Draft version October 22, 2020 Typeset using LATEX twocolumn style in AASTeX63

Constraints from detections of binary mergers on the 12C(α, γ)16O rate

R. Farmer,1, 2 M. Renzo,3 S. E. de Mink,2, 1 M. Fishbach,4 and S. Justham5, 6, 1

1Anton Pannekoek Institute for Astronomy and GRAPPA, University of Amsterdam, NL-1090 GE Amsterdam, The Netherlands 2Center for Astrophysics — Harvard & Smithsonian, 60 Garden Street, Cambridge, MA 02138, USA 3Center for Computational Astrophysics, Flatiron Institute, New York, NY 10010, USA 4Department of Astronomy and Astrophysics, University of Chicago, Chicago, IL 60637, USA 5School of Astronomy & Space Science, University of the Chinese Academy of Sciences, Beijing 100012, China 6National Astronomical Observatories, Chinese Academy of Sciences, Beijing 100012, China

(Dated: October 22, 2020)

ABSTRACT Gravitational wave detections are starting to allow us to probe the physical processes in the evolution of very massive through the imprints they leave on their final remnants. Stellar evolution theory predicts the existence of a gap in the black hole mass distribution at high mass due to the effects of pair- instability. Previously, we showed that the location of the gap is robust against model uncertainties, but it does depend sensitively on the uncertain 12C(α, γ)16O rate. This rate is of great astrophysical significance and governs the production of oxygen at the expense of carbon. We use the open source MESA stellar evolution code to evolve massive helium stars to probe the location of the mass gap. We find that the maximum black hole mass below the gap varies between 40 M to 90 M , depending on the strength of the uncertain 12C(α, γ)16O reaction rate. With the first ten gravitational-wave detections of black holes, we constrain the astrophysical S-factor for 12C(α, γ)16O, at 300keV, to S300 > 175 keV barns at 68% confidence. With (50) detected binary black hole mergers, we expect to constrain the S-factor to within 10–30 keV barns.O We also highlight a role for independent constraints from electromagnetic transient± surveys. The unambiguous detection of pulsational pair instability supernovae would imply that S300 > 79 keV barns. Degeneracies with other model uncertainties need to be investigated further, but probing nuclear stellar astrophysics poses a promising science case for the future gravitational wave detectors. 1. INTRODUCTION and 16O in the Universe (Boothroyd & Sackmann 1988; The 12C(α, γ)16O reaction is one of the most impor- Weaver & Woosley 1993; Thielemann et al. 1996) tant nuclear reaction rates (Burbidge et al. 1957) in the Thus improving our understanding of this key rate is evolution of stars yet also one of the most uncertain of critical importance to stellar astrophysics. The diffi- (Holt et al. 2019). Reducing the uncertainty on this culty in measuring the rate occurs due to the negligible rate has been dubbed “the holy grail of nuclear astro- cross section of the reaction at temperatures relevant for physics” (deBoer et al. 2017; Bemmerer et al. 2018). It helium burning in stars (An et al. 2015, 2016). Thus nu- plays a key role in governing the evolution and com- clear experiments can only provide data for much higher energies (i.e. temperatures) from which we extrapolate arXiv:2006.06678v2 [astro-ph.HE] 21 Oct 2020 position of stars beyond the main sequence, from the C/O ratio in white dwarfs (Salaris et al. 1997; Straniero down to astrophysically relevant energies. However, the et al. 2003; Fields et al. 2016), whether a will form a cross-section has a complex energy dependence and thus or a black hole (Brown et al. 2001; Woosley is not easily extrapolated to lower temperatures (de- et al. 2002; Heger et al. 2002; Tur et al. 2007; West et al. Boer et al. 2017; Friˇsˇci´cet al. 2019). Recent lab mea- 2013; Sukhbold & Adams 2020), and the amount of 12C surements, with high beam luminosities, though indirect studies of the excited states of 16O, and improved theo- retical modeling of the 12C(α, γ)16O rate have begun to [email protected] reduce the uncertainty on the rate (Hammer et al. 2005; An et al. 2016; Hammache et al. 2016; deBoer et al. 2017; Shen et al. 2020). New experiments will soon be 2 better able to probe the 12C(α, γ)16O reaction rate at star’s evolution before pulsations begin, and how this is astrophysically relevant temperatures (Holt et al. 2018; altered by the 12C(α, γ)16O reaction rate. In Section4 Friˇsˇci´cet al. 2019). we show how the maximum black hole mass below the Astrophysical studies using white dwarfs have at- gap is affected by the nuclear redaction rates and place tempted to place constraints on the 12C(α, γ)16O reac- constraints on the 12C(α, γ)16O reaction rate in Section tion rate, using astroseismology of white dwarfs (Met- 5. In Section6 we discuss how these results will improve calfe et al. 2001, 2002; Metcalfe 2003). However, these with future gravitational wave detections. In Section7 measurements are sensitive to other physics choices, we discuss potentially other observables that can be used like semiconvection and convective overshoot mixing to constrain the 12C(α, γ)16O. Finally, in Sections8& (Straniero et al. 2003), that are poorly constrained. 9, we discuss and summarize our results. Thus a cleaner signal is needed to provide a more ro- bust estimate from stellar astrophysical sources. Merging black holes detected by LIGO/Virgo (LIGO 2. METHOD Scientific Collaboration et al. 2015; Acernese et al. 2015) There are many channels for the formation of a source can provide such a signal, via the location of the pair- detectable by ground-based gravitational-wave detec- instability mass gap (Takahashi 2018; Farmer et al. tors. We consider here the case where the progenitors 2019). A gap is predicted to form in the mass distri- of the merging black holes have come from an isolated bution of black holes, due to pair-instability supernovae binary system. There are multiple stellar pathways for (PISN) completely disrupting massive stars, leaving be- this to produce a successful binary black hole merger, in- hind no remnant (Fowler & Hoyle 1964; Barkat et al. cluding common-envelope evolution (Tutukov & Yungel- 1967; Woosley 2017). The lower edge of the gap is set son 1993; Dominik et al. 2012; Belczynski et al. 2016a), by mass loss experienced by a star during pulsational chemically homogeneous evolution (de Mink & Mandel pair instability supernovae (Rakavy & Shaviv 1967; Fra- 2016; Marchant et al. 2016; Mandel & de Mink 2016), or ley 1968; Woosley et al. 2002). These objects undergo stars which interact in dynamic environments (Kulkarni multiple mass-loss phases before collapsing into a black et al. 1993; Portegies Zwart & McMillan 2000; Gerosa hole (Woosley et al. 2002; Chen et al. 2014; Yoshida & Berti 2019) In each case, we expect the stars to lose et al. 2016; Woosley 2017; Marchant et al. 2019; Farmer their hydrogen envelopes after the end of the star’s main et al. 2019). sequence, leaving behind the helium core of the star. In Farmer et al.(2019) we evolved hydrogen-free he- We use the stellar evolution code, version 11701 lium cores and found that the lower edge of the PSIN MESA (Paxton et al. 2011, 2013, 2015, 2018, 2019), to fol- black hole mass gap was robust to changes in the metal- low through the various stages of nuclear burning in- licity and other uncertain physical processes, e.g wind side these helium cores until they either collapse to mass loss and chemical mixing. Over the range of metal- form a black hole or explode as a PISN. We follow the licities considered the maximum black hole mass de- evolution of helium cores with initial masses between creased by 3 M . We also showed that the choices for M = 30 M and M = 200 M in steps of many other uncertain physical processes inside stars do He,init He,init 1 M , at a metallicity of Z = 10−5. We use the de- not greatly affect the location of the PISN mass gap. fault model choices from Farmer et al.(2019) for set- The existence of a gap in the mass distribution of ting all other input parameters. See Appendix merging binary black holes (BBH) would provide strong MESA A for further details of our usage of , and our in- constraints on their progenitors, and hence of the post MESA put files with all the parameters we set can be found at main-sequence evolution of stars, which includes the ef- 16 https://zenodo.org/record/3559859. fect of 12C(α, γ) O on a star’s evolution (Takahashi After a star has formed its helium core, it begins burn- 2018; Sukhbold & Adams 2020). The existence of a gap ing helium in its central region, converting 4He into 12C in the mass distribution can also be used as a “stan- and then 16O. The final ratio of the mass fractions of dardizable sirens” for cosmology and used to place con- 12C/16O depends on the relative strengths of the 3α re- straints on the Hubble constant (Schutz 1986; Holz & action rate, which produces 12C, and 12C(α, γ)16O re- Hughes 2005; Farr et al. 2019). action rate, which converts the 12C into 16O. We define Here we investigate how the maximum black hole the end of core helium burning to occur when the central mass below the PISN mass gap is sensitive to the 4 −4 16 mass fraction of He drops below 10 . The core is now 12C(α, γ) O nuclear reaction rate and thus can be used dominated by 12C and 16O, with only trace mass frac- to place constraints on the reaction rate. In Section2 we tions of other nuclei. This core then begins a phase of discuss our methodology. In Section3 we describe the contraction, and thermal neutrino losses begin to dom- 3 inate the total energy loss from the star (Fraley 1968; 2019; Renzo et al. 2020a). We follow the ejected ma- Heger et al. 2003, 2005; Farmer et al. 2016). terial until the bound portion of the star relaxes back As the core contracts the central density and tem- into hydrostatic equilibrium, after it has radiated away perature increases which, for sufficiently massive cores, the energy of the pulse. We remove the material that causes the core to begin producing copious amounts of has become unbound from our computational grid by electron-positron pairs (e±). The production of the e± generating a new stellar model with the same entropy removes photons which were providing pressure support, and chemical distribution as the remaining bound ma- softening the equation of state in the core, and causes terial. We evolve this new star assuming hydrostatic the core to contract further. We then follow the dy- equilibrium until either another pulse occurs or the core 9.6 namical collapse of the star, which can be halted by temperature (Tc) exceeds Tc > 10 K, as the star is ap- the ignition of oxygen leading to either a PPISN or a proaching core collapse. At which point we switch back PISN. We follow the core as it contracts and bounces to using the hydrodynamic solver, We define the final (Marchant et al. 2019), generating shock waves that we core collapse to occur when any part of the star begins follow through the star until they reach the outer lay- collapsing with a velocity v > 8000km s−1, so that any ers of the star. These shocks then cause mass loss from pulse that is in the process of being ejected during core the star to occur, as material becomes unbound. In this collapse is resolvable. case we find that the star can eject between 0.1 M and Stars with core masses above MHe,init ' 120 M at- 20 M of material in a pulsational mass loss episode tempt to undergo a PISN, however sufficient energy is (∼Woosley 2017; Farmer et al. 2019; Renzo et al. 2020a). released during the pulse that the core heats to the point Stars at the boundary between core collapse and PPISN where photo-distintegrations become the dominant en- may generate weak pulses, due to only a small amount ergy sink. These reactions then reduce the available of material becoming dynamicaly unstable, and there- energy, which was powering the outward moving shock, fore do not drive any appreciable mass loss (Woosley and prevents the envelope from becoming unbound. The 2017). We use the term PPISN only for an event which star then collapses without significant mass loss. We as- ejects mass (Renzo et al. 2020a). PISN are stars for sume that this forms a black hole (Bond et al. 1984; which the energy liberated by the thermonuclear explo- Woosley 2017). sion of oxygen (and carbon) exceeds the total binding We define the mass of the black hole formed to be the energy, resulting in total disruption after only one mass mass of the bound material of the star at collapse. Given loss episode. the uncertain black hole formation mechanism (Fryer As a star evolves into the pair instability region we 1999; Fryer et al. 2001, 2012), or weak shock generation switch to using MESA’s Riemen contact solver, HLLC, (Nadezhin 1980; Lovegrove & Woosley 2013; Fern´andez (Toro et al. 1994; Paxton et al. 2018), to follow the et al. 2018), our black holes masses are upper limits. We hydrodynamical evolution of each pulse. This switch take the bound mass not the total mass, as some stars occurs when the volumetric pressure-weighted average are under going a mass ejection from a pulsation at the adiabatic index Γ1 4/3 < 0.01, which occurs slightly time of core collapse (Renzo et al. 2020a). before the starh entersi − the pair instability region. The adiabatic index, Γ1 is defined as 2.1. Nuclear reaction rates d ln P Γ1 = (1) Nuclear reaction rates play a key role in the evolution d ln ρ s and final fate of a star. However, they are also uncertain where P , ρ are the local pressure and density and is and this uncertainty varies as function of temperature evaluated at a constant entropy s . We used the continu- (Iliadis et al. 2010a,b; Longland et al. 2010). Varying ity equation to transform the volumetric integral of Γ 1 nuclear reaction rates within their known uncertainties into an integral over the mass domain, thus (Stothers has been shown to a have large impact on the structure 1999): of a star (Hoffman et al. 1999; Iliadis et al. 2002). 3 P To sample nuclear reaction rates within their known Γ1P d r Γ1 ρ dm Γ1 (2) uncertainties, we use the STARLIB (Sallaska et al. 2013) h i ≡ P d3r ≡ P dm R R ρ (version 67a) library. STARLIB provides both the me- We follow the dynamicalR evolutionR of the star, until dian reaction rate and uncertainty in that reaction as all shocks have reached the surface of the star. These a function of temperature. We sample each reaction at shocks may unbind a portion of the outer stellar enve- a fixed number of standard deviations from the median lope, resulting in mass loss (Yoshida et al. 2016; Woosley (Evans et al. 2000). We assume that the temperature- 4 dependent uncertainty in a reaction follows a log-normal 200 distribution Longland et al.(2010). 175 CC For each reaction rate tested, we create a sampled reaction rate at 60 points log-spaced in temperature be- 150 tween, 0.01 and 10 109 K(Fields et al. 2016, 2018). ] × The rate of a reaction per particle pair is given by: [M 125 Intial

