This site uses cookies. By continuing to use this site you agree to our use of cookies. To find out more, see our Privacy and Cookies policy.

Articles

SHATTERING FLARES DURING CLOSE ENCOUNTERS OF NEUTRON STARS

Published 2013 October 18 © 2013. The American Astronomical Society. All rights reserved.
, , Citation David Tsang 2013 ApJ 777 103 DOI 10.1088/0004-637X/777/2/103

0004-637X/777/2/103

ABSTRACT

We demonstrate that resonant shattering flares can occur during close passages of neutron stars in eccentric or hyperbolic encounters. We provide updated estimates for the rate of close encounters of compact objects in dense stellar environments, which we find are substantially lower than given in previous works. While such occurrences are rare, we show that shattering flares can provide a strong electromagnetic counterpart to the gravitational wave bursts expected from such encounters, allowing triggered searches for these events to occur.

Export citation and abstract BibTeX RIS

1. INTRODUCTION

The major expected source of gravitational waves (GWs) for the advanced LIGO (Harry et al. 2010) class of GW detectors is compact binary systems. The long inspiral signals from such binaries will be detected by matched filtering with theoretical templates, which allows signal-to-noise (S/N) to be built up over many orbits (Cutler & Flanagan 1994).

GWs are also emitted as broad band bursts when compact objects undergo close passages, either during single parabolic or hyperbolic encounters, or during repeated eccentric encounters (Kocsis et al. 2006). Such events occur rarely, but are more likely in dense stellar environments, such as globular clusters or galactic nuclear clusters. The brief duration of such bursts does not allow a large integrated buildup of S/N, and they may be difficult to detect without some electromagnetic trigger.

Recently, Tsang et al. (2012) showed that during binary inspiral of neutron stars (NSs), resonant tidal excitation of the interface mode—a natural mode of a NS peaked at the crust–core boundary—could result in an isotropic resonant shattering flare, and that these were consistent with short gamma-ray burst (sGRB) precursors observed seconds before some sGRBs (Troja et al. 2010). Coincident timing of such precursor flares and the GW inspiral signal can be used to provide strong constraints on the NS equation of state (Tsang et al. 2012).

In this paper, we show that resonant shattering flares can also occur during close passages with other compact objects, such as another NS or a black hole (BH), and that such flares could serve as electromagnetic counterparts to GW bursts, allowing triggered searches for these bursts.

2. TIDAL ENERGY TRANSFER DURING PARABOLIC AND ECCENTRIC ENCOUNTERS

Tidal energy transfer during close encounters can be determined in a Newtonian approximation through the procedure outlined in Press & Teukolsky (1977). While a fully relativistic formulation would be preferable, the Newtonian formulation is sufficiently accurate for a periapse distance much larger than the NS radius; relativistic effects would only slightly modify the frequencies and increase the strength of the interaction.

In general, the energy transfer rate to a star is given by

Equation (1)

where the fluid velocity v ≡ ∂ξ/∂t is the time derivative of the Lagrangian displacement ξ and U is the gravitational potential. To examine the response of a NS with mass M1, it is convenient to decompose the potential due to a star with mass M2 into spherical harmonics Ylm(θ, ϕ),

Equation (2)

Equation (3)

where Φ is the true anomaly of the system, * denotes the complex conjugate, R(t) is the distance between the stars, and (r, θ, ϕ) is the comoving coordinate system centered with M1. Note that for the purposes of mode excitation, we are only concerned with the tidal (l > 2) component of the potential. Assuming the normalization for the spherical harmonics given in Jackson (1999), the constants Wlm are

Equation (4)

where (−)k is defined to be zero when k is a non-integer.

The energy transferred to a particular mode (assuming no nonlinear effects, such as crust fracture) during a periapse passage for a parabolic or highly eccentric encounter can be estimated by Equation (40) of Press & Teukolsky (1977),

Equation (5)

where Rmin is the separation of the stellar centers of mass at the periapse, R1 is the radius of the NS, Qnl is the overlap integral for the NS displacement eigenmode ξnlm, with radial mode number n,

Equation (6)

and

Equation (7)

with

Equation (8)

Here, $\hat{\eta } \equiv [M_1/(M_1 + M_2)]^{1/2} (R_{\rm min}/R_1)^{3/2}$ and $\hat{\omega }_{nlm} \equiv \omega _{nlm} (M_1 + M_2)^{-1/2}R_{\rm min}^{3/2}$ are the Keplerian frequency at the NS surface and the mode frequency, respectively, both scaled by the Keplerian frequency at periapse.

