linares-Romero, 75–115. doi:10.1063/1.4817032
Chavanis, P.-H. (2014). Models of universe with a polytropic equation of state: I. The early universe. Eur. Phys. J. Plus 129, 38. doi:10.1140/epjp/i2014-14038-x
Chavanis, P.-H. (2015a). Cosmology with a stiff matter era. Phys. Rev. D. 92, 103004. doi:10.1103/PhysRevD.92.103004
Chavanis, P.-H. (2015b). Is the universe logotropic? Eur. Phys. J. Plus 130, 130. doi:10.1140/epjp/i2015-15130-5
Chavanis, P.-H. (2015c). Partially relativistic self-gravitating Bose-Einstein condensates with a stiff equation of state. Eur. Phys. J. Plus 130, 181. doi:10.1140/epjp/i2015-15181-6
Chavanis, P.-H. (2015d). “Self-gravitating Bose-Einstein condensates,” in Quantum aspects of black holes. Editor X. Calmet, 151–194. doi:10.1007/978-3-319-10852-0_6
Chavanis, P.-H. (2016a). Collapse of a self-gravitating Bose-Einstein condensate with attractive self-interaction. Phys. Rev. D. 94, 083007. doi:10.1103/PhysRevD.94.083007
Chavanis, P.-H. (2016b). The Logotropic Dark Fluid as a unification of dark matter and dark energy. Phys. Lett. B 758, 59–66. doi:10.1016/j.physletb.2016.04.042
Chavanis, P.-H. (2017a). Derivation of a generalized Schrödinger equation from the theory of scale relativity. Eur. Phys. J. Plus 132, 286. doi:10.1140/epjp/i2017-11528-3
Chavanis, P.-H. (2017b). Dissipative self-gravitating Bose-Einstein condensates with arbitrary nonlinearity as a model of dark matter halos. Eur. Phys. J. Plus 132, 248. doi:10.1140/epjp/i2017-11544-3
Chavanis, P.-H. (2018a). Derivation of a generalized Schrödinger equation for dark matter halos from the theory of scale relativity. Phys. Dark Universe 22, 80–95. doi:10.1016/j.dark.2018.09.004
Chavanis, P.-H. (2018b). Phase transitions between dilute and dense axion stars. Phys. Rev. D. 98, 023009. doi:10.1103/PhysRevD.98.023009
Chavanis, P.-H. (2019a). Derivation of the core mass-halo mass relation of fermionic and bosonic dark matter halos from an effective thermodynamical model. Phys. Rev. D. 100, 123506. doi:10.1103/PhysRevD.100.123506
Chavanis, P.-H. (2019b). Mass-radius relation of self-gravitating Bose-Einstein condensates with a central black hole. Eur. Phys. J. Plus 134, 352. doi:10.1140/epjp/i2019-12734-7
Chavanis, P.-H. (2019c). New predictions from the logotropic model. Phys. Dark Universe 24, 100271. doi:10.1016/j.dark.2019.100271
Chavanis, P.-H. (2019d). Predictive model of BEC dark matter halos with a solitonic core and an isothermal atmosphere. Phys. Rev. D. 100, 083022. doi:10.1103/PhysRevD.100.083022
Chavanis, P.-H. (2020a). Core mass-halo mass relation of bosonic and fermionic dark matter halos harboring a supermassive black hole. Phys. Rev. D. 101, 063532. doi:10.1103/PhysRevD.101.063532
Chavanis, P.-H. (2020b). Jeans instability of dissipative self-gravitating Bose–Einstein condensates with repulsive or attractive
Chavanis, P.-H. (2020c). Quantum tunneling rate of dilute axion stars close to the maximum mass. Phys. Rev. D. 102, 083531. doi:10.1103/PhysRevD.102.083531
Chavanis, P.-H. (2020d). Statistical mechanics of self-gravitating systems in general relativity: I. The quantum Fermi gas. Eur. Phys. J. Plus 135, 290. doi:10.1140/epjp/s13360-020-00268-0
Chavanis, P.-H. (2020e). Statistical mechanics of self-gravitating systems in general relativity: II. The classical Boltzmann gas. Eur. Phys. J. Plus 135, 310. doi:10.1140/epjp/s13360-020-00291-1
Chavanis, P.-H. (2021a). Jeans mass-radius relation of self-gravitating Bose-Einstein condensates and typical parameters of the dark matter particle. Phys. Rev. D. 103, 123551. doi:10.1103/PhysRevD.103.123551
Chavanis, P.-H. (2021b). Landau equation for self-gravitating classical and quantum particles: application to dark matter. Eur. Phys. J. Plus 136, 703. doi:10.1140/epjp/s13360-021-01617-3
Chavanis, P.-H. (2022a). A heuristic wave equation parameterizing BEC dark matter halos with a quantum core and an isothermal atmosphere. Eur. Phys. J. B 95, 48. doi:10.1140/epjb/s10051-022-00299-9
Chavanis, P.-H. (2022b). Cosmological models based on a complex scalar field with a power-law potential associated with a polytropic equation of state. Phys. Rev. D. 106, 043502. doi:10.1103/PhysRevD.106.043502
Chavanis, P.-H. (2022c). K-essence Lagrangians of polytropic and logotropic unified dark matter and dark energy models. Astronomy 1, 126–221. doi:10.3390/astronomy1030011
Chavanis, P.-H. (2022d). Kinetic theory of collisionless relaxation for systems with long-range interactions. Phys. A Stat. Mech. Appl. 606, 128089. doi:10.1016/j.physa.2022.128089
Chavanis, P.-H. (2022e). New logotropic model based on a complex scalar field with a logarithmic potential. Phys. Rev. D. 106, 063525. doi:10.1103/PhysRevD.106.063525
Chavanis, P.-H. (2022f). Predictions from the logotropic model: the universal surface density of dark matter halos and the present proportions of dark matter and dark energy. Phys. Dark Universe 37, 101098. doi:10.1016/j.dark.2022.101098
Chavanis, P.-H. (2022g). Predictive model of fermionic dark matter halos with a quantum core and an isothermal atmosphere. Phys. Rev. D. 106, 043538. doi:10.1103/PhysRevD.106.043538
Chavanis, P.-H. (2023a). Generalized equations in quantum mechanics and Brownian theory. Symmetry 15, 2195. doi:10.3390/sym15122195