, PISN

He 100 1/2 8 NA M NA σν = h i πµ (3/2) 75   (kBT ) ∞ CC σ (E) E exp(−E/kB T ) dE (3) 50 PPISN Z0 where µ is the reduced mass of the particles, E = 3 2 1 0 1 2 3 − − − σC12 µν2/2 is the center of mass energy, ν is the average ve- locity of the particles, NA is Avogadro’s number, and Figure 1. Final fate of a star as function of the initial 12 16 kB is the Boltzmann constant (e.g Lippuner & Roberts helium core mass and C(α, γ) O rate. σC12 denotes how 12 16 2017; deBoer et al. 2017; Holt et al. 2019). We can fac- far the C(α, γ) O is from the median STARLIB rate, mea- tor out the energy dependent cross-section σ (E) E, by sured in standard deviations. Blue regions indicate stars which undergo core collapse (CC) below the pair instability replacing it with the astrophysical S-factor: supernovae (PISN) mass gap, green regions form black holes after a pulsational pair instability supernovae (PPISN), while 2πη S(E) = σ (E) E exp (4) white regions are completely disrupted in a PISN, and mod- and els in the orange region form black holes from core collapse µ e2 for stars above the PISN mass gap. There are 2210 models, η = Z Z (5) in the grid spaced by 1 M and 0.5σ . 2E 1 2 C12 r ~ where η is the Summerfield parameter, Z the proton 1,2 See Section7 for a discussion of the implications of this charge of each particle, e is the electric charge, is the ~ for the black hole formation and EM transient rate. reduced Planck’s constant. The exp2πη term accounts for (approximately) the influence of the Coulomb barrier on the cross-section. As the S-factor depends on energy, 200 we quote it at the typical energy for a reaction. For CC 12C(α, γ)16O the typical energy is E = 300 keV. 175 1 10− 3. PRE-SN CARBON BURNING 150 ]

In Fig.1 we show the outcome for our grid of 2210  [M

125 C

evolutionary models as a function of the initial helium 12

16 Initial 12 , PISN core mass and the C(α, γ) O reaction. We parame- X He 100 2 12 16 10− terize the C(α, γ) O in terms of the number of sigmas M 12 16 (σC12) from the median STARLIB C(α, γ) O reaction 75 rate: 50 CC PPISN

µ(T )σC12 σν = σν mediane (6) h i h i 3 2 1 0 1 2 3 − − − σ where σν median is the median reaction rate provided C12 h i by STARLIB, and µ (T ) is the temperature-dependent un- Figure 2. The mass fraction of 12C in the core after core certainty in the reaction (which is assumed to follow a helium burning, but before carbon burning, for all initial log-normal distribution). masses as a function of σ . White lines denote the bound- 16 C12 For higher initial core mass (for a given 12C(α, γ) O aries between the different end fates. Text labels denote the rate), the final fate of a star transitions from core col- final fate of the star. Grey boxes denote models that do not lapse, to PPISN, to PISN, and then to core collapse evolve beyond core helium burning, defined as when the mass 4 −4 again (Bond et al. 1984). As the reaction rate increases fraction of He at the center of the star drops below 10 . (i.e, large values of σC12) the boundary between the dif- ferent end fates shifts to lower initial helium core mass. 5

Helium shell Center Carbon Off-center Carbon Explosive Oxygen Center Oxygen

Reduced

Core Collapse

Median Pulsations t n a n

Enhanced Pair m e

Instability r

SNe o N

(A) (B) (C) (D) (E)

Figure 3. A schematic of the time progression of the major fuel burning, for a star with MHe,init = 60 M as a function 12 16 of σC12, at reduced (σC12 = −3), median (σC12 = 0), and enhanced (σC12 = 3) C(α, γ) O reaction rates. Yellow regions denote helium burning, orange regions denote carbon burning, and red regions denote oxygen burning. Hatched regions indicate convection mixing regions. Letters match those points marked in Fig.4. Also shown is the core’s C/O mass ratio at the end of core helium burning. To understand the reason for these trends, it is in- proceeding to ignite oxygen explosively. For the highest sightful to consider the 12C mass fraction in the core 12C(α, γ)16O shown, no carbon is burnt before oxygen of the stellar models, after core helium burning has fin- ignites (bottom panel). The C/O ratios shown are de- ished. We define the end of core helium burning, when fined at the end of core helium burning. the mass fraction of 4He at the center of the star drops Why changing the carbon burning behavior changes below 10−4. Figure2 shows that the 12C mass frac- the final outcome for a star can be seen in Fig.4. tion in the core of the stars considered here decreases Here we show the time evolution of the helium cores, for 12 16 from 30% to 0.001% as the C(α, γ) O rate is stars with MHe,init = 60 M during carbon burning and ≈ ≈ 12 16 increased from σC12 = 3 to σC12 = 3, independent up to the ignition of oxygen, for different C(α, γ) O of initial mass. This change− in the 12C mass fraction rates. The top row shows a Kippenhan diagram of the is what drives the changes in the star’s later phases of time evolution of the net nuclear energy minus neutrino evolution and thus final fate. losses and the mixing regions inside the star. The mid- After core helium burning has ceased the core be- dle row shows the evolution of Γ1 4/3. Regions where − gins contracting, increasing its density and temperature. Γ1 4/3 < 0 are locally unstable. The bottom row However, at the same time, thermal neutrino losses in- shows− the temperature and density structure inside the crease which acts to cool the core. The next fuel to burn star at points in time marked on the top row of Fig. is, 12C via 12C+12C to 22Ne, 23Na, and 24Mg (Arnett 4. The points in time marked, show how the stars are & Truran 1969; Farmer et al. 2015). As the 12C + 12C evolving on different timescales. Timescales vary from a reaction rate depends on the number density of carbon few thermal timescales (left column) to a few dynamical squared, small changes in the number density of 12C timescales (middle and right columns). can have a large impact on the power generated by the When the 12C(α, γ)16O rate is small (and thus the 12C + 12C reaction. 12C mass fraction is 30%, with the rest of the core In Fig.3 we show a simplified picture of the steps a being made of 16O), 12≈C ignites vigorously at the cen- star takes to its final fate depending on its 12C(α, γ)16O ter in a radiative region (Fig.4 top-left) and burns reaction rate. The top panel shows a star which would outwards until it begins to drive a convective 12C burn- undergo core collapse, first by igniting carbon both at ing shell. The star will then ignite oxygen in the core the center and then in a off-center shell. This star (non-explosively) and proceed through silicon and iron then avoids igniting oxygen explosively, instead pro- burning before collapsing in a core collapse. ceeds through its evolution to core collapse. As the As the 12C(α, γ)16O rate increases the initial 12C ig- 12C(α, γ)16O increases, the carbon stops igniting at the nition point, defined where the nuclear energy gener- center and only ignites in a shell (middle panel), before ated is greater than the energy lost in neutrinos, moves 6

Core Collapse (σ = 3.0) PPISN (σ = 0.0) PISN (σ = 3.0) C12 − C12 C12 1 yr 1 yr 1 month 1 yr 1 day 1 min 10 days 1 min Time Time Time

A B C E A C D E A D E 60 60 60 15 15 15 Helium Helium Helium 45 10 45 10 45 10 ) ) ) | | |    5 | 5 | 5 | ( ( ( ] ] ]

M M M [ 30 0 [ 30 0 [ 30 0 m m m 5 5 5 − − − signed log signed log signed log 15 Carbon 10 15 Carbon 10 15 10 − − − Oxygen 15 Oxygen 15 Oxygen 15 − − − 0 0 0 2400 3000 3600 4200 4800 2400 2800 3200 3600 4000 2800 3200 3600 4000 Model Number Model Number Model Number

A B C E A C D E A D E 60 0.06 60 0.06 60 0.06

0.04 0.04 0.04 45 45 45 0.02 0.02 0.02 3 3 3 / / / ] ] ]

4 4 4 M M M [ 30 0.00 [ 30 0.00 [ 30 0.00 − − − m m m 1 1 1 Γ Γ Γ 0.02 0.02 0.02 − − − 15 15 15 0.04 0.04 0.04 − − −

0 0.06 0 0.06 0 0.06 2400 3000 3600 4200 4800 − 2400 2800 3200 3600 4000 − 2800 3200 3600 4000 − Model Number Model Number Model Number

9.6 9.6 9.6 E 9.5 9.5 E 9.5 D 9.4 9.4 9.4 D E 9.3 Γ1 < 4/3 9.3 Γ1 < 4/3 9.3 Γ1 < 4/3 ) ) ) K K K / / C / T C T T ( 9.2 ( 9.2 ( 9.2 Time Time Time log log log 9.1 9.1 9.1 A B A 9.0 A 9.0 9.0

8.9 8.9 8.9 P P P P P P rad ≈ gas rad ≈ gas rad ≈ gas 8.8 8.8 8.8 3 4 5 6 7 3 4 5 6 7 3 4 5 6 7 3 3 3 log ρ/ g cm− log ρ/ g cm− log ρ/ g cm−   

Figure 4. The time evolution of the internal structure of a star for MHe,init = 60 M for different assump- tions for the 12C(α, γ)16O reaction rate. The top row shows the signed logarithm of the net specific power, i.e. −1 −1 sign (nuc − ν ) log10 max (1.0, |nuc − ν |) / erg g s , where nuc is the specific power from nuclear reactions and ν is the specific power lost via neutrinos. Purple regions denote strong neutrino cooling and red regions denote regions of strong nuclear burning. Hatched regions indicate convective mixing regions. Text labels state the primary fuel burned in that region. Points marked on the top x axis correspond to those marked in Fig.3, given with approximate timescales. The middle row shows the evolution of Γ1 (Equation1), regions with Γ 1 −4/3 < 0 are locally unstable. The bottom row shows the density-temperature structure of the inner region of the stars at the points marked in the top panel, light colors denote later phases. The dashed line shows where the gas pressure is approximately equal to the radiation pressure. The solid black line that encloses Γ1 < 4/3 shows the approximate location of the pair instability region. outwards in mass coordinate (Fig.4 top-center). As Stars with a convective 12C burning shell can resist the the 12C abundance decreases, the star requires a higher collapse caused by the production of e±, and thus main- density to burn 12C vigorously, thus the star must con- tain hydrostatic equilibrium until core collapse. Figure tract further which increases the neutrino losses. No 4 (middle-left) shows that when the shell forms it pre- convective carbon shell forms before the oxygen in the vents the center of the star from reaching Γ1 4/3 < 0. core ignites explosively, proceeding to a PPISN. Once Therefore, the instability is only local and never− becomes the 12C(α, γ)16O rate increases sufficiently such that global: only a small region around the carbon shell be- the core is depleted in 12C after core helium burning, comes unstable. For stars without the convective carbon no 12C burning region forms and the core proceeds to shell (middle-center and middle-right), a significant frac- ignite oxygen explosively (Fig.4 top-right) as a PISN tion of the entire star becomes unstable resulting in a leaving behind no black hole remnant. global instability. 7

