Location via proxy:   [ UP ]  
[Report a bug]   [Manage cookies]                
11institutetext: Donostia International Physics Center, Paseo Manuel de Lardizabal 4, E-20118 Donostia-San Sebastián, Spain22institutetext: University of the Basque Country UPV/EHU, Department of Theoretical Physics, Bilbao, E-48080, Spain33institutetext: IKERBASQUE, Basque Foundation for Science, E-48013, Bilbao, Spain44institutetext: Dipartimento di Fisica ‘G. Occhialini’, Università degli Studi di Milano-Bicocca, Piazza della Scienza 3, I-20216 Milano, Italy 55institutetext: INFN, Sezione di Milano-Bicocca Piazza della Scienza 3, I-20126 Milano, Italy66institutetext: INAF - Osservatorio Astronomico di Brera, via Brera 20, I-20121 Milano, Italy77institutetext: Max-Planck-Institut für Astronomie, Königstuhl 17, D-69117 Heidelberg, Germany88institutetext: Institute of Astronomy, Pontificia Universidad Católica de Chile, Avenida Vicuña Mackena 4690, Santiago, Chile99institutetext: Universität Heidelberg, Seminarstrasse 2, D-69117 Heidelberg, Germany1010institutetext: Department of Astronomy, MongManWai Building, Tsinghua University, Beijing 100084, China

Demographics of Tidal Disruption Events with L-Galaxies

I. Volumetric TDE rates and the abundance of Nuclear Star Clusters
M. Polkas markos.polkas@ dipc.org1122    S. Bonoli1,3    E. Bortolas4,5    D. Izquierdo-Villalba 4455    A. Sesana4,5,6    L. Broggi4,5    N. Hoyer1,7,8,9    D. Spinoso10
(Received 2023; accepted 2023)

Stars can be ripped apart by tidal forces in the vicinity of a massive black hole (MBH), causing luminous flares known as tidal disruption events (TDEs). These events could be contributing to the mass growth of intermediate-mass MBHs, and new samples from transient surveys can provide useful information on this unexplored growth channel. This work aims to study the demographics of TDEs by modeling the co-evolution of MBHs and their galactic environments in a cosmological framework. We use the semi-analytic galaxy formation model L-GalaxiesBH, which follows the evolution of galaxies as well as of MBHs, including multiple scenarios for MBH seeds and growth, spin evolution, and binary MBH dynamics. Time-dependent TDE rates are associated with each MBH depending on the stellar environment, following the solutions to the 1-D Fokker Planck equation solved with PhaseFlow. Our model produces volumetric rates that are in agreement with the latest optical and previous X-ray samples. This agreement requires a high occupation fraction of nuclear star clusters with MBHs since these star reservoirs host the majority of TDEs at all mass regimes. We predict that TDE rates are an increasing function of MBH mass up to  105.5Msimilar-toabsentsuperscript105.5subscriptMdirect-product{\sim}\,10^{5.5}\,\mathrm{M_{\odot}}∼ 10 start_POSTSUPERSCRIPT 5.5 end_POSTSUPERSCRIPT roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT, beyond which the distribution flattens and eventually drops for > 107Mabsentsuperscript107subscriptMdirect-product{>}\,10^{7}\,\mathrm{M_{\odot}}> 10 start_POSTSUPERSCRIPT 7 end_POSTSUPERSCRIPT roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT. In general, volumetric rates are predicted to be redshift-independent at z< 1𝑧1z\,{<}\,1italic_z < 1. We discuss how the spin distribution of MBHs around the event horizon suppression can be constrained via TDE rates and what is the average contribution of TDEs to the MBH growth. In our work, the majority of low-mass galaxies host nuclear star clusters that have their loss-cone depleted by z= 0𝑧 0z\,=\,0italic_z = 0, explaining why TDEs are rare in these systems. This highlights that time-dependent TDE rates are essential for any model to be in good agreement with observations at all mass regimes.

Key Words.:
black holes – tidal-disruption-events – nuclear-star-clusters – semi-analytic – galaxy evolution

1 Introduction

A tidal disruption event (TDE) occurs when a star wanders too close to a massive111In the occurring field of TDEs from stellar mass black holes (Kremer et al. 2022, 2023; Vynatheya et al. 2024; Xin et al. 2024) the events are frequently referred to as micro-TDEs, so we preserve the term TDE for events from massive black holes. black hole (MBH) so that the MBH gravitational pull overcomes the star’s self-gravity. As a result, the star gets spagettified, and a part of it settles into a disk-like configuration producing a distinct, multi-wavelength electromagnetic flare. TDEs can outshine their host galaxies, with luminosities of 1042 1045superscript1042superscript104510^{42}\,{-}\,10^{45}10 start_POSTSUPERSCRIPT 42 end_POSTSUPERSCRIPT - 10 start_POSTSUPERSCRIPT 45 end_POSTSUPERSCRIPTerg s-1 which decline over weeks to years timescales. Most TDEs can be identified by the characteristic post-peak decrease of their accretion rate, which drops for most of the events as t5/3proportional-toabsentsuperscript𝑡53\,{\propto}\,t^{-5/3}∝ italic_t start_POSTSUPERSCRIPT - 5 / 3 end_POSTSUPERSCRIPT, as predicted by the standard fallback theory (Rees 1988; Phinney 1989).

The first observations of TDE-like transients, initially in the X-ray (e.g. Bade et al. 1996; Komossa & Bade 1999; Esquej et al. 2008) and then in the optical/UV sky (Gezari et al. 2006, 2008; van Velzen et al. 2011), sparked the interest on their overall rate per galaxy (Magorrian & Tremaine 1999; Syer & Ulmer 1999; Wang & Merritt 2004). Such interest has grown even further at the present date, as the number of observed TDEs is growing faster than ever (we now have identified approximately 100 TDE candidates), mainly owing to the advent of wide-field transient optical surveys, such as Pan-STARRS1 (Chornock et al. 2014), the Palomar Transient Factory (PTF, e.g. Cenko et al. 2012; Arcavi et al. 2014), the ongoing Zwicky Transient Facility (ZTF, e.g. Lin et al. 2022; Hammerstein et al. 2023; Yao et al. 2023), and ASAS-SN (e.g. Krolik et al. 2016; Hinkle et al. 2021; Liu et al. 2023). In the X-rays, eROSITA has already provided a sample of candidates (Sazonov et al. 2021) adding to previously compiled inventories from Swift, XMM-Newton, and Chandra (Kawamuro et al. 2016; Auchettl et al. 2017; Goldtooth et al. 2023). Finally, individual dust-shrouded TDE candidates have been detected in mid-infrared (Mattila et al. 2018; Kool et al. 2020; Wang et al. 2022; Panagiotou et al. 2023), and a recent analysis of NEOWISE data has yielded the first sample at this wavelength (Masterson et al. 2024).

The collection of these growing samples has allowed us to assess the TDE rates. In particular, the works of van Velzen (2018), Lin et al. (2022) and Yao et al. (2023) used a relatively wide sample of observed TDEs to infer an overall volumetric rate of  107 106similar-toabsentsuperscript107superscript106{\sim}\,10^{-7}\,{-}\,10^{-6}∼ 10 start_POSTSUPERSCRIPT - 7 end_POSTSUPERSCRIPT - 10 start_POSTSUPERSCRIPT - 6 end_POSTSUPERSCRIPT Mpc-3 yr-1 dlog10Lb1𝑑subscript10superscriptsubscript𝐿𝑏1d\log_{10}\,L_{b}^{-1}italic_d roman_log start_POSTSUBSCRIPT 10 end_POSTSUBSCRIPT italic_L start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT, with Lbsubscript𝐿𝑏L_{b}italic_L start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT being the peak luminosity within the band of a given survey, corresponding to a number of TDEs per galaxy of the order of 105 104superscript105superscript10410^{-5}\,{-}\,10^{-4}10 start_POSTSUPERSCRIPT - 5 end_POSTSUPERSCRIPT - 10 start_POSTSUPERSCRIPT - 4 end_POSTSUPERSCRIPT yr-1. Although these results necessarily depend on the shape of the TDE luminosity function, which remains uncertain, the obtained rates are in decent agreement with (although somewhat lower than) the TDE rates predicted by theoretical and numerical studies (Syer & Ulmer 1999; Magorrian & Tremaine 1999; Donley et al. 2002; Wang & Merritt 2004; Esquej et al. 2008; Merritt 2009; Brockamp et al. 2011; Stone & Metzger 2016; Pfister et al. 2021; Broggi et al. 2022). Nevertheless, both observational and theoretical estimates suffer from several limitations. On one side, the available sample of observed TDEs is still relatively small. On the other hand, for simplicity, theoretical models typically neglect important ingredients such as the time evolution of TDE rates under the evolution of the host galaxy. Most models are also affected by our poor knowledge of the MBH mass function at low masses ( 107Mless-than-or-similar-toabsentsuperscript107subscriptMdirect-product{\lesssim}\,10^{7}\,\mathrm{M_{\odot}}≲ 10 start_POSTSUPERSCRIPT 7 end_POSTSUPERSCRIPT roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT, see Shankar et al. 2013; Greene et al. 2020) and the occupation fraction of nuclear star clusters (NSCs, see recent studies Hoyer et al. 2023; Ashok et al. 2023; Hoyer et al. 2024).

Still, currently available samples have allowed us to perform statistical studies on the host galaxies of observed TDEs (see French et al. 2020b for a review). In particular, ZTF observations have shed light on the fact that TDEs are over-represented in ultra-luminous Infrared Galaxies (ULIRGs, Tadhunter et al. 2017; Reynolds et al. 2022) as well as in the rarer post-starburst/green-valley galaxies (Hammerstein et al. 2021, 2023), and in particular in the quiescent Balmer-strong (E+A) ones (Arcavi et al. 2014; French et al. 2016; Law-Smith et al. 2017; Graur et al. 2018; Dodd et al. 2021). The stellar mass distribution of TDE-hosts is relatively flat compared to the stellar mass function and is concentrated in the dwarf-to-massive transition regime,  109 1011Msimilar-toabsentsuperscript109superscript1011subscriptMdirect-product{\sim}\,10^{9}\,{-}\,10^{11}\,\mathrm{M_{\odot}}∼ 10 start_POSTSUPERSCRIPT 9 end_POSTSUPERSCRIPT - 10 start_POSTSUPERSCRIPT 11 end_POSTSUPERSCRIPT roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT (Wevers et al. 2019). Interestingly, the occupation fraction of NSCs peaks in the same mass range (e.g. Sánchez-Janssen et al. 2019; Hoyer et al. 2021). That being said, MBHs and NSCs frequently have been related in many works (Antonini et al. 2015; Naiman et al. 2015; Trani et al. 2018; Askar et al. 2022; Atallah et al. 2023; Lee et al. 2023; Tremmel et al. 2024), yet the exact nature of their connection remains unknown. What seems to be evident from theoretical studies (Merritt 2009; Stone & Metzger 2016; Pfister et al. 2020), is that the presence of an NSC, the densest stellar system possible in the universe (Neumayer et al. 2020), may enhance the rate of TDEs on the central MBH. Therefore, the majority of TDEs are expected to be related to intermediate-mass black holes (typically defined as 102.5 106Msuperscript102.5superscript106subscriptMdirect-product10^{2.5}\,{-}\,10^{6}\,\mathrm{M_{\odot}}10 start_POSTSUPERSCRIPT 2.5 end_POSTSUPERSCRIPT - 10 start_POSTSUPERSCRIPT 6 end_POSTSUPERSCRIPT roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT) in the center of NSCs. At the same time, the very existence of MBHs below 105Msuperscript105subscriptMdirect-product10^{5}\,\mathrm{M_{\odot}}10 start_POSTSUPERSCRIPT 5 end_POSTSUPERSCRIPT roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT is being challenged by the tentative and scarce observational evidence, especially towards the low end of the mass range (Mezcua 2017, but see recent robust evidence for an IMBH in Omega Centauri Häberle et al. 2024).

At this intermediate MBH-mass scale, the majority (>>>90%) of black holes are believed to be inactive (Greene et al. 2020; Mezcua & Domínguez Sánchez 2024), making samples inferred from Active Galactic Nuclei (AGN) observations incomplete. Also, mass measurements through spectral information become troublesome at low masses (Kormendy & Ho 2013). In fact, beyond the local universe where accurate dynamical measurements of MBH mass can be made, TDE observations serve as the only direct detection method of the dormant MBH population, as opposed to the indirect method of detecting relic AGN activity.

Finally, TDEs offer a viable channel of black hole growth (Hills 1975) that could in principle be dominant for MBHs that do not grow efficiently through gas accretion. After the first observations, the TDE growth channel was revisited (Magorrian & Tremaine 1999), with Milosavljević et al. (2006) proposing that low-mass MBHs (< 2× 106Mabsent2superscript106subscriptMdirect-product\,<\,2\,{\times}\,10^{6}\,\mathrm{M_{\odot}}< 2 × 10 start_POSTSUPERSCRIPT 6 end_POSTSUPERSCRIPT roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT) may acquire the majority of their mass through TDEs. Furthermore, Alexander & Bar-Or (2017) used this channel to set a lower limit on the masses of MBHs that can exist in the local universe since all MBH seeds should grow either through gas or TDEs. The initial growth through TDEs has been also studied within zoom-in cosmological simulations (Pfister et al. 2021; Lee et al. 2023) and, recently, high-resolution simulations achieved growing a black hole seed by a factor of >20absent20>20> 20 within 1 Gyr (Rizzuto et al. 2023). Yet, the argument of effective TDE growth has been questioned for low-mass galaxies, since NSCs which contribute mainly to the TDE rates are not dense enough to fuel black hole growth by runaway tidal capture of stars (Miller & Davies 2012; Stone et al. 2017), a physical process that is more efficient at growing intermediate-mass MBH in globular clusters (Portegies Zwart & McMillan 2002; Arca-Sedda 2016) and hierarchically-merging star clusters (see recent advancements in e.g. Rantala et al. 2024). Nevertheless, a statistical study on the frequency of TDEs across a realistic population of stellar environments and the role of stellar accretion in the growth of MBHs has not been performed so far.

In this paper, we set the foundation to address many of the aforementioned theoretical uncertainties by combining, for the first time, a semi-analytic model of galaxy/black hole evolution in a wide range of stellar environments with time-dependent TDE rates provided by a fast 1D-Fokker Planck solver. The paper is structured as follows; In Sect. 2 we describe our method for coupling the physics of TDEs in a given local environment with a variety of stellar environments. In Sect. 3 we describe our most important findings and compare our TDE rate predictions with new constraints. In particular, we focus on the cosmological evolution of TDE rates and the contribution from active MBHs. In Sect. 4, we complement our analysis by addressing the impact of the parameter choice and discussing the implications for the occupation fraction of NSCs, the MBH spin model, and MBH growth. In Sect. 5, we discuss some caveats and subjects that we aim to investigate explicitly in the future. Finally, we summarize the key aspects of our work in Sect.  6. Throughout the paper, we adopt a Lambda Cold Dark Matter (Λ(\Lambda( roman_ΛCDM) cosmology with parameters Ωm= 0.315subscriptΩm0.315\Omega_{\rm m}\,{=}\,0.315roman_Ω start_POSTSUBSCRIPT roman_m end_POSTSUBSCRIPT = 0.315, ΩΛ= 0.685subscriptΩΛ0.685\Omega_{\rm\Lambda}\,{=}\,0.685roman_Ω start_POSTSUBSCRIPT roman_Λ end_POSTSUBSCRIPT = 0.685, Ωb= 0.0493subscriptΩb0.0493\Omega_{\rm b}\,{=}\,0.0493roman_Ω start_POSTSUBSCRIPT roman_b end_POSTSUBSCRIPT = 0.0493, σ8= 0.826subscript𝜎80.826\sigma_{8}\,{=}\,0.826italic_σ start_POSTSUBSCRIPT 8 end_POSTSUBSCRIPT = 0.826 and H0= 67.4kms1Mpc1subscriptH067.4kmsuperscripts1superscriptMpc1\rm H_{0}\,{=}\,67.4\,\rm km\,s^{-1}\,Mpc^{-1}roman_H start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 67.4 roman_km roman_s start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT roman_Mpc start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT (Planck Collaboration et al. 2020).

2 Model Description

In this work, we combine the semi-analytical model of galaxy formation L-Galaxies with the 1D Fokker-Planck solver PhaseFlow to estimate the time evolution of TDE rates and compare it with observations. In particular, we use a version of L-Galaxies, dubbed as L-GalaxiesBH throughout this work, developed to study a wide variety of physical processes that drive the evolution of the MBH population and its co-evolution with galaxies. In this section, we first describe the L-Galaxies models and the additional physics included to model the TDE statistics. We then describe PhaseFlow and how it is linked to L-GalaxiesBH to assign TDE rates to MBHs.

2.1 The L-Galaxies semi-analytic models: Dark matter merger trees & baryonic physics

The L-Galaxies semi-analytic model is a well-tested model that tracks the cosmological evolution of the baryonic component of the Universe on top of dark matter merger trees (Croton et al. 2006; Guo et al. 2011; Henriques et al. 2015, 2020). It has been developed on, and is mainly being applied to, the dark matter merger trees of the Millennium-I (MS, Springel et al. 2005) and Millennium-II (MSII, Boylan-Kolchin et al. 2009) cosmological N-body simulations. In this work, we use the merger trees of the MSII which offer a higher mass resolution compared to the MS simulation. Specifically, the MSII has a dark matter particle resolution of 6.89× 106h1M6.89superscript106superscript1subscriptMdirect-product6.89\,{\times}\,10^{6}h^{-1}\,\mathrm{M_{\odot}}6.89 × 10 start_POSTSUPERSCRIPT 6 end_POSTSUPERSCRIPT italic_h start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT in a box size of 100h1100superscript1100h^{-1}100 italic_h start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPTMpc, enabling a good tracing of the cosmological evolution of halos where MBHs of 104 108Msuperscript104superscript108subscriptMdirect-product10^{4}\,{-}\,10^{8}\,\mathrm{M_{\odot}}10 start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT - 10 start_POSTSUPERSCRIPT 8 end_POSTSUPERSCRIPT roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT are placed. Originally, the MSII was run by using the WMAP1 & 2dFGRS “concordance” ΛΛ\Lambdaroman_ΛCDM framework (Spergel et al. 2003). However, the version of L-Galaxies used here applies the Angulo & White (2010) methodology to re-scale it to the cosmology of Planck 2018 data release (Planck Collaboration et al. 2020). This re-scaling modifies by a factor of 0.96 and 1.12 the MSII box size and particle mass, respectively.

Regarding the baryonic component, the current paradigm of galaxy evolution assumes that, as soon as a dark matter halo collapses, an amount of baryons, equal to the baryon fraction multiplied by the halo mass, also collapses within it (White & Frenk 1991). During this process, the infalling baryons are heated up and distributed inside the dark matter halo in the form of a diffuse, spherical, and quasi-static hot gas atmosphere. With time, this gas is allowed to cool down and migrate towards the center of the halo at a rate that depends on redshift and halo mass (White & Rees 1978; Sutherland & Dopita 1993). Due to angular momentum conservation, the cooled gas settles in a disk-like structure characterized by a radially exponential distribution. Once the disk becomes massive enough, star formation is triggered giving rise to a stellar component distributed in a disk with specific angular momentum inherited from the cold gas (Croton et al. 2006). Right after any star formation event, massive and short-lived stars explode polluting the interstellar medium and injecting energy in their environment, which can warm up and/or eject part of the cold gas of the disk. As a consequence of the ongoing stellar disk growth, some galaxies are prone to become unstable, with the subsequent disk instabilities leading to the formation of a stellar bulge component (Efstathiou et al. 1982). Besides supernova explosions, L-Galaxies assumes that MBHs in the center of the galaxy can prevent the gas from cooling in massive galaxies by injecting kinetic energy into the surrounding medium via quiescent gas accretion directly from the hot gas component (dubbed as “radio mode” accretion, see Croton et al. 2006), thus hampering the supply of cold gas to a galaxy’s disk. On top of secular processes, L-Galaxies models the interactions between galaxies, occurring after a given time of the fusion of their parent DM halos. Such interactions include major and minor galaxy mergers and alter the structure of the remnant galaxy by triggering bursts of star formation and leading to the formation of a stellar bulge or pure elliptical structure. Finally, L-Galaxies also takes into account environmental processes such as ram pressure stripping or galaxy disruptions in its galaxy formation paradigm (Henriques et al. 2015).

To improve the time resolution offered by the outputs of the MSII simulation, L-Galaxies does an internal time interpolation between two consecutive snapshots, with the time resolution being dtstep 5 20similar-to𝑑subscript𝑡step520dt_{\rm step}\,{\sim}\,5\,{-}\,20italic_d italic_t start_POSTSUBSCRIPT roman_step end_POSTSUBSCRIPT ∼ 5 - 20 Myr, depending on redshift.

2.1.1 Massive Black Holes in L-Galaxies

The version of L-Galaxies used in this work, L-GalaxiesBH, is based on the one presented in Henriques et al. (2015) with the modifications included in Izquierdo-Villalba et al. (2019, 2020, 2022) and Spinoso et al. (2023) to incorporate detailed massive black hole physics.

In brief, with respect to the model presented in Henriques et al. (2015), this new version adds a detailed model for the assembly (mass and spin) of nuclear MBHs, the dynamical evolution of MBH binaries, and the production of wandering MBHs (see Izquierdo-Villalba et al. 2020, 2022). Concerning the genesis of the first MBHs, L-GalaxiesBH models the spatial variations of the intergalactic metallicity and the Lyman-Werner background222The Lyman-Werner (LW) band is a specific energy-interval (i.e. hν=[11.2 13.6]eV𝜈delimited-[]11.213.6eVh\nu=[11.2\,-\,13.6]{\rm eV}italic_h italic_ν = [ 11.2 - 13.6 ] roman_eV) in the UV spectrum. LW photons are responsible for the photo-dissociation of molecular hydrogen (e.g. Haiman et al. 1997). produced by star formation events to account for the formation of heavy (i.e. Mseed 105Msimilar-tosubscript𝑀seedsuperscript105subscriptMdirect-productM_{\rm seed}\,{\sim}\,10^{5}\,\mathrm{M_{\odot}}italic_M start_POSTSUBSCRIPT roman_seed end_POSTSUBSCRIPT ∼ 10 start_POSTSUPERSCRIPT 5 end_POSTSUPERSCRIPT roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT) and intermediate-mass (Mseed 103 104Msimilar-tosubscript𝑀seedsuperscript103superscript104subscriptMdirect-productM_{\rm seed}\,{\sim}\,10^{3}\,{-}\,10^{4}\,\mathrm{M_{\odot}}italic_M start_POSTSUBSCRIPT roman_seed end_POSTSUBSCRIPT ∼ 10 start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT - 10 start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT) MBH-seeds, respectively via the collapse of pristine massive gas clouds and runaway stellar mergers within dense high-redshift333not to be confused with the ones introduced later in this work, for which we follow a different treatment NSCs. The formation of light seeds (Mseed 102Msimilar-tosubscript𝑀seedsuperscript102subscriptMdirect-productM_{\rm seed}\,{\sim}\,10^{2}\,\mathrm{M_{\odot}}italic_M start_POSTSUBSCRIPT roman_seed end_POSTSUBSCRIPT ∼ 10 start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT) after the explosion of the first metal-free stars (PopIII) is accounted for by grafting/inheriting the evolved counterparts of light-seed modeled self-consistently by the GametQSO/dust model (see e.g. Valiante et al. 2016). We note that concerning the black hole seeding model presented in Spinoso et al. (2023), we adopt a slightly higher amplitude of the “grafting probability”, setting the parameter Gp=0.25subscriptGp0.25\rm G_{p}=0.25roman_G start_POSTSUBSCRIPT roman_p end_POSTSUBSCRIPT = 0.25 (see Spinoso et al. 2023, for the definition of this parameter). This choice is motivated by the normalization of the z= 0𝑧 0z\,{=}\,0italic_z = 0 black hole mass function in the current work.

Once the first MBH seeds are formed, galaxy mergers and disk instabilities funnel new gas to the galactic nuclei, making it available for the growth of nuclear MBHs. The gas reaching the center is progressively consumed by the MBH first in an Eddington limited phase, followed by a sub-Eddington one (Bonoli et al. 2009; Izquierdo-Villalba et al. 2020). Episodes of gas accretion, on top of triggering BH growth, also modify its spin (Izquierdo-Villalba et al. 2020). The dynamical evolution of massive black hole binaries in a post-merger galaxy is two-fold (Begelman et al. 1980). During the pairing phase, the MBH(s) from the satellite galaxy migrate towards the galactic center via dynamical friction (following Binney & Tremaine 1987), forming a hard binary upon arrival at the nucleus. Then the hardening phase, where interactions with a circumbinary disc (gas-torque model by Dotti et al. 2015 and preferential growth as in Duffell et al. 2020) or the surrounding stars (following the model by Quinlan & Hernquist 1997; Sesana & Khan 2015) assist the gravitational-wave-driven evolution, along potential triplet interactions in the cases that a third MBHs comes in during the inspiral (modeled based on the results of Bonetti et al. 2018). The latter, along with gravitational recoil after merging of MBHs (as described in Lousto et al. 2012), can result in wandering MBHs. While L-GalaxiesBH can track the evolution of wandering MBHs, we do not include that in this work for simplicity (see the discussion on TDEs from wandering MBHs in Sect. 5).

As an example of the capability of L-GalaxiesBH to produce a realistic population of MBHs, in Fig. 1 we show the MBH mass function ϕ(M)italic-ϕsubscript𝑀\phi(M_{\bullet})italic_ϕ ( italic_M start_POSTSUBSCRIPT ∙ end_POSTSUBSCRIPT ), where Msubscript𝑀M_{\bullet}italic_M start_POSTSUBSCRIPT ∙ end_POSTSUBSCRIPT is the black hole mass (throughout this work), and the MBH median spin distribution χ~(M)subscript~𝜒subscript𝑀\tilde{\chi}_{\bullet}(M_{\bullet})over~ start_ARG italic_χ end_ARG start_POSTSUBSCRIPT ∙ end_POSTSUBSCRIPT ( italic_M start_POSTSUBSCRIPT ∙ end_POSTSUBSCRIPT ) at z= 0𝑧 0z\,{=}\,0italic_z = 0 and z= 5.5𝑧5.5z\,{=}\,5.5italic_z = 5.5 for the version of L-GalaxiesBH adopted in the current work. We compare our results with available data. Regarding the black hole mass function, as noted by Shankar et al. (2019), all observationally-derived values seem to converge at the high mass-end (M> 107.5Msubscript𝑀superscript107.5subscriptMdirect-productM_{\bullet}\,{>}\,10^{7.5}\,\mathrm{M_{\odot}}italic_M start_POSTSUBSCRIPT ∙ end_POSTSUBSCRIPT > 10 start_POSTSUPERSCRIPT 7.5 end_POSTSUPERSCRIPT roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT Merloni & Heinz 2008; Cao 2010; Gallo & Sesana 2019; Shankar et al. 2009, 2013; Mutlu-Pakdil et al. 2016; Vika et al. 2009; Aversa et al. 2015). However, constraints in the intermediate-mass range M 105106Msimilar-tosubscript𝑀superscript105superscript106subscriptMdirect-productM_{\bullet}\,{\sim}\,10^{5}-10^{6}\,\mathrm{M_{\odot}}italic_M start_POSTSUBSCRIPT ∙ end_POSTSUBSCRIPT ∼ 10 start_POSTSUPERSCRIPT 5 end_POSTSUPERSCRIPT - 10 start_POSTSUPERSCRIPT 6 end_POSTSUPERSCRIPT roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT are much less stringent. L-GalaxiesBH agrees with the most optimistic estimates from Shankar et al. (2009) and over-estimates with respect to all other available constraints. Regarding the spin distribution, the model agrees with the constraints from Reynolds (2021) at high masses, yet it predicts high-to-maximal spin for MBHs in the intermediate mass range, where there are no observational constraints. TDEs can potentially offer as a new probe for both the MBH mass and spin distribution in this range (see discussion in Sect. 4.3).

Refer to caption
Figure 1: MBH mass function (top) and median spin for X-ray bright MBHs (bottom) as a function of the MBH mass Msubscript𝑀M_{\bullet}italic_M start_POSTSUBSCRIPT ∙ end_POSTSUBSCRIPT predicted by the L-GalaxiesBH model used in this work; data are shown for z= 0𝑧 0z\,{=}\,0italic_z = 0 (red solid line) and z= 5.5𝑧5.5z\,{=}\,5.5italic_z = 5.5 (yellow dashed line), with shaded areas in the bottom panel referring to the 1σ𝜎\sigmaitalic_σ dispersion at a given mass range. In the top panel, the grey dashed line corresponds to the MBH mass function value equal to 1dex11desuperscriptx11\mathrm{dex}^{-1}1 roman_d roman_e roman_x start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT per MSII simulation volume; the results are compared with observational data at z= 0𝑧 0z\,{=}\,0italic_z = 0: MH08, Vika09, Sh09, Cao10, Sh13, Arvs15, GS19, MuPa16 refer to the model-dependent constraints on the MBH mass function derived respectively by Merloni & Heinz (2008),Cao (2010),Gallo & Sesana (2019),Shankar et al. (2009),Shankar et al. (2013),Mutlu-Pakdil et al. (2016),Vika et al. (2009),Aversa et al. (2015). For spin constraints, we display the upper and lower limits from X-ray reflection spectroscopy (Reynolds 2021). For a closer comparison to observational results, the average spin values shown here are for MBHs with a predicted hard X-ray luminosity of logLHX> 40subscript𝐿𝐻𝑋40\log\,L_{HX}\,{>}\,40roman_log italic_L start_POSTSUBSCRIPT italic_H italic_X end_POSTSUBSCRIPT > 40 erg s-1.

2.1.2 The Stellar Environment of Massive Black Holes

As mentioned above, the novelty of this work is the inclusion of TDEs within a full galaxy evolution context. To encompass this ambitious task it is necessary to model the stellar environment in which MBHs are embedded, on top of their formation and evolution. In this section, we describe how disks, bulges, and NSCs are included in L-GalaxiesBH. Together with black hole masses, the properties of the nuclear stellar environment will be the input for predicting time-evolving TDE rates with PhaseFlow.

Bulge and Disk profiles in L-Galaxies

Bulges in L-GalaxiesBH grow after galaxy interactions (major and minor mergers) and disk instabilities occurring in massive stellar disks. The specific properties of these events fully determine the final mass and extension (i.e. effective size) of the bulge. We stress that the scale length for disks Reff,dsubscript𝑅effdR_{\rm eff,d}italic_R start_POSTSUBSCRIPT roman_eff , roman_d end_POSTSUBSCRIPT and the effective size of bulges Reff,bsubscript𝑅effbR_{\rm eff,b}italic_R start_POSTSUBSCRIPT roman_eff , roman_b end_POSTSUBSCRIPT (equivalent to the scale radius of a Sérsic profile) are computed self-consistently inside L-GalaxiesBH by tracing the spin evolution of the galaxy components and applying energy conservation during mergers and disk instabilities. The profile of each bulge is assumed to follow a Sérsic model, whose steepness (i.e. Sérsic index) is associated with each bulge by using the observational results of Gadotti (2009), approximately a Gaussian distribution peaking at Sérsic index ns= 34subscript𝑛𝑠34n_{s}\,=\,3-4italic_n start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT = 3 - 4, as implemented in Izquierdo-Villalba et al. (2019). As already mentioned, galactic disks arise as a consequence of gas cooling and star formation events occurring in the center or dark matter halos. Together with galaxy encounters, these events determine the extent of the disk radial profile. Taking this into account, in this work we assume that pure disks with no bulge are characterized by the Sérsic index ns= 1subscript𝑛𝑠1n_{s}\,{=}\,1italic_n start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT = 1, regardless of redshift, and a scale radius of Rgal= 1.68Reff,dsubscript𝑅gal1.68subscript𝑅effdR_{\rm gal}\,{=}\,1.68\,R_{\rm eff,d}italic_R start_POSTSUBSCRIPT roman_gal end_POSTSUBSCRIPT = 1.68 italic_R start_POSTSUBSCRIPT roman_eff , roman_d end_POSTSUBSCRIPT. For the rest of this work, we refer to the sum of the disk and bulge mass as galaxy stellar mass Msubscript𝑀M_{\ast}italic_M start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT (the halo and intracluster stellar mass are neglected during our analysis), while Mgalsubscript𝑀galM_{\rm gal}italic_M start_POSTSUBSCRIPT roman_gal end_POSTSUBSCRIPT is reserved for the integrated mass of a Sérsic profile (either a bulge or a disk) with radius Rgalsubscript𝑅galR_{\rm gal}italic_R start_POSTSUBSCRIPT roman_gal end_POSTSUBSCRIPT. As we will see later, TDE events due to encounters between the nuclear MBHs and stars belonging to the bulge or disk component will be calculated assuming that these density profiles extend all the way to the center of the galaxy.