Chavanis, P.-H. (2023b). “The maximum mass of dilute axion stars,” in The sixteenth Marcel Grossmann meeting on recent developments in theoretical and experimental general relativity, astrophysics, and relativistic field theories. Editors R. Ruffini, and G. Vereshchagin, 2149–2173. doi:10.1142/9789811269776_0168
Chavanis, P.-H. (2023c). Maximum mass of relativistic self-gravitating Bose-Einstein condensates with repulsive or attractive |φ| 4 self-interaction. Phys. Rev. D. 107, 103503. doi:10.1103/PhysRevD.107.103503
Chavanis, P.-H. (2023d). “The self-gravitating Fermi gas in Newtonian gravity and general relativity,” in The sixteenth Marcel Grossmann meeting on recent developments in theoretical and experimental general relativity, astrophysics, and relativistic field theories. Editors R. Ruffini, and G. Vereshchagin, 2230–2251. doi:10.1142/9789811269776_0174
Chavanis, P.-H. (2024a). On the connection between Nelson’s stochastic quantum mechanics and Nottale’s theory of scale relativity. Axioms 13, 606. doi:10.3390/axioms13090606
Chavanis, P.-H. (2024b). A mass scale law connecting cosmophysics to microphysics. Phys. Dark Universe 44, 101420. doi:10.1016/j.dark.2024.101420
Chavanis, P.-H. (2024c). A simple model of magnetic universe without singularity associated with a quadratic equation of state. Eur. Phys. J. Plus 139, 889. doi:10.1140/epjp/s13360-024-05617-x
Chen, J., Du, X., Lentz, E. W., Marsh, D. J. E., and Niemeyer, J. C. (2021). New insights into the formation and growth of boson stars in dark matter halos. Phys. Rev. D. 104, 083022. doi:10.1103/PhysRevD.104.083022
Chen, J., Du, X., Lentz, E. W., and Marsh, D. J. E. (2022). Relaxation times for Bose-Einstein condensation by self-interaction and gravity. Phys. Rev. D. 106, 023009. doi:10.1103/PhysRevD.106.023009
Chiao, R. Y., Garmire, E., and Townes, C. H. (1964). Self-trapping of optical beams. Phys. Rev. Lett. 13, 479–482. doi:10.1103/PhysRevLett.13.479
Colpi, M., Shapiro, S. L., and Wasserman, I. (1986). Boson stars: gravitational equilibria of self-interacting scalar fields. Phys. Rev. Lett. 57, 2485–2488. doi:10.1103/PhysRevLett.57.2485
Cotner, E. (2016). Collisional interactions between self-interacting nonrelativistic boson stars: effective potential analysis and numerical simulations. Phys. Rev. D. 94, 063503. doi:10.1103/PhysRevD.94.063503
Davidson, S., and Schwetz, T. (2016). Rotating drops of axion dark matter. Phys. Rev. D. 93, 123509. doi:10.1103/PhysRevD.93.123509
Davies, E. Y., and Mocz, P. (2020). Fuzzy dark matter soliton cores around supermassive black holes. Mon. Not. R. Astron. Soc. 492, 5721–5729. doi:10.1093/mnras/staa202
Dawoodbhoy, T., Shapiro, P. R., and Rindler-Daller, T. (2021). Core-envelope haloes in scalar field dark matter with repulsive self-interaction: fluid dynamics beyond the de Broglie wavelength. Mon. Not. R. Astron. Soc. 506, 2418–2444. doi:10.1093/mnras/stab1859
de Broglie, L. (1927b). La mécanique ondulatoire et la structure atomique de la matière et du rayonnement. J. de Physique le Radium 8, 225–241. doi:10.1051/jphysrad:0192700805022500
de Broglie, L. (1927c). Sur le rôle des ondes continues Ψ en Mécanique ondulatoire. Compt. Rend. Acad. Sci. Paris 185, 380.
de la Peña, L., Cetto, A. M., and Valdés Hernández, A. (2015). The emerging quantum: the physics behind quantum mechanics. doi:10.1007/978-3-319-07893-9
Delgado, V., and Muñoz Mateo, A. (2023). Self-interacting superfluid dark matter droplets. Mon. Notices R. Astronomical Soc. 518, 4064–4072. doi:10.1093/mnras/stac3386
Deng, H., Hertzberg, M. P., Namjoo, M. H., and Masoumi, A. (2018). Can light dark matter solve the core-cusp problem? Phys. Rev. D. 98, 023513. doi:10.1103/PhysRevD.98.023513
Derrick, G. H. (1964). Comments on nonlinear wave equations as models for elementary particles. J. Math. Phys. 5, 1252–1254. doi:10.1063/1.1704233
Desjacques, V., Kehagias, A., and Riotto, A. (2018). Impact of ultralight axion self-interactions on the large scale structure of the Universe. Phys. Rev. D. 97, 023529. doi:10.1103/PhysRevD.97.023529
Destri, C., de Vega, H. J., and Sanchez, N. G. (2013). Fermionic warm dark matter produces galaxy cores in the observed scales because of quantum mechanics. New Astron. 22, 39–50. doi:10.1016/j.newast.2012.12.003
Di, H., and Shi, H. (2023). Can planet 9 be an axion star? Phys. Rev. D. 108, 103038. doi:10.1103/PhysRevD.108.103038
Di, H., Shao, L., Yi, Z., and Kong, S.-B. (2024). Novel standard candle: collapsing axion stars. Phys. Rev. D. 110, 103031. doi:10.1103/PhysRevD.110.103031
Di, H. (2024). Stimulated decay of collapsing axion stars and fast radio bursts. Eur. Phys. J. C 84, 283. doi:10.1140/epjc/s10052-024-12654-4
Dirac, P. A. M. (1931). Quantised singularities in the electromagnetic field. Proc. R. Soc. Lond. Ser. A 133, 60–72. doi:10.1098/rspa.1931.0130
Donato, F., Gentile, G., Salucci, P., Frigerio Martins, C., Wilkinson, M. I., Gilmore, G., et al. (2009). A constant dark matter halo surface density in galaxies. Mon. Not. R. Astron. Soc. 397, 1169–1176. doi:10.1111/j.1365-2966.2009.15004.x
Eby, J., Leembruggen, M., Suranyi, P., and Wijewardhana, L. C. R. (2016). Collapse of axion stars. J. High Energy Phys. 2016, 66. doi:10.1007/JHEP12(2016)066
Eby, J., Leembruggen, M., Leeney, J., Suranyi, P., and Wijewardhana, L. C. R. (2017). Collisions of dark matter axion stars with astrophysical sources. J. High Energy Phys. 2017, 99. doi:10.1007/JHEP04(2017)099
Eby, J., Leembruggen, M., Street, L., Suranyi, P., and Wijewardhana, L. C. R. (2018a). Approximation methods in the study of boson stars. Phys. Rev. D. 98, 123013. doi:10.1103/PhysRevD.98.123013
