Element formation in radiation-hydrodynamics simulations of kilonovae

Fabio Magistrelli [email protected] Theoretisch-Physikalisches Institut, Friedrich-Schiller-Universität Jena, 07743, Jena, Germany Sebastiano Bernuzzi [email protected] Theoretisch-Physikalisches Institut, Friedrich-Schiller-Universität Jena, 07743, Jena, Germany Albino Perego [email protected] Dipartimento di Fisica, Università di Trento, Via Sommarive 14, 38123 Trento, Italy INFN-TIFPA, Trento Institute for Fundamental Physics and Applications, via Sommarive 14, I-38123 Trento, Italy David Radice [email protected] Institute for Gravitation & the Cosmos, The Pennsylvania State University, University Park, PA 16802 Department of Physics, The Pennsylvania State University, University Park, PA 16802 Department of Astronomy & Astrophysics, The Pennsylvania State University,University Park, PA 16802
(October 3, 2024)
Abstract

Understanding the details of r𝑟ritalic_r-process nucleosynthesis in binary neutron star mergers (BNSM) ejecta is key to interpreting kilonova observations and identifying the role of BNSMs in the origin of heavy elements. We present a self-consistent two-dimensional, ray-by-ray radiation-hydrodynamic evolution of BNSM ejecta with an online nuclear network (NN) up to the days timescale. For the first time, an initial numerical-relativity ejecta profile composed of the dynamical component, spiral-wave and disk winds is evolved including detailed r𝑟ritalic_r-process reactions and nuclear heating effects. A simple model for the jet energy deposition is also included. Our simulation highlights that the common approach of relating in post-processing the final nucleosynthesis yields to the initial thermodynamic profile of the ejecta can lead to inaccurate predictions. Moreover, we find that neglecting the details of the radiation-hydrodynamic evolution of the ejecta in nuclear calculations can introduce deviations up to one order of magnitude in the final abundances of several elements, including very light and second r𝑟ritalic_r-process peak elements. The presence of a jet affects element production only in the innermost part of the polar ejecta, and it does not alter the global nucleosynthesis results. Overall, our analysis shows that employing an online NN improves the reliability of nucleosynthesis and kilonova light curves predictions.

\tensordelimiter

? \tensorformatlrc

thanks: Alfred P. Sloan fellow

1 Introduction

Mass ejecta from binary neutron star mergers (BNSMs) are primary sites for rapid neutron capture (r𝑟ritalic_r-process) nucleosynthesis (Eichler et al., 1989), see e.g. (Cowan et al., 2021; Perego et al., 2021; Arcones & Thielemann, 2023). The heavy, neutron-rich elements produced in these environments undergo radioactive decays, powering an electromagnetic (EM) transient known as a kilonova. Moreover, some of these mergers can produce short gamma-ray bursts (sGRBs), offering further insights into their astrophysical properties. The unambiguous detection of gravitational waves (GW170817) and EM counterparts (kilonova AT2017gfo and GRB 170817A) from a BNSM in August 2017 has confirmed theoretical predictions and triggered intense work on the subject (Abbott et al., 2017a, b, c; Coulter et al., 2017; Tanvir et al., 2017; Savchenko et al., 2017).

Numerical simulations are the main tool to explore the physics of BNSMs and identify the mechanisms underlying the associated gravitational and EM emissions (Rosswog et al., 2014; Roberts et al., 2017; Nedora et al., 2021; Radice et al., 2022; Zappa et al., 2023; Combi & Siegel, 2023a, b; Schianchi et al., 2024; Radice & Bernuzzi, 2023; Kiuchi et al., 2024; Musolino et al., 2024). Simulations are used to carry out extensive investigations on r𝑟ritalic_r-process nucleosynthesis and its relationship with kilonovae. Lanthanides and actinides production is consistently found in low electron fraction ejected material (Ye0.2less-than-or-similar-tosubscript𝑌𝑒0.2Y_{e}\lesssim 0.2italic_Y start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT ≲ 0.2) (Korobkin et al., 2012; Lippuner & Roberts, 2015; Perego et al., 2021). Low values of Yesubscript𝑌𝑒Y_{e}italic_Y start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT can be found in the equatorial component of dynamical ejecta (Wanajo et al., 2014; Radice et al., 2016; Kiuchi et al., 2023) and in the late disk wind, e.g. (Beloborodov, 2003; Siegel & Metzger, 2017; Sprouse et al., 2024; Kiuchi et al., 2024). The high opacity associated with these elements commonly links the outer regions of the (low-latitude) ejecta with the red component of the kilonova light curves (Metzger & Fernández, 2014; Perego et al., 2017). However, small expansion timescales can inhibit neutron captures and thus produce a blue-UV precursor powered by decays of free neutrons (Metzger et al., 2015; Radice et al., 2018a; Combi & Siegel, 2023a). Higher values of Ye0.3greater-than-or-equivalent-tosubscript𝑌𝑒0.3Y_{e}\gtrsim 0.3italic_Y start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT ≳ 0.3 associated with most of the disk and the majority of polar ejecta prevent strong r𝑟ritalic_r-processes, thus leading to low opacity matter contributing to the blue part of the kilonova spectra (Metzger & Fernández, 2014; Kasen et al., 2017; Tanaka et al., 2017; Curtis et al., 2024). In the polar regions, this is enhanced by interactions with neutrinos emitted by the central remnant until its possible collapse to a black hole (Radice & Bernuzzi, 2024). The passage of a relativistic jet can in principle affect the nucleosynthesis by injecting extra energy in the system and thus reigniting suppressed reactions (Janiuk, 2014). Understanding the details of the nuclear evolution of the ejecta and how this affects its dynamics is crucial for accurately interpreting observational data and assessing the role of BNSMs in explaining the origin of r𝑟ritalic_r-process elements in the universe.

One of the main open challenges towards the accurate prediction of r𝑟ritalic_r-process nucleosynthesis in BNSMs ejecta is capturing the dependence of the nuclear composition outcomes on the initial thermodynamic profile and detailed hydrodynamic evolution of the ejecta. Traditional approaches assume homologously expanding ejecta and neglect radiation transport, hydrodynamics and ejecta self-interaction, e.g. (Korobkin et al., 2012; Radice et al., 2016; Rosswog et al., 2017; Perego et al., 2022; Curtis et al., 2024), although some recent efforts have considered detailed radiation transport, e.g. (Collins et al., 2023; Shingles et al., 2023). Alternatively, nuclear networks are employed in a post-processing step on hydrodynamics profiles to obtain yields and heating rates (Just et al., 2015; Martin et al., 2015; Roberts et al., 2017). The latter are sometimes fitted as time-domain functions and re-used in hydrodynamical simulations to improve ejecta evolutions, e.g. (Rosswog et al., 2014; Wu et al., 2022; Ricigliano et al., 2024).

In this work, we present, for the first time, a self-consistent simulation of r𝑟ritalic_r-process nucleosynthesis for a fiducial BNSM ejecta using a 2-dimensional, ray-by-ray radiation hydrodynamics code coupled with an online nuclear network (NN). In Section 2, we present our model and discuss how the NN is coupled with the radiation-hydrodynamic equations to update the ejecta composition and calculate the associated nuclear power in each hydrodynamic step. We also describe our implementation of thermalization processes and of an extra energy source modeling a sGRB. Section 3 presents the nucleosynthesis and kilonova light curves predictions for the ejecta profiles we extract from a fiducial BNSM numerical relativity simulation. In Section 4, we conclude discussing the implications of our findings and emphasizing the importance of implementing an online NN for self-consistent predictions of r𝑟ritalic_r-process nucleosynthesis patterns.

2 Method

2.1 Numerical-relativity profiles

Our simulation is initialized with initial ejecta profiles extracted from ab-initio numerical relativity simulations that include microphysics, M0 neutrino transport111See Zappa et al. (2023) for the importance of including neutrino heating effects. and magnetic-field induced turbulence (Radice et al., 2018b; Perego et al., 2019; Bernuzzi, 2020; Nedora et al., 2021). The considered binary has mass ratio q=1.43𝑞1.43q=1.43italic_q = 1.43 and produces a short-lived remnant that collapses to a black hole (BH) at about 10101010 ms postmerger. The simulation is run with the LS220 equation of state (EoS) and with a typical resolution of 185185185185 m up to 28.528.528.528.5 ms after merger (Nedora et al., 2021). Throughout the simulation, the unbound material is identified via the Bernoulli criterion and collected at an extraction radius of Rext295similar-to-or-equalssubscript𝑅ext295R_{\text{ext}}\simeq 295italic_R start_POSTSUBSCRIPT ext end_POSTSUBSCRIPT ≃ 295 km. The ejecta consists of the dynamical component of mass 1.24×102Msimilar-toabsent1.24superscript102subscriptMdirect-product{\sim}1.24\times 10^{-2}~{}{{\rm M_{\odot}}}∼ 1.24 × 10 start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT and a short spiral-wave wind (0.80×102Msimilar-toabsent0.80superscript102subscriptMdirect-product{\sim}0.80\times 10^{-2}~{}{{\rm M_{\odot}}}∼ 0.80 × 10 start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT). In order to prepare the subsequent long-term evolution, the dependence of the ejecta properties on the azimuthal coordinate is integrated out, while the polar dependence is accounted for by discretizing the polar angle in 51515151 angular sections. The time-dependent ejecta profile is mapped to a Lagrangian profile by positioning the latest shell at Rextsubscript𝑅extR_{\text{ext}}italic_R start_POSTSUBSCRIPT ext end_POSTSUBSCRIPT and progressively layering previously ejected shells on top as prescribed by the constraint m(r)=4πRextrρ(r)r2𝑑r𝑚𝑟4𝜋superscriptsubscriptsubscript𝑅ext𝑟𝜌𝑟superscript𝑟2differential-d𝑟m(r)=4\pi\int_{R_{\text{ext}}}^{r}\rho(r)\,r^{2}dritalic_m ( italic_r ) = 4 italic_π ∫ start_POSTSUBSCRIPT italic_R start_POSTSUBSCRIPT ext end_POSTSUBSCRIPT end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_r end_POSTSUPERSCRIPT italic_ρ ( italic_r ) italic_r start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_d italic_r, with ρ𝜌\rhoitalic_ρ mass density and r𝑟ritalic_r position of the radius including the mass m𝑚mitalic_m (Wu et al., 2022). An analytical disk wind profile of mass 2.75×102Msimilar-toabsent2.75superscript102subscriptMdirect-product{\sim}2.75\times 10^{-2}~{}{{\rm M_{\odot}}}∼ 2.75 × 10 start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT is constructed from a similar simulation but performed on an equal mass binary with the M1 neutrino transport (Radice et al., 2022), which better describes the optically thick regime (Zappa et al., 2023). The simulation of the q=1𝑞1q=1italic_q = 1 binary is run with the SFHo EoS for 270similar-toabsent270{\sim}270∼ 270 ms postmerger222 We take the results of this simulation, where the BH collapse occurs around t10 mssimilar-to-or-equals𝑡10 mst\simeq 10\textrm{ ms}italic_t ≃ 10 ms postmerger, as a representative history of disk ejecta. The general properties of disks and disk ejecta are found to be relatively stable against variations in the properties of the compact binary or numerical prescriptions (e.g. EoS, effective viscosity, neutrino treatment), as long as the BH collapse occurs in a timescale of 𝒪(10) ms𝒪10 ms\mathcal{O}(10)\textrm{ ms}caligraphic_O ( 10 ) ms (Fernández & Metzger, 2013; Just et al., 2021, 2023; Kiuchi et al., 2023, 2024; Camilletti et al., 2024).. The disk profile is constructed by interpolating the spherically averaged density, temperature, and electron fraction of the unbound matter over time and using second-order Padé approximants. We assume a constant velocity and entropy, and rescale the density evolution to get a total ejection of 40%percent4040\%40 % of the total disk mass (of the LS220 simulation) assuming a sin2θproportional-toabsentsuperscript2𝜃\propto\sin^{2}\!\theta∝ roman_sin start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_θ angular distribution (Perego et al., 2017; Fernández et al., 2019). The complete initial profiles of two selected angular sections are shown in the two bottom panels of Fig. 1.