Nuclear Star Clusters in L-Galaxies

NSCs observed in the centers of a great fraction of dwarf and massive galaxies in the local Universe are the densest stellar structures known (see Neumayer et al. 2020, for a review). This inevitably suggests that they might be an ideal nursery for TDEs. In the following paragraphs, we describe a simple phenomenological model that we are introducing in L-GalaxiesBH to incorporate NSCs in galaxies (“nucleation”). An extensive and self-consistent model of the birth and evolution of NSCs will be presented in a future paper (Hoyer et al. in prep).

NSC Mass:

The mass of an NSC, MNSCsubscript𝑀NSCM_{\rm NSC}italic_M start_POSTSUBSCRIPT roman_NSC end_POSTSUBSCRIPT, is a fundamental property to be determined. To this end, we connect the NSC mass with the total galaxy stellar mass Msubscript𝑀M_{\ast}italic_M start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT of the host system via the following relation derived from observations of clusters in the local universe:

log10(MNSC/M)=A+Blog10(M/109.4M)subscript10subscript𝑀NSCsubscript𝑀direct-product𝐴𝐵subscript10subscript𝑀superscript109.4subscript𝑀direct-product\log_{10}(M_{\rm NSC}/M_{\odot})=A+B\log_{10}(M_{\ast}/10^{9.4}M_{\odot})roman_log start_POSTSUBSCRIPT 10 end_POSTSUBSCRIPT ( italic_M start_POSTSUBSCRIPT roman_NSC end_POSTSUBSCRIPT / italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT ) = italic_A + italic_B roman_log start_POSTSUBSCRIPT 10 end_POSTSUBSCRIPT ( italic_M start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT / 10 start_POSTSUPERSCRIPT 9.4 end_POSTSUPERSCRIPT italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT ) (1)

with

A=6.684&B={0.94,ifM>109.4M0.55,ifM109.4M.𝐴6.684𝐵cases0.94ifsubscript𝑀superscript109.4subscript𝑀direct-product0.55ifsubscript𝑀superscript109.4subscript𝑀direct-product\displaystyle A=6.684\;\;\&\;\;B=\begin{cases}0.94,&{\rm if\,}M_{\ast}>10^{9.4% }M_{\odot}\\ 0.55,&{\rm if\,}M_{\ast}\leq 10^{9.4}M_{\odot}\end{cases}.italic_A = 6.684 & italic_B = { start_ROW start_CELL 0.94 , end_CELL start_CELL roman_if italic_M start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT > 10 start_POSTSUPERSCRIPT 9.4 end_POSTSUPERSCRIPT italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT end_CELL end_ROW start_ROW start_CELL 0.55 , end_CELL start_CELL roman_if italic_M start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT ≤ 10 start_POSTSUPERSCRIPT 9.4 end_POSTSUPERSCRIPT italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT end_CELL end_ROW .

The high mass end of this relation is obtained from the work of Pechetti et al. (2020), while the lower mass end is adapted to the results from Hoyer et al. (2023). We also introduce a 0.5dex0.5dex0.5\,\mathrm{dex}0.5 roman_dex uniform scatter to these median values, comparable to the scatter of 0.23dex0.23dex0.23\,\mathrm{dex}0.23 roman_dex measured by Pechetti et al. (2020) to their relation as well as to the uncertainty on the assumed mass-to-light ratio of about 0.3dex0.3dex0.3\,\mathrm{dex}0.3 roman_dex (Roediger & Courteau 2015).

To avoid the formation of too many small clusters, when applying the above mass scaling relation at arbitrarily low galaxy stellar masses and at all redshifts we impose a minimum mass limit M,NSCmin=5×Mjeans(z)superscriptsubscript𝑀NSCmin5subscript𝑀jeans𝑧M_{\ast,\rm NSC}^{\rm min}={5}\,{\times}\,M_{\rm jeans}(z)italic_M start_POSTSUBSCRIPT ∗ , roman_NSC end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_min end_POSTSUPERSCRIPT = 5 × italic_M start_POSTSUBSCRIPT roman_jeans end_POSTSUBSCRIPT ( italic_z ) where Mjeans(z)subscript𝑀jeans𝑧M_{\rm jeans}(z)italic_M start_POSTSUBSCRIPT roman_jeans end_POSTSUBSCRIPT ( italic_z ) is the redshift-dependent Jeans mass for cold gas in the absence of a heat bath (Rees & Ostriker 1977). To avoid unphysically massive NSCs, we also impose a maximum mass limit of M,NSCmaxsubscript𝑀NSCM_{\ast,\rm NSC\,\max}italic_M start_POSTSUBSCRIPT ∗ , roman_NSC roman_max end_POSTSUBSCRIPT equal to 95% that of the galaxy component (bulge, or disk in the absence of bulge) Mgalsubscript𝑀galM_{\rm gal}italic_M start_POSTSUBSCRIPT roman_gal end_POSTSUBSCRIPT. These factors are arbitrary and are kept fixed throughout this work.

Nucleation:

In this work, we make the simple assumption that only galaxies with an MBH can host an NSC, although we are fully aware that the frequency of co-existence of MBHs and NSCs is observationally not yet fully established, with only a handful of NSCs and MBHs being detected in the same galaxy (Seth et al. 2008; Graham & Spitler 2009; Neumayer & Walcher 2012; Georgiev et al. 2016; Nguyen et al. 2018; Kimbrell et al. 2021; Nguyen et al. 2022; Ashok et al. 2023; Thater et al. 2023).444Notice the existence of a few nearby galaxies hosting an NSC but lacking an SMBH: M 33 (Gebhardt et al. 2001; Merritt et al. 2001), M 110 (Valluri et al. 2005), or NGC 7793 (Neumayer & Walcher 2012). In our toy model, galaxies hosting an MBH can also host an NSC depending on a simple step-function probability:

P(M,z)=P0,forM,NSCmin<M<M,NSCcutoffformulae-sequence𝑃subscript𝑀𝑧subscript𝑃0forsuperscriptsubscript𝑀NSCminsubscript𝑀subscript𝑀NSCcutoffP(M_{\ast},z)=P_{0},\;\;{\rm for}\;\;M_{\ast,\rm NSC}^{\rm min}<M_{\ast}<M_{% \ast,\rm NSC\,cut-off}italic_P ( italic_M start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT , italic_z ) = italic_P start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT , roman_for italic_M start_POSTSUBSCRIPT ∗ , roman_NSC end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_min end_POSTSUPERSCRIPT < italic_M start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT < italic_M start_POSTSUBSCRIPT ∗ , roman_NSC roman_cut - roman_off end_POSTSUBSCRIPT (2)

where M,NSCcutoffsubscript𝑀NSCcutoffM_{\ast,\rm NSC\,cut-off}italic_M start_POSTSUBSCRIPT ∗ , roman_NSC roman_cut - roman_off end_POSTSUBSCRIPT is a cut-off mass and P0subscript𝑃0P_{0}italic_P start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT is a free parameter. For our fiducial model, we assume P0= 1subscript𝑃01P_{0}\,=\,1italic_P start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 1. For the case of M,NSCcutoffsubscript𝑀NSCcutoffM_{\ast,\rm NSC\,cut-off}italic_M start_POSTSUBSCRIPT ∗ , roman_NSC roman_cut - roman_off end_POSTSUBSCRIPT our fiducial model uses the value 109.75Msuperscript109.75subscriptMdirect-product10^{9.75}\,\mathrm{M_{\odot}}10 start_POSTSUPERSCRIPT 9.75 end_POSTSUPERSCRIPT roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT motivated by the theoretical work of Antonini et al. (2015) and the observations of the Local Volume and close galaxy clusters (see e.g., Hoyer et al. 2021).

After formation, we assume that NSCs do not change in mass if the galaxy is evolving secularly (see discussion in Sect. 2.3.2) or experiences only minor mergers (in this case, the NSC of the central galaxy acquires the NSC of the merged satellite, following the dynamics of the companion MBH). Indeed, NSCs are extremely compact stellar systems and are expected to be difficult to be destroyed from external tidal fields during mergers (see the works of Bassino et al. 1994; Pfeffer et al. 2014; Mayes et al. 2021, for the detection of stripped nuclei).

Following these assumptions, the model predictions for the NSC occupation fraction at z= 0𝑧 0z\,{=}\,0italic_z = 0 for all galaxies hosting an MBH and for the full population are shown in Fig. 2. Regarding galaxies hosting an MBH, we see that below the cut-off galaxy stellar mass M,NSCcutoffsubscript𝑀NSCcutoffM_{\ast,\rm NSC\,cut-off}italic_M start_POSTSUBSCRIPT ∗ , roman_NSC roman_cut - roman_off end_POSTSUBSCRIPT, all galaxies also host an NSC, by construction. Above the cut-off mass, NSCs are present only in galaxies that were smaller at the time of NSC formation and then evolved secularly in stellar mass (see discussion in Sect. 2.3.2). When looking at the occupation fraction of the full galaxy population, instead, we see that dwarf galaxies are less likely to host an NSC because of the lower MBH occupation fraction. This is what is driving the total NSC occupation fraction to increase with galaxy stellar mass until approximately the cut-off mass, naturally following the logistic function that is fitted to the observations of Hoyer et al. (2021) up to M=109.5Msubscript𝑀superscript109.5subscriptMdirect-productM_{\ast}=10^{9.5}\,\mathrm{M_{\odot}}italic_M start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT = 10 start_POSTSUPERSCRIPT 9.5 end_POSTSUPERSCRIPT roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT.

Refer to caption
Figure 2: The NSC occupation fraction for our fiducial model as a function of galaxy stellar mass, for all galaxies (solid line) and all galaxies hosting an MBH (dashed line). All M< 109Msubscript𝑀superscript109subscriptMdirect-productM_{\ast}\,{<}\,10^{9}\,\mathrm{M_{\odot}}italic_M start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT < 10 start_POSTSUPERSCRIPT 9 end_POSTSUPERSCRIPT roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT galaxies hosting an MBH have also a 100% probability of hosting an NSC at creation. The data represents the NSC occupation fraction for the Virgo (orange circles), Fornax (white squares), Coma (grey triangles) clusters, and the Local Volume (green rhombuses) as presented in Hoyer et al. (2021). Our model fits the logistic function for NSC occupation at M< 109.5Msubscript𝑀superscript109.5subscriptMdirect-productM_{\ast}\,{<}\,10^{9.5}\,\mathrm{M_{\odot}}italic_M start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT < 10 start_POSTSUPERSCRIPT 9.5 end_POSTSUPERSCRIPT roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT of the same work (thin grey line). We stress that this agreement occurs naturally from the occupation of MBHs per galaxy (see discussion in Sect.4.2).

2.2 From galaxy properties to time-dependent TDE rates with PhaseFlow

Black hole masses and the properties of the nuclear stellar component of galaxies, modeled in L-GalaxiesBH as described above, provide the starting point for the calculation of time-dependent TDE rates.

The main, ubiquitous generation mechanism for TDEs is thought to be two-body relaxation (Chandrasekhar 1942; Frank & Rees 1976; Binney & Tremaine 2008). In simple terms, stars in the vicinity of an MBH deflect each other’s orbits so that a number of them may eventually reach a very small pericentre, closer than its tidal disruption radius, defined as

rt(Mm)1/3rsubscript𝑟tsuperscriptsubscript𝑀subscript𝑚13subscript𝑟r_{\rm t}\approx\left(\frac{M_{\bullet}}{m_{\star}}\right)^{1/3}r_{\star}italic_r start_POSTSUBSCRIPT roman_t end_POSTSUBSCRIPT ≈ ( divide start_ARG italic_M start_POSTSUBSCRIPT ∙ end_POSTSUBSCRIPT end_ARG start_ARG italic_m start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT end_ARG ) start_POSTSUPERSCRIPT 1 / 3 end_POSTSUPERSCRIPT italic_r start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT (3)

and be disrupted by the MBH of mass Msubscript𝑀M_{\bullet}italic_M start_POSTSUBSCRIPT ∙ end_POSTSUBSCRIPT; here msubscript𝑚m_{\star}italic_m start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT is the stellar mass and rsubscript𝑟r_{\star}italic_r start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT its radius. The occurrence of these events can be described in terms of the loss cone theory (see e.g., Merritt 2013; Stone et al. 2020), adopting a Fokker-Planck approach to treat two-body relaxation. In particular, in the present work, we make use of the PhaseFlow code (Vasiliev 2017), which is part of the AGAMA toolkit (Vasiliev 2019). PhaseFlow evolves in time a phase space density profile assuming a spherical and isotropic distribution function f(E)𝑓𝐸f(E)italic_f ( italic_E ), by solving the Poisson and orbit-averaged Fokker–Planck 1-D equations for the stellar distribution function, its gravitational potential, and its density. In general, two-body relaxation would induce an additional, explicit dependence of the distribution function on the angular momentum J𝐽Jitalic_J of the orbit, which allows for the computation of the rate of stars being tidally disrupted. The rate of stars with energy E𝐸Eitalic_E whose pericentre becomes smaller than the tidal disruption radius rtsubscript𝑟𝑡r_{t}italic_r start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT can be computed as the rate of stars whose angular momentum becomes smaller than 2GMrt2𝐺subscript𝑀subscript𝑟𝑡\sqrt{2\,G\,M_{\bullet}\,r_{t}}square-root start_ARG 2 italic_G italic_M start_POSTSUBSCRIPT ∙ end_POSTSUBSCRIPT italic_r start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT end_ARG. At each energy, PhaseFlow assumes the steady profile in angular momentum arising from relaxation (Cohn & Kulsrud 1978), and directly associates the isotropic profile f(E)𝑓𝐸f(E)italic_f ( italic_E ) to a (per energy) TDE rate. This results in an extra sink term in the energy-only Fokker-Planck equation and a growth term for the MBH mass. PhaseFlow has been extensively used in the framework of inferring TDE rates, including addressing the impact of the stellar mass function on TDE rates (Bortolas 2022) as well as predicting realistically partial disruption event rates (Bortolas et al. 2023).

2.2.1 PhaseFlow set-up

In our implementation, we assume that the stellar system surrounding the MBH is composed by a bi-chromatic population of stars made up of main-sequence stars of 0.38 M (encompassing 93absent93\approx 93≈ 93% of the total stellar mass) and 16M16subscriptMdirect-product16\,\mathrm{M_{\odot}}16 roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT stellar black holes555The choice for these values for the mass of main sequence stars and stellar black holes comes from the fact that those are close to the average masses for those objects assuming an evolved Kroupa (2001) stellar mass function; in addition, the second moment of the mass function, which sets the relaxation rate of the system and thus the TDE rates, attains a value which is very close to the actual one if we were to assume a complete and evolved stellar mass function (see e.g. Kochanek 2016; Stone & Metzger 2016; Pfister et al. 2022; Bortolas 2022). (encompassing the remaining \approx7%). Stars are considered to be destroyed if their separation to the MBH gets below rtsubscript𝑟tr_{\rm t}italic_r start_POSTSUBSCRIPT roman_t end_POSTSUBSCRIPT (Eq. 3), where we used r=0.44Rsubscript𝑟0.44subscript𝑅direct-productr_{\star}=0.44R_{\odot}italic_r start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT = 0.44 italic_R start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT, which is the expected radius of a 0.38 MsubscriptMdirect-product\rm M_{\odot}roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT star (see e.g. Bortolas 2022, Eq. 3, for more details). Once accretion occurs, 30% of the stellar mass666At a full disruption, the MBH captures 50% of the stellar mass. We selected a subsequent mass loss due to radiation and winds removing 40% of the bound material based on results from observations (Mockler & Ramirez-Ruiz 2021) and simulations (Bu et al. 2023). Simulations including radiative transfer from Steinberg & Stone (2024) yield a lower percentage of 15%. is added to the MBH, while the remainder is assumed to be lost in radiation and the interstellar medium. Stellar black holes are instead captured by the MBH if they get closer than 8GM/c28𝐺subscript𝑀superscript𝑐28GM_{\bullet}/c^{2}8 italic_G italic_M start_POSTSUBSCRIPT ∙ end_POSTSUBSCRIPT / italic_c start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT from the MBH, where c𝑐citalic_c is the speed of light in vacuum and G𝐺Gitalic_G the gravitational constant. During the evolution of the system, the stellar populations undergo the traditional dynamical phenomena expected in the vicinity to an MBH, i.e. they develop a Bahcall & Wolf (1976) cusp, stellar black holes segregate in the center dominating relaxation in the closest vicinity to the MBH and finally, once the system reaches a dynamical equilibrium, expands and lowers its TDE rates as a result of dynamical heating. All these phenomena are captured by PhaseFlow.

Given the fast performance of the code, we have generated multi-dimensional tables spanning a range of initial central MBH masses and a range of host-environment properties (mass, scale radius, compactness for disks, bulges, and NSCs), encompassing all values predicted by L-GalaxiesBH, as described in what follows. In particular, all runs in the multi-dimensional grid were initially evolved for a Hubble time using 300 bins in the phase-volume, the variable used in the code to parameterize the distribution function 777The phase-volume h(E)𝐸h(E)italic_h ( italic_E ) is the volume in phase space spanned by all the orbits with energy E𝐸Eitalic_E; unlike the orbital energy it is invariant under adiabatic changes of the potential, like in the case of a central MBH accreting via TDEs, and this makes h(E)𝐸h(E)italic_h ( italic_E ) a very good choice for this problem..

2.2.2 The PhaseFlow-generated grid of TDE rates

We model rates depending on their reservoir origin, which could be either a bulge/disk or an NSC. Rates are saved in a large multi-dimensional grid, which includes the parameter space of the MBH, the stellar environment properties, and the time dimension. Rates are not static but instead vary with time (not necessarily a steady state is reached). This adds substantial realism to our computation, as many systems dominating the overall TDE rate are often characterized by a very large initial rate which drops by orders of magnitude with time. Assuming the static rate (which would coincide with the t=0𝑡0t=0italic_t = 0 rate) thus results in a non-negligible overestimation of TDE rates, especially for the case of NSCs.

The parameter space mapped is the following:

MBH mass:

The first input parameter is the central MBH mass. The grid covers the range:

log10(M/M)[2.5,8.0]subscript10subscript𝑀subscript𝑀direct-product2.58.0\log_{10}(M_{\bullet}/M_{\odot})\in[2.5,8.0]roman_log start_POSTSUBSCRIPT 10 end_POSTSUBSCRIPT ( italic_M start_POSTSUBSCRIPT ∙ end_POSTSUBSCRIPT / italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT ) ∈ [ 2.5 , 8.0 ] (4)

in 34 equally spaced logarithmic steps. Here, the upper limit lies at the high-mass end of the event-horizon suppression defined by the range of values of the Hills mass. The Hills mass is defined as the MBH mass for which the tidal radius is within the horizon radius (rt<rgsubscript𝑟tsubscript𝑟gr_{\rm t}<r_{\rm g}italic_r start_POSTSUBSCRIPT roman_t end_POSTSUBSCRIPT < italic_r start_POSTSUBSCRIPT roman_g end_POSTSUBSCRIPT, Hills 1975). The Hills mass depends on the black hole spin and the infalling star properties and its orbit (Kesden 2012; Mummery 2024), but it is in the range M=107 108Msubscript𝑀superscript107superscript108subscriptMdirect-productM_{\bullet}=10^{7}\,{-}\,10^{8}\,\mathrm{M_{\odot}}italic_M start_POSTSUBSCRIPT ∙ end_POSTSUBSCRIPT = 10 start_POSTSUPERSCRIPT 7 end_POSTSUPERSCRIPT - 10 start_POSTSUPERSCRIPT 8 end_POSTSUPERSCRIPT roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT for an 0.38M0.38subscriptMdirect-product0.38\,\mathrm{M_{\odot}}0.38 roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT star (see also Sect. 2.3.4).

Galaxy stellar component:

To predict TDE rates, PhaseFlow needs the galaxy stellar mass Mgalsubscript𝑀galM_{\rm gal}italic_M start_POSTSUBSCRIPT roman_gal end_POSTSUBSCRIPT, the scale radius Rgalsubscript𝑅galR_{\rm gal}italic_R start_POSTSUBSCRIPT roman_gal end_POSTSUBSCRIPT and Sérsic index nssubscript𝑛𝑠n_{s}italic_n start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT. We map stellar mass values in a broad range around the MBH mass Msubscript𝑀M_{\bullet}italic_M start_POSTSUBSCRIPT ∙ end_POSTSUBSCRIPT values, following this scaling:

log10(Mgal/M)[1,4]subscript10subscript𝑀galsubscript𝑀14\log_{10}(M_{\rm gal}/M_{\bullet})\in[1,4]roman_log start_POSTSUBSCRIPT 10 end_POSTSUBSCRIPT ( italic_M start_POSTSUBSCRIPT roman_gal end_POSTSUBSCRIPT / italic_M start_POSTSUBSCRIPT ∙ end_POSTSUBSCRIPT ) ∈ [ 1 , 4 ] (5)

in 16 equally spaced logarithmic bins. The scale radius can instead vary in the range:

log10(Rgal/Rscl)[2,1]subscript10subscript𝑅galsubscript𝑅scl21\log_{10}(R_{\rm gal}/R_{\rm scl})\in[-2,1]roman_log start_POSTSUBSCRIPT 10 end_POSTSUBSCRIPT ( italic_R start_POSTSUBSCRIPT roman_gal end_POSTSUBSCRIPT / italic_R start_POSTSUBSCRIPT roman_scl end_POSTSUBSCRIPT ) ∈ [ - 2 , 1 ] (6)

with 13 equally spaced logarithmic bins, where Rsclsubscript𝑅sclR_{\rm scl}italic_R start_POSTSUBSCRIPT roman_scl end_POSTSUBSCRIPT depends on the stellar mass as (Shen et al. 2003):

log10(Rscl/kpc)=0.14log10(Mgal/M)1.21.subscript10subscript𝑅sclkpc0.14subscript10subscript𝑀galsubscript𝑀direct-product1.21\log_{10}(R_{\rm scl}/{\rm kpc})=0.14\log_{10}(M_{\rm gal}/M_{\odot})-1.21\;.roman_log start_POSTSUBSCRIPT 10 end_POSTSUBSCRIPT ( italic_R start_POSTSUBSCRIPT roman_scl end_POSTSUBSCRIPT / roman_kpc ) = 0.14 roman_log start_POSTSUBSCRIPT 10 end_POSTSUBSCRIPT ( italic_M start_POSTSUBSCRIPT roman_gal end_POSTSUBSCRIPT / italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT ) - 1.21 . (7)

Finally, for the Sérsic index we assume all integer values between one and seven, which is the range we probe in L-GalaxiesBH, as described in Sect. 2.1. This gives a grand total of 34×16×13×7= 49 50434161374950434\times 16\times 13\times 7\,=\,49{\,}50434 × 16 × 13 × 7 = 49 504 combinations of parameters for the galaxy stellar component. For galaxies with both a disk and bulge component, the contribution of the disk to TDE rates is ignored, as it is generally significantly lower. Note, however, that small galaxies in L-GalaxiesBH are often pure disks, and their contribution to the total volumetric rate is non-negligible.

NSCs:

To estimate the TDE rates in an NSC environment with PhaseFlow, the NSC mass MNSCsubscript𝑀NSCM_{\rm NSC}italic_M start_POSTSUBSCRIPT roman_NSC end_POSTSUBSCRIPT, effective radius RNSC,effsubscript𝑅NSCeffR_{\rm NSC,eff}italic_R start_POSTSUBSCRIPT roman_NSC , roman_eff end_POSTSUBSCRIPT and density profile are needed. Given the mass range for the galaxy component considered in the above grid (Eq. 5), the NSC mass range follows by applying to this the scaling relation presented in Eq. 1 (from combining the works of Pechetti et al. 2020 and Hoyer et al. 2023). From Pechetti et al. (2020) we also use the definition of the NSC effective radius, which correlates with the galaxy stellar mass Msubscript𝑀M_{\ast}italic_M start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT as:

log10(RNSC,eff/pc)=0.53+0.29log10(M/109M).subscript10subscript𝑅NSCeffpc0.530.29subscript10subscript𝑀superscript109subscript𝑀direct-product\log_{10}(R_{\rm NSC,eff}/{\rm pc})=0.53+0.29\log_{10}(M_{\ast}/10^{9}M_{\odot% }).roman_log start_POSTSUBSCRIPT 10 end_POSTSUBSCRIPT ( italic_R start_POSTSUBSCRIPT roman_NSC , roman_eff end_POSTSUBSCRIPT / roman_pc ) = 0.53 + 0.29 roman_log start_POSTSUBSCRIPT 10 end_POSTSUBSCRIPT ( italic_M start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT / 10 start_POSTSUPERSCRIPT 9 end_POSTSUPERSCRIPT italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT ) . (8)

We stress that this relation is especially holding towards the high-mass end but flattens out at low masses (Neumayer et al. 2020). To explore how the results presented in this work depend on the choice of fixed compactness of the NSC, we construct a second grid using Reff,NSCsubscript𝑅effNSCR_{\rm eff,NSC}italic_R start_POSTSUBSCRIPT roman_eff , roman_NSC end_POSTSUBSCRIPT equal to 1/3131/31 / 3 of the radius predicted by the scaling relation. We dub these additional runs as compactNSC. The fiducial model follows Eq. 8 and will be compared with the compactNSC one in Sect. 4.4.

The initial NSC density profile is assumed to follow the functional forms of Saha (1992) and Zhao (1996):

ρNSC(r)=ρ0(ra)γ[1+(ra)α]γβαexp[(rrcut)ξ].subscript𝜌NSC𝑟subscript𝜌0superscript𝑟𝑎𝛾superscriptdelimited-[]1superscript𝑟𝑎𝛼𝛾𝛽𝛼superscript𝑟subscript𝑟cut𝜉\rho_{\rm NSC}(r)=\rho_{0}\left(\frac{r}{a}\right)^{-\gamma}\left[1+\left(% \frac{r}{a}\right)^{\alpha}\right]^{\frac{\gamma-\beta}{\alpha}}\exp{\left[% \left(-\frac{r}{r_{\rm cut}}\right)^{\xi}\right]}.italic_ρ start_POSTSUBSCRIPT roman_NSC end_POSTSUBSCRIPT ( italic_r ) = italic_ρ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( divide start_ARG italic_r end_ARG start_ARG italic_a end_ARG ) start_POSTSUPERSCRIPT - italic_γ end_POSTSUPERSCRIPT [ 1 + ( divide start_ARG italic_r end_ARG start_ARG italic_a end_ARG ) start_POSTSUPERSCRIPT italic_α end_POSTSUPERSCRIPT ] start_POSTSUPERSCRIPT divide start_ARG italic_γ - italic_β end_ARG start_ARG italic_α end_ARG end_POSTSUPERSCRIPT roman_exp [ ( - divide start_ARG italic_r end_ARG start_ARG italic_r start_POSTSUBSCRIPT roman_cut end_POSTSUBSCRIPT end_ARG ) start_POSTSUPERSCRIPT italic_ξ end_POSTSUPERSCRIPT ] . (9)

where ρ0subscript𝜌0\rho_{0}italic_ρ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT is a normalization constant that ensures the total mass of the cluster is MNSCsubscript𝑀NSCM_{\rm NSC}italic_M start_POSTSUBSCRIPT roman_NSC end_POSTSUBSCRIPT, a=RNSC,eff/4.6𝑎subscript𝑅NSCeff4.6a\,=\,R_{\rm NSC,eff}/4.6italic_a = italic_R start_POSTSUBSCRIPT roman_NSC , roman_eff end_POSTSUBSCRIPT / 4.6 is its scale radius, while α= 4𝛼4\alpha\,=\,4italic_α = 4, β= 2𝛽2\beta\,=\,2italic_β = 2, and γ= 0.5𝛾0.5\gamma\,=\,0.5italic_γ = 0.5 are respectively the outer, intermediate and inner density slopes. Finally, rcut= 12asubscript𝑟cut12𝑎r_{\rm cut}\,=\,12aitalic_r start_POSTSUBSCRIPT roman_cut end_POSTSUBSCRIPT = 12 italic_a is a cutoff radius and ξ= 2𝜉2\xi\,=\,2italic_ξ = 2 represents the cutoff strength. The selected values have been chosen based on the results by Antonini et al. (2012), who explored the formation of NSCs through the infall of star clusters (see e.g. Capuzzo-Dolcetta 1993; Hartmann et al. 2011; Arca-Sedda & Capuzzo-Dolcetta 2014; Sánchez-Janssen et al. 2019; Fahrion et al. 2022; Carlsten et al. 2022; Leaman & van de Ven 2022; Hoyer et al. 2023, on the connection between NSCs and globular clusters). This profile should thus be a valid resemblance to the profile of a recently-formed NSC.

Taking into account the scaling relations described above, it is clear that all NSC properties ultimately depend only on the galaxy stellar mass Msubscript𝑀M_{\ast}italic_M start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT. Thus, the grid for the TDEs due to NSCs spans only the values of Msubscript𝑀M_{\bullet}italic_M start_POSTSUBSCRIPT ∙ end_POSTSUBSCRIPT and Msubscript𝑀M_{\ast}italic_M start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT, in the same ranges as mentioned for the galaxy stellar component above: total 34×16= 544341654434\times 16\,=\,54434 × 16 = 544 pairs of parameters.