Carbon burning begins at the center and moves out- oxygen ignites the core is now outside the instability wards (either vigorously or not), thus depleting the cen- region (Γ1 4/3 > 0) and as such does not undergo ter of 12C. Therefore, the carbon shell (if it forms) can pulsational− mass loss (Takahashi 2018). not move inwards as there is insufficient fuel for it to For stars without a convective carbon shell (bottom- burn. The region undergoing carbon burning can not center and bottom-right) the core continues to contract move outwards either, as the convection zone is mix- homogeneously. When oxygen ignites it does so inside ing the energy released from the nuclear reactions over the Γ1 4/3 < 0 region. As the temperature increases, a significant portion of the star. This prevents layers due to− the oxygen burning, the production of e± in- above the carbon burning region from reaching sufficient creases causing a positive feedback loop. This leads to temperatures and densities needed for vigorous carbon the explosive ignition needed to drive a pulse. Stars un- burning (Farmer et al. 2015; Takahashi 2018). dergoing a PPISN (bottom-center) have slightly lower The convective carbon shell can only be sustained then core entropies than stars undergoing a PISN (bottom- if it can bring fresh fuel in via convective mixing from right) due to the small amount of non-convective carbon the rest of the core. Thus when a convective carbon shell burning that occurs before oxygen burning begins. forms it also allows additional fuel to be mixed into the Further decreases in the carbon abundance leave little burning region from the outer layers of the core. This carbon fuel to burn. Thus as the star collapses due to the prolongs the lifetime of the carbon burning shell and production of e±, the oxygen is free to ignite violently. prevents the collapse due to e± from occurring, until the This causes the star to undergo a PISN, completely dis- carbon shell convective region is depleted in 12C which rupting the star. This can also be seen in Figure2 where may not occur before the star undergoes core collapse. the boundaries between the different final fates move to As the carbon fraction decreases, the carbon shell lower masses as σC12 increases, as the pulses become burning becomes less energetic (due to the 12C + 12C more energetic for a given initial mass. reaction depending on the density of carbon squared). 12 Therefore, as σC12 increases (and C fraction decreases) 3.1. Black holes above the PISN mass gap less energy is released from the carbon burning, thus the Figure1 shows population of black holes that form fraction of the core where Γ 4/3 < 0 increases There 1 above the PISN mass gap. These black holes form due becomes a critical point where− the carbon burning is to the failure of the PISN explosion to fully unbind the insufficient to prevent the violent ignition of oxygen in star (e.g., Bond et al. 1984). As the helium core mass is the core, thus pulsations begin. Around this critical increased (at constant σ ) a PISN explosion increases region, convective carbon burning can still occur, how- C12 in energy, due to a greater fraction of the oxygen be- ever the burning region can undergo flashes, where the ing burnt in the oxygen ignition. This increased energy 12C ignites but is then quenched. This leads to weaker leads to an increase in the maximum core temperature and shorter lived convection zones which do not mix in the star reaches, before the inward collapse is reverted sufficient 12C to sustain a continuous carbon burning and the star becomes unbound. This increased temper- shell. This leads to very weak pulses removing only a ature can be seen in the increased production of 56Ni as few tenths of a solar mass of material. For stars with the initial mass increases (Woosley et al. 2002; Renzo these weak convection zones the carbon shell only de- et al. 2020a). Eventually the core reaches sufficent tem- lays the ignition of oxygen; once carbon is sufficiently peratures that the energy extracted from the core by depleted the oxygen can ignite explosively. Eventually photo disintegrations is sufficent to prevent the star from no carbon shell convection zone is formed at all (Fig.4 becoming unbound. middle-center), this leads to larger pulses removing solar Figure1 shows that as σ increases the initial he- masses of material. C12 lium core mass needed to form a black hole above the The bottom row of Fig.4 shows the temperature- gap decreases. This is due to the increased production density profile inside the star at moments marked in of oxygen as σ increases. At these masses we do the top row of Fig.4. As the stars evolve the core C12 not see the formation of a convective carbon shell, even contracts and heats up, eventually the central regions of for σ = 3, though some radiative carbon burning the star enter the instability region (Γ 4/3 < 0). Once C12 1 occurs (similar− to point C in the middle panels of Fig- a convective carbon shell forms (bottom-left)− the core ure4). Instead the cores with greater total amounts of stops contracting homogeneously (along the line where oxygen can liberate greater amounts of energy from the the radiation pressure is equal to the gas pressure) and oxygen burning. This burning raises the temperature moves to higher densities. This is due to the continued in the core, allowing additional nuclosynethesis to oc- loss of entropy to neutrinos from the core. Thus when cur with the silicon and iron group elements produced 8 from the oxygen burning. For a fixed initial helium core 175 mass as σC12 increases, the peak core temperature in- BHs creases. Once a core reaches log T/K 9.86, then the ≈ 150 rate of photodisintegrations is sufficient to prevent the Mass Gap 125 star from unbinding itself. This is what sets the upper ] edge of the mass gap. [M 100 BH

4. EDGES OF THE PISN MASS GAP M 75 Figure5 shows the location of the PISN black hole Mass Gap 50 mass gap as a function of the temperature-dependent BHs uncertainty in 12C(α, γ)16O. As the rate increases (with 25 increasing σC12) both the lower and upper edge of the PISN mass gap shift to lower masses, from 90 M to 3 2 1 0 1 2 3 ≈ − − − σC12 40 M for the lower edge and 175 M to 120 M ≈for the upper edge. The width of≈ the region remains≈ ap- GW150914 GW151012 GW151226 GW170104 GW170608 GW170729 GW170809 GW170814 GW170817 GW170818 GW170823 +5 proximately constant at 83−8 M . The typical quoted Figure 5. The location of the PISN mass gap as a value for the maximum mass of a black hole below function of the temperature-dependent uncertainty in the 12C(α, γ)16O reaction rate. The white region denotes the the PISN mas gap is 45 55 M (Yoshida et al. 2016; Woosley 2017; Leung et al.− 2019; Marchant et al. 2019; mass gap, while the grey horizontal bar region denotes the mass range where we can not place a black hole for any value Farmer et al. 2019) The gray box in Fig.5 shows the re- of the 12C(α, γ)16O rate. The side plot shows the inferred gion of black hole masses, between MBH 90 120 M , ≈ − masses of the currently known black holes from LIGO/Virgo where we can not place a black hole from a first genera- O2 (Abbott et al. 2019a), with their 90% confidence inter- tion core collapse or PPISN model. Thus black holes de- vals. The red region shows the 90% confidence range for the tected in this mass region would need to come from alter- inferred location of the lower edge of the PISN mass gap from nate formation mechanisms, for instance; either second the O1/O2 data (Abbott et al. 2019b). generation mergers (Rodriguez et al. 2016, 2019; Gerosa & Berti 2019), primordial black holes (Carr et al. 2016; where RLow and RHigh are from Kunz et al.(2002). The Ali-Ha¨ımoudet al. 2017), or accretion on to the black maximum black mass for the deBoer et al.(2017) rate hole (Di Carlo et al. 2020a; Roupas & Kazanas 2019; is computed using the “adopted”, “lower”, and “upper” van Son et al. 2020). rates from Table XXV of deBoer et al.(2017). The The detection of the upper edge of the PISN mass gap lower edge of the black hole mass gap over the different (MBH > 120 M ) would provide a strong constraint on sources is between 47 51 M , with an uncertainty on 16 − the 12C(α, γ) O rate. This edge has smaller numerical the maximum black hole mass of < 5%. The upper edge uncertainties associated with it, as it is defined only by varies between 130 136 M , with a similar uncertainty − a combination of fundamental physics (nuclear reaction on the maximum black hole mass of < 5%. rates and the equation of state of an ionized gas) and The small variations seen in the edges of the PISN does not depend on the complexities of modeling the mass gap, are due to the fact that the different sources of hydrodynamical pulses which define the lower edge of the 12C(α, γ)16O have been slowly converging over time the PISN mass gap. Mergers in this mass range are on a S-factor between 140 160 keV barns (See figure − expected to be rare due to the difficulty in producing 7). See Figure 26 of deBoer et al.(2017) for a review sufficiently massive stars in close binaries (Belczynski of how the uncertainty in the different energy levels has et al. 2016b), however they may be detectable by third improved since the 1970’s. generation gravitational wave detectors (Mangiagli et al. 2019). 4.2. Sensitivity to other reaction rates Table2 shows how the maximum black hole mass 12 16 4.1. Other sources of the C(α, γ) O rate varies as a function of both the 12C(α, γ)16O rate and Table1 shows the maximum black hole mass as a func- other reaction rates that either create carbon (the 3α tion of different sources for the 12C(α, γ)16O reaction. reaction), destroy carbon (12C+ 12C), and oxygen burn- The Sallaska et al.(2013) STARLIB rate is based on that ing (16O + 16O). For each rate varied we compute the of Kunz et al.(2002), however STARLIB assumes rate location of the mass gap for the number of standard probability density is lognormal thus its median value deviations from the median for that rate, and for vari- 12 16 is RLow RHigh (eqaution 17 Sallaska et al.(2013)), ations in the C(α, γ) O reaction. This is to probe × p 9

Table 1. Location of the edges of the PISN mass gap for different much larger change when reducing the rate than when 12 16 sources of the C(α, γ) O reaction rate. Uncertainties quoted the rate is increased. are 1σ where applicable. To test variations in the 12C + 12C we use the rate provided by Tumino et al.(2018), which provides a tem- Source Lower [ M ] Upper [ M ] perature dependent 1σ uncertainty on the reaction rate. 1.7× Caughlan & Fowler(1988) 49 135 However, the uncertainty is only available up to 3 GK, Angulo et al.(1999)( NACRE) 49 130 at higher temperatures we revert to MESA’s standard a Kunz et al.(2002) 50 134 12C + 12C reaction rate, which does not have a provided Cyburt et al.(2010)( REACLIB) 50 136 uncertainty estimate (Caughlan & Fowler 1988). b +7 +13 Sallaska et al.(2013) ( STARLIB) 47−2 130−7 It is difficult to determine the trend in black hole mass c +0 +5 12 12 deBoer et al.(2017) 51−4 134−5 compared to the C + C given a number of models due not converge. We might expect variations around aBased on the “adopted” fitting coefficents in Table 5 of Kunz 15%. The larger change occurs when the σC12 0 et al.(2002) ≈ ≥ than when σC12 = 3. This is due to the change in the b p − The STARLIB median rate is based on RLow × RHigh (eqaution power generated during the carbon burning. When the 17 Sallaska et al.(2013)), where the rates R Low and RHigh come 12C + 12C rate is increased then stars with low 12C frac- from Table 5 of Kunz et al.(2002) tions (σC12 0), which would not generate a convective c Based on the “adopted”, “lower”, and “upper” rates from Table carbon shell≥ (when 12C+ 12C is small) can now generate XXV of deBoer et al.(2017) sufficient power to alter the core structure and poten- tially drive the formation of convective carbon burning shell. See however Tan et al.(2020) for a discussion on why the Tumino et al.(2018) rate may have been over- for correlations between the rates. In the case of the estimated, and Fruet et al.(2020) for a discussion on rates from Caughlan & Fowler(1988) we follow the un- new measurement techniques of the 12C + 12C rate. certainty provided by STARLIB and we multiply (divide) As the 16O + 16O rate increases the maximum black the rate by a fixed factor of 10 due to the lack of knowl- hole mass decreases. This change is asymmetric, with a edge of the uncertainty in this rate. Table2 then shows larger change occurring when the rate decreases than the fractional change in the location of each edge of the when the rate increases. The 0% change seen when mass gap, as an indication of how sensitive the edges of σC12 = 3, is due to those stars lacking pulsations. As − the mass gap are to other uncertainties. In general the the star has a high 12C fraction, and thus a carbon shell, maximum fractional error is 5 15% from considering it does not undergo explosive oxygen burning only sta- other reactions, in the location≈ of− the PISN mass gap. ble core oxygen burning. As there are no pulsations no We would expect that varying a rate between its up- mass is lost. This rate does have an effect on the core per and lower limits would produce relative changes with structure of the star, which might lead to variations in opposing signs, however in some cases this does not oc- the mass lost during the final collapse into a black hole. cur. This is due to numerical difficulties in the evolution For larger values of σC12 there is up to a 10% variation of the models. Some of these models would fail to reach in the location of the edges of the PISN mass gap. core collapse, and thus we could not measure the peak of More work is needed to understand the correlations the black hole mass distribution. Instead, we report the between the different reaction rates and their effect on largest black hole mass only for the models that success- our ability to constrain the edges of the PISN mass gap, fully reach core collapse. This means that the relative e.g. West et al.(2013). We need to improve our under- change we measure includes changes in the initial mass standing of how the uncertainty in the rates at differ- as well as the black hole mass. Table2 should therefore ent temperatures alters the behavior of the carbon shell be taken as a representation of the changes expected and the final black hole mass. This could be achieved for different rates, but it does not show the complete with a Monto-Carlo sampling of the reactions rates, e.g. picture. Rauscher et al.(2016); Fields et al.(2016, 2018), how- For the STARLIB median 12C(α, γ)16O reaction rate, ever this comes at a much greater computational cost. the 3α reaction produces a fractional uncertainty of 16 1 10%, independent of the 12C(α, γ) O rate. By in- 5. CONSTRAINING THE 12C(α, γ)16O REACTION − creasing the 3α, within its one sigma uncertainties, rate RATE WITH GRAVITATIONAL WAVES we decrease the maximum black hole mass. There is a Because of the sensitivity of the edges of the PISN mass gap to the 12C(α, γ)16O reaction rate, we can use 10