2.2 Ray-by-ray Radiation-Hydrodynamics

The ejecta profile is evolved with the system of Lagrangian radiation-hydrodynamic equations as implemented in Morozova et al. (2015); Wu et al. (2022). In particular, the energy equation in spherical symmetry reads

ϵt=Pρlnρt4πr2QvmLm+ϵ˙nucl,italic-ϵ𝑡𝑃𝜌𝜌𝑡4𝜋superscript𝑟2𝑄𝑣𝑚𝐿𝑚subscript˙italic-ϵnucl\frac{\partial\epsilon}{\partial t}=\frac{P}{\rho}\,\frac{\partial\ln\rho}{% \partial t}-4\pi r^{2}Q\frac{\partial v}{\partial m}-\frac{\partial L}{% \partial m}+\dot{\epsilon}_{\text{nucl}}\,,divide start_ARG ∂ italic_ϵ end_ARG start_ARG ∂ italic_t end_ARG = divide start_ARG italic_P end_ARG start_ARG italic_ρ end_ARG divide start_ARG ∂ roman_ln italic_ρ end_ARG start_ARG ∂ italic_t end_ARG - 4 italic_π italic_r start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_Q divide start_ARG ∂ italic_v end_ARG start_ARG ∂ italic_m end_ARG - divide start_ARG ∂ italic_L end_ARG start_ARG ∂ italic_m end_ARG + over˙ start_ARG italic_ϵ end_ARG start_POSTSUBSCRIPT nucl end_POSTSUBSCRIPT , (1)

with m𝑚mitalic_m mass coordinate, ϵitalic-ϵ\epsilonitalic_ϵ specific internal energy, t𝑡titalic_t time, P𝑃Pitalic_P pressure, v𝑣vitalic_v (radial) velocity, L𝐿Litalic_L luminosity and Q𝑄Qitalic_Q von Neumann-Richtmyer artificial viscosity (Von Neumann & Richtmyer, 1950). The specific energy deposition ϵ˙nuclsubscript˙italic-ϵnucl\dot{\epsilon}_{\text{nucl}}over˙ start_ARG italic_ϵ end_ARG start_POSTSUBSCRIPT nucl end_POSTSUBSCRIPT accounts for the local production and partial thermalization of the energy released by all possible nuclear reactions occurring in the expanding material. To calculate the luminosity, we employ the same analytic, time-independent opacity introduced in Wu et al. (2022).

The set of hydrodynamic equations is closed by the EoS. In the high-temperature regime, we implement the tabulated Helmholtz EoS introduced in Timmes & Swesty (2000) and also used in Lippuner & Roberts (2017). For lower temperatures, where the contribution of positrons is negligible, we switch to the Paczynski EoS introduced in Paczynski (1986) and utilized in Morozova et al. (2015); Wu et al. (2022). This analytical EoS describes a mixture of ideal gases, namely a Boltzmann gas of non-degenerate, non-relativistic ions, an ideal gas of arbitrarily degenerate and relativistic electrons, and a photon ideal gas. The inclusion of an online nuclear network, as described in Sec. 2.3, improves the realism of these EoS, as the mean molecular weight and the electron fraction, which enter the EoS, are self-consistently calculated on the fly.

To account for the angular dependency of the ejecta properties, we employ the spherically symmetric hydrodynamic equations in a ray-by-ray fashion. For each angular section, we first map the profile into an effective 1D problem by multiplying the total mass by the scaling factor λθ=4π/ΔΩsubscript𝜆𝜃4𝜋ΔΩ\lambda_{\theta}=4\pi/\Delta\Omegaitalic_λ start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT = 4 italic_π / roman_Δ roman_Ω, where ΔΩ2πsinθdθsimilar-to-or-equalsΔΩ2𝜋𝜃𝑑𝜃\Delta\Omega\simeq 2\pi\sin{\theta}\,d\thetaroman_Δ roman_Ω ≃ 2 italic_π roman_sin italic_θ italic_d italic_θ represents the solid angle included in the angular section. This ensures that all the intensive quantities (including density) remain fixed. Then, we independently evolve the different angular sections by discretizing the 1D hydrodynamic equations over nshsubscript𝑛shn_{\text{sh}}italic_n start_POSTSUBSCRIPT sh end_POSTSUBSCRIPT spherical fluid elements (mass shells). We ran tests for some selected sections with grid resolutions nsh=300,600,1000subscript𝑛sh3006001000n_{\text{sh}}=300,600,1000italic_n start_POSTSUBSCRIPT sh end_POSTSUBSCRIPT = 300 , 600 , 1000. We found nsh=600subscript𝑛sh600n_{\text{sh}}=600italic_n start_POSTSUBSCRIPT sh end_POSTSUBSCRIPT = 600 sufficiently high to resolve the details of the nucleosynthesis and decided to use this resolution for the complete simulation. Note that non-radial flows of matter and radiation are neglected, as well as higher dimensional fluid instabilities (e.g. Kelvin-Helmholtz), convection and the possibility of shells surpassing each other. This could in principle lead to over or underestimating the pressure work exchanged between the radial shells and introduce artificial shocks in the system. Finally, we map the problem back to the axisymmetric scenario by keeping the intensive quantities unchanged, rescaling the extensive ones by the scaling factor 1/λθ1subscript𝜆𝜃1/\lambda_{\theta}1 / italic_λ start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT and combining the results. In particular, we calculate the global mass fractions and abundances with a mass weighted average over all the mass shells and angular sections. Kilonova light curves are recombined accounting for the angle of view as in Martin et al. (2015); Perego et al. (2017).

2.3 Nuclear Network Coupling

We calculate the nuclear composition and specific energy deposition, ϵ˙nuclsubscript˙italic-ϵnucl\dot{\epsilon}_{\text{nucl}}over˙ start_ARG italic_ϵ end_ARG start_POSTSUBSCRIPT nucl end_POSTSUBSCRIPT, in a self-consistent way with an online implementation of the nuclear network (NN) SkyNet (Lippuner & Roberts, 2017). The NN includes 7836 isotopes up to 337Cn and uses the JINA REACLIB (Cyburt et al., 2010) and the same setup as in Lippuner & Roberts (2015); Perego et al. (2022). The simulation is started by initializing matter composition from the initial temperature T0(m)subscript𝑇0𝑚T_{0}(m)italic_T start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( italic_m ) and density ρ0(m)subscript𝜌0𝑚\rho_{0}(m)italic_ρ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( italic_m ) assuming nuclear statistical equilibrium (NSE). We thus impose, to each isotope considered in the NN, the abundance (Cowan et al., 2021; Perego et al., 2021)

Yi,0=ni,0nb,0=Yp,0ZiYn,0(AiZi)Gi(T)Ai3/22Ai(ρmb)Ai1×(2π2mbkBT)3(Ai1)/2eBEi/kBT,subscript𝑌𝑖0subscript𝑛𝑖0subscript𝑛𝑏0superscriptsubscript𝑌𝑝0subscript𝑍𝑖superscriptsubscript𝑌𝑛0subscript𝐴𝑖subscript𝑍𝑖subscript𝐺𝑖𝑇superscriptsubscript𝐴𝑖32superscript2subscript𝐴𝑖superscript𝜌subscript𝑚𝑏subscript𝐴𝑖1superscript2𝜋superscriptPlanck-constant-over-2-pi2subscript𝑚𝑏subscript𝑘B𝑇3subscript𝐴𝑖12superscript𝑒subscriptBE𝑖subscript𝑘B𝑇\begin{split}Y_{i,0}=\frac{n_{i,0}}{n_{b,0}}=Y_{p,0}^{Z_{i}}\,Y_{n,0}^{(A_{i}-% Z_{i})}\,\frac{G_{i}(T)\,A_{i}^{3/2}}{2^{A_{i}}}\left(\frac{\rho}{m_{b}}\right% )^{\!\!A_{i}-1}\\ \times\left(\frac{2\pi\hbar^{2}}{m_{b}k_{\text{B}}T}\right)^{\!\!3(A_{i}-1)/2}% e^{\text{BE}_{i}/k_{\text{B}}T}\,,\end{split}start_ROW start_CELL italic_Y start_POSTSUBSCRIPT italic_i , 0 end_POSTSUBSCRIPT = divide start_ARG italic_n start_POSTSUBSCRIPT italic_i , 0 end_POSTSUBSCRIPT end_ARG start_ARG italic_n start_POSTSUBSCRIPT italic_b , 0 end_POSTSUBSCRIPT end_ARG = italic_Y start_POSTSUBSCRIPT italic_p , 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_Z start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_POSTSUPERSCRIPT italic_Y start_POSTSUBSCRIPT italic_n , 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_A start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT - italic_Z start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) end_POSTSUPERSCRIPT divide start_ARG italic_G start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_T ) italic_A start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 3 / 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 start_POSTSUPERSCRIPT italic_A start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_POSTSUPERSCRIPT end_ARG ( divide start_ARG italic_ρ end_ARG start_ARG italic_m start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT end_ARG ) start_POSTSUPERSCRIPT italic_A start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT - 1 end_POSTSUPERSCRIPT end_CELL end_ROW start_ROW start_CELL × ( divide start_ARG 2 italic_π roman_ℏ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG italic_m start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT italic_k start_POSTSUBSCRIPT B end_POSTSUBSCRIPT italic_T end_ARG ) start_POSTSUPERSCRIPT 3 ( italic_A start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT - 1 ) / 2 end_POSTSUPERSCRIPT italic_e start_POSTSUPERSCRIPT BE start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT / italic_k start_POSTSUBSCRIPT B end_POSTSUBSCRIPT italic_T end_POSTSUPERSCRIPT , end_CELL end_ROW (2)

with mbsubscript𝑚𝑏m_{b}italic_m start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT and nbsubscript𝑛𝑏n_{b}italic_n start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT baryon mass and number density, Aisubscript𝐴𝑖A_{i}italic_A start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT and Zisubscript𝑍𝑖Z_{i}italic_Z start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT mass and atomic numbers of the i𝑖iitalic_i-th isotope and nisubscript𝑛𝑖n_{i}italic_n start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT and BEisubscriptBE𝑖\text{BE}_{i}BE start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT its number density and total binding energy333 Several mass shells have T<5𝑇5T<5italic_T < 5 GK at the end of the numerical relativity simulation (see Fig. 1). In the case of low entropy tidal ejecta, their initial composition should be close to the one set by cold NSE conditions inside isolated neutron stars. In all the other cases, we expect matter to reach hot NSE conditions inside our extraction radius, before matter expansion causes an NSE freeze-out. We plan to attest the impact of the early out-of-NSE evolution on our results by developing and comparing alternative initialization procedures for cold fluid elements in a future work.. Here, Gi(T)subscript𝐺𝑖𝑇G_{i}(T)italic_G start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_T ) is the internal partition function of the nucleus i𝑖iitalic_i at temperature T𝑇Titalic_T. The two conditions iAiYi=1subscript𝑖subscript𝐴𝑖subscript𝑌𝑖1\sum_{i}A_{i}Y_{i}=1∑ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_A start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_Y start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = 1 and iZiYi=Yesubscript𝑖subscript𝑍𝑖subscript𝑌𝑖subscript𝑌𝑒\sum_{i}Z_{i}Y_{i}=Y_{e}∑ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_Z start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_Y start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = italic_Y start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT, with Yesubscript𝑌𝑒Y_{e}italic_Y start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT electron fraction, constrain the NSE composition, which thus depends on ρ𝜌\rhoitalic_ρ, T𝑇Titalic_T and Yesubscript𝑌𝑒Y_{e}italic_Y start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT. See Lippuner & Roberts (2017) for details about SkyNet ​​’s implementation of Eq. (2). The isotopic mass fractions can be calculated from the element abundances via Xi=AiYisubscript𝑋𝑖subscript𝐴𝑖subscript𝑌𝑖X_{i}=A_{i}Y_{i}italic_X start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = italic_A start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_Y start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT.

At the initial time t0subscript𝑡0t_{0}italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT of each hydrodynamic step ΔtΔ𝑡\Delta troman_Δ italic_t, an independent instance of SkyNet is called in each mass shell. The NN evolves the composition with its own time sub-steps until t0+Δtsubscript𝑡0Δ𝑡t_{0}+\Delta titalic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT + roman_Δ italic_t. The (non-thermalized) power released by nuclear reactions is averaged over ΔtΔ𝑡\Delta troman_Δ italic_t to get the total nuclear power that will be then thermalized into ϵ˙nuclsubscript˙italic-ϵnucl\dot{\epsilon}_{\text{nucl}}over˙ start_ARG italic_ϵ end_ARG start_POSTSUBSCRIPT nucl end_POSTSUBSCRIPT as described below. During the sub-steps, the temperature is constant (no self-heating) and fixed by T(t0)𝑇subscript𝑡0T(t_{0})italic_T ( italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ), and it will only be updated by the hydrodynamic step. The density evolution is prescribed with a log-interpolation between ρ(t0)𝜌subscript𝑡0\rho(t_{0})italic_ρ ( italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) and ρ(t0+Δt)𝜌subscript𝑡0Δ𝑡\rho(t_{0}+\Delta t)italic_ρ ( italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT + roman_Δ italic_t ). The latter is calculated from the density and radial velocity profiles at t=t0𝑡subscript𝑡0t=t_{0}italic_t = italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT. Note that in our setup mass shells cannot mix.