Eby, J., Leembruggen, M., Suranyi, P., and Wijewardhana, L. C. R. (2018b). Stability of condensed fuzzy dark matter halos. J. Cosmol. Astropart. Phys. 2018, 058. doi:10.1088/1475-7516/2018/10/058
Eby, J., Leembruggen, M., Street, L., Suranyi, P., and Wijewardhana, L. C. R. (2019a). Global view of QCD axion stars. Phys. Rev. D. 100, 063002. doi:10.1103/PhysRevD.100.063002
Eby, J., Mukaida, K., Takimoto, M., Wijewardhana, L. C. R., and Yamada, M. (2019b). Classical nonrelativistic effective field theory and the role of gravitational interactions. Phys. Rev. D. 99, 123503. doi:10.1103/PhysRevD.99.123503
Eby, J., Shirai, S., Stadnik, Y. V., and Takhistov, V. (2022). Probing relativistic axions from transient astrophysical sources. Phys. Lett. B 825, 136858. doi:10.1016/j.physletb.2021.136858
Eddington, A. (1931). On the value of the cosmical constant. Proc. R. Soc. Lond. Ser. A 133, 605–615. doi:10.1098/rspa.1931.0170
Eggemeier, B., and Niemeyer, J. C. (2019). Formation and mass growth of axion stars in axion miniclusters. Phys. Rev. D. 100, 063528. doi:10.1103/PhysRevD.100.063528
Eggemeier, B., Niemeyer, J. C., and Easther, R. (2021). Formation of inflaton halos after inflation. Phys. Rev. D. 103, 063525. doi:10.1103/PhysRevD.103.063525
Eggemeier, B., Schwabe, B., Niemeyer, J. C., and Easther, R. (2022). Gravitational collapse in the postinflationary Universe. Phys. Rev. D. 105, 023516. doi:10.1103/PhysRevD.105.023516
El-Zant, A. A., Freundlich, J., Combes, F., and Halle, A. (2020). The effect of fluctuating fuzzy axion haloes on stellar dynamics: a stochastic model. Mon. Not. R. Astron. Soc. 492, 877–894. doi:10.1093/mnras/stz3478
Fan, J. (2016). Ultralight repulsive dark matter and BEC. Phys. Dark Universe 14, 84–94. doi:10.1016/j.dark.2016.10.005
Ferreira, E. G. M. (2021). Ultra-light dark matter. Astron. Astrophys. Rev. 29, 7. doi:10.1007/s00159-021-00135-6
Feynman, R. P. (1958). Excitations in liquid helium. Physica 24, S18–S26. doi:10.1016/S0031-8914(58)80495-4
Foidl, H., Rindler-Daller, T., and Zeilinger, W. W. (2023). Halo formation and evolution in scalar field dark matter and cold dark matter: new insights from the fluid approach. Phys. Rev. D. 108, 043012. doi:10.1103/PhysRevD.108.043012
Friedberg, R., Lee, T. D., and Pang, Y. (1987). Scalar soliton stars and black holes. Phys. Rev. D. 35, 3658–3677. doi:10.1103/PhysRevD.35.3658
Fujikura, K., Hertzberg, M. P., Schiappacasse, E. D., and Yamaguchi, M. (2021). Microlensing constraints on axion stars including finite lens and source size effects. Phys. Rev. D. 104, 123012. doi:10.1103/PhysRevD.104.123012
Galazo García, R., Brax, P., and Valageas, P. (2024). Solitons and halos for self-interacting scalar dark matter. Phys. Rev. D. 109, 043516. doi:10.1103/PhysRevD.109.043516
Galazo-García, R., Brax, P., and Valageas, P. (2022). Self-similar solutions for fuzzy dark matter. Phys. Rev. D. 105, 123528. doi:10.1103/PhysRevD.105.123528
Gamba, A., Ambrosi, D., Coniglio, A., de Candia, A., di Talia, S., Giraudo, E., et al. (2003). Percolation, morphogenesis, and Burgers dynamics in blood vessels formation. Phys. Rev. Lett. 90, 118101. doi:10.1103/PhysRevLett.90.118101
Gan, X., Wang, L.-T., and Xiao, H. (2024). Detecting axion dark matter with black hole polarimetry. Phys. Rev. D. 110, 063039. doi:10.1103/PhysRevD.110.063039
Gillessen, S., Eisenhauer, F., Fritz, T. K., Bartko, H., Dodds-Eden, K., Pfuhl, O., et al. (2009). The orbit of the star S2 around SGR A* from very large telescope and Keck data. Astrophys. J. 707, L114–L117. doi:10.1088/0004-637X/707/2/L114
Glennon, N., and Prescod-Weinstein, C. (2021). Modifying PyUltraLight to model scalar dark matter with self-interactions. Phys. Rev. D. 104, 083532. doi:10.1103/PhysRevD.104.083532
Glennon, N., Musoke, N., Nadler, E. O., Prescod-Weinstein, C., and Wechsler, R. H. (2024). Dynamical friction in self-interacting ultralight dark matter. Phys. Rev. D. 109, 063501. doi:10.1103/PhysRevD.109.063501
Goodman, J. (2000). Repulsive dark matter. New Astron. 5, 103–107. doi:10.1016/S1384-1076(00)00015-4
Guerra, D., Macedo, C. F. B., and Pani, P. (2019). Axion boson stars. J. Cosmol. Astropart. Phys. 2019, 061. doi:10.1088/1475-7516/2019/09/061
Guth, A. H., Hertzberg, M. P., and Prescod-Weinstein, C. (2015). Do dark matter axions form a condensate with long-range correlation? Phys. Rev. D. 92, 103513. doi:10.1103/PhysRevD.92.103513
Guzmán, F. S., and Matos, T. (2000). Letter to the editor: scalar fields as dark matter in spiral galaxies. Class. Quantum Gravity 17, L9–L16. doi:10.1088/0264-9381/17/1/102
Guzmán, F. S., and Ureña-López, L. A. (2004). Evolution of the Schrödinger-Newton system for a self-gravitating scalar field. Phys. Rev. D. 69, 124033. doi:10.1103/PhysRevD.69.124033
Guzmán, F. S., and Ureña-López, L. A. (2006). Gravitational cooling of self-gravitating Bose condensates. Astrophys. J. 645, 814–819. doi:10.1086/504508
Haghani, Z., and Harko, T. (2022). Compact stars in the Einstein dark energy model. Phys. Rev. D. 105, 064059. doi:10.1103/PhysRevD.105.064059
Harko, T. (2019). Jeans instability and turbulent gravitational collapse of Bose–Einstein condensate dark matter halos. Eur. Phys. J. C 79, 787. doi:10.1140/epjc/s10052-019-7285-3
Harrison, B. K., Thorne, K. S., Wakano, M., and Wheeler, J. A. (1965). Gravitation theory and gravitational collapse.