Once the runs with PhaseFlow are completed and the grid in the parameter space has been fully mapped, we store the resulting event rates, ΓΓ\Gammaroman_Γ, which are a function of the aforementioned parameters and time t𝑡titalic_t. For the bulge/disk contribution, the rates are encapsulated in the function Γgal(Mgal,Rgal,ns;M,t)subscriptΓgalsubscript𝑀galsubscript𝑅galsubscript𝑛𝑠subscript𝑀𝑡\Gamma_{\rm gal}(M_{\rm gal},R_{\rm gal},n_{s};M_{\bullet},t)roman_Γ start_POSTSUBSCRIPT roman_gal end_POSTSUBSCRIPT ( italic_M start_POSTSUBSCRIPT roman_gal end_POSTSUBSCRIPT , italic_R start_POSTSUBSCRIPT roman_gal end_POSTSUBSCRIPT , italic_n start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ; italic_M start_POSTSUBSCRIPT ∙ end_POSTSUBSCRIPT , italic_t ), while for the NSCs, the rates are given by the function ΓNSC(MNSC;M,t)subscriptΓNSCsubscript𝑀NSCsubscript𝑀𝑡\Gamma_{\rm NSC}(M_{\rm NSC};M_{\bullet},t)roman_Γ start_POSTSUBSCRIPT roman_NSC end_POSTSUBSCRIPT ( italic_M start_POSTSUBSCRIPT roman_NSC end_POSTSUBSCRIPT ; italic_M start_POSTSUBSCRIPT ∙ end_POSTSUBSCRIPT , italic_t ). We store the rates differently for late times:

log10(t/yr)[7.0,10.146]subscript10𝑡yr7.010.146\log_{10}(t/{\rm yr})\in[7.0,10.146]roman_log start_POSTSUBSCRIPT 10 end_POSTSUBSCRIPT ( italic_t / roman_yr ) ∈ [ 7.0 , 10.146 ] (10)

and for early times

log10(t/yr)[1.0,7.0]subscript10𝑡yr1.07.0\log_{10}(t/{\rm yr})\in[1.0,7.0]roman_log start_POSTSUBSCRIPT 10 end_POSTSUBSCRIPT ( italic_t / roman_yr ) ∈ [ 1.0 , 7.0 ] (11)

with 50 and 60 evenly spaced logarithmic bins, respectively for the two time ranges. We consider separately TDE rates at times below 107yrsuperscript107yr10^{7}{\rm yr}10 start_POSTSUPERSCRIPT 7 end_POSTSUPERSCRIPT roman_yr given the time resolution of L-GalaxiesBH which spans between dtstep5 20Myrsimilar-to𝑑subscript𝑡step520Myrdt_{\rm step}\rm\sim 5\,{-}\,20\,Myritalic_d italic_t start_POSTSUBSCRIPT roman_step end_POSTSUBSCRIPT ∼ 5 - 20 roman_Myr, depending on redshift. Specifically, for MBH mass M<106Msubscript𝑀superscript106subscriptMdirect-productM_{\bullet}<10^{6}\,\mathrm{M_{\odot}}italic_M start_POSTSUBSCRIPT ∙ end_POSTSUBSCRIPT < 10 start_POSTSUPERSCRIPT 6 end_POSTSUPERSCRIPT roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT the peak TDE rate due to NSCs falls below the time resolution of L-GalaxiesBH dtstep10similar-to𝑑subscript𝑡step10dt_{\rm step}\sim 10italic_d italic_t start_POSTSUBSCRIPT roman_step end_POSTSUBSCRIPT ∼ 10Myr, therefore the initial phase of high TDE rates and growth (dubbed here as prompt phase) would not be treated properly by our model (see Fig. 3). To resolve that, when the TDE phase due to NSCs is initialized for these small MBHs, we draw a random time between 1111 and dtstep𝑑subscript𝑡stepdt_{\rm step}italic_d italic_t start_POSTSUBSCRIPT roman_step end_POSTSUBSCRIPT and assign rates from the prompt phase tables according to this time. At the same time, we add the integrated mass during the unresolved prompt phase to the MBH.

In the following section, we outline how we merge L-GalaxiesBH with the ΓgalsubscriptΓgal\Gamma_{\rm gal}roman_Γ start_POSTSUBSCRIPT roman_gal end_POSTSUBSCRIPT and ΓNSCsubscriptΓNSC\Gamma_{\rm NSC}roman_Γ start_POSTSUBSCRIPT roman_NSC end_POSTSUBSCRIPT rates constructed from the multi-dimensional grid described above. To guide the reader, Fig. 3 shows several examples of the time-evolving TDE rates for a variety of black hole masses and for a fixed set of parameters. In the figure, for each black hole mass, we assume that the corresponding galaxy stellar mass is given from the relation log10(M/107.43M)=0.62log10(Mgal/1010.5M)subscript10subscript𝑀superscript107.43subscriptMdirect-product0.62subscript10subscript𝑀galsuperscript1010.5subscriptMdirect-product\log_{10}\,(M_{\bullet}/10^{7.43}{\rm M}_{\odot})=0.62\log_{10}(M_{\rm gal}/10% ^{10.5}{\rm M}_{\odot})roman_log start_POSTSUBSCRIPT 10 end_POSTSUBSCRIPT ( italic_M start_POSTSUBSCRIPT ∙ end_POSTSUBSCRIPT / 10 start_POSTSUPERSCRIPT 7.43 end_POSTSUPERSCRIPT roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT ) = 0.62 roman_log start_POSTSUBSCRIPT 10 end_POSTSUBSCRIPT ( italic_M start_POSTSUBSCRIPT roman_gal end_POSTSUBSCRIPT / 10 start_POSTSUPERSCRIPT 10.5 end_POSTSUPERSCRIPT roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT ), which is the best-fit relation we get in L-GalaxiesBH at z= 0𝑧 0z\,{=}\,0italic_z = 0 for MBHs up to  108Msimilar-toabsentsuperscript108subscriptMdirect-product{\sim}\,10^{8}\,\mathrm{M_{\odot}}∼ 10 start_POSTSUPERSCRIPT 8 end_POSTSUPERSCRIPT roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT. The scale radius assigned to the Sérsic profile is here assumed to be an average value of Rgal=Rsclsubscript𝑅galsubscript𝑅sclR_{\rm gal}\,{=}\,R_{\rm scl}italic_R start_POSTSUBSCRIPT roman_gal end_POSTSUBSCRIPT = italic_R start_POSTSUBSCRIPT roman_scl end_POSTSUBSCRIPT. The disk corresponds to ns= 1subscript𝑛𝑠1n_{s}\,{=}\,1italic_n start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT = 1 and the bulge to ns= 4subscript𝑛𝑠4n_{s}\,{=}\,4italic_n start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT = 4. The matching NSC profiles are created by using the scaling relations presented in Eq. 1 and Eq. 8.

As shown from Fig. 3, the main general feature of our model is that the typical, average rates increase with increasing MBH mass, as the stellar reservoirs assigned are increasingly more massive for more massive MBHs. However, there is a strong differentiation between the galaxy component and the NSC component; while bulges and disks profiles often obtain a constant event rate given the large relaxation timescale of these systems, NSCs matching to M< 106Msubscript𝑀superscript106subscriptMdirect-productM_{\bullet}\,{<}\,10^{6}\,\mathrm{M_{\odot}}italic_M start_POSTSUBSCRIPT ∙ end_POSTSUBSCRIPT < 10 start_POSTSUPERSCRIPT 6 end_POSTSUPERSCRIPT roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT show decay in their initial TDE rate due to their fast relaxation dynamics. Typically, the decay happens earlier for smaller MBHs and smaller reservoirs. Notably, NSCs with Msubscript𝑀M_{\bullet}italic_M start_POSTSUBSCRIPT ∙ end_POSTSUBSCRIPT from 104Msuperscript104subscriptMdirect-product10^{4}\,\mathrm{M_{\odot}}10 start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT to 106Msuperscript106subscriptMdirect-product10^{6}\,\mathrm{M_{\odot}}10 start_POSTSUPERSCRIPT 6 end_POSTSUPERSCRIPT roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT start with similar TDE rates at 107yrssuperscript107yrs10^{7}\,\mathrm{yrs}10 start_POSTSUPERSCRIPT 7 end_POSTSUPERSCRIPT roman_yrs, but maintain it shorter times compared to the Hubble time, as the systems quickly reach their equilibrium in the form of a Bahcall & Wolf (1976) cusp and subsequently expand due to the dynamical heating resulting from efficient relaxation. That proves the importance of including the analysis of the initial rates that fall below L-GalaxiesBH time resolution.

Refer to caption
Refer to caption
Figure 3: Top panels: Stellar density profiles for a range of MBH masses as indicated in the inset legend (with the same color-coding applying to all panels of the figures). Disks are assigned nS= 1subscript𝑛𝑆1n_{S}\,=\,1italic_n start_POSTSUBSCRIPT italic_S end_POSTSUBSCRIPT = 1 Sérsic profiles (solid lines, left) while bulges nS= 4subscript𝑛𝑆4n_{S}\,=\,4italic_n start_POSTSUBSCRIPT italic_S end_POSTSUBSCRIPT = 4 profiles (dotted lines, left). NSCs are instead assigned spheroid profiles from Eq. 9 (right). The galaxy and NSC host properties scale with Msubscript𝑀M_{\bullet}italic_M start_POSTSUBSCRIPT ∙ end_POSTSUBSCRIPT as described in the text and are created to be representative of the average environment encountered in L-GalaxiesBH. Middle & bottom panels: Tidal disruption event rate evolution with PhaseFlow when initiating for different MBH mass with the associated profiles from the top panels, displayed separately for bulges/disks (galaxy component, middle panel) and NSCs (bottom). The grey region below 10101010Myr indicates the region where we trace unresolved growth and high-rates below the time-resolution dtstep𝑑subscript𝑡stepdt_{\rm step}italic_d italic_t start_POSTSUBSCRIPT roman_step end_POSTSUBSCRIPT for the black holes in NSCs with mass M<106Msubscript𝑀superscript106subscriptMdirect-productM_{\bullet}<10^{6}\,\mathrm{M_{\odot}}italic_M start_POSTSUBSCRIPT ∙ end_POSTSUBSCRIPT < 10 start_POSTSUPERSCRIPT 6 end_POSTSUPERSCRIPT roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT. This initial phase we refer to as prompt phase.

2.3 TDE rates in L-Galaxies

The final step consists of joining in a self-consistent way the TDE rates calculated by PhaseFlow and the properties and stellar environments of MBH predicted by L-GalaxiesBH. We schematically show this in Fig. 4, and we describe the details below.

Refer to caption
Figure 4: Flowchart of the current scheme for estimating TDEs within a single time-step of L-GalaxiesBH. The decision-making tree is colored red and blue for the galaxy (bulge or disk in the absence of bulge) and NSC components, respectively. With the density profile ρgalsubscript𝜌gal\rho_{\rm gal}italic_ρ start_POSTSUBSCRIPT roman_gal end_POSTSUBSCRIPT we refer to all galaxy component parameters {Mgal,Rgal,nS}subscript𝑀galsubscript𝑅galsubscript𝑛𝑆\,\left\{M_{\rm gal},R_{\rm gal},n_{S}\right\}{ italic_M start_POSTSUBSCRIPT roman_gal end_POSTSUBSCRIPT , italic_R start_POSTSUBSCRIPT roman_gal end_POSTSUBSCRIPT , italic_n start_POSTSUBSCRIPT italic_S end_POSTSUBSCRIPT } that are used to initialize the Sérsic profile. TDE rates from the galaxy as a whole and its NSC are effectively treated independently (each process has its clock ΔtTDΔsubscript𝑡TD\Delta t_{\rm TD}roman_Δ italic_t start_POSTSUBSCRIPT roman_TD end_POSTSUBSCRIPT) and can be active at the same time, while they sum together to produce the instantaneous rate of each MBH.

2.3.1 Event rates and mass accretion

As soon as an MBH forms inside a galaxy in L-GalaxiesBH, we set a TDE rate depending on the existence of a stellar bulge/disk and/or an NSC. Effectively, we look up the multi-dimensional grid generated with PhaseFlow for the rate ΓisubscriptΓ𝑖\Gamma_{i}roman_Γ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT, with i=gal𝑖gali\,{=}\,{\rm gal}italic_i = roman_gal for the galaxy component and i=NSC𝑖NSCi\,{=}\,{\rm NSC}italic_i = roman_NSC for the NSC. From this, we derive the time-evolving rates for each MBH and stellar environment. For galaxy and NSC properties we look up for the closest values in the grid, while we interpolate in the MBH and time dimension.888In only a few instances values predicted by L-GalaxiesBH are outside the parameter space covered by the grid. In these rare cases, we use the closest valid value.

When TDEs start being produced within a galaxy, either from the relaxation of the bulge/disk or the NSC, we start a clock, ΔtTD,iΔsubscript𝑡TDi\Delta t_{\rm TD,i}roman_Δ italic_t start_POSTSUBSCRIPT roman_TD , roman_i end_POSTSUBSCRIPT. In the successive steps of the L-GalaxiesBH run, the PhaseFlow grid is checked at the corresponding subsequent times. As described in the next section, we reset this clock to zero only when the MBH and/or the galaxy have substantially changed, which can happen, for example, after mergers.

Next, we define the fraction of the stellar mass accreted by the MBH after it has been tidally disrupted:
f,TDE={0.3,ifM108M1,otherwise.subscript𝑓TDEcases0.3ifsubscript𝑀superscript108subscriptMdirect-product1otherwisef_{\ast,\rm TDE}=\left\{\begin{array}[]{l}0.3,\;{\rm if}\;M_{\bullet}\leq 10^{% 8}{\rm M}_{\odot}\\ 1,\;{\rm otherwise}\end{array}\right..italic_f start_POSTSUBSCRIPT ∗ , roman_TDE end_POSTSUBSCRIPT = { start_ARRAY start_ROW start_CELL 0.3 , roman_if italic_M start_POSTSUBSCRIPT ∙ end_POSTSUBSCRIPT ≤ 10 start_POSTSUPERSCRIPT 8 end_POSTSUPERSCRIPT roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT end_CELL end_ROW start_ROW start_CELL 1 , roman_otherwise end_CELL end_ROW end_ARRAY .
In reality, around the Hills mass both direct captures and disruptions could take place, depending on the orbit of the infalling star 999as mentioned earlier, M= 108Msubscript𝑀superscript108subscriptMdirect-productM_{\bullet}\,{=}\,10^{8}\,\mathrm{M_{\odot}}italic_M start_POSTSUBSCRIPT ∙ end_POSTSUBSCRIPT = 10 start_POSTSUPERSCRIPT 8 end_POSTSUPERSCRIPT roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT is the approximate mass limit for event-horizon suppression; beyond this MBH mass stars are not disrupted, therefore f,TDE=1subscript𝑓TDE1f_{\ast,\,\mathrm{TDE}}=1italic_f start_POSTSUBSCRIPT ∗ , roman_TDE end_POSTSUBSCRIPT = 1 instead of 0.30.30.30.3., but we don’t expect this to change our results both on the TDE rates and the BH growth which, in this mass range, is anyway dominated by gas growth 101010For the TDE rate 105superscript10510^{-5}10 start_POSTSUPERSCRIPT - 5 end_POSTSUPERSCRIPTyr-1 a 108Msuperscript108subscriptMdirect-product10^{8}\,\mathrm{M_{\odot}}10 start_POSTSUPERSCRIPT 8 end_POSTSUPERSCRIPT roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT will grow over 10101010Gyr only 4×1044superscript1044\times 10^{-4}4 × 10 start_POSTSUPERSCRIPT - 4 end_POSTSUPERSCRIPT its mass, approximately equal to the gas growth at 103superscript10310^{-3}10 start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT of the Eddington-limited accretion episode over 15151515Myr..

Provided the TDE rate and the accretion fraction f,TDEsubscript𝑓TDEf_{\ast,\,\mathrm{TDE}}italic_f start_POSTSUBSCRIPT ∗ , roman_TDE end_POSTSUBSCRIPT per event, the mass accreted by the central MBH from each stellar component111111Note that we consider an NSC as a decomposed part of the central bulge/disk, that we subtract separately from. during each time-step of L-GalaxiesBH is given by:

dMi,acc=dtstepΓif,TDEm𝑑subscript𝑀𝑖absentacc𝑑subscript𝑡stepsubscriptΓ𝑖subscript𝑓TDEsubscript𝑚dM_{i,\bullet-\mathrm{acc}}=dt_{\rm step}\,\Gamma_{i}\,f_{\ast,\rm TDE}\,m_{\ast}italic_d italic_M start_POSTSUBSCRIPT italic_i , ∙ - roman_acc end_POSTSUBSCRIPT = italic_d italic_t start_POSTSUBSCRIPT roman_step end_POSTSUBSCRIPT roman_Γ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_f start_POSTSUBSCRIPT ∗ , roman_TDE end_POSTSUBSCRIPT italic_m start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT (12)

and the mass subtracted (i=gal𝑖gali=\rm galitalic_i = roman_gal or i=NSC𝑖NSCi=\rm NSCitalic_i = roman_NSC) is:

dMi,loss=dtstep[Γi(+ΓNSCifi=gal)]m.𝑑subscript𝑀𝑖loss𝑑subscript𝑡stepdelimited-[]subscriptΓ𝑖subscriptΓNSCif𝑖galsubscript𝑚dM_{i,\mathrm{loss}}=dt_{\rm step}\left[\,\Gamma_{i}\,(+\,\Gamma_{\rm NSC}\;{% \rm if}\,i=\rm gal)\,\right]\,m_{\ast}.italic_d italic_M start_POSTSUBSCRIPT italic_i , roman_loss end_POSTSUBSCRIPT = italic_d italic_t start_POSTSUBSCRIPT roman_step end_POSTSUBSCRIPT [ roman_Γ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( + roman_Γ start_POSTSUBSCRIPT roman_NSC end_POSTSUBSCRIPT roman_if italic_i = roman_gal ) ] italic_m start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT . (13)

Notice that, if both a bulge/disk and an NSC are present, we sum the two TDE rate contributions. The results of the mass growth of MBHs through this mechanism are discussed briefly in Sect. 4.1.

2.3.2 Conditions for changing/resetting TDE rates

As mentioned above, PhaseFlow is able to capture the time evolution of the stellar density profile as a result of relaxation and MBH growth due to star accretion. Thus, once TDEs start taking place within a galaxy in L-GalaxiesBH, we follow their time evolution based on the grid provided by PhaseFlow.

However, dramatic events, such as mergers, disk instabilities, and starbursts, can lead to major changes in the stellar environment surrounding the MBH. In these cases, we reset the clock for TDE rates (ΔtTD,iΔsubscript𝑡TDi\Delta t_{\rm TD,i}roman_Δ italic_t start_POSTSUBSCRIPT roman_TD , roman_i end_POSTSUBSCRIPT), and the new, unrelaxed stellar system starts again evolving towards a new relaxed state until the next violent dynamical event.

The conditions under which we consider it relevant to check the state of the TDE process taking place around an MBH are the following ones:

Changes in MBH mass:

If the mass accreted by the MBH via gas accretion is three times larger than the mass accumulated from TDEs, then the ΔtTD,NSCΔsubscript𝑡TDNSC\Delta t_{\rm TD,NSC}roman_Δ italic_t start_POSTSUBSCRIPT roman_TD , roman_NSC end_POSTSUBSCRIPT clock is reset again and the NSC mass and properties are adapted to the galaxy stellar mass at that time121212This condition guarantees the MBH does not move by more than three mass bins away from its initial mass on the PhaseFlow grid, at which point the assigned TDE rates we expect to be significantly different. For bulges and disks, rates are less time-dependent (see discussion in Sec. 2.2), so we omit this condition.

Changes in the galaxy component mass:

If during one L-GalaxiesBH time-step the bulge or disk stellar mass increases by more than 20%, the clock ΔtTD,galΔsubscript𝑡TDgal\Delta t_{\rm TD,gal}roman_Δ italic_t start_POSTSUBSCRIPT roman_TD , roman_gal end_POSTSUBSCRIPT is reset. Such variations in stellar mass in a short amount of time is a typical change in post-starburst galaxies (Kaviraj et al. 2007; van Velzen 2018; Wild et al. 2016, 2020). Note that events that do not modify the total stellar mass can also lead to a reset of the rates (e.g., the transfer of mass from the disk to the bulge during disk instabilities).

Changes in NSC mass:

We do not expect all changes of the galaxy component to affect the TDE rate evolution in NSCs, since these can continue relaxing at their own pace unaffected by the changes at larger galactic scales. However, as described in Sect. 2.1, an NSC can undergo significant mass changes following the host galaxy changes. Following the same approach described above, we select that NSC that grow more than >{>}\,>20% within one L-GalaxiesBH time-step will satisfy the condition of resetting the clock ΔtTD,NSCΔsubscript𝑡TDNSC\Delta t_{\rm TD,NSC}roman_Δ italic_t start_POSTSUBSCRIPT roman_TD , roman_NSC end_POSTSUBSCRIPT. This threshold is selected to be equal to that of the galaxy component in order to minimize the free parameters of the model.

Changes in NSC to MBH mass ratio:

While there are no clear constraints on the limits of the mass ratio between an NSC and the MBH at its center, namely 𝒟=MNSC/M𝒟subscript𝑀NSCsubscript𝑀\mathcal{D}\,=\,M_{\rm NSC}/M_{\bullet}caligraphic_D = italic_M start_POSTSUBSCRIPT roman_NSC end_POSTSUBSCRIPT / italic_M start_POSTSUBSCRIPT ∙ end_POSTSUBSCRIPT, we decided to put a lower limit on this ratio so that an NSC gets to be (re)generated only if 𝒟>𝒟nuc𝒟subscript𝒟nuc\mathcal{D}\,{>}\,\mathcal{D}_{\rm nuc}caligraphic_D > caligraphic_D start_POSTSUBSCRIPT roman_nuc end_POSTSUBSCRIPT. Only few physically-motivated arguments can provide some insight for the allowed value of 𝒟𝒟\mathcal{D}caligraphic_D: a) MBHs cannot be arbitrarily massive if they form in the cluster, e.g. simulations of runaway stellar collisions by Kritos et al. (2023) indicate that 𝒟> 1𝒟1\mathcal{D}\,>\,1caligraphic_D > 1 and b) the NSC destruction from binary MBHs e.g. total and partial disruption of clusters from intermediate-mass MBH binaries takes place for values 𝒟< 5and<15𝒟5and15\mathcal{D}\,{<}\,5\;{\rm and}\;<15caligraphic_D < 5 roman_and < 15 according to Khan & Holley-Bockelmann (2021). Given the theoretical uncertainties of the NSC-MBH symbiosis and provided local NSCs are observed with even lower values (𝒟 0.01similar-to𝒟0.01\mathcal{D}\,\sim\,0.01caligraphic_D ∼ 0.01 Neumayer & Walcher 2012), we select conservatively 𝒟nuc= 0.1subscript𝒟nuc0.1\mathcal{D}_{\rm nuc}\,=\,0.1caligraphic_D start_POSTSUBSCRIPT roman_nuc end_POSTSUBSCRIPT = 0.1. We stress that this parameter (along with M,NSCcutoffsubscript𝑀NSCcutoffM_{\ast,\rm NSCcut-off}italic_M start_POSTSUBSCRIPT ∗ , roman_NSCcut - roman_off end_POSTSUBSCRIPT) can be later substituted from a physically motivated formulation when the NSC model is implemented in L-GalaxiesBH by Hoyer et al. (in prep).

We have checked that variations within an order of magnitude of the thresholds mentioned above do not significantly affect the results on the volumetric rates presented in this work. However, we will further discuss some of these assumptions in Sect. 4.4.

2.3.3 Treatment of TDEs in AGNs

When MBHs experience phases of gas accretion, they shine as active galactic nuclei (AGN). These phases can be coeval with events of stars disruptions. The two processes may not be completely unrelated to each other. For instance, it has been proposed that TDE rates can be enhanced due to the alignment of retrograde orbits of NSC stars with the MBH accretion disk (Generozov & Perets 2023; Nasim et al. 2023) or due to the turn-off of the disk (Wang et al. 2024) or due to the quadrupole moment of the disk modifying the loss cone(Kaur & Stone 2024). However, by construction, our model does not account for such possible rate enhancement since gas physics are not included in PhaseFlow. Regarding the impact of TDEs on the AGN disk, the few TDE interpretations of fast flares in AGN light curves suggest a (temporal) post-flare increase (Blanchard et al. 2017; Liu et al. 2020) and others a decrease (Cannizzaro et al. 2022; Cao et al. 2023) of the AGN luminosity. Given the high uncertainty, we assume that the disk recovers quickly (compared to the time to the next TDE) or is not interrupted at all by TDEs so that both the TDE and AGN luminosity are kept as independent properties of each MBH. We do not expect this assumption to affect the global rates, due to the small AGN fraction at z 0similar-to𝑧 0z\,{\sim}\,0italic_z ∼ 0 (<{<}<10% of MBHs are active) unless the TDE rates are significantly boosted in the presence of an accretion disk (by a factor of 10 or greater).

2.3.4 From simulated to observable TDE rates

Once TDEs are carefully linked to the MBHs and their galaxies evolving within L-GalaxiesBH, we need to transform the individual rates to observable events to finally compare our predictions with the events picked up by time-domain surveys. Here we keep the following minimum number of assumptions to translate simulated TDE rates to observed ones:

  • \bullet

    We assume that all TDE events are full disruptions. In reality, for massive stars, only those with a pericenter that is a fraction of the tidal radius are fully disrupted (otherwise they are partial), and the energy output of the events may depend on the penetration factor of the star. Moreover, provided that a handful of partial disruption events have been identified (Payne et al. 2021; Malyali et al. 2023), we discuss the impact of such a choice in Sect. 5.

  • \bullet

    We assume that all TDEs give a luminosity that falls within the sensitivity of current surveys, i.e., all events are detectable, if there is no AGN contribution. If there is no AGN contribution. In the presence of an AGN, we ignore TDEs in MBHs with AGN bolometric luminosity Lbol> 1042subscript𝐿bolsuperscript1042L_{\rm bol}\,{>}\,10^{42}italic_L start_POSTSUBSCRIPT roman_bol end_POSTSUBSCRIPT > 10 start_POSTSUPERSCRIPT 42 end_POSTSUPERSCRIPTerg/s (fiducial model) unless noted otherwise. Note also that we are not including any obscuration and line-of-sight effects.

  • \bullet

    Regarding the event horizon suppression, the predicted TDE rates for each MBH are considered either visible or not depending on the Hills mass of the black hole itself. The Hills mass depends on (i) the mass Msubscript𝑀M_{\bullet}italic_M start_POSTSUBSCRIPT ∙ end_POSTSUBSCRIPT and spin χsubscript𝜒\chi_{\bullet}italic_χ start_POSTSUBSCRIPT ∙ end_POSTSUBSCRIPT of the MBH, (ii) the age of the loss-cone since the cluster/galaxy component was last reset ΔtTD,iΔsubscript𝑡TDi\Delta t_{\rm TD,i}roman_Δ italic_t start_POSTSUBSCRIPT roman_TD , roman_i end_POSTSUBSCRIPT and (iii) the mass of the infalling star. For this last point, we draw the stellar mass msubscript𝑚m_{\ast}italic_m start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT of the disrupted star from a truncated Kroupa mass function (m< 1.5Msubscript𝑚1.5subscriptMdirect-productm_{\ast}\,{<}\,1.5\,\mathrm{M_{\odot}}italic_m start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT < 1.5 roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT). Given the values above, each MBH is assigned a Hills mass MHills(m,ΔtTD,i,χ)subscript𝑀𝐻𝑖𝑙𝑙𝑠subscript𝑚Δsubscript𝑡TDisubscript𝜒M_{Hills}(m_{\ast},\Delta t_{\rm TD,i},\chi_{\bullet})italic_M start_POSTSUBSCRIPT italic_H italic_i italic_l italic_l italic_s end_POSTSUBSCRIPT ( italic_m start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT , roman_Δ italic_t start_POSTSUBSCRIPT roman_TD , roman_i end_POSTSUBSCRIPT , italic_χ start_POSTSUBSCRIPT ∙ end_POSTSUBSCRIPT ) based on the tabulated results of Huang & Lu (2024, Table D1) who have performed a series of simulations of solar-metallicity stars disrupted by MBHs131313These authors may have used slightly different environmental set-ups, but we expect the values to be quantitatively close.. If M>MHillssubscript𝑀subscript𝑀𝐻𝑖𝑙𝑙𝑠M_{\bullet}\,{>}\,M_{Hills}italic_M start_POSTSUBSCRIPT ∙ end_POSTSUBSCRIPT > italic_M start_POSTSUBSCRIPT italic_H italic_i italic_l italic_l italic_s end_POSTSUBSCRIPT, we omit this MBH from the sum of the bulk rates. Otherwise, we assume that all events are full TDEs and have the rate assigned to the MBH.141414From theory, e.g. MacLeod et al. (2012) and Kochanek (2016), for a given Msubscript𝑀M_{\bullet}italic_M start_POSTSUBSCRIPT ∙ end_POSTSUBSCRIPT, TDE rates scale with the initial mass function dN/dm𝑑𝑁𝑑subscript𝑚dN\,{/}dm_{\ast}italic_d italic_N / italic_d italic_m start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT and power-law m1/6superscriptsubscript𝑚16m_{\ast}^{1/6}italic_m start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 1 / 6 end_POSTSUPERSCRIPT. This will matter mostly beyond the maximum Hills mass for m= 0.38Msubscript𝑚0.38subscriptMdirect-productm_{\ast}\,{=}\,0.38\,\mathrm{M_{\odot}}italic_m start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT = 0.38 roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT (M> 108Msubscript𝑀superscript108subscriptMdirect-productM_{\bullet}\,{>}\,10^{8}\,\mathrm{M_{\odot}}italic_M start_POSTSUBSCRIPT ∙ end_POSTSUBSCRIPT > 10 start_POSTSUPERSCRIPT 8 end_POSTSUPERSCRIPT roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT), a mass range that TDE rates are shown only indicatively. Although the dependence on the initial mass function is captured with the sampling of stars, we do not rescale rates with this power-law dependence for massive monochromatic stars when adding them to the bulk rates.

Note also that, given the mass resolution of the dark matter simulation used and of the multidimensional grid of TDE rates, we omit from the analysis the TDE rates from galaxies with masses M< 105.5Msubscript𝑀superscript105.5subscriptMdirect-productM_{\ast}\,{<}\,10^{5.5}\,\mathrm{M_{\odot}}italic_M start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT < 10 start_POSTSUPERSCRIPT 5.5 end_POSTSUPERSCRIPT roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT and/or MBHs M< 102.5Msubscript𝑀superscript102.5subscriptMdirect-productM_{\bullet}\,{<}\,10^{2.5}\,\mathrm{M_{\odot}}italic_M start_POSTSUBSCRIPT ∙ end_POSTSUBSCRIPT < 10 start_POSTSUPERSCRIPT 2.5 end_POSTSUPERSCRIPT roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT. Finally, note that given all the assumptions above, the volumetric rates presented below can be considered optimistic.

3 Results

In this section, we present our main results on the predicted TDE rates for the fiducial choice of parameters and how they compare with the TDE rates observed by the latest time-domain campaigns. We also discuss the implications of our model for the time evolution of TDE rates and the general properties of the MBHs and galaxies hosting TDEs. The interpretation of the results in view of the population of NSCs and MBHs, the implications for MBH spin and growth through TDEs are discussed later in Sect. 4.

Refer to caption
Refer to caption
Figure 5: Volumetric TDE rates per black hole (Msubscript𝑀M_{\bullet}italic_M start_POSTSUBSCRIPT ∙ end_POSTSUBSCRIPT, left) and galaxy (Msubscript𝑀M_{\ast}italic_M start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT, right) log mass at z= 0.0𝑧0.0z\,{=}\,0.0italic_z = 0.0 for the fiducial model (solid line). Our model is compared against the constraints (here plotted with the error range) from Yao et al. (2023) for all MBHs (left) and all galaxy types (right). For reference, we display the previous constraints from van Velzen (2018) and the luminosity function from Sazonov et al. (2021) transformed to a mass function (see the details in the text). We also display the model lines of Yao et al. (2023) for the TDE rates as a function of a) MBH mass, assuming the Shankar et al. (2016, maroon shorter dash-dotted line; Sh16) and the Gallo & Sesana (2019, cyan long dash-dotted line; G19) MBH mass functions, both including the event-horizon suppression b) galaxy stellar mass, obtained by multiplying the galaxy stellar mass function with the power-law dependence of rates on M0.41superscriptsubscript𝑀0.41M_{\ast}^{-0.41}italic_M start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 0.41 end_POSTSUPERSCRIPT (grey dotted-dashed, SMFxPL).

3.1 Volumetric TDE rates

In Fig. 5, we present our predictions for the volumetric rates as a function of MBH and galaxy stellar mass for our fiducial model, comparing them with the recent constraints of Yao et al. (2023) derived from a sample of 33 TDEs from the ZTF survey. As shown in the left panel of Fig. 5, our model can reproduce the overall shape of the volumetric rates, including the flattening at intermediate MBH masses and the drop for M> 106.5Msubscript𝑀superscript106.5subscriptMdirect-productM_{\bullet}\,{>}\,10^{6.5}\,\mathrm{M_{\odot}}italic_M start_POSTSUBSCRIPT ∙ end_POSTSUBSCRIPT > 10 start_POSTSUPERSCRIPT 6.5 end_POSTSUPERSCRIPT roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT. Yao et al. (2023) also derived a simple model for the volumetric rates as a function of MBH mass using the Gallo & Sesana (2019) and Shankar et al. (2016) mass functions and including the event-horizon suppression (dotted-dashed lines in the left panel of the figure), and our predictions are in overall agreement with these models.

