<<

Draft version June 12, 2019 Typeset using LATEX twocolumn style in AASTeX62

Explosions Driven by the Coalescence of a Compact Object with the Core of a Massive- Companion Inside a Common Envelope: Circumstellar Properties, Light Curves, and Population Statistics

Sophie Lund Schrøder,1, 2 Morgan MacLeod,2, ∗ Abraham Loeb,2 Alejandro Vigna-Gomez´ ,3, 4, 5 and Ilya Mandel4, 5, 3

1DARK, Niels Bohr Institute, University of Copenhagen, Blegdamsvej 17, 2100 Copenhagen, Denmark 2Harvard-Smithsonian Center for Astrophysics, 60 Garden Street, Cambridge, MA, 02138, USA 3Birmingham Institute for and School of Physics and Astronomy, University of Birmingham,Birmingham, B15 2TT, United Kingdom 4Monash Centre for Astrophysics, School of Physics and Astronomy, Monash University, Clayton, Victoria 3800, Australia 5The ARC Center of Excellence for Gravitational Wave Discovery – OzGrav

ABSTRACT We model explosions driven by the coalescence of a black hole or with the core of its massive-star companion. Upon entering a common envelope phase, a compact object may spiral all the way to the core. The concurrent release of energy is likely to be deposited into the surrounding common envelope, powering a merger-driven explosion. We use hydrodynamic models of binary coalescence to model the common envelope density distribution at the time of coalescence. We find toroidal profiles of material, concentrated in the binary’s equatorial plane and extending to many times the massive star’s original radius. We use the spherically-averaged properties of this circumstellar material (CSM) to estimate the emergent light curves that result from the interaction between the blast wave and the CSM. We find that typical merger-driven explosions are brightened by up to three magnitudes by CSM interaction. From population synthesis models we discover that the brightest merger-driven explosions, MV ∼ −18 to −19, are those involving black holes because they have the most massive and extended CSM. Black hole coalescence events are also common; they represent about 50% of all merger-driven explosions and approximately 0.3% of the core-collapse rate. Merger-driven explosions offer a window into the highly-uncertain physics of common envelope interactions in binary systems by probing the properties of systems that merge rather than eject their envelopes.

Keywords: binaries: close, methods: numerical

1. INTRODUCTION immersed within the envelope of the massive star and Binary and multiple systems are ubiquitous among spirals closer to the massive star’s core (e.g. Taam et al. massive . Of these systems, a large fraction are at 1978; Iben & Livio 1993; Terman et al. 1995; Taam & separations so close that they will interact over the stars’ Sandquist 2000; Armitage & Livio 2000; Ivanova et al. lifetimes (Sana et al. 2012; de Mink et al. 2014; Moe & Di 2013; De Marco & Izzard 2017). Stefano 2017). As these multiple-star systems evolve to Common envelope interactions can lead to either the leave behind compact object stellar remnants, the stage ejection of the shared, gaseous envelope and a surviving, arXiv:1906.04189v1 [astro-ph.HE] 10 Jun 2019 is set for interactions between the evolving stars and the binary pair or to the merger of the donor star core with stellar remnants. In some cases, a phase of escalating, the companion. The distinction between these cases for unstable mass transfer from a massive star donor onto a a given binary remains highly uncertain, but is of signif- compact-object companion leads to a common envelope icant interest given its importance (e.g. Belczynski et al. phase (Paczynski 1976), in which the compact object is 2002; Kalogera et al. 2004, 2007; Belczynski et al. 2008; Andrews et al. 2015; Tauris et al. 2017; Vigna-G´omez et al. 2018), for example, to estimating rates of compact Corresponding author: Sophie Lund Schrøder object mergers and associated gravitational-wave tran- [email protected] sients (Abbott et al. 2016a,b; Abbott 2017a,b,c; The ∗ NASA Einstein Fellow LIGO Scientific Collaboration & the Virgo Collabora- 2 tion 2018a,b). In the case of massive stars interacting pernova for many years (Chugai 1997, 2001; Chugai & with lower-mass, compact object companions, the the- Danziger 1994; Chugai et al. 2004; Smith & McCray oretical expectation is that only donor stars with the 2007; Chevalier & Irwin 2011; Moriya et al. 2013a; Ofek most-extended, weakly bound hydrogen envelopes are et al. 2013b; Dessart et al. 2015; Morozova et al. 2017), susceptible to ejection, while the remainder of systems and even a population of objects that appear to tran- are likely to merge (Kruckow et al. 2016). sition from type I to type II (Margutti et al. 2017). What happens when a compact object merges with Specifically type IIn, with their narrow line features, are the helium core of a massive star? At least two possible known to be interacting with a slow moving CSM (Kiewe outcomes have been suggested. Perhaps a neutron star et al. 2012; Taddia et al. 2013). But the origin of the embedded in a stellar envelope could burn stably, form- material around the pre-SN star is not yet clear. Ideas ing a Thorne-Zytkow object (Thorne & Zytkow 1977; include late stage stellar winds or small outbursts of Podsiadlowski et al. 1995; Podsiadlowski 2007; Levesque gas (Quataert & Shiode 2012; Shiode & Quataert 2014; et al. 2014; Moriya 2018). Alternatively, it is possible Fuller 2017) or formations of a disk-like torus (Andrews for the angular momentum of the merger to lead to the & Smith 2018; McDowell et al. 2018). With this pa- formation of a disk around the compact object (be it per we add to the calculations by Chevalier(2012) the a neutron star or a black hole), from which material light curves expected from an engine-driven explosion forms a rapidly accreting neutrino-cooled disk (Houck & inside a merger ejecta profile. We show how the atyp- Chevalier 1991; Chevalier 1993, 1996; Fryer et al. 1996; ical CSM distribution leads to light curves powered in Fryer & Woosley 1998; Popham et al. 1999; MacFadyen part by CSM interaction, resembling type IIn and IIL & Woosley 1999; Fryer et al. 1999; MacFadyen et al. and lacking a plateau phase (e.g. Das & Ray 2017; Mo- 2001; Zhang & Fryer 2001; Lee & Ramirez-Ruiz 2006; rozova & Stone 2018; Eldridge et al. 2018). Chen & Beloborodov 2007; Barkov & Komissarov 2011; We briefly review the engine-driven explosion model Chevalier 2012; Fryer et al. 2014; Song & Liu 2019). in2. We then self-consistently model the circum-merger of the surrounding core material liberates on density distribution from the common envelope phase 2 54 the order of ηM c ∼ η10 erg, where η is an effi- using a three-dimensional hydrodynamic simulation in ciency factor of order 0.1 (Frank et al. 2002). Much of Section3. To explore the impact of this CSM on the this energy emerges in neutrinos, which stream freely resultant engine-driven explosions, we produce a num- away from the accretion object (Chevalier 1993; Fryer ber of spherically-symmetric radiative transfer models of et al. 1996; Zhang & Fryer 2001). A fraction, however, blast waves interacting with the (spherically-averaged) emerges as Poynting flux or mechanical energy (either common-envelope ejecta profiles in Section4. Next, a disk wind or collimated outflow) and can feed back we map the expected populations of compact object– on the surroundings, powering a blast wave with energy core mergers and their resultant transients in Section similar to a (Fryer & Woosley 1998; Zhang & 5. Finally, in Section6 we study the imprint of CSM Fryer 2001; Chevalier 2012; Dexter & Kasen 2013; Fryer on merger-driven explosions, compare to known super- et al. 2014; Song & Liu 2019; Gilkis et al. 2019; Soker novae, and discuss possible identification strategies. In 2019; Soker et al. 2019). Section7 we conclude. Chevalier(2012) observed that, in the case that the merger was initiated by a preceding common envelope 2. MERGER-DRIVEN EXPLOSIONS phase, the blast wave necessarily interacts with the 2.1. Inspiral, Merger, and Central Engine dense surrounding medium of the common envelope Following its inspiral through the common envelope, ejecta (Soker & Gilkis 2018; Soker 2019; Gilkis et al. a compact object can tidally disrupt and merge with 2019; Soker et al. 2019). The resulting transient is de- the helium core of the massive star (Zhang & Fryer scribed as a ”common envelope jets supernova” by Soker 2001; Chevalier 2012; Soker & Gilkis 2018; Soker et al. et al.(2019). One of the key points that Chevalier(2012) 2019). The disrupted material forms an and Soker et al.(2019) mention is that the distribution surrounding the now-central compact object. The local of common envelope ejecta is crucial in shaping the ob- densities of more than 102 g cm−3 imply that accretion served light curve (see Kleiser et al. 2018, for a similar occuring on a dynamical timescale is very rapid of the discussion in the context of rapidly-fading supernovae). order of 10−3 to 10−1M s−1 (Fryer & Woosley 1998; Here, we pursue this line of examination. Zhang & Fryer 2001), making these conditions very sim- Light curves from supernova explosions of single stars ilar to those of a classical collapsar scenario of rapid ac- within a dense circumstellar material (CSM) have been cretion onto a newly formed black hole (MacFadyen & considered as the origin of the variations in type II su- Woosley 1999; MacFadyen et al. 2001; Siegel et al. 2018). 3

Neutrinos mediate the bulk of the accretion siadlowski et al. 1995; Podsiadlowski 2007), potentially at these accretion rates (MacFadyen & Woosley 1999), showing unique surface features (Levesque et al. 2014). so accretion can occur onto either neutron stars or black If the core of the helium star is near its original den- holes under these conditions (Fryer et al. 1996). At sity when neutron star enters it, the flow convergence higher-still accretion rates, neutrino energy deposition rate is high enough that it seems very difficult to avoid can overturn the accretion flow (e.g. Pejcha & Thomp- a neutrino-cooled accretion state (Chevalier 1993; Fryer son 2012). This rapid accretion, over approximately the & Woosley 1998). However, it is interesting to note that dynamical time of the core (Zhang & Fryer 2001), 103 s, because the neutron star enters the core from the out- will lead a neutron star companion object to quickly side in (rather than inside-out as in a collapsar) there collapse to a black hole. is a possibility that feedback from lower accretion-rate These conditions of hypercritical accretion from the common envelope inspiral (e.g. Fryer et al. 1996) would core onto an embedded black hole set the stage for an prevent the object from ever reaching the hypercriti- explosion powered by this central, accreting engine. In cal, neutrino-cooled accreting branch (see, for example, the context of collapsing helium stars stripped of their the trajectories of infall and accretion in MacLeod & hydrogen envelopes, the result is a long gamma ray burst Ramirez-Ruiz 2015; Holgado et al. 2018). Thus, more (GRB), in which a relativistic jet, perhaps powered by work is needed to conclusively distinguish between these the coupling between the magnetic field in the accretion alternatives. disk coupling to the rotational energy of the black hole (via the Blandford & Znajek(1977) process (e.g. Barkov 2.2. Model Adopted: Engine Mass and Energetics & Komissarov 2008)) tunnels out of the helium star Before merger the donor star consists of a core of mass and is accelerated to high Lorentz factor (MacFadyen Mcore and an envelope of mass Menv. The total mass & Woosley 1999; MacFadyen et al. 2001). of the star is M∗ = Mcore + Menv. The compact-object In the context of a helium core surrounded by an ex- companion that ends up in the center of the donor’s core tended hydrogen envelope, the beamed power of the jet has mass M2 = qM∗. When the two stars merge, M2 is not expected to be able to tunnel out of the enve- accretes a fraction, facc, of Mcore. This either causes it lope under most circumstances. This is because the jet to grow if it is already a black hole, or, if it is a neutron head must displace the ambient stellar material and, star, collapse to a black hole, then grow to a final mass, therefore, expands at a rate which balances the ram pressure of this interaction with the momentum flux of MBH,final = M2 + Macc = M2 + faccMcore. (1) the jet (e.g. Matzner 2003). Following the estimates The released energy from the accretion is, therefore, of Quataert & Kasen(2012)’s Section 2.2, a typical jet ∆E = ηM c2, where η ∼ 0.1. Much of this en- whose working surface expands at a few percent the acc acc ergy is radiated by neutrino emission. However, some speed of light (as it displaces the surrounding stellar energy emerges mechanically, from a magnetohydrody- envelope) reaches only approximately 0.03c × 103s ∼ namic disk wind (e.g. Feng et al. 2018). The mechanical 1012 cm before the core-accretion event ends and the jet power is an uncertain fraction of the accretion energy, shuts off. Lacking the pressure to continue driving its such that ∆E = f ∆E . Dexter & Kasen(2013) expansion, the jet is choked by the surrounding gas dis- mec mec acc estimate f ∼ 10−3 given an inflow-outflow model in tribution, and expands laterally, distributing its power mec which the accreting material decreases as a power law more isotropically into the envelope material and power- with radius, implying ing an outburst (e.g. Murguia-Berthier et al. 2014; Senno     et al. 2016). We note that Quataert & Kasen(2012); 50 fmec  η  Macc ∆Emec ∼ 3 × 10 −3 erg. (2) Woosley & Heger(2012); Dexter & Kasen(2013) dis- 10 0.1 2M cuss an alternate scenario in which the jet might escape if long-lived accretion from the envelope material per- The energy that emerges in a jet via the Blandford & sists over a duration of hundreds of days. The typical Znajek(1977) process can be estimated as (Zhang & energies range from 1050 − 1052 erg (Fryer & Woosley Fryer 2001), 1998; Zhang & Fryer 2001), as we discuss in Section 2.2. M 2  B 2  t  An alternative process of stable burning has been sug- 52 2 BH acc ∆EBZ ∼ 10 a 15 2 erg, gested when the engulfed object is a neutron star. In 3M 10 G 10 s (3) these cases a Thorne-Zytkow Object is said to form – a where B is the magnetic field in the disk, t is the star with a neutron-star core, which might stably burn acc timescale of accretion, and a is the black hole’s dimen- hydrogen for up to 105 yr (Thorne & Zytkow 1977; Pod- sionless spin. Because Macc & MBH, values of order 4