Table 2. The relative change in the location of the upper and lower edge of the PISN mass gap, when varying both the 12C(α, γ)16O reaction rate and either the 3α, 12C+12C, or 16O+16O reaction rate, with respect to our default choices.

Rate Uncertainty 12C(α, γ)16O

σC12 = −3 σC12 = 0 σC12 = +3 Lower Upper Lower Upper Lower Upper 3α +1σ −1.0% −2.7% 1.7%a −0.8% −4.3%a −0.8% Sallaska et al.(2013) −1σ 6.2% 4.2% 8.9%a 4.6% −7.6%a 0.8% 12C + 12C +1σ −1.0% 0.5%a 16.6% ∼ 0.0% −16.8%a ∼ 0.0% Tumino et al.(2018) −1σ 1.9% 0.5%a −1.4% ∼ 0.0% −1.2%a ∼ 0.0% 16O + 16O ×10 0.0%b −1.8% 0.0% −2.3% −2.6% −3.3% Caughlan & Fowler(1988) ×0.1 0.0%b 4.4% 5.7% 9.1% 6.9% 9.2% aVariations have the same sign as numerical difficulties prevent comparison between similar models. b No variation seen as most burning occurs at temperatures where the rate has reverted to Caughlan & Fowler (1988) and thus shows no variation.

Core Shell Explosive tures. See AppendixB for the sensitivity of our results to different temperature ranges. We focus here on the 3.0 lower edge of the mass gap, as it has been inferred from

3.0 the existing LIGO/Virgo data (Abbott et al. 2019b). 2.5 The currently most massive black hole in O1/O2, as inferred by LIGO/Virgo is GW170729 at MBH = 2.0 2.0 +16.6 50.6−10.2 M (Abbott et al. 2019b), which could be used as an estimate for the location of the PISN mass gap, 1.5 Lab measurements 1.0 assuming that it is from a first generation black hole. Normalised Rate 1.0 0.0 There are also several other candidates for the most mas-

-1.0 sive black hole. This includes IMBHC-170502 has which 0.5 -2.0 has been inferred to have individual black hole compo- -3.0 nents with masses 94 M and 62 M (Udall et al. 0.25 0.00 0.1 1.0 10.0 ≈ ≈ p(rate) T [GK] 2019). GW151205 has also been proposed to have one +28 component with an inferred mass of MBH = 68 M 12 16 −17 Figure 6. The right pannel shows the C(α, γ) O rate as (Nitz et al. 2020). By having a component mass inside a function of temperature, normalized to the median STARLIB µ(T )σC12 the classical PISN mass gap it was suggested that this rate hσνi/hσνimedian = e . The left panel shows the posterior of our distribution. The solid line, labeled 0.0, in- was the result of dynamical mergers. dicates the normalized median rate (i.e., σC12=0.0). The However, we must be careful in not over-interpreting dashed lines show the reaction rates above and below the single events, which may be susceptible to noise fluctu- median STARLIB rate, labeled by the appropriate value of ations (Fishbach et al. 2020) which can make a black σC12. We assume that the black hole mass distribution is hole have a higher apparent mass than it truly does. given by model B of Abbott et al.(2019b). The dark blue For instance, considering GW170729 jointly with the 12 16 region shows the 50% confidence range in the C(α, γ) O other O1/O2 detections, lowers its mass to M = rate, while the lighter blue shows the 90% confidence interval. BH 38.9+7.3 M , which places it below the PISN mass gap. Note the upper rate limit is unbounded when adopting the −4.5 current O2 LIGO/Virgo posteriors on the maximum black Thus we must consider the entire population of bi- hole mass. The red region (T > 10 GK) shows the approx- nary black hole mergers as a whole, when measuring the imate lower edge of the energy range for lab measurements maximum inferred black hole mass below the gap. The of the 12C(α, γ)16O rate (Holt et al. 2019). 10 detections in O1/O2 places the maximum black hole mass below the PISN mass gap at 42 44 M depend- − the measured location of the gap to derive a value for ing on the choice of model parameters (Abbott et al. the 12C(α, γ)16O at astrophysically relevant tempera- 2019b). The current 90% confidence interval on this 11 value is 10 M . With a large enough population cussion on how this limit may be improved with future ≈ ± ( (50)) of black holes we can place limits of 1 M gravitational wave detections. Figure7 also shows the onO the location of the gap (Fishbach & Holz 2017≈ ±; Ab- lower limit on the S-factor if PPISN can be shown to bott et al. 2019b). exist though other means, for instance electromagnetic We assume that all binary black hole mergers so far observations of the SN. The existence of PPISN would detected come from isolated binaries or first generation imply that S300 > 79 keV barns, this is discussed further black hole mergers, thus the maximum mass black holes in Section 7.2. below the gap come from PPISN. We also assume that only uncertainties in 12C(α, γ)16O matter. Thus we can NACRE NACRE2 use the posterior distribution over the maximum black Hammer 05 Nuclear hole mass for the population of black holes as the esti- STARLIB mate of the maximum black hole mass below the mass Deboer 17 Shen 20 gap. Woosley 93 Astro Figure6 shows the uncertainty in the STARLIB Metcalfe 03 12 16 Mmax,BH 40M O1/O2 C(α, γ) O reaction rate as a function of tempera- ∼ Mmax,BH = 45M ture. Over the temperatures we are sensitive to, less Mmax,BH = 50M O3 Mmax,BH = 60M than 1.0 GK where helium burns non-explosively, the Mmax,BH = 70M uncertainty is approximately constant. Thus we need PPISN EM Transients only find a single temperature-independent σC12 to fit 0 50 100 150 200 250 300 350 to the maximum black hole mass below the gap. S300 [keV barns] th We fit a 4 order polynomial to the lower edge of Figure 7. Constraints on the S-factor at 300keV for the the PISN mass gap (Fig.5) to map from maximum 12C(α, γ)16O reaction rate. The blue region presents val- black hole mass to σC12. This is then combined with the ues for commonly-used nuclear reaction rate libraries, while posterior of the maximum black hole mass to generate a the red region shows values inferred from astrophysical mea- surements, by observations of white dwarfs (Metcalfe 2003) posterior distribution over σC12. The blue boxes in Fig. 6 show our 50% and 90% confidence interval on σ . and from the galactic chemical enrichment model of Woosley C12 et al.(1993). The green region show our posteriors on the At 50% confidence, we can limit σ to be between 0.5 16 C12 12C(α, γ) O derived from the O1/O2 maximum black hole and 2.5σC12, while at 90% confidence we can only place a mass inferred from model B of Abbott et al.(2019b), and for lower limit of σC12 > 1.5. This is due to the posterior 50 simulated binary black-hole observations, approximately − distribution from LIGO/Virgo allowing the maximum the number expected in O3, assuming that the O3 detections black hole mass to be MBH,max < 40 M , which is below follow the power law model B from Abbott et al.(2019b). the lower limit we find for the edge of the mass gap. Error bars show the 68% confidence intervals on our inferred By taking the STARLIB astrophysical S-factor, at S-factor at 300keV. The yellow region shows the constraints placed if PPISN supernovae exist (Section 7.2). Error bars 300keV, to be 165 keV barns (Sallaska et al. 2013) we in the blue and red regions are 1σ uncertainties. We assume can scale the S-factor from the σC12. As the nor- that the uncertainty from Shen et al.(2020) is the same as malized rate is approximately flat (Fig.6) the uncer- that of deBoer et al.(2017). tainty is flat. Using Equation4 and Fig.6 we have µ(T )σC12 e = σν / σν median S(300 keV) , thus the S- h i h i ∝ factor can be linearly scaled from its STARLIB value to 5.1. GW190521 a new value, given by a different σ . C12 The detection of GW190521 (Abbott et al. 2020a) Figure7 shows a comparison between the values with component masses of 85+21 M and 66+17 M for the S-factor from nuclear laboratory experiments, −14 −18 places both black holes firmly in the mass gap, as constraints placed by asterosiesmology inferred from the O1/O2 observations (Abbott et al. and galactic chemical enrichment models, and this 2020b). However, if we assume that the primary black study. For our assumptions we find the S-factor 16 hole in GW190521 was a first generation black hole, for the 12C(α, γ) O rate at 300keV to be S > 300 this leads to an inference of σ = 2.4+0.6, and 175 keV barns, at 68% confidence. At 95% confidence C12 thus S(300 keV) = 73+11 keV barns. We− can not in- we find a limit of S > 82 keV barns and at 99% con- 300 fer a lower limit as the 90% confidence interval for the fidence we find a limit of S > 68 keV barns. The 300 mass extends above M 95 M where our models S-factors computed here are consistent with experimen- BH can not place a black hole,≈ below the mass gap. If tally derived values, though we only currently place a σ 2.5, then PPISN would be suppressed due lower limit on the S-factor. See section6 for a dis- C12 to the≈ carbon − burning shell. Instead, if we assume 12 that only the secondary object was a first generation 45, 50, 60, 70 M , and we choose α = 1.5, 1.5, 2, 3, +1.8 − − − − black hole we would infer a σC12 = 1.7−0.5, and thus to remain consistent with the joint α, Mmax posterior +84 − S(300 keV) = 87−12 keV barns. from O1/O2 (Abbott et al. 2019b). However, it is unlikely that GW190521 is a pair of We generate 50 mock detections from each of the four first generation black holes, instead it is likely to be a mass distributions described above. We assume that pair of second generation black holes, perhaps inside an the underlying merger rate density is constant in red- AGN disk (Abbott et al. 2020b). Indeed, (Graham et al. shift, and that sources are detected if they pass an SNR 2020) claims a tentative detection of EM counterpart threshold of 8 in a single detector. We neglect the (ZTF19abanrhr) due to the merger product ramming spins of the black holes in this study, as the maximum into an AGN disk. However, the redshift of the AGN mass is well-measured independently of the underlying does not agree with the GW-inferred redshift. Further spin distribution. For each mock detection, we simu- investigation of this event and the likelihood of a double late the measurement uncertainty on the source-frame second generation merger in an AGN disk is warranted. masses according to the prescription in Fishbach et al. The presence of black holes in the expected mass gap (2020), which is calibrated to the simulation study of Vi- region can make the analysis of the location of the gap tale et al.(2017). This gives a typical 1 σ measurement more difficult, though folding in a prior based on the uncertainty of 30% on the source-frame masses. We 12C(α, γ)16O may help to identify outliers that have then perform a∼ hierarchical Bayesian analysis (Mandel formed through alternative formation mechanisms. 2010; Mandel et al. 2019) on each set of 50 detections to recover the posterior over the population parameters, 6. PROSPECTS FOR CONSTRAINING α, M , β and M . This provides a projection of 12C(α, γ)16O FROM FUTURE GW DETECTIONS max min how well Mmax, the maximum mass below the PISN With the release of the O3 data from LIGO/Virgo, mass gap, can be measured with 50 observations. We it is predicted that there will be (50) BBH detections sample from the hierarchical Bayesian likelihood using O (Abbott et al. 2018). Thus we can ask how well can we PyMC3 (Salvatier et al. 2016). Finally, we translate the expect to do with additional observations? This depends projected measurement of Mmax under each of the simu- strongly on what value the lower edge of the PISN mass lated populations to a measurement of the 12C(α, γ)16O gap is found to be. reaction rate according to Fig.5. We assume that the detections in O3 will follow a Figure7 shows our inferred S-factors (at 300keV) for power law in primary mass (M1) of the black hole in the different choices of the maximum black hole mass below binary, with the form previously assumed for the O1/O2 the PISN mass gap. As the maximum black hole mass data (Abbott et al. 2019b). Thus we have: increases, the S-factor decreases (as stars have more 12C in their core and thus have reduced mass loss from pul- α p(M1) M (7) sations). The 68% confidence interval also reduces in ∝ 1 size as the maximum black hole mass increases. The for Mmin < M1 < Mmax, where Mmin is the minimum possible black hole mass below the PISN mass gap, and predicted accuracy with which LIGO/Virgo is expected to infer the maximum black hole mass decreases as the Mmax is the maximum possible black hole mass. We also mass increases, as we require a steeper power law in- assume that the secondary mass M2 follows: dex α to be consistent with the O1/O2 observations. β This leads to fewer mergers near the gap. However, the p(M2 M1) M2 (8) | ∝ maximum black hole mass becomes more sensitive to where M < M < M . This is equivalent to model min 2 1 the choice of σ . This can be seen in the gradient of B of Abbott et al.(2019b). We set M = 5 M and C12 min the lower edge of the PISN mass gap in Fig.5, which β = 4 to be consistent with the current O1/O2 obser- increases as σ decreases. We caution that we have vations (Abbott et al. 2019b). There is a strong cor- C12 likely under-estimated the size of the uncertainty range, relation between the values of α and M . Thus for max especially at the higher black hole masses, due to the different choices of M we choose a value of α that max effect of uncertainties in other reaction rates and mass remains consistent with the O1/O2 data (e.g, Fig. 3 of lost during the formation of the black hole. With the Abbott et al.(2019b)). As M increases, we require a max predicted accuracy expected for LIGO/Virgo during O3 larger α value. We consider 4 possible locations for the in inferring the maximum black hole mass, we will be lower edge of the PISN mass gap (M ) to explore how max limited by the accuracy of our models, not the data, in the uncertainty both from the LIGO/Virgo measure- 16 constraining the 12C(α, γ) O reaction rate. ments and from our stellar models affects our determi- 12 16 nation of the C(α, γ) O rate. We considered Mmax = 13