Modes with frequency much higher than periapse Keplerian frequency ($\hat{\omega }_{nlm} \gg 1$) cannot be strongly excited. In contrast, if the periapse distance is too small, then the stars may collide, or a tidal disruption may occur.

We can calculate the energy transfer to the interface mode ΔEi by utilizing the i-mode frequencies and overlap integrals for various equations of state given by Tsang et al. (2012) and summarized in Table 1. If ΔEi > Eb, the mode energy required for the crust to reach the breaking strain, then a shattering flare can occur. ΔEi/Eb for various equations of state are shown as a function of periapse distance Rmin for parabolic encounters between a 1.4 M NS (Figure 1) and a 10 M BH and a 1.4 M NS (Figure 2).

Figure 1.

Figure 1. Top: ratio of the maximum energy transfer through tidal resonance to the interface mode, ΔEi, to the mode energy required to reach the breaking strain, Eb, for a parabolic or highly eccentric 10 M BH–1.4 M NS encounter as a function of periapse distance (Rmin) for various equations of state. Below a critical periapse distance, the mode energy exceeds the breaking energy (ΔEi/Eb > 1) and a resonant shattering flare could occur. The upper axis shows the Keplerian orbital frequency at the periapse, $f_K(R_{\rm min}) \equiv [G(M_1 + M_2)/R_{\rm min}^3]^{1/2}/2\pi$. Bottom: sky and observer inclination averaged signal-to-noise ratio (S/N) for a single advanced LIGO gravitational wave detector for a gravitational wave burst at 50 Mpc from a parabolic 10–1.4 M close encounter with periapse Rmin.

Standard image High-resolution image
Figure 2.

Figure 2. Top: ratio of the maximum energy transfer through tidal resonance to the interface mode, ΔEi, to the mode energy required to reach the breaking strain, Eb, for a parabolic or highly eccentric 1.4 M NS–1.4 M NS encounter as a function of periapse distance (Rmin) for various equations of state. Below a critical periapse distance, the mode energy exceeds the breaking energy (ΔEi/Eb > 1) and a resonant shattering flare could occur. The upper axis shows the Keplerian orbital frequency at the periapse, $ f_K(R_{\rm min}) \equiv [G(M_1 + M_2)/R_{\rm min}^3]^{1/2}/2\pi$. Bottom: sky and observer inclination averaged signal-to-noise ratio (S/N) for a single advanced LIGO gravitational wave detector for a gravitational wave burst at 50 Mpc from a parabolic 1.4–1.4 M close encounter with periapse Rmin.

Standard image High-resolution image

Table 1. Resonant Mode Properties for the l = 2 i-mode for Various Equations of State from Tsang et al. (2012)

EOS R1.4 fi Qi Eb
(km) (Hz) (erg)
SLy4 11.7 188 0.041 5 × 1046
SkI6 12.5 67.3 0.017 3 × 1045
Rs 13.0 32.0 0.059 1 × 1046

Notes. R1.4 is the radius of a 1.4 M NS, fi is the i-mode frequency, Qi is the overlap integral for the i-mode and the tidal field, while Eb is the mode energy required to reach the breaking strain in the crust.

Download table as:  ASCIITypeset image

3. RESONANT SHATTERING

The process that produces a resonant shattering flare is outlined in Figure 3. During a close encounter (or at orbital resonance), energy is extracted from the kinetic energy of the orbit through resonant tidal coupling. The interface mode is excited strongly, which drives the mode to an amplitude at which the breaking strain of the crust is exceeded.

Figure 3.

Figure 3. Schematic of the resonant shattering process. The gravitational potential of the system is the ultimate source of the energy powering the resonant flare. During close passage, or at resonance for circular orbits, tidal resonance transfers energy from the orbit (A) to the i-mode (B) at a rate ≳ 1050 erg s−1. The i-mode grows quickly until the maximum strain at the base of the crust exceeds the breaking strain at mode energy ∼1047 erg. A fracture occurs, releasing ∼1043 erg of low-frequency seismic energy (C) per fracture, however, the mode continues to be driven by the resonance. As more fractures occur, more energy is deposited into seismic energy in the crust. When the total seismic energy in the NS crust exceeds the elastic limit of the curst Eelastic ∼ 1046 erg, the crust shatters, scattering the mode energy and elastic energy to high-frequency oscillations (D). High-frequency oscillations can couple strongly to the magnetic field (Blaes et al. 1989; Thompson & Blaes 1998) by strongly vibrating their footprints (E). Strong perturbations of the magnetic field at the neutron star surface drive strong electric fields, which can accelerate charged particles, triggering pair production and a relativistic fireball with luminosity 1047–1048 erg s−1.