unity for a are expected regardless of the initial spin aRL 1500 state. In what follows, we will assume that the com- ) R ( R n pact object accretes the entire core mass (facc = 1.0) o 1000 i t a

50 r a

and we explore explosion energies between 3 × 10 erg p e and 1052 erg. We further assume that regardless of the s 500 M = 30M R = 1000R precise mechanism or energetics, the energy is shared 50 40 30 20 10 0 roughly spherically with the hydrogen envelope (Cheva- time, t t1 (years) lier 2012; Soker & Gilkis 2018; Soker 2019). 6 4 a = 1500.0R a = 1000.0R 3. CIRCUMSTELLAR MATERIAL EXPELLED

DURING BINARY COALESCENCE ) R 3

As a basis for our analysis, we model binary coales- 0 0 8 1 (

cence and the circumbinary ejecta that results from the y

merger of two example mass ratio binaries. Here we ] 3 m

describe our numerical method, the unstable mass ex- 4 c

g [ change that leads the binary to merge, and the CSM 10 ) (

a = 500.0R a = 100.0R 0 4 1 mass distribution that this runaway mass transfer cre- g o ates. The distribution of CSM at large scales, r ∼ l 1015 cm, is of particular importance for the outburst ) R 3

0 0 12 light curves. We therefore focus our numerical modeling 1 ( on the early phases of runaway, unstable mass exchange y that expels this largest-scale CSM. 4 3.1. Hydrodynamic Models of Binary Coalescence 14 4 0 4 4 0 4 Our models are simulated within the Eulerian hydro- x (103R ) x (103R ) dynamic code Athena++ (Stone, J.M., in preparation), and are based on the methodology described in MacLeod Figure 1. Runaway overflow leading to binary coalescence for binary mass ratio q = 0.1. The upper panel et al.(2018a,b,c). We use a spherical polar coordinate shows binary separation as a function of time during our system surrounding the donor star in a binary pair. We simulation. The lower panel shows four snapshots (marked model the interaction of this donor star with a softened with vertical lines in the upper panel) of the gas density dis- point mass representing an unresolved companion ob- tribution in the orbital plane. We show these slices through ject. The domain extends over the full three-dimensional the orbital plane in a rotated x0 − y0 coordinate system so solid angle from 0.1 to 100 times the donor-star’s origi- that the companion star lies along the +x-axis. The dimen- nal radius. sionless simulations are rescaled to donor star mass of 30M and radius of 1000R for specificity in these images. Follow- The donor star is modeled by polytropic envelope, ing Roche lobe overflow, mass is pulled from the donor star with structural index Γ = 1.35. The donor has a core and expelled from the dragging the binary to mass of 25% its total mass. The gas in the simulation tighter separations. domain follows an ideal-gas equation of state, with in- dex γ = 1.35. These choices are intended to approxi- mately represent a convective, isentropic envelope of a The calculations themselves are carried out in dimen- massive star in which radiation pressure is important in sionless units where the donor’s mass, radius, and grav- the equation of state (in which case Γ = γ → 4/3, e.g. itational constant are all set to unity. They may, there- MacLeod et al. 2017a; Murguia-Berthier et al. 2017). fore, be rescaled to a physical binary of any mass or size. We initialize the calculation at a separation slightly Below, we report on two models which have secondary to smaller than the analytic Roche limit separation, where donor star mass ratios of q ≡ M2/M∗ = 0.1 and q = 0.3. the donor star overflows its Roche lobe (Eggleton 1983), As we will see in Section5, q ∼ 0.1 is a very common and halt the calculation when the companion star has mass ratio, while q = 0.3 is near the upper end of the plunged to 10% of the donor’s original radius – the inner range of events that result in mergers. boundary of our computational domain. The binary is initialized in a circular orbit and the donor star is ini- 3.2. Unstable Mass Transfer Leading to Binary Merger tially rotating as a solid body with rotational frequency Mass transfer is unstable in our model binary system matching the orbital frequency at the Roche limit sepa- in that it runs away to ever increasing rates and drives ration. the binary toward merger. This process begins with 5

Roche lobe overflow of the donor star into the vicinity near the L2 , near the lower-mass com- of its companion. In general, mass transfer proceeds panion object (Shu et al. 1979; Pejcha et al. 2016a,b; unstably when the loss of material from the donor causes Metzger & Pejcha 2017; MacLeod et al. 2018b). By the the donor star to increasingly overflow its Roche lobe – final panel, where the separation is 100R , the core of either because it grows in radius, or because its Roche the original donor and companion object are mutually lobe shrinks. In binary systems such conditions are often immersed in a significantly extended common envelope realized in binary pairs where a more massive donor star that originated from the donor star (Paczynski 1976). transfers mass onto a less massive accretor, causing the Once immersed, some binary systems deposit enough binary separation to shrink. energy into their environments to expel this envelope. In Figure1, we show the binary system separation as Others do not, and the companion object merges with a function of time in our model system, and snapshots the core – powering the sort of engine-driven explosions of the gas density distribution in the orbital plane. In discussed in Section2. these figures, we have rescaled our dimensionless simula- Because orbital tightening and coalescence of the bi- tions to a fiducial donor star mass of 30M and radius nary system is a direct result of angular momentum loss of 1000R . In the upper panel, time is zeroed at the to ejected material, the amount of ejecta relates directly time at which the companion plunges within the orig- to the binary properties. First using semi-analytic scal- inal donor-star’s radius, t1, where a(t1) = R∗. Over ings (MacLeod et al. 2017b), then hydrodynamic simula- the preceding 50 years, the binary separation contin- tion results (MacLeod et al. 2018b), we have found that uously shrinks, at first gradually, but with increasing the expelled mass at the onset of coalescence (defined as rapidity (MacLeod et al. 2018a). After the companion mass at r > R∗ at t = t1) is always on the order of 25% object plunges within the donor’s envelope it spirals to the mass of the merging companion object (Section 4.2 the inner boundary of our computational domain (at of MacLeod et al. 2018b). In the calculation shown in 10% the donor’s radius) within about 5 orbital cycles, Figure1, which has a mass ratio q = 0.1, at t1 the ejecta or two years, approximately the orbital period at the mass (measured as the mass at radius greater than the donor star’s surface. donor’s original radius) is 16% the companion’s mass, or Mass loss from the donor star at the expense of or- approximately 0.49M . At the termination of our cal- bital energy drives this rapid decrease in binary separa- culation, when the separation has decreased by a further tion and the pair’s coalescence. Turning our attention factor of ten, the ejecta mass has increased to roughly to the lower panels of Figure1, we note that as the 150% of the companion object’s mass, or 4.45M . By donor star overflows its Roche lobe, material is pulled, comparison, our calculation with q = 0.3 expels nearly primarily from the vicinity of the L1 Lagrange point, identical percentages of mass relative to the more mas- toward the companion object. As the orbital separa- sive black hole: 1.44M at a separation equal to the tion decreases, from 1500R to 1000R to 500R , the donor’s radius and 13.9M in our final snapshot (sep- breadth and intensity of this mass transfer stream in- aration 10% the donor’s radius). In the following, we crease dramatically. MacLeod et al.(2018a) studied the analyze the distribution of this material in the circum- dynamics of this runaway, unstable Roche lobe overflow stellar environment. in detail, and found that the mass loss rate from the donor increases by orders of over this period. 3.3. Resultant Circumstellar Distribution However, MacLeod et al.(2018a) also show that the an- Next, we analyze the three-dimensional distribution alytic model of Paczy´nski& Sienkiewicz(1972) coupled of debris expelled by the merger episode. To do so, we to a point-mass binary orbit evolution model captures analyze the final snapshot of our hydrodynamic simu- the key features of these stages once the specific angular lation, when the separation has tightened to one-tenth momentum of the ejecta has been measured (e.g. Huang the donor’s original radius, or 100R in our fiducial, 1963). 30M , 1000R model. Because the binary separation These high mass exchange rates quickly exceed the is tightening extremely rapidly at this phase, material Eddington limit mass accretion rate that material can ejected subsequently in the merger does not affect the accrete onto a compact object companion, and much largest-scale gas distribution prior to the compact ob- of the material pulled from the donor is lost to the cir- ject’s merger with the core. This, therefore, is the CSM cumbinary environment (as seen in the snapshots of Fig- that any explosive outburst will interact with as it ex- ure1, though in the case of the simulation this is because pands, particularly when we consider the crucial scales accretion onto the companion object is not modeled, see of interest of 1014 to 1015 cm that lie near the photo- MacLeod et al. 2018b). Much of this mass loss occurs sphere of the explosive transient. 6

40 10 40 10 )

11 e g

11 a r

20 20 e v a

l ] 12 a 3

12 h t u ) ) m c m

i R R g z [ 3 3 a

( 0 0 0 13 ) 0 13 ] 1 1 3 ( ( (

0 z y 1 m g c

o l 14 g [

14 ) ( 0

20 20 1

15 g o 15 l

16 40 16 40

40 20 0 20 40 0 20 40 x (103R ) R (103R )

Figure 2. Three dimensional distribution of ejecta near the time of merger. As in Figure1, we have scaled to a fiducial donor mass of 30M and radius of 1000R . When the compact object merges with the donor’s core, it is surrounded by an extensive, thick torus of debris expelled by the merger itself.