7. OTHER OBSERVABLES interface between the stars undergoing CC and the light- 7.1. Formation rates est PPISN progenitors, depending on the strength of the mass loss in the lighest PPISN progenitors (Renzo et al. Given that the initial mass function (IMF) strongly 2020b). We chose not to show Figure8 in terms of the favors less massive progenitors, this would imply that carbon core mass, which is the more applicable quantity PPISN and PISN would be more common at higher to show when comparing between different metalicities 12 16 values of σC12 (higher C(α, γ) O rates), all else be- (Farmer et al. 2019), as the highest σC12 models have ing equal. This is potentially detectable given a suffi- X 12C 0 in their cores. This leads to the carbon ciently large population of binary black hole mergers or core mass≈ being ill-defined. PPISN/PISN transients, and could provide additional  12 16 constraints on the C(α, γ) O rate. A number of 100 σ = 3.0 σ = 0.5 − upcoming surveys, including LSST, JWST, and WFIRST, σ = 2.5 σ = 1.0 − σ = 2.0 σ = 1.5 − are expected to find significant numbers of PPISN and σ = 1.5 σ = 2.0 − σ = 1.0 σ = 2.5 PISN transients (Young et al. 2008; Hummel et al. 2012; 80 − ) σ = 0.5 σ = 3.0 ] − Whalen et al. 2013a,b; Villar et al. 2018; Reg˝oset al. σ = 0.0 [M

2020). BH 60 M To provide a rough estimate for this, we make the ( simplified assumption that the helium core masses fol- low a Salpeter-like IMF with a power law α = 2.35, 40 so that we can compare the relative difference− in for- mation rates for stars with σC12 = 1. We take the Black hole mass ± 20 smallest helium cores to be MHe,init = 30 M (the least massive stars modeled in our grid), and the maximum 0 helium core mass that makes a black hole as: 73 M for 40 60 80 100 σC12 = 1, and 61 M for σC12 = +1. This folds to- MHe,Initial [M ] − gether the formation rates of black holes from both core GW150914 GW151012 GW151226 GW170104 GW170608 GW170729 GW170809 GW170814 GW170817 GW170818 GW170823 collapse and PPISN, as gravitional waves can not dis- Figure 8. The final black hole mass as a function of the ini- tinguish these objects from their gravitional wave signal tial helium core mass. The colors represent different choices alone. This leads to a relative increase in the forma- for σC12. Circles denote models which undergo pulsational tion rates of black holes of 10% for the σC12 = 1 mass loss (PPISN) while plus symbols denote models which ≈ − over σC12 = +1 models. There is a larger variation pos- do not undergo mass lost due to pulsations (CC). The right sible, if we consider how the lowest-mass helium core pannel shows the inferred masses of the currently known that makes a black hole varies with 12C(α, γ)16O. For black holes from LIGO/Virgo O2 (Abbott et al. 2019a), with their 90% confidence intervals. σC12 = 1 this puts the lower limit at 5 M and 8 M (Sukhbold± & Adams 2020), which would lead to a factor 2 difference in the formation rates of black holes. This We can compare the formation rates of PPISN (and is mostly due to the change in the relative number of thus potentially detectable supernovae) only for σC12 > low mass black holes. 2.0. As an indication of the variation, we consider − Figure8 shows the black hole masses as a function of the relative change (due to the IMF) in formation rates the initial helium core mass, for different choices of σ . only for σC12 = 1. We find the helium core masses for C12 ± At the lowest values of σC12, no star undergoes PPISN, stars that we would classify as undergoing a PPISN, as thus the black hole mass scales lineraly with the initial 41 61 M for σC12 = +1, and 55 73 M for σC12 = 1. − − − helium core mass. As σC12 increases the final black hole This leads to a increase by a factor of 2 for σC12 = +1 mass shifts to lower masses away from a linear relation- over σC12 = 1. However, there can be difficulty in − ship with the helium core mass, due to pulsational mass determining which stars are PPISN due to the small loss. There is also a turn over in the initial-black hole amounts of mass lost for the lightest PPISN progenitors mass realtionship, where the most massive black holes (Renzo et al. 2020a). do not form from the most massive stars undergoing For the rate of PISN, we find the initial helium core PPISN (Farmer et al. 2019). This turn over may be de- masses to be 61 124 M for +1σC12, and 73 142 M − − tectable in the infered black hole mass distribution (see for σC12 = 1, These ranges then lead to a relative − model C of Abbott et al. 2019b). There may also be a increase of 30% for the rate of PISN for σC12 = +1 ≈ small bump in the black hole mass distribution at the compared to σC12 = 1. These variations are driven by the change in the lower− masses needed for a PPISN 14 or PISN as the 12C(α, γ)16O rate increases (Takahashi are formation mechanisms which may place a black hole 2018). in the mass gap. If a star can retain its hydrogen enve- lope until collapse, though a combination of weak stel- 7.2. Observations of supernovae lar winds (Woosley 2017) or for a stellar merger (Vigna- G´omezet al. 2019; Spera et al. 2019), then we could find In Fig.1 we show that when σC12 < 2 no PPISN − black holes up to 60 M , see van Son et al.(2020) for are formed. This provides an intriguing observational ∼ test for the 12C(α, γ)16O reaction rate. If the exis- an overview. Black holes formed in dense stellar clus- tence of PPISN can be confirmed though photometric ters, or AGN disks (McKernan et al. 2018), can merge observations (for instance candidate PPISN SN2006gy multiple times (Rodriguez et al. 2016; Stone et al. 2017; (Woosley et al. 2007), SN2006jc (Pastorello et al. 2007), Di Carlo et al. 2019; Yang et al. 2019). SN iPTF14hls (Arcavi et al. 2017; Woosley 2018; Vigna- The first generation black holes in these environments G´omezet al. 2019), SN iPTF16eh (Lunnan et al. 2018), will be limited by the PISN mass gap, however higher or SN2016iet (Gomez et al. 2019), or SN2016aps (Nicholl generation mergers would not be limited and could pop- et al. 2020)) then we can place a lower limit on the ulate the gap (Di Carlo et al. 2020b). Their effect on 12 16 12 16 the inference of C(α, γ) O will depend on whether C(α, γ) O reaction rate of σC12 > 2.0. The outer layers of the star that− are expelled in a the kick the resulting black hole receives is small enough pulsational mass loss event, or even the small amount that the black hole stays in the cluster (Rodriguez et al. of material that might be ejected in the final collapse 2018; Fragione & Kocsis 2018), and thus on their uncer- to a black hole, provides some information on the final tain contribution to the total rate of black hole mergers. composition of the star, though this will depend on the If they are distinguishable from mergers due to isolated star’s metallicity and assumed wind mass loss rate. In binary evolution, for instance via their spin (Fishbach general the lowest mass stars that undergo a PPISN, in et al. 2017; Gerosa & Berti 2017; Bouffanais et al. 2019; our grid, have surface layers dominated by 4He, while Arca Sedda et al. 2020), then they could be removed 12 16 the higher mass stars have 16O rich outer layers (Renzo from the population used to infer the C(α, γ) O rate. et al. 2020a). It may also be possible to fit the maximum black hole As the initial mass increases, and the pulses get mass below the gap assuming the population contains stronger and thus remove more mass they can expose both isolated binaries and hierarchical mergers without 16O rich layers, with traces of 20Ne, 23Na, 24Mg, and 28Si needing to subtract out the hierarchical mergers (Kim- (Renzo et al. 2020a). However, the 12C fraction in the ball et al. 2020) Also, if a channel that produces black outer layers follows the same trend seen in the core 12C holes in the mass gap is rare, we may still be able to fraction. As the 12C(α, γ)16O reaction rate increases determine the location of the mass gap for the more more 12C is converted into 16O in helium shell burning. dominant channel. Therefore, the measured abundances could provide ad- In this work we have considered four reaction rates, 12 16 12 12 16 16 ditional constraints on 12C(α, γ)16O, and may provide 3α, C(α, γ) O, C + C, and O + O. There constraints in a reduced temperature region (that asso- are many other reaction rates that can alter the evolu- ciated with shell helium burning, 0.3 < T/ GK < 1.0). tion of a star (Rauscher et al. 2016; Fields et al. 2018). Further work is needed to quantify the amount of mix- We expect their effect to be small for the final black ing in the outer layers of the star (Renzo et al. 2020b), hole mass, compared with the reaction rates considered understanding which parts of the ejecta (and thus which here, though they would play a role in the nuclosynethic layers of the star) would be measured in spectroscopy of yields from the PPISN and PISN ejecta. In Farmer a PPISN (Renzo et al. 2020a), as well as investigating et al.(2019) we investigated other uncertainties, e.g the effect of a larger nuclear network to follow in greater mass loss, metallicity, convective mixing, and neutrino physics, have only a small effect of 10%, on the lo- detail the nucleosynthetic yields from the explosive oxy- ≈ gen burning (Weaver & Woosley 1993; Heger & Woosley cation of the mass gap. How convection is treated in 2002; Woosley & Heger 2007; West et al. 2013). this hydrodynamical regime can have a small impact on the final black hole mass for stars at the boundary be- 8. DISCUSSION tween core collapse and PPISN i.e, where the pulses are weak. However the edge of the PISN mass gap is not af- This work assumes that all black holes found by fected, due to the stronger pulses experienced by a star LIGO/Virgo so far have come from stars that lose their near the mass gap (Renzo et al. 2020b). In this work hydrogen envelope before their collapse to a black hole. we considered only non-rotating models. Rotation may Thus the maximum mass a black hole can have is lim- play a role in the final black hole mass, depending on ited by PISN and mass loss from PPISN. However there 15 how material with high angular momentum is accreted It is crucial in determining the final properties and fate onto the black hole during the collapse (Fryer & Warren of a star, and as we explicitly show in this work the 2004; Rockefeller et al. 2006; Batta et al. 2017). predictions for the location of the pair-instability mass By assuming the entire bound mass of the star col- gap. lapses into a black hole, we place an upper limit on the We show that the physical reason why this reaction black hole mass possible for a PPISN. If the star was to rate is so important is that it determines the relative lose mass during the collapse, then we would over esti- fraction of carbon and oxygen in the core at the end of mate the inferred S-factor for the 12C(α, γ)16O reaction helium burning. In models for which we adopted a lower rate. If the star forms a proto-neutron star during its reaction rate, enough carbon is left to ignite such that collapse it might eject 10% of the proto-neutron star it effectively delays the ignition of oxygen. As carbon- ≈ mass as neutrinos, 0.1 M (Fryer 1999; Fryer et al. carbon burning occurs in a shell, the core inside the 2012). The envelope≈ of the star may then respond to shell contracts to higher densities. This increases the the change in the gravitational potential, generating a effects of electron degeneracy and gas pressure, which weak shock that unbinds material with binding energies stabilizes the core. The formation of electron-positron less than 1047ergs (Nadezhin 1980; Lovegrove & Woosley pairs is then suppressed due to the increased occupation 2013; Fern´andezet al. 2018). However, stars undergoing fraction of the low energy states for electrons. Oxygen PPISN will have already expelled their weakly bound can then ignite stably even for higher core masses. outer envelopes, thus mass lost via a weak shock is lim- In contrast, in models for which we assume higher re- ited to a few tenths of a solar mass. If a jet is pro- action rates, almost all carbon is depleted at the end duced, by accretion onto the compact object, then there of helium burning. The star then skips carbon-carbon maybe an ejection > 1 M of material (Gilkis et al. burning and oxygen ignites explosively. The net effect 12 16 2016; Quataert et al. 2019). Assuming 1 M of mass is that increasing the C(α, γ) O reaction rate pushes loss during the collapse, then our inferred S-factor at the mass regime for pair pulsations and pair-instability 68% confidence decreases from S300 > 175 keV barns to supernovae to lower masses. This allows for lower mass S300 > 159 keV barns. Further work is needed to un- black holes and thus shifts the location of the pair in- derstand the collapse mechanism of these massive cores, stability mass gap to lower masses. and whether we can extrapolate from models of stars Our results can be summarized as follows: that core collapse with 5 10 M cores to those with − 1. The location of the gap is sensitive to the 50 M cores. ≈Previous studies of PPISN and PISN progenitors have reaction rate for alpha captures on to car- found the location of the PISN mass gap consistent bon, but the width of the mass gap is not. with ours, 50 53 M (Yoshida et al. 2016), 48 M The lower edge of the mass gap varies between +49 +44 − ≈ 47 M and 130 M for the upper edge, for (Woosley 2017), 50 M (Leung et al. 2019). The small −4 −12 12 16 variations in the≈ location of the gap can be attributed 3 sigma variations in the C(α, γ) O reaction ± +5 to differences in chosen metallicity, mass loss rates, and rate. The width is 83−8 M (Figure5). 12 16 source of the C(α, γ) O reaction rate. Takahashi 2. We can place a lower limit on the (2018) showed how increasing the reaction rate decreases 12C(α, γ)16O reaction rate using the first ten the initial mass needed for PISN, in agreement with our gravitational-wave detections of black holes. findings and that the range of initial masses that can Considering only variations in this reaction, we form a PPISN is reduced as the reaction rate decreases. constrain the astrophysical S-factor, which is a measure of the strength of the reaction rate, to 9. SUMMARY S300 > 175 keV barns at 68% confidence (Fig- With the rapid increase in the number of ure7). gravitational-wave detections, the hope is that they can 3. With (50) detections, as expected after be used to start drawing lessons about the uncertain the completionO of the third observing run, physics of their massive-star progenitors. In an earlier we expect to place constraints of 10– paper (Farmer et al. 2019) we speculated that measure- 16 30 keV barns on the 12C(α, γ) O S-factor.± We ments of the edge of the predicted black-hole mass gap show how the constraints depend on the actual lo- due to pair-instability could be used to constraint the cation of the gap (Figure7). nuclear reaction rate of carbon capturing alpha particles producing oxygen (12C(α, γ)16O). This reaction rate is 4. We find other stellar model uncertainties to very uncertain but has large astrophysical significance. be subdominant, although this needs to be 16