Standard image High-resolution image

The crust fractures, depositing ∼epsilonbμΔr3 ∼ 1043 erg of seismic energy into the crust, where epsilonb ∼ 0.1 is the breaking strain (Horowitz & Kadau 2009), μ ∼ 1030 erg cm−3 is the shear modulus, and Δr ∼ 105 cm is roughly the thickness of the crust. These broad spectrum seismic waves are peaked at characteristic frequency ∼(μ/ρ)1/2/(2πΔr) ∼ 200 Hz, where ρ is the density of the crust. Low-frequency seismic waves cannot couple efficiently to the magnetic field (Blaes et al. 1989), and the energy builds up in the NS crust as the interface mode is driven further and more fractures occur. This seismic energy builds until the crust reaches the elastic limit $E_{\rm elastic} = \int dV \epsilon _{b}^2 \mu \sim 10^{46}$ erg when it shatters, scattering the mode and seismic energy to high-frequency oscillations which can then couple to the magnetic field. Strong perturbations of the magnetic field result in strong transverse electric fields, which can accelerate particles to high energy, sparking a pair-photon fireball. The luminosities of resonant shattering flares are expected to be up to ∼1047–1048 erg s−1 (Tsang et al. 2012) if the precursor flare timescales are assumed.

Troja et al. (2010) found precursors occurring in 3 of the 49 soft gamma repeaters analyzed, implying that not every binary merger should result in a detectable shattering flare. We note that the extraction of seismic energy from the crust by the magnetic field is limited by the strength of the magnetic field at the surface of the NS. The maximum luminosity that can be extracted from the crust by the magnetic field can be estimated by

Equation (9)

where v is the maximum velocity of the perturbation to the field line, R is the NS radius, and Bsurf is the local surface field strength, which can be significantly higher than the large scale dipole field. Thus, only shattering flares from NSs with sufficiently strong surface fields can be detected.

4. ELECTROMAGNETIC COUNTERPARTS TO GRAVITATIONAL WAVE BURSTS

To calculate the expected GW S/N due to parabolic encounters, we follow the procedure outlined in Kocsis et al. (2006). The strain caused by a GW burst due to a parabolic encounter is given by (Flanagan & Hughes 1998)

Equation (10)

where z is the redshift, dL is the luminosity distance, and dE/df is the total GW energy emitted by encounter per unit frequency, which is given for a parabolic (e = 1) encounter in the non-relativistic limit by Equation (46) from Turner (1977). The S/N for a sky- and orientation-averaged signal on a single detector is given by (Dalal et al. 2006; Nissanke et al. 2010)

Equation (11)

where Sn(f) is the spectral noise density for a given detector. In Figures 1 and 2, the S/N is shown for the NS–NS and BH–NS encounters assuming a single encounter at 50 Mpc (z ≃ 0.011) for advanced LIGO, with spectral noise density given by Harry et al. (2010).

Blind detection (S/N ≳ 6 coincident at each detector; see, e.g., Aasi et al. 2013) of a single GW burst from a NS close encounter would be extremely challenging at reasonable distances, with fairly low S/N even for close passages, in particular for NS–NS encounters. Using X-ray or gamma-ray detections of resonant shattering flares as electromagnetic counterparts, triggered GW searches could be performed, significantly lowering the S/N threshold for GW burst detection (Kochanek & Piran 1993; Nissanke et al. 2010; Kelley et al. 2013; Dietz et al. 2013). Networks of detectors can also be used to enhance burst detection through coincident and coherent methods (Schutz 2011; Nissanke et al. 2013; Aasi et al. 2013).

Kocsis & Levin (2012) also show that repeated GW bursts from eccentric captures can be combined with the final chirp to boost the integrated S/N by roughly an order of magnitude, and would optimistically allow detection of bursts from BH–NS eccentric captures out to ∼300 Mpc, and NS–NS encounters to ∼150 Mpc. The pattern of these repeated bursts can be modeled for given orbital parameters. Resonant shattering flares can be seen significantly farther than the GW bursts. If they occur for a given system, then they will happen for sufficiently close passages, which are also those that contribute the largest component of the GW burst signal. If repeated flares are seen, then these could also be used to characterize the orbit and target a burst search to accumulate S/N over multiple passages. However, significant changes to the current GW templates may be necessary to detect eccentric captures and mergers (East et al. 2013; Huerta & Brown 2013).