Helfer, T., Marsh, D. J. E., Clough, K., Fairbairn, M., Lim, E. A., and Becerril, R. (2017). Black hole formation from axion stars. J. Cosmol. Astropart. Phys. 2017, 055. doi:10.1088/1475-7516/2017/03/055
Hertzberg, M. P., Li, Y., and Schiappacasse, E. D. (2020). Merger of dark matter axion clumps and resonant photon emission. J. Cosmol. Astropart. Phys. 2020, 067. doi:10.1088/1475-7516/2020/07/067
Hinshaw, G., Weiland, J. L., Hill, R. S., Odegard, N., Larson, D., Bennett, C. L., et al. (2009). Five-year Wilkinson microwave anisotropy probe observations: data processing, sky maps, and basic results. Astrophys. J. 180, 225–245. doi:10.1088/0067-0049/180/2/225
Hu, W., Barkana, R., and Gruzinov, A. (2000). Fuzzy cold dark matter: the wave properties of ultralight particles. Phys. Rev. Lett. 85, 1158–1161. doi:10.1103/PhysRevLett.85.1158
Hui, L., Ostriker, J. P., Tremaine, S., and Witten, E. (2017). Ultralight scalars as cosmological dark matter. Phys. Rev. D. 95, 043541. doi:10.1103/PhysRevD.95.043541
Hwang, J.-c., and Noh, H. (2022). Axion as a fuzzy dark matter candidate: proofs in different gauges. J. Cosmol. Astropart. Phys. 2022, 001. doi:10.1088/1475-7516/2022/03/001
Indjin, M., Liu, I. K., Proukakis, N. P., and Rigopoulos, G. (2024). Virialized profiles and oscillations of self-interacting fuzzy dark matter solitons. Phys. Rev. D. 109, 103518. doi:10.1103/PhysRevD.109.103518
Jeans, J. H. (1902). The stability of a spherical nebula. Philosophical Trans. R. Soc. Lond. Ser. A 199, 1–53. doi:10.1098/rsta.1902.0012
Ji, S. U., and Sin, S. J. (1994). Late-time phase transition and the galactic halo as a Bose liquid. II. The effect of visible matter. Phys. Rev. D. 50, 3655–3659. doi:10.1103/PhysRevD.50.3655
Jiao, Y., Hammer, F., Wang, H., Wang, J., Amram, P., Chemin, L., et al. (2023). Detection of the Keplerian decline in the Milky Way rotation curve. Astron. Astrophys. 678, A208. doi:10.1051/0004-6361/202347513
Kamionkowski, M., and Liddle, A. R. (2000). The dearth of halo dwarf galaxies: is there power on short scales? Phys. Rev. Lett. 84, 4525–4528. doi:10.1103/PhysRevLett.84.4525
Katz, J. (2003). Thermodynamics of self-gravitating systems. Found. Phys. 33, 223–269. doi:10.1023/A:1023776921610
Kauffmann, G., White, S. D. M., and Guiderdoni, B. (1993). The formation and evolution of galaxies within merging dark matter haloes. Mon. Not. R. Astron. Soc. 264, 201–218. doi:10.1093/mnras/264.1.201
Khlopov, M. I., Malomed, B. A., and Zeldovich, I. B. (1985). Gravitational instability of scalar fields and formation of primordial black holes. Mon. Not. R. Astron. Soc. 215, 575–589. doi:10.1093/mnras/215.4.575
Khoury, J. (2022). Dark matter superfluidity. SciPost Phys. Lect. Notes 42, 42. doi:10.21468/SciPostPhysLectNotes.42
King, I. R. (1965). The structure of star clusters. II. Steady-state velocity distributions. Astron. J. 70, 376. doi:10.1086/109750
Kirkpatrick, K., Mirasola, A. E., and Prescod-Weinstein, C. (2022). Analysis of Bose-Einstein condensation times for self-interacting scalar dark matter. Phys. Rev. D. 106, 043512. doi:10.1103/PhysRevD.106.043512
Klypin, A., Kravtsov, A. V., Valenzuela, O., and Prada, F. (1999). Where are the missing galactic satellites? Astrophys. J. 522, 82–92. doi:10.1086/307643
Kormendy, J., and Freeman, K. C. (2004). “Scaling laws for dark matter halos in late-type and dwarf spheroidal galaxies,” in Dark matter in galaxies. Editors S. Ryder, D. Pisano, M. Walker, and K. Freeman doi:10.48550/arXiv.astro-ph/0407321
Korshynska, K., Bidasyuk, Y. M., Gorbar, E. V., Jia, J., and Yakimenko, A. I. (2023). Dynamical galactic effects induced by solitonic vortex structure in bosonic dark matter. Eur. Phys. J. C 83, 451. doi:10.1140/epjc/s10052-023-11548-1
Krut, A., Argüelles, C., and Chavanis, P.-H. (2025). Thermodynamics of self-gravitating fermions as a robust theory for dark matter halos: stability analysis applied to the Milky Way. arXiv:2503.10870.
Lancaster, L., Giovanetti, C., Mocz, P., Kahn, Y., Lisanti, M., and Spergel, D. N. (2020). Dynamical friction in a fuzzy dark matter universe. J. Cosmol. Astropart. Phys. 2020, 001. doi:10.1088/1475-7516/2020/01/001
Lee, J.-W., and Koh, I.-G. (1996). Galactic halos as boson stars. Phys. Rev. D. 53, 2236–2239. doi:10.1103/PhysRevD.53.2236
Lee, T. D., and Pang, Y. (1987). Fermion soliton stars and black holes. Phys. Rev. D. 35, 3678–3694. doi:10.1103/PhysRevD.35.3678
Lee, B. K. K., Chu, M.-C., and Lin, L.-M. (2021). Could the GW190814 secondary component be a bosonic dark matter admixed compact star? Astrophys. J. 922, 242. doi:10.3847/1538-4357/ac2735
Lee, T. D. (1987b). Soliton stars and the critical masses of black holes. Phys. Rev. D. 35, 3637–3639. doi:10.1103/PhysRevD.35.3637
Lee, J.-W. (2018). Brief history of ultra-light scalar dark matter models. Eur. Phys. J. Web Conf. 168, 06005. doi:10.1051/epjconf/201816806005
Levkov, D. G., Panin, A. G., and Tkachev, I. I. (2017). Relativistic axions from collapsing Bose stars. Phys. Rev. Lett. 118, 011301. doi:10.1103/PhysRevLett.118.011301
Levkov, D. G., Panin, A. G., and Tkachev, I. I. (2018). Gravitational Bose-Einstein condensation in the kinetic regime. Phys. Rev. Lett. 121, 151301. doi:10.1103/PhysRevLett.121.151301
Li, B., Rindler-Daller, T., and Shapiro, P. R. (2014). Cosmological constraints on Bose-Einstein-condensed scalar field dark matter. Phys. Rev. D. 89, 083536. doi:10.1103/PhysRevD.89.083536
Liddle, A. R., and Madsen, M. S. (1992). The structure and formation of boson stars. Int. J. Mod. Phys. D 1, 101–143. doi:10.1142/S0218271892000057
Liebling, S. L., and Palenzuela, C. (2017). Dynamical boson stars. Living Rev. Relativ. 20, 5. doi:10.1007/s41114-017-0007-y