In Figure2, we show the large-scale density distribu- locities are similar to other sources of loss 4 tion out to 50 times the initial donor radius (5 × 10 R , like winds or non-terminal outbursts, in that they are or approximately 3.5 × 1015 cm). The panels show a similar to the ’s escape velocity and are much slice through the orbital, x − y plane, and an aziumthal less than the later explosion’s blast wave velocity. average, plotted in z − R, perpendicular to the orbital Though the axisymmetric torus structure discussed plane. Figure2 shows that a thick, extended circumbi- above is clearly structured in polar angle, for the sake nary torus of expelled material from the donor’s enve- of computational efficiency, we model the interaction lope has formed around the merging pair of stars. This of the explosive blast wave with a one-dimensional torus is roughly azimuthally symmetric, but has dis- (spherically-symmetric) density distribution derived tinct structure in polar angle, with relatively evacuated from these models. In future work, it may be extremely poles and dense equator, representative of the fact that interesting to relax this simplification. To derive one- material is flung away from the merging binary in the dimensional profiles, we spherically-average our model equatorial plane. Pejcha et al.(2016a,b, 2017) analyzed results about the donor star’s core. the thermodynamics of similar outflows and show that These 1D density profiles are shown in Figure3, in heating, arising continuously from internal shocks, and which we compare the unperturbed envelope profile to radiative diffusion and cooling very likely regulate the the cases disturbed by binaries of q = 0.1 and q = 0.3. torus scale height. Thus, the precise scale height ob- Where the hydrostatic profile has a distinct limb at the served in Figure2, modeled under the simplification of donor’s radius, the post-merger profiles show a roughly an ideal-gas equation of state, would likely be modified power-law slope in radius, with approximate scaling of by the inclusion of more detailed physics. r−3, as shown in the lower panel. Comparing the q = MacLeod et al.(2018b) analyzed the kinematics of this 0.1 and q = 0.3 results shows that in the higher mass- torus material and found that the radial velocities of the ratio coalescence, more of the envelope material has been most extended material are low relative to the escape ve- expelled beyond the donor’s original radius, yielding a locity of the original donor star (roughly 100 km s−1 for shallower density fall-off with radius and a profile with our fiducial model). Thus, the majority of these (ear- more mass at large radii. In both cases, we see that the lier) ejecta around bound to the merging binary. Some of distribution of ejecta extends to roughly 1015 cm, with the material at smaller radii (the later ejecta) is moving of order a solar mass (q = 0.3) or a tenth of a solar mass more rapidly, at velocities similar to the escape veloc- (q = 0.1) on these scales. ity. It therefore collides with the earlier, slow moving ejecta (MacLeod et al. 2018b). Qualitatively, these ve- 4. MERGER-DRIVEN LIGHT CURVES 7

6 10− ing transient light curves. In doing so, we summarize ρave, q = 0.1 8 and build on a considerable literature that describes 10− ρave, q = 0.3 ] 3 10 how CSM interaction can form a significant contribu- − HSE profile 10−

m tion or even dominate the radiative power of a transient c 10 12 g −

[ under certain conditions (e.g. Chugai & Danziger 1994;

ρ 14 10− Smith & McCray 2007; Chevalier & Irwin 2011; Cheva- 16 10− lier 2012; Chevalier & Irwin 2012; Ginzburg & Balberg 1013 1014 1015 2012; Moriya & Maeda 2012; Moriya et al. 2013a,b; Pan 1034 et al. 2013; Ofek et al. 2014; Ginzburg & Balberg 2014; ]

g Morozova et al. 2015b, 2017, 2018; Morozova & Stone [

3 33

r 10 2018; Kleiser et al. 2018; Chandra 2018). ρ π 4 32 10 4.1.1. Thin Shell of CSM

31 In the simplest version of a CSM interaction, an addi- 10 13 14 15 10 10 10 tional internal energy ∆E is added to ejecta by sweeping r [cm] up a thin shell of CSM, Figure 3. Spherically-averaged density distributions, com- paring the initial, hydrostatic polytrope (labeled HSE), with dMCSM merger-simulation snapshots for q = 0.1 and q = 0.3. As ∆E ≈ E, (4) Mtot in the previous Figures, we scale our models to a donor of 30M and 1000R . The existence of significant quantities where Mtot is the sum of the explosive ejecta and the of mass near the type II supernova radius of swept-up CSM mass internal to the shell, dM is the 1015 cm implies that interaction with this medium will play CSM CSM shell mass, and E is the kinetic energy of the explo- an important role in explosive transient light curves. sive ejecta. If this deposition of internal energy occurs in optically thin regions, all of this energy is radiated, In this section we use analytic and numerical mod- and ∆Erad ≈ ∆E. If the CSM shell lies interior to els to understand the properties of merger-driven light the photosphere radius, the heated ejecta must continue curves. We find that the CSM distribution plays a cru- to expand, with gas internal energy decaying along an cial role in shaping these light curves. In Section 4.1 we adiabat, before they are free to radiate. If we assume provide some analytic context for the potential role of that gas specific internal energy decays adiabatically as −1 CSM. In Section 4.2, we describe our numerical method r prior to reaching the photosphere (which is the case 4/3 for 1D radiative hydrodynamics calculations and light when radiation pressure dominates and P ∝ ρ along curve generation. Finally, in Sections 4.3 and 4.4 we de- an adiabat), then ∆Erad ≈ (r/Rph)∆E. scribe the key features and variations across parameter space of these numerical model light curves. 4.1.2. Continuous Distributions of CSM The differential radiated energy due to sweeping up 4.1. Analytic Context for Contribution to Radiative a CSM mass dM at radius r interior to the photo- Efficiency from CSM Interaction CSM sphere radius is In order to provide context for the interpretation of our numerical light curve models, in this section we dE E r rad ≈ , (5) analyze simplified analytic models of emission powered dMCSM Mtot Rph by CSM interaction. In general, CSM interaction en- hances the intrinsic luminosity of “cooling emission” where Erad is the contribution to the radiated energy from heated, expanding material like supernova ejecta. due to CSM interaction alone. As before, we are also As ejecta shock-heat by collisions with CSM, their ki- assuming that gas specific internal energy decays adi- netic energy is dissipated and converted into radiation. abatically as r−1. Given a continuous distribution of Depending on the location of the interaction (within or CSM material distributed between R∗ and Rph, we can outside the photosphere), this radiation may either es- integrate this expression over radius. In what follows, cape immediately or adiabatically decay with expansion we will assume that E is constant, which is justified only in the outflow prior to being free to stream out. if MCSM  Mtot and Erad  E. Otherwise, the losses In particular, we focus on different radial density pro- in kinetic energy to thermal energy or radiation must 2 files of CSM material and the role this plays in shap- be taken into account. We replace dMCSM = 4πr ρdr, 8 where ρ is the CSM density, to write where, in the last line, we have used MCSM from equa- tion (8). First, we emphasize that the ratio of R to E Z Rph r 4πr2ρ ph rad ≈ dr, R now affects the radiated luminosity arising from CSM E R M ∗ R∗ ph tot interaction, which is not the case for n = 2. Therefore, 4π Z Rph when R  R , the radiated energy from this CSM ≈ ρr3dr. (6) ph ∗ RphMtot R∗ profile is considerably less than that of the n = 2 pro- file. A final way to interpret this result is in terms of This integral shows that the dependence of ρ(r) will be the CSM mass at radii similar to the photosphere ra- critical in determining the CSM contribution to the ra- dius, r ∼ R . From equation (8), this is approximately diated luminosity. ph M / ln(R /R ). Equation (10) shows that this is Let us write a general, power law density form for the CSM ph ∗ the fraction of the CSM mass that contributes signifi- CSM that applies from the stellar radius, R , to the ∗ cantly to the radiated energy. eventual photosphere radius, R , ph In our q = 0.1 merger model, the scaling of the CSM is  r −n steeper, approximately ρ ∝ r−4. We reevaluate equation ρ(r) = ρph , (7) Rph (6) for n = 4 to find, where ρph is the density at the photosphere radius, and 3 Z Rph Erad 4πρphRph 1 we have chosen R as a characteristic radius to normal- ≈ dr, ph E M r ize the power law. We will further adopt the approxima- tot R∗ 3 4πρphR R  tion that Rph  R∗, under which the total CSM mass ≈ ph ln ph , can be written, Mtot R∗      R∗ Rph MCSM 1 n = 2, ≈ ln (for n = 4). (11)  Rph R∗ Mtot 3  MCSM ≈ 4πρphRph × ln(Rph/R∗) n = 3, (8)  Thus, for R  R , the CSM contribution to radiated  ph ∗ Rph/R∗ n = 4, luminosity is less for n = 4 than either n = 3 or n = 2 for several representative values of n. given a CSM mass. This result can be interperted in The most frequently considered form of ρ is that of light of equation8, which shows that the fraction of a steady, spherical wind, n = 2. Then, equation (6) CSM mass with r ∼ Rph is R∗/Rph for n = 4. becomes 4.1.3. Interpretation Z Rph Erad 4πρphRph ≈ rdr, In the preceding subsection, we have shown that for E M tot R∗ CSM density profiles that are sufficiently steep, n ≥ 3, 3 4πρphR ≈ ph , the CSM contribution to the radiated luminosity Erad/E 2Mtot depends on the ratio of the stellar radius over the pho- 1 MCSM tosphere radius – the radial extent of the CSM. This ≈ (for n = 2), (9) 2 Mtot important ratio varies in explosions with different size thus retrieving the often quoted result of the increase in stars of similar mass, or over the time evolution of a radiated energy scaling with the CSM mass as a fraction given transient as the photosphere radius increases. If of the total mass (Chevalier & Irwin 2011; Ginzburg & Erad/E becomes too small, then CSM interaction does Balberg 2012; Pan et al. 2013). not contribute significantly to the light curve of the tran- In our q = 0.3 model, the one-dimensional profile ap- sient at a given phase and the bulk of the radiated lumi- proximates n = 3, which yeilds constant mass per log- nosity comes instead from the adiabatically-expanding arithmic increase in radius. For n = 3, equation (6) blast wave. In this case, the transient assumes more evaluates to typical supernovae type IIP properties. 2 R These scalings indicate that we expect the CSM to be E 4πρphR Z ph rad ≈ ph dr, an important contribution to the q = 0.3 merger case E M tot R∗ light curve, because the mass in the CSM is a signifi- 3 4πρphR cant fraction of the total envelope mass, and with n = 3 ≈ ph , Mtot there is only logarithmic dependence on Rph/R∗, equa- 1 MCSM tion (10). In the q = 0.1 merger case, in which n = 4, ≈   (for n = 3), (10) we expect preferential contribution from CSM interac- ln Rph Mtot R∗ tion at early times in the transient light curves or for 9 particularly extended donors, because either situation radius, R0 < Rph, and the CSM contribution to the maximizes the ratio R∗/Rph, see equation (11). eventual radiated luminosity is reduced accordingly, see Finally, we have so far discussed the case in which the the discussion of Section 4.1.3. Nonetheless, the trun- CSM distribution extends out to, and perhaps beyond cated profiles do retain more than 95% of the CSM mass the photosphere radius. This is not necessarily realized. in all parameter variations. We assume a roughly solar In the case where the CSM terminates at a radius R0, isotopic composition that matches the hydrogen enve- for which R0 < Rph, the radiated luminosity from CSM lope of a presupernova stellar model of an initially 15M interaction is computed much as before, but only inte- star evolved with the MESA code that is included in the grating the mass distribution out to R0. This reduces SNEC distribution. the radiated luminosity due to CSM interaction by a fac- To drive the explosion of our models, we adopt a tor similar to R0/Rph < 1. In the sections that follow, thermal bomb at the inner edge of the envelope do- we use this framework to interpret 1D radiative transfer main. This broadly mimics the energy deposition of the models of explosions interacting with our model CSM quenched jet into the hydrogen envelope, as described distributions. in Section2. Based on Morozova et al.(2015b), we de- posit the energy over the innermost ∆Mbomb = 0.1M 4.2. 1D Radiation Hydrodynamic Models over a duration ∆tbomb = 0.1 s (Morozova et al. 2015b, has shown that the model light curves are not particu- While the analytic approach highlighted above is use- larly sensitive to these parameters, see their Figure 5). ful, it is necessarily simplified. To extend these calcu- The code also offers the option of adding Nickel to the lations of the CSM imprint on transient light curves composition. Though Nickel and lanthanide production to slightly more realistic scenarios, we need to per- is possible in merger-driven explosions (e.g. Siegel et al. form the associated integrations numerically. We uti- 2018; Grichener & Soker 2018), the quantities are uncer- lize the publicly available spherically symmetric (1D) tain. Because decay of radioactive 56Ni mainly powers Lagrangian hydrodynamics code SuperNova Explosion the late-time emission, here we choose to focus on Ni-free Code (SNEC) to calculate bolometric and filtered light models of the early light curve dominated by the CSM curves (Morozova et al. 2015a,b, 2017). The code uses and the hydrogen envelope (Morozova et al. 2015b). equilibrium-diffusion radiation transport to follow the Lacking radioactive material in the ejecta, our models time dependent radiation hydrodynamics of the expand- decline rapidly after the ejecta become fully transparent. ing blast wave. We choose to set the equation of state Emission from the photosphere in SNEC is assumed to using the built-in version of the Paczynski(1983) equa- follow a thermal blackbody, and thus neglects some line- tion of state, which includes contributions to the total blanketing effects that may be important for iron-rich pressure from radiation, ions and electrons based on the ejecta in the U and B bands. composition. Finally, in AppendixA, we test the sensitivity of our We map our one-dimensional, spherically-averaged model results to these choices by varying the inner mass profiles, shown in Figure3, into SNEC. Mass grid cells (or equivalently, radius) at which energy is deposited, as are customized by the user’s choice of binning in mass, well as the mass-resolution of the SNEC calculation. and we have found that the density profiles’ steep decline is best simulated with increasingly fine mass resolution at larger radii. This means that the shock break out is 4.3. Imprint of CSM not well resolved (Ensman & Burrows 1992; Morozova We begin to explore the imprint of the CSM mass et al. 2015a), but light curve calculations after the first distribution on the explosion light curve in Figure4, day are robust as shown by Morozova et al.(2015b). We in which we plot , along with photosphere run the simulations for this paper with 456 grid cells. radii, effective temperature, and gas bulk velocity at the Even so, the code does not function well for the lowest photosphere radius for our fiducial case of a 30M and densities. We therefore are required to restrict the CSM 1000R donor star in its initial, hydrostatic equilibrium −12 −3 density profile to ρ > 10 g cm (Morozova & Stone state (labeled HSE), and following merger with a 3M 14 2018). The outer radius is therefore 2.6 × 10 cm for (q = 0.1) or 9M (q = 0.3) black hole. In each case, the the fiducial simulation with q = 0.1 and 1.0 × 1015 cm explosion energy is taken to be 1051 erg, and is injected for the simulation with q = 0.3. We note that this at 0.1 times the radius of the original donor star, the in- restriction is not ideal because it limits the potential nermost radius resolved in our hydrodynamical models. interaction-driven luminosity of our models, see equa- As predicted by the analytic scalings in Section 4.1, tions (10) and (11). In practice, this implies that the the models with CSM are significantly more luminous at CSM maximum radius is often less than the photosphere peak (by a factor of roughly 100) than the hydrostatic 10