5. EVENT RATES

Close encounters of NSs with other compact objects are much more likely to occur in dense stellar environments, such as globular clusters and galactic nuclei. While it is beyond the scope of this paper to perform an extremely detailed evaluation of the event rates for close encounters of compact objects, we will briefly discuss the event rates for such encounters in both of these environments and provide updated estimates for some of the rates in the literature.

5.1. Globular Clusters

Kocsis et al. (2006) calculated the parabolic encounter rate for compact objects in globular clusters using simplified globular cluster models, predicting a rate of ≳ 1 detection per year for advanced LIGO in optimistic scenarios. However, their detection rates are dominated by rare distant events involving close encounters of ≳ 20 M BHs.

Lee et al. (2010) examine various dynamical pathways to sGRBs in globular clusters, including binary interactions and tidal capture. They calculate a high rate of close encounters for two NSs, $\Gamma _{\rm NS-NS}^{({\rm GC})} \sim 55$ yr−1 Gpc−3, using as a calibration the estimate of ∼104 NSs in the collapsed core of M15, from the Fokker–Planck calculations of Dull et al. (1997). This would require an extremely high NS retention fraction. Subsequent more careful calculations by Murphy et al. (2011) have determined the number of NSs in the core of M15 to be closer to ∼103, which is consistent with ∼1%–10% retention fraction estimates from pulsar kick velocities (Drukier 1996; Hansen & Phinney 1997; Davies & Hansen 1998). This reduces the estimates of Lee et al. (2010) to $\Gamma _{\rm NS-NS}^{({\rm GC})} \sim 0.5$ yr−1 Gpc−3, however, Samsing et al. (2013) have recently showed that considering chaotic resonances in binary–single interactions can significantly increase the expected rate of eccentric binaries.

5.2. Galactic Nuclei

While there are many globular clusters per galaxy, high kick velocities at NS birth significantly lower their retention fraction. The deeper potentials of galactic nuclei may provide dense stellar environments where the NS retention fraction is higher and close encounters are more likely to occur. O'Leary et al. (2009) and Kocsis & Levin (2012) both provide estimates of ∼1–100 BH close passages per year within a few Gpc detectable by advanced LIGO, with GW-detectable NS– BH encounters estimated to be ∼1% of this. However, as we discuss in Appendix A below, they scale by a factor ξ, representing the contribution due to the variance of the nuclear cluster density across galaxies. They take this factor to be ξ ≳ 30–100, but we find that ξ is more correctly evaluated to be significantly lower, even with the most optimistic assumptions.

In Appendix A, we have re-evaluated the rates for single–single eccentric captures of compact objects in nuclear star clusters containing massive central BHs, assuming a simplified isothermal density distribution, as in Kocsis & Levin (2012). We find that the 10 M BH–BH eccentric capture encounter rate is $\Gamma _{\rm BH-BH}^{({\rm GN}, {\rm EC})}\sim 0.02$ yr−1 Gpc−3, which is significantly lower than the previously estimated values. However, a more top-heavy initial mass function (IMF; Bartko et al. 2010), along with enhanced segregation and spatial flattening for heavier BHs, may help to increase this value.

The NS–NS and NS–BH (10 M) rates can be estimated in a similar fashion for an isothermal ∝r−2 density distribution to be $\Gamma _{\rm NS-NS}^{({\rm GN},{\rm EC})} \sim 0.04\hbox{--}6$ yr−1 Gpc−3, and $\Gamma _{\rm NS-BH}^{({\rm GN},{\rm EC})} \sim 0.05\hbox{--} 0.6$ yr−1 Gpc−3 with the range mainly due to uncertainty in the IMF and mass loss models for NS progenitors (O'Connor & Ott 2011). However, there is reason to suspect that the slope of the density distribution is somewhat flattened due to interaction with segregated BHs (see, e.g., O'Leary et al. 2009 and references therein). Taking a NS density distribution ∝r−3/2 as a lower bound for our rates, we find in this case $\Gamma ^{({\rm GN},{\rm EC})}_{\rm NS-NS} \simeq 0.003\hbox{--} 0.3$ yr−1 Gpc−3.