We note that the higher normalization with respect to Yao et al. (2023) at intermediate masses (M=105.5107Msubscript𝑀superscript105.5superscript107subscriptMdirect-productM_{\bullet}=10^{5.5}-10^{7}\,\mathrm{M_{\odot}}italic_M start_POSTSUBSCRIPT ∙ end_POSTSUBSCRIPT = 10 start_POSTSUPERSCRIPT 5.5 end_POSTSUPERSCRIPT - 10 start_POSTSUPERSCRIPT 7 end_POSTSUPERSCRIPT roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT) is due to the presence in our model of many black holes in this mass range in dwarf galaxies (M=106109Msubscript𝑀superscript106superscript109subscriptMdirect-productM_{*}=10^{6}-10^{9}\,\mathrm{M_{\odot}}italic_M start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT = 10 start_POSTSUPERSCRIPT 6 end_POSTSUPERSCRIPT - 10 start_POSTSUPERSCRIPT 9 end_POSTSUPERSCRIPT roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT), which is a stellar mass range still not constrained by current datasets (however see discussion on partial disruption events in Sect. 4). Also, Yao et al. (2023) assign to the detected events black hole masses based on the Msubscript𝑀M_{\bullet}italic_M start_POSTSUBSCRIPT ∙ end_POSTSUBSCRIPTσsubscript𝜎\sigma_{\ast}italic_σ start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT relation from (Kormendy & Ho 2013), which predicts smaller MBH masses in the low stellar mass regime with respect to the predictions of our model and the observational results of, e.g., Reines & Volonteri (2015). Alternatively, we could obtain a lower normalization in the volumetric rates at intermediate masses by decreasing the MBH and/or the NSC occupation fraction or allowing smaller NSC masses in the model (as discussed in Sect. 4.2). Moving to the low-mass regime (M<105.5Msubscript𝑀superscript105.5subscriptMdirect-productM_{\bullet}<10^{5.5}\,\mathrm{M_{\odot}}italic_M start_POSTSUBSCRIPT ∙ end_POSTSUBSCRIPT < 10 start_POSTSUPERSCRIPT 5.5 end_POSTSUPERSCRIPT roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT), we find that the turnover is due to lower average NSC masses and greater ages around small MBHs (which translates into significantly reduced TDE rates by z= 0𝑧 0z\,{=}\,0italic_z = 0).

Our results are also broadly in agreement with the volumetric rates as a function of MBH derived from the eROSITA X-ray luminosity function assuming an Eddington-limited accretion and a simple volumetric correction151515computed simply as M[M]=kbolLX/Ledd,1subscript𝑀delimited-[]subscript𝑀direct-productsubscript𝑘bolsubscript𝐿𝑋subscript𝐿edd1M_{\bullet}[M_{\odot}]\,{=}\,k_{\rm bol}\,L_{X}\,{/}\,L_{\rm edd,1}italic_M start_POSTSUBSCRIPT ∙ end_POSTSUBSCRIPT [ italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT ] = italic_k start_POSTSUBSCRIPT roman_bol end_POSTSUBSCRIPT italic_L start_POSTSUBSCRIPT italic_X end_POSTSUBSCRIPT / italic_L start_POSTSUBSCRIPT roman_edd , 1 end_POSTSUBSCRIPT where Ledd,1subscript𝐿edd1L_{\rm edd,1}italic_L start_POSTSUBSCRIPT roman_edd , 1 end_POSTSUBSCRIPT the Eddington luminosity for a solar-mass compact object and LXsubscript𝐿𝑋L_{X}italic_L start_POSTSUBSCRIPT italic_X end_POSTSUBSCRIPT the soft X-ray luminosity of kbol=15subscript𝑘bol15k_{\rm bol}=15italic_k start_POSTSUBSCRIPT roman_bol end_POSTSUBSCRIPT = 15. The two classes of optical-UV and X-ray TDEs might not be eventually different as inferred from recent studies (Guolo et al. 2024), so we anticipate the two mass functions to be similar (which is the case within errors with the simple transformation we performed). Finally, in the left panel of Fig. 5 we also plot the previous constraints from van Velzen (2018) for reference. These are higher than both the model predictions and the Yao et al. (2023) constraints at all MBH masses below event-horizon suppression. This could be due to the smaller number statistics (17 TDEs) and the heterogeneous nature of the van Velzen (2018) sample (see discussion for the low-end of the luminosity function, Yao et al. 2023).

In the right panel of Fig. 5 we present again the model predictions of the volumetric rates, but now as a function of the host-galaxy stellar mass, Msubscript𝑀M_{\ast}italic_M start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT. As shown, our predictions at z= 0𝑧 0z\,{=}\,0italic_z = 0 display a broad agreement with the shape of the distribution from Yao et al. (2023).

To describe the distribution of observed TDEs, Yao et al. (2023) derive a model using the galaxy stellar mass function coupled with a power-law dependence of the rates on galaxy stellar mass (M0.4superscriptsubscript𝑀0.4M_{\ast}^{-0.4}italic_M start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 0.4 end_POSTSUPERSCRIPT, originally from the rate dependence on MBH mass MBsuperscriptsubscript𝑀𝐵M_{\bullet}^{B}italic_M start_POSTSUBSCRIPT ∙ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_B end_POSTSUPERSCRIPT, with B=0.25𝐵0.25B=-0.25italic_B = - 0.25). A similar functional form has been used previously, e.g., for various local galaxies (Wang & Merritt 2004), and for galaxy central cores and cusps (Stone & Metzger 2016). In the range M=109.5 1010.5Msubscript𝑀superscript109.5superscript1010.5subscriptMdirect-productM_{\ast}=10^{9.5}\,{-}\,10^{10.5}\,\mathrm{M_{\odot}}italic_M start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT = 10 start_POSTSUPERSCRIPT 9.5 end_POSTSUPERSCRIPT - 10 start_POSTSUPERSCRIPT 10.5 end_POSTSUPERSCRIPT roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT our model prediction can be fitted with a similar power-law, but we predict a different functional form at the low mass end.

Refer to caption
Refer to caption
Figure 6: Average TDE rates per log mass of MBH Msubscript𝑀M_{\bullet}italic_M start_POSTSUBSCRIPT ∙ end_POSTSUBSCRIPT (left) and of MBH-host galaxy Msubscript𝑀M_{\ast}italic_M start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT (right) for the fiducial model at z= 0.0𝑧0.0z\,{=}\,0.0italic_z = 0.0 (solid red line, tagged as “all”). We average separately over NSC rates (light-blue lines) and galaxy component rates (maroon lines), and subsequently splitting to just restarted (young systems with ΔtTD< 30MyrΔsubscript𝑡TD30Myr\Delta t_{\rm TD}\,{<}\,30\,\mathrm{Myr}roman_Δ italic_t start_POSTSUBSCRIPT roman_TD end_POSTSUBSCRIPT < 30 roman_Myr, dashed line) and after a long time (old systems with ΔtTD> 3GyrΔsubscript𝑡TD3Gyr\Delta t_{\rm TD}\,{>}\,3\,\mathrm{Gyr}roman_Δ italic_t start_POSTSUBSCRIPT roman_TD end_POSTSUBSCRIPT > 3 roman_Gyr, dotted lines). For comparison, we present the results of (Pfister et al. 2020) for three different scaling relation pairs of MBH-galaxy stellar mass & NSC mass-size adopted in their work (DD, DP, and KD as defined in Table 2 of Pfister et al. 2020) to bracket the uncertainties following these hypotheses.

3.2 Per-galaxy TDE rates

In Fig. 6, we show the per-galaxy TDE rates as a function of MBH mass (left panel) and stellar mass of galaxies hosting an MBH (right panel). We further split the contributions to the total TDE rates from NSCs and the galaxy component161616As mentioned in the model description, the galaxy component contributing to TDEs is assumed to be the bulge, or the disk in the absence of bulge., further subdivided into old (ΔtTD> 3GyrΔsubscript𝑡TD3Gyr\Delta t_{\rm TD}\,{>}\,3\,\mathrm{Gyr}roman_Δ italic_t start_POSTSUBSCRIPT roman_TD end_POSTSUBSCRIPT > 3 roman_Gyr) and young (ΔtTD< 30MyrΔsubscript𝑡TD30Myr\Delta t_{\rm TD}\,{<}\,30\,\mathrm{Myr}roman_Δ italic_t start_POSTSUBSCRIPT roman_TD end_POSTSUBSCRIPT < 30 roman_Myr) systems. We immediately see that the contribution of NSCs to the total rates dominates across all masses. The rates from the galaxy component increase both with MBH and galaxy stellar mass up to the event-horizon suppression mass. However, they are on average more than three orders of magnitude lower when compared to the contribution from the NSC at fixed MBH/host galaxy stellar mass. This result has important implications for the expected nucleation fraction of galaxies, which we will further discuss in Sect. 4.2. Indeed, based on these results, TDEs should be predominately observed in the presence of NSCs. Current observations of TDEs are primarily from distant enough sources where NSCs remain unresolved.

When looking at the difference between old and young systems, we see a different behavior for the NSCs and the galaxy contributions. Recently formed NSCs give the highest contributions to the per-galaxy rates, with approximately constant values of  105yr1similar-toabsentsuperscript105superscriptyr1{\sim}\,10^{-5}\,\mathrm{yr}^{-1}∼ 10 start_POSTSUPERSCRIPT - 5 end_POSTSUPERSCRIPT roman_yr start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT per MBH/galaxy (see the bottom panel of Fig. 3, where we showed that M=104107Msubscript𝑀superscript104superscript107subscriptMdirect-productM_{\bullet}=10^{4}-10^{7}\,\mathrm{M_{\odot}}italic_M start_POSTSUBSCRIPT ∙ end_POSTSUBSCRIPT = 10 start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT - 10 start_POSTSUPERSCRIPT 7 end_POSTSUPERSCRIPT roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT have similarly high rates at t=107𝑡superscript107t=10^{7}italic_t = 10 start_POSTSUPERSCRIPT 7 end_POSTSUPERSCRIPTyr). Instead, old systems below M< 107Msubscript𝑀superscript107subscriptMdirect-productM_{\bullet}\,{<}\,10^{7}\,\mathrm{M_{\odot}}italic_M start_POSTSUBSCRIPT ∙ end_POSTSUBSCRIPT < 10 start_POSTSUPERSCRIPT 7 end_POSTSUPERSCRIPT roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT exhibit a power-law dependence on MBH mass, namely N˙perMBHMBproportional-tosubscript˙𝑁perMBHsuperscriptsubscript𝑀𝐵\dot{N}_{\rm per-MBH}\,{\propto}\,M_{\bullet}^{B}over˙ start_ARG italic_N end_ARG start_POSTSUBSCRIPT roman_per - roman_MBH end_POSTSUBSCRIPT ∝ italic_M start_POSTSUBSCRIPT ∙ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_B end_POSTSUPERSCRIPT with B= 1𝐵1B\,{=}\,1italic_B = 1 (compared to B= 0𝐵 0B\,{=}\,0italic_B = 0 for young systems under the same definition). The old NSCs have on average small mass compared to their MBH. For these systems, the relaxation time becomes significantly short, thus the loss cone gets quickly depleted. In the low mass regime (M<105Msubscript𝑀superscript105subscriptMdirect-productM_{\bullet}<10^{5}\,\mathrm{M_{\odot}}italic_M start_POSTSUBSCRIPT ∙ end_POSTSUBSCRIPT < 10 start_POSTSUPERSCRIPT 5 end_POSTSUPERSCRIPT roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT), the majority of systems have rates from old rather than young NSCs and that is why the global per-MBH rates drop towards lower masses. This has specific implications for the dwarf galaxy regime (yet to be probed with observations), where per-MBH/per-galaxy TDE rates are higher by a factor of >>>10 for young NSCs over old NSCs. Therefore, our work suggests that a promising opportunity of sampling TDEs is selecting recently interacting systems, with a predicted per-galaxy (volumetric) rate of  1× 105similar-toabsent1superscript105{\sim}\,1\,{\times}\,10^{-5}∼ 1 × 10 start_POSTSUPERSCRIPT - 5 end_POSTSUPERSCRIPTyr-1( 1×107similar-toabsent1superscript107{\sim}\,1\times 10^{-7}∼ 1 × 10 start_POSTSUPERSCRIPT - 7 end_POSTSUPERSCRIPTyr-1Mpc-3dex-1). This prediction holds down to the least massive dwarfs, as marked by the flattening of the rates as a function of galaxy stellar mass (for both per-galaxy and volumetric TDE rates, see respectively the right panels of Figs. 5 and 6). This revision of rates in dwarf galaxies may have a significant impact on theoretical predictions of TDEs in intermediate-mass MBHs. The galaxy contribution shows a different behavior with age. First of all, we note again that TDEs from the galaxy component do not strongly depend on time and that bulges lead to larger rates compared to pure disks (see discussion of Fig. 3 in Sect. 2.2). In Fig. 6, we see that old systems have higher per-galaxy rates with respect to the young galactic components at the lower MBH M< 106.5Msubscript𝑀superscript106.5subscriptMdirect-productM_{\bullet}\,{<}\,10^{6.5}\,\mathrm{M_{\odot}}italic_M start_POSTSUBSCRIPT ∙ end_POSTSUBSCRIPT < 10 start_POSTSUPERSCRIPT 6.5 end_POSTSUPERSCRIPT roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT and galaxy M< 108Msubscript𝑀superscript108subscriptMdirect-productM_{\ast}\,{<}\,10^{8}\,\mathrm{M_{\odot}}italic_M start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT < 10 start_POSTSUPERSCRIPT 8 end_POSTSUPERSCRIPT roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT masses. We attribute this to the morphological differences between old and young galaxies in this mass range: L-GalaxiesBH predicts that young systems are mostly pure disks (with an average bulge-to-total mass ratio of B/T=0.075delimited-⟨⟩𝐵𝑇0.075\langle B/T\rangle=0.075⟨ italic_B / italic_T ⟩ = 0.075), while old systems are preferentially bulge-dominated (B/T=0.3delimited-⟨⟩𝐵𝑇0.3\langle B/T\rangle=0.3⟨ italic_B / italic_T ⟩ = 0.3). At the high mass end, instead, we see that young galaxies have higher per-galaxy rates than old ones as a function of galaxy stellar mass. In this case, the difference can be attributed to the differences in the MMgalsubscript𝑀subscript𝑀galM_{\bullet}-M_{\rm gal}italic_M start_POSTSUBSCRIPT ∙ end_POSTSUBSCRIPT - italic_M start_POSTSUBSCRIPT roman_gal end_POSTSUBSCRIPT ratio for young and old systems: at fixed galaxy stellar mass as an effect of early on-set of MBH growth, old systems are preferentially hosting heavier MBHs, whose galaxy-component rates are lower. In both cases the per-galaxy rates follow a power-law of the form N˙per-MBHMBproportional-tosubscript˙𝑁per-MBHsuperscriptsubscript𝑀𝐵\dot{N}_{\text{per-MBH}}\,{\propto}\,M_{\bullet}^{B}over˙ start_ARG italic_N end_ARG start_POSTSUBSCRIPT per-MBH end_POSTSUBSCRIPT ∝ italic_M start_POSTSUBSCRIPT ∙ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_B end_POSTSUPERSCRIPT with B0.7𝐵0.7B\approx 0.7italic_B ≈ 0.7 for the mass range M=106108Msubscript𝑀superscript106superscript108subscriptMdirect-productM_{\bullet}=10^{6}-10^{8}\,\mathrm{M_{\odot}}italic_M start_POSTSUBSCRIPT ∙ end_POSTSUBSCRIPT = 10 start_POSTSUPERSCRIPT 6 end_POSTSUPERSCRIPT - 10 start_POSTSUPERSCRIPT 8 end_POSTSUPERSCRIPT roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT and M=1081011Msubscript𝑀superscript108superscript1011subscriptMdirect-productM_{\ast}=10^{8}-10^{11}\,\mathrm{M_{\odot}}italic_M start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT = 10 start_POSTSUPERSCRIPT 8 end_POSTSUPERSCRIPT - 10 start_POSTSUPERSCRIPT 11 end_POSTSUPERSCRIPT roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT.

Our model predictions in Fig. 6 are then compared with the ones of Pfister et al. (2020) which, motivated by the computation of TDE rates in observed galaxy profiles, computed the per-galaxy TDE rates for a mock catalog of galaxies hosting NSCs and with a well-characterized bulge component. Before making a comparison, it is worth noticing that Pfister et al. (2020) derives MBH masses using, also in the low-mass regime, MMsubscript𝑀subscript𝑀M_{\bullet}-M_{\rm*}italic_M start_POSTSUBSCRIPT ∙ end_POSTSUBSCRIPT - italic_M start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT scaling relations observationally derived for massive spheroidal galaxies (Kormendy & Ho 2013; Davis et al. 2019). Moreover, although they use PhaseFlow to compute the TDE rates, they assume a monochromatic star distribution of m= 1Msubscript𝑚1subscriptMdirect-productm_{\ast}\,=\,1\,\mathrm{M_{\odot}}italic_m start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT = 1 roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT, while this work uses m= 0.38Msubscript𝑚0.38subscriptMdirect-productm_{\ast}\,=\,0.38\,\mathrm{M_{\odot}}italic_m start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT = 0.38 roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT and stellar black holes (16M16subscriptMdirect-product16\,\mathrm{M_{\odot}}16 roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT), a difference that can change by a factor of two the number of stars entering the loss cone, thus available for TDEs (Stone & Metzger 2016).

The per-galaxy rates for NSCs (solid red lines) are consistent with the predictions of Pfister et al. (2020) for M> 106.5Msubscript𝑀superscript106.5subscriptMdirect-productM_{\bullet}\,{>}\,10^{6.5}\,\mathrm{M_{\odot}}italic_M start_POSTSUBSCRIPT ∙ end_POSTSUBSCRIPT > 10 start_POSTSUPERSCRIPT 6.5 end_POSTSUPERSCRIPT roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT and M> 1010Msubscript𝑀superscript1010subscriptMdirect-productM_{\ast}\,{>}\,10^{10}\,\mathrm{M_{\odot}}italic_M start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT > 10 start_POSTSUPERSCRIPT 10 end_POSTSUPERSCRIPT roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT but have a different trend (they diverge) at lower masses. We suspect that the main difference to the rates towards lower masses is the use of steep Sérsic profiles nNSCsubscript𝑛NSCn_{\rm NSC}italic_n start_POSTSUBSCRIPT roman_NSC end_POSTSUBSCRIPT for low mass NSCs, using the scaling relation of Pechetti et al. (2020) (for MNSC< 106Msubscript𝑀NSCsuperscript106subscriptMdirect-productM_{\rm NSC}\,{<}\,10^{6}\,\mathrm{M_{\odot}}italic_M start_POSTSUBSCRIPT roman_NSC end_POSTSUBSCRIPT < 10 start_POSTSUPERSCRIPT 6 end_POSTSUPERSCRIPT roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT the relation gives Sérsic indices with values above 4). This anti-correlation between NSC mass and Sérsic index for lower masses appears to be weaker in the recent results of Hoyer et al. (2023). We instead assume shallower, possibly more realistic, profiles, as described in Sect. 2.2 and no correlation of the steepness of the profile with mass. Moreover, Pfister et al. (2020) do not allow for small structures with M>MNSCsubscript𝑀subscript𝑀NSCM_{\bullet}\,{>}\,M_{\rm NSC}italic_M start_POSTSUBSCRIPT ∙ end_POSTSUBSCRIPT > italic_M start_POSTSUBSCRIPT roman_NSC end_POSTSUBSCRIPT, while there is evidence of such objects and they are naturally included in our model (see Sect. 4.4 for the impact of this parameter).

Regarding the TDE rates from the galaxy component, Pfister et al. (2020) also finds an increase both with MBH and galaxy stellar mass until the event-horizon suppression mass. Nevertheless, the average rate they predict is three to four orders of magnitude higher than our predictions for all galaxy stellar mass ranges. This is probably because Pfister et al. (2020) assumes on average more centrally concentrated galaxies compared to our work. Also, as mentioned earlier, at the low galaxy stellar mass end, they assign MBH masses using the same steep relation as for a more massive system: this implies that, at fixed MBH mass, their stellar mass can be up to 2dex2dex2\,\mathrm{dex}2 roman_dex larger than ours.
The results just discussed highlight that the demographical analysis of TDE rates requires a realistic cosmological environment with carefully constructed morphological galaxy properties and number distributions as a function of galaxy stellar mass, that L-GalaxiesBH provides.

3.2.1 Redshift evolution of TDE rates

Having captured the z= 0𝑧 0z\,{=}\,0italic_z = 0 global rates, we now discuss their redshift evolution and how that compares with the evolution of the underlying MBH properties and their environment.

Refer to caption
Figure 7: Redshift evolution [z= 0𝑧 0z\,{=}\,0italic_z = 0 (top), z= 1𝑧1z\,{=}\,1italic_z = 1 (middle), z= 2𝑧2z\,{=}\,2italic_z = 2 (bottom)] of the volumetric TDE rates per MBH (left) and stellar (right) mass originating from all low-luminosity or inactive MBHs (cut-off at log10LAGN[erg/s]<42subscript10subscript𝐿AGNdelimited-[]ergs42\log_{10}L_{\rm AGN}\,{\rm[erg/s]}<42roman_log start_POSTSUBSCRIPT 10 end_POSTSUBSCRIPT italic_L start_POSTSUBSCRIPT roman_AGN end_POSTSUBSCRIPT [ roman_erg / roman_s ] < 42 for the fiducial model, solid lines) and TDE rates from the galaxy components only (bulges/disks; solid lines tagged as “gal”). We also display the rates only from AGN-hosts with high (purple dashed-dotted), moderate (green dashed), and low (blue dotted) luminosity (log10LAGN[erg/s][42,45]subscript10subscript𝐿AGNdelimited-[]ergs4245\log_{10}L_{\rm AGN}\,{\rm[erg/s]}\in[42,45]roman_log start_POSTSUBSCRIPT 10 end_POSTSUBSCRIPT italic_L start_POSTSUBSCRIPT roman_AGN end_POSTSUBSCRIPT [ roman_erg / roman_s ] ∈ [ 42 , 45 ], [40.5,42]40.542[40.5,42][ 40.5 , 42 ] and [39,40.5]3940.5[39,40.5][ 39 , 40.5 ] respectively). The z= 0𝑧 0z\,{=}\,0italic_z = 0 constraints Yao et al. (2023) are plotted in all panels for reference (with grey when they do not apply). These fractions remain qualitatively the same for these redshifts for most of the parameter variations.

In Fig. 7 we display the volumetric TDE rates for redshifts z= 0,1,2𝑧 012z\,=\,0,1,2italic_z = 0 , 1 , 2. Regarding the case of TDE rates as a function of MBH mass, there does not seem to be a significant change in the rates between z= 0𝑧 0z\,{=}\,0italic_z = 0 and z= 1𝑧1z\,{=}\,1italic_z = 1. This is due to the very mild evolution evolution predicted by the model for z< 1𝑧1z\,{<}\,1italic_z < 1. This result is in line with the general assumptions that the volumetric TDE rates do not evolve significantly with redshift (see e.g.van Velzen & Farrar 2014; van Velzen 2018; Yao et al. 2023 see, however, the approach of Kochanek 2016). Despite this, there is a slight decrease of half an order of magnitude when comparing resolved TDEs at z= 0𝑧 0z\,{=}\,0italic_z = 0, 1111 and z= 2𝑧2z\,{=}\,2italic_z = 2, which will be an important feature to study with deep sky surveys such as LSST (Hambleton et al. 2023; Bricman & Gomboc 2020; Bučar Bricman et al. 2023) and UV transient surveys like QUVIK (Zajaček et al. 2024). However, it is worth mentioning that our predictions are model-dependent. We assume that MBHs are always associated to an NSC at all redshifts (see Sect. 2.1). If this was not true at early times, high-redshift TDE rates would be lower.

The volumetric rates shown in Fig. 7 can be divided into contributions from the galaxy component (“gal”) and the NSC. The first, although it steadily increases from z= 2𝑧2z\,{=}\,2italic_z = 2 to z= 0𝑧 0z\,{=}\,0italic_z = 0 with the increase of bulges in galaxies, still provides a negligible contribution to the volumetric TDE rates at all masses and all redshifts. In practice, the volumetric rates are entirely due to galaxies hosting NSCs (this contribution is not shown in the figure as it essentially coincides with the total “fiducial” volumetric TDE rates). While the per-galaxy rates for NSCs remain the same at high masses towards higher redshifts, there is a slight enhancement (factor less than three) compared to z= 0𝑧 0z\,{=}\,0italic_z = 0 for M< 106Msubscript𝑀superscript106subscriptMdirect-productM_{\bullet}\,<\,10^{6}\,\mathrm{M_{\odot}}italic_M start_POSTSUBSCRIPT ∙ end_POSTSUBSCRIPT < 10 start_POSTSUPERSCRIPT 6 end_POSTSUPERSCRIPT roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT MBH and M< 109Msubscript𝑀superscript109subscriptMdirect-productM_{\ast}\,<\,10^{9}\,\mathrm{M_{\odot}}italic_M start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT < 10 start_POSTSUPERSCRIPT 9 end_POSTSUPERSCRIPT roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT host-galaxy stellar masses, because of continuous reset of TDE rates during frequent mergers at cosmic noon. This somewhat counteracts the decrease of the number density of inactive MBHs towards higher redshifts and results in a moderate evolution of the volumetric TDE rates.

3.3 TDE rates in active MBHs

Searches for TDEs in AGN at optical (Dgany et al. 2023, ZTF) and X-ray (Homan et al. 2023, eROSITA) wavelengths have not yielded many representative cases. However, observations of luminous ambiguous nuclear transients (Frederick et al. 2021; Hinkle et al. 2022; Oates et al. 2024), TDEs and TDE-like flares in ongoing/previous AGN (Li et al. 2023; Huang et al. 2023; Makrygianni et al. 2023; Charalampopoulos et al. 2024), and changing-look AGN (Ricci & Trakhtenbrot 2023; Zhang 2022) hint that the two TDE and AGN activity could be related. So far, few theoretical works (Karas & Šubr 2007; Chan et al. 2019; McKernan et al. 2022; Prasad et al. 2024) and recent simulations (Ryu et al. 2023b) have explored the co-existence of TDEs and AGN. Predicting TDEs in AGN hosts is beyond the scope of this paper, as it requires dedicated modeling of the electromagnetic emission of individual events, considering the relative brightness of TDEs and their host. This section overviews TDE occurrence in AGN under the assumptions outlined in Sect. 2.3.3, aiding future searches for transient events in AGN.

In Fig. 7 we show the volumetric global rates at z= 0𝑧 0z\,{=}\,0italic_z = 0, 1111 and 2222, for MBHs triggering AGN with low, moderate and high bolometric luminosity bins: log10(Lbol[erg/s])[39.0,40.5]subscript10subscript𝐿boldelimited-[]ergs39.040.5\log_{10}\,(L_{\rm bol}\,[\mathrm{erg/s}])\in[39.0,40.5]roman_log start_POSTSUBSCRIPT 10 end_POSTSUBSCRIPT ( italic_L start_POSTSUBSCRIPT roman_bol end_POSTSUBSCRIPT [ roman_erg / roman_s ] ) ∈ [ 39.0 , 40.5 ], [40.5,42.0]40.542.0[40.5,42.0][ 40.5 , 42.0 ] and [42.0,45.0]42.045.0[42.0,45.0][ 42.0 , 45.0 ]. Note that for high AGN luminosity, the detectable rate of events will depend on the luminosity of individual TDEs. In other words, only TDEs with peak luminosity significantly higher than the AGN luminosity at a given wavelength would allow distinguishing the characteristic time-decay of a TDE light curve, thus singling out the event from AGN stochastic variability or unrelated flaring activity. The first thing to notice in Fig. 7 is that the strong redshift evolution of the AGN luminosity function leaves an imprint on the AGN hosting TDEs. Indeed, the lower the redshift, the greater the importance of inactive MBHs in the volumetric TDE rates. This is because a large fraction of massive black holes consume their high-z𝑧zitalic_z gas reservoir and end up in an inactive phase in the low-z𝑧zitalic_z Universe.

We observe that at high MBH and host-galaxy stellar masses (M> 108Msubscript𝑀superscript108subscriptMdirect-productM_{\bullet}\,{>}\,10^{8}\,\mathrm{M_{\odot}}italic_M start_POSTSUBSCRIPT ∙ end_POSTSUBSCRIPT > 10 start_POSTSUPERSCRIPT 8 end_POSTSUPERSCRIPT roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT and M> 1010.5Msubscript𝑀superscript1010.5subscriptMdirect-productM_{\ast}\,{>}\,10^{10.5}\,\mathrm{M_{\odot}}italic_M start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT > 10 start_POSTSUPERSCRIPT 10.5 end_POSTSUPERSCRIPT roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT) and at all redshifts, the volumetric rate of TDEs occurring in luminous AGN (all luminosity bins) is comparable to the one occurring in inactive or low activity AGN (fiducial) curve and almost redshift-independent at Gpc3yr1dex1superscriptGpc3superscriptyr1superscriptdex1\mathrm{Gpc}^{-3}\,\mathrm{yr}^{-1}\,\mathrm{dex}^{-1}roman_Gpc start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT roman_yr start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT roman_dex start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT. Specifically for redshift z= 2𝑧2z\,{=}\,2italic_z = 2, the TDE rates of luminous AGN (high) dominate over the rates of low-activity MBHs at host-galaxy stellar masses M>108Msubscript𝑀superscript108subscriptMdirect-productM_{\ast}\,{>}10^{8}\,\mathrm{M_{\odot}}italic_M start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT > 10 start_POSTSUPERSCRIPT 8 end_POSTSUPERSCRIPT roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT and M 106.5Msimilar-tosubscript𝑀superscript106.5subscriptMdirect-productM_{\bullet}\,\sim\,10^{6.5}\,\mathrm{M_{\odot}}italic_M start_POSTSUBSCRIPT ∙ end_POSTSUBSCRIPT ∼ 10 start_POSTSUPERSCRIPT 6.5 end_POSTSUPERSCRIPT roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT MBHs. Therefore monitoring flares in AGN at z> 1𝑧1z\,{>}\,1italic_z > 1 might be an effective way to identify TDEs, assuming that a significant fraction of TDEs will be brighter than the AGN emission itself.

4 Discussion

Here we address the implications of the inclusion of TDEs for the modeling of MBHs with L-GalaxiesBH, and explore how the results presented in this work are affected by the parameters and assumptions of our model (Section 2).

4.1 MBH growth via Tidal Disruption Events

As mentioned in the Introduction, the role of TDEs in the mass content of MBHs has been so far theoretical, with no means to calibrate this mechanism at analytical prescriptions of this MBH growth channel. Since our work already is in broad agreement with the TDE rates at z= 0𝑧 0z\,{=}\,0italic_z = 0, we can draw some first conclusions on the contribution of TDE rates to the growth of the MBH population. Here we limit the discussion to the impact of TDE on the growth of the MBH population by z= 0𝑧 0z\,{=}\,0italic_z = 0. Another important aspect regarding growth is the time resolution; according to Broggi et al. (2022) MBHs residing in NSCs will double their mass within a characteristic timescale that depends on the MBH mass (in their set-up, they have a power-law dependence 1.291.291.291.29) and falls below time resolution for MBHs M<105Msubscript𝑀superscript105subscriptMdirect-productM_{\bullet}<10^{5}\,\mathrm{M_{\odot}}italic_M start_POSTSUBSCRIPT ∙ end_POSTSUBSCRIPT < 10 start_POSTSUPERSCRIPT 5 end_POSTSUPERSCRIPT roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT. We include the prompt phase by integrating the rates taking place below the time-resolution dtstep𝑑subscript𝑡stepdt_{\rm step}italic_d italic_t start_POSTSUBSCRIPT roman_step end_POSTSUBSCRIPT. A more detailed investigation of the conditions for efficient growth via TDEs and of the sub-categories of galaxy types where and when this growth channel is important will be the subject of follow-up work (Polkas et al. in prep).