explored further. Variations in other nuclear re- redshift, and potentially lead to detections above actions such as helium burning (3α), carbon burn- the mass gap. Improved progenitor models will be ing (12C + 12C), and oxygen burning (16O + 16O) needed to maximize the science return as the ob- contribute uncertainties of the order of 10% to the servational constraints improve, but the future is edge of the mass gap (Table2). See Farmer et al. promising. (2019) and Renzo et al.(2020b) for a discussion of ACKNOWLEDGMENTS the effect of physical uncertainties and numerics. We acknowledge helpful discussions with B. Paxton, F. Timmes, P. Marchant, E. C. Laplace, and L. van 5. The unambiguous detection of pulsational Son. RF is supported by the Netherlands Organiza- pair-instability supernovae in electromag- tion for Scientific Research (NWO) through a top mod- netic transient surveys would place an in- ule 2 grant with project number 614.001.501 (PI de dependent constraint on the 12C(α, γ)16O re- Mink). SdM acknowledges funding by the European action rate. For the lowest adopted reaction Union’s Horizon 2020 research and innovation program rates (< 2σ) we no longer see pulsations due from the European Research Council (ERC) (Grant to pair instability.− The detection of pulsational agreement No. 715063), and by the Netherlands Orga- pair-instability would thus imply a lower limit of nization for Scientific Research (NWO) as part of the S > 79 keV barns for the 12C(α, γ)16O reaction 300 Vidi research program BinWaves with project number rate (Figure1, Section 7.2). This will be of interest 639.042.728. SdM also acknowledges the Black Hole Ini- for automated wide-field transient searches such as tiative at Harvard University, which is funded by grants the Legacy Survey of Space and Time (LSST). from the John Templeton Foundation and the Gordon and Betty Moore Foundation to Harvard University. 6. Constraining nuclear stellar astrophysics is This work was also supported by the Cost Action Pro- an interesting science case for third genera- gram ChETEC CA16117. This work was carried out tion gravitational wave detectors. Future de- on the Dutch national e-infrastructure with the support tectors such as the Einstein Telescope and Cosmic of SURF Cooperative. This research has made use of Explorer will be able to probe detailed features in NASA’s Astrophysics Data System. the black-hole mass distribution as a function of

APPENDIX

A. MESA screening corrections are provided by Chugunov et al. When solving the stellar structure equations MESA (2007). Weak reaction rates are based on the following uses a set of input microphysics. This includes ther- tabulations; Langanke & Mart´ınez-Pinedo(2000), Oda mal neutrino energy losses from the fitting formula of et al.(1994), and Fuller et al.(1985). Itoh et al.(1996). The equation of state (EOS) which Reverse nuclear reaction rates are computed from de- is a blend of the OPAL (Rogers & Nayfonov 2002), tailed balance based on the NACRE/REACLIB reaction SCVH (Saumon et al. 1995), PTEH (Pols et al. 1995), rate, instead of consistently from the STARLIB rate. This HELM (Timmes & Swesty 2000), and PC (Potekhin & is due to limitations in MESA. However, for the rates 12 16 12 12 Chabrier 2010) EOSes. Radiative opacities are primar- we are interested in 3α, C(α, γ) O, C + C, and 16 16 ily drawn from OPAL (Iglesias & Rogers 1993, 1996), O + O their reverse reactions have negligible im- with low-temperature data from Ferguson et al.(2005) pact on the star’s evolution. The nuclear partition func- and the high-temperature, Compton-scattering domi- tions used to calculate the inverse rates are taken from nated regime from Buchler & Yueh(1976). Electron Rauscher & Thielemann(2000). conduction opacities are taken from Cassisi et al.(2007). We treat wind mass loss rates as in Marchant et al. MESA’s default nuclear reaction rates come from a com- (2019), where we assume the mass loss rate of Hamann bination of NACRE (Angulo et al. 1999) and REACLIB (Cy- et al.(1982, 1995); Hamann & Koesterke(1998) with burt et al. 2010)(default snapshot1). The MESA nuclear a wind efficiency of 0.1 to account for wind clumpiness. For further discussion of the effect of wind mass loss, see Farmer et al.(2019). 1 Dated 2017-10-20. Available from http://reaclib.jinaweb.org/library.php?action=viewsnapshots 17

B. CALIBRATION OF 12C(α, γ)16O REACTION range, with a smaller dependence on the shell helium RATE burning region, and we are not sensitive to changes in 12 16 To test which temperature range we are most sen- the C(α, γ) O reaction rate in the explosive helium sitive to, we ran models where we used the STARLIB burning temperature range. median 12C(α, γ)16O rate, but changed the rate in one The changes in the maximum black hole mass occur of three temperature regions to that of the 3σ value due to the changes in the carbon fraction and where (for that temperature region only). These temperatures− those changes occur. Changes during core helium burn- were chosen based on the type of helium burning encoun- ing primarily effect the core carbon fraction (see Sec. tered at that temperature: core helium burning 0.1 < 3). The maximum black hole mass does not mass de- 12 16 T/ GK < 0.3, shell helium burning 0.3 < T/ GK < 1.0, pend on the C(α, γ) O in the explosive helium burn- and explosive helium burning T > 1 GK. For the de- ing regime, as no helium-rich region reaches T > 1 GK. fault case of the median σC12, the maximum black hole Software: mesaPlot (Farmer 2018), mesaSDK mass was 46 M , and for σC12 = 3 (over the whole (Townsend 2019), ipython/jupyter (P´erez& Granger − temperature range) it was 95 M . When only changing 2007; Kluyver et al. 2016), matplotlib (Hunter 2007), the rate during the core helium burning, the maximum NumPy (van der Walt et al. 2011), PyMC3 (Salvatier et al. black hole mass became 79 M , for helium shell burn- 2016), MESA(Paxton et al. 2011, 2013, 2015, 2018, 2019), ing it was 55 M , and explosive helium burning it was and pyMesa (Farmer & Bauer 2018). 46 M . Thus we are most sensitive to changes in the 12C(α, γ)16O in the core helium burning temperature