Lin, S.-C., Schive, H.-Y., Wong, S.-K., and Chiueh, T. (2018). Self-consistent construction of virialized wave dark matter halos. Phys. Rev. D. 97, 103523. doi:10.1103/PhysRevD.97.103523
Liu, I. K., Proukakis, N. P., and Rigopoulos, G. (2023). Coherent and incoherent structures in fuzzy dark matter haloes. Mon. Not. R. Astron. Soc. 521, 3625–3647. doi:10.1093/mnras/stad591
Lora, V., Magaña, J., Bernal, A., Sánchez-Salcedo, F. J., and Grebel, E. K. (2012). On the mass of ultra-light bosonic dark matter from galactic dynamics. J. Cosmol. Astropart. Phys. 2012, 011. doi:10.1088/1475-7516/2012/02/011
Lynden-Bell, D., and Wood, R. (1968). The gravo-thermal catastrophe in isothermal spheres and the onset of red-giant structure for stellar systems. Mon. Not. R. Astron. Soc. 138, 495–525. doi:10.1093/mnras/138.4.495
Lynden-Bell, D. (1967). Statistical mechanics of violent relaxation in stellar systems. Mon. Not. R. Astron. Soc. 136, 101–121. doi:10.1093/mnras/136.1.101
Madelung, E. (1927). Quantentheorie in hydrodynamischer form. Z. Fur Phys. 40, 322–326. doi:10.1007/BF01400372
Marsh, D. J. E., and Niemeyer, J. C. (2019). Strong constraints on fuzzy dark matter from ultrafaint dwarf galaxy Eridanus II. Phys. Rev. Lett. 123, 051103. doi:10.1103/PhysRevLett.123.051103
Marsh, D. J. E. (2015). Nonlinear hydrodynamics of axion dark matter: relative velocity effects and quantum forces. Phys. Rev. D. 91, 123520. doi:10.1103/PhysRevD.91.123520
Maseizik, D., Eby, J., Seong, H., and Sigl, G. (2024). Detectability of accretion-induced bosenovae in the Milky Way. arXiv e-prints, arXiv:2410.13082. doi:10.48550/arXiv.2410.13082
Matos, T., Avilez, A., Bernal, T., and Chavanis, P.-H. (2019). Energy balance of a Bose gas in a curved space-time. General Relativ. Gravit. 51, 159. doi:10.1007/s10714-019-2644-9
Matos, T., Ureña-López, L. A., and Lee, J.-W. (2024). Short review of the main achievements of the scalar field, fuzzy, ultralight, wave, BEC dark matter model. Front. Astronomy Space Sci. 11, 1347518. doi:10.3389/fspas.2024.1347518
Membrado, M., Abad, J., Pacheco, A. F., and Saudo, J. (1989a). Newtonian boson spheres. Phys. Rev. D. 40, 2736–2738. doi:10.1103/PhysRevD.40.2736
Membrado, M., Pacheco, A. F., and Sañudo, J. (1989b). Hartree solutions for the self-Yukawian boson sphere. Phys. Rev. A Gen. Phys. 39, 4207–4211. doi:10.1103/PhysRevA.39.4207
Michel, F., and Moss, I. G. (2018). Relativistic collapse of axion stars. Phys. Lett. B 785, 9–13. doi:10.1016/j.physletb.2018.07.063
Mocz, P., Vogelsberger, M., Robles, V. H., Zavala, J., Boylan-Kolchin, M., Fialkov, A., et al. (2017). Galaxy formation with BECDM - I. Turbulence and relaxation of idealized haloes. Mon. Not. R. Astron. Soc. 471, 4559–4570. doi:10.1093/mnras/stx1887
Mocz, P., Lancaster, L., Fialkov, A., Becerra, F., and Chavanis, P.-H. (2018). Schrödinger-Poisson–Vlasov-Poisson correspondence. Phys. Rev. D. 97, 083519. doi:10.1103/PhysRevD.97.083519
Mocz, P., Fialkov, A., Vogelsberger, M., Becerra, F., Amin, M. A., Bose, S., et al. (2019). First star-forming structures in fuzzy cosmic filaments. Phys. Rev. Lett. 123, 141301. doi:10.1103/PhysRevLett.123.141301
Mocz, P., Fialkov, A., Vogelsberger, M., Becerra, F., Shen, X., Robles, V. H., et al. (2020). Galaxy formation with BECDM - II. Cosmic filaments and first galaxies. Mon. Not. R. Astron. Soc. 494, 2027–2044. doi:10.1093/mnras/staa738
Mocz, P., Fialkov, A., Vogelsberger, M., Boylan-Kolchin, M., Chavanis, P.-H., Amin, M. A., et al. (2023). Cosmological structure formation and soliton phase transition in fuzzy dark matter with axion self-interactions. Mon. Not. R. Astron. Soc. 521, 2608–2615. doi:10.1093/mnras/stad694
Moore, B., Quinn, T., Governato, F., Stadel, J., and Lake, G. (1999). Cold collapse and the core catastrophe. Mon. Not. R. Astron. Soc. 310, 1147–1152. doi:10.1046/j.1365-8711.1999.03039.x
Musoke, N., Hotchkiss, S., and Easther, R. (2020). Lighting the dark: evolution of the postinflationary universe. Phys. Rev. Lett. 124, 061301. doi:10.1103/PhysRevLett.124.061301
Navarro, J. F., Frenk, C. S., and White, S. D. M. (1996). The structure of cold dark matter halos. Astrophys. J. 462, 563. doi:10.1086/177173
Nelson, E. (1966). Derivation of the Schrödinger equation from Newtonian mechanics. Phys. Rev. 150, 1079–1085. doi:10.1103/PhysRev.150.1079
Niemeyer, J. C., and Easther, R. (2020). Inflaton clusters and inflaton stars. J. Cosmol. Astropart. Phys. 2020, 030. doi:10.1088/1475-7516/2020/07/030
Niemeyer, J. C. (2020). Small-scale structure of fuzzy and axion-like dark matter. Prog. Part. Nucl. Phys. 113, 103787. doi:10.1016/j.ppnp.2020.103787
Nori, M., Macciò, A. V., and Baldi, M. (2023). Fuzzy aquarius: evolution of a Milky-Way like system in the fuzzy dark matter scenario. Mon. Notices R. Astronomical Soc. 522, 1451–1463. doi:10.1093/mnras/stad1081
Nottale, L. (2011). Scale relativity and fractal space-time: a new approach to unifying relativity and quantum mechanics. doi:10.1142/p752
Oñorbe, J., Boylan-Kolchin, M., Bullock, J. S., Hopkins, P. F., Kereš, D., Faucher-Giguère, C.-A., et al. (2015). Forged in FIRE: cusps, cores and baryons in low-mass dwarf galaxies. Mon. Not. R. Astron. Soc. 454, 2092–2106. doi:10.1093/mnras/stv2072
Oppenheimer, J. R., and Volkoff, G. M. (1939). On massive neutron cores. Phys. Rev. 55, 374–381. doi:10.1103/PhysRev.55.374
Ourabah, K. (2020). Jeans instability in dark matter halos. Phys. Scr. 95, 055005. doi:10.1088/1402-4896/ab7650
Padilla, L. E., Rindler-Daller, T., Shapiro, P. R., Matos, T., and Vázquez, J. A. (2021). Core-halo mass relation in scalar field dark matter models and its consequences for the formation of supermassive black holes. Phys. Rev. D. 103, 063012. doi:10.1103/PhysRevD.103.063012
Padilla, L. E., Hidalgo, J. C., and Malik, K. A. (2022). New mechanism for primordial black hole formation during reheating. Phys. Rev. D. 106, 023519. doi:10.1103/PhysRevD.106.023519