21 v

n 1.0

q=0.1 e

20 q=0.3 M

/ 0.8 ]

HSE ) e

15 m

r 10 c o [

c 0.6

19 h p M

0.4 R − q=0.1 h

18 p 0.2 q=0.3 M

( HSE 14 V 10 17 0.0 M 104 105

16 ]

s ] 3 /

K 10 m [ k

f [ f

15 e h p T 4 102 v 14 10

13 101 0 20 40 60 80 100 120 140 160 0 30 60 90 120 150 0 30 60 90 120 150 Time [days] Time [days] Time [days]

51 Figure 4. Fiducial, 30M , 1000R model star undergoing explosion of 10 erg in three scenarios: in its initial hydrostatic state (HSE), after merging with a 3M black hole (q = 0.1), and after merging with a 9M black hole (q = 0.3). The large panel shows bolometric luminosity, while the smaller panels track the photosphere’s properties – its mass element within the ejecta, radius, effective temperature, and bulk velocity at its location. CSM interaction brightens the merger models significantly as compared to the HSE case. Points mark the time of peak V-band brightness. model. The CSM interaction models also show delayed of days observed in some IIn supernovae with extended time of peak brightness, modified colors, and light curve CSM distributions like that observed for SN 1988Z and shapes, as we discuss in what follows. Morozova et al. SN 2005ip (Smith 2017). (2017) have recently discussed how rapid mass loss im- Comparing the two mergers, the q = 0.3 scenario mediately pre-supernova can transform light curves from with larger MCSM has a later and more luminous peak, a IIP shape (at low pre-supernova mass loss rates) to along with a higher luminosity during the plateau. Be- the more luminous type IIL (at higher mass loss rates). cause the CSM mass is related to the merger, we find This occurs when additional internal energy is added to MCSM/Mtot ∼ 1.5q, see Section3. The models of Figure the ejecta by shock-heating due to the CSM mass as it 4 for a hydrostatic explosion, q = 0.1, and q = 0.3, thus is swept up. Because the CSM lies outside the stellar provide a context for interpreting the apparent varia- radius, this new internal energy does not adiabatically tions in CSM contribution. While the CSM mass plays decay as much prior to being radiated from the tran- a primary role in determining the light curve brightness, sient’s photosphere. As a result, the radiative efficiency the distribution of CSM is crucial in shaping the light of the models, Erad/E, ranges from 1.4% for the HSE curves. In the case of the hydrostatic explosion (labeled model, to 8.5% for the q = 0.1 model, to 23% for the HSE in Figure4), there is no contribution from CSM q = 0.3 model. We note that these radiative efficiencies interaction. The q = 0.3 model shows a light curve are with a factor of two of those prediced by the scaling that is always elevated by approximately three magni- models of Section 4.1, for n = 4 and n = 3, respectively, tudes above the HSE model due to CSM interaction (at equations (10) and (11). t > 30 d). By contrast, the q = 0.1 model is signifi- Many features of our model light curves with merger- cantly elevated above the HSE model only earlier in the ejecta are similar to Morozova et al.(2017)’s model lightcurve, and converges to the HSE plateau luminosity suites including dense CSM distributions of varying around 100 d. The distinction between these cases lies mass and power-law slope. In particular, elevated early in the slope of the CSM density and in the outer CSM “plateau” luminosities that decay down to the unper- radius. turbed plateau are representative of significant CSM at For q = 0.3, the CSM has ρ ∝ r−3 (n = 3). The total radii less than the transient’s eventual maximum photo- radiative efficiency due to the CSM, Erad/E, equation sphere radius of approximately 1015 cm. Comparing to (10), only decreases with the logarithm of the ratio of Figure3, we note that large masses of relatively close-in the expanding ejetca photosphere radius, seen in the CSM are the distinguishing features of our models. The right hand panels of Figure4, to initial stellar radius resultant light curves, therefore, have typical duration (which is the base of the CSM distribution). Further, of hundreds of days like normal IIP, not the thousands the outermost radius of the CSM at the moment of en- 11 ergy injection in our numerical model is R = 1015 cm. 0 E = 3 1050erg This is larger than the ejecta photosphere radius early -20 E = 1 × 1051erg in the light curve, and similar to the ejecta photosphere E = 3 × 1051erg radius later in the light curve, implying that there is not -18 × 52

V E = 1 10 erg a significant adiabatic degradation of the CSM contribu- × M tion before light can escape from the expanding ejecta. -16 We can compare these trends to the q = 0.1 model, in which case the overall CSM mass is lower and ρ ∝ r−4 (n = 4). As the photosphere grows with time across the -14 transient duration, the contribution of CSM interaction Time [days] M = 15M -20 decreases approximately as R∗/Rph, see equation (11). M∗ = 30M ¯ Of similar importance, the outermost CSM radius at the M∗ = 60M ¯ 14 ∗ ¯ moment of energy deposition, R0 = 2.6 × 10 cm is a -18 factor of a few less than the ejecta photosphere radius V M late in the plateau phase. This effect also decreases the -16 contribution of the CSM to the late-stage light curve compared to an n = 4 CSM of infinite extension, e.g. equation (11). As a result, the light curve converges to -14 a similar magnitude as the plateau of the hydrostatic Time [days] R = 150R explosion late in the light curve. -20 R∗ = 500R ¯ In addition to brightening the explosion, CSM interac- R∗ = 1000R¯ tion modifies the object’s colors at timescales of days to -18 ∗ ¯ weeks on which transients are typically discovered. CSM V interaction yields bluer colors at time of peak (effective M -16 temperatures of several 104 K on timescales of tens of days). The photosphere cools to more typical IIP tem- peratures of thousands of Kelvin only after 50 to 100 -14 days (for the q = 0.1 and q = 0.3 models, respectively). 0 20 40 60 80 100 120 140 160 These higher temperatures are directly representative of Time [days] the extra internal energy injection due to shock heating of the ejecta by the CSM density distribution. Figure 5. Bolometric light curves for q = 0.1 transients with varying energy (top panel) mass (center panel) and ra- dius (lower panel). Unless specifically modified, we adopt our fiducial values of a 30M and 1000R donor star and 4.4. Implications of Varying Energetics and Donor 1051 erg explosion. The location of the V -band peak is Star Properties marked with a dot. Varying energetics and donor star prop- erties create light curves of different duration, peak bright- We expect mergers between compact objects and giant ness, and degree of CSM contribution. stars to occur at a wide range of donor star and com- pact object properties because the binaries from which they form have broad distributions of mass, semi-major Varying explosion energy with other properties kept axis, and mass ratio. Further, the energy of the central fixed yields the qualitatively expected variation in light engine is unknown, and may, in fact, vary from merger curve luminosity and duration – higher energy explo- to merger. Here we explore the implications of the pa- sions give rise to faster ejecta, with more luminous but rameter space of merger properties on the resultant light shorter duration transients. We note that the relative curves. contribution of the CSM interaction, shown by the early Figure5 shows V-band light curves for a range of mod- bump in the lightcurve, decreases in the more energetic els, all q = 0.1, in which we vary energy (top panel), supernovae. When the explosion energy changes, one mass (center panel) and radius (bottom panel) around consequence is that the photosphere radius during the 51 our fiducial, 30M , 1000R case with 10 erg explo- 5/12 plateau phase changes, roughly as Rph ∝ E (Popov sion energy. In all of these cases, because q = 0.1, the 1993; Kasen & Woosley 2009). For higher energies, −4 CSM density profile is roughly ρ ∝ r , and equation the larger photosphere radii imply smaller contributions (11) predicts the approximate contribution of CSM in- from CSM interaction, because the photosphere is fur- teraction to the radiated energy. 12 ther outside the outermost CSM radius, R0. In Figure 21 q = 0. 3 4.95 5, we observe that the light curve shape transforms as E = 1 1052erg the energy increases. This is reflective of the decreasing × contribution of CSM interaction to the light curve as 20 4.80 E increases. Consequently, the radiative efficiency de-

50 ] creases from 10% for the 3 × 10 erg explosion to 4.7% M = 15M 4.65 K [ ∗ ¯ M = 60M