To calculate the specific energy deposition in Eq. (1), we independently thermalize the different contributions to ϵ˙nucl=ϵ˙nuclγ+ϵ˙nuclα+ϵ˙nuclβ+ϵ˙nucl othsubscript˙italic-ϵnuclsuperscriptsubscript˙italic-ϵnucl𝛾superscriptsubscript˙italic-ϵnucl𝛼superscriptsubscript˙italic-ϵnucl𝛽superscriptsubscript˙italic-ϵnucl oth\dot{\epsilon}_{\text{nucl}}=\dot{\epsilon}_{\text{nucl}}^{\gamma}+\dot{% \epsilon}_{\text{nucl}}^{\alpha}+\dot{\epsilon}_{\text{nucl}}^{\beta}+\dot{% \epsilon}_{\text{nucl}}^{\text{ oth}}over˙ start_ARG italic_ϵ end_ARG start_POSTSUBSCRIPT nucl end_POSTSUBSCRIPT = over˙ start_ARG italic_ϵ end_ARG start_POSTSUBSCRIPT nucl end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_γ end_POSTSUPERSCRIPT + over˙ start_ARG italic_ϵ end_ARG start_POSTSUBSCRIPT nucl end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_α end_POSTSUPERSCRIPT + over˙ start_ARG italic_ϵ end_ARG start_POSTSUBSCRIPT nucl end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_β end_POSTSUPERSCRIPT + over˙ start_ARG italic_ϵ end_ARG start_POSTSUBSCRIPT nucl end_POSTSUBSCRIPT start_POSTSUPERSCRIPT oth end_POSTSUPERSCRIPT coming from γ𝛾\gammaitalic_γ rays, α𝛼\alphaitalic_α particles, electrons, and other nuclear reactions products. We compute the fraction of energy thermalized by the emission of γ𝛾\gammaitalic_γ-rays as

ϵ˙nuclγ(t)=jfjγ(t)EjγYj(t)τjmp,superscriptsubscript˙italic-ϵnucl𝛾𝑡subscript𝑗superscriptsubscript𝑓𝑗𝛾𝑡delimited-⟨⟩superscriptsubscript𝐸𝑗𝛾subscript𝑌𝑗𝑡subscript𝜏𝑗subscript𝑚𝑝\dot{\epsilon}_{\text{nucl}}^{\gamma}(t)=\sum_{j}f_{j}^{\gamma}(t)\frac{% \langle E_{j}^{\gamma}\rangle Y_{j}(t)}{\tau_{j}\,m_{p}}\,,over˙ start_ARG italic_ϵ end_ARG start_POSTSUBSCRIPT nucl end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_γ end_POSTSUPERSCRIPT ( italic_t ) = ∑ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT italic_f start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_γ end_POSTSUPERSCRIPT ( italic_t ) divide start_ARG ⟨ italic_E start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_γ end_POSTSUPERSCRIPT ⟩ italic_Y start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( italic_t ) end_ARG start_ARG italic_τ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT italic_m start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT end_ARG , (3)

with j𝑗jitalic_j isotope index and mpsubscript𝑚𝑝m_{p}italic_m start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT proton mass. The average lifetimes τjsubscript𝜏𝑗\tau_{j}italic_τ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT and the mean energy Ejγdelimited-⟨⟩superscriptsubscript𝐸𝑗𝛾\langle E_{j}^{\gamma}\rangle⟨ italic_E start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_γ end_POSTSUPERSCRIPT ⟩ released by each of these nuclei in γ𝛾\gammaitalic_γ-rays via a radioactive decay are taken from the ENDF/B-VIII.0 database444 https://www-nds.iaea.org/exfor/endf.htm (Brown et al., 2018). The thermalization factor fjγ(t)superscriptsubscript𝑓𝑗𝛾𝑡f_{j}^{\gamma}(t)italic_f start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_γ end_POSTSUPERSCRIPT ( italic_t ) is calculated, as in Hotokezaka & Nakar (2019); Combi & Siegel (2023a), starting from the detailed composition of the ejecta and the same effective opacity tables from the NIST-XCOM catalogue555 https://www.nist.gov/pml/xcom-photon-cross-sections-database (Berger et al., 2010) used in Barnes et al. (2016). For the energy injected in electrons and α𝛼\alphaitalic_α particles and their thermalization factors, we use the analytic expressions of Kasen & Barnes (2019). Neutrinos from β𝛽\betaitalic_β-decays do not deposit energy into the system, while fission fragments and daughter nuclei tend to thermalize very efficiently (Barnes et al., 2016). We approximate this behavior by assuming that half of the remaining (non-thermalized) heating rate from SkyNet is deposited into the ejecta.

2.4 Jet Energy Deposition

We consider a jet model following the “thermal bomb” prescription of Morozova et al. (2015). Essentially, an extra energy term is added to the RHS of Eq. (1) for the innermost shells of the ejecta during a chosen time interval. The thermal bomb parameters are fixed assuming that the released isotropic energy from the sGRB is proportional to the kinetic energy of the structured jet described by Ghirlanda et al. (2019), Eiso(θ)=E0/[1+(θ/θj)5.5]subscript𝐸iso𝜃subscript𝐸0delimited-[]1superscript𝜃subscript𝜃𝑗5.5E_{\text{iso}}(\theta)=E_{0}/[1+(\theta/\theta_{j})^{5.5}]italic_E start_POSTSUBSCRIPT iso end_POSTSUBSCRIPT ( italic_θ ) = italic_E start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT / [ 1 + ( italic_θ / italic_θ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 5.5 end_POSTSUPERSCRIPT ], with θjsubscript𝜃𝑗\theta_{j}italic_θ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT jet’s opening angle. The total energy Ejsubscript𝐸𝑗E_{j}italic_E start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT released by the jet is related to Eisosubscript𝐸isoE_{\text{iso}}italic_E start_POSTSUBSCRIPT iso end_POSTSUBSCRIPT by Eiso=(4π/ΔΩj)Ejsubscript𝐸iso4𝜋ΔsubscriptΩ𝑗subscript𝐸𝑗E_{\text{iso}}=(4\pi/\Delta\Omega_{j})E_{j}italic_E start_POSTSUBSCRIPT iso end_POSTSUBSCRIPT = ( 4 italic_π / roman_Δ roman_Ω start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ) italic_E start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT, where ΔΩj=2π(1cosθj)ΔsubscriptΩ𝑗2𝜋1subscript𝜃𝑗\Delta\Omega_{j}=2\pi(1-\cos\theta_{j})roman_Δ roman_Ω start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT = 2 italic_π ( 1 - roman_cos italic_θ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ) is the solid angle covered by the jet. We fix θj=15subscript𝜃𝑗15\theta_{j}=15italic_θ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT = 15 degrees, within the range also explored in Hamidani et al. (2020) for the sGRB associated with GW170817. We simulate a jet with the same duration as the observed sGRB from Abbott et al. (2017c), Δtj=100 msΔsubscript𝑡𝑗100 ms\Delta t_{j}=100\textrm{ ms}roman_Δ italic_t start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT = 100 ms. We launch the jet at t=200 ms𝑡200 mst=200\textrm{ ms}italic_t = 200 ms and fix the parameter E0=1051 ergsubscript𝐸0superscript1051 ergE_{0}=10^{51}\textrm{ erg}italic_E start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 10 start_POSTSUPERSCRIPT 51 end_POSTSUPERSCRIPT erg, to be compatible with the isotropic luminosity range given by Hamidani et al. (2020). Their Figure 9 shows that, in the case of GRB 170817A, choosing the isotropic luminosity to be Liso1052 erg/ssimilar-tosubscript𝐿isosuperscript1052 erg/sL_{\text{iso}}\sim 10^{52}\textrm{ erg/s}italic_L start_POSTSUBSCRIPT iso end_POSTSUBSCRIPT ∼ 10 start_POSTSUPERSCRIPT 52 end_POSTSUPERSCRIPT erg/s constrains the jet to be launched no more than 300 ms300 ms300\textrm{ ms}300 ms after merger.

3 Results

3.1 Fluid Elements Composition Evolution

Our simulation allows to self-consistently monitor the nuclear composition of the matter during its expansion. Figure 1 shows the time evolution of the mass fraction of protons, neutrons, rare earths and r𝑟ritalic_r-process peak elements (panels (a𝑎aitalic_a-c𝑐citalic_c) and (hhitalic_h-l𝑙litalic_l)), of the parameter h(t)𝑡h(t)italic_h ( italic_t ) (see below, panels (d𝑑ditalic_d) and (m𝑚mitalic_m)) and of the electron fraction (panels (e𝑒eitalic_e) and (n𝑛nitalic_n)), together with the initial thermodynamic conditions of the ejecta (panels (f𝑓fitalic_f,g𝑔gitalic_g) and (o𝑜oitalic_o,p𝑝pitalic_p)), as a function of the Lagrangian mass coordinate. The left and right panels refer to two different angular sections, θ=12, 90𝜃1290\theta=12,\,90italic_θ = 12 , 90 degrees. Mass shells with, respectively, m50×106Mless-than-or-similar-to𝑚50superscript106subscriptMdirect-productm\lesssim 50\times 10^{-6}~{}{{\rm M_{\odot}}}italic_m ≲ 50 × 10 start_POSTSUPERSCRIPT - 6 end_POSTSUPERSCRIPT roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT and m1.4×103Mless-than-or-similar-to𝑚1.4superscript103subscriptMdirect-productm\lesssim 1.4\times 10^{-3}~{}{{\rm M_{\odot}}}italic_m ≲ 1.4 × 10 start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT correspond to the wind, while larger values to the dynamical ejecta. In the latter, the entropy slightly increases towards the outer shells, where a high entropy, low mass tail is observed at all angles. Material with low initial electron fractions, Ye,0<0.22subscript𝑌𝑒00.22Y_{e,0}<0.22italic_Y start_POSTSUBSCRIPT italic_e , 0 end_POSTSUBSCRIPT < 0.22, is found in the outer regions of the ejecta and, for m2.5×103Msimilar-to𝑚2.5superscript103subscriptMdirect-productm\sim 2.5\times 10^{-3}~{}{{\rm M_{\odot}}}italic_m ∼ 2.5 × 10 start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT, around the equatorial plane. During the evolution, the electron fraction rises to Ye>0.22subscript𝑌𝑒0.22Y_{e}>0.22italic_Y start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT > 0.22 for almost every mass shell at around t1 ssimilar-to𝑡1 st\sim 1\textrm{ s}italic_t ∼ 1 s due to the β𝛽\betaitalic_β-decays following the neutron captures and neutron freeze-out.

Refer to caption
Figure 1: Evolution of the ejecta composition for every mass shell of the angular sections θ12,90similar-to-or-equals𝜃1290\theta\simeq 12,90italic_θ ≃ 12 , 90 degrees (first and second column respectively). The x-axis indicates the Lagrangian mass coordinate, linear in the shell index. The vertical black, dotted lines indicate the point where the analytical disk wind is attached to the dynamical ejecta profile. The color maps in panels (a𝑎aitalic_a-c𝑐citalic_c) and (hhitalic_h-l𝑙litalic_l) represent the mass fractions of neutrons (in gray), protons (red), first (blue), second (orange), and third (purple) r-process peak elements, and rare earths (brown). Panels (d𝑑ditalic_d, m𝑚mitalic_m) and (e𝑒eitalic_e, n𝑛nitalic_n) show the evolution of the parameter h(t)𝑡h(t)italic_h ( italic_t ) defined in (4) (blue to red) and of the electron fraction (green to pink), respectively. In (e𝑒eitalic_e) and (n𝑛nitalic_n) we also plot in blue the amount of mass δm𝛿𝑚\delta mitalic_δ italic_m included in each shell, rescaled by the total mass of the angular section Mej(θ)subscript𝑀ej𝜃M_{\text{ej}}(\theta)italic_M start_POSTSUBSCRIPT ej end_POSTSUBSCRIPT ( italic_θ ), and in black the position of the photosphere. Subplots (f𝑓fitalic_f-g𝑔gitalic_g) and (o𝑜oitalic_o-p𝑝pitalic_p) show the initial values of Ye,s,T,τsubscript𝑌𝑒𝑠𝑇𝜏Y_{e},s,T,\tauitalic_Y start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT , italic_s , italic_T , italic_τ as a function of the included mass. The horizontal dotted lines indicate the limit values of Ye,0=0.22subscript𝑌𝑒00.22Y_{e,0}=0.22italic_Y start_POSTSUBSCRIPT italic_e , 0 end_POSTSUBSCRIPT = 0.22, s0=100kB/baryonsubscript𝑠0100subscript𝑘Bbaryons_{0}=100\,k_{\text{B}}/\text{baryon}italic_s start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 100 italic_k start_POSTSUBSCRIPT B end_POSTSUBSCRIPT / baryon, T0=7 GKsubscript𝑇07 GKT_{0}=7\text{ GK}italic_T start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 7 GK, and τb=5 mssubscript𝜏𝑏5 ms\tau_{b}=5\textrm{ ms}italic_τ start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT = 5 ms.