Padilla, L. E., Hidalgo, J. C., Gomez-Aguilar, T. D., Malik, K. A., and German, G. (2024). Primordial black hole formation during slow-reheating: a review. Front. Astronomy Space Sci. 11, 1361399. doi:10.3389/fspas.2024.1361399
Padmanabhan, T. (1990). Statistical mechanics of gravitating systems. Phys. Rep. 188, 285–362. doi:10.1016/0370-1573(90)90051-3
Painter, C. A., Boylan-Kolchin, M., Mocz, P., and Vogelsberger, M. (2024). An attractive model: simulating fuzzy dark matter with attractive self-interactions. Mon. Not. R. Astron. Soc. 533, 2454–2472. doi:10.1093/mnras/stae1912
Patkós, A. (2023). Electromagnetic energy loss of axion stars. Phys. Rev. D. 107, 055017. doi:10.1103/PhysRevD.107.055017
Pils, K., and Rindler-Daller, T. (2022). Orbits and adiabatic contraction in scalar-field dark matter halos: revisiting the cusp-core problem in dwarf galaxies. Mon. Not. R. Astron. Soc. 514, 1990–2009. doi:10.1093/mnras/stac1471
Poincaré, H. (1885). Sur l'équilibre d'une masse fluide animée d'un mouvement de rotation. Acta Math. 7, 259–380. doi:10.1007/bf02402204
Pontzen, A., and Governato, F. (2014). Cold dark matter heats up. Nature 506, 171–178. doi:10.1038/nature12953
Proukakis, N. P., Rigopoulos, G., and Soto, A. (2023). Unified description of corpuscular and fuzzy bosonic dark matter. Phys. Rev. D. 108, 083513. doi:10.1103/PhysRevD.108.083513
Proukakis, N. P., Rigopoulos, G., and Soto, A. (2024). Hybrid model of condensate and particle dark matter: linear perturbations in the hydrodynamic limit. Phys. Rev. D. 110, 023504. doi:10.1103/PhysRevD.110.023504
Raby, S. (2016). Axion star collisions with neutron stars and fast radio bursts. Phys. Rev. D. 94, 103004. doi:10.1103/PhysRevD.94.103004
Randall, S. W., Markevitch, M., Clowe, D., Gonzalez, A. H., and Bradač, M. (2008). Constraints on the self-interaction cross section of dark matter from numerical simulations of the merging galaxy cluster 1E 0657-56. Astrophys. J. 679, 1173–1180. doi:10.1086/587859
Rindler-Daller, T., and Shapiro, P. R. (2012). Angular momentum and vortex formation in Bose-Einstein-condensed cold dark matter haloes. Mon. Not. R. Astron. Soc. 422, 135–161. doi:10.1111/j.1365-2966.2012.20588.x
Rindler-Daller, T., and Shapiro, P. R. (2014). “Finding new signature effects on galactic dynamics to constrain Bose-Einstein-condensed cold dark matter,” in Accelerated cosmic expansion. Editors C. Moreno Gonzalez, J. E. Madriz Aguilar, and L. M. Reyes Barrera, 163, 163–182. doi:10.1007/978-3-319-02063-1_12
Robles, V. H., and Matos, T. (2013). Exact solution to finite temperature SFDM: natural cores without feedback. Astrophys. J. 763, 19. doi:10.1088/0004-637X/763/1/19
Romano-Díaz, E., Shlosman, I., Hoffman, Y., and Heller, C. (2008). Erasing dark matter cusps in cosmological galactic halos with baryons. Astrophys. J. 685, L105–L108. doi:10.1086/592687
Ruffini, R., and Bonazzola, S. (1969). Systems of self-gravitating particles in general relativity and the concept of an equation of state. Phys. Rev. 187, 1767–1783. doi:10.1103/PhysRev.187.1767
Ruffini, R., Argüelles, C. R., and Rueda, J. A. (2015). On the core-halo distribution of dark matter in galaxies. Mon. Notices R. Astronomical Soc. 451, 622–628. doi:10.1093/mnras/stv1016
Schive, H.-Y., Chiueh, T., and Broadhurst, T. (2014a). Cosmic structure as the quantum interference of a coherent dark wave. Nat. Phys. 10, 496–499. doi:10.1038/nphys2996
Schive, H.-Y., Liao, M.-H., Woo, T.-P., Wong, S.-K., Chiueh, T., Broadhurst, T., et al. (2014b). Understanding the core-halo relation of quantum wave dark matter from 3D simulations. Phys. Rev. Lett. 113, 261302. doi:10.1103/PhysRevLett.113.261302
Schobesberger, S. O., Rindler-Daller, T., and Shapiro, P. R. (2021). Angular momentum and the absence of vortices in the cores of fuzzy dark matter haloes. Mon. Not. R. Astron. Soc. 505, 802–829. doi:10.1093/mnras/stab1153
Schunck, F. E., and Mielke, E. W. (2003). General relativistic boson stars. Cl. Quantum Grav. 20, R301–R356. doi:10.1088/0264-9381/20/20/201
Schwabe, B., Niemeyer, J. C., and Engels, J. F. (2016). Simulations of solitonic core mergers in ultralight axion dark matter cosmologies. Phys. Rev. D. 94, 043513. doi:10.1103/PhysRevD.94.043513
Seidel, E., and Suen, W.-M. (1991). Oscillating soliton stars. Phys. Rev. Lett. 66, 1659–1662. doi:10.1103/PhysRevLett.66.1659
Seidel, E., and Suen, W.-M. (1994). Formation of solitonic stars through gravitational cooling. Phys. Rev. Lett. 72, 2516–2519. doi:10.1103/PhysRevLett.72.2516
Shapiro, P. R., Dawoodbhoy, T., and Rindler-Daller, T. (2022). Cosmological structure formation in scalar field dark matter with repulsive self-interaction: the incredible shrinking Jeans mass. Mon. Not. R. Astron. Soc. 509, 145–173. doi:10.1093/mnras/stab2884
Sharma, A., Kartvelishvili, G., and Khoury, J. (2022). Finite temperature description of an interacting Bose gas. Phys. Rev. D. 106, 045025. doi:10.1103/PhysRevD.106.045025
Shukla, S., Brachet, M. E., and Pandit, R. (2024a). Neutron-superfluid vortices and proton-superconductor flux tubes: development of a minimal model for pulsar glitches. Phys. Rev. D. 110, 083002. doi:10.1103/PhysRevD.110.083002
Shukla, S., Verma, A. K., Brachet, M. E., and Pandit, R. (2024b). Gravity- and temperature-driven phase transitions in a model for collapsed axionic condensates. Phys. Rev. D. 109, 063009. doi:10.1103/PhysRevD.109.063009
Sikivie, P., and Yang, Q. (2009). Bose-Einstein condensation of dark matter axions. Phys. Rev. Lett. 103, 111301. doi:10.1103/PhysRevLett.103.111301
Simon-Petit, A., Perez, J., and Plum, G. (2019). The status of isochrony in the formation and evolution of self-gravitating systems. Mon. Not. R. Astron. Soc. 484, 4963–4971. doi:10.1093/mnras/stz351