REFERENCES

Abbott, B. P., Abbott, R., Abbott, T. D., et al. 2018, Barkat, Z., Rakavy, G., & Sack, N. 1967, Physical Review Living Reviews in Relativity, 21, 3, Letters, 18, 379, doi: 10.1103/PhysRevLett.18.379 doi: 10.1007/s41114-018-0012-9 Batta, A., Ramirez-Ruiz, E., & Fryer, C. 2017, ApJL, 846, —. 2019a, Physical Review X, 9, 031040, L15, doi: 10.3847/2041-8213/aa8506 doi: 10.1103/PhysRevX.9.031040 Belczynski, K., Holz, D. E., Bulik, T., & O’Shaughnessy, R. —. 2019b, ApJL, 882, L24, doi: 10.3847/2041-8213/ab3800 2016a, Nature, 534, 512, doi: 10.1038/nature18322 Abbott, R., Abbott, T. D., Abraham, S., et al. 2020a, Belczynski, K., Heger, A., Gladysz, W., et al. 2016b, A&A, arXiv e-prints, arXiv:2009.01075. 594, A97, doi: 10.1051/0004-6361/201628980 https://arxiv.org/abs/2009.01075 Bemmerer, D., Cowan, T. E., Grieger, M., et al. 2018, in —. 2020b, arXiv e-prints, arXiv:2009.01190. European Physical Journal Web of Conferences, Vol. 178, https://arxiv.org/abs/2009.01190 European Physical Journal Web of Conferences, 01008, Acernese, F., Agathos, M., Agatsuma, K., et al. 2015, doi: 10.1051/epjconf/201817801008 Classical and Quantum Gravity, 32, 024001, Bond, J. R., Arnett, W. D., & Carr, B. J. 1984, ApJ, 280, doi: 10.1088/0264-9381/32/2/024001 825, doi: 10.1086/162057 Ali-Ha¨ımoud,Y., Kovetz, E. D., & Kamionkowski, M. 2017, Boothroyd, A. I., & Sackmann, I. J. 1988, ApJ, 328, 653, PhRvD, 96, 123523, doi: 10.1103/PhysRevD.96.123523 doi: 10.1086/166323 An, Z.-D., Ma, Y.-G., Fan, G.-T., et al. 2016, ApJL, 817, Bouffanais, Y., Mapelli, M., Gerosa, D., et al. 2019, ApJ, L5, doi: 10.3847/2041-8205/817/1/L5 886, 25, doi: 10.3847/1538-4357/ab4a79 An, Z.-D., Chen, Z.-P., Ma, Y.-G., et al. 2015, PhRvC, 92, Brown, G. E., Heger, A., Langer, N., et al. 2001, NewA, 6, 045802, doi: 10.1103/PhysRevC.92.045802 457, doi: 10.1016/S1384-1076(01)00077-X Angulo, C., Arnould, M., Rayet, M., et al. 1999, NuPhA, Buchler, J. R., & Yueh, W. R. 1976, ApJ, 210, 440, 656, 3, doi: 10.1016/S0375-9474(99)00030-5 doi: 10.1086/154847 Arca Sedda, M., Mapelli, M., Spera, M., Benacquista, M., Burbidge, E. M., Burbidge, G. R., Fowler, W. A., & Hoyle, & Giacobbo, N. 2020, ApJ, 894, 133, F. 1957, Reviews of Modern Physics, 29, 547, doi: 10.3847/1538-4357/ab88b2 doi: 10.1103/RevModPhys.29.547 Arcavi, I., Hosseinzadeh, G., Howell, D. A., et al. 2017, Carr, B., K¨uhnel,F., & Sandstad, M. 2016, PhRvD, 94, Nature, 551, 64, doi: 10.1038/nature24291 083504, doi: 10.1103/PhysRevD.94.083504 Arnett, W. D., & Truran, J. W. 1969, ApJ, 157, 339, Cassisi, S., Potekhin, A. Y., Pietrinferni, A., Catelan, M., & doi: 10.1086/150072 Salaris, M. 2007, ApJ, 661, 1094, doi: 10.1086/516819 18

Caughlan, G. R., & Fowler, W. A. 1988, Atomic Data and Fields, C. E., Timmes, F. X., Farmer, R., et al. 2018, The Nuclear Data Tables, 40, 283, Astrophysical Journal Supplement Series, 234, 19, doi: 10.1016/0092-640X(88)90009-5 doi: 10.3847/1538-4365/aaa29b Chen, K.-J., Woosley, S., Heger, A., Almgren, A., & Fishbach, M., Farr, W. M., & Holz, D. E. 2020, ApJL, 891, Whalen, D. J. 2014, The Astrophysical Journal, 792, 28. L31, doi: 10.3847/2041-8213/ab77c9 http://stacks.iop.org/0004-637X/792/i=1/a=28 Fishbach, M., & Holz, D. E. 2017, ApJL, 851, L25, Chugunov, A. I., Dewitt, H. E., & Yakovlev, D. G. 2007, doi: 10.3847/2041-8213/aa9bf6 PhRvD, 76, 025028, doi: 10.1103/PhysRevD.76.025028 Fishbach, M., Holz, D. E., & Farr, B. 2017, ApJL, 840, L24, Cyburt, R. H., Amthor, A. M., Ferguson, R., et al. 2010, doi: 10.3847/2041-8213/aa7045 The Astrophysical Journal Supplement Series, 189, 240, Fowler, W. A., & Hoyle, F. 1964, The Astrophysical doi: 10.1088/0067-0049/189/1/240 Journal Supplement Series, 9, 201, doi: 10.1086/190103 de Mink, S. E., & Mandel, I. 2016, MNRAS, 460, 3545, Fragione, G., & Kocsis, B. 2018, PhRvL, 121, 161103, doi: 10.1093/mnras/stw1219 doi: 10.1103/PhysRevLett.121.161103 deBoer, R. J., G¨orres,J., Wiescher, M., et al. 2017, Fraley, G. S. 1968, Ap&SS, 2, 96, doi: 10.1007/BF00651498 Reviews of Modern Physics, 89, 035007, Friˇsˇci´c,I., Donnelly, T. W., & Milner, R. G. 2019, PhRvC, doi: 10.1103/RevModPhys.89.035007 100, 025804, doi: 10.1103/PhysRevC.100.025804 Di Carlo, U. N., Giacobbo, N., Mapelli, M., et al. 2019, Fruet, G., Courtin, S., Heine, M., et al. 2020, Phys. Rev. MNRAS, 487, 2947, doi: 10.1093/mnras/stz1453 Lett., 124, 192701, doi: 10.1103/PhysRevLett.124.192701 Fryer, C. L. 1999, ApJ, 522, 413, doi: 10.1086/307647 Di Carlo, U. N., Mapelli, M., Bouffanais, Y., et al. 2020a, Fryer, C. L., Belczynski, K., Wiktorowicz, G., et al. 2012, MNRAS, 497, 1043, doi: 10.1093/mnras/staa1997 ApJ, 749, 91, doi: 10.1088/0004-637X/749/1/91 Di Carlo, U. N., Mapelli, M., Giacobbo, N., et al. 2020b, Fryer, C. L., & Warren, M. S. 2004, ApJ, 601, 391, MNRAS, doi: 10.1093/mnras/staa2286 doi: 10.1086/380193 Dominik, M., Belczynski, K., Fryer, C., et al. 2012, ApJ, Fryer, C. L., Woosley, S. E., & Heger, A. 2001, ApJ, 550, 759, 52, doi: 10.1088/0004-637X/759/1/52 372, doi: 10.1086/319719 Evans, M., Hastings, N., & Peacock, B. 2000, Statistical Fuller, G. M., Fowler, W. A., & Newman, M. J. 1985, ApJ, Distributions, 3rd edn. (Hoboken, NJ: Wiley) 293, 1, doi: 10.1086/163208 Farmer, R. 2018, rjfarmer/mesaplot, Gerosa, D., & Berti, E. 2017, PhRvD, 95, 124046, doi: 10.5281/zenodo.1441329 doi: 10.1103/PhysRevD.95.124046 Farmer, R., & Bauer, E. B. 2018, rjfarmer/pyMesa: Add —. 2019, PhRvD, 100, 041301, support for 10398, v1.0.3, Zenodo, doi: 10.1103/PhysRevD.100.041301 doi: 10.5281/zenodo.1205271 Gilkis, A., Soker, N., & Papish, O. 2016, ApJ, 826, 178, Farmer, R., Fields, C. E., Petermann, I., et al. 2016, ApJS, doi: 10.3847/0004-637X/826/2/178 227, 22, doi: 10.3847/1538-4365/227/2/22 Gomez, S., Berger, E., Nicholl, M., et al. 2019, ApJ, 881, Farmer, R., Fields, C. E., & Timmes, F. X. 2015, ApJ, 807, 87, doi: 10.3847/1538-4357/ab2f92 184, doi: 10.1088/0004-637X/807/2/184 Graham, M. J., Ford, K. E. S., McKernan, B., et al. 2020, Farmer, R., Renzo, M., de Mink, S. E., Marchant, P., & PhRvL, 124, 251102, Justham, S. 2019, ApJ, 887, 53, doi: 10.1103/PhysRevLett.124.251102 doi: 10.3847/1538-4357/ab518b Hamann, W. R., & Koesterke, L. 1998, A&A, 335, 1003 Farr, W. M., Fishbach, M., Ye, J., & Holz, D. E. 2019, Hamann, W.-R., Koesterke, L., & Wessolowski, U. 1995, ApJL, 883, L42, doi: 10.3847/2041-8213/ab4284 A&A, 299, 151 Ferguson, J. W., Alexander, D. R., Allard, F., et al. 2005, Hamann, W.-R., Schoenberner, D., & Heber, U. 1982, ApJ, 623, 585, doi: 10.1086/428642 A&A, 116, 273 Fern´andez,R., Quataert, E., Kashiyama, K., & Coughlin, Hammache, F., Oulebsir, N., Roussel, P., et al. 2016, in E. R. 2018, MNRAS, 476, 2366, Journal of Physics Conference Series, Vol. 665, Journal of doi: 10.1093/mnras/sty306 Physics Conference Series, 012007, Fields, C. E., Farmer, R., Petermann, I., Iliadis, C., & doi: 10.1088/1742-6596/665/1/012007 Timmes, F. X. 2016, ApJ, 823, 46, Hammer, J. W., Fey, M., Kunz, R., et al. 2005, NuPhA, doi: 10.3847/0004-637X/823/1/46 752, 514, doi: 10.1016/j.nuclphysa.2005.02.056 19