The color maps in panels (d𝑑ditalic_d) and (m𝑚mitalic_m) of Fig. 1 show the evolution of the parameter

h(t)log10(ρ(t)ρ𝚂𝚔𝚢𝙽𝚎𝚝(t|ρ0,τb)).𝑡subscript10𝜌𝑡subscript𝜌𝚂𝚔𝚢𝙽𝚎𝚝conditional𝑡subscript𝜌0subscript𝜏𝑏h(t)\equiv\log_{10}\left(\frac{\rho(t)}{\rho_{\tt SkyNet}(t\,|\,\rho_{0},\tau_% {b})}\right)\,.italic_h ( italic_t ) ≡ roman_log start_POSTSUBSCRIPT 10 end_POSTSUBSCRIPT ( divide start_ARG italic_ρ ( italic_t ) end_ARG start_ARG italic_ρ start_POSTSUBSCRIPT typewriter_SkyNet end_POSTSUBSCRIPT ( italic_t | italic_ρ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT , italic_τ start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT ) end_ARG ) . (4)

The density evolution for the analytical expansion ρ𝚂𝚔𝚢𝙽𝚎𝚝subscript𝜌𝚂𝚔𝚢𝙽𝚎𝚝\rho_{\tt SkyNet}italic_ρ start_POSTSUBSCRIPT typewriter_SkyNet end_POSTSUBSCRIPT is parametrized by the initial density ρ0subscript𝜌0\rho_{0}italic_ρ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT and expansion timescale τbsubscript𝜏𝑏\tau_{b}italic_τ start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT as defined in Eq. (1) of Lippuner & Roberts (2015). Here, τbsubscript𝜏𝑏\tau_{b}italic_τ start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT is calculated from the expansion timescale measured at the extraction radius as prescribed in (Radice et al., 2016). The same homologous expansion is used by Perego et al. (2022); Wu et al. (2022) for the time-domain fitting of the heating rates. Positive (negative) values of hhitalic_h indicate regions of the tM𝑡𝑀t-Mitalic_t - italic_M plane where the mass shells are characterized by higher (lower) densities than the one predicted by the simple analytical model. Shells where hhitalic_h is constant in time are expanding homologously. This happens for most of the shells in our simulation for t100 sgreater-than-or-equivalent-to𝑡100 st\gtrsim 100\textrm{ s}italic_t ≳ 100 s, in agreement with the results of Rosswog et al. (2014). If h00h\neq 0italic_h ≠ 0, the evolution happens on a curve ρ𝚂𝚔𝚢𝙽𝚎𝚝(t|ρ0,τb)subscript𝜌𝚂𝚔𝚢𝙽𝚎𝚝conditional𝑡subscriptsuperscript𝜌0subscriptsuperscript𝜏𝑏\rho_{\tt SkyNet}(t\,|\,\rho^{\prime}_{0},\tau^{\prime}_{b})italic_ρ start_POSTSUBSCRIPT typewriter_SkyNet end_POSTSUBSCRIPT ( italic_t | italic_ρ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT , italic_τ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT ) different from the one associated with the values of ρ0,τbsubscript𝜌0subscript𝜏𝑏\rho_{0},\tau_{b}italic_ρ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT , italic_τ start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT of the considered mass shell. We identify a deviation of approximately two to three orders of magnitude between our numerical solution and the specific analytical expansion ρ𝚂𝚔𝚢𝙽𝚎𝚝(t|ρ0,τb)subscript𝜌𝚂𝚔𝚢𝙽𝚎𝚝conditional𝑡subscript𝜌0subscript𝜏𝑏\rho_{\tt SkyNet}(t\,|\,\rho_{0},\tau_{b})italic_ρ start_POSTSUBSCRIPT typewriter_SkyNet end_POSTSUBSCRIPT ( italic_t | italic_ρ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT , italic_τ start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT ) in all the significative timescales. As further explained below, such a different expansion in the early evolution (t1less-than-or-similar-to𝑡1t\lesssim 1italic_t ≲ 1 s, when the production of heavy nuclei is still ongoing) can lead to inconsistent nucleosynthesis predictions. In Appendix A we discuss an example of how a full evolution of a fluid element can differ from the one obtained by an independent NN assuming ρ𝚂𝚔𝚢𝙽𝚎𝚝(t|ρ0,τb)subscript𝜌𝚂𝚔𝚢𝙽𝚎𝚝conditional𝑡subscript𝜌0subscript𝜏𝑏\rho_{\tt SkyNet}(t\,|\,\rho_{0},\tau_{b})italic_ρ start_POSTSUBSCRIPT typewriter_SkyNet end_POSTSUBSCRIPT ( italic_t | italic_ρ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT , italic_τ start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT ). A comparison with the results from a run using the fitted heating rates of Wu et al. (2022) instead of the data from the coupled NN reveals that the disagreement between the predicted density evolution and the simulated one builds up mostly because of hydrodynamic effects. The relative differences in the nuclear heating are not big enough to impact the hydrodynamic evolution while competing with the pressure work in Eq. (1) (see Appendix B).

At t100 mssimilar-to𝑡100 mst\sim 100\textrm{ ms}italic_t ∼ 100 ms, a sudden increase of hhitalic_h is observed in the disk ejecta. The analysis of the released nuclear energy reveals a correspondent absorption of energy from the fluid to the NN (ϵ˙nucl<0subscript˙italic-ϵnucl0\dot{\epsilon}_{\text{nucl}}<0over˙ start_ARG italic_ϵ end_ARG start_POSTSUBSCRIPT nucl end_POSTSUBSCRIPT < 0). In the same region, the abrupt change in the hhitalic_h parameter reveals ongoing shocks. A series of shock-induced photodissociation processes could explain the negative heating rates, in analogy with supernovae simulations, see e.g. (Janka, 2012; Burrows, 2013). Such kind of effect is genuinely due to the coupling between the NN and hydrodynamics, and it cannot be reproduced by offline analyses. In the polar regions, a further compression at t200 mssimilar-to𝑡200 mst\sim 200\textrm{ ms}italic_t ∼ 200 ms, followed by a more rapid expansion, is caused by the extra energy deposited by the jet.

Panels (a𝑎aitalic_a-c𝑐citalic_c) and (hhitalic_h-l𝑙litalic_l) of Fig. 1 show the evolution of the cumulative mass fractions of groups of selected isotopes. Consistently throughout the ejecta, shells with Ye,00.22less-than-or-similar-tosubscript𝑌𝑒00.22Y_{e,0}\lesssim 0.22italic_Y start_POSTSUBSCRIPT italic_e , 0 end_POSTSUBSCRIPT ≲ 0.22 produce first peak elements (embedded in a sea of free neutrons) within t104 sless-than-or-similar-to𝑡superscript104 st\lesssim 10^{-4}\textrm{ s}italic_t ≲ 10 start_POSTSUPERSCRIPT - 4 end_POSTSUPERSCRIPT s. On a time scale of a second, these nuclei are converted to second and third peak elements and a small fraction of rare earths. At T07 GKgreater-than-or-equivalent-tosubscript𝑇07 GKT_{0}\gtrsim 7\text{ GK}italic_T start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ≳ 7 GK, most of the ejecta is initially composed of free neutrons and protons (their ratio depending on the initial electron fraction) and light elements, which will act as seeds for r𝑟ritalic_r-process nucleosynthesis. In regions where T07 GKless-than-or-similar-tosubscript𝑇07 GKT_{0}\lesssim 7\text{ GK}italic_T start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ≲ 7 GK, the initial temperature is low enough in certain mass shells (see the arrows in panel (i)) to make the binding energy term in Eq. (2) strongly affect the NSE composition. Hence, a significant fraction of first peak elements is already formed at t=0𝑡0t=0italic_t = 0 and can be directly used as r-process seeds. This can ease the heavy elements production and, in some cases, partially induce strong r𝑟ritalic_r-processes even for Ye0.22greater-than-or-equivalent-tosubscript𝑌𝑒0.22Y_{e}\gtrsim 0.22italic_Y start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT ≳ 0.22 (see for example the high Ye,0subscript𝑌𝑒0Y_{e,0}italic_Y start_POSTSUBSCRIPT italic_e , 0 end_POSTSUBSCRIPT, low s0subscript𝑠0s_{0}italic_s start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT third peak production in the θ=90𝜃90\theta=90italic_θ = 90 degrees case).

High enough values of the initial entropy (s0100kB/baryongreater-than-or-equivalent-tosubscript𝑠0100subscript𝑘Bbaryons_{0}\gtrsim 100\,k_{\text{B}}/\text{baryon}italic_s start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ≳ 100 italic_k start_POSTSUBSCRIPT B end_POSTSUBSCRIPT / baryon) can lead to the production of rare earths, lanthanides and actinides even for intermediate values of the initial electron fraction, 0.22Ye0.35less-than-or-similar-to0.22subscript𝑌𝑒less-than-or-similar-to0.350.22\lesssim Y_{e}\lesssim 0.350.22 ≲ italic_Y start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT ≲ 0.35. This outcome is compatible with an α𝛼\alphaitalic_α-rich freeze-out, characterized by a large neutrons to seed ratio (Just et al., 2015; Cowan et al., 2021). However, not all shells with high initial entropy are sites of strong r𝑟ritalic_r-process nucleosynthesis. This is clearly shown, for example, by the θ=12𝜃12\theta=12italic_θ = 12 degree angular section in the region 74.07m/(106M)74.60less-than-or-similar-to74.07𝑚superscript106subscriptMdirect-productless-than-or-similar-to74.6074.07\lesssim m/(10^{-6}~{}{{\rm M_{\odot}}})\lesssim 74.6074.07 ≲ italic_m / ( 10 start_POSTSUPERSCRIPT - 6 end_POSTSUPERSCRIPT roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT ) ≲ 74.60, see panels (a𝑎aitalic_a) and (b𝑏bitalic_b). When strong r𝑟ritalic_r-processes are activated, free neutrons are efficiently consumed, reducing the final abundance of free protons. Starting from m74.07×106Msimilar-to𝑚74.07superscript106subscript𝑀direct-productm\sim 74.07\times 10^{-6}M_{\odot}italic_m ∼ 74.07 × 10 start_POSTSUPERSCRIPT - 6 end_POSTSUPERSCRIPT italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT and moving outward in the ejecta, while the initial T0subscript𝑇0T_{0}italic_T start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT, s0subscript𝑠0s_{0}italic_s start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT and τbsubscript𝜏𝑏\tau_{b}italic_τ start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT profiles are almost constant (panels (f𝑓fitalic_f) and (g𝑔gitalic_g)), strong r𝑟ritalic_r-processes get progressively suppressed. The suppression is smooth even given the jump in the initial conditions around m74.60×106Msimilar-to𝑚74.60superscript106subscript𝑀direct-productm\sim 74.60\times 10^{-6}M_{\odot}italic_m ∼ 74.60 × 10 start_POSTSUPERSCRIPT - 6 end_POSTSUPERSCRIPT italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT, which also does not change the second peak production in panel (c𝑐citalic_c). This discussion gives an example of how only the detailed evolution of the local thermo- and hydrodynamics conditions determine with accuracy the subsequent nucleosynthesis yields.