Note that in evaluating the above rates, we have made very optimistic assumptions about systematic versus intrinsic variation of the relevant observations, as in O'Leary et al. (2009). Possibly more realistic estimates for this intrinsic scatter reduce these rates by a factor of ∼4.

In the high density, high relative-velocity region near the center of nuclear star clusters, there may be a significant rate of hyperbolic passages where the periapse is sufficiently close to trigger a shattering flare, but insufficient to result in eccentric capture by GW emission. In Appendix B, we have calculated the rate of encounters that result in a shattering flare during the first close passage, and find this to be higher than the eccentric capture rate for the fiducial model used. We find, for the optimistic assumptions about intrinsic variation used above, and assuming our canonical isothermal model, $\Gamma ^{({\rm GN}, {\rm SF})}_{\rm NS-NS} \simeq 0.2\hbox{--} 60$ yr−1 Gpc−3. For a more flattened density profile ∝r−3/2, we have $\Gamma ^{({\rm GN}, {\rm SF})}_{\rm NS-NS} \simeq 0.005\hbox{--}0.5$ yr−1. Taking our less generous estimates for the intrinsic scatter across galaxies significantly reduces these rates by a factor of ∼6.

5.3. Other Possible Event Rate Contributions

In the above discussion, we have primarily considered single–single interactions in determining the event rates in dense clusters. Binary–single or binary–binary interactions have larger cross sections and could increase the rates for such events (see, e.g., Miller & Lauburg 2009). Recently, Katz & Dong (2012) and Kushnir et al. (2013) demonstrated that Kozai–Lidov type interactions can drive the inner binaries of hierarchical triple systems toward extreme eccentricity, with collisions occurring when the periapse distance is driven below the stellar radius. They claim that the rates for such Kozai-oscillation driven collisions between white dwarfs in field triple star systems can be comparable to the SN Type Ia rate. Similar interactions could potentially drive close encounters of NSs or BHs in triple systems. However, the periapse for shattering flares is two orders of magnitude smaller than those considered by Katz & Dong (2012), and NSs and BHs are substantially more rare than white dwarfs, particularly outside of dense clusters. Within clusters, such triple systems would need to reach these extreme eccentricities quickly, before other encounters ionize away the softer less-bound outer companion.

6. DISCUSSION

We have calculated the energy transfer to the interface mode through tidal interaction for NSs during close encounters with other compact objects, and we have shown that resonant shattering flares can occur during parabolic or eccentric encounters if the periapse is sufficiently close and the local surface field of the NS is sufficiently high. Such flares are similar to resonant shattering flares during binary inspirals, and should have luminosity ∼1047–1048 erg s−1 (Tsang et al. 2012).

Broad band GW bursts are also generated by such encounters. While they are rare, there is intense interest in such GW burst events which are detectable by the next generation of GW detectors. GW bursts with high S/N are also those for which shattering flares may occur and act as an electromagnetic counterpart to trigger burst searches. Highly eccentric captures of NSs in dense stellar environments are expected to result in repeated GW bursts (Kocsis & Levin 2012), but may also lead to repeated shattering flares at each periapse passage. Radioactive emission from ejecta (so-called kilonova; Metzger et al. 2010) and radio flares is also expected to be stronger for eccentric mergers than for their circularized counterparts (East & Pretorius 2012), which may allow easier identification and localization of a shattering flare through coincident detections.

We have also reviewed and updated the rates for compact object encounters presented in the literature for dense stellar environments, and find that these rates should be revised significantly downward. While these estimated rates for close encounters involving NSs within the horizon of advanced LIGO are low, shattering flares from such encounters can be detected significantly farther, and may occur at rates of up to $\Gamma ^{({\rm GN}, {\rm SF})}_{\rm NS-NS} \simeq 0.2\hbox{--}60$ yr−1 Gpc−3, subject to large model uncertainties. More conservative assumptions substantially lower this rate.

This research was supported by generous funding from the Lorne Trottier Chair in Astrophysics and Cosmology as well as the Canadian Institute for Advanced Research, and is brought to you by the letter ξ. I would like to thank Bence Kocsis, Ryan O'Leary, and Cole Miller for detailed and enlightening discussions involving the rates, as well as Samaya Nissanke, Tony Piro, Enrico Ramirez-Ruiz, Kostas Gourgouliatos, Andrew Cumming, Sterl Phinney, and Gil Holder for insightful conversation.