In Fig. 8, we show the fractional cumulative mass content of MBHs, namely fBHAMsubscript𝑓BHAMf_{\rm BHAM}italic_f start_POSTSUBSCRIPT roman_BHAM end_POSTSUBSCRIPT171717Here, the definition of the total black hole accreted mass (BHAM) includes the mass content in seed mass, although strictly speaking, this is not accreted onto the MBH. at z= 0𝑧 0z\,{=}\,0italic_z = 0 as a function of their mass, for the runs with and without prompt phase. As demonstrated in these plots, the mass growth through TDEs is mostly insignificant for the high mass ranges (i.e. fBHAM<subscript𝑓BHAMabsentf_{\rm BHAM}\,<\,italic_f start_POSTSUBSCRIPT roman_BHAM end_POSTSUBSCRIPT <1% for M> 105Msubscript𝑀superscript105subscriptMdirect-productM_{\bullet}\,{>}\,10^{5}\,\mathrm{M_{\odot}}italic_M start_POSTSUBSCRIPT ∙ end_POSTSUBSCRIPT > 10 start_POSTSUPERSCRIPT 5 end_POSTSUPERSCRIPT roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT), especially when compared to the mass growth induced by cold gas accretion. However, for MBHs that remain close to their seed mass M= 102.5 105Msubscript𝑀superscript102.5superscript105subscriptMdirect-productM_{\bullet}\,=\,10^{2.5}\,{-}\,10^{5}\,\mathrm{M_{\odot}}italic_M start_POSTSUBSCRIPT ∙ end_POSTSUBSCRIPT = 10 start_POSTSUPERSCRIPT 2.5 end_POSTSUPERSCRIPT - 10 start_POSTSUPERSCRIPT 5 end_POSTSUPERSCRIPT roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT, TDEs offer a competitive channel of growth, composing fBHAM= 110%subscript𝑓BHAM1percent10f_{\rm BHAM}\,=\,1-10\%italic_f start_POSTSUBSCRIPT roman_BHAM end_POSTSUBSCRIPT = 1 - 10 % of their mass. Notably, for the lightest mass bin M< 103Msubscript𝑀superscript103subscriptMdirect-productM_{\bullet}\,{<}\,10^{3}\,\mathrm{M_{\odot}}italic_M start_POSTSUBSCRIPT ∙ end_POSTSUBSCRIPT < 10 start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT, TDE growth is as important as the cold gas accretion, the latter being less significant (if at all existent) in dwarf galaxies (see Fig. 7 and Fig. 10 of Izquierdo-Villalba et al. 2019; Spinoso et al. 2023, respectively). Regarding this result, we stress however the theoretical uncertainties regarding the position and residence time of MBH seeds in the centers of galaxies and NSCs that may diverge from the physics included in L-GalaxiesBH regarding more massive MBHs.

For the relative massive MBHs at z= 0𝑧 0z\,{=}\,0italic_z = 0 (M> 106subscript𝑀superscript106M_{\bullet}\,{>}\,10^{6}italic_M start_POSTSUBSCRIPT ∙ end_POSTSUBSCRIPT > 10 start_POSTSUPERSCRIPT 6 end_POSTSUPERSCRIPTM), their seeds grow through gas accretion to M=104 106Msubscript𝑀superscript104superscript106subscriptMdirect-productM_{\bullet}=10^{4}\,{-}\,10^{6}\,\mathrm{M_{\odot}}italic_M start_POSTSUBSCRIPT ∙ end_POSTSUBSCRIPT = 10 start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT - 10 start_POSTSUPERSCRIPT 6 end_POSTSUPERSCRIPT roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT by z= 23𝑧23z\,{=}\,2-3italic_z = 2 - 3, allowing for important cumulative TDE growth (after the initial gas growth) until z= 0𝑧 0z\,{=}\,0italic_z = 0. This is shown by the excess which brings the star-accretion curve over the seed mass one (for MBHs with M= 104.5 107.5Msubscript𝑀superscript104.5superscript107.5subscriptMdirect-productM_{\bullet}\,{=}\,10^{4.5}\,-\,10^{7.5}\,\mathrm{M_{\odot}}italic_M start_POSTSUBSCRIPT ∙ end_POSTSUBSCRIPT = 10 start_POSTSUPERSCRIPT 4.5 end_POSTSUPERSCRIPT - 10 start_POSTSUPERSCRIPT 7.5 end_POSTSUPERSCRIPT roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT). Both Msubscript𝑀M_{\ast}italic_M start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT and MNSCsubscript𝑀NSCM_{\rm NSC}italic_M start_POSTSUBSCRIPT roman_NSC end_POSTSUBSCRIPT will be greater for MBHs of this mass range, resulting in constant high rates with time compared to those of surviving dormant seeds living in isolation. The latter cannot maintain a constant high rate as indicated by the per-MBH rates for old NSC systems in Fig. 6. More massive MBHs (M 108Msimilar-tosubscript𝑀superscript108subscriptMdirect-productM_{\bullet}\,{\sim}\,10^{8}\,\mathrm{M_{\odot}}italic_M start_POSTSUBSCRIPT ∙ end_POSTSUBSCRIPT ∼ 10 start_POSTSUPERSCRIPT 8 end_POSTSUPERSCRIPT roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT) form early in the most massive galaxies (M 1011Msimilar-tosubscript𝑀superscript1011subscriptMdirect-productM_{\ast}\,{\sim}\,10^{11}\,\mathrm{M_{\odot}}italic_M start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT ∼ 10 start_POSTSUPERSCRIPT 11 end_POSTSUPERSCRIPT roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT) of the simulation which will stop hosting NSCs earlier and grow less their MBH through TDEs.

Refer to caption
Figure 8: Average fraction of mass accreted by different MBHs at z= 0𝑧 0z\,{=}\,0italic_z = 0 (fBHAMsubscript𝑓BHAMf_{\rm BHAM}italic_f start_POSTSUBSCRIPT roman_BHAM end_POSTSUBSCRIPT) via stellar captures (purple squares), cold gas accretion (blue right triangles), hot gas accretion (orange left triangles) and accumulated seed mass (green circles), without (top panel) and with (bottom panel) an initial prompt phase (see main text). The margin of ±1plus-or-minus1\pm 1± 1 standard deviation of log10fBHAMsubscript10subscript𝑓BHAM\log_{10}f_{\rm BHAM}roman_log start_POSTSUBSCRIPT 10 end_POSTSUBSCRIPT italic_f start_POSTSUBSCRIPT roman_BHAM end_POSTSUBSCRIPT for each type of channel is highlighted with the same-color shaded area.

Our results add realism to estimates from earlier works. Unlike claims using the full loss-cone theory (e.g. Milosavljević et al. 2006; Alexander & Bar-Or 2017), the growth via TDEs in our model is not such that all M< 104Msubscript𝑀superscript104subscriptMdirect-productM_{\bullet}\,{<}\,10^{4}\,\mathrm{M_{\odot}}italic_M start_POSTSUBSCRIPT ∙ end_POSTSUBSCRIPT < 10 start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT MBHs would reach the supermassive regime as a result of stellar accretion. We attribute this behavior to a) the vast majority of low-mass MBHs being hosted in lower-mass galaxies with less massive NSCs and b) the time-dependent nature of the rates; non-interacting ancient NSCs relax quickly and can even become less important than the bulge/disk rates (see input rates Fig. 3). As mentioned above, however, we will study in a follow-up work the specific properties of the environment and redshifts that lead to more efficient growth via TDEs.

4.2 Implications for MBH and NSC populations

Refer to caption
Figure 9: Top and middle panels: occupation fractions of MBH (M> 105Msubscript𝑀superscript105subscriptMdirect-productM_{\bullet}\,{>}\,10^{5}\,\mathrm{M_{\odot}}italic_M start_POSTSUBSCRIPT ∙ end_POSTSUBSCRIPT > 10 start_POSTSUPERSCRIPT 5 end_POSTSUPERSCRIPT roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT, top) and NSC (all masses, middle) that result into the volumetric TDE rates as a function of galaxy-host mass (bottom). We show the maxOcc and lowOcc models with orange solid and black dashed lines, respectively. In the top panels we display the local MBH occupation fraction derived from a compilation of detections Greene (2012, cyan squares) and X-ray constraints by Miller & Davies (2012, light-blue region), as well as an occupation for slightly higher redshifts 0.01<z< 0.10.01𝑧0.10.01\,{<}\,z\,{<}\,0.10.01 < italic_z < 0.1 derived via optical-line AGN (Trump et al. 2015, blue circles). In the middle panel, the data points and logistic function (the dashed-dotted thin line), both from Hoyer et al. (2021), have the same meaning as in Fig. 2. The enhanced MBH occupation is still compatible with the observations of the Coma Cluster (grey triangles), even though it yields a greater fraction of nucleated galaxies. We stress that all NSCs are hosting an MBH in our model, hence adding NSCs without MBHs would increase this fraction.Bottom panel: Both maxOcc and lowOcc model (solid and dashed line) reproduce the TDE-rate distribution constraint of Yao et al. (2023) within errors. However, in the lack of evidence of TDEs in dwarf galaxies the lowOcc model is preferred.

Complete galaxy samples are starting to constrain the AGN occupation fraction across different galaxy stellar masses, hinting at significantly smaller AGN fractions in dwarf galaxies (i.e. < 102absentsuperscript102{<}\,10^{-2}< 10 start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT) than in massive systems (see e.g. Mezcua et al. 2018; Mezcua & Domínguez Sánchez 2020; Bykov et al. 2023; Zou et al. 2023; Siudek et al. 2023). This could be explained if the number of AGN in dwarf galaxies remains small because of the lack of cold gas in the center of these systems (Urquhart et al. 2022) which cannot sustain significant MBH growth. Therefore, AGN cannot be used as good tracers of low-mass MBHs (nor of the full population of MBHs). Alternatively, the abundance of MBHs in the local universe and beyond can be unveiled by exploiting the population of TDEs in all-sky transient surveys.

As shown in Section 2.1.1 and Section 2.1 our fiducial model is in fair agreement with both the MBH mass function and the NSC occupation fraction. However, there are uncertainties at the low-mass end of the MBH/host-galaxy stellar mass function and the frequency of MBHs hosted by NSCs is currently unknown. To address these uncertainties, we compare our fiducial model to a run with the highest number of MBHs permitted by our model, hereafter maxOcc model. They differ in the amplitude of the MBH seeding probability by one order of magnitude, namely

Gp=0.1(lowOcc)vs 1.0(maxOcc),subscriptGp0.1𝑙𝑜𝑤𝑂𝑐𝑐vs1.0𝑚𝑎𝑥𝑂𝑐𝑐{\rm G_{p}}\,{=}0.1\,(lowOcc)\;{\rm vs}\;1.0\,(maxOcc),roman_G start_POSTSUBSCRIPT roman_p end_POSTSUBSCRIPT = 0.1 ( italic_l italic_o italic_w italic_O italic_c italic_c ) roman_vs 1.0 ( italic_m italic_a italic_x italic_O italic_c italic_c ) ,

resulting in different MBH occupation fractions of M> 105Msubscript𝑀superscript105subscriptMdirect-productM_{\bullet}\,{>}\,10^{5}\,\mathrm{M_{\odot}}italic_M start_POSTSUBSCRIPT ∙ end_POSTSUBSCRIPT > 10 start_POSTSUPERSCRIPT 5 end_POSTSUPERSCRIPT roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT at z= 0𝑧 0z\,{=}\,0italic_z = 0, as shown in the top panel of Fig. 9. The large occupation fraction of MBHs181818For the sake of brevity, we do not show the black hole mass function of the maxOcc model. This model tends to overpredict by a factor of three the number of MBHs in the mass range M= 106 107.5Msubscript𝑀superscript106superscript107.5subscriptMdirect-productM_{\bullet}\,{=}\,10^{6}\,{-}\,10^{7.5}\,\mathrm{M_{\odot}}italic_M start_POSTSUBSCRIPT ∙ end_POSTSUBSCRIPT = 10 start_POSTSUPERSCRIPT 6 end_POSTSUPERSCRIPT - 10 start_POSTSUPERSCRIPT 7.5 end_POSTSUPERSCRIPT roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT, compared to observational constraints at z= 0𝑧 0z\,{=}\,0italic_z = 0 from Shankar et al. (2009). Besides this, the overall shape of the mass function remains relatively similar concerning the fiducial model at high masses. in galaxies has a direct effect on the abundance of NSC. Specifically, the maxOcc and lowOcc models display respectively NSC occupation (central panel of Fig. 9) up to a factor 1.6 larger and lower than the one shown in the fiducial case (see Fig. 2).

The bottom panel of Fig. 9 shows that the changes in the MBH and NSC occupation result in appreciable differences in the observed TDE rates per galaxy. Although the maxOcc model also shows good agreement with the results of Yao et al. (2023) it lies on the upper limits of the observational constraints. Quantitatively, both of the models suggest: p1less-than-or-similar-to𝑝1p\,{\lesssim}\,-1italic_p ≲ - 1 for dn/dlog10MMpproportional-to𝑑subscript𝑛𝑑subscript10subscript𝑀superscriptsubscript𝑀𝑝dn_{\bullet}\,{/}\,d\log_{10}M_{\bullet}\,{\propto}\,M_{\bullet}^{p}italic_d italic_n start_POSTSUBSCRIPT ∙ end_POSTSUBSCRIPT / italic_d roman_log start_POSTSUBSCRIPT 10 end_POSTSUBSCRIPT italic_M start_POSTSUBSCRIPT ∙ end_POSTSUBSCRIPT ∝ italic_M start_POSTSUBSCRIPT ∙ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_p end_POSTSUPERSCRIPT, instead of p0greater-than-or-equivalent-to𝑝0p\gtrsim 0italic_p ≳ 0, as found by Yao et al. 2023. Also, both the maxOcc and fiducial models support the interpretation of van Velzen (2018) for a flat (100%similar-toabsentpercent100\sim\!100\%∼ 100 %) occupation fraction of MBHs down to M 1010Msimilar-tosubscriptMsuperscript1010subscriptMdirect-product\rm M_{\ast}\,{\sim}\,10^{10}\,\mathrm{M_{\odot}}roman_M start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT ∼ 10 start_POSTSUPERSCRIPT 10 end_POSTSUPERSCRIPT roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT, while lowOcc model does not (see middle panel of Fig. 9). The noticable difference in the dwarf galaxy regime (M=107 1010Msubscript𝑀superscript107superscript1010subscriptMdirect-productM_{\ast}=10^{7}\,-\,10^{10}\,\mathrm{M_{\odot}}italic_M start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT = 10 start_POSTSUPERSCRIPT 7 end_POSTSUPERSCRIPT - 10 start_POSTSUPERSCRIPT 10 end_POSTSUPERSCRIPT roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT) is the main driver in boosting the volumetric TDEs at all galaxy stellar masses. Conversely, the fact that we do not detect TDEs at M<109Msubscript𝑀superscript109subscriptMdirect-productM_{*}<10^{9}\,\mathrm{M_{\odot}}italic_M start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT < 10 start_POSTSUPERSCRIPT 9 end_POSTSUPERSCRIPT roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT may be primarily interpreted as a lower occupation fraction of MBHs than the one predicted by L-GalaxiesBH, indicative of a less efficient seeding mechanism.

Regarding the NSC occupation, while the fiducial model (see Fig. 2) matches the observations of Virgo (and Fornax at lower masses) clusters (Muñoz et al. 2015; Eigenthaler et al. 2018; Sánchez-Janssen et al. 2019), the maxOcc resembles the higher NSC occupation in the Coma cluster (den Brok et al. 2014; Zanatta et al. 2021) and the lowOcc that of the Local Volume. Most of the clusters in our model are less massive than their MBH mass (see Appendix A for mass distribution of NSCs). In particular, only 30%and 60%absentpercent30andpercent60\approx 30\%\;{\rm and}\;60\%≈ 30 % roman_and 60 % of MBHs, hosted in galaxies with mass M 107Mand 109.5Msimilar-tosubscript𝑀superscript107subscriptMdirect-productandsuperscript109.5subscriptMdirect-productM_{\ast}\,{\sim}\,10^{7}\,\mathrm{M_{\odot}}\;{\rm and}\;10^{9.5}\,\mathrm{M_{% \odot}}italic_M start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT ∼ 10 start_POSTSUPERSCRIPT 7 end_POSTSUPERSCRIPT roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT roman_and 10 start_POSTSUPERSCRIPT 9.5 end_POSTSUPERSCRIPT roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT at z= 0𝑧 0z\,{=}\,0italic_z = 0, are also found in an NSC more massive than them, namely 𝒟=MNSC/M> 1𝒟subscript𝑀NSCsubscript𝑀1\mathcal{D}=M_{\rm NSC}/M_{\bullet}\,{>}\,1caligraphic_D = italic_M start_POSTSUBSCRIPT roman_NSC end_POSTSUBSCRIPT / italic_M start_POSTSUBSCRIPT ∙ end_POSTSUBSCRIPT > 1. For reference, the median mass ratio between NSCs and their associated MBHs with a measured mass is 𝒟 4similar-to𝒟4\mathcal{D}\,\sim\,4caligraphic_D ∼ 4 (Greene et al. 2020), although detections are made in the most massive systems hosting MBHs M> 106Msubscript𝑀superscript106subscriptMdirect-productM_{\bullet}\,{>}\,10^{6}\,\mathrm{M_{\odot}}italic_M start_POSTSUBSCRIPT ∙ end_POSTSUBSCRIPT > 10 start_POSTSUPERSCRIPT 6 end_POSTSUPERSCRIPT roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT.

We underline that there is a degeneracy between the occupation of MBHs per galaxy and the occupation of NSCs per MBH. In particular, simultaneous increase and decrease of each number density can result in the same observed normalization of the volumetric TDE rates. Interestingly, denser environments (galaxy clusters) are found to have higher-than-average occupation fractions of both MBHs (Tremmel et al. 2024) and NSCs (Sánchez-Janssen et al. 2019; Hoyer et al. 2021; Zanatta et al. 2021; Leaman & van de Ven 2022) at a given galaxy stellar mass, in line with maxOcc model predictions. Based on our model and the novel TDE rates per-galaxy of our study we speculate that the following conditions should be fulfilled by future models, for them to explain the observed TDE rates: i) a tight connection between the two families of nuclear compact objects (in our model all NSCs host an MBH) and ii) the dominant contributor to TDE rates to be MBHs hosted in M 109 1010.5Msimilar-tosubscript𝑀superscript109superscript1010.5subscriptMdirect-productM_{\ast}\,{\sim}\,10^{9}\,{-}\,10^{10.5}\,\mathrm{M_{\odot}}italic_M start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT ∼ 10 start_POSTSUPERSCRIPT 9 end_POSTSUPERSCRIPT - 10 start_POSTSUPERSCRIPT 10.5 end_POSTSUPERSCRIPT roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT galaxies in dense environments, where NSC and MBH occupation reach 100% and the per-galaxy rate is peaking. An explicit study of the environment dependence, involving realistic star clusters and NSC modeling will be needed to further understand the connection between the two classes of objects.

4.3 Spin Evolution & Implications for TDE rates

Refer to caption
Refer to caption
Figure 10: Volumetric rates at z= 0𝑧 0z\,{=}\,0italic_z = 0 per black hole mass (top) and galaxy stellar mass (bottom) for different model variations (names of the models in legend as in text), compared with constraints from Yao et al. (2023, blue shaded area) and the fiducial model (red solid line). The shaded orange area is constructed by assuming no spin and maximally spinning MBHs in the post-processing treatment of the event-horizon suppression of the fiducial model (orange solid line).

MBHs close to the Hills mass (M 108Msimilar-tosubscript𝑀superscript108subscriptMdirect-productM_{\bullet}\,{\sim}\,10^{8}\,\mathrm{M_{\odot}}italic_M start_POSTSUBSCRIPT ∙ end_POSTSUBSCRIPT ∼ 10 start_POSTSUPERSCRIPT 8 end_POSTSUPERSCRIPT roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT) should exhibit near-to-zero TDEs by z= 0𝑧 0z\,{=}\,0italic_z = 0, and if they do they should be spinning (Kesden 2012). Also, because of their rarity, any TDE observed in this mass range will reside statistically at higher redshift (van Velzen 2018; Mummery & Balbus 2020; Hammerstein et al. 2023). Still, with the handful of massive systems observed with TDEs in the ZTF sample (Yao et al. 2023) we can test large volume statistics of the model’s MBH spin distribution.

As shown in Sect. 2.1, L-GalaxiesBH predicts a median spin of χ 0.75subscript𝜒0.75\chi_{\bullet}\,{\approx}\,0.75italic_χ start_POSTSUBSCRIPT ∙ end_POSTSUBSCRIPT ≈ 0.75 for MBH with masses M> 107Msubscript𝑀superscript107subscriptMdirect-productM_{\bullet}{>}\,10^{7}\,\mathrm{M_{\odot}}italic_M start_POSTSUBSCRIPT ∙ end_POSTSUBSCRIPT > 10 start_POSTSUPERSCRIPT 7 end_POSTSUPERSCRIPT roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT, with a standard deviation of σχ 0.2subscript𝜎subscript𝜒0.2\sigma_{\chi_{\bullet}}\,\approx\,0.2italic_σ start_POSTSUBSCRIPT italic_χ start_POSTSUBSCRIPT ∙ end_POSTSUBSCRIPT end_POSTSUBSCRIPT ≈ 0.2. Note that when initially applied on the lower-resolution Millennium-I simulation (Izquierdo-Villalba et al. 2020), the spin model predicts that M 108Mless-than-or-similar-tosubscript𝑀superscript108subscriptMdirect-productM_{\bullet}{\lesssim}\,10^{8}\,\mathrm{M_{\odot}}italic_M start_POSTSUBSCRIPT ∙ end_POSTSUBSCRIPT ≲ 10 start_POSTSUPERSCRIPT 8 end_POSTSUPERSCRIPT roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT MBHs remain high-spinning (χ> 0.8subscript𝜒0.8\chi_{\bullet}\,{>}\,0.8italic_χ start_POSTSUBSCRIPT ∙ end_POSTSUBSCRIPT > 0.8) down to z 0.1similar-to𝑧0.1z\,{\sim}\,0.1italic_z ∼ 0.1. Overall, the agreement191919We stress that we have identified the slight underprediction to be due to the lack of massive NSC abundance by the model as discussed in our results (see Sect. 4.4, model tagged as noMstcut). of the shape of the predicted volumetric rates at high MBH masses with observations (Fig. 5) suggests that the model makes a reasonable prediction of the spin distribution at z< 0.5𝑧0.5z\,{<}\,0.5italic_z < 0.5.

To guide the reader and bracket the effect of the spin model on the event-horizon suppression, Fig. 10 displays the volumetric TDEs for the same MBH population of the fiducial model but assuming χ=0subscript𝜒0\chi_{\bullet}=0italic_χ start_POSTSUBSCRIPT ∙ end_POSTSUBSCRIPT = 0 and χ=0.998subscript𝜒0.998\chi_{\bullet}=0.998italic_χ start_POSTSUBSCRIPT ∙ end_POSTSUBSCRIPT = 0.998 when computing the event horizon suppression (Sect. 2.3.4) as a shaded region around the model line. As shown, a non-spinning MBH population in the mass range M= 107.5 108.5Msubscript𝑀superscript107.5superscript108.5subscriptMdirect-productM_{\bullet}\,{=}\,10^{7.5}\,{-}\,10^{8.5}\,\mathrm{M_{\odot}}italic_M start_POSTSUBSCRIPT ∙ end_POSTSUBSCRIPT = 10 start_POSTSUPERSCRIPT 7.5 end_POSTSUPERSCRIPT - 10 start_POSTSUPERSCRIPT 8.5 end_POSTSUPERSCRIPT roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT (left border of the shaded region) can be excluded while a maximally spinning population in the same range, the model rates lie within the error margins of the constraints. Our results are consistent with the recent findings from Mummery et al. (2024) who derived MBH spin χ= 0.3 0.75subscript𝜒0.30.75\chi_{\bullet}\,{=}\,0.3\,{-}\,0.75italic_χ start_POSTSUBSCRIPT ∙ end_POSTSUBSCRIPT = 0.3 - 0.75 by modeling the late-time TDE emission of 10 MBHs with mass M> 107Msubscript𝑀superscript107subscriptMdirect-productM_{\bullet}\,{>}\,10^{7}\,\mathrm{M_{\odot}}italic_M start_POSTSUBSCRIPT ∙ end_POSTSUBSCRIPT > 10 start_POSTSUPERSCRIPT 7 end_POSTSUPERSCRIPT roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT.

We are careful with our interpretation since for converting simulated TDE rates to observable ones we have neglected some aspects of the disruption of the stars themselves. For example, we have used a truncated (m< 1.5Msubscript𝑚1.5subscriptMdirect-productm_{\ast}\,{<}\,1.5\,\mathrm{M_{\odot}}italic_m start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT < 1.5 roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT) Kroupa IMF for the event-horizon suppression calculations. Nevertheless, the scenario of a top-heavy initial stellar mass functions for the stellar nuclear environment is frequently referenced as an alternative to high-spinning MBHs (Mockler et al. 2022). Our results may also be affected by the distribution of the stellar orbit inclinations as recently indicated by Singh & Kesden (2024). Furthermore, the effects of the De Sitter precession of the streams of the disrupted star on the outcome of a TDE may play an important role202020in the scenario where radiation predominantly comes from stream collisions (stream self-crossing becomes ineffective Bonnerot et al. 2022; Jankovič et al. 2024, for reference). Nevertheless, our model suggests that basic assumptions for the star population and the detectability of individual events (Sect. 2.3.4), coupled with the spin distribution of L-GalaxiesBH, can fairly reproduce the observed event-horizon suppression.

4.4 The impact of parameters choice

As seen before, the occupation fraction of MBHs in galaxies seems to play an important role in the normalization of the volumetric TDEs, MBH spin distribution regulates the shape at the event horizon suppression. Here, we examine how TDE rates are influenced by other free parameters related to the NSC model (nucleation, regeneration, and compactness) and by the conditions used to reset the TDEs. We highlight that we do not seek to explore the full parameter space but rather pinpoint the most important choices for the model. Also, we stress that our main results for MBH growth do not change qualitatively among runs featuring single-parameter variations.

More compact NSCs.

NSCs are the main stellar environments in which TDEs occur (compared to the galaxy contribution, see both per-galaxy & volumetric TDE rates in Fig. 6 and Fig. 7). When computing the number of TDEs of NSCs inside PhaseFlow we assumed an effective radius of NSCs according to specific scaling relations (see Sect. 2.2). However, a large number of uncertainties are still present in these relations. For example, NSCs hosting an MBH are expected to dynamically evolve (evaporate/expand), so the observed scale radius may be greater than the initial radius (Merritt 2009). Therefore, we have also explored how the compactness of the NSC affects the number of TDEs and consequently, the volumetric TDE rates (hereafter compactNSC model). In Fig. 10, we show the volumetric rates for a run where NSCs are initialized with a scale radius a𝑎aitalic_a reduced to 1/3131/31 / 3 of what is assumed in our fiducial model (see Sect. 2.2.2). This choice is motivated by observations showing NSCs who lie 0.5dexsimilar-toabsent0.5dex\sim 0.5\,\mathrm{dex}∼ 0.5 roman_dex below the mass-radius relation for NSCs at z= 0𝑧 0z\,{=}\,0italic_z = 0 (see e.g. Pechetti et al. 2020). We observe that the model overpredicts the rates for MBH masses M= 106 107.5Msubscript𝑀superscript106superscript107.5subscriptMdirect-productM_{\bullet}\,{=}\,10^{6}\,{-}\,10^{7.5}\,\mathrm{M_{\odot}}italic_M start_POSTSUBSCRIPT ∙ end_POSTSUBSCRIPT = 10 start_POSTSUPERSCRIPT 6 end_POSTSUPERSCRIPT - 10 start_POSTSUPERSCRIPT 7.5 end_POSTSUPERSCRIPT roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT by more than an order of magnitude. The overestimation disfavors NSCs that are significantly more compact than the observed relations used in our fiducial model. We stress that with the current implementation of NSCs, the model lacks the most massive NSCs of mass MNSC= 108 109Msubscript𝑀NSCsuperscript108superscript109subscriptMdirect-productM_{\rm NSC}\,{=}\,10^{8}\,{-}\,10^{9}\,\mathrm{M_{\odot}}italic_M start_POSTSUBSCRIPT roman_NSC end_POSTSUBSCRIPT = 10 start_POSTSUPERSCRIPT 8 end_POSTSUPERSCRIPT - 10 start_POSTSUPERSCRIPT 9 end_POSTSUPERSCRIPT roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT which are naturally more compact and will contribute at TDE rates of M> 107.5Msubscript𝑀superscript107.5subscriptMdirect-productM_{\bullet}\,{>}\,10^{7.5}\,\rm M_{\odot}italic_M start_POSTSUBSCRIPT ∙ end_POSTSUBSCRIPT > 10 start_POSTSUPERSCRIPT 7.5 end_POSTSUPERSCRIPT roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT MBH.

Variation of 𝒟nucsubscript𝒟nuc\mathcal{D}_{\rm nuc}caligraphic_D start_POSTSUBSCRIPT roman_nuc end_POSTSUBSCRIPT.

This is a key parameter in our model since it controls the frequency of NSC regeneration. Considering lower values than our fiducial case, observations presented in (Neumayer et al. 2020) suggest a value of 𝒟MNSC/M𝒟subscript𝑀NSCsubscript𝑀\mathcal{D}\,{\equiv}\,M_{\rm NSC}/M_{\bullet}caligraphic_D ≡ italic_M start_POSTSUBSCRIPT roman_NSC end_POSTSUBSCRIPT / italic_M start_POSTSUBSCRIPT ∙ end_POSTSUBSCRIPT as low as 0.010.010.010.01. However, it remains unclear if these configurations are just transients. Despite that, we test a value of 0.010.010.010.01, and no significant changes were found. For the sake of brevity and clarity, we do not display the results of this run. Instead, we boost 𝒟nucsubscript𝒟nuc\mathcal{D}_{\rm nuc}caligraphic_D start_POSTSUBSCRIPT roman_nuc end_POSTSUBSCRIPT towards higher values, from 0.10.10.10.1 to 1111 (hereafter Dnuc1 model) and show our results in Fig. 10. For this case, the rates are heavily suppressed at high MBH mass (M> 107.5Msubscript𝑀superscript107.5subscriptMdirect-productM_{\bullet}\,{>}\,10^{7.5}\,\mathrm{M_{\odot}}italic_M start_POSTSUBSCRIPT ∙ end_POSTSUBSCRIPT > 10 start_POSTSUPERSCRIPT 7.5 end_POSTSUPERSCRIPT roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT) since the NSCs with values 𝒟nuc= 0.1 1subscript𝒟nuc0.11\mathcal{D}_{\rm nuc}\,{=}\,0.1\,{-}\,1caligraphic_D start_POSTSUBSCRIPT roman_nuc end_POSTSUBSCRIPT = 0.1 - 1 are not nucleated in this run. This is a piece of evidence that small and possibly unresolved NSCs (𝒟<1𝒟1\mathcal{D}<1caligraphic_D < 1) are the main contributors to TDE rates in the event-horizon suppression mass range. However, only the replacement of this parameter with a physical NSC treatment (including the destruction of NSCs) can offer hindsight on why this is the case.

Variation of cut-off galaxy stellar mass for nucleation.