For low entropies, s0100kB/baryonless-than-or-similar-tosubscript𝑠0100subscript𝑘Bbaryons_{0}\lesssim 100\,k_{\text{B}}/\text{baryon}italic_s start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ≲ 100 italic_k start_POSTSUBSCRIPT B end_POSTSUBSCRIPT / baryon, and intermediate values of the initial electron fractions, 0.22Ye,00.35less-than-or-similar-to0.22subscript𝑌𝑒0less-than-or-similar-to0.350.22\lesssim Y_{e,0}\lesssim 0.350.22 ≲ italic_Y start_POSTSUBSCRIPT italic_e , 0 end_POSTSUBSCRIPT ≲ 0.35, the fewer free neutrons are rapidly captured by seed nuclei, and only second peak elements are produced at the end of the nucleosynthesis. Note again the exception represented by the shells at around m2.704×103Mgreater-than-or-equivalent-to𝑚2.704superscript103subscriptMdirect-productm\gtrsim 2.704\times 10^{-3}~{}{{\rm M_{\odot}}}italic_m ≳ 2.704 × 10 start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT roman_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT in the equatorial section. At the high Ye,00.35greater-than-or-equivalent-tosubscript𝑌𝑒00.35Y_{e,0}\gtrsim 0.35italic_Y start_POSTSUBSCRIPT italic_e , 0 end_POSTSUBSCRIPT ≳ 0.35 reached in the polar section, weak r𝑟ritalic_r-processes occur at low entropy and only first peak elements can be produced.

Extremely high values of the initial entropy, s0200kB/baryongreater-than-or-equivalent-tosubscript𝑠0200subscript𝑘Bbaryons_{0}\gtrsim 200\,k_{\text{B}}/\text{baryon}italic_s start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ≳ 200 italic_k start_POSTSUBSCRIPT B end_POSTSUBSCRIPT / baryon, can be reached in the outermost shells of the ejecta. Despite the large neutrons to seed ratio, neutron capture is not effective in the fast ejecta tail (v0.6greater-than-or-equivalent-to𝑣0.6v\gtrsim 0.6italic_v ≳ 0.6). Nucleosynthesis is therefore hindered. In this rarefied environment, most of the neutrons remain free after the r𝑟ritalic_r-process freeze-out and start β𝛽\betaitalic_β-decaying on a timescale of t10similar-to𝑡10t\sim 10italic_t ∼ 10 minutes, while crossing the photosphere, whose position is tracked in panels (e𝑒eitalic_e) and (n𝑛nitalic_n). This could power a UV/blue kilonova precursor on the hours timescales, as discussed in Sec. 3.3. After 104 ssuperscript104 s10^{4}\text{ s}10 start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT s, no heavy element is produced, and free protons dominate the final matter composition. This is reflected in the high final electron fractions of these shells. If the strong rise in entropy is combined with an initial electron fraction Ye,00.35greater-than-or-equivalent-tosubscript𝑌𝑒00.35Y_{e,0}\gtrsim 0.35italic_Y start_POSTSUBSCRIPT italic_e , 0 end_POSTSUBSCRIPT ≳ 0.35, all neutrons are initially bound in stable first peak elements. In these outer regions, the composition remains essentially frozen. We can thus predict a negligible nuclear contribution to the kilonova light curves from these shells throughout the ejecta evolution, regardless of the photosphere position.

The qualitative changes in the nucleosynthesis patterns caused by the jet are negligible. The only quantitative effect observed on the mass fractions depicted in Fig. 1 is a very slight delay in the production of second r𝑟ritalic_r-process peak elements in the innermost shells of the polar sections. No significant effects arise in the regions near the equatorial plane, as the jet energy is negligible at low latitudes, and we do not account for the coupling between different angular sections.

In summary, our results indicate that it is not possible to predict the final nucleosynthesis yields based on the initial thermodynamic conditions of the ejecta solely.

3.2 Global Nucleosynthesis Yields

Refer to caption
Figure 2: Global r𝑟ritalic_r-process nucleosynthesis yields at t5×104 ssimilar-to-or-equals𝑡5superscript104 st\simeq 5\times 10^{4}\textrm{ s}italic_t ≃ 5 × 10 start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT s. Top panel: in blue and black, the final global abundances YA,fdelimited-⟨⟩subscript𝑌𝐴𝑓\langle Y_{A,f}\rangle⟨ italic_Y start_POSTSUBSCRIPT italic_A , italic_f end_POSTSUBSCRIPT ⟩ as a function of the mass number A𝐴Aitalic_A, obtained respectively from our SNEC simulation and with a post-process analysis similar to Perego et al. (2022). In orange and green, the results obtained from our SNEC simulation excluding the disk ejecta or including only the matter emitted at θ=12𝜃12\theta=12italic_θ = 12 degrees (at all times). Red dots represent the Solar r𝑟ritalic_r-abundances (Prantzos et al., 2020) scaled to match the average value of the second peak. The histogram shows, for the complete simulation, the global cumulative mass fractions X(A)delimited-⟨⟩𝑋𝐴\langle X(A)\rangle⟨ italic_X ( italic_A ) ⟩ of selected groups of elements; in particular, first (blue), second (orange) and third (purple) r𝑟ritalic_r-process peaks and rare-earths (brown). We report the initial value of the average electron fraction of the ejecta in the top left corner of the plot. Bottom panel: in red, relative difference against the model with prescribed homologous expansion, see Eq. (5); in black and purple, analogous comparisons between our global results and the post-process analysis, and between the results of the complete simulation with and without the jet. The differences are only shown for mass numbers A^^𝐴\hat{A}over^ start_ARG italic_A end_ARG such that YA^,f>1010delimited-⟨⟩subscript𝑌^𝐴𝑓superscript1010\langle Y_{\hat{A},f}\rangle>10^{-10}⟨ italic_Y start_POSTSUBSCRIPT over^ start_ARG italic_A end_ARG , italic_f end_POSTSUBSCRIPT ⟩ > 10 start_POSTSUPERSCRIPT - 10 end_POSTSUPERSCRIPT.

We calculate the global abundances Yidelimited-⟨⟩subscript𝑌𝑖\langle Y_{i}\rangle⟨ italic_Y start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ⟩ as mass-weighted averages over all the mass shells of all the angular sections. In Fig. 2, we show the matter composition at t5×104 ssimilar-to-or-equals𝑡5superscript104 st\simeq 5\times 10^{4}\textrm{ s}italic_t ≃ 5 × 10 start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT s. Except for small modifications due to long lived α𝛼\alphaitalic_α-decaying isotopes with A220greater-than-or-equivalent-to𝐴220A\gtrsim 220italic_A ≳ 220, this is a good representation of the final r𝑟ritalic_r-process nucleosynthesis yields. All the r𝑟ritalic_r-process peaks and in part rare earths are produced. This is expected, as matter is mostly ejected along the equatorial plane, where the neutron-rich dynamical ejecta keeps the global, initial average electron fraction at Ye,00.24similar-to-or-equalsdelimited-⟨⟩subscript𝑌𝑒00.24\langle Y_{e,0}\rangle\simeq 0.24⟨ italic_Y start_POSTSUBSCRIPT italic_e , 0 end_POSTSUBSCRIPT ⟩ ≃ 0.24. The production of heavy elements is partially hindered in the polar sections, consistently with their higher Yesubscript𝑌𝑒Y_{e}italic_Y start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT and the picture outlined in Fig. 1. By comparing the results of our complete model with the ones obtained by excluding the late disk wind, we establish that the relative production of heavy elements is already saturated in the dynamical ejecta. Most of the lighter elements with A130less-than-or-similar-to𝐴130A\lesssim 130italic_A ≲ 130 (in particular elements between the first and second peaks and isotopes of the iron group) are relatively more produced by the disk wind. These results are in agreement with Martin et al. (2015); Cowan et al. (2021); Curtis et al. (2023); Chiesa et al. (2024).

Refer to caption
Figure 3: Evolution of the abundances of few selected isotopes and of the cumulative abundances of lanthanides and actinides. Panel (a𝑎aitalic_a) shows global results obtained with mass-weighted averages all over the ejecta. Panel (b𝑏bitalic_b) is restricted to an average over the innermost shell of all the angular sections in order to highlight the jet effects. Panel (c𝑐citalic_c) shows the ratio ΔYi=Yi/YippΔdelimited-⟨⟩subscript𝑌𝑖delimited-⟨⟩subscript𝑌𝑖delimited-⟨⟩superscriptsubscript𝑌𝑖pp\Delta\langle Y_{i}\rangle=\langle Y_{i}\rangle/\langle Y_{i}^{\text{pp}}\rangleroman_Δ ⟨ italic_Y start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ⟩ = ⟨ italic_Y start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ⟩ / ⟨ italic_Y start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT pp end_POSTSUPERSCRIPT ⟩ between our results and the prediction obtained with a post-process procedure similar to the one of Perego et al. (2022), see Appendix C for details. Panel (d𝑑ditalic_d) shows a similar comparison between the results obtained with and without the jet. The vertical dashed lines in panels (b𝑏bitalic_b) and (d𝑑ditalic_d) indicate the time when the jet is launched. In panels (c,d𝑐𝑑c,ditalic_c , italic_d), the results are only shown for times t^^𝑡\hat{t}over^ start_ARG italic_t end_ARG such that Yi(t^)>1012delimited-⟨⟩subscript𝑌𝑖^𝑡superscript1012\langle Y_{i}\rangle(\hat{t})>10^{-12}⟨ italic_Y start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ⟩ ( over^ start_ARG italic_t end_ARG ) > 10 start_POSTSUPERSCRIPT - 12 end_POSTSUPERSCRIPT.

In the bottom panel of Fig. 2 we show the final value of the relative difference

ΔYAch=|YAYAch|min(YA,YAch)Δdelimited-⟨⟩superscriptsubscript𝑌𝐴chdelimited-⟨⟩subscript𝑌𝐴delimited-⟨⟩superscriptsubscript𝑌𝐴chdelimited-⟨⟩subscript𝑌𝐴delimited-⟨⟩superscriptsubscript𝑌𝐴ch\Delta\langle Y_{A}^{\text{ch}}\rangle=\frac{\left|\langle Y_{A}\rangle-% \langle Y_{A}^{\text{ch}}\rangle\right|}{\min{\left(\langle Y_{A}\rangle,% \langle Y_{A}^{\text{ch}}\rangle\right)}}roman_Δ ⟨ italic_Y start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ch end_POSTSUPERSCRIPT ⟩ = divide start_ARG | ⟨ italic_Y start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT ⟩ - ⟨ italic_Y start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ch end_POSTSUPERSCRIPT ⟩ | end_ARG start_ARG roman_min ( ⟨ italic_Y start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT ⟩ , ⟨ italic_Y start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ch end_POSTSUPERSCRIPT ⟩ ) end_ARG (5)

between our results and the predictions YAchdelimited-⟨⟩superscriptsubscript𝑌𝐴ch\langle Y_{A}^{\text{ch}}\rangle⟨ italic_Y start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ch end_POSTSUPERSCRIPT ⟩ obtained with a SNEC simulation of the ejecta with coupled NN and a prescribed homologous expansion. We evolve the system with the full hydrodynamic equations up to t0=10subscript𝑡010t_{0}=10italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 10 ms, and then force the density to evolve as ρ(t,m)=ρ(t0,m)(t/t0)3𝜌𝑡𝑚𝜌subscript𝑡0𝑚superscript𝑡subscript𝑡03\rho(t,m)=\rho(t_{0},m)(t/t_{0})^{-3}italic_ρ ( italic_t , italic_m ) = italic_ρ ( italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT , italic_m ) ( italic_t / italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT. Neglecting the details of the hydrodynamic evolution introduces an error of 10%greater-than-or-equivalent-toabsentpercent10\gtrsim 10\%≳ 10 % for almost all values of mass numbers. The biggest discrepancies are observed for some light and second peak elements. We also compare our results against a post-process analysis similar to the method described in Perego et al. (2022), but adapted to our initialization method. The agreement is overall worse in the new case, and the position of the third r𝑟ritalic_r-process peak is shifted. The extra source of error is introduced here by effectively coarsening the resolution on the initial thermodynamic conditions (see Appendix C for details). Figure 3 shows the evolution of the abundances for free protons and neutrons and a few selected isotopes. The r𝑟ritalic_r-process nucleosynthesis takes place between t102 ssimilar-to𝑡superscript102 st\sim 10^{-2}\textrm{ s}italic_t ∼ 10 start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT s and t1 ssimilar-to𝑡1 st\sim 1\text{ s}italic_t ∼ 1 s, when the first drop in Yndelimited-⟨⟩subscript𝑌𝑛\langle Y_{n}\rangle⟨ italic_Y start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ⟩ occurs and the abundances of lanthanides and actinides saturate at Yi104similar-to-or-equalsdelimited-⟨⟩subscript𝑌𝑖superscript104\langle Y_{i}\rangle\simeq 10^{-4}⟨ italic_Y start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ⟩ ≃ 10 start_POSTSUPERSCRIPT - 4 end_POSTSUPERSCRIPT. The plateau at Yn104similar-to-or-equalssubscript𝑌𝑛superscript104Y_{n}\simeq 10^{-4}italic_Y start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ≃ 10 start_POSTSUPERSCRIPT - 4 end_POSTSUPERSCRIPT in panel (a𝑎aitalic_a) indicates the presence of regions of the ejecta where the material undergoes incomplete neutron burning. The second drop in the abundance of neutrons at t10similar-to𝑡10t\sim 10italic_t ∼ 10 min, connected with an increase in the proton fraction, is due to free neutrons β𝛽\betaitalic_β-decays. Our analysis shows qualitative features similar to those discussed in Perego et al. (2022). However, a direct comparison between the two models shows discrepancies of several orders of magnitude in the abundance evolution of various isotopes, see panel (c𝑐citalic_c). In particular, note the six orders of magnitude difference in the neutron abundance for the post-nucleosynthesis plateau, and the overproduction of 56Fe from the post-processing of about two and one orders of magnitude at timescales between 102103superscript102superscript10310^{-2}-10^{3}10 start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT - 10 start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT s and t104similar-to𝑡superscript104t\sim 10^{4}italic_t ∼ 10 start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT s, respectively. At late times, the post-processing has underproduced 4He, 2H and 3H, and free protons of about 1, 3, and 6 orders of magnitude.