APPENDIX A: ECCENTRIC CAPTURE RATES IN NUCLEAR STAR CLUSTERS

Here, we will carefully estimate the eccentric capture rate of compact objects in nuclear star clusters. We begin by following the general procedure outlined in Appendix C of Kocsis & Levin (2012), and calculate the eccentric capture rate for a single fiducial galaxy.

A.1. The Rate for a Single Galaxy

The cross section for eccentric capture is basically the cross section for which the energy emitted by GW (or lost due to tidal interactions) during a close encounter exceeds the kinetic energy of the objects at infinity. The maximum periapse for capture is given by Quinlan & Shapiro (1989) as

Equation (A1)

where η is the symmetric mass ratio, mtot is the total mass of the two objects, and vrel is the relative speed of the two objects at infinity. We only consider GW capture here since, for compact objects, the tidal capture cross section is much smaller than the GW capture cross section.

We can then calculate the cross section, using the standard formula for gravitational focusing

Equation (A2)

where we have scaled this to fiducial values that will be used later. The rate of eccentric captures for a single galaxy with a central supermassive black hole (SMBH) of mass MS and velocity dispersion σdisp(MS) is then

Equation (A3)

where n1 and n2 are the number densities of each type of object as a function of radius, $r_i \equiv {GM}_S{/}\sigma _{\rm disp}^2$ is the radius of influence of the BH, and rmin is approximately the radius inside which there is only a single object.

We take the density of objects to be the same as in an isothermal distribution, and the velocity distribution to be Maxwellian at each radius with relative speed $\langle v_{\rm rel}^2\rangle = 2 v_{\rm circ}^2 \equiv 2{GM}_S/r = 2\sigma _{\rm disp}^2 (r/r_i)^{-1}$, which captures the behavior well near the central BH, where the contribution to the rate is the largest. The number densities are then given by

Equation (A4)

where rdyn defines the dynamical radius inside which twice the mass of the central BH is contained such that Ni ≡ 2κiMS/mi and κi are the number of objects and total mass fraction of type i = 1, 2 within rdyn, respectively. We take these scalings of n(r) and vrel(r) for simplicity to highlight the sources of uncertainty and provide a rate estimate for the simplest case. For how different and more realistic scalings may alter the basic single galaxy rate, see the detailed discussion in O'Leary et al. (2009) and Appendix C of Kocsis & Levin (2012).

It is also convenient to define the geometric means of the number $\tilde{N}$ and mass fraction $\tilde{\kappa }$, such that

Equation (A5)

as well as the fiducial scalings σ84 ≡ σdisp/(84 km s−1), η0.25 ≡ η/0.25, and m20mtot/(20 M). Approximating $r_{\rm min} \simeq \tilde{N}^{-1} r_{\rm dyn}$, we find

Equation (A6)

Equation (A7)

where we have averaged over the Maxwellian distribution, $\langle v_{\rm rel}(r_i)^{-11/7} \rangle = 1.15 \times \sigma _{\rm disp}^{-11/7}$, and changed integration variables to xr/rdyn. Assuming that $(\tilde{N}r_i/r_{\rm dyn})^{3/14} \gg 1$, we have

Equation (A8)

where M4e6MS/(4 × 106M) is scaled to the Milky Way, and $\tilde{\kappa }_{2.5} \equiv \tilde{\kappa }/(2.5\%)$ as in Kocsis & Levin (2012). Applying the M–σ relation, $M_{4e6} =\sigma _{84}^4$ (Tremaine et al. 2002), we finally have

Equation (A9)

Thus far, this agrees relatively well with Kocsis & Levin (2012) and O'Leary et al. (2009).

A.2. Averaging Over Many Galaxies

From the scatter in the inferred nuclear star cluster relaxation time Tr for the low σdisp galaxies given in Figure 1 of Merritt et al. (2007), O'Leary et al. (2009) and Kocsis & Levin (2012) claim that the variance of the central number density scales the average rate per galaxy by a factor $\xi = \overline{n^2}/\overline{n}^2 \gtrsim 30$, increasing their total rate substantially.

Here, in calculating the average rate over many galaxies, we will carefully consider the effect of variation in both the M–σ relation and the scaling of the central density implied by Merritt et al. (2007)—related to the parameter ri/rdyn—and show that such a substantial increase in the inferred rate is not warranted.