52 ) ∗ ¯ for the the 10 erg explosion. With this smaller CSM 19 k k a a e 52 e p p

contribution, the 10 erg explosion light curve shows a , , 4.50 f f V relatively typical IIP shape, with a small, early bump e T M (

18 0 due to the CSM. 50 1

E = 3 10 erg 4.35 g

× o

Varying donor star properties, in the form of mass l and radius, similarly changes light curve duration, peak R = 150R brightness, and shape. More massive donor stars yield 17 ∗ ¯ 4.20 Erad/E = 0. 2 HSE higher ejecta masses, but constant MCSM/Mtot ∝ q. At fixed energy, the ejecta velocities are lower and light Erad/E = 0. 01 4.05 16 curve durations are correspondingly longer. As mass 10 18 30 50 varies in Figure5, the plateau photosphere radius varies tpeak [days] only mildly because the higher ejecta masses are bal- anced by lower ejecta velocities. Therefore, though these Figure 6. Luminosity and timescale of the optical light models show different characteristic timescales, they all curves of merger-driven explosions. Here we additionally have very similar peak magnitudes and degrees of CSM summarize radiative efficiency with size of the marker and effective temperature at peak with color. The central point contribution to their overall radiated luminosity (total is our fiducial model; lines connect the isolated variations of radiative efficiencies range from 11.7% to 9.0%). energy (dashed), mass (dotted), radius (dot-dash), and mass Varying donor star radius changes not only the extent ratio (solid). of the donor itself but the extent of the CSM, which extends to tens of stellar radii. This, in turn, varies the crucial ratio of maximum CSM radius to transient Together, Figure6 shows that merger-driven explo- photosphere radius (because varying donor radius has sions occupy a somewhat restricted phase space of lumi- little effect on Rph). When the donor is more compact, nosity and timescale. Typical models are more luminous for example 150R , the CSM extends to approximately than standard type IIP, but less luminous than super- 1014 cm, and largely affects only the early lightcurve. luminous supernovae. In all of the models bearing sig- Progressively larger donors of 500R and 1000R scale nificant CSM-interaction features, effective temperature the radial size of the CSM distribution. This scaling varies systematically with time of peak brightness with yields more CSM material at radii closer to the tran- longer-duration transients appearing redder and shorter- sient’s photosphere radius at later times (for example duration transients appearing bluer. near peak), in turn implying higher radiative efficien- cies, and brighter transients. For example, the 150R 5. POPULATION SYNTHESIS OF model has a radiative efficiency of only 1.9%, while the MERGER-DRIVEN EXPLOSIONS 500R model radiates 5.3% of the explosion energy and We use population synthesis models of stellar binary the 1000R model radiates 8.5% of the explosion energy. evolution to explore the statistical properties of binary Figure6 summarizes the parameter space of merger- systems at the time of a common envelope phase lead- driven explosions in luminosity, timescale, effective tem- ing to merger between a compact object and a giant perature, and radiative efficiency. The majority of star’s core. We then use these models to estimate the merger-driven explosions have radiative efficiency on the population of observable merger-driven explosions. order of 10%, much higher than the hydrostatic model with no CSM. The q = 0.3 model has even higher ra- 5.1. Population Model diative efficiency of 25%. However, the more compact We analyze rapid population synthesis models from 150R donor model and the highest explosion energy the Compact Object Mergers: Population, Astrophysics model, 1052 erg, both show relatively minimal CSM- and Statistics (COMPAS) suite (Stevenson et al. 2017; interaction features in Figure5 and have somewhat Barrett et al. 2018; Vigna-G´omezet al. 2018). These lower radiative efficiency (because R /R and R /R ∗ ph 0 ph models employ approximate tracks and are reduced, see section 4.1). parameterized physics in order to facilitate exploration 13 of the statistical properties of binary stellar evolution star or black hole remnant. Because neutron star kicks – including rare outcomes like the formation of double tend to be large in magnitude, a relatively small frac- compact object binaries (for a full description of the ap- tion of systems containing newly-formed neutron stars – proach, see Stevenson et al. 2017; Barrett et al. 2018; less than 4% (Vigna-G´omez et al. 2018) – remain bound Vigna-G´omezet al. 2018). In particular, we adopt the following the supernova. Of those that remain binaries, model parameters of Vigna-G´omezet al.(2018)’s “Fidu- a large fraction will undergo a common envelope phase cial” case and we capitalize on a recent development by during a reverse episode of mass transfer onto the com- Vigna-Gomez et. al. (in preparation) to record the char- pact object, initiated by the expansion of the intially less acteristics and outcomes of all common envelope phases massive companion after it completes core hydrogen fu- experienced by modeled binaries. sion. This may result in either a merger or a common Initial distributions of binary properties are sampled envelope ejection. Those that eject their envelopes may at the zero-age (ZAMS) in COMPAS. go on to form a double compact object binary, as dis- In the models we study, the mass of the primary star cussed by Vigna-G´omezet al.(2018). is drawn from an initial mass function in the form The population of common envelope phases involving dN/dm ∝ m−2.3 (Salpeter 1955) with masses between compact objects in these models is depicted in Figure7. 5 ≤ m/M ≤ 100. The mass of the secondary star In what follows, we report on and show only events that is then chosen from a flat distribution in mass ratio involve post-main-sequence donor stars more massive with 0.1 < qZAMS ≤ 1 (Sana et al. 2012). The ini- than 10M . We show the distribution of these sources tial separation is drawn from a flat-in-the-log distribu- in the Hertzsprung-Russell Diagram (HRD) as well as tion, dN/da ∝ a−1, with separations between 0.01 < in mass, radius, and mass ratio. We highlight the dis- aZAMS/AU < 1000 (Opik¨ 1924; Sana et al. 2012). All tinction between all common envelope phases involving stars in our model population adopt solar metalicity compact objects (labeled CO CE), and events resulting (Z = 0.0142). A total of 106 binary systems are simu- in mergers between a neutron star and the donor core lated. (labeled NS-Core) and a black hole and the donor core Common envelope phases are identified by conditions (labeled BH-Core). for dynamically unstable mass transfer in COMPAS. A number of interesting trends emerge from these When a common envelope episode occurs, an energy distributions. While common envelope phases occur criterion is used to evaluate the outcome. In particu- throughout the donor star’s post-main sequence evolu- lar, the final change in orbital energy is related to the tion, and therefore also the HRD, particular criteria are energy needed to unbind the giant star’s hydrogen en- most likely to result in a merger. Merging sources tend velope from its core, ∆Eorb = −αEbind, where α = 1 is to have the more compact radii compared to the overall 4 an efficiency parameter (Webbink 1984). If the maximal distribution of common envelope phases (Teff & 10 K). change in orbital energy (defined on the basis of the min- Neutron stars interact with a broad range of stellar imal separation at which the core fills its Roche lobe) is companion masses, while black holes common envelope insufficient to unbind the envelope, |∆Eorb| < α|Ebind|, phases tend to involve massive M & 30M and thus lu- then a merger between the companion and the core is as- minous donors. Typical mass ratios of compact-object sumed to result. This scaling implies that more compact common envelope phases range from 0.02 . q . 0.6; stars have higher binding energies and, for a given com- those resulting in mergers tend to have q . 0.2. The panion mass, are more likely to result in merger. More upper limits of these ranges reflect the conditions of dy- extend stars (nearer to the tip of their giant-branch evo- namical mass transfer stability and envelope ejection, lution) have lower binding energies and their common respectively. Of the mergers, the black holes form the envelope phases are more likely to result in envelope higher mass-ratio population, 0.1 . q . 0.2, while neu- ejection (de Kool 1990; Kruckow et al. 2016). tron stars typically have q . 0.1.

5.2. Compact Object-Core Mergers 5.3. Event Rate The most common evolutionary channel leading to a We can estimate the event rate of compact object - compact - giant star merger and a merger-driven explo- giant star mergers using the results of these popula- sion is as follows. A binary pair in an initially relatively tion synthesis models. We simulate 106 binary systems, 7 wide orbit evolves, likely going through a dynamically or approximately 1.96 × 10 M of binary mass. Each stable mass transfer from the initially more-massive star solar mass of modeled stars represents 3.8M of stars onto its companion. That initially more massive star formed (Vigna-G´omezet al. 2018). From our models, undergoes core collapse, leaving behind either a neutron common envelope phases involving compact objects and 14

106

103

5

10 ] ] R [

L s [ u i L

d 2

a 10 r 104

CO CE CO CE NS-Core NS-Core BH-Core 101 BH-Core 103 3 × 104 2 × 104 104 6 × 103 4 × 103 101 102 Teff [K] mass [M ] 10 3 ]

1 CO CE

x NS-Core e

d 4 BH-Core 1 10 M [

X

0 5 1 10 g N o l d d M d 10 6 101 102 101 102 103 10 1 100 mass [M ] radius [R ] q

Figure 7. Distributions of binary properties at the onset of common envelope phases involving black holes or neutron stars interacting with evolved, massive star donors with mass greater than 10M . Here we distinguish between all common envelope phases involving compact objects (labeled CO CE), and cases in which a neutron star merges with the donor’s helium core (NS-Core) or a black hole merges with the donor’s helium core (BH-Core). Mergers occur in roughly 22% of the compact object common envelope phases, and are split relatively equally between black hole and neutron star events. The companion mass distribution, especially for black hole mergers, favors massive companions. Histograms are plotted in units of events per solar mass of stars formed per logarithmic bin in x-value (mass, radius, or mass ratio).

donors more massive than 10M occur with a frequency 5.4. Outburst Population −4 −1 of 1.5 × 10 M , where the unit denotes mergers per Having assessed the population of donor stars and solar mass of stars formed. Of these, approximately compact object companions that undergo mergers, we 22% result in mergers. Among the mergers, 49% in- now extend the results of our light curve models to esti- volve black holes, while 51% involve neutron stars. The mate the properties of the population of observable tran- rate of black hole-core mergers and neutron-star merg- sients. Guided by the results of Sections 4.4 and 4.1, we −5 −1 ers are thus each approximately 1.6 × 10 M . These note that CSM interaction is most important when the events occur in a spread of ages between 3 and 40 Myr, binary mass ratio is larger (yielding more merger ejecta and are thus strongly correlated with recent star for- and higher CSM mass) and when the radius is extended mation. By comparison, core collapse supernovae occur (yielding less adiabatic degradation of CSM-interaction with a frequency of 5.8 × 10−3M −1 in the model sys- energy, proportional to R∗/Rph). Comparison to the tems. Merger-driven explosions therefore represent on population properties in Figure7 shows that the sys- the order of 0.6% of all core collapse events. tems that tend to have high mass ratios and large radii are predominantly the BH-Core merger group, in which 15 a black hole merges with its giant star companion. By NS-Core (Popov) contrast, the typical radii, R∗ ∼ 100R , and mass ra- ] 2.0 BH-Core (Popov) tios, q < 0.1, for the neutron star-core mergers are such 1 NS-Core (CSM) g that we expect less dramatic signatures of CSM interac- a BH-Core (CSM) m tion (see Figure5). 1 1.5 To map our parameter variations onto the modeled ) population, we estimate the following scalings of MV M 5 0

with changing model parameters from the results of Fig- 1 1.0 ( [ ure6, k a e p , V N M MV,peak ≈ −18.9 − 2.42 log (R∗/1000R ) d 10 d 0.5 M − 0.229 log10(M∗/30M ) d − 1.41 log (E/1051 erg) 10 0.0 14 15 16 17 18 19 − 2.20 log10(q/0.1) (12) MV, peak An important caveat is that, given our limited model parameter coverage, these numerical scalings represent Figure 8. Transformation of merger-driven explosions by CSM interaction. We show peak V-band magnitudes of the individual dependencies on binary properties about the population of merger-driven explosions using the Popov our fiducial model rather than the full parameter co- (1993) model (equation (13); dashed lines), then apply our variance. We will compare these luminosities to those results (equation (12); filled histograms) to derive the peak of standard type IIP supernovae (Popov 1993), magnitudes including CSM interaction with merger-expelled ejecta. Black hole merger-driven explosions, in particular, MV ≈ −11.42 − 1.67 log10(R∗/R ) form a distinct and luminous group that comprises 49% of the merger-driven explosion transients. The y-axis is shown + 1.25 log (M∗/M ) 10 in units of events per magnitude per 105M of stars formed. 50 − 2.08 log10(E/10 erg). (13)