Recently, the careful observational sampling of massive galaxies presented in Ashok et al. (2023) finds a possible higher nucleation fraction in the high-mass end than the one reported in the observational constraints used for this work (i.e. Hoyer et al. 2021). In our phenomenological model of NSCs, we impose a maximum galaxy stellar mass M,NSCcutoff= 109.75Msubscript𝑀NSCcutoffsuperscript109.75subscriptMdirect-productM_{\rm*,NSC-cutoff}\,{=}\,10^{9.75}\,\mathrm{M_{\odot}}italic_M start_POSTSUBSCRIPT ∗ , roman_NSC - roman_cutoff end_POSTSUBSCRIPT = 10 start_POSTSUPERSCRIPT 9.75 end_POSTSUPERSCRIPT roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT (Sect. 2.1), above which either the conditions are not adequate to trigger the NSC formation or, equivalently, the destruction of NSC becomes important. The specific value of this cut-off is extremely important for M> 1011Msubscript𝑀superscript1011subscriptMdirect-productM_{\ast}\,{>}\,10^{11}\,\mathrm{M_{\odot}}italic_M start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT > 10 start_POSTSUPERSCRIPT 11 end_POSTSUPERSCRIPT roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT galaxies since higher values of the cut-off will both allow for a higher number density of NSCs, but also more massive NSCs (since the scaling relation extends to high-mass galaxies). On the other hand, a sufficiently low M,NSCcutoffsubscript𝑀NSCcutoffM_{\rm*,NSC-cutoff}italic_M start_POSTSUBSCRIPT ∗ , roman_NSC - roman_cutoff end_POSTSUBSCRIPT can turn off the nucleation around M> 108Msubscript𝑀superscript108subscriptMdirect-productM_{\bullet}\,{>}\,10^{8}\,\mathrm{M_{\odot}}italic_M start_POSTSUBSCRIPT ∙ end_POSTSUBSCRIPT > 10 start_POSTSUPERSCRIPT 8 end_POSTSUPERSCRIPT roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT MBHs, converting the galaxy component to the only source of TDEs in this mass range.

To explore how the volumetric rates are affected by the nucleation mass cut-off, in Fig. 10 we present the results of a run where no cut-off is applied, i.e all galaxies at all masses can host an NSC as long as 𝒟nuc> 0.1subscript𝒟nuc0.1\mathcal{D}_{\rm nuc}\,{>}\,0.1caligraphic_D start_POSTSUBSCRIPT roman_nuc end_POSTSUBSCRIPT > 0.1 (hereafter, noMstcut model). As compared to the fiducial model, this parameter variation shows that the inclusion of more massive clusters at more massive galaxies won’t affect the low-mass regime of the volumetric rate distribution and increase the rates at the event horizon suppression with error margins. These “more massive” NSCs can be similarly introduced by increasing the scatter on the MMNSCsubscript𝑀subscript𝑀NSCM_{\ast}\,{-}\,M_{\rm NSC}italic_M start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT - italic_M start_POSTSUBSCRIPT roman_NSC end_POSTSUBSCRIPT scaling relation when assigning NSC masses in the model. We anticipate the inclusion of a realistic scenario for NSC evolution and destruction will increase the TDE rates in massive galaxies.

Variation of TDE rates reset frequency.

To assess the importance of the conditions used to reset the TDE rates (see Sect. 2.3.2) we consider that only major events are capable of resetting the clock of TDEs (ΔtTD,iΔsubscript𝑡TDi\Delta t_{\rm TD,i}roman_Δ italic_t start_POSTSUBSCRIPT roman_TD , roman_i end_POSTSUBSCRIPT). In particular, major events refer here to (i) major galaxy mergers (satellite/central baryonic mass ratio >{>}>10%) and (ii) disk instability events that displace more than >{>}>10% of the stellar disk mass to the bulge. Hereafter, we refer to this model as OnlyMajor and we present its volumetric TDE rates in Fig. 10. Overall, the predictions of the model are  1dexsimilar-toabsent1dex{\sim}\,1\,\rm dex∼ 1 roman_dex below the current observational constraints regardless of the MBH and galaxy stellar masses. Yet, the suppression of rates induced by the OnlyMajor model for galaxies with M 1011Msimilar-tosubscript𝑀superscript1011subscriptMdirect-productM_{\ast}\,{\sim}\,10^{11}\,\mathrm{M_{\odot}}italic_M start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT ∼ 10 start_POSTSUPERSCRIPT 11 end_POSTSUPERSCRIPT roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT leads to a better agreement with observations than for the fiducial model. This points out that minor events occurring in massive galaxies should take place away from the galactic nucleus, hindering their possibility of resetting the TDE rates. While this might be true for massive systems, it is unlikely that for small galaxies catastrophic events such as major mergers or important disk instabilities are the unique processes able to significantly modify the stellar environment around nuclear MBHs and thus reset the relaxation process.

The underprediction of the volumetric TDE rates for small MBHs and galaxies shown in OnlyMajor model highlights that NSC structures (that dominate the overall TDE rates, see Sect. 3.1) should be dynamically influenced by their host galaxy interactions quite often. For instance, the accretion of non-major satellites and weak disk instabilities should result in a relatively frequent refilling of the loss cone of their hosted MBH. An important caveat to this argument is that catastrophic events are the type of interactions in which the basic assumptions of spherical symmetry and isotropy of stellar profiles (assumed in this work) are most frequently violated and a rate enhancement may occur (see Sect. 5).

5 Limitations and Prospects

Here we discuss the main limitations which cannot be addressed with our methods. In addition, we discuss further details on the nature of TDE events which could potentially be captured in the framework of our model. We make some suggestions for improvements regarding both the use of PhaseFlow and L-GalaxiesBH, as well as motivate further research projects as an extension to this work.

5.1 Caveats regarding input TDE rates

As pointed out in Merritt (2013), the 1D Fokker Planck approach (see Sect. 2.2) may be inaccurate over the relaxation timescale. A more general treatment that accounts for the evolution of energy and angular momentum may be reliable on longer timescales, over which the angular momentum profile alters the overall evolution of the stellar profile (see e.g.  Broggi et al. 2022). Moreover, there are effects that cannot be taken into account in PhaseFlow, at least in its current form. Noticeable ones are: resonant relaxation of Keplerian orbits (Rauch & Tremaine 1996), large scattering (instead of diffusing into the loss-cone Weissbein & Sari 2017) or loss-cone shielding (Teboul et al. 2024); all of which have been shown to have an impact on TDE rates for specific systems. On top of this, relativistic effects may affect the outcome of the interactions and thus the prediction of the rate (e.g. Stone et al. 2019). Addressing all these caveats would require a different solution to the problem than the one provided with two-body energy relaxation as performed by PhaseFlow.

In addition to the limitations due to the negligence of anisotropies, the current method is limited to working with spherical systems (while many systems are observed rotating and therefore flattened). Although Merritt (2013) argues that non-axisymmetry affects rate estimation within a factor of two, we stress that rates for axisymmetric profiles have been noted to deviate from the ones produced with PhaseFlow (see e.g. Vasiliev & Merritt 2013).Kim et al. (2018) found evidence of elongated bulges being correlated with nuclear starbursts212121This relates to the open questions of the effects of stellar bars on central nuclear activity., which may be a possible way to allow an increase in rates for a specific type of bulges. The bursty behavior of high TDE rates in certain galaxy types has been explained through radial anisotropies (Stone et al. 2018), occurring through the formation of eccentric disks (Madigan et al. 2018; Rantala & Naab 2024), as opposed to overdensity in the nucleus (as assumed in this work). Nevertheless, it is worth noticing that observations have been favoring the latter (French et al. 2020a). All these are subjects that have mainly been explored by expensive N-body simulations with tailored initial conditions and go far beyond the general/average viewpoint we have drawn for the global TDE rates.

The stellar initial mass function is also an important caveat to expand on. All systems (young and old) are assigned TDE rates from an evolved population of stars, represented by the average of the Kroupa mass function, and heavy black holes. The particle mass and the fraction of total mass of this heavy component (16M16subscriptMdirect-product16\,\mathrm{M_{\odot}}16 roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT at 7%percent77\%7 % of the cluster’s mass respectively) drive the magnitude of mass segregation in energy (a process that has its own timescale and may damp the rates, see Bortolas 2022). Since the initial phase is affected by mass segregation and low-mass systems clusters during the initial phase have higher per-galaxy TDE rates (see Fig. 6), the choice of these parameters should be investigated further in the future when studying the growth and the rates at these systems.

Regarding the complexity of the initial mass function, the theoretical calculation of rates only demands the first two moments of mass (see Stone & Metzger 2016), therefore we expect that a more complete mass distribution will not impact severely the total number of events (but only a factor of a few). Moreover, in our model, the NSC profiles are motivated by the assembly of NSCs through the accumulation of globular clusters (Antonini et al. 2012), which will have themselves evolved stellar populations (see recent advancement in simulations of the formation of NSCs by Gray et al. 2024). However, for clusters that form for the first time in a galaxy at high redshift, the assumed actual initial mass function could be lacking stellar black holes that need some time to form as remnants of the first massive stars, but also it could be more top-heavy than is currently assumed (Chon et al. 2021; Sneppen et al. 2022; Cameron et al. 2023). Going beyond the bi-chromatic distribution into a realistic initial mass function is limited for now from the computational feasibility of a huge parameter space for the input tables 222222The computational cost scales with the number of stellar particle mass bins times the number of different initial mass functions we would like to test..

5.2 Caveats regarding the galaxy environment

Regarding the galaxy evolution model used in this work, one of the main aspects affecting the TDE demographics is the galaxy sizes predicted by L-GalaxiesBH (initially introduced by Guo et al. 2011). Despite being traced self-consistently the expected sizes do not capture some particular galaxy classes discovered in the last decade (e.g. ultra-compact dwarfs). The inclusion of more refined environmental effects included in other L-Galaxies versions (e.g. Ayromlou et al. 2021, that adds ram-pressure stripping), may be necessary to capture the nature of these types of objects.

Regarding galaxy profiles, there are also some improvements in the modeling of stellar dynamics that can be included in our semi-analytical approach. Specifically, the Sérsic indices assigned to bulges by L-GalaxiesBH are motivated by z= 0𝑧 0z\,{=}\,0italic_z = 0 observations. However, this does not take into account the evolutionary history of galaxies. For instance, systems in a post-starburst and early quiescent phase (Setton et al. 2022), as well as “red nuggets” (e.g. Saracco et al. 2010; Barro et al. 2013; Ito et al. 2024; Baggen et al. 2023; Lohmann et al. 2023; Pandya et al. 2024), have been shown to display a larger Sérsic index than ordinary galaxies. Finally, the bulge velocity dispersion is not included self-consistently in L-GalaxiesBH. This hinders the possibility of exploring the effect of this quantity on the TDE rates. For instance, observations of “σ𝜎\sigmaitalic_σ-drop” bulges (Comerón et al. 2008) suggest that some bulges will have a low enough velocity dispersion to significantly shorten relaxation timescales and thus change the expected TDEs (Cen 2020).

Regarding the reset mechanism (Sect. 2.3.2), internal processes such as phase-mixing, gravitational heating from perturbers (e.g. giant molecular clouds, Merritt 2013), and bar interactions with the nuclear region will certainly affect the event rates. These effects cannot be captured by our model since they are not quantified in L-GalaxiesBH but rather would require modeling through expensive, dedicated hydrodynamical simulations that also include TDEs.

Finally, regarding the NSC phenomenological scheme itself, we do not explore nor the redshift-dependence of the scaling relations neither the conditions under which a galaxy will host an NSC. For example, NSCs can both lose mass over time due to stellar evolution, two-body relaxation, tidal shocks and MBH interactions and gain mass through in-situ star formation or cluster mergers. This makes uncertain the evolution of the NSC-galaxy stellar mass scaling relation in Eq. 1. If we assume that, at higher redshifts and in dwarf galaxies, NSCs were more massive and more compact than the observed values at z0similar-to𝑧0z\sim 0italic_z ∼ 0, the TDE rates will increase accordingly and could result in a disagreement with the observational constraints. Conversely, suppose we assume that most high-mass NSCs in the present-day Universe accumulated their mass over time via in-situ star formation. In that case, they might have been lower-mass at higher redshift compared to today, which would result in lower high-redshift TDE rates (causing a stronger redshift-dependence in Fig. 7) and could contribute to the overprediction of TDE rates in massive galaxies (right panel of Fig. 5).

5.3 Rate enhancement

The highest per-galaxy rate predicted by our fiducial model corresponds to  104similar-toabsentsuperscript104{\sim}\,10^{-4}∼ 10 start_POSTSUPERSCRIPT - 4 end_POSTSUPERSCRIPTyr-1, compatible with the earliest results of Syer & Ulmer (1999). Despite that, a sample of galaxies like ULIRGs or E+A(k+a) post-starbursts seem to feature rates of 103superscript10310^{-3}10 start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT to 101superscript10110^{-1}10 start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT per year, values that are not reproduced in any individual galaxy of L-GalaxiesBH, and rarely occur in the input grid of TDE rates (see Sect. 2.2.2). The large rate seen in post-starburst galaxies could be the result of a TDE enhancement driven by particular conditions. For instance, the recent theoretical work of French et al. (2017) supports the amplification of TDEs in the core of post-starburst galaxies, driven primarily by the compactness of these systems and not by the initial stellar mass function variations (Bortolas 2022). Another interpretation is that as soon as the AGN phase is over, the loss cone region around the MBH can be filled, thus boosting the expected TDE rates by up to two orders of magnitude compared to standard nuclei (Wang et al. 2024). Moreover, a significant fraction of MBHs in L-GalaxiesBH are found in binaries (in MS by z=0 galaxies with M=10101011Msubscript𝑀superscript1010superscript1011subscriptMdirect-productM_{*}=10^{10}-10^{11}\,\mathrm{M_{\odot}}italic_M start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT = 10 start_POSTSUPERSCRIPT 10 end_POSTSUPERSCRIPT - 10 start_POSTSUPERSCRIPT 11 end_POSTSUPERSCRIPT roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT host MBH binaries with primary’s mass M>107Msubscript𝑀superscript107subscriptMdirect-productM_{\bullet}>10^{7}\,\mathrm{M_{\odot}}italic_M start_POSTSUBSCRIPT ∙ end_POSTSUBSCRIPT > 10 start_POSTSUPERSCRIPT 7 end_POSTSUPERSCRIPT roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT, see Fig. 8 in Izquierdo-Villalba et al. 2023) at which the rates of the secondary MBH can be enhanced to be similar to those of the most massive one (see Li et al. 2017; Ryu et al. 2022; Mockler et al. 2023, and discussion further below). There are therefore three plausible explanations for the scarce number of cases with TDEs rates per MBH as high as > 104yr1absentsuperscript104superscriptyr1\rm{>}\,10^{-4}\,yr^{-1}> 10 start_POSTSUPERSCRIPT - 4 end_POSTSUPERSCRIPT roman_yr start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT in our model232323These explanations rely on the assumption that high TDEs rates per MBH (> 103yr1absentsuperscript103𝑦superscript𝑟1{>}\,10^{-3}\,yr^{-1}> 10 start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT italic_y italic_r start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT) are not due to observationally unresolved NSCs (discussed later in this section).: i) the stellar profile of dwarf and intermediate-mass galaxies is not correctly captured at all times, ii) at least one enhancement mechanism due to additional physics is determining the global rates and should be included, or iii) the theoretical uncertainty regarding the initial NSC profiles and the accuracy of the initial relaxation phase hinders the possibility of probing the initial prompt phase of violent relaxation for these objects. To examine the last explanation, a cautious approach for the initial “bursty” phase when resetting the stellar environment around the MBH (see Sect. 2.22.3.2) may resolve this problem. This will be the subject of exploration in future work.

If the rates in the aforementioned over-represented classes of galaxies are indeed dominated by NSC rates, a more careful treatment of NSCs at different galaxy types could yield exceptionally high rates. Denser, cuspier, and more massive clusters are all conditions that have a non-negligible impact on the evolution of time-dependent TDE rates (Broggi et al. 2022). We have already noted that the introduced phenomenological model does not produce the most massive NSCs of mass MNSC= 108 109Msubscript𝑀NSCsuperscript108superscript109subscriptMdirect-productM_{\rm NSC}\,{=}\,10^{8}\,{-}\,10^{9}\,\mathrm{M_{\odot}}italic_M start_POSTSUBSCRIPT roman_NSC end_POSTSUBSCRIPT = 10 start_POSTSUPERSCRIPT 8 end_POSTSUPERSCRIPT - 10 start_POSTSUPERSCRIPT 9 end_POSTSUPERSCRIPT roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT (see the mass function in Appendix A). In addition to that, the formation channel via accretion of clusters (assumed in this work) appears to operate in dwarf galaxies (MNSC< 108 109Msubscript𝑀NSCsuperscript108superscript109subscriptMdirect-productM_{\rm NSC}\,{<}\,10^{8}\,{-}\,10^{9}\,\mathrm{M_{\odot}}italic_M start_POSTSUBSCRIPT roman_NSC end_POSTSUBSCRIPT < 10 start_POSTSUPERSCRIPT 8 end_POSTSUPERSCRIPT - 10 start_POSTSUPERSCRIPT 9 end_POSTSUPERSCRIPT roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT, Johnston et al. 2020; Fahrion et al. 2021, 2022); yet central star formation may dominate cluster formation in larger galaxies (Aharon & Perets 2015; van Donkelaar et al. 2024). Consequently, the spheroidal profile used for this work could not be applied in such cases (see also Kacharov et al. 2018 and Pinna et al. 2021 for differences with host galaxy morphology). However, for the inclusion of cuspier profiles, mass segregation in angular momentum—which cannot be captured by PhaseFlow—will prevent the model from computing correctly the TDE rates and a more sophisticated approach may be needed.

5.4 Other classes of star-MBH interactions:

In this work, we have focused our results on the rates of full TDEs and their contribution to MBH growth. However, there are many possible directions for future work exploring physically different types of interactions between stars and MBHs.

5.4.1 Partial Tidal Disruption Events

Partial tidal disruption events (partial TDEs) might be equally important for the growth of black holes as full TDEs. For example, Zhong et al. (2022) find that partial TDEs are super-Eddington 2.5 times more frequently than full TDEs, while the repetitive nature of these events can spoon-feed the MBH with increased efficiency MacLeod et al. (2012); Ryu et al. (2020b). Partial TDE rates are as high as those of full disruptions both theoretically (Krolik et al. 2020) and from recent observations (Somalwar et al. 2023b). Recently, Bortolas et al. (2023) by tracing the fate of the remnants of such events and the inferred rates using PhaseFlow showed that partial TDEs can be more frequent than full TDEs up to a few tens of times for nucleated galaxies. However, the radiative signatures of partial TDEs, and how they compare with those of full TDEs, need to be addressed before discussing further their statistics (e.g. a TDE can be falsely attributed as a rising AGN event, see e.g. Chen & Shen 2021). If indeed partial outnumber full TDEs, the volumetric TDE rates may be even lower, possibly indicating a model-observation tension towards the low-mass regime (M<106Msubscript𝑀superscript106subscriptMdirect-productM_{\bullet}<10^{6}\,\mathrm{M_{\odot}}italic_M start_POSTSUBSCRIPT ∙ end_POSTSUBSCRIPT < 10 start_POSTSUPERSCRIPT 6 end_POSTSUPERSCRIPT roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT). One of the dimmest events of Yao et al. (2023), AT2020vdq with an assigned mass of M 105.5Msimilar-tosubscript𝑀superscript105.5subscriptMdirect-productM_{\bullet}\,{\sim}\,10^{5.5}\,\,\mathrm{M_{\odot}}italic_M start_POSTSUBSCRIPT ∙ end_POSTSUBSCRIPT ∼ 10 start_POSTSUPERSCRIPT 5.5 end_POSTSUPERSCRIPT roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT, has recently been flagged as a partial disruption (Somalwar et al. 2023a), a class of events that is not included in our predictions. If these and other events at M< 106Msubscript𝑀superscript106subscriptMdirect-productM_{\bullet}\,<\,10^{6}\,\mathrm{M_{\odot}}italic_M start_POSTSUBSCRIPT ∙ end_POSTSUBSCRIPT < 10 start_POSTSUPERSCRIPT 6 end_POSTSUPERSCRIPT roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT are indeed partial disruptions that would mean a tension between our model and observations (in Fig. 5). We continue this discussion quantitatively in Appendix  B for the interested readers, where we introduce some relevant definitions.

In general, different orbit distributions will result in different frequencies of full TDEs, partial TDEs, and direct captures (Park & Hayasaki 2020; Cufari et al. 2022), but also different rates for certain subcategories of full TDEs, e.g. relativistic TDEs (Ryu et al. 2023a; Amaro Seoane 2023) and TDEs with eccentric debris disks (Zanazzi & Ogilvie 2020; Wevers et al. 2022). Tapping into the statistics of the orbits can serve as an additional time-dependent output in the grid-like exploration of the parameter space which then can be provided as L-GalaxiesBH input, addressing the aforementioned issues. We plan to improve this model aspect in future works.

5.4.2 Rates from binary and wandering MBHs

The rich physics shaping the evolution of two (or more) MBHs toward coalescence following galaxy mergers, may significantly affect TDE rates. In the pairing phase, MBHs move within the potential of the bulge (not lying anymore in the dense center) and produce different rates of TDEs than central MBHs, with possible enhancement of rates during passages of stellar overdensities (Merritt 2013). After the formation of a binary (on parsec scales), the eccentric Kozai-Lidov mechanism can boost the rates of the secondary MBH to be similar to the most massive one (Chen et al. 2011; Mockler et al. 2023). Also at smaller separations (a multiple of the sum of the two MBHs tidal radii), the rates are theoretically expected to be boosted (Li et al. 2017; Ryu et al. 2022). An important enhancement would allow for an indirect inference of binaries through TDEs (Mockler et al. 2023), something to be investigated in follow-up work dedicated to rates from binary MBHs.

Finally, the inclusion of wandering MBHs, assuming that a significant amount of them can maintain a massive star cluster after ejection, may contribute to the overall TDE signal and add to the off-nucleus signal from MBHs in globular clusters (expected to be lower than the nuclear MBHs Ramirez-Ruiz & Rosswog 2009), especially when considering the smallest MBHs. The inclusion of such rates could lead to a possible explanation of the non-detection of a host galaxy (e.g. the long-duration TDE “scary barbie” Subrayan et al. 2023) or other off-nuclear transients (e.g. the Fast Blue Optical Transient “Finch” Chrimes et al. 2024).

5.4.3 Electromagnetically-dark events

Extreme Mass Ratio Inspirals (EMRIs) and Direct Plunges (as defined in Broggi et al. 2022) of stellar remnants have their own evolution within an NSC. Given the relatively large mass accreted for each EMRI and plunge event, MBH growth via dark accretion of compact objects might even be more important than accretion through TDEs. However, while the distribution of stars in NSCs can be constrained by observations, there is no direct way to probe the distribution of non-luminous objects242424This is the main reason these events were not included in our analysis because they do not fall into the same category as the optically detected TDEs. Yet, our model is already capable of investigating these phenomena, provided that it we track a population of stellar black holes within the NSC.: this adds large theoretical uncertainties in calculating the cosmological rates. Also, regimes where general relativity effects are dominant over two-body relaxation should be taken into consideration.

Also of great interest, binary stars have a large tidal radius (Merritt 2013, 10-100 times that of a single star), while a binary system may be disrupted consecutively, either by one or both MBHs (Wu & Yuan 2018). For realistic EMRI and plunge rates, triple interactions between stellar remnant binaries and the MBH must be considered (Bonetti et al. 2019; Bonetti & Sesana 2020). Our method is currently not ideal for the inclusion of binary stars and remnants. We aim to investigate the aforementioned type of interactions, along with their contribution to the gravitational wave background to be probed by Laser Interferometry Space Antenna (LISA), in a future dedicated study.

6 Conclusions

In this work, we have included for the very first time tidal disruption events (TDEs) from massive black holes in a semi-analytic model of galaxy formation and evolution. To this end, we have used the L-Galaxies model (Henriques et al. 2015) in the version presented in Izquierdo-Villalba et al. (2020, 2022) and Spinoso et al. (2023), which includes many physical processes that drive the formation and evolution of massive black holes (MBHs). For this work, we also included a phenomenological model for nuclear star clusters (NSCs) that uses the observed NSC - galaxy stellar mass relations and reproduces the observed NSC occupation fraction. Time-dependent TDE rates for each galaxy are then obtained by solving the 1D Fokker-Planck equation with PhaseFlow for a variety of stellar environments surrounding the MBHs. Finally, we have made simple post-processing assumptions for transforming theoretical to observable rates. The key findings of the model are summarized in the following:

  • \bullet

    Our model predicts volumetric TDE rates that are in agreement with the observed ones. Rates from disruptions of stars belonging to the stellar disk or bulges alone can not explain observations. Instead, we predict that the majority of TDEs should take place in NSCs, independently of galaxy or black hole mass. To explain the rates within the local cosmological volume (z< 0.5𝑧0.5z\,{<}\,0.5italic_z < 0.5), a steep black hole mass function (M1proportional-toabsentsuperscriptsubscript𝑀1{\propto}\,M_{\bullet}^{-1}∝ italic_M start_POSTSUBSCRIPT ∙ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT at M 105 106.5Msimilar-tosubscript𝑀superscript105superscript106.5subscriptMdirect-productM_{\bullet}\,{\sim}\,10^{5}\,{-}\,10^{6.5}\,\mathrm{M_{\odot}}italic_M start_POSTSUBSCRIPT ∙ end_POSTSUBSCRIPT ∼ 10 start_POSTSUPERSCRIPT 5 end_POSTSUPERSCRIPT - 10 start_POSTSUPERSCRIPT 6.5 end_POSTSUPERSCRIPT roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT) and a relatively high occupation fraction of black holes also in the dwarf regime are needed.

  • \bullet

    The TDE rates per-MBH do not follow a single power law as obtained in other works from scaling relations (e.g. N˙per-MBHMBproportional-tosubscript˙𝑁per-MBHsuperscriptsubscript𝑀𝐵\dot{N}_{\text{per-MBH}}\,{\propto}\,M_{\bullet}^{B}over˙ start_ARG italic_N end_ARG start_POSTSUBSCRIPT per-MBH end_POSTSUBSCRIPT ∝ italic_M start_POSTSUBSCRIPT ∙ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_B end_POSTSUPERSCRIPT with 0.4<B< 00.4𝐵 0-0.4\,{<}\,B\,{<}\,0- 0.4 < italic_B < 0, see Wang & Merritt 2004; Stone & Metzger 2016; Pfister et al. 2020). Instead, our model produces a positive power-law dependence, with B 0.7𝐵0.7B\,{\approx}\,0.7italic_B ≈ 0.7 for the TDE rates from the galaxy component that peak at M= 108Msubscript𝑀superscript108subscriptMdirect-productM_{\bullet}\,{=}\,10^{8}\,\mathrm{M_{\odot}}italic_M start_POSTSUBSCRIPT ∙ end_POSTSUBSCRIPT = 10 start_POSTSUPERSCRIPT 8 end_POSTSUPERSCRIPT roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT, before the event-horizon suppression completely takes over. For the rates from the NSC component (which dominate the total rates), we predict a power law index of B= 0𝐵 0B\,{=}\,0italic_B = 0 and B= 1𝐵1B\,{=}\,1italic_B = 1 for young and old NSCs, respectively, turning over at M 107Msimilar-tosubscript𝑀superscript107subscriptMdirect-productM_{\bullet}\,{\sim}\,10^{7}\,\,\mathrm{M_{\odot}}italic_M start_POSTSUBSCRIPT ∙ end_POSTSUBSCRIPT ∼ 10 start_POSTSUPERSCRIPT 7 end_POSTSUPERSCRIPT roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT.

  • \bullet

    The spin distribution, a result of the MBH spin model introduced in L-GalaxiesBH by Izquierdo-Villalba et al. (2019), naturally explains the event horizon suppression observed in the optical TDE samples. The model points out that MBHs of mass M= 107 108.5Msubscript𝑀superscript107superscript108.5subscriptMdirect-productM_{\bullet}\,{=}\,10^{7}\,{-}\,10^{8.5}\,\,\mathrm{M_{\odot}}italic_M start_POSTSUBSCRIPT ∙ end_POSTSUBSCRIPT = 10 start_POSTSUPERSCRIPT 7 end_POSTSUPERSCRIPT - 10 start_POSTSUPERSCRIPT 8.5 end_POSTSUPERSCRIPT roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT have median spin χ¯ 0.75subscript¯𝜒0.75\bar{\chi}_{\bullet}\,\approx\,0.75over¯ start_ARG italic_χ end_ARG start_POSTSUBSCRIPT ∙ end_POSTSUBSCRIPT ≈ 0.75 with a standard deviation of σχ 0.2subscript𝜎subscript𝜒0.2\sigma_{\chi_{\bullet}}\,\approx\,0.2italic_σ start_POSTSUBSCRIPT italic_χ start_POSTSUBSCRIPT ∙ end_POSTSUBSCRIPT end_POSTSUBSCRIPT ≈ 0.2.

  • \bullet

    A great fraction of TDEs (similar-to\sim100% for the most massive hosts M> 1011Msubscript𝑀superscript1011subscriptMdirect-productM_{\ast}\,{>}\,10^{11}\,\mathrm{M_{\odot}}italic_M start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT > 10 start_POSTSUPERSCRIPT 11 end_POSTSUPERSCRIPT roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT) takes place around MBHs which are also experiencing some level of gas accretion. We thus predict TDEs to be detectable on top of low-luminosity AGN (bolometric luminosity 3×10403superscript10403\times 10^{40}3 × 10 start_POSTSUPERSCRIPT 40 end_POSTSUPERSCRIPT-1042erg/ssuperscript1042ergs10^{42}\,\mathrm{erg/s}10 start_POSTSUPERSCRIPT 42 end_POSTSUPERSCRIPT roman_erg / roman_s), although AGN are generally excluded from TDE searches. Also, we predict TDE-like flares to have a volumetric rate of 0.31Gpc3yr1dex10.31superscriptGpc3superscriptyr1superscriptdex10.3-1\,\mathrm{Gpc}^{-3}\,\mathrm{yr}^{-1}\,\mathrm{dex}^{-1}0.3 - 1 roman_Gpc start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT roman_yr start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT roman_dex start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT for Seyfert galaxies with AGN luminosity Lbol= 1042subscript𝐿bolsuperscript1042L_{\rm bol}\,{=}\,10^{42}italic_L start_POSTSUBSCRIPT roman_bol end_POSTSUBSCRIPT = 10 start_POSTSUPERSCRIPT 42 end_POSTSUPERSCRIPT-1045erg/ssuperscript1045ergs10^{45}\,\mathrm{erg/s}10 start_POSTSUPERSCRIPT 45 end_POSTSUPERSCRIPT roman_erg / roman_s in the broad mass range M= 1081010.5Msubscript𝑀superscript108superscript1010.5subscriptMdirect-productM_{\ast}\,{=}\,10^{8}-10^{10.5}\,\,\mathrm{M_{\odot}}italic_M start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT = 10 start_POSTSUPERSCRIPT 8 end_POSTSUPERSCRIPT - 10 start_POSTSUPERSCRIPT 10.5 end_POSTSUPERSCRIPT roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT, with the rate increasing by a factor of a few tens at redshifts z 1𝑧1z\,{\geq}\,1italic_z ≥ 1.

To conclude, the model presented here, a combination of L-GalaxiesBH and PhaseFlow, can overall reproduce the latest observed TDE rates, while respecting the constraints on the galaxy and black hole mass functions. Our results highlight the importance of using a realistic cosmological framework to obtain a comprehensive view of TDE demographics. Furthermore, our results underline the importance of including time-dependent TDE rates (as demonstrated in Broggi et al. 2022), which allow us to account for the relaxation timescale in the statistics, unlike studies that use instantaneous TDE rates.