Sin, S.-J. (1994). Late-time phase transition and the galactic halo as a Bose liquid. Phys. Rev. D. 50, 3650–3654. doi:10.1103/PhysRevD.50.3650
Slepian, Z., and Goodman, J. (2012). Ruling out bosonic repulsive dark matter in thermal equilibrium. Mon. Not. R. Astron. Soc. 427, 839–849. doi:10.1111/j.1365-2966.2012.21901.x
Spano, M., Marcelin, M., Amram, P., Carignan, C., Epinat, B., and Hernandez, O. (2008). GHASP: an Hα kinematic survey of spiral and irregular galaxies - V. Dark matter distribution in 36 nearby spiral galaxies. Mon. Not. R. Astron. Soc. 383, 297–316. doi:10.1111/j.1365-2966.2007.12545.x
Spergel, D. N., and Steinhardt, P. J. (2000). Observational evidence for self-interacting cold dark matter. Phys. Rev. Lett. 84, 3760–3763. doi:10.1103/PhysRevLett.84.3760
Suárez, A., and Chavanis, P.-H. (2015a). Hydrodynamic representation of the Klein-Gordon-Einstein equations in the weak field limit: general formalism and perturbations analysis. Phys. Rev. D. 92, 023510. doi:10.1103/PhysRevD.92.023510
Suárez, A., and Chavanis, P.-H. (2015b). Hydrodynamic representation of the Klein-Gordon-Einstein equations in the weak field limit. J. Phys. Conf. Ser. (IOP) 654, 012008. doi:10.1088/1742-6596/654/1/012008
Suárez, A., and Chavanis, P.-H. (2017). Cosmological evolution of a complex scalar field with repulsive or attractive self-interaction. Phys. Rev. D. 95, 063515. doi:10.1103/PhysRevD.95.063515
Suárez, A., and Chavanis, P.-H. (2018). Jeans-type instability of a complex self-interacting scalar field in general relativity. Phys. Rev. D. 98, 083529. doi:10.1103/PhysRevD.98.083529
Suárez, A., and Matos, T. (2011). Structure formation with scalar-field dark matter: the fluid approach. Mon. Notices R. Astronomical Soc. 416, 87–93. doi:10.1111/j.1365-2966.2011.19012.x
Suárez, A., and Matos, T. (2014). Bose-Einstein condensate dark matter phase transition from finite temperature symmetry breaking of Klein-Gordon fields. Class. Quantum Gravity 31, 045015. doi:10.1088/0264-9381/31/4/045015
Suárez, A., Robles, V. H., and Matos, T. (2014). “A review on the scalar field/Bose-Einstein condensate dark matter model,” in Accelerated cosmic expansion. Editors C. Moreno Gonzalez, J. E. Madriz Aguilar, and L. M. Reyes Barrera, 107. doi:10.1007/978-3-319-02063-1_9
Takabayasi, T. (1952). On the formulation of quantum mechanics associated with classical pictures. Prog. Theor. Phys. 8, 143–182. doi:10.1143/ptp/8.2.143
Tangphati, T., Sakallı, İ., Banerjee, A., and Pradhan, A. (2024). Behaviors of quark stars in the Rainbow Gravity framework. Phys. Dark Universe 46, 101610. doi:10.1016/j.dark.2024.101610
Tkachev, I. I. (1986). Coherent scalar-field oscillations forming compact astrophysical object. Sov. Astron. Lett. 12, 305–308.
Uhlemann, C., Kopp, M., and Haugg, T. (2014). Schrödinger method as N-body double and UV completion of dust. Phys. Rev. D. 90, 023517. doi:10.1103/PhysRevD.90.023517
Veltmaat, J., Niemeyer, J. C., and Schwabe, B. (2018). Formation and structure of ultralight bosonic dark matter halos. Phys. Rev. D. 98, 043509. doi:10.1103/PhysRevD.98.043509
Veltmaat, J., Schwabe, B., and Niemeyer, J. C. (2020). Baryon-driven growth of solitonic cores in fuzzy dark matter halos. Phys. Rev. D. 101, 083518. doi:10.1103/PhysRevD.101.083518
Verma, A. K., Pandit, R., and Brachet, M. E. (2021). Formation of compact objects at finite temperatures in a dark-matter-candidate self-gravitating bosonic system. Phys. Rev. Res. 3, L022016. doi:10.1103/PhysRevResearch.3.L022016
Verma, A. K., Pandit, R., and Brachet, M. E. (2022). Rotating self-gravitating Bose-Einstein condensates with a crust: a model for pulsar glitches. Phys. Rev. Res. 4, 013026. doi:10.1103/PhysRevResearch.4.013026
Visinelli, L., Baum, S., Redondo, J., Freese, K., and Wilczek, F. (2018). Dilute and dense axion stars. Phys. Lett. B 777, 64–72. doi:10.1016/j.physletb.2017.12.010
Visinelli, L. (2021). Boson stars and oscillatons: a review. Int. J. Mod. Phys. D 30, 2130006–2130293. doi:10.1142/S0218271821300068
Wallstrom, T. C. (1994). Inequivalence between the Schrödinger equation and the Madelung hydrodynamic equations. Phys. Rev. A. 49, 1613–1617. doi:10.1103/PhysRevA.49.1613
Whitney, H. (1955). On singularities of mappings of Euclidean spaces. I. Mappings of the plane into the plane. Ann. Math. 62, 374–410. doi:10.2307/1970070
Widrow, L. M., and Kaiser, N. (1993). Using the Schroedinger equation to simulate collisionless matter. Astrophys. J. Lett. 416, L71. doi:10.1086/187073
Yavetz, T. D., Li, X., and Hui, L. (2022). Construction of wave dark matter halos: numerical algorithm and analytical constraints. Phys. Rev. D. 105, 023512. doi:10.1103/PhysRevD.105.023512
Yin, Z., and Visinelli, L. (2024). Axion star condensation around primordial black holes and microlensing limits. J. Cosmol. Astropart. Phys. 2024, 013. doi:10.1088/1475-7516/2024/10/013
Zagorac, J. L., Kendall, E., Padmanabhan, N., and Easther, R. (2023). Soliton formation and the core-halo mass relation: an eigenstate perspective. Phys. Rev. D. 107, 083513. doi:10.1103/PhysRevD.107.083513
Appendix 1: Energy functional
The GPP Equations 8 and 9 conserve the total mass
and the total energy
which is the sum of the kinetic energy
Using the Madelung transformation, these functionals can be written in terms of hydrodynamic variables as
where the first two terms in Equation A4 are the classical kinetic energy
An extremum of energy at fixed mass determines a stationary solution of the GPP equations. However, only minima of energy at fixed mass are dynamically stable (maxima or saddle points of energy at fixed mass are linearly unstable). A more detailed discussion of these results is given in Section 3.3 of Chavanis (2017b) and in Appendix B of Chavanis (2021a).