We note that the Popov(1993) model accurately pre- stellar-wind mass loss. Type IIn supernovae have long dicts the peak V-band luminosity of our hydrostatic been acknowledged to have CSM due to the persistent model (Figures4 and6). narrow lines in their spectra. This otherwise diverse In Figure8, we apply these scalings to the popula- class of supernovae occupies approximately 10% of the tion of compact-object core mergers. Again we divide overall core collapse rate (e.g. Kiewe et al. 2012). More the population on the basis of whether a neutron star recently, evidence has been emerging that a majority (up or a black hole is merging with the core. We assume to 70%) of type II supernovae show evidence of having that all events have 1051 erg explosion energy for the at least 0.1M of CSM imprinted on their light curves sake of this illustration. We find that CSM interaction (Morozova et al. 2018; Morozova & Stone 2018). For (as predicted by equation (12)) brightens all merger- example, Forster et al.(2018) has argued for systematic driven explosions relative to their hydrostatic equiva- evidence that most type II shock breakouts are delayed lents (as estimated from equation (13)). The neutron by interaction with dense CSM. A shared feature of the star-core mergers are brighter by approximately 1 mag- CSM in many type IIP and IIL supernovae is that it nitude than their Popov-model equivalents. However, is very close to the donor star, indicating its loss in the the black hole mergers are brightened significantly more, years immediately prior to the explosion (e.g. Ofek et al. by approximately 3 magnitudes. In this diagram, we ob- 2013c; Smith & Arnett 2014). serve that the black hole merger-driven explosions form One proposed explanation for the presence of pre- a distinct population more luminous than the non-CSM- supernova CSM ejection lies in the phenomenological interacting IIP population. comparison to (LBV) outbursts, 6. DISCUSSION which are non-terminal outbursts of massive O-type stars. Though the precise cause of these outbursts re- 6.1. Production of Supernovae-like Transients With mains uncertain (e.g. Justham et al. 2014), as does their Massive, Close CSM potential correlation with the evolutionary trend of the It has recently become apparent that a large fraction core toward collapse, in at least one dramatic example, of type II supernovae show signs of interaction with CSM SN2009ip, both LBV outbursts and a terminal super- of densities much larger than that implied by nominal 16 were observed in the same object over the course 20 B of a decade (Smith et al. 2010; Prieto et al. 2013; Ofek V R et al. 2013a; Smith et al. 2014; Margutti et al. 2014; I Mauerhan et al. 2014). 19 Another possible explanation links the CSM to the vigorous convection due to accelerating nuclear burn- 18 ing in the pre-supernova core. In this case, convection launches waves at the interface between the con- 17

vective core and an overlying radiative layer. These waves propagate through the radiative zone and dis- 16 sipate near the base of the convective hydrogen enve- lope (Quataert & Shiode 2012; Shiode & Quataert 2014; SN1979c Fuller 2017). The luminosity of these dissipating waves 15 20 0 20 40 60 80 100 can be highly super Eddington in the year prior to core- Time [days] collapse, driving extensive mass loss (Quataert et al. 2016) or outbursts (Fuller 2017). V 19 R An explosion driven by the merger itself also naturally I links merger ejecta and CSM with the explosive fate of 18 the star, as we have described in the preceding sections. However, a merger-driven model cannot explain the full 17 diversity of type II supernovae or their CSM properties. In practice, some combination of these processes must be 16 at play in order to explain the abundance and diversity of CSM observed in type II supernovae. Absolute Magnitude 15 6.2. Comparison to Observed Supernovae 14 SN1998s We compare our model light curves to two represen- tative, well-studied supernovae. Photometric similarity 20 0 20 40 60 80 100 120 140 Time [days] is insufficient to demonstrate the origin of a given tran- sient, as we discuss further in Section 6.3. In this section, Figure 9. Upper panel: SN 1979c plotted on top of absolute we contextualize our model merger-driven explosions by magnitude from simulations with M∗ = 30M and R∗ = 51 showing that they bear similarities to transients already 1000R and ESN = 3 × 10 erg. Lower panel: SN 1998s plotted on top of absolute magnitude from simulations with in the supernovae archives. M∗ = 30M and radius scaled to R∗ = 1000R (solid lines) and R∗ = 500R (dashed lines). 6.2.1. 1979c SN1979c, classified as a type IIL, was discovered in April 1979, several weeks after explosion (Mattei et al. terial can also explain X-ray data seen from 1995 and 1979). It has been observed extensively at radio wave- 2007. lengths (Weiler et al. 1986; Montes et al. 2000; Bartel In the top panel of figure9 we have plotted optical & Bietenholz 2008; Marcaide et al. 2009), and early data from the first 100 days of observation (from the modeling suggested a very large progenitor radius of open supernova catalog; Guillochon et al. 2017). On top 5 R ∼ 6000R and CSM extending out to R ∼ 10 R we plot absolute magnitudes from our simulations with 51 (Bartunov & Blinnikov 1992). M∗ = 30M and R∗ = 1000R and E = 3 × 10 erg. Observations from the following 20 years gave rise to The plot is not a fit, but shows that the outcome of more theoretical discussion. Bartel & Bietenholz(2003) our simulations can closely replicate observed transients. suggested that the remnant is expanding into low den- This, in addition to the potential for a remnant black −n +0.10 sity CSM with ρ ∼ r , with n = 1.94−0.05 decreas- hole (Patnaude et al. 2011) make SN1979c an interest- ing to n < 1.5 at larger radii. Later Kasen & Bildsten ing candidate for further investigation under the merger- (2010) suggested that the light curve was brightened by driven hypothesis. the spindown of a at the center of the SN rem- nant. Patnaude et al.(2011) note that, rather than a 6.2.2. 1998s magnetar, a 5 − 10M BH accreting from fallback ma- 17

SN1998s is one of the most studied type IIn super- the blast wave. The Large Synoptic Survey Telescope novae (Shivvers et al. 2015). From spectral lines, two (LSST) will discover on the order of 105 core collapse shells of CSM were identified. Fassia et al.(2001) found events per year (LSST Science Collaboration et al. 2009, that the inner CSM was within 90AU from the center chapter 8), implying hundreds to thousands of merger- and the outer CSM extended from 185 AU to over 1800 driven explosions detected in a given observing year. AU. The classification of SN1998s as a IIn is a direct re- Among this flood of optical transients, the challenge sult of the very early spectral observations; the narrow will be unambiguous identification of merger-driven ex- line features disappeared and morphed into the broad plosions rather than detection. A full consideration is lines of a type IIL or IIb within weeks Smith(2017). beyond the scope of our initial study, but we speculate SN1998S has later been interpreted as having a red on several potential signatures here. As discussed in Sec- supergiant progenitor with possibly asymmetric CSM tion3, the ejecta from the pre-merger common envelope consisting of the two separate shells, caused by separate phase are densest in the equatorial plane of the binary. mass-loss events. (Kangas et al. 2016) claims that this When the supernova explodes into this aspherical den- type of SN is very common and that many IIL and IIP sity distribution, the blast wave will be shaped by these share spectral features with IIn in early spectra. asymmetrical surroundings (Blondin et al. 1996). Emis- In the bottom panel of figure9 we plot the V,R and sion from the photosphere will, as a result, be polarized I band data from SN1998s (data from the open super- by one to several percent, as has been described in the nova catalog; Guillochon et al. 2017) plotted on top of case of SN2009ip (Mauerhan et al. 2014). absolute magnitude from simulations with M∗ = 30M The interacting binary progenitor of the explosion and E = 1 × 1051 erg, solid lines for radius scaled may also offer clues in the identification of merger- to R∗ = 1000R , and dashed lines for R∗ = 500R . driven supernovae, as in ongoing progenitor-monitoring The light curve we see from from our simulations with efforts described by Kochanek et al.(2008); Adams et al. R∗ = 1000R is similar to SN1998s, though the rate of (2017). Drawing parallels to low-mass, Galactic stel- decline perhaps fits better with our R∗ = 500R sim- lar merger events like V1309 Sco (Mason et al. 2010; ulation. Just as with SN1979c, the overall light curve Tylenda et al. 2011), increasing rates of non-conservative shape, duration, and brightness are well-approximated mass transfer (seen in the panels of Figure1) may en- by our models. The presence of nearby CSM is also con- shroud the merging binary in dust and cause an optical sistent with a merger-driven explosion. However, the fading of the progenitor star prior to merger. In V1309 explanation for two distinct shells of CSM is not imme- Sco, such a phase of optical dimming was observed in diately apparent given our model predictions, and may the phase of 100 to 1000 orbital periods prior to coa- be in tension with the merger-driven hypothesis for this lescence. In the last orbits leading into the merger (the transient (though see the discussion of Clayton et al. portion captured by Figure1), V1309 Sco brightened 2017). in optical bands as more-and-more emission arose from the outflow from the binary (Pejcha 2014; Pejcha et al. 6.3. Identification in Optical Surveys 2016a,b, 2017). Future work is needed to extend these scenarios to detailed predictions for pre-explosive behav- Having shown that merger-driven explosion models ior in massive star coalescence. can reproduce the basic, photometric properties of sev- Multiwavelength, particularly X-ray, signatures, while eral observed supernovae, we now focus on the prospects less frequently available than optical photometry, pro- for their more secure identification. vide a powerful tool for probing early CSM interaction The prevalence of merger-driven explosions (of order (e.g. Chevalier & Irwin 2012; Margutti et al. 2017; Mo- 0.5% of the core-collapse rate) begs questions about rozova & Stone 2018). These data can probe the CSM their prior detection in existing datasets and their im- distribuiton in great detail, including the density distri- prints on future surveys. Current surveys, such as the bution through the hard to soft emission ratio (Moro- Zwicky Transient Factory (Bellm & Kulkarni 2017) and zova & Stone 2018). If the CSM is as steep as predicted All-Sky Automated Survey for Supernovae (e.g. Holoien in the merger-driven models (steeper than ρ ∝ r−3), it et al. 2019) are presently discovering hundreds of new will accelerate the leading edge of the ejecta to high ve- core-collapse supernovae per year. This discovery rate locities and produce hard x-ray emission (Morozova & suggests that one or more merger-driven explosions is Stone 2018). currently being discovered per year. Efforts at early Finally, merger-driven explosions will leave a black discovery and spectroscopy of these transients aim to hole as the remnant of the rapid accretion phase fol- reveal CSM properties through “flash spectroscopy” in lowing merger of the compact object with the stellar which the CSM is ionized prior to being swept up by 18 core. Though black hole formation is common in core- brightens neutron star mergers by approximately collapse events, it is typically believed to accompany im- one magnitude, but brightens the population of plosion rather than explosions and luminous supernovae black hole mergers by approximately three magni- (e.g. Sukhbold et al. 2018). If detected, the coexistence tudes relative to type IIP models with the same en- of a supernova-like transient and a remnant black hole ergy injection and pre-supernova stellar mass and would thus be consistent with the merger-driven explo- radius (Figure8). sion scenario. In theory we might distinguish neutron star and black hole central x-ray sources on the basis 4. The most luminous transients, those involving of their x-ray specta. In practice, this identification can black hole mergers, are at least as common as their be ambiguous when the surrounding, absorbing medium less luminous neutron star counterparts. Black is substantial. One such example of a transient harbor- hole mergers have MV,peak ∼ −18 to −19 with ing an embedded x-ray source is AT 2018cow (Margutti tpeak ∼ 20 to 30 d. The implication for optical sur- et al. 2019). veys is that the brightest, easiest-to-detect events comprise a significant fraction of the entire popu- 7. SUMMARY AND CONCLUSION lation. In this paper, we have presented models for merger- The calculations presented in this paper have demon- driven explosions that arise from the plunge of a com- strated that merger-driven explosions provide a nat- pact object within the helium core of its giant star com- ural mechanism for the production of supernovae-like panion following a common envelope phase (Chevalier transients with close-in, slow-moving CSM. Future work 2012). When a compact object merges with the helium could improve on the treatment of the stellar model (a core of a massive, post main-sequence star, the condi- polytropic envelope in our approximation) and the de- tions for rapid, neutrino-cooled accretion are met (e.g. tails of energy injection into this envelope. At present, Zhang & Fryer 2001). The accompanying release of en- we inject energy spherically into the envelope at one ergy may deposit approximately 1051 erg into the sur- tenth the star’s overall radius. In practice, the unknown rounding hydrogen envelope, leading to a merger-driven location and asymmetry of energy injection might play explosion (Chevalier 2012). Some key findings of our a key role in shaping transient light curves, colors, and investigation are: peak luminosities with respect to the estimates of our current models. 1. The binary coalescence leading to the merger of We compare our models to two representative super- the compact object with the core expels slow- novae, SN1979c and SN1998s in Figure9. However, we moving material into the surrounding environ- note that more work is needed to provide unambiguous ment, forming a dense, toroidal CSM (Figures1 confirmations of merger-driven explosions. In Section and2). The spherically-averaged density profile 6.3, we discuss additional strategies for the identification −3 −4 has a steep radial slope of ρ ∝ r or ρ ∝ r of merger-driven explosions including their asymmetry (Figure3). and polarization due to the toroidal CSM, the proper- ties of their progenitor binaries, and their early spectra 2. Using 1D radiation hydrodynamic models of the and X-ray emission. In future work, these signatures can explosions, we find that the CSM distribution be investigated through multi-dimensional calculations is crucial in shaping the transient light curves. of the explosive evolution and emergent light curve, as Merger-driven explosions are brightened by up to well as more detailed modeling of the progenitor sys- three magnitudes relative to their counterparts in tem’s plunge toward merger. hydrostatic stars (Figure4), with timescale and light curve shape that vary with donor-star mass and radius and explosion energy (Figures5 and6). We thank R. Margutti, E. Ramirez-Ruiz, and M. Rees for advice and helpful discussions in the development of 3. From population models, we find that black hole this work. We gratefully acknowledge the support of E. and neutron star mergers with giant star compan- Ostriker and J. Stone in the development and analysis of ions occur with similar frequency, each with a rate the hydrodynamic models, and V. Morozova for support −5 −1 per mass of stars formed of 1.6 × 10 M . The with SNEC. S. S. acknowledges support by the Danish combined rate is 0.6% of the core-collapse rate in National Research Foundation (DNRF132). Part of the our models. Merger-driven explosions occur across simulations used in this paper were performed on the a roughly flat distribution of donor-star masses University of Copenhagen high-performance computing from 10M to 100M (Figure7). CSM interaction cluster funded by a grant from VILLUM FONDEN 19