Finally, while we do not find a significant contribution of TDEs to the average MBH mass growth, TDE growth might be important for specific classes of MBHs and seeds, which will be investigated in a follow-up work (Polkas et al. in prep). We also plan to address the caveats and limitations of our model in future works to improve its ability to draw predictions for TDEs and the associated growth of MBHs, in particular in view of the upcoming flow of data from LSST.

Acknowledgements.
This work used, as a foundation, the 2015 public version of the Munich model of galaxy formation and evolution: L-Galaxies. The source code and a full description of the model are available at https://lgalaxiespublicrelease.github.io/.
It also uses the publically available 1D-Fokker Planck solver PhaseFlow available at http://eugvas.net/software/phaseflow/.
We thank Rosa Valiante for kindly allowing us to use her catalogues of GQd for the PopIII sub-grid physics.
M.P. and S.B. acknowledge support from the Spanish Ministerio de Ciencia e Innovación through project PID2021-124243NB-C21.
A.S., D.I.V. and E.B. acknowledge the financial support provided under the European Union’s H2020 ERC Consolidator Grant “Binary Massive Black Hole Astrophysics” (B Massive, Grant Agreement: 818691).
E.B. acknowledges support from the European Union’s Horizon Europe programme under the Marie Skłodowska-Curie grant agreement No 101105915 (TESIFA), from the European Consortium for Astroparticle Theory in the form of an Exchange Travel Grant, and the European Union’s Horizon 2020 Programme under the AHEAD2020 project (grant agreement 871158).
N.H. is a fellow of the International Max Planck Research School for Astronomy and Cosmic Physics at the University of Heidelberg (IMPRS-HD). N.H. received financial support from the European Union’s HORIZON-MSCA-2021-SE-01 Research and Innovation programme under the Marie Sklodowska-Curie grant agreement number 101086388 - Project acronym: LACEGAL.
D.S. acknowledges the support of the National Key R&D Program of China (grant no. 2018YFA0404503), the National Science Foundation of China (grant no. 12073014), the science research grants from the China Manned Space Project with No. CMS-CSST2021-A05, and Tsinghua University Initiative Scientific Research Program (No. 20223080023).

References

  • Aharon & Perets (2015) Aharon, D. & Perets, H. B. 2015, ApJ, 799, 185
  • Alexander & Bar-Or (2017) Alexander, T. & Bar-Or, B. 2017, Nature Astronomy, 1, 0147
  • Amaro Seoane (2023) Amaro Seoane, P. 2023, arXiv e-prints, arXiv:2307.13043
  • Angulo & White (2010) Angulo, R. E. & White, S. D. M. 2010, MNRAS, 405, 143
  • Antonini et al. (2015) Antonini, F., Barausse, E., & Silk, J. 2015, ApJ, 812, 72
  • Antonini et al. (2012) Antonini, F., Capuzzo-Dolcetta, R., Mastrobuono-Battisti, A., & Merritt, D. 2012, ApJ, 750, 111
  • Arca-Sedda (2016) Arca-Sedda, M. 2016, MNRAS, 455, 35
  • Arca-Sedda & Capuzzo-Dolcetta (2014) Arca-Sedda, M. & Capuzzo-Dolcetta, R. 2014, MNRAS, 444, 3738
  • Arcavi et al. (2014) Arcavi, I., Gal-Yam, A., Sullivan, M., et al. 2014, ApJ, 793, 38
  • Ashok et al. (2023) Ashok, A., Seth, A., Erwin, P., et al. 2023, ApJ, 958, 100
  • Askar et al. (2022) Askar, A., Davies, M. B., & Church, R. P. 2022, MNRAS, 511, 2631
  • Atallah et al. (2023) Atallah, D., Trani, A. A., Kremer, K., et al. 2023, MNRAS, 523, 4227
  • Auchettl et al. (2017) Auchettl, K., Guillochon, J., & Ramirez-Ruiz, E. 2017, ApJ, 838, 149
  • Aversa et al. (2015) Aversa, R., Lapi, A., de Zotti, G., Shankar, F., & Danese, L. 2015, ApJ, 810, 74
  • Ayromlou et al. (2021) Ayromlou, M., Kauffmann, G., Yates, R. M., Nelson, D., & White, S. D. M. 2021, MNRAS, 505, 492
  • Bade et al. (1996) Bade, N., Komossa, S., & Dahlem, M. 1996, A&A, 309, L35
  • Baggen et al. (2023) Baggen, J. F. W., van Dokkum, P., Labbé, I., et al. 2023, ApJ, 955, L12
  • Bahcall & Wolf (1976) Bahcall, J. N. & Wolf, R. A. 1976, ApJ, 209, 214
  • Baldry et al. (2008) Baldry, I. K., Glazebrook, K., & Driver, S. P. 2008, MNRAS, 388, 945
  • Barro et al. (2013) Barro, G., Faber, S. M., Pérez-González, P. G., et al. 2013, ApJ, 765, 104
  • Bassino et al. (1994) Bassino, L. P., Muzzio, J. C., & Rabolli, M. 1994, ApJ, 431, 634
  • Begelman et al. (1980) Begelman, M. C., Blandford, R. D., & Rees, M. J. 1980, Nature, 287, 307
  • Binney & Tremaine (1987) Binney, J. & Tremaine, S. 1987, Galactic dynamics
  • Binney & Tremaine (2008) Binney, J. & Tremaine, S. 2008, Galactic Dynamics: Second Edition
  • Blanchard et al. (2017) Blanchard, P. K., Nicholl, M., Berger, E., et al. 2017, ApJ, 843, 106
  • Bonetti et al. (2018) Bonetti, M., Haardt, F., Sesana, A., & Barausse, E. 2018, MNRAS, 477, 3910
  • Bonetti & Sesana (2020) Bonetti, M. & Sesana, A. 2020, Phys. Rev. D, 102, 103023
  • Bonetti et al. (2019) Bonetti, M., Sesana, A., Haardt, F., Barausse, E., & Colpi, M. 2019, MNRAS, 486, 4044
  • Bonnerot et al. (2022) Bonnerot, C., Pessah, M. E., & Lu, W. 2022, ApJ, 931, L6
  • Bonoli et al. (2009) Bonoli, S., Marulli, F., Springel, V., et al. 2009, Monthly Notices of the Royal Astronomical Society, 396, 423
  • Bortolas (2022) Bortolas, E. 2022, MNRAS, 511, 2885
  • Bortolas et al. (2023) Bortolas, E., Ryu, T., Broggi, L., & Sesana, A. 2023, MNRAS, 524, 3026
  • Boylan-Kolchin et al. (2009) Boylan-Kolchin, M., Springel, V., White, S. D. M., Jenkins, A., & Lemson, G. 2009, MNRAS, 398, 1150
  • Bricman & Gomboc (2020) Bricman, K. & Gomboc, A. 2020, ApJ, 890, 73
  • Brockamp et al. (2011) Brockamp, M., Baumgardt, H., & Kroupa, P. 2011, MNRAS, 418, 1308
  • Broggi et al. (2022) Broggi, L., Bortolas, E., Bonetti, M., Sesana, A., & Dotti, M. 2022, MNRAS, 514, 3270
  • Bu et al. (2023) Bu, D.-F., Chen, L., Mou, G., Qiao, E., & Yang, X.-H. 2023, MNRAS, 521, 4180
  • Bučar Bricman et al. (2023) Bučar Bricman, K., van Velzen, S., Nicholl, M., & Gomboc, A. 2023, ApJS, 268, 13
  • Bykov et al. (2023) Bykov, S. D., Gilfanov, M. R., & Sunyaev, R. A. 2023, MNRAS[arXiv:2310.00303]
  • Cameron et al. (2023) Cameron, A. J., Katz, H., Witten, C., et al. 2023, arXiv e-prints, arXiv:2311.02051
  • Cannizzaro et al. (2022) Cannizzaro, G., Levan, A. J., van Velzen, S., & Brown, G. 2022, MNRAS, 516, 529
  • Cao (2010) Cao, X. 2010, ApJ, 725, 388
  • Cao et al. (2023) Cao, X., You, B., & Wei, X. 2023, MNRAS, 526, 2331
  • Capuzzo-Dolcetta (1993) Capuzzo-Dolcetta, R. 1993, ApJ, 415, 616
  • Carlsten et al. (2022) Carlsten, S. G., Greene, J. E., Beaton, R. L., & Greco, J. P. 2022, ApJ, 927, 44
  • Carter & Luminet (1982) Carter, B. & Luminet, J. P. 1982, Nature, 296, 211
  • Cen (2020) Cen, R. 2020, ApJ, 888, L14
  • Cenko et al. (2012) Cenko, S. B., Bloom, J. S., Kulkarni, S. R., et al. 2012, MNRAS, 420, 2684
  • Chan et al. (2019) Chan, C.-H., Piran, T., Krolik, J. H., & Saban, D. 2019, ApJ, 881, 113
  • Chandrasekhar (1942) Chandrasekhar, S. 1942, Principles of stellar dynamics
  • Charalampopoulos et al. (2024) Charalampopoulos, P., Kotak, R., Wevers, T., et al. 2024, arXiv e-prints, arXiv:2401.11773
  • Chen & Shen (2021) Chen, J.-H. & Shen, R.-F. 2021, ApJ, 914, 69
  • Chen et al. (2011) Chen, X., Sesana, A., Madau, P., & Liu, F. K. 2011, ApJ, 729, 13
  • Chon et al. (2021) Chon, S., Omukai, K., & Schneider, R. 2021, MNRAS, 508, 4175
  • Chornock et al. (2014) Chornock, R., Berger, E., Gezari, S., et al. 2014, ApJ, 780, 44
  • Chrimes et al. (2024) Chrimes, A. A., Jonker, P. G., Levan, A. J., et al. 2024, MNRAS, 527, L47
  • Cohn & Kulsrud (1978) Cohn, H. & Kulsrud, R. M. 1978, The Astrophysical Journal, 226, 1087
  • Comerón et al. (2008) Comerón, S., Knapen, J. H., & Beckman, J. E. 2008, A&A, 485, 695
  • Coughlin & Nixon (2022) Coughlin, E. R. & Nixon, C. J. 2022, ApJ, 936, 70
  • Croton et al. (2006) Croton, D. J., Springel, V., White, S. D. M., et al. 2006, MNRAS, 365, 11
  • Cufari et al. (2022) Cufari, M., Coughlin, E. R., & Nixon, C. J. 2022, ApJ, 924, 34
  • Dai et al. (2015) Dai, L., McKinney, J. C., & Miller, M. C. 2015, ApJ, 812, L39
  • Davis et al. (2019) Davis, B. L., Graham, A. W., & Cameron, E. 2019, ApJ, 873, 85
  • den Brok et al. (2014) den Brok, M., Peletier, R. F., Seth, A., et al. 2014, MNRAS, 445, 2385
  • Dgany et al. (2023) Dgany, Y., Arcavi, I., Makrygianni, L., Pellegrino, C., & Howell, D. A. 2023, ApJ, 957, 57
  • Dodd et al. (2021) Dodd, S. A., Law-Smith, J. A. P., Auchettl, K., Ramirez-Ruiz, E., & Foley, R. J. 2021, ApJ, 907, L21
  • Donley et al. (2002) Donley, J. L., Brandt, W. N., Eracleous, M., & Boller, T. 2002, AJ, 124, 1308
  • Dotti et al. (2015) Dotti, M., Merloni, A., & Montuori, C. 2015, MNRAS, 448, 3603
  • Duffell et al. (2020) Duffell, P. C., D’Orazio, D., Derdzinski, A., et al. 2020, ApJ, 901, 25
  • Efstathiou et al. (1982) Efstathiou, G., Lake, G., & Negroponte, J. 1982, MNRAS, 199, 1069
  • Eigenthaler et al. (2018) Eigenthaler, P., Puzia, T. H., Taylor, M. A., et al. 2018, ApJ, 855, 142
  • Esquej et al. (2008) Esquej, P., Saxton, R. D., Komossa, S., et al. 2008, A&A, 489, 543
  • Fahrion et al. (2022) Fahrion, K., Bulichi, T.-E., Hilker, M., et al. 2022, A&A, 667, A101
  • Fahrion et al. (2021) Fahrion, K., Lyubenova, M., van de Ven, G., et al. 2021, A&A, 650, A137
  • Frank & Rees (1976) Frank, J. & Rees, M. J. 1976, MNRAS, 176, 633
  • Frederick et al. (2021) Frederick, S., Gezari, S., Graham, M. J., et al. 2021, ApJ, 920, 56
  • French et al. (2016) French, K. D., Arcavi, I., & Zabludoff, A. 2016, ApJ, 818, L21
  • French et al. (2017) French, K. D., Arcavi, I., & Zabludoff, A. 2017, ApJ, 835, 176
  • French et al. (2020a) French, K. D., Arcavi, I., Zabludoff, A. I., et al. 2020a, ApJ, 891, 93
  • French et al. (2020b) French, K. D., Wevers, T., Law-Smith, J., Graur, O., & Zabludoff, A. I. 2020b, Space Sci. Rev., 216, 32
  • Gadotti (2009) Gadotti, D. A. 2009, MNRAS, 393, 1531
  • Gallo & Sesana (2019) Gallo, E. & Sesana, A. 2019, ApJ, 883, L18
  • Gebhardt et al. (2001) Gebhardt, K., Lauer, T. R., Kormendy, J., et al. 2001, AJ, 122, 2469
  • Generozov & Perets (2023) Generozov, A. & Perets, H. B. 2023, MNRAS, 522, 1763
  • Georgiev et al. (2016) Georgiev, I. Y., Böker, T., Leigh, N., Lützgendorf, N., & Neumayer, N. 2016, MNRAS, 457, 2122
  • Gezari et al. (2008) Gezari, S., Basa, S., Martin, D. C., et al. 2008, ApJ, 676, 944
  • Gezari et al. (2006) Gezari, S., Martin, D. C., Milliard, B., et al. 2006, ApJ, 653, L25
  • Goldtooth et al. (2023) Goldtooth, A., Zabludoff, A. I., Wen, S., et al. 2023, PASP, 135, 034101
  • Graham & Spitler (2009) Graham, A. W. & Spitler, L. R. 2009, MNRAS, 397, 2148
  • Graur et al. (2018) Graur, O., French, K. D., Zahid, H. J., et al. 2018, ApJ, 853, 39
  • Gray et al. (2024) Gray, E. I., Read, J. I., Taylor, E., et al. 2024, arXiv e-prints, arXiv:2405.19286
  • Greene (2012) Greene, J. E. 2012, Nature Communications, 3, 1304
  • Greene et al. (2020) Greene, J. E., Strader, J., & Ho, L. C. 2020, ARA&A, 58, 257
  • Guillochon & Ramirez-Ruiz (2013) Guillochon, J. & Ramirez-Ruiz, E. 2013, ApJ, 767, 25
  • Guo et al. (2011) Guo, Q., White, S., Boylan-Kolchin, M., et al. 2011, MNRAS, 413, 101
  • Guolo et al. (2024) Guolo, M., Gezari, S., Yao, Y., et al. 2024, ApJ, 966, 160
  • Häberle et al. (2024) Häberle, M., Neumayer, N., Seth, A., et al. 2024, arXiv e-prints, arXiv:2405.06015
  • Haiman et al. (1997) Haiman, Z., Rees, M. J., & Loeb, A. 1997, ApJ, 476, 458
  • Hambleton et al. (2023) Hambleton, K. M., Bianco, F. B., Street, R., et al. 2023, PASP, 135, 105002
  • Hammerstein et al. (2021) Hammerstein, E., Gezari, S., van Velzen, S., et al. 2021, ApJ, 908, L20
  • Hammerstein et al. (2023) Hammerstein, E., van Velzen, S., Gezari, S., et al. 2023, ApJ, 942, 9
  • Hartmann et al. (2011) Hartmann, M., Debattista, V. P., Seth, A., Cappellari, M., & Quinn, T. R. 2011, MNRAS, 418, 2697
  • Henriques et al. (2015) Henriques, B. M. B., White, S. D. M., Thomas, P. A., et al. 2015, MNRAS, 451, 2663
  • Henriques et al. (2020) Henriques, B. M. B., Yates, R. M., Fu, J., et al. 2020, MNRAS, 491, 5795
  • Hills (1975) Hills, J. G. 1975, Nature, 254, 295
  • Hinkle et al. (2021) Hinkle, J. T., Holoien, T. W. S., Auchettl, K., et al. 2021, MNRAS, 500, 1673
  • Hinkle et al. (2022) Hinkle, J. T., Holoien, T. W. S., Shappee, B. J., et al. 2022, ApJ, 930, 12
  • Homan et al. (2023) Homan, D., Krumpe, M., Markowitz, A., et al. 2023, A&A, 672, A167
  • Hoyer et al. (2024) Hoyer, N., Arcodia, R., Bonoli, S., et al. 2024, A&A, 682, A36
  • Hoyer et al. (2021) Hoyer, N., Neumayer, N., Georgiev, I. Y., Seth, A. C., & Greene, J. E. 2021, MNRAS, 507, 3246
  • Hoyer et al. (2023) Hoyer, N., Neumayer, N., Seth, A. C., Georgiev, I. Y., & Greene, J. E. 2023, MNRAS, 520, 4664
  • Huang & Lu (2024) Huang, H.-T. & Lu, W. 2024, MNRAS, 527, 1865
  • Huang et al. (2023) Huang, S., Jiang, N., Lin, Z., Zhu, J., & Wang, T. 2023, MNRAS, 525, 4057
  • Ito et al. (2024) Ito, K., Valentino, F., Brammer, G., et al. 2024, ApJ, 964, 192
  • Izquierdo-Villalba et al. (2020) Izquierdo-Villalba, D., Bonoli, S., Dotti, M., et al. 2020, MNRAS, 495, 4681
  • Izquierdo-Villalba et al. (2019) Izquierdo-Villalba, D., Bonoli, S., Spinoso, D., et al. 2019, MNRAS, 488, 609
  • Izquierdo-Villalba et al. (2022) Izquierdo-Villalba, D., Sesana, A., Bonoli, S., & Colpi, M. 2022, MNRAS, 509, 3488
  • Izquierdo-Villalba et al. (2023) Izquierdo-Villalba, D., Sesana, A., & Colpi, M. 2023, MNRAS, 519, 2083
  • Jankovič et al. (2024) Jankovič, T., Bonnerot, C., & Gomboc, A. 2024, MNRAS, 529, 673
  • Johnston et al. (2020) Johnston, E. J., Puzia, T. H., D’Ago, G., et al. 2020, MNRAS, 495, 2247
  • Kacharov et al. (2018) Kacharov, N., Neumayer, N., Seth, A. C., et al. 2018, MNRAS, 480, 1973
  • Karas & Šubr (2007) Karas, V. & Šubr, L. 2007, A&A, 470, 11
  • Kaur & Stone (2024) Kaur, K. & Stone, N. C. 2024, arXiv e-prints, arXiv:2405.18500
  • Kaviraj et al. (2007) Kaviraj, S., Kirkby, L. A., Silk, J., & Sarzi, M. 2007, MNRAS, 382, 960
  • Kawamuro et al. (2016) Kawamuro, T., Ueda, Y., Shidatsu, M., et al. 2016, PASJ, 68, 58
  • Kesden (2012) Kesden, M. 2012, Phys. Rev. D, 85, 024037
  • Khan & Holley-Bockelmann (2021) Khan, F. M. & Holley-Bockelmann, K. 2021, MNRAS, 508, 1174
  • Kim et al. (2018) Kim, E., Kim, S. S., Choi, Y.-Y., et al. 2018, MNRAS, 479, 562
  • Kimbrell et al. (2021) Kimbrell, S. J., Reines, A. E., Schutte, Z., Greene, J. E., & Geha, M. 2021, ApJ, 911, 134
  • Kıroğlu et al. (2023) Kıroğlu, F., Lombardi, J. C., Kremer, K., et al. 2023, ApJ, 948, 89
  • Kochanek (2016) Kochanek, C. S. 2016, MNRAS, 461, 371
  • Komossa & Bade (1999) Komossa, S. & Bade, N. 1999, A&A, 343, 775
  • Kool et al. (2020) Kool, E. C., Reynolds, T. M., Mattila, S., et al. 2020, MNRAS, 498, 2167
  • Kormendy & Ho (2013) Kormendy, J. & Ho, L. C. 2013, ARA&A, 51, 511
  • Kremer et al. (2022) Kremer, K., Lombardi, J. C., Lu, W., Piro, A. L., & Rasio, F. A. 2022, ApJ, 933, 203
  • Kremer et al. (2023) Kremer, K., Mockler, B., Piro, A. L., & Lombardi, J. C. 2023, MNRAS, 524, 6358
  • Kritos et al. (2023) Kritos, K., Berti, E., & Silk, J. 2023, Phys. Rev. D, 108, 083012
  • Krolik et al. (2020) Krolik, J., Piran, T., & Ryu, T. 2020, ApJ, 904, 68
  • Krolik et al. (2016) Krolik, J., Piran, T., Svirski, G., & Cheng, R. M. 2016, ApJ, 827, 127
  • Kroupa (2001) Kroupa, P. 2001, MNRAS, 322, 231
  • Law-Smith et al. (2017) Law-Smith, J., Ramirez-Ruiz, E., Ellison, S. L., & Foley, R. J. 2017, ApJ, 850, 22
  • Leaman & van de Ven (2022) Leaman, R. & van de Ven, G. 2022, MNRAS, 516, 4691
  • Lee et al. (2023) Lee, S., Kim, J.-h., & Oh, B. K. 2023, ApJ, 943, 77
  • Li et al. (2023) Li, J., Wang, Z.-X., Zheng, D., et al. 2023, Research in Astronomy and Astrophysics, 23, 025012
  • Li et al. (2017) Li, S., Liu, F. K., Berczik, P., & Spurzem, R. 2017, ApJ, 834, 195
  • Lin et al. (2022) Lin, Z., Jiang, N., Kong, X., et al. 2022, ApJ, 939, L33
  • Liu et al. (2023) Liu, C., Mockler, B., Ramirez-Ruiz, E., et al. 2023, ApJ, 944, 184
  • Liu et al. (2020) Liu, Z., Li, D., Liu, H.-Y., et al. 2020, ApJ, 894, 93
  • Lohmann et al. (2023) Lohmann, F. S., Schnorr-Müller, A., Trevisan, M., Ricci, T. V., & Clerici, K. S. 2023, MNRAS, 524, 5266
  • Lousto et al. (2012) Lousto, C. O., Zlochower, Y., Dotti, M., & Volonteri, M. 2012, Phys. Rev. D, 85, 084015
  • MacLeod et al. (2012) MacLeod, M., Guillochon, J., & Ramirez-Ruiz, E. 2012, ApJ, 757, 134
  • Madigan et al. (2018) Madigan, A.-M., Halle, A., Moody, M., et al. 2018, ApJ, 853, 141
  • Magorrian & Tremaine (1999) Magorrian, J. & Tremaine, S. 1999, MNRAS, 309, 447
  • Mainetti et al. (2017) Mainetti, D., Lupi, A., Campana, S., et al. 2017, A&A, 600, A124
  • Makrygianni et al. (2023) Makrygianni, L., Trakhtenbrot, B., Arcavi, I., et al. 2023, ApJ, 953, 32
  • Malyali et al. (2023) Malyali, A., Liu, Z., Rau, A., et al. 2023, MNRAS, 520, 3549
  • Masterson et al. (2024) Masterson, M., De, K., Panagiotou, C., et al. 2024, ApJ, 961, 211
  • Mattila et al. (2018) Mattila, S., Pérez-Torres, M., Efstathiou, A., et al. 2018, Science, 361, 482
  • Mayes et al. (2021) Mayes, R. J., Drinkwater, M. J., Pfeffer, J., et al. 2021, MNRAS, 506, 2459
  • McKernan et al. (2022) McKernan, B., Ford, K. E. S., Cantiello, M., et al. 2022, MNRAS, 514, 4102
  • Merloni & Heinz (2008) Merloni, A. & Heinz, S. 2008, MNRAS, 388, 1011
  • Merritt (2009) Merritt, D. 2009, ApJ, 694, 959
  • Merritt (2013) Merritt, D. 2013, Classical and Quantum Gravity, 30, 244005
  • Merritt et al. (2001) Merritt, D., Ferrarese, L., & Joseph, C. L. 2001, Science, 293, 1116
  • Mezcua (2017) Mezcua, M. 2017, International Journal of Modern Physics D, 26, 1730021
  • Mezcua et al. (2018) Mezcua, M., Civano, F., Marchesi, S., et al. 2018, MNRAS, 478, 2576
  • Mezcua & Domínguez Sánchez (2020) Mezcua, M. & Domínguez Sánchez, H. 2020, ApJ, 898, L30
  • Mezcua & Domínguez Sánchez (2024) Mezcua, M. & Domínguez Sánchez, H. 2024, MNRAS, 528, 5252
  • Miller & Davies (2012) Miller, M. C. & Davies, M. B. 2012, ApJ, 755, 81
  • Milosavljević et al. (2006) Milosavljević, M., Merritt, D., & Ho, L. C. 2006, ApJ, 652, 120
  • Mockler et al. (2023) Mockler, B., Melchor, D., Naoz, S., & Ramirez-Ruiz, E. 2023, ApJ, 959, 18
  • Mockler & Ramirez-Ruiz (2021) Mockler, B. & Ramirez-Ruiz, E. 2021, ApJ, 906, 101
  • Mockler et al. (2022) Mockler, B., Twum, A. A., Auchettl, K., et al. 2022, ApJ, 924, 70
  • Moustakas et al. (2013) Moustakas, J., Coil, A. L., Aird, J., et al. 2013, ApJ, 767, 50
  • Muñoz et al. (2015) Muñoz, R. P., Eigenthaler, P., Puzia, T. H., et al. 2015, ApJ, 813, L15
  • Mummery (2024) Mummery, A. 2024, MNRAS, 527, 6233
  • Mummery & Balbus (2020) Mummery, A. & Balbus, S. A. 2020, MNRAS, 497, L13
  • Mummery et al. (2024) Mummery, A., van Velzen, S., Nathan, E., et al. 2024, MNRAS, 527, 2452
  • Mutlu-Pakdil et al. (2016) Mutlu-Pakdil, B., Seigar, M. S., & Davis, B. L. 2016, ApJ, 830, 117
  • Naiman et al. (2015) Naiman, J. P., Ramirez-Ruiz, E., Debuhr, J., & Ma, C. P. 2015, ApJ, 803, 81
  • Nasim et al. (2023) Nasim, S. S., Fabj, G., Caban, F., et al. 2023, MNRAS, 522, 5393
  • Neumayer et al. (2020) Neumayer, N., Seth, A., & Böker, T. 2020, A&A Rev., 28, 4
  • Neumayer & Walcher (2012) Neumayer, N. & Walcher, C. J. 2012, Advances in Astronomy, 2012, 709038
  • Nguyen et al. (2022) Nguyen, D. D., Bureau, M., Thater, S., et al. 2022, MNRAS, 509, 2920
  • Nguyen et al. (2018) Nguyen, D. D., Seth, A. C., Neumayer, N., et al. 2018, ApJ, 858, 118
  • Oates et al. (2024) Oates, S. R., Kuin, N. P. M., Nicholl, M., et al. 2024, MNRAS, 530, 1688
  • Panagiotou et al. (2023) Panagiotou, C., De, K., Masterson, M., et al. 2023, ApJ, 948, L5
  • Pandya et al. (2024) Pandya, V., Zhang, H., Huertas-Company, M., et al. 2024, ApJ, 963, 54
  • Park & Hayasaki (2020) Park, G. & Hayasaki, K. 2020, ApJ, 900, 3
  • Payne et al. (2021) Payne, A. V., Shappee, B. J., Hinkle, J. T., et al. 2021, ApJ, 910, 125
  • Pechetti et al. (2020) Pechetti, R., Seth, A., Neumayer, N., et al. 2020, ApJ, 900, 32
  • Pérez-González et al. (2008) Pérez-González, P. G., Trujillo, I., Barro, G., et al. 2008, ApJ, 687, 50
  • Pfeffer et al. (2014) Pfeffer, J., Griffen, B. F., Baumgardt, H., & Hilker, M. 2014, MNRAS, 444, 3670
  • Pfister et al. (2021) Pfister, H., Dai, J. L., Volonteri, M., et al. 2021, MNRAS, 500, 3944
  • Pfister et al. (2022) Pfister, H., Toscani, M., Wong, T. H. T., et al. 2022, MNRAS, 510, 2025
  • Pfister et al. (2020) Pfister, H., Volonteri, M., Dai, J. L., & Colpi, M. 2020, MNRAS, 497, 2276
  • Phinney (1989) Phinney, E. S. 1989, in The Center of the Galaxy, ed. M. Morris, Vol. 136, 543
  • Pinna et al. (2021) Pinna, F., Neumayer, N., Seth, A., et al. 2021, ApJ, 921, 8
  • Planck Collaboration et al. (2020) Planck Collaboration, Aghanim, N., Akrami, Y., et al. 2020, A&A, 641, A6
  • Portegies Zwart & McMillan (2002) Portegies Zwart, S. F. & McMillan, S. L. W. 2002, ApJ, 576, 899
  • Prasad et al. (2024) Prasad, C., Wang, Y., Perna, R., Ford, K. E. S., & McKernan, B. 2024, MNRAS, 531, 1409
  • Quinlan & Hernquist (1997) Quinlan, G. D. & Hernquist, L. 1997, New A, 2, 533
  • Ramirez-Ruiz & Rosswog (2009) Ramirez-Ruiz, E. & Rosswog, S. 2009, ApJ, 697, L77
  • Rantala & Naab (2024) Rantala, A. & Naab, T. 2024, MNRAS, 527, 11458
  • Rantala et al. (2024) Rantala, A., Naab, T., & Lahén, N. 2024, arXiv e-prints, arXiv:2403.10602
  • Rauch & Tremaine (1996) Rauch, K. P. & Tremaine, S. 1996, New A, 1, 149
  • Rees (1988) Rees, M. J. 1988, Nature, 333, 523
  • Rees & Ostriker (1977) Rees, M. J. & Ostriker, J. P. 1977, MNRAS, 179, 541
  • Reines & Volonteri (2015) Reines, A. E. & Volonteri, M. 2015, ApJ, 813, 82
  • Reynolds (2021) Reynolds, C. S. 2021, ARA&A, 59, 117
  • Reynolds et al. (2022) Reynolds, T. M., Mattila, S., Efstathiou, A., et al. 2022, A&A, 664, A158
  • Ricci & Trakhtenbrot (2023) Ricci, C. & Trakhtenbrot, B. 2023, Nature Astronomy, 7, 1282
  • Rizzuto et al. (2023) Rizzuto, F. P., Naab, T., Rantala, A., et al. 2023, MNRAS, 521, 2930
  • Roediger & Courteau (2015) Roediger, J. C. & Courteau, S. 2015, MNRAS, 452, 3209
  • Ryu et al. (2023a) Ryu, T., Krolik, J., & Piran, T. 2023a, ApJ, 946, L33
  • Ryu et al. (2020a) Ryu, T., Krolik, J., Piran, T., & Noble, S. C. 2020a, ApJ, 904, 98
  • Ryu et al. (2020b) Ryu, T., Krolik, J., Piran, T., & Noble, S. C. 2020b, ApJ, 904, 100
  • Ryu et al. (2023b) Ryu, T., McKernan, B., Ford, K. E. S., et al. 2023b, MNRAS[arXiv:2310.00610]
  • Ryu et al. (2022) Ryu, T., Trani, A. A., & Leigh, N. W. C. 2022, MNRAS, 515, 2430
  • Saha (1992) Saha, P. 1992, MNRAS, 254, 132
  • Sánchez-Janssen et al. (2019) Sánchez-Janssen, R., Côté, P., Ferrarese, L., et al. 2019, ApJ, 878, 18
  • Saracco et al. (2010) Saracco, P., Longhetti, M., & Gargiulo, A. 2010, MNRAS, 408, L21
  • Sazonov et al. (2021) Sazonov, S., Gilfanov, M., Medvedev, P., et al. 2021, MNRAS, 508, 3820
  • Sesana & Khan (2015) Sesana, A. & Khan, F. M. 2015, MNRAS, 454, L66
  • Seth et al. (2008) Seth, A., Agüeros, M., Lee, D., & Basu-Zych, A. 2008, ApJ, 678, 116
  • Setton et al. (2022) Setton, D. J., Verrico, M., Bezanson, R., et al. 2022, ApJ, 931, 51
  • Shankar et al. (2019) Shankar, F., Bernardi, M., Richardson, K., et al. 2019, MNRAS, 485, 1278
  • Shankar et al. (2016) Shankar, F., Bernardi, M., Sheth, R. K., et al. 2016, MNRAS, 460, 3119
  • Shankar et al. (2009) Shankar, F., Weinberg, D. H., & Miralda-Escudé, J. 2009, ApJ, 690, 20
  • Shankar et al. (2013) Shankar, F., Weinberg, D. H., & Miralda-Escudé, J. 2013, MNRAS, 428, 421
  • Shen et al. (2003) Shen, S., Mo, H. J., White, S. D. M., et al. 2003, MNRAS, 343, 978
  • Singh & Kesden (2024) Singh, T. & Kesden, M. 2024, Phys. Rev. D, 109, 043016
  • Siudek et al. (2023) Siudek, M., Mezcua, M., & Krywult, J. 2023, MNRAS, 518, 724
  • Sneppen et al. (2022) Sneppen, A., Steinhardt, C. L., Hensley, H., et al. 2022, ApJ, 931, 57
  • Somalwar et al. (2023a) Somalwar, J. J., Ravi, V., & Lu, W. 2023a, arXiv e-prints, arXiv:2310.03795
  • Somalwar et al. (2023b) Somalwar, J. J., Ravi, V., Yao, Y., et al. 2023b, arXiv e-prints, arXiv:2310.03782
  • Spergel et al. (2003) Spergel, D. N., Verde, L., Peiris, H. V., et al. 2003, ApJS, 148, 175
  • Spinoso et al. (2023) Spinoso, D., Bonoli, S., Valiante, R., Schneider, R., & Izquierdo-Villalba, D. 2023, MNRAS, 518, 4672
  • Springel et al. (2005) Springel, V., White, S. D. M., Jenkins, A., et al. 2005, Nature, 435, 629
  • Steinberg & Stone (2024) Steinberg, E. & Stone, N. C. 2024, Nature, 625, 463
  • Stone et al. (2013) Stone, N., Sari, R., & Loeb, A. 2013, MNRAS, 435, 1809
  • Stone et al. (2018) Stone, N. C., Generozov, A., Vasiliev, E., & Metzger, B. D. 2018, MNRAS, 480, 5060
  • Stone et al. (2019) Stone, N. C., Kesden, M., Cheng, R. M., & van Velzen, S. 2019, General Relativity and Gravitation, 51, 30
  • Stone et al. (2017) Stone, N. C., Küpper, A. H. W., & Ostriker, J. P. 2017, MNRAS, 467, 4180
  • Stone & Metzger (2016) Stone, N. C. & Metzger, B. D. 2016, MNRAS, 455, 859
  • Stone et al. (2020) Stone, N. C., Vasiliev, E., Kesden, M., et al. 2020, Space Sci. Rev., 216, 35
  • Subrayan et al. (2023) Subrayan, B. M., Milisavljevic, D., Chornock, R., et al. 2023, ApJ, 948, L19
  • Sutherland & Dopita (1993) Sutherland, R. S. & Dopita, M. A. 1993, ApJS, 88, 253
  • Syer & Ulmer (1999) Syer, D. & Ulmer, A. 1999, MNRAS, 306, 35
  • Tadhunter et al. (2017) Tadhunter, C., Spence, R., Rose, M., Mullaney, J., & Crowther, P. 2017, Nature Astronomy, 1, 0061
  • Teboul et al. (2024) Teboul, O., Stone, N. C., & Ostriker, J. P. 2024, MNRAS, 527, 3094
  • Thater et al. (2023) Thater, S., Lyubenova, M., Fahrion, K., et al. 2023, A&A, 675, A18
  • Trani et al. (2018) Trani, A. A., Mapelli, M., & Ballone, A. 2018, ApJ, 864, 17
  • Tremmel et al. (2024) Tremmel, M., Ricarte, A., Natarajan, P., et al. 2024, The Open Journal of Astrophysics, 7, 26
  • Trump et al. (2015) Trump, J. R., Sun, M., Zeimann, G. R., et al. 2015, ApJ, 811, 26
  • Urquhart et al. (2022) Urquhart, R., McDermott, L. I., Strader, J., et al. 2022, ApJ, 940, 111
  • Valiante et al. (2016) Valiante, R., Schneider, R., Volonteri, M., & Omukai, K. 2016, MNRAS, 457, 3356
  • Valluri et al. (2005) Valluri, M., Ferrarese, L., Merritt, D., & Joseph, C. L. 2005, ApJ, 628, 137
  • van Donkelaar et al. (2024) van Donkelaar, F., Mayer, L., Capelo, P. R., et al. 2024, MNRAS, 529, 4104
  • van Velzen (2018) van Velzen, S. 2018, ApJ, 852, 72
  • van Velzen & Farrar (2014) van Velzen, S. & Farrar, G. R. 2014, ApJ, 792, 53
  • van Velzen et al. (2011) van Velzen, S., Farrar, G. R., Gezari, S., et al. 2011, ApJ, 741, 73
  • Vasiliev (2017) Vasiliev, E. 2017, ApJ, 848, 10
  • Vasiliev (2019) Vasiliev, E. 2019, MNRAS, 482, 1525
  • Vasiliev & Merritt (2013) Vasiliev, E. & Merritt, D. 2013, ApJ, 774, 87
  • Vika et al. (2009) Vika, M., Driver, S. P., Graham, A. W., & Liske, J. 2009, MNRAS, 400, 1451
  • Vynatheya et al. (2024) Vynatheya, P., Ryu, T., Pakmor, R., de Mink, S. E., & Perets, H. B. 2024, A&A, 685, A45
  • Wang & Merritt (2004) Wang, J. & Merritt, D. 2004, ApJ, 600, 149
  • Wang et al. (2024) Wang, M., Ma, Y., Wu, Q., & Jiang, N. 2024, ApJ, 960, 69
  • Wang et al. (2022) Wang, Y., Jiang, N., Wang, T., et al. 2022, ApJ, 930, L4
  • Weissbein & Sari (2017) Weissbein, A. & Sari, R. 2017, MNRAS, 468, 1760
  • Wevers et al. (2022) Wevers, T., Nicholl, M., Guolo, M., et al. 2022, A&A, 666, A6
  • Wevers et al. (2019) Wevers, T., Stone, N. C., van Velzen, S., et al. 2019, MNRAS, 487, 4136
  • White & Frenk (1991) White, S. D. M. & Frenk, C. S. 1991, ApJ, 379, 52
  • White & Rees (1978) White, S. D. M. & Rees, M. J. 1978, MNRAS, 183, 341
  • Wild et al. (2016) Wild, V., Almaini, O., Dunlop, J., et al. 2016, MNRAS, 463, 832
  • Wild et al. (2020) Wild, V., Taj Aldeen, L., Carnall, A., et al. 2020, MNRAS, 494, 529
  • Wong et al. (2022) Wong, T. H. T., Pfister, H., & Dai, L. 2022, ApJ, 927, L19
  • Wu & Yuan (2018) Wu, X.-J. & Yuan, Y.-F. 2018, MNRAS, 479, 1569
  • Xin et al. (2024) Xin, C., Haiman, Z., Perna, R., Wang, Y., & Ryu, T. 2024, ApJ, 961, 149
  • Yao et al. (2023) Yao, Y., Ravi, V., Gezari, S., et al. 2023, ApJ, 955, L6
  • Zajaček et al. (2024) Zajaček, M., Czerny, B., Jaiswal, V. K., et al. 2024, Space Sci. Rev., 220, 29
  • Zanatta et al. (2021) Zanatta, E., Sánchez-Janssen, R., Chies-Santos, A. L., de Souza, R. S., & Blakeslee, J. P. 2021, MNRAS, 508, 986
  • Zanazzi & Ogilvie (2020) Zanazzi, J. J. & Ogilvie, G. I. 2020, MNRAS, 499, 5562
  • Zhang (2022) Zhang, X.-G. 2022, MNRAS, 517, L71
  • Zhao (1996) Zhao, H. 1996, MNRAS, 278, 488
  • Zhong et al. (2022) Zhong, S., Li, S., Berczik, P., & Spurzem, R. 2022, ApJ, 933, 96
  • Zou et al. (2023) Zou, F., Brandt, W. N., Ni, Q., et al. 2023, ApJ, 950, 136