Appendix 2: Gaussian ansatz
We can obtain an approximate analytical solution of the GPP Equations 8 and 9 by using a Gaussian ansatz for the wavefunction and by developing a mechanical analogy (see Chavanis (2011c) and Section 8 of Chavanis (2017b) for details). We write the wavefunction as
where
We note that
With the Gaussian ansatz, we find that the energy functional (Equation A4) can be written as a function of
The first term in Equation A9 is the classical kinetic energy
This equation is similar to the equation of motion of a particle of mass
where
In the following, we shall specifically consider a standard BEC with the self-interaction potential from Equation 10 or with the equation of state from Equation 22. In that case,
with the coefficients
The equilibrium states are determined by the condition
This relation can also be obtained from the static quantum virial theorem (see Section 8.3 of Chavanis, 2017b for details)
The mass-radius relation in the noninteracting limit, the radius in the TF limit (in the repulsive case), and the maximum mass
Only minima of
where
This relation directly connects the stability of the self-gravitating BEC to the slope of the mass-radius relation. The equilibrium state is stable when
In the noninteracting case
Recalling that
In the case of an attractive self-interaction, expanding the effective potential from Equation A12 to third order close to the maximum mass
where
This equation has been studied in (Chavanis, 2016a; Chavanis, 2020c), leading to the results reported in Section 4.3. The mass-radius relation close to
The upper sign corresponds to the stable branch
where we have introduced the dynamical time
which is a special case of the general identity from Equation A17 valid close to the critical point where
The Gaussian ansatz approach has been generalized in Chavanis (2018b) to the case of a self-interaction potential comprising an attractive
with
It is also possible to generalize the theoretical formalism developed in (Chavanis, 2017b) to a multispecies system of self-gravitating BECs described by the coupled GPP equations
The mass
Making a Gaussian ansatz (for each species) and adapting the procedure described in Section 8 of Chavanis (2017b) to the present situation we readily obtain the energy function
with the effective potential
This potential displays a wealth of regimes.
Appendix 3: TF approximation
We assume that the bosons have a repulsive self-interaction
where the constant
which is equivalent to the Lane-Emden equation of index
In
The density vanishes at
The radius is independent of the mass. The central density is related to the mass by
In
The density vanishes at
where
In
The density vanishes at
The radius is independent of the mass. The central density is related to the mass by
Remark: In
and its solution is
where
where
The radius of the rotating configuration lies in the interval
The mass and the moment of inertia are related to the radius and to the angular velocity by
Using the Gaussian ansatz or the Ledoux formula, the squared pulsation is given by (Chavanis, 2017b)
Using Equations A43–A45 we can show that
Appendix 4: Mass-radius relation in dimensions
We briefly discuss the mass-radius relation of nonrelativistic self-gravitating BECs in
4.1 Repulsive self-interaction
When
When
When
These results show that the dimension of space
4.2 Attractive self-interaction
When
When
When
When
Appendix 5: The gravitational instability of an infinite homogeneous BEC when the Coriolis force is acting and a magnetic field is present
The effect of a uniform rotation and a constant magnetic field on the Jeans instability of an infinite homogeneous self-gravitating system has been studied by Chandrasekhar (Chandrasekhar and Fermi, 1953; Chandrasekhar, 1954; Chandrasekhar, 1955) for a classical fluid. For a quantum fluid, we just have to make the substitution
in the linearized hydrodynamic equations (see Section 8.1). In this manner, the results of Chandrasekhar (Chandrasekhar and Fermi, 1953; Chandrasekhar, 1954; Chandrasekhar, 1955) can be generalized straightforwardly. The dispersion relation taking into account the Coriolis force, a magnetic field and the quantum potential is given by Equation 7 of Chandrasekhar (1954) where now
As shown by Chandrasekhar (1954), the medium is stable if, and only if,
where
Different cases can happen depending on the signs of
which is different from the usual Jeans wavenumber
Appendix 6: Theories of modified gravity
There exist several theories of modified gravity. One of them is the so-called Eddington-inspired Born-Infeld (EiBI) theory (Bañados and Ferreira, 2010). In the Newtonian limit, the EiBI theory generates a modified Poisson equation of the form
where
The case
Another theory of modified gravity consists in replacing the Poisson equation by the screened Poisson (or Yukawa) equation. This type of equations can also describe interactions different from the gravitational one. Therefore, we write it under the generic form
where
This is similar to the Cahn-Hilliard expansion in soft matter physics. The first term leads to a self-interaction potential
with
with
Furthermore, one can show that the identity from Equation A17 remains valid in this more general situation. Under certain conditions, there is an extremum mass
In the noninteracting case
Stable equilibrium states only exist for
at which
In the nongravitational case
Stable equilibrium states only exist for
In the TF limit
Keywords: self-gravitating systems, Bose–Einstein condensates, scalar field, dark matter, axion stars, Klein–Gordon–Einstein equations, Schrödinger–Poisson equations, Gross–Pitaevskii–Poisson equations
Citation: Chavanis P-H (2025) A review of basic results on the Bose–Einstein condensate dark matter model. Front. Astron. Space Sci. 12:1538434. doi: 10.3389/fspas.2025.1538434
Received: 02 December 2024; Accepted: 06 February 2025;
Published: 20 August 2025.
Edited by:
Tonatiuh Matos, Center for Research and Advanced Studies, National Polytechnic Institute of Mexico (CINVESTAV), MexicoReviewed by:
Izzet Sakalli, Eastern Mediterranean University, TürkiyeDario Bettoni, University of León, Spain
Copyright © 2025 Chavanis. This is an open-access article distributed under the terms of the Creative Commons Attribution License (CC BY). The use, distribution or reproduction in other forums is permitted, provided the original author(s) and the copyright owner(s) are credited and that the original publication in this journal is cited, in accordance with accepted academic practice. No use, distribution or reproduction is permitted which does not comply with these terms.
*Correspondence: Pierre-Henri Chavanis, Y2hhdmFuaXNAaXJzYW1jLnVwcy10bHNlLmZy