The initial global abundance of 4He is of the order of 102superscript10210^{-2}10 start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT because of shells undergoing α𝛼\alphaitalic_α-rich freeze out, see panel (a𝑎aitalic_a). This isotope is partially consumed at early times to construct heavier elements, e.g. 88Sr. In the innermost shells, it is produced again during the r𝑟ritalic_r-process nucleosynthesis by α𝛼\alphaitalic_α-decays of the freshly produced heavy, neutron-rich elements, see panel (b𝑏bitalic_b). These regions are crossed by the jet, which produces a rapid increase in the abundances of protons and deuterium at t200 mssimilar-to-or-equals𝑡200 mst\simeq 200\textrm{ ms}italic_t ≃ 200 ms. 88Sr is initially strongly suppressed, but it rises enough at early times to effectively act as a seed for r𝑟ritalic_r-process nucleosynthesis, see panel (a𝑎aitalic_a). At late times this element is produced again as part of the first peak by β𝛽\betaitalic_β-decaying neutron-rich isotopes. This is particularly evident in panel (b𝑏bitalic_b). We also show the abundance of 56Fe as representative of the most bounded nuclei around the iron peak. Its initial formation is favored by its high binding energy. Like strontium, 56Fe acts as a seed nucleus, but it is only partially produced again by later nuclear reactions.

The additional energy released by the jet significantly affects only a fraction of the inner shells at the highest latitudes. The mass involved is a too small fraction of the ejecta to yield visible effects on the global results presented in Fig. 2 and 3. In particular, the final nuclear yields are compatible to few percent between the complete simulations with and without the jet (see the bottom panel of Fig. 2). Nevertheless, the examination of the nucleosynthesis within the inner shells reveals some modifications of the final abundances for the selected elements of Fig. 3, see panel (d𝑑ditalic_d). The sudden acceleration induced by the jet slows neutron captures down, therefore leaving, at the end of the nucleosynthesis, fewer lanthanides, more free β𝛽\betaitalic_β-decaying neutrons and, consequently, free protons. The faster dynamics also inhibits the burning of light elements, leading to an increase in the abundance of hydrogen isotopes at the end of the simulation.

3.3 Light curves

Refer to caption
Figure 4: AB apparent magnitudes predicted by our simulation for the Gemini bands u𝑢uitalic_u, g𝑔gitalic_g, i𝑖iitalic_i and Kssubscript𝐾𝑠K_{s}italic_K start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT for an observation angle of θ=20𝜃20\theta=20italic_θ = 20 degrees and a distance of 40404040 Mpc. Left panel: results of the complete simulation (solid lines) against the predictions obtained by excluding the jet (dashed lines) or the disk component (dash-dotted lines). Central panel: isotropized θ=12,90𝜃1290\theta=12,90italic_θ = 12 , 90 degrees angular sections (solid and dashed lines respectively). Right panel: light curves obtained assuming a constant thermalization factor fth=0.5subscript𝑓th0.5f_{\text{th}}=0.5italic_f start_POSTSUBSCRIPT th end_POSTSUBSCRIPT = 0.5 with (solid lines) or without (with the same heating rate fits of Wu et al. (2022), dashed lines) a coupled NN.

In Fig. 4 we show the kilonova light curves predicted by our model for a few selected UV/visible/IR bands for an observation angle of θ=20𝜃20\theta=20italic_θ = 20 degrees and a distance of 40404040 Mpc.

The disk is predicted to contribute to the blue component of the kilonova due to its lower opacity, see e.g. (Metzger & Fernández, 2014; Siegel, 2019; Curtis et al., 2024). However, at all angles, the photosphere enters the disk region only at late times (t10similar-to𝑡10t\sim 10italic_t ∼ 10 days), when the temperature drop has already made the disk spectra red. Essentially, the dynamical ejecta acts as a curtain, preventing the disk from significantly boosting the blue part of the kilonova. Correspondingly, in the left panel of Fig. 4, all the curves show an analogous decrease in their luminosity, when the disk is excluded, only due to the removal of part of the ejected mass. The central panel of Fig. 4 shows that, around the equator, the opaque lanthanide curtain associated with the dynamical ejecta produces a red kilonova. The blue components of the spectra become dominant at higher latitudes, suggesting a main contribution to the blue kilonova from the angular sections around the disk edge, where a still significative mass of relatively neutron-poor material is ejected.

To assess the impact of the online NN on the kilonova predictions, we plot in the right panel of Fig. 4 the results obtained without evolving the matter composition, but taking as input the heating rate fits employed already in Wu et al. (2022)666 Without the composition information given by the NN we cannot calculate the thermalization factors as described in Sec. 2.3. We therefore follow Wu et al. (2022) imposing a constant thermalization factor fth=0.5subscript𝑓th0.5f_{\text{th}}=0.5italic_f start_POSTSUBSCRIPT th end_POSTSUBSCRIPT = 0.5 and compare the results against a complete simulation with a coupled NN but with the same simple thermalization.. Around t4×102similar-to𝑡4superscript102t\sim 4\times 10^{-2}italic_t ∼ 4 × 10 start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT days, the simulation with a coupled NN shows a bump in the blue light curves, compatible with the UV precursor predicted in Metzger et al. (2015); Combi & Siegel (2023a) to be powered by β𝛽\betaitalic_β-decaying free neutrons. This feature is not reproduced by the simulation with no coupled NN. The reason is that the fits of Wu et al. (2022) do not capture the free neutron contribution to the heating rate (see their Fig. 2 at t102similar-to𝑡superscript102t\sim 10^{-2}italic_t ∼ 10 start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT days). Most importantly, the simulation without a coupled NN fails in predicting the position and brightness of the main peak of the high frequency filters. It predicts instead a faster dimming at late times of the low frequency ones. Thus, even if a coupled NN does have negligible impact on the hydrodynamic evolution, it introduces significative corrections in the predicted light curves. While the heating rate competes with the pdV work and has thus a negligible effect on the hydrodynamics (see also Appendix B), it acts more directly on the luminosity of the optically thin regions above the photosphere. Moreover, relatively small changes in the temperature of the photosphere can lead to an observable shift on the frequency peak of its blackbody radiation.

The jet slightly increases the early (t0.11similar-to𝑡0.11t\sim 0.1-1italic_t ∼ 0.1 - 1 days) light curves in all bands. The effect is stronger and can be seen earlier for higher frequencies (see the left panel of Fig. 4). This is due to the extra energy input increasing the temperature and thus leading to a bluer and brighter emission. A similar effect is discussed in Nativi et al. (2020), where the jet is found to clear a significant fraction of the lanthanide curtain.

4 Conclusions

In this work we investigated the nucleosynthesis process in a BNSM ejecta by means of a 2-dimensional, ray-by-ray radiation-hydrodynamics simulation incorporating an online NN.

Our results challenge the widely used approach of predicting r𝑟ritalic_r-process nucleosynthesis from isolated fluid elements. Figure 1 illustrates that it is not possible to fully predict the detailed distribution of the (final) nucleosynthesis yields solely from the initial thermodynamic conditions of a set of fluid elements. Moreover, the most commonly employed approximations for homologous expansions fail to capture the asymptotic hydrodynamic evolution of the unbound material. Therefore, even estimates for the nuclear heating rate obtained by post-processing isolated fluid elements, see e.g. (Rosswog et al., 2014; Wanajo, 2018; Wu et al., 2022; Rosswog & Korobkin, 2024; Collins et al., 2023; Shingles et al., 2023; Ricigliano et al., 2024), may lead to inconsistent results. To get a consistent picture of the ongoing nuclear processes, the effects related to radiation transfer and ejecta self-interaction must be taken into account. On the other hand, the corrections introduced by an online NN on the nuclear energy supplied or removed to the expanding material do not introduce significative changes in the hydrodynamic evolution of the ejecta. However, nuclear reactions have a more direct impact on their EM emission (see Fig. 4). Providing a radiation-hydrodynamic simulation of the ejecta from compact binary mergers with an online NN thus appears crucial also for improving the realism of the predicted kilonova light curves.

Some of the outer layers of the ejecta primarily consist of β𝛽\betaitalic_β-decaying free neutrons (see in Fig. 1), in agreement with de Jesús Mendoza-Temis et al. (2015); Radice et al. (2018a). In our simulations, having a very rapidly (τ5less-than-or-similar-to𝜏5\tau\lesssim 5italic_τ ≲ 5 ms) expanding ejecta is not sufficient to prevent the production of third-peak elements. Very high entropies, s100kB/baryongreater-than-or-equivalent-to𝑠100subscript𝑘Bbaryons\gtrsim 100\,k_{\text{B}}/\text{baryon}italic_s ≳ 100 italic_k start_POSTSUBSCRIPT B end_POSTSUBSCRIPT / baryon, are also required to leave a strong abundance of free neutrons after the r𝑟ritalic_r-process freeze-out.

From a qualitative point of view, our averaged results for nucleosynthesis yields in Fig. 2 align with other predictions in the literature, e.g. (Cowan et al., 2021; Perego et al., 2021). Our analysis reveals a connection between the production of light (below second peak) and third-peak elements with late disk and dynamical ejecta respectively, in agreement with Martin et al. (2015); Cowan et al. (2021). Figure 4 shows that the outer lanthanides curtain shields the radiation produced by the inner disk ejecta, thus preventing them to give the significant contribution to the blue component of the kilonova predicted by Metzger & Fernández (2014); Siegel (2019); Curtis et al. (2024). The polar part of our ejecta evolves from higher Yesubscript𝑌𝑒Y_{e}italic_Y start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT and produces few heavy r𝑟ritalic_r-process elements; this is consistent with the link discussed in Kasen et al. (2017); Tanaka et al. (2017); Nicholl et al. (2017); Perego et al. (2017) between the blue kilonova component and these sections of the unbound material.

We compared the final results for the nucleosynthesis yields with a SNEC run where a homologous expansion is prescribed from t=10𝑡10t=10italic_t = 10 ms. Neglecting the details of the radiation-hydrodynamics evolution of the ejecta introduces deviations 10%greater-than-or-equivalent-toabsentpercent10\gtrsim 10\%≳ 10 % for all the mass numbers (see the bottom panel of Fig. 2). In particular, the final abundances of some light (A50less-than-or-similar-to𝐴50A\lesssim 50italic_A ≲ 50) and second r𝑟ritalic_r-process peak elements differ of almost one order of magnitude. We also compared the time evolution of the nuclear abundances with a post-processing method analogous to a previous work that utilized the same NN (Perego et al., 2022). We found significant quantitative differences for some of the analyzed elements (a qualitative agreement is found only for some specific shells in our simulation). In particular, we found discrepancies of several orders of magnitude in the evolution of 4He, 56Fe, deuterium, tritium, and free protons, see Fig. 3. The differences are due both to the coupling of the NN with the radiation-hydrodynamic evolution and to the grid introduced by the post-processing method on the initial thermodynamic conditions.