Appendix A TDE hosts mass functions

Refer to caption
Figure 11: Stellar Mass Functions of all galaxies (solid black) and TDE-host galaxies with the highest TDE rates (Γtot> 105subscriptΓtotsuperscript105\Gamma_{\rm tot}{>}\,10^{-5}roman_Γ start_POSTSUBSCRIPT roman_tot end_POSTSUBSCRIPT > 10 start_POSTSUPERSCRIPT - 5 end_POSTSUPERSCRIPT, grey dashed) for the fiducial model at z= 0𝑧 0z\,{=}\,0italic_z = 0 and z= 1𝑧1z\,{=}\,1italic_z = 1. The first is compared with compiled mass constraints from Henriques et al. (2015) and Henriques et al. (2020), Baldry et al. (2008), Pérez-González et al. (2008), and Moustakas et al. (2013) (inset legend). With a thin colored line at lower masses, the NSC Mass Function of the fiducial model is plotted at the same redshifts. For comparison, we display the model NSC mass function from Antonini et al. (2015) at z= 0𝑧 0z\,{=}\,0italic_z = 0 (Ant15 z=0).

Our model can explain the constraints of Yao et al. (2023) on the galaxy stellar mass distribution of the volumetric rates. We discuss briefly the reasons for the relatively good agreement and the issues to be addressed regarding the TDE host mass range in future works.

As observed in Fig. 11 (grey dashed), we reproduce the flat galaxy stellar mass function expected for TDE host galaxies as inferred from observations (Law-Smith et al. 2017). Galaxy stellar mass functions are all in good agreement with one another and with observations at z= 0𝑧 0z\,{=}\,0italic_z = 0, 1111, as expected by the success of the L-Galaxies models.

In the same plot, we present the NSC mass function at z= 0𝑧 0z\,{=}\,0italic_z = 0 (dotted-dashed) which is lower compared to the theoretical predictions of Antonini et al. (2015) by an order of magnitude at lower masses, obviously having a different slope. The number density provided by this function should be a theoretical lower limit to reproduce the TDE rates, yet NSCs without MBHs should also be considered. The model predicts that NSCs do not evolve significantly with redshift from z= 1𝑧1z\,{=}\,1italic_z = 1, as they are tightly related to the nuclear MBHs (that also do not evolve).

At z= 0𝑧 0z\,{=}\,0italic_z = 0 the peak at hosts with high rates within M= 109.5subscript𝑀superscript109.5M_{\ast}\,=\,10^{9.5}italic_M start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT = 10 start_POSTSUPERSCRIPT 9.5 end_POSTSUPERSCRIPT - 1010.5Msuperscript1010.5subscriptMdirect-product10^{10.5}\,\,\mathrm{M_{\odot}}10 start_POSTSUPERSCRIPT 10.5 end_POSTSUPERSCRIPT roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT is directly related to hosting massive NSC MNSC= 107subscript𝑀NSCsuperscript107M_{\rm NSC}\,=\,10^{7}italic_M start_POSTSUBSCRIPT roman_NSC end_POSTSUBSCRIPT = 10 start_POSTSUPERSCRIPT 7 end_POSTSUPERSCRIPT-108Msuperscript108subscriptMdirect-product10^{8}\,\,\mathrm{M_{\odot}}10 start_POSTSUPERSCRIPT 8 end_POSTSUPERSCRIPT roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT and recently seeded (high per-galaxy rates, left panel in Fig. 11 for NSC distribution) while the evolving tail (M< 109Msubscript𝑀superscript109subscriptMdirect-productM_{\ast}\,{<}\,10^{9}\,\,\mathrm{M_{\odot}}italic_M start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT < 10 start_POSTSUPERSCRIPT 9 end_POSTSUPERSCRIPT roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT from z= 1𝑧1z\,{=}\,1italic_z = 1 to z= 0𝑧 0z\,{=}\,0italic_z = 0 is from the contribution of intermediate-mass MBHs and smaller NSC (right NSC distribution in Fig. 11), and systems with TDE rates decaying with time. The model predicts mild redshift evolution, a falsifiable argument with the observations of TDEs up to z= 1𝑧1z\,{=}\,1italic_z = 1 from the forthcoming LSST samples.

Appendix B Useful definitions for Tidal Disruption Event rate interpretation

An interaction between a massive black hole and a star reaches its proximity zone of less than one tidal radius rtsubscript𝑟𝑡r_{t}italic_r start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT (Eq. 3 in the main text), at fixed masses Msubscript𝑀M_{\bullet}italic_M start_POSTSUBSCRIPT ∙ end_POSTSUBSCRIPT and msubscript𝑚m_{\ast}italic_m start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT respectively, will result in a variety of outcomes, with different energetics and observable signatures depending on many parameters. How close to the horizon the star is destined to make a passage is of particular importance (Carter & Luminet 1982; Stone et al. 2013; Dai et al. 2015). It is useful to quantify this with the penetration parameter that is defined as

β=rt/rp,𝛽subscript𝑟𝑡subscript𝑟𝑝\beta=r_{t}/r_{p},italic_β = italic_r start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT / italic_r start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT ,

where rpsubscript𝑟𝑝r_{p}italic_r start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT the pericenter of the orbit and

rt=6.9×1012cmηTD2/3(M/106M)1/3m1/3rsubscript𝑟𝑡6.9superscript1012cmsuperscriptsubscript𝜂TD23superscriptsubscript𝑀superscript106subscript𝑀direct-product13superscriptsubscript𝑚13subscript𝑟r_{t}=6.9\times 10^{12}{\rm cm}\;\eta_{\rm TD}^{2/3}(M_{\bullet}/10^{6}M_{% \odot})^{1/3}m_{\ast}^{-1/3}r_{\ast}italic_r start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT = 6.9 × 10 start_POSTSUPERSCRIPT 12 end_POSTSUPERSCRIPT roman_cm italic_η start_POSTSUBSCRIPT roman_TD end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 / 3 end_POSTSUPERSCRIPT ( italic_M start_POSTSUBSCRIPT ∙ end_POSTSUBSCRIPT / 10 start_POSTSUPERSCRIPT 6 end_POSTSUPERSCRIPT italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 1 / 3 end_POSTSUPERSCRIPT italic_m start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 1 / 3 end_POSTSUPERSCRIPT italic_r start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT

the tidal radius depends on the mass/radius of the star disrupted msubscript𝑚m_{\ast}italic_m start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT and rsubscript𝑟r_{\ast}italic_r start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT (units of 1M1subscriptMdirect-product1\,\mathrm{M_{\odot}}1 roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT and 1111R). The coefficient ηTDsubscript𝜂TD\eta_{\rm TD}italic_η start_POSTSUBSCRIPT roman_TD end_POSTSUBSCRIPT depends on the polytropic index of the star and is of the order unity (Merritt 2013). Penetration factor values range from βmin= 0.5subscript𝛽min0.5\beta_{\rm min}\,=\,0.5italic_β start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT = 0.5-0.60.60.60.6 as derived from hydrodynamical simulations (Guillochon & Ramirez-Ruiz 2013) and at which point stars are merely disrupted, to βmaxsubscript𝛽max\beta_{\rm max}italic_β start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT at which point stars enter whole into the MBH’s horizon. Analytical calculation on horizon suppression for a Schwarzchild MBH yields (Kesden 2012; Mummery 2024)

βmax=11.8M,62/3m1/3r.subscript𝛽max11.8superscriptsubscript𝑀623superscriptsubscript𝑚13subscript𝑟\beta_{\rm max}=11.8M_{\bullet,6}^{-2/3}m_{\ast}^{-1/3}r_{\ast}.italic_β start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT = 11.8 italic_M start_POSTSUBSCRIPT ∙ , 6 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 2 / 3 end_POSTSUPERSCRIPT italic_m start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 1 / 3 end_POSTSUPERSCRIPT italic_r start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT . (14)

Also, below a value βcsubscript𝛽𝑐\beta_{c}italic_β start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT the star is not fully disrupted. Therefore, full TDEs in the interval βc<β<βmaxsubscript𝛽c𝛽subscript𝛽max\beta_{\rm c}\,{<}\,\beta\,{<}\,\beta_{\rm max}italic_β start_POSTSUBSCRIPT roman_c end_POSTSUBSCRIPT < italic_β < italic_β start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT while direct captures occur for β>βmax(M)𝛽subscript𝛽subscript𝑀\beta\,{>}\,\beta_{\max}(M_{\bullet})italic_β > italic_β start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT ( italic_M start_POSTSUBSCRIPT ∙ end_POSTSUBSCRIPT ). The rest of the events are found in the interval βmin<βcsubscript𝛽subscript𝛽𝑐\beta_{\min}\,{<}\,\beta_{c}italic_β start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT < italic_β start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT and are considered partial TDEs. The latter are identified by the re-flaring of their source and less frequently from their differentiation in the fading power law (9/494-9/4- 9 / 4 instead of 5/353-5/3- 5 / 3). Ryu et al. (2020a) has shown that the physical tidal radius at which the star is fully disrupted has both stellar and MBH mass dependence. For disrupted stars of mass 0.3 and 1 M, we adopt

βc=βc,011+C0(M/106M)2/3.subscript𝛽csubscript𝛽c011subscript𝐶0superscriptsubscript𝑀superscript106subscript𝑀direct-product23\beta_{\rm c}=\beta_{\rm c,0}\frac{1}{1+C_{0}(M_{\bullet}/10^{6}M_{\odot})^{2/% 3}}.italic_β start_POSTSUBSCRIPT roman_c end_POSTSUBSCRIPT = italic_β start_POSTSUBSCRIPT roman_c , 0 end_POSTSUBSCRIPT divide start_ARG 1 end_ARG start_ARG 1 + italic_C start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( italic_M start_POSTSUBSCRIPT ∙ end_POSTSUBSCRIPT / 10 start_POSTSUPERSCRIPT 6 end_POSTSUPERSCRIPT italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 2 / 3 end_POSTSUPERSCRIPT end_ARG . (15)

where βc,0=2.0subscript𝛽c02.0\beta_{\rm c,0}=2.0italic_β start_POSTSUBSCRIPT roman_c , 0 end_POSTSUBSCRIPT = 2.0,C0=0.166subscript𝐶00.166C_{0}=0.166italic_C start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 0.166 & βc,0=2.65subscript𝛽c02.65\beta_{\rm c,0}=2.65italic_β start_POSTSUBSCRIPT roman_c , 0 end_POSTSUBSCRIPT = 2.65, C0=0.204subscript𝐶00.204C_{0}=0.204italic_C start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 0.204 respectively. We set the simple form βc,0(m)=1.6(m+0.45)subscript𝛽c0subscript𝑚1.6subscript𝑚0.45\beta_{\rm c,0}(m_{\ast})=1.6(m_{\ast}+0.45)italic_β start_POSTSUBSCRIPT roman_c , 0 end_POSTSUBSCRIPT ( italic_m start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT ) = 1.6 ( italic_m start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT + 0.45 ) and fix C0=0.2subscript𝐶00.2C_{0}=0.2italic_C start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 0.2 for our simple calculations here (although there is not a linear transition between stars with different polytropic indices). For reference, Guillochon & Ramirez-Ruiz (2013) and Mainetti et al. (2017) values from simulations for convective & radiation-dominated stars around a M=106Msubscript𝑀superscript106subscriptMdirect-productM_{\bullet}=10^{6}\,\mathrm{M_{\odot}}italic_M start_POSTSUBSCRIPT ∙ end_POSTSUBSCRIPT = 10 start_POSTSUPERSCRIPT 6 end_POSTSUPERSCRIPT roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT MBH translate into βc,0=1.08subscript𝛽c01.08\beta_{\rm c,0}=1.08italic_β start_POSTSUBSCRIPT roman_c , 0 end_POSTSUBSCRIPT = 1.08 & 2.22.22.22.2 when fixing C0=0.2subscript𝐶00.2C_{0}=0.2italic_C start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 0.2 respectively. Note that there is a continuous transition from TDEs to direct captures as βc(m)βmax(M,m,χ,t),subscript𝛽csubscript𝑚subscript𝛽subscript𝑀subscript𝑚subscript𝜒𝑡\beta_{\rm c}(m_{\ast})\rightarrow\beta_{\max}(M_{\bullet},m_{\ast},\chi_{% \bullet},t),italic_β start_POSTSUBSCRIPT roman_c end_POSTSUBSCRIPT ( italic_m start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT ) → italic_β start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT ( italic_M start_POSTSUBSCRIPT ∙ end_POSTSUBSCRIPT , italic_m start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT , italic_χ start_POSTSUBSCRIPT ∙ end_POSTSUBSCRIPT , italic_t ) , with the Hills mass being rather a range of values for the distribution of star masses. This transition is computed solely in the analysis here; in the main text TDEs are considered as β>βc=1𝛽subscript𝛽c1\beta\,{>}\,\beta_{\rm c}=1italic_β > italic_β start_POSTSUBSCRIPT roman_c end_POSTSUBSCRIPT = 1 (all stars within the tidal radius are disrupted) with βmaxsubscript𝛽max\beta_{\rm max}\rightarrow\inftyitalic_β start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT → ∞ for M< 108Msubscript𝑀superscript108subscriptMdirect-productM_{\bullet}\,{<}\,10^{8}\,\mathrm{M_{\odot}}italic_M start_POSTSUBSCRIPT ∙ end_POSTSUBSCRIPT < 10 start_POSTSUPERSCRIPT 8 end_POSTSUPERSCRIPT roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT (direct captures are not considered separately) and βmax<βcsubscript𝛽subscript𝛽c\beta_{\max}\,{<}\,\beta_{\rm c}italic_β start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT < italic_β start_POSTSUBSCRIPT roman_c end_POSTSUBSCRIPT for higher MBH masses (all events are considered direct captures).

The number of orbits per penetration factor bin dN(β)/dβ𝑑𝑁𝛽𝑑𝛽dN(\beta)/d\betaitalic_d italic_N ( italic_β ) / italic_d italic_β takes the analytical form of a power-law for the full loss-cone regime (the system just started to relax). By assuming uniformity of the orbits over rpsubscript𝑟𝑝r_{p}italic_r start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT gives a power-law dependence of β2proportional-toabsentsuperscript𝛽2\,{\propto}\,\beta^{-2}∝ italic_β start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT (Stone et al. 2020). When the relativistic gravity is taken into account, this power law drops to β10/3superscript𝛽103\beta^{-10/3}italic_β start_POSTSUPERSCRIPT - 10 / 3 end_POSTSUPERSCRIPT (Coughlin & Nixon 2022) for maximally-spinning MBHs (Kerr metric). Also, we consider the scenario where orbits are distributed uniformly over surface 2πrpdrp2𝜋subscript𝑟𝑝𝑑subscript𝑟𝑝2\pi r_{p}dr_{p}2 italic_π italic_r start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT italic_d italic_r start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT and the probability function scales as β3proportional-toabsentsuperscript𝛽3\,{\propto}\,\beta^{-3}∝ italic_β start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT (Bricman & Gomboc 2020). At the empty loss-cone regime (the system has relaxed) there is a weak (logarithmic) dependence of rates on β𝛽\betaitalic_β, so essentially we can assume β0proportional-toabsentsuperscript𝛽0\,{\propto}\,\beta^{0}∝ italic_β start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT for this case. For simplicity, we assume that the distribution of orbits destined to do partial TDEs follows from one of the full TDEs as inferred for the disruption of normal stars from Zhong et al. (2022).

We can now write the fraction of full TDEs as the integral of a power-law probability βSβ1superscript𝛽subscript𝑆𝛽1\beta^{-S_{\beta}-1}italic_β start_POSTSUPERSCRIPT - italic_S start_POSTSUBSCRIPT italic_β end_POSTSUBSCRIPT - 1 end_POSTSUPERSCRIPT in the value intervals mentioned above:

fFTDE(m,M)=βc(m,M)Sββmax(m,M)SββminSββmax(m,M)Sβsubscript𝑓FTDEsubscript𝑚subscript𝑀subscript𝛽csuperscriptsubscript𝑚subscript𝑀subscript𝑆𝛽subscript𝛽superscriptsubscript𝑚subscript𝑀subscript𝑆𝛽superscriptsubscript𝛽subscript𝑆𝛽subscript𝛽superscriptsubscript𝑚subscript𝑀subscript𝑆𝛽f_{\rm FTDE}(m_{\ast},M_{\bullet})=\frac{\beta_{\rm c}(m_{\ast},M_{\bullet})^{% -S_{\beta}}-\beta_{\max}(m_{\ast},M_{\bullet})^{-S_{\beta}}}{\beta_{\min}^{-S_% {\beta}}-\beta_{\max}(m_{\ast},M_{\bullet})^{-S_{\beta}}}italic_f start_POSTSUBSCRIPT roman_FTDE end_POSTSUBSCRIPT ( italic_m start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT , italic_M start_POSTSUBSCRIPT ∙ end_POSTSUBSCRIPT ) = divide start_ARG italic_β start_POSTSUBSCRIPT roman_c end_POSTSUBSCRIPT ( italic_m start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT , italic_M start_POSTSUBSCRIPT ∙ end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT - italic_S start_POSTSUBSCRIPT italic_β end_POSTSUBSCRIPT end_POSTSUPERSCRIPT - italic_β start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT ( italic_m start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT , italic_M start_POSTSUBSCRIPT ∙ end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT - italic_S start_POSTSUBSCRIPT italic_β end_POSTSUBSCRIPT end_POSTSUPERSCRIPT end_ARG start_ARG italic_β start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - italic_S start_POSTSUBSCRIPT italic_β end_POSTSUBSCRIPT end_POSTSUPERSCRIPT - italic_β start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT ( italic_m start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT , italic_M start_POSTSUBSCRIPT ∙ end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT - italic_S start_POSTSUBSCRIPT italic_β end_POSTSUBSCRIPT end_POSTSUPERSCRIPT end_ARG (16)

where the generic power-law Sβsubscript𝑆𝛽S_{\beta}italic_S start_POSTSUBSCRIPT italic_β end_POSTSUBSCRIPT is in the range Sβ[1,2]subscript𝑆𝛽12S_{\beta}\in[1,2]italic_S start_POSTSUBSCRIPT italic_β end_POSTSUBSCRIPT ∈ [ 1 , 2 ] for our calculations, for a non-empty loss cone (for all types of events). For the empty loss-cone regime, Sβ=1subscript𝑆𝛽1S_{\beta}=-1italic_S start_POSTSUBSCRIPT italic_β end_POSTSUBSCRIPT = - 1 and events at all β𝛽\betaitalic_β should be equally rare.

Now, we can estimate a general correction for our calculated rates of full disruption stars to the rate of all disruptions by using Eq.16, for the regime where βminβcβmaxmuch-greater-thansubscript𝛽minsubscript𝛽𝑐much-greater-thansubscript𝛽\beta_{\rm min}\,{\gg}\,\beta_{c}\,{\gg}\,\beta_{\max}italic_β start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT ≫ italic_β start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ≫ italic_β start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT get that full/total TDE rate converges as fFTDE(βmin/βc)Sβsubscript𝑓FTDEsuperscriptsubscript𝛽minsubscript𝛽csubscript𝑆𝛽f_{\rm FTDE}\rightarrow(\beta_{\rm min}\,{/}\,\beta_{\rm c})^{S_{\beta}}italic_f start_POSTSUBSCRIPT roman_FTDE end_POSTSUBSCRIPT → ( italic_β start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT / italic_β start_POSTSUBSCRIPT roman_c end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT italic_S start_POSTSUBSCRIPT italic_β end_POSTSUBSCRIPT end_POSTSUPERSCRIPT. For βmin=0.6subscript𝛽min0.6\beta_{\rm min}=0.6italic_β start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT = 0.6 and M 106Mless-than-or-similar-tosubscript𝑀superscript106subscriptMdirect-productM_{\bullet}\,{\lesssim}\,10^{6}\,\mathrm{M_{\odot}}italic_M start_POSTSUBSCRIPT ∙ end_POSTSUBSCRIPT ≲ 10 start_POSTSUPERSCRIPT 6 end_POSTSUPERSCRIPT roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT yields almost invariably252525This refers to both convection-dominated and radiation-dominated stars since the fraction of full TDE rates is already is over-estimated by a factor of (βc)Sβsuperscriptsubscript𝛽𝑐subscript𝑆𝛽(\beta_{c})^{S_{\beta}}( italic_β start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT italic_S start_POSTSUBSCRIPT italic_β end_POSTSUBSCRIPT end_POSTSUPERSCRIPT when assuming the TDE rates of 0.38 M stars hold for more massive stars PhaseFlow. Also, the weak dependence of βcsubscript𝛽𝑐\beta_{c}italic_β start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT on MBH mass in the regime is neglected. a correction factor of

1/fFTDE[2,4]1subscript𝑓FTDE241/f_{\rm FTDE}\in[2,4]1 / italic_f start_POSTSUBSCRIPT roman_FTDE end_POSTSUBSCRIPT ∈ [ 2 , 4 ]

for Sβ[1,2]subscript𝑆𝛽12S_{\beta}\in[1,2]italic_S start_POSTSUBSCRIPT italic_β end_POSTSUBSCRIPT ∈ [ 1 , 2 ]. This correction could be applied only to young systems relaxing only for the last <100Myrabsent100Myr<100\,\mathrm{Myr}< 100 roman_Myr that have not emptied their loss-cone. However, the physics of circularization could be different for TDEs varying in β𝛽\betaitalic_β, especially for the unknown regime of lower MBH masses (Kıroğlu et al. 2023, M=102 104Msubscript𝑀superscript102superscript104subscriptMdirect-productM_{\bullet}=10^{2}\,{-}\,10^{4}\,\mathrm{M_{\odot}}italic_M start_POSTSUBSCRIPT ∙ end_POSTSUBSCRIPT = 10 start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - 10 start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT) and the rates may be modified accordingly (Wong et al. 2022). These simulations262626Both studies use general relativistic hydrodynamics and initial conditions from the stellar code MESA. show a critical value greater than the one adopted from studies at M= 106Msubscript𝑀superscript106subscriptMdirect-productM_{\bullet}\,=\,10^{6}\,\mathrm{M_{\odot}}italic_M start_POSTSUBSCRIPT ∙ end_POSTSUBSCRIPT = 10 start_POSTSUPERSCRIPT 6 end_POSTSUPERSCRIPT roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT, namely βc 10/ 3>47/ 19.1similar-tosubscript𝛽𝑐1034719.1\beta_{c}\,{\sim}\,10\,{/}\,3\,{>}47\,{/}\,19.1italic_β start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ∼ 10 / 3 > 47 / 19.1, making the fraction of full disruptions even smaller and this correction greater. Furthermore, this correction factor can be a few tens (>1dexabsent1dex>1\,\mathrm{dex}> 1 roman_dex), as demonstrated by the self-consistent work on partial disruption events by Bortolas et al. (2023). By visual inspection, in the optical sample of Yao et al. (2023) there are more than a few light curves that show a re-flaring activity, instead of a monotonic drop of luminosity. Caution should be drawn when comparing model event rates with samples of TDEs. In the case that indeed events with β<1𝛽1\beta<1italic_β < 1 are included in the observational sample, the predictions of our model would be >1absent1>1> 1dex apart from observations and a more thorough discussion should be done on the optimistic assumptions we made when translating PhaseFlow rates to observations (see Sect. 2.3.4)