(project number 16599). M.M. is grateful for support this work were provided by the NASA High-End Com- for this work provided by NASA through Einstein Post- puting (HEC) Program through the NASA Advanced doctoral Fellowship grant number PF6-170169 awarded Supercomputing (NAS) Division at Ames Research Cen- by the Chandra X-ray Center, which is operated by the ter. Smithsonian Astrophysical Observatory for NASA un- Software: Athena++, Stone et al. (in preparation) der contract NAS8-03060. This work was supported in http://princetonuniversity.github.io/athena, Astropy part by the Black Hole Initiative at Harvard University, (Astropy Collaboration et al. 2013), SNEC (Morozova which is funded by a JTF grant. Resources supporting et al. 2015a), COMPAS (Stevenson et al. 2017; Barrett et al. 2018; Vigna-G´omezet al. 2018)

APPENDIX

A. VALIDATION OF LIGHT CURVE CALCULATIONS In this appendix, we discuss the validation of several numerical choices in the 1D radiation hydrodynamics calcula- tions with SNEC that we use to produce model light curves. In mapping the 3D hydrodynamics calculation of the merger (Section3) to the 1D explosive calculation, we need to make an assumption about the location (described by radius or mass coordinate) where the explosion energy is deposited. We have argued in Section2 that this deposition location is somewhere within the hydrogen envelope. Here we explore the sensitivity to that choice as follows. Beginning with our fiducial case of a 30M donor star and 3M 51 black hole in a q = 0.1 merger, we deposit 10 erg of thermal energy spread over 0.1M at different mass coordinate locations. Our default assumption is Min = 10.75M , which corresponds to the enclosed mass of at 0.1R∗ of 7.75M plus a 3M black hole. This model in Figure 10 corresponds to the fiducial simulation presented in Figure4, which is labeled q = 0.1. We then vary the mass coordinate at which thermal energy is deposited, moving outward in the star’s Hydrogen envelope. Material inside Min acts as a gravitational point mass for the remainder of the calculation. We find that for Min = 10.75M , Min = 14M , and Min = 18M (corresponding to radius coordinates of 0.1, 0.16, and 0.24 times the donor star’s original radius) the model light curves are very similar indicating only weak dependence on how the energy is spatially deposited within the hydrogen envelope. All of these radii are significantly outside the star’s more compact Helium core. We note that for Min > 20M , the case in which > 2/3 of the donor star forms a black hole while only < 1/3 is expelled, we do observe departures in the model light curves, with the bulk of the thermal energy radiated early, and not coupling efficiently to driving envelope expansion. We also test the dependence of our model results on spatial resolution within the 1D SNEC calculations. Our fiducial case divides the mass into 456 elements. Figure 11 compares this case to models with twice and four times as many zones (912, 1824, respectively). These tests confirm that our results are converged to within 1% across the light curve duration with any of these resolution choices.

REFERENCES

Abbott, B. P., Abbott, R., Abbott, T. D., et al. 2016a, Barkov, M. V., & Komissarov, S. S. 2008, MNRAS, 385, PhRvL, 116, 061102 L28 —. 2016b, PhRvL, 116, 241103 —. 2011, MNRAS, 415, 944 Abbott, B. P. e. a. 2017a, PhRvL, 118, 221101 Barrett, J. W., Gaebel, S. M., Neijssel, C. J., et al. 2018, —. 2017b, PhRvL, 119, 161101 MNRAS, 477, 4685 —. 2017c, ApJ, 851, L35 Bartel, N., & Bietenholz, M. F. 2003, ApJ, 591, 301 Adams, S. M., Kochanek, C. S., Gerke, J. R., & Stanek, —. 2008, ApJ, 682, 1065 K. Z. 2017, MNRAS, 469, 1445 Bartunov, O. S., & Blinnikov, S. I. 1992, Soviet Astronomy Andrews, J. E., & Smith, N. 2018, MNRAS, 477, 74 Letters, 18, 43 Andrews, J. J., Farr, W. M., Kalogera, V., & Willems, B. Belczynski, K., Kalogera, V., & Bulik, T. 2002, ApJ, 572, 2015, ApJ, 801, 32 407 Armitage, P. J., & Livio, M. 2000, ApJ, 532, 540 Belczynski, K., Kalogera, V., Rasio, F. A., et al. 2008, The Astropy Collaboration, Robitaille, T. P., Tollerud, E. J., Astrophysical Journal Supplement Series, 174, 223 et al. 2013, A&A, 558, A33 Bellm, E., & Kulkarni, S. 2017, Nature Astronomy, 1, 0071 20

v

Min = 10.75M n 1.0 ¯ e Min = 14.00M -20 ¯ M / ]

Min = 18.00M )

e 15 m

¯ r

10 c o [ c

h p M 0.5 R − h

-18 p M

( 14 V 10 0.0 M 4 Min = 10.75M 10 5 ¯ 10 Min = 14.00M

-16 ¯ ]

M = 18.00M s ] in / ¯ 3 K 10 m [ k

f [ f

e h p T 4 102 v -14 10

101 0 20 40 60 80 100 120 140 160 0 30 60 90 120 150 0 30 60 90 120 150 Time [days] Time [days] Time [days]

Figure 10. V-band absolute magnitude (left) and photosphere properties (right) for a 30M donor star involved in a q = 0.1 merger, in which we vary the inner mass coordinate of the 1051 erg of energy deposition and subsequent ejection within the Hydrogen envelope. Material inside Min does not explode and acts as a gravitational point mass, while material outside Min is expelled. We find that within a factor of two in Min, model light curves are very similar, varying primarily in total plateau duration (which results from differing ejecta masses).

20

18 V M 16

14 0 25 50 75 100 125 150

x 1.02 4 , V

M 1.00 / fiducial V 2x M 0.98 4x 0 25 50 75 100 125 150 Time [days]

Figure 11. Absolute V-band magnitude for a 30M donor star involved in a q = 0.1 merger. We show models with a fiducial resolution of 456 mass zones, and for mass resolutions of twice and four times as many zones. We find that the light curves are converged to within 1% for the bulk of the model light curves, with the largest variations at shock breakout and near the end of the light curve.

Blandford, R. D., & Znajek, R. L. 1977, MNRAS, 179, 433 Chugai, N. N. 1997, Astronomy Reports, 41, 672 Blondin, J. M., Lundqvist, P., & Chevalier, R. A. 1996, —. 2001, MNRAS, 326, 1448 ApJ, 472, 257 Chugai, N. N., & Danziger, I. J. 1994, MNRAS, 268, 173 Chandra, P. 2018, SSRv, 214, 27 Chugai, N. N., Blinnikov, S. I., Cumming, R. J., et al. 2004, Chen, W.-X., & Beloborodov, A. M. 2007, ApJ, 657, 383 MNRAS, 352, 1213 Chevalier, R. A. 1993, ApJ, 411, L33 Clayton, M., Podsiadlowski, P., Ivanova, N., & Justham, S. —. 1996, ApJ, 459, 322 2017, MNRAS, 470, 1788 —. 2012, ApJ, 752, L2 Chevalier, R. A., & Irwin, C. M. 2011, ApJ, 729, L6 Das, S., & Ray, A. 2017, ApJ, 851, 138 —. 2012, ApJ, 747, L17 de Kool, M. 1990, ApJ, 358, 189 21