The inclusion of an extra energy term mimicking a jet only affects the dynamic evolution of the polar regions of the ejecta, without altering the global, qualitative predictions of the r𝑟ritalic_r-process nucleosynthesis.

Our work will be extended and improved in several directions. We aim at implementing an improved thermalization of charged particles and a more realistic opacity treatment based on the complete, tracked information about matter composition. A following paper will report a systematic investigation of kilonova light curves using hundreds-of-milliseconds long numerical-relativity profiles and those improved thermalization and opacity models. Finally, the methods developed here will be ported to 3-dimensional simulations.

FM acknowledges support from the Deutsche Forschungsgemeinschaft (DFG) under Grant No. 406116891 within the Research Training Group RTG 2522/1. FM also acknowledges ECT* for organizing the MICRA2023 workshop and TIFPA for their invitation to Trento in December 2023. Both occasions were sites of useful discussions. SB acknowledges funding from the EU Horizon under ERC Consolidator Grant, no. InspiReM-101043372 and from the Deutsche Forschungsgemeinschaft, DFG, project MEMI number BE 6301/2-1. The work of AP is partially funded by the European Union under NextGenerationEU, PRIN 2022 Prot. n. 2022KX2Z3B. DR acknowledges funding from the U.S. Department of Energy, Office of Science, Division of Nuclear Physics under Award Number(s) DE-SC0021177 and DE-SC0024388, and from the National Science Foundation under Grants No. PHY-2011725, PHY-2020275, PHY-2116686, and AST-2108467. Simulations were performed on the ARA and DRACO clusters at Friedrich Schiller University Jena, on the supercomputer SuperMUC-NG at the Leibniz- Rechenzentrum (LRZ, www.lrz.de) Munich, and on the national HPE Apollo Hawk at the High Performance Computing Center Stuttgart (HLRS). The ARA cluster is funded in part by DFG grants INST 275/334-1 FUGG and INST 275/363-1 FUGG, and ERC Starting Grant, grant agreement no. BinGraSp-714626. The authors acknowledge the Gauss Centre for Supercomputing e.V. (www.gauss-centre.eu) for funding this project by providing computing time on the GCS Supercomputer SuperMUC-NG at LRZ (allocations pn36ge and pn36jo). The authors acknowledge HLRS for funding this project by pro- viding access to the supercomputer HPE Apollo Hawk under the grant number INTRHYGUE/44215.

References

  • Abbott et al. (2017a) Abbott, B. P., et al. 2017a, Phys. Rev. Lett., 119, 161101, doi: 10.1103/PhysRevLett.119.161101
  • Abbott et al. (2017b) —. 2017b, Astrophys. J., 848, L12, doi: 10.3847/2041-8213/aa91c9
  • Abbott et al. (2017c) —. 2017c, Astrophys. J. Lett., 848, L13, doi: 10.3847/2041-8213/aa920c
  • Arcones & Thielemann (2023) Arcones, A., & Thielemann, F.-K. 2023, Astron. Astrophys. Rev., 31, 1, doi: 10.1007/s00159-022-00146-x
  • Barnes et al. (2016) Barnes, J., Kasen, D., Wu, M.-R., & Martinez-Pinedo, G. 2016, Astrophys. J., 829, 110, doi: 10.3847/0004-637X/829/2/110
  • Beloborodov (2003) Beloborodov, A. M. 2003, Astrophys. J., 588, 931, doi: 10.1086/374217
  • Berger et al. (2010) Berger, M., et al. 2010. http://physics.nist.gov/xcom
  • Bernuzzi (2020) Bernuzzi, S. 2020, Gen. Rel. Grav., 52, 108, doi: 10.1007/s10714-020-02752-5
  • Brown et al. (2018) Brown, D. A., et al. 2018, Nucl. Data Sheets, 148, 1, doi: 10.1016/j.nds.2018.02.001
  • Burrows (2013) Burrows, A. 2013, Rev. Mod. Phys., 85, 245, doi: 10.1103/RevModPhys.85.245
  • Camilletti et al. (2024) Camilletti, A., Perego, A., Guercilena, F. M., Bernuzzi, S., & Radice, D. 2024, Phys. Rev. D, 109, 063023, doi: 10.1103/PhysRevD.109.063023
  • Chiesa et al. (2024) Chiesa, L., Perego, A., & Guercilena, F. M. 2024, Astrophys. J. Lett., 962, L24, doi: 10.3847/2041-8213/ad236e
  • Collins et al. (2023) Collins, C. E., Bauswein, A., Sim, S. A., et al. 2023, Mon. Not. Roy. Astron. Soc., 521, 1858, doi: 10.1093/mnras/stad606
  • Combi & Siegel (2023a) Combi, L., & Siegel, D. M. 2023a, Astrophys. J., 944, 28, doi: 10.3847/1538-4357/acac29
  • Combi & Siegel (2023b) —. 2023b, Phys. Rev. Lett., 131, 231402, doi: 10.1103/PhysRevLett.131.231402
  • Coulter et al. (2017) Coulter, D. A., et al. 2017, Science, 358, 1556, doi: 10.1126/science.aap9811
  • Cowan et al. (2021) Cowan, J. J., Sneden, C., Lawler, J. E., et al. 2021, Rev. Mod. Phys., 93, 15002, doi: 10.1103/RevModPhys.93.015002
  • Curtis et al. (2024) Curtis, S., Bosch, P., Mösta, P., et al. 2024, Astrophys. J. Lett., 961, L26, doi: 10.3847/2041-8213/ad0fe1
  • Curtis et al. (2023) Curtis, S., Miller, J. M., Frohlich, C., et al. 2023, Astrophys. J. Lett., 945, L13, doi: 10.3847/2041-8213/acba16
  • Cyburt et al. (2010) Cyburt, R. H., Amthor, A. M., Ferguson, R., et al. 2010, Astrophys. J. Suppl., 189, 240, doi: 10.1088/0067-0049/189/1/240
  • de Jesús Mendoza-Temis et al. (2015) de Jesús Mendoza-Temis, J., Wu, M.-R., Martinez-Pinedo, G., et al. 2015, Phys. Rev., C92, 055805, doi: 10.1103/PhysRevC.92.055805
  • Eichler et al. (1989) Eichler, D., Livio, M., Piran, T., & Schramm, D. N. 1989, Nature, 340, 126, doi: 10.1038/340126a0
  • Fernández & Metzger (2013) Fernández, R., & Metzger, B. D. 2013, Mon. Not. Roy. Astron. Soc., 435, 502, doi: 10.1093/mnras/stt1312
  • Fernández et al. (2019) Fernández, R., Tchekhovskoy, A., Quataert, E., Foucart, F., & Kasen, D. 2019, Mon. Not. Roy. Astron. Soc., 482, 3373, doi: 10.1093/mnras/sty2932
  • Ghirlanda et al. (2019) Ghirlanda, G., et al. 2019, Science, 363, 968, doi: 10.1126/science.aau8815
  • Hamidani et al. (2020) Hamidani, H., Kiuchi, K., & Ioka, K. 2020, Mon. Not. Roy. Astron. Soc., 491, 3192, doi: 10.1093/mnras/stz3231
  • Hotokezaka & Nakar (2019) Hotokezaka, K., & Nakar, E. 2019, doi: 10.3847/1538-4357/ab6a98
  • Janiuk (2014) Janiuk, A. 2014, Astron. Astrophys., 568, A105, doi: 10.1051/0004-6361/201423822
  • Janka (2012) Janka, H.-T. 2012, Ann. Rev. Nucl. Part. Sci., 62, 407, doi: 10.1146/annurev-nucl-102711-094901
  • Just et al. (2015) Just, O., Bauswein, A., Pulpillo, R. A., Goriely, S., & Janka, H. T. 2015, Mon. Not. Roy. Astron. Soc., 448, 541, doi: 10.1093/mnras/stv009
  • Just et al. (2021) Just, O., Goriely, S., Janka, H.-T., Nagataki, S., & Bauswein, A. 2021, Mon. Not. Roy. Astron. Soc., 509, 1377, doi: 10.1093/mnras/stab2861
  • Just et al. (2023) Just, O., Vijayan, V., Xiong, Z., et al. 2023, Astrophys. J. Lett., 951, L12, doi: 10.3847/2041-8213/acdad2
  • Kasen & Barnes (2019) Kasen, D., & Barnes, J. 2019, Astrophys. J., 876, 128, doi: 10.3847/1538-4357/ab06c2
  • Kasen et al. (2017) Kasen, D., Metzger, B., Barnes, J., Quataert, E., & Ramirez-Ruiz, E. 2017, Nature, doi: 10.1038/nature24453
  • Kiuchi et al. (2023) Kiuchi, K., Fujibayashi, S., Hayashi, K., et al. 2023, Phys. Rev. Lett., 131, 011401, doi: 10.1103/PhysRevLett.131.011401
  • Kiuchi et al. (2024) Kiuchi, K., Reboul-Salze, A., Shibata, M., & Sekiguchi, Y. 2024, Nature Astron., 8, 298, doi: 10.1038/s41550-024-02194-y
  • Korobkin et al. (2012) Korobkin, O., Rosswog, S., Arcones, A., & Winteler, C. 2012, Mon. Not. Roy. Astron. Soc., 426, 1940, doi: 10.1111/j.1365-2966.2012.21859.x
  • Lippuner & Roberts (2015) Lippuner, J., & Roberts, L. F. 2015, Astrophys. J., 815, 82, doi: 10.1088/0004-637X/815/2/82
  • Lippuner & Roberts (2017) —. 2017, Astrophys. J. Suppl., 233, 18, doi: 10.3847/1538-4365/aa94cb
  • Martin et al. (2015) Martin, D., Perego, A., Arcones, A., et al. 2015, Astrophys. J., 813, 2, doi: 10.1088/0004-637X/813/1/2
  • Metzger et al. (2015) Metzger, B. D., Bauswein, A., Goriely, S., & Kasen, D. 2015, Mon. Not. Roy. Astron. Soc., 446, 1115, doi: 10.1093/mnras/stu2225
  • Metzger & Fernández (2014) Metzger, B. D., & Fernández, R. 2014, Mon.Not.Roy.Astron.Soc., 441, 3444, doi: 10.1093/mnras/stu802
  • Morozova et al. (2015) Morozova, V., Piro, A. L., Renzo, M., et al. 2015, Astrophys. J., 814, 63, doi: 10.1088/0004-637X/814/1/63
  • Musolino et al. (2024) Musolino, C., Duqué, R., & Rezzolla, L. 2024, Astrophys. J. Lett., 966, L31, doi: 10.3847/2041-8213/ad3bb3
  • Nativi et al. (2020) Nativi, L., Bulla, M., Rosswog, S., et al. 2020, Mon. Not. Roy. Astron. Soc., 500, 1772, doi: 10.1093/mnras/staa3337
  • Nedora et al. (2021) Nedora, V., Bernuzzi, S., Radice, D., et al. 2021, Astrophys. J., 906, 98, doi: 10.3847/1538-4357/abc9be
  • Nicholl et al. (2017) Nicholl, M., et al. 2017, Astrophys. J., 848, L18, doi: 10.3847/2041-8213/aa9029
  • Paczynski (1986) Paczynski, B. 1986, Astrophys. J., 308, L43
  • Perego et al. (2019) Perego, A., Bernuzzi, S., & Radice, D. 2019, Eur. Phys. J., A55, 124, doi: 10.1140/epja/i2019-12810-7
  • Perego et al. (2017) Perego, A., Radice, D., & Bernuzzi, S. 2017, Astrophys. J., 850, L37, doi: 10.3847/2041-8213/aa9ab9
  • Perego et al. (2021) Perego, A., Thielemann, F. K., & Cescutti, G. 2021, in Handbook of Gravitational Wave Astronomy, 1, doi: 10.1007/978-981-15-4702-7_13-1
  • Perego et al. (2022) Perego, A., et al. 2022, Astrophys. J., 925, 22, doi: 10.3847/1538-4357/ac3751
  • Prantzos et al. (2020) Prantzos, N., Abia, C., Cristallo, S., Limongi, M., & Chieffi, A. 2020, Mon. Not. Roy. Astron. Soc., 491, 1832, doi: 10.1093/mnras/stz3154
  • Radice & Bernuzzi (2023) Radice, D., & Bernuzzi, S. 2023, Astrophys. J., 959, 46, doi: 10.3847/1538-4357/ad0235
  • Radice & Bernuzzi (2024) —. 2024, J. Phys. Conf. Ser., 2742, 012009, doi: 10.1088/1742-6596/2742/1/012009
  • Radice et al. (2022) Radice, D., Bernuzzi, S., Perego, A., & Haas, R. 2022, Mon. Not. Roy. Astron. Soc., 512, 1499, doi: 10.1093/mnras/stac589
  • Radice et al. (2016) Radice, D., Galeazzi, F., Lippuner, J., et al. 2016, Mon. Not. Roy. Astron. Soc., 460, 3255, doi: 10.1093/mnras/stw1227
  • Radice et al. (2018a) Radice, D., Perego, A., Hotokezaka, K., et al. 2018a, Astrophys. J. Lett., 869, L35, doi: 10.3847/2041-8213/aaf053
  • Radice et al. (2018b) —. 2018b, Astrophys. J., 869, 130, doi: 10.3847/1538-4357/aaf054
  • Ricigliano et al. (2024) Ricigliano, G., Perego, A., Borhanian, S., et al. 2024, Mon. Not. Roy. Astron. Soc., 529, 647, doi: 10.1093/mnras/stae572
  • Roberts et al. (2017) Roberts, L. F., Lippuner, J., Duez, M. D., et al. 2017, Mon. Not. Roy. Astron. Soc., 464, 3907, doi: 10.1093/mnras/stw2622
  • Rosswog et al. (2017) Rosswog, S., Feindt, U., Korobkin, O., et al. 2017, Class. Quant. Grav., 34, 104001, doi: 10.1088/1361-6382/aa68a9
  • Rosswog & Korobkin (2024) Rosswog, S., & Korobkin, O. 2024, Annalen Phys., 536, 2200306, doi: 10.1002/andp.202200306
  • Rosswog et al. (2014) Rosswog, S., Korobkin, O., Arcones, A., Thielemann, F. K., & Piran, T. 2014, Mon. Not. Roy. Astron. Soc., 439, 744, doi: 10.1093/mnras/stt2502
  • Savchenko et al. (2017) Savchenko, V., et al. 2017, Astrophys. J., 848, L15, doi: 10.3847/2041-8213/aa8f94
  • Schianchi et al. (2024) Schianchi, F., Gieg, H., Nedora, V., et al. 2024, Phys. Rev. D, 109, 044012, doi: 10.1103/PhysRevD.109.044012
  • Shingles et al. (2023) Shingles, L. J., Collins, C. E., Vijayan, V., et al. 2023, Astrophys. J. Lett., 954, L41, doi: 10.3847/2041-8213/acf29a
  • Siegel (2019) Siegel, D. M. 2019, Eur. Phys. J. A, 55, 203, doi: 10.1140/epja/i2019-12888-9
  • Siegel & Metzger (2017) Siegel, D. M., & Metzger, B. D. 2017, Phys. Rev. Lett., 119, 231102, doi: 10.1103/PhysRevLett.119.231102
  • Sprouse et al. (2024) Sprouse, T. M., Lund, K. A., Miller, J. M., McLaughlin, G. C., & Mumpower, M. R. 2024, Astrophys. J., 962, 79, doi: 10.3847/1538-4357/ad1819
  • Tanaka et al. (2017) Tanaka, M., et al. 2017, Publ. Astron. Soc. Jap., doi: 10.1093/pasj/psx121
  • Tanvir et al. (2017) Tanvir, N. R., et al. 2017, Astrophys. J., 848, L27, doi: 10.3847/2041-8213/aa90b6
  • Timmes & Swesty (2000) Timmes, F. X., & Swesty, F. D. 2000, Astrophys. J. Suppl., 126, 501, doi: 10.1086/313304
  • Von Neumann & Richtmyer (1950) Von Neumann, J., & Richtmyer, R. D. 1950, Journal of Applied Physics, 21, 232, doi: 10.1063/1.1699639
  • Wanajo (2018) Wanajo, S. 2018, Astrophys. J., 868, 65, doi: 10.3847/1538-4357/aae0f2
  • Wanajo et al. (2014) Wanajo, S., Sekiguchi, Y., Nishimura, N., et al. 2014, Astrophys. J., 789, L39, doi: 10.1088/2041-8205/789/2/L39
  • Wu et al. (2022) Wu, Z., Ricigliano, G., Kashyap, R., Perego, A., & Radice, D. 2022, Mon. Not. Roy. Astron. Soc., 512, 328, doi: 10.1093/mnras/stac399
  • Zappa et al. (2023) Zappa, F., Bernuzzi, S., Radice, D., & Perego, A. 2023, Mon. Not. Roy. Astron. Soc., 520, 1481, doi: 10.1093/mnras/stad107
\twocolumngrid