Variation in the M–σ relation. We begin by taking the generous assumption that the intrinsic scatter in the M–σ relation is ∼0.5 dex in MS (Tremaine et al. 2002). We take the distribution of MS for a fixed σdisp to be log-normal such that $M_{4e6} = C_{M\sigma }\times \sigma _{84}^4$, where CMσ is a random variable with log-normal probability distribution with geometric mean 〈CMσ〉 = 1 and scale factor $\delta _{M\sigma } = \ln \sqrt{10}$.

Our single galaxy rate (Equation (A8)) has scaling such that for a fixed SMBH mass bin, $\Gamma _{\rm gal}(M_S, C_{M\sigma }) \propto C_{M\sigma }^{-31/28}$. Using the properties of log-normal random variables, the scaling to the average rate over this distribution is given as

Equation (A10)

where $\overline{f}$ denotes averaging of f over the distribution.

Variation in ri/rdyn. The ratio of the radius of influence, ri, to the dynamical radius, rdyn, determines the relative density of the nuclear cluster, and varies for different N-body models from ∼0.1 to 1. (Binney & Tremaine 2008). For simplicity we will take this ratio to be a log-normal distributed random variable independent of CMσ.

Figure 1 of Merritt et al. (2007) showed an estimate of the relaxation time at the dynamical radius as a function of the velocity dispersion σdisp for nuclei of early-type galaxies in the ACS Virgo Cluster Survey (Côté et al. 2004). While the majority of the scatter in this distribution is for low-luminosity unresolved nuclear star clusters, an indication that much of this scatter may be due to observational uncertainty, for the purposes of this discussion we will assume, as in O'Leary et al. (2009), that this scatter is intrinsic. We will again assume, for simplicity, that the relaxation time Tr for fixed σdisp is distributed log-normally, with a generous estimate of 1.5 dex standard deviation for the scatter in log Tr, such that the scale factor is $\delta _{T_r} \simeq 1.5 \ln 10$.

The nuclear relaxation time is given by

Equation (A11)

(Spitzer 1987), where ln Λ is the Coulomb logarithm. For fixed σdisp we can rewrite the Tr in terms of the random variables CMσ and ri/rdyn:

Equation (A12)

For independent log-normal random variables CMσ and ri/rdyn, this gives

Equation (A13)

where δdyn is the standard deviation of ln ri/rdyn. This then gives the scaling due to variation in ri/rdyn to the average rate of

Equation (A14)

Final rate. The central BH mass function can be estimated by

Equation (A15)

(Shankar et al. 2004), where α ≃ −1.11, β ≃ 0.5, and M* ≃ 6.4 × 107M, assuming the local Hubble constant Ho = 70 km s−1 Mpc−1.

Integrating our mass dependence $M_{4e6}^{9/28}$ over the mass function, we can obtain the effective density

Equation (A16)

with which to multiply our single galaxy rate evaluated at M4e6 = 1. The mass function (Equation (A15)) is only constructed to be valid between 106MMS ≲ 5 × 109M, however, we expect significant contribution from smaller galaxies. The halo mass function increases for lower mass, however, for dwarf galaxies the stellar mass (and therefore SMBH mass) to halo mass ratio drops significantly and the expected number density at that mass should also fall. If we integrate down to only MS, min ∼ 106M, then this gives us an effective density ngal, eff ≃ 0.043 Mpc−3, while extending this mass function down to a cutoff of MS, min ∼ 104M yields ngal, eff ≃ 0.067 Mpc−3. With this in mind, we take the fiducial value of the effective density to be ngal, eff = ngal, 5 × 0.05 Mpc−3.

Scaling to the Milky Way where O'Leary et al. (2009) assume a fiducial value of ri/rdyn ≃ 0.5, we have our final rate of eccentric captures in galactic nuclei,

Equation (A17)

Equation (A18)

within d2 Gpc × 2 Gpc, where we take 2 Gpc for the fiducial value as it is roughly the advanced LIGO horizon distance for 10 M BH–BH eccentric captures. Here, we have also included an additional factor (H/r)−2, which may increase the density for nuclear clusters where significant flattening has occurred (B. Kocsis 2013, private communication). Assuming the generous fiducial values for ξMσ and ξdyn, and no significant flattening, the rate for eccentric capture of 10 M BH–BH encounters is

Equation (A19)

Less generous estimates for the intrinsic scatter, δMσ ≃ 0.3ln 10 (Tremaine et al. 2002) and $\delta _{T_r} \simeq \ln 10$ give ξMσ ≃ 1.33 and ξdyn = 2.52, which reduces the above rate by a factor of ∼4.