De Marco, O., & Izzard, R. G. 2017, Publications of the Kasen, D., & Bildsten, L. 2010, ApJ, 717, 245 Astronomical Society of Australia, 34, e001 Kasen, D., & Woosley, S. E. 2009, ApJ, 703, 2205 de Mink, S. E., Sana, H., Langer, N., Izzard, R. G., & Kiewe, M., Gal-Yam, A., Arcavi, I., et al. 2012, ApJ, 744, Schneider, F. R. N. 2014, ApJ, 782, 7 10 Dessart, L., Audit, E., & Hillier, D. J. 2015, MNRAS, 449, Kleiser, I. K. W., Kasen, D., & Duffell, P. C. 2018, 4304 MNRAS, 475, 3152 Dexter, J., & Kasen, D. 2013, ApJ, 772, 30 Kochanek, C. S., Beacom, J. F., Kistler, M. D., et al. 2008, Eggleton, P. P. 1983, ApJ, 268, 368 ApJ, 684, 1336 Eldridge, J. J., Xiao, L., Stanway, E. R., Rodrigues, N., & Kruckow, M. U., Tauris, T. M., Langer, N., et al. 2016, Guo, N. Y. 2018, PASA, 35, 49 A&A, 596, A58 Ensman, L., & Burrows, A. 1992, ApJ, 393, 742 Lee, W. H., & Ramirez-Ruiz, E. 2006, ApJ, 641, 961 Fassia, A., Meikle, W. P. S., Chugai, N., et al. 2001, Levesque, E. M., Massey, P., Zytkow, A. N., & Morrell, N. MNRAS, 325, 907 2014, MNRAS, 443, L94 Feng, E.-H., Shen, R.-F., & Lin, W.-P. 2018, ApJ, 867, 130 LSST Science Collaboration, Abell, P. A., Allison, J., et al. Forster, F., Moriya, T. J., Maureira, J. C., et al. 2018, 2009, arXiv e-prints, arXiv:0912.0201 Nature Astronomy, 2, 808 MacFadyen, A. I., & Woosley, S. E. 1999, ApJ, 524, 262 Frank, J., King, A., & Raine, D. J. 2002, Accretion Power MacFadyen, A. I., Woosley, S. E., & Heger, A. 2001, ApJ, in Astrophysics: Third Edition 550, 410 Fryer, C. L., Benz, W., & Herant, M. 1996, ApJ, 460, 801 MacLeod, M., Antoni, A., Murguia-Berthier, A., Macias, Fryer, C. L., Rueda, J. A., & Ruffini, R. 2014, ApJ, 793, P., & Ramirez-Ruiz, E. 2017a, ApJ, 838, 56 L36 MacLeod, M., Macias, P., Ramirez-Ruiz, E., et al. 2017b, Fryer, C. L., & Woosley, S. E. 1998, ApJ, 502, L9 ApJ, 835, 282 Fryer, C. L., Woosley, S. E., & Hartmann, D. H. 1999, ApJ, MacLeod, M., Ostriker, E. C., & Stone, J. M. 2018a, ApJ, 526, 152 863, 5 Fuller, J. 2017, MNRAS, 470, 1642 —. 2018b, ApJ, 868, 136 Gilkis, A., Soker, N., & Kashi, A. 2019, MNRAS, 482, 4233 MacLeod, M., & Ramirez-Ruiz, E. 2015, ApJ, 798, L19 Ginzburg, S., & Balberg, S. 2012, ApJ, 757, 178 MacLeod, M., Vick, M., Lai, D., & Stone, J. M. 2018c, —. 2014, ApJ, 780, 18 arXiv e-prints, arXiv:1812.07594 Grichener, A., & Soker, N. 2018, arXiv e-prints, Marcaide, J. M., Mart´ı-Vidal,I., Perez-Torres, M. A., et al. arXiv:1810.03889 2009, A&A, 503, 869 Guillochon, J., Parrent, J., Kelley, L. Z., & Margutti, R. Margutti, R., Milisavljevic, D., Soderberg, A. M., et al. 2017, ApJ, 835, 64 2014, ApJ, 780, 21 Holgado, A. M., Ricker, P. M., & Huerta, E. A. 2018, ApJ, 857, 38 Margutti, R., Kamble, A., Milisavljevic, D., et al. 2017, Holoien, T. W. S., Brown, J. S., Vallely, P. J., et al. 2019, ApJ, 835, 140 MNRAS, 484, 1899 Margutti, R., Metzger, B. D., Chornock, R., et al. 2019, Houck, J. C., & Chevalier, R. A. 1991, ApJ, 376, 234 ApJ, 872, 18 Huang, S.-S. 1963, ApJ, 138, 471 Mason, E., Diaz, M., Williams, R. E., Preston, G., & Iben, Icko, J., & Livio, M. 1993, Publications of the Bensby, T. 2010, A&A, 516, A108 Astronomical Society of the Pacific, 105, 1373 Mattei, J., Johnson, G. E., Rosino, L., Rafanelli, P., & Ivanova, N., Justham, S., Chen, X., et al. 2013, Astronomy Kirshner, R. 1979, International Astronomical Union and Astrophysics Review, 21, 59 Circular, 3348, 1 Justham, S., Podsiadlowski, P., & Vink, J. S. 2014, ApJ, Matzner, C. D. 2003, MNRAS, 345, 575 796, 121 Mauerhan, J., Williams, G. G., Smith, N., et al. 2014, Kalogera, V., Belczynski, K., Kim, C., O’Shaughnessy, R., MNRAS, 442, 1166 & Willems, B. 2007, PhR, 442, 75 McDowell, A. T., Duffell, P. C., & Kasen, D. 2018, ApJ, Kalogera, V., Kim, C., Lorimer, D. R., et al. 2004, ApJ, 856, 29 601, L179 Metzger, B. D., & Pejcha, O. 2017, MNRAS, 471, 3200 Kangas, T., Mattila, S., Kankare, E., et al. 2016, MNRAS, Moe, M., & Di Stefano, R. 2017, The Astrophysical Journal 456, 323 Supplement Series, 230, 15 22

Montes, M. J., Weiler, K. W., Van Dyk, S. D., et al. 2000, Popham, R., Woosley, S. E., & Fryer, C. 1999, ApJ, 518, ApJ, 532, 1124 356 Moriya, T. J. 2018, MNRAS, 475, L49 Popov, D. V. 1993, ApJ, 414, 712 Moriya, T. J., Blinnikov, S. I., Tominaga, N., et al. 2013a, Prieto, J. L., Brimacombe, J., Drake, A. J., & Howerton, S. MNRAS, 428, 1020 2013, ApJ, 763, L27 Moriya, T. J., & Maeda, K. 2012, ApJ, 756, L22 Quataert, E., Fern´andez,R., Kasen, D., Klion, H., & Moriya, T. J., Maeda, K., Taddia, F., et al. 2013b, Paxton, B. 2016, MNRAS, 458, 1214 MNRAS, 435, 1520 Quataert, E., & Kasen, D. 2012, MNRAS, 419, L1 Morozova, V., Ott, C. D., & Piro, A. L. 2015a, SNEC: Quataert, E., & Shiode, J. 2012, MNRAS, 423, L92 SuperNova Explosion Code, , , ascl:1505.033 Salpeter, E. E. 1955, ApJ, 121, 161 Morozova, V., Piro, A. L., Renzo, M., et al. 2015b, ApJ, Sana, H., de Mink, S. E., de Koter, A., et al. 2012, Science, 814, 63 337, 444 Morozova, V., Piro, A. L., & Valenti, S. 2017, ApJ, 838, 28 Senno, N., Murase, K., & M´esz´aros,P. 2016, PhRvD, 93, —. 2018, ApJ, 858, 15 083003 Morozova, V., & Stone, J. M. 2018, ApJ, 867, 4 Shiode, J. H., & Quataert, E. 2014, ApJ, 780, 96 Murguia-Berthier, A., MacLeod, M., Ramirez-Ruiz, E., Shivvers, I., Groh, J. H., Mauerhan, J. C., et al. 2015, ApJ, Antoni, A., & Macias, P. 2017, ApJ, 845, 173 806, 213 Shu, F. H., Lubow, S. H., & Anderson, L. 1979, ApJ, 229, Murguia-Berthier, A., Montes, G., Ramirez-Ruiz, E., De 223 Colle, F., & Lee, W. H. 2014, ApJ, 788, L8 Siegel, D. M., Barnes, J., & Metzger, B. D. 2018, arXiv Ofek, E. O., Lin, L., Kouveliotou, C., et al. 2013a, ApJ, e-prints, arXiv:1810.00098 768, 47 Smith, N. 2017, Interacting Supernovae: Types IIn and Ibn, Ofek, E. O., Fox, D., Cenko, S. B., et al. 2013b, ApJ, 763, 403 42 Smith, N., & Arnett, W. D. 2014, ApJ, 785, 82 Ofek, E. O., Sullivan, M., Cenko, S. B., et al. 2013c, Smith, N., Mauerhan, J. C., & Prieto, J. L. 2014, MNRAS, Nature, 494, 65 438, 1191 Ofek, E. O., Arcavi, I., Tal, D., et al. 2014, ApJ, 788, 154 Smith, N., & McCray, R. 2007, ApJ, 671, L17 Opik,¨ E. 1924, Publications of the Tartu Astrofizica Smith, N., Miller, A., Li, W., et al. 2010, AJ, 139, 1451 Observatory, 25 Soker, N. 2019, arXiv e-prints, arXiv:1902.01187 Paczynski, B. 1976, in IAU Symposium, Vol. 73, Structure Soker, N., & Gilkis, A. 2018, MNRAS, 475, 1198 and Evolution of Close Binary Systems, ed. P. Eggleton, Soker, N., Grichener, A., & Gilkis, A. 2019, MNRAS, 484, S. Mitton, & J. Whelan, 75 4972 Paczynski, B. 1983, ApJ, 267, 315 Song, C.-Y., & Liu, T. 2019, ApJ, 871, 117 Paczy´nski,B., & Sienkiewicz, R. 1972, AcA, 22, 73 Stevenson, S., Vigna-G´omez,A., Mandel, I., et al. 2017, Pan, T., Patnaude, D., & Loeb, A. 2013, MNRAS, 433, 838 Nature Communications, 8, 14906 Patnaude, D. J., Loeb, A., & Jones, C. 2011, New Sukhbold, T., Woosley, S. E., & Heger, A. 2018, ApJ, 860, Astronomy, 16, 187 93 Pejcha, O. 2014, ApJ, 788, 22 Taam, R. E., Bodenheimer, P., & Ostriker, J. P. 1978, ApJ, Pejcha, O., Metzger, B. D., & Tomida, K. 2016a, MNRAS, 222, 269 455, 4351 Taam, R. E., & Sandquist, E. L. 2000, Annual Review of —. 2016b, MNRAS, 461, 2527 Astronomy and Astrophysics, 38, 113 Pejcha, O., Metzger, B. D., Tyles, J. G., & Tomida, K. Taddia, F., Stritzinger, M. D., Sollerman, J., et al. 2013, 2017, ApJ, 850, 59 A&A, 555, A10 Pejcha, O., & Thompson, T. A. 2012, ApJ, 746, 106 Tauris, T. M., Kramer, M., Freire, P. C. C., et al. 2017, Podsiadlowski, P. 2007, in Astronomical Society of the ApJ, 846, 170 Pacific Conference Series, Vol. 367, Massive Stars in Terman, J. L., Taam, R. E., & Hernquist, L. 1995, ApJ, Interactive Binaries, ed. N. St. -Louis & A. F. J. Moffat, 445, 367 541 The LIGO Scientific Collaboration, & the Virgo Podsiadlowski, P., Cannon, R. C., & Rees, M. J. 1995, Collaboration. 2018a, arXiv e-prints, arXiv:1811.12940 MNRAS, 274, 485 —. 2018b, arXiv e-prints, arXiv:1811.12907 23

Thorne, K. S., & Zytkow, A. N. 1977, ApJ, 212, 832 Weiler, K. W., Sramek, R. A., Panagia, N., van der Hulst, Tylenda, R., Hajduk, M., Kami´nski,T., et al. 2011, A&A, J. M., & Salvati, M. 1986, ApJ, 301, 790 528, A114 Woosley, S. E., & Heger, A. 2012, ApJ, 752, 32 Vigna-G´omez,A., Neijssel, C. J., Stevenson, S., et al. 2018, MNRAS, 481, 4009 Zhang, W., & Fryer, C. L. 2001, ApJ, 550, 357 Webbink, R. F. 1984, ApJ, 277, 355