Appendix A Hydrodynamic effects on NN

To get an idea of how the prescribed thermodynamic trajectory and the hydrodynamical one differ, we compare in Fig. 5 the density ρ𝜌\rhoitalic_ρ, temperature T𝑇Titalic_T and electron fraction Yesubscript𝑌𝑒Y_{e}italic_Y start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT evolutions for the innermost shell of the 12,90129012,9012 , 90 degrees angular sections of our SNEC profile. We find orders of magnitude discrepancies in ρ𝜌\rhoitalic_ρ and T𝑇Titalic_T, while Yesubscript𝑌𝑒Y_{e}italic_Y start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT is generally more compatible between the two evolutions. Note, in the left column, how the jet affects the more polar trajectory, reducing the density and causing a discontinuity in the derivatives of T(t)𝑇𝑡T(t)italic_T ( italic_t ) and Ye(t)subscript𝑌𝑒𝑡Y_{e}(t)italic_Y start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT ( italic_t ). Mind that Fig. 5 is only meant to work as an example, but is not representative of the general behavior of the ejecta. For instance, note in Fig. 1 how hhitalic_h settles at different final values (i.e. the expansion becomes homologous) at different times for different fluid elements. Better o worse agreement between the compared evolutions can be found looking at other mass shells and angular sections.

Refer to caption
Figure 5: Evolution of the density (black), temperature (orange) and electron fraction (green) in the innermost shell of the 12121212 degrees (left column) and 90909090 degrees (right column) angular sections of the ejecta as calculated from the complete simulation (solid lines) and an independent SkyNet run (dashed line). The bottom panels show the relative differences calculated with an expression analogous to Eq. (5). The evolution of the temperature from SkyNet is plotted until it is actively updated by the NN and used in the nuclear calculations. The vertical dashed line at t=200𝑡200t=200italic_t = 200 ms indicates the launch of the jet.

Appendix B NN effects on Hydrodynamics

Refer to caption
Figure 6: Top panel: evolution of total pressure work pdV (in blue) and nuclear heating ϵ=ϵ˙nucldtitalic-ϵsubscript˙italic-ϵnucl𝑑𝑡\epsilon=\dot{\epsilon}_{\text{nucl}}\,dtitalic_ϵ = over˙ start_ARG italic_ϵ end_ARG start_POSTSUBSCRIPT nucl end_POSTSUBSCRIPT italic_d italic_t (orange) averaged all over the ejecta and integrated over time from the SNEC runs with (solid lines) and without (dashed lines) a coupled NN and a constant thermalization factor fth=0.5subscript𝑓th0.5f_{\text{th}}=0.5italic_f start_POSTSUBSCRIPT th end_POSTSUBSCRIPT = 0.5. Bottom panel: difference in the nuclear heating between the two simulations normalized over pdV (black line) or ϵitalic-ϵ\epsilonitalic_ϵ (red line) from the simulation using the heating rate fits.

We discuss here why the introduction of an online NN in a hydrodynamic simulation of the ejecta can affect the kilonova predictions while having negligible effects on the hydrodynamic evolution. To do that, we run a SNEC simulation using the heating rate fits of Wu et al. (2022) instead of the information from a coupled NN. In Fig. 6, we compare the total pressure work done on the ejecta and released nuclear energy, as a function of time, against a SNEC run with an online NN. In both simulations we assume a constant thermalization factor fth=0.5subscript𝑓th0.5f_{\text{th}}=0.5italic_f start_POSTSUBSCRIPT th end_POSTSUBSCRIPT = 0.5. The same assumption is made in Wu et al. (2022), as the composition information needed by the detailed thermalization and coming from the NN are missing.

As discussed in Sec. 3.1, the hydrodynamical evolution of the density profile from our complete run does not agree with the considered common prescription for homologous expansion. The same conclusion holds if the run using the heating rate fits is used. The introduction of a coupled NN introduces global corrections in the heating rates of 10%similar-toabsentpercent10\sim 10\%∼ 10 %. If specific fluid elements are considered, these corrections can grow up to a factor 10101010. Anyway, as shown in Fig. 6, the global corrections are typically 𝒪(104)𝒪superscript104\mathcal{O}(10^{-4})caligraphic_O ( 10 start_POSTSUPERSCRIPT - 4 end_POSTSUPERSCRIPT ) the total pressure work done by the mass shells during their expansion. As a consequence, the effect of the online NN on their thermodynamic trajectories through Eq. (1) is negligible. However, the kilonova light curves are affected by the NN, as the heating rate changes act more directly on the properties of the photosphere and on the radiation emitted by the shells outside of it (see Sec. 3.3).

Appendix C Post-processing

Perego et al. (2022) compute the nucleosynthesis in post-processing using our same NN and imposing homologously expanding ejecta profiles. A grid on the initial entropy s0subscript𝑠0s_{0}italic_s start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT, expansion timescale τ0subscript𝜏0\tau_{0}italic_τ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT and electron fraction Ye,0subscript𝑌𝑒0Y_{e,0}italic_Y start_POSTSUBSCRIPT italic_e , 0 end_POSTSUBSCRIPT is introduced to have a series of ready-to-use tabulated results. To adapt their post-process procedure to our initialization method, we first distribute the mass shells of the initial ejecta profiles (the same we run with SNEC ​​, averaged over the azimuthal angle) on a grid based on the initial thermodynamic conditions. We take the same grid in s0subscript𝑠0s_{0}italic_s start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT, τ0subscript𝜏0\tau_{0}italic_τ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT and Ye,0subscript𝑌𝑒0Y_{e,0}italic_Y start_POSTSUBSCRIPT italic_e , 0 end_POSTSUBSCRIPT used in Perego et al. (2022) and introduce a fine grid (150 values) on the temperature T0subscript𝑇0T_{0}italic_T start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT to be consistent with our NN initialization procedure. We run then an independent instance of SkyNet on each of the grid points. To be consistent with the homology assumption of Radice et al. (2016), we base our grid on the value τ0=RE/v0subscript𝜏0subscript𝑅𝐸subscript𝑣0\tau_{0}=R_{E}/v_{0}italic_τ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = italic_R start_POSTSUBSCRIPT italic_E end_POSTSUBSCRIPT / italic_v start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT, with REsubscript𝑅𝐸R_{E}italic_R start_POSTSUBSCRIPT italic_E end_POSTSUBSCRIPT extraction radius and v0subscript𝑣0v_{0}italic_v start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT initial velocity, initialize each NN at t0=τ0subscript𝑡0subscript𝜏0t_{0}=\tau_{0}italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = italic_τ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT and T=T0𝑇subscript𝑇0T=T_{0}italic_T = italic_T start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT imposing NSE, and evolve it with the prescription ρ(t)=ρ0((t+τ0)/τ0)3𝜌𝑡subscript𝜌0superscript𝑡subscript𝜏0subscript𝜏03\rho(t)=\rho_{0}((t+\tau_{0})/\tau_{0})^{-3}italic_ρ ( italic_t ) = italic_ρ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( ( italic_t + italic_τ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) / italic_τ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT.

The introduction of the grid on s0subscript𝑠0s_{0}italic_s start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT, τ0subscript𝜏0\tau_{0}italic_τ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT, Ye,0subscript𝑌𝑒0Y_{e,0}italic_Y start_POSTSUBSCRIPT italic_e , 0 end_POSTSUBSCRIPT coarsens the resolution on the initial thermodynamic conditions, thus introducing another layer of approximation in the post-processing on top of the assumption of homologous expansion. As discussed in Sec. 3.2, this is also a relevant source of error (note anyway that a step that reduces the number of NN to run is crucial to define a post-process procedure able to give results in a reasonable amount of time).