Heger, A., Fryer, C. L., Woosley, S. E., Langer, N., & Langanke, K., & Mart´ınez-Pinedo,G. 2000, Nuclear Hartmann, D. H. 2003, ApJ, 591, 288, Physics A, 673, 481, doi: 10.1016/S0375-9474(00)00131-7 doi: 10.1086/375341 Leung, S.-C., Nomoto, K., & Blinnikov, S. 2019, ApJ, 887, Heger, A., & Woosley, S. E. 2002, ApJ, 567, 532, 72, doi: 10.3847/1538-4357/ab4fe5 doi: 10.1086/338487 LIGO Scientific Collaboration, Aasi, J., Abbott, B. P., Heger, A., Woosley, S. E., Rauscher, T., Hoffman, R. D., & et al. 2015, Classical and Quantum Gravity, 32, 074001, Boyes, M. M. 2002, NewAR, 46, 463, doi: 10.1088/0264-9381/32/7/074001 doi: 10.1016/S1387-6473(02)00184-7 Lippuner, J., & Roberts, L. F. 2017, ApJS, 233, 18, Heger, A., Woosley, S. E., & Spruit, H. C. 2005, ApJ, 626, doi: 10.3847/1538-4365/aa94cb 350, doi: 10.1086/429868 Longland, R., Iliadis, C., Champagne, A. E., et al. 2010, Hoffman, R. D., Woosley, S. E., Weaver, T. A., Rauscher, Nuclear Physics A, 841, 1, T., & Thielemann, F. K. 1999, ApJ, 521, 735, doi: 10.1016/j.nuclphysa.2010.04.008 doi: 10.1086/307568 Lovegrove, E., & Woosley, S. E. 2013, ApJ, 769, 109, Holt, R. J., Filippone, B. W., & Pieper, S. C. 2018, arXiv doi: 10.1088/0004-637X/769/2/109 e-prints, arXiv:1809.10176. Lunnan, R., Fransson, C., Vreeswijk, P. M., et al. 2018, https://arxiv.org/abs/1809.10176 Nature Astronomy, doi: 10.1038/s41550-018-0568-z —. 2019, PhRvC, 99, 055802, Mandel, I. 2010, PhRvD, 81, 084029, doi: 10.1103/PhysRevC.99.055802 doi: 10.1103/PhysRevD.81.084029 Holz, D. E., & Hughes, S. A. 2005, ApJ, 629, 15, Mandel, I., & de Mink, S. E. 2016, MNRAS, 458, 2634, doi: 10.1086/431341 doi: 10.1093/mnras/stw379 Hummel, J. A., Pawlik, A. H., Milosavljevi´c,M., & Bromm, Mandel, I., Farr, W. M., & Gair, J. R. 2019, MNRAS, 486, V. 2012, ApJ, 755, 72, doi: 10.1088/0004-637X/755/1/72 1086, doi: 10.1093/mnras/stz896 Hunter, J. D. 2007, Computing In Science & Mangiagli, A., Bonetti, M., Sesana, A., & Colpi, M. 2019, Engineering, 9, 90 ApJL, 883, L27, doi: 10.3847/2041-8213/ab3f33 Iglesias, C. A., & Rogers, F. J. 1993, ApJ, 412, 752, Marchant, P., Langer, N., Podsiadlowski, P., Tauris, T. M., doi: 10.1086/172958 & Moriya, T. J. 2016, A&A, 588, A50, —. 1996, ApJ, 464, 943, doi: 10.1086/177381 doi: 10.1051/0004-6361/201628133 Iliadis, C., Champagne, A., Jos´e,J., Starrfield, S., & Marchant, P., Renzo, M., Farmer, R., et al. 2019, ApJ, 882, Tupper, P. 2002, ApJS, 142, 105, doi: 10.1086/341400 36, doi: 10.3847/1538-4357/ab3426 Iliadis, C., Longland, R., Champagne, A. E., & Coc, A. McKernan, B., Ford, K. E. S., Bellovary, J., et al. 2018, 2010a, Nuclear Physics A, 841, 251, ApJ, 866, 66, doi: 10.3847/1538-4357/aadae5 doi: 10.1016/j.nuclphysa.2010.04.010 Metcalfe, T. S. 2003, ApJL, 587, L43, doi: 10.1086/375044 Iliadis, C., Longland, R., Champagne, A. E., Coc, A., & Metcalfe, T. S., Salaris, M., & Winget, D. E. 2002, ApJ, Fitzgerald, R. 2010b, Nuclear Physics A, 841, 31, 573, 803, doi: 10.1086/340796 doi: 10.1016/j.nuclphysa.2010.04.009 Metcalfe, T. S., Winget, D. E., & Charbonneau, P. 2001, Itoh, N., Hayashi, H., Nishikawa, A., & Kohyama, Y. 1996, ApJ, 557, 1021, doi: 10.1086/321643 The Astrophysical Journal Supplement Series, 102, 411, Nadezhin, D. K. 1980, Ap&SS, 69, 115, doi: 10.1086/192264 doi: 10.1007/BF00638971 Kimball, C., Talbot, C., Berry, C. P. L., et al. 2020, arXiv Nicholl, M., Blanchard, P. K., Berger, E., et al. 2020, e-prints, arXiv:2005.00023. Nature Astronomy, doi: 10.1038/s41550-020-1066-7 https://arxiv.org/abs/2005.00023 Nitz, A. H., Dent, T., Davies, G. S., et al. 2020, The Kluyver, T., Ragan-Kelley, B., P´erez,F., et al. 2016, in Astrophysical Journal, 891, 123, Positioning and Power in Academic Publishing: Players, doi: 10.3847/1538-4357/ab733f Agents and Agendas: Proceedings of the 20th Oda, T., Hino, M., Muto, K., Takahara, M., & Sato, K. International Conference on Electronic Publishing, IOS 1994, Atomic Data and Nuclear Data Tables, 56, 231, Press, 87 doi: 10.1006/adnd.1994.1007 Kulkarni, S. R., Hut, P., & McMillan, S. 1993, Nature, 364, Pastorello, A., Smartt, S. J., Mattila, S., et al. 2007, 421, doi: 10.1038/364421a0 Nature, 447, 829, doi: 10.1038/nature05825 Kunz, R., Fey, M., Jaeger, M., et al. 2002, ApJ, 567, 643, Paxton, B., Bildsten, L., Dotter, A., et al. 2011, ApJS, 192, doi: 10.1086/338384 3, doi: 10.1088/0067-0049/192/1/3 20

Paxton, B., Cantiello, M., Arras, P., et al. 2013, ApJS, 208, Salvatier, J., Wieckiˆa,T. V., & Fonnesbeck, C. 2016, 4, doi: 10.1088/0067-0049/208/1/4 PyMC3: Python probabilistic programming framework. Paxton, B., Marchant, P., Schwab, J., et al. 2015, ApJS, http://ascl.net/1610.016 220, 15, doi: 10.1088/0067-0049/220/1/15 Saumon, D., Chabrier, G., & van Horn, H. M. 1995, ApJS, Paxton, B., Schwab, J., Bauer, E. B., et al. 2018, ApJS, 99, 713, doi: 10.1086/192204 234, 34, doi: 10.3847/1538-4365/aaa5a8 Schutz, B. F. 1986, Nature, 323, 310, doi: 10.1038/323310a0 Paxton, B., Smolec, R., Schwab, J., et al. 2019, ApJS, 243, Shen, Y. P., Guo, B., deBoer, R. J., et al. 2020, Phys. Rev. 10, doi: 10.3847/1538-4365/ab2241 Lett., 124, 162701, doi: 10.1103/PhysRevLett.124.162701 P´erez,F., & Granger, B. E. 2007, Computing in Science & Spera, M., Mapelli, M., Giacobbo, N., et al. 2019, MNRAS, Engineering, 9, 21 485, 889, doi: 10.1093/mnras/stz359 Pols, O. R., Tout, C. A., Eggleton, P. P., & Han, Z. 1995, Stone, N. C., Metzger, B. D., & Haiman, Z. 2017, MNRAS, MNRAS, 274, 964, doi: 10.1093/mnras/274.3.964 464, 946, doi: 10.1093/mnras/stw2260 Portegies Zwart, S. F., & McMillan, S. L. W. 2000, ApJ, Stothers, R. B. 1999, MNRAS, 305, 365, 528, L17, doi: 10.1086/312422 doi: 10.1046/j.1365-8711.1999.02444.x Potekhin, A. Y., & Chabrier, G. 2010, Contributions to Straniero, O., Dom´ınguez,I., Imbriani, G., & Piersanti, L. Plasma Physics, 50, 82, doi: 10.1002/ctpp.201010017 2003, ApJ, 583, 878, doi: 10.1086/345427 Sukhbold, T., & Adams, S. 2020, MNRAS, 492, 2578, Quataert, E., Lecoanet, D., & Coughlin, E. R. 2019, doi: 10.1093/mnras/staa059 MNRAS, 485, L83, doi: 10.1093/mnrasl/slz031 Takahashi, K. 2018, ApJ, 863, 153, Rakavy, G., & Shaviv, G. 1967, ApJ, 148, 803, doi: 10.3847/1538-4357/aad2d2 doi: 10.1086/149204 Tan, W. P., Boeltzig, A., Dulal, C., et al. 2020, Phys. Rev. Rauscher, T., Nishimura, N., Hirschi, R., et al. 2016, Lett., 124, 192702, doi: 10.1103/PhysRevLett.124.192702 MNRAS, 463, 4153, doi: 10.1093/mnras/stw2266 Thielemann, F.-K., Nomoto, K., & Hashimoto, M.-A. 1996, Rauscher, T., & Thielemann, F.-K. 2000, Atomic Data and ApJ, 460, 408, doi: 10.1086/176980 Nuclear Data Tables, 75, 1, doi: 10.1006/adnd.2000.0834 Timmes, F. X., & Swesty, F. D. 2000, ApJS, 126, 501, Reg˝os,E., Vink´o,J., & Ziegler, B. L. 2020, ApJ, 894, 94, doi: 10.1086/313304 doi: 10.3847/1538-4357/ab8636 Toro, E. F., Spruce, M., & Speares, W. 1994, Shock Waves, Renzo, M., Farmer, R., Justham, S., et al. 2020a, A&A, 4, 25, doi: 10.1007/BF01414629 640, A56, doi: 10.1051/0004-6361/202037710 Townsend, R. 2019, MESA SDK for Linux: 20190911, Renzo, M., Farmer, R. J., Justham, S., et al. 2020b, doi: 10.5281/zenodo.2603170 MNRAS, 493, 4333, doi: 10.1093/mnras/staa549 Tumino, A., Spitaleri, C., La Cognata, M., et al. 2018, Rockefeller, G., Fryer, C. L., & Li, H. 2006, arXiv e-prints, Nature, 557, 687, doi: 10.1038/s41586-018-0149-4 astro. https://arxiv.org/abs/astro-ph/0608028 Tur, C., Heger, A., & Austin, S. M. 2007, ApJ, 671, 821, Rodriguez, C. L., Amaro-Seoane, P., Chatterjee, S., et al. doi: 10.1086/523095 2018, PhRvD, 98, 123005, Tutukov, A. V., & Yungelson, L. R. 1993, MNRAS, 260, doi: 10.1103/PhysRevD.98.123005 675, doi: 10.1093/mnras/260.3.675 Rodriguez, C. L., Chatterjee, S., & Rasio, F. A. 2016, Udall, R., Jani, K., Lange, J., et al. 2019, arXiv e-prints, PhRvD, 93, 084029, doi: 10.1103/PhysRevD.93.084029 arXiv:1912.10533. https://arxiv.org/abs/1912.10533 Rodriguez, C. L., Zevin, M., Amaro-Seoane, P., et al. 2019, van der Walt, S., Colbert, S. C., & Varoquaux, G. 2011, PhRvD, 100, 043027, doi: 10.1103/PhysRevD.100.043027 Computing in Science Engineering, 13, 22, Rogers, F. J., & Nayfonov, A. 2002, ApJ, 576, 1064, doi: 10.1109/MCSE.2011.37 doi: 10.1086/341894 van Son, L. A. C., De Mink, S. E., Broekgaarden, F. S., Roupas, Z., & Kazanas, D. 2019, A&A, 632, L8, et al. 2020, ApJ, 897, 100, doi: 10.1051/0004-6361/201937002 doi: 10.3847/1538-4357/ab9809 Salaris, M., Dom´ınguez,I., Garc´ıa-Berro,E., et al. 1997, Vigna-G´omez,A., Justham, S., Mandel, I., de Mink, S. E., ApJ, 486, 413, doi: 10.1086/304483 & Podsiadlowski, P. 2019, ApJL, 876, L29, Sallaska, A. L., Iliadis, C., Champange, A. E., et al. 2013, doi: 10.3847/2041-8213/ab1bdf The Astrophysical Journal Supplement Series, 207, 18, Villar, V. A., Nicholl, M., & Berger, E. 2018, ApJ, 869, 166, doi: 10.1088/0067-0049/207/1/18 doi: 10.3847/1538-4357/aaee6a 21

Vitale, S., Lynch, R., Raymond, V., et al. 2017, PhRvD, 95, Woosley, S. E., & Heger, A. 2007, PhR, 442, 269, 064053, doi: 10.1103/PhysRevD.95.064053 doi: 10.1016/j.physrep.2007.02.009 Weaver, T. A., & Woosley, S. E. 1993, PhR, 227, 65, Woosley, S. E., Heger, A., & Weaver, T. A. 2002, Rev. doi: 10.1016/0370-1573(93)90058-L Mod. Phys., 74, 1015, doi: 10.1103/RevModPhys.74.1015 West, C., Heger, A., & Austin, S. M. 2013, ApJ, 769, 2, doi: 10.1088/0004-637X/769/1/2 Woosley, S. E., Timmes, F. X., & Weaver, T. A. 1993, Whalen, D. J., Fryer, C. L., Holz, D. E., et al. 2013a, Journal of Physics G Nuclear Physics, 19, S183, ApJL, 762, L6, doi: 10.1088/2041-8205/762/1/L6 doi: 10.1088/0954-3899/19/S/016 Whalen, D. J., Even, W., Frey, L. H., et al. 2013b, ApJ, Yang, Y., Bartos, I., Gayathri, V., et al. 2019, PhRvL, 123, 777, 110, doi: 10.1088/0004-637X/777/2/110 181101, doi: 10.1103/PhysRevLett.123.181101 Woosley, S. E. 2017, ApJ, 836, 244, Yoshida, T., Umeda, H., Maeda, K., & Ishii, T. 2016, doi: 10.3847/1538-4357/836/2/244 —. 2018, ApJ, 863, 105, doi: 10.3847/1538-4357/aad044 MNRAS, 457, 351, doi: 10.1093/mnras/stv3002 —. 2019, ApJ, 878, 49, doi: 10.3847/1538-4357/ab1b41 Young, D. R., Smartt, S. J., Mattila, S., et al. 2008, A&A, Woosley, S. E., Blinnikov, S., & Heger, A. 2007, Nature, 489, 359, doi: 10.1051/0004-6361:20078662 450, 390, doi: 10.1038/nature06333