A.3. Neutron Star Rates

We are now (finally) ready to estimate the rates for NS eccentric captures in galactic nuclei. There is large uncertainty in the NS production rate in nuclear star clusters, mostly due to two factors. First, the IMF is unknown and could range from the standard Salpeter IMF to an extremely top-heavy IMF (e.g., Bartko et al. 2010). Second, there is great uncertainty in the effects of mass-loss for determining the fraction of stars with MZAMS ≳ 8 M that will become NSs, and the fraction that will collapse to form BHs (see, e.g., O'Connor & Ott 2011 for discussion).

With these considerations, we will assume that between ∼1% and 10% of the stars in a nuclear star cluster will become NSs. The low end of this range is for a standard IMF, with low mass loss and metallicity, while the high end roughly corresponds to a top-heavy IMF with more mass loss in the NS progenitors. Assuming an average stellar mass in nuclear star clusters of ∼0.5 M, this gives κNS ∼ 0.03–0.3.

NSs have only had time to segregate in galaxies for which σdisp ≲ 50 km s−1 (Miller & Lauburg 2009). Unfortunately, these low mass galaxies also have escape velocities ∼2σdisp, and thus NS kick velocities of ∼100 km s−1 will significantly reduce the retention fraction after NS formation, much like in globular clusters. So we will assume the scaling above with no significant enhancements due to segregation or flattening.

Substituting mtot = 2.8 M, η = 0.25, and κNS = 0.03–0.3 into (A18) and continuing to use the very generous assumptions above for ξMσ ≃ 2.25 and ξdyn ≃ 6.1, we get a rate for eccentric capture of

Equation (A20)

For 10 M BH–1.4 M NS encounters we take mtot = 11.4 M, η ≃ 0.11, κBH ≃ 0.025, and κNS ≃ 0.03–0.3, and get a rate

Equation (A21)

However, in systems where the BHs dominate the core, a cusp of more massive objects tends to flatten out the distribution of lighter objects compared to an isothermal density profile. O'Leary et al. (2009) find that the distribution can approach a power law index of 1.5 for the lighter objects in their Fokker–Planck calculation. Repeating the above calculations for nNSr−3/2 as a lower bound, we find

Equation (A22)

noting that the rate in a single galaxy is no longer dominated by the contribution due to the innermost objects.

APPENDIX B: RATES FOR SHATTERING DURING FIRST PASSAGE

Eccentric captures, described above, provide multiple close passages, however, the cross section for eccentric capture only exceeds that for shattering flares from the first passage below vrel ∼ 1000 km s−1. Near the center of nuclear star clusters—where the density and relative velocity dispersion are the highest—the rate for hyperbolic passages that result in shattering flares but are not bound through GW emission can be significant. In this Appendix, we calculate this rate for shattering flares during first passage.

We begin by assuming that the speed at infinite separation is small compared to the speed at periapse $v_{\rm rel} \ll \sqrt{2G m_{\rm tot}/r_p}$. The maximum periapse distance for shattering is then only a function of the stellar masses, and the equation of state

Equation (B1)

which can be determined through setting ΔEi = Eb as in Section 2 of the main text above. We can then calculate the single galaxy rate as for eccentric captures, giving

Equation (B2)

Equation (B3)

where we have used the M − σ relation $M_{4e6} = C_{M\sigma } \sigma _{84}^4$ and the fact that $\langle v_{\rm rel}^{-1} \rangle \simeq 1.38 \times \langle v_{\rm rel} \rangle ^{-1}$ over a Maxwellian distribution.

Again, for the same extremely generous assumptions for the intrinsic variation in M–σ and Tr(σ) as we did above, we can evaluate

Equation (B4)

To calculate the effective density, we take

Equation (B5)

which gives us our rate for shattering flare encounters of

Equation (B6)

For κNS ≃ 0.03–0.3 we then have the range

Equation (B7)

for the fiducial values above, including our optimistic estimates for ξMσ ≃ 2.82 and ξdyn ≃ 9.99.

Similar to the case for eccentric captures, less generous estimates for the intrinsic scatter significantly reduce these rates. Taking δMσ ≃ 0.3ln 10 and $\delta _{T_r} \simeq \ln 10$, we have ξMσ ≃ 1.45 and ξdyn ≃ 3.24, which reduces the above rates by a factor of ∼6.

If we repeat the above calculations for the flattened distribution nNSr−3/2, then we obtain a reduced rate estimate of

Equation (B8)
Please wait… references are loading.
10.1088/0004-637X/777/2/103