(Translated by https://www.hiragana.jp/)
ZTF SN Ia DR2: The secondary maximum in Type Ia supernovae
11institutetext: School of Physics, Trinity College Dublin, College Green, Dublin 2, Ireland 22institutetext: GSI Helmholtzzentrum für Schwerionenforschung, Planckstraße 1, 64291 Darmstadt, Germany 33institutetext: Univ Lyon, Univ Claude Bernard Lyon 1, CNRS, IP2I Lyon/IN2P3, UMR 5822, F-69622, Villeurbanne, France 44institutetext: Department of Physics, Lancaster University, Lancaster LA1 4YB, UK 55institutetext: Oskar Klein Centre, Department of Physics, Stockholm University, SE-10691 Stockholm, Sweden 66institutetext: Institut für Physik, Humboldt-Universität zu Berlin, Newtonstr. 15, 12489 Berlin, Germany 77institutetext: National Research Council of Canada, Herzberg Astronomy & Astrophysics Research Centre, 5071 West Saanich Road, Victoria, BC V9E 2E7, Canada 88institutetext: Institute of Astronomy and Kavli Institute for Cosmology, University of Cambridge, Madingley Road, Cambridge CB3 0HA, UK 99institutetext: IPAC, California Institute of Technology, 1200 E. California Blvd, Pasadena, CA 91125, USA 1010institutetext: 247-17 Caltech, Pasadena, CA 91125, USA 1111institutetext: Institute of Space Sciences (ICE, CSIC), Campus UAB, Carrer de Can Magrans, s/n, E-08193 Barcelona, Spain 1212institutetext: Institut d’Estudis Espacials de Catalunya (IEEC), E-08034 Barcelona, Spain 1313institutetext: Lawrence Berkeley National Laboratory, 1 Cyclotron Road MS 50B-4206, Berkeley, CA, 94720, USA 1414institutetext: Department of Astronomy, University of California, Berkeley, 501 Campbell Hall, Berkeley, CA 94720, USA 1515institutetext: Department of Physics, Drexel University, Disque Hall, Office No. 808 32 S. 32nd St., Philadelphia, PA 19104 1616institutetext: Caltech Optical Observatories, California Institute of Technology, Pasadena, CA 91125, USA 1717institutetext: Aix Marseille Université, CNRS/IN2P3, CPPM, Marseille, France 1818institutetext: Oskar Klein Centre, Department of Astronomy, Stockholm University, SE-10691 Stockholm, Sweden

ZTF SN Ia DR2: The secondary maximum in Type Ia supernovae

M. Deckers 1, E-mail: deckersm@tcd.ie    K. Maguire1    L. Shingles2    G. Dimitriadis1    M. Rigault3    M. Smith4    A. Goobar5    J. Nordin6    J. Johansson5    M. Amenouche7    U. Burgaz1    S. Dhawan8    M. Ginolin3    L. Harvey1    W. D. Kenworthy5    Y. -L. Kim4    R. R. Laher9    N. Luo1    S. R. Kulkarni10    F. J. Masci9    T. E. Müller-Bravo11,12    P. E. Nugent13,14    N. Pletskova15    J. Purdum16    B. Racine17    J. Sollerman18    J. H. Terwel1
(Received 15 April 2024; accepted 26 June, 2024)

Type Ia supernova (SN Ia) light curves have a secondary maximum that exists in the r𝑟ritalic_r, i𝑖iitalic_i, and near-infrared filters. The secondary maximum is relatively weak in the r𝑟ritalic_r band, but holds the advantage that it is accessible, even at high redshift. We used Gaussian Process fitting to parameterise the light curves of 893 SNe Ia from the Zwicky Transient Facility’s (ZTF) second data release (DR2), and we were able to extract information about the timing and strength of the secondary maximum. We found >5σしぐまabsent5𝜎>5\sigma> 5 italic_σしぐま correlations between the light curve decline rate (Δでるたm15(g)Δでるたsubscript𝑚15𝑔\Delta m_{15}(g)roman_Δでるた italic_m start_POSTSUBSCRIPT 15 end_POSTSUBSCRIPT ( italic_g )) and the timing and strength of the secondary maximum in the r𝑟ritalic_r band. Whilst the timing of the secondary maximum in the i𝑖iitalic_i band also correlates with Δでるたm15(g)Δでるたsubscript𝑚15𝑔\Delta m_{15}(g)roman_Δでるた italic_m start_POSTSUBSCRIPT 15 end_POSTSUBSCRIPT ( italic_g ), the strength of the secondary maximum in the i𝑖iitalic_i band shows significant scatter as a function of Δでるたm15(g)Δでるたsubscript𝑚15𝑔\Delta m_{15}(g)roman_Δでるた italic_m start_POSTSUBSCRIPT 15 end_POSTSUBSCRIPT ( italic_g ). We found that the transparency timescales of 97 per cent of our sample are consistent with double detonation models, and that SNe Ia with small transparency timescales (<<< 32 d) reside predominantly in locally red environments. We measured the total ejected mass for the normal SNe Ia in our sample using two methods, and both were consistent with medians of 1.3± 0.3plus-or-minus1.30.31.3\ \pm\ 0.31.3 ± 0.3 and 1.2± 0.2plus-or-minus1.20.21.2\ \pm\ 0.21.2 ± 0.2 Msubscript𝑀direct-productM_{\odot}italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT. We find that the strength of the secondary maximum is a better standardisation parameter than the SALT light curve stretch (x1subscript𝑥1x_{1}italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT). Finally, we identified a spectral feature in the r𝑟ritalic_r band as Fe ii, which strengthens during the onset of the secondary maximum. The same feature begins to strengthen at <<< 3 d post maximum light in 91bg-like SNe. Finally, the correlation between x1subscript𝑥1x_{1}italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT and the strength of the secondary maximum was best fit with a broken line, with a split at x10=0.5± 0.2superscriptsubscript𝑥10plus-or-minus0.50.2x_{1}^{0}\ =\ -0.5\ \pm\ 0.2italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT = - 0.5 ± 0.2, suggestive of the existence of two populations of SNe Ia.

Key Words.:
supernovae: general – surveys

1 Introduction

Although there is general consensus that Type Ia supernovae (SNe Ia) originate from the thermonuclear explosions of white dwarfs (WDs) in binary systems, there is still disagreement about the explosion mechanisms and progenitor scenarios (see Hillebrandt et al., 2013; Maoz et al., 2014; Ruiter, 2020; Jha et al., 2019; Liu et al., 2023, for comprehensive reviews). The light curves of SNe Ia around peak light have been extensively studied because in order to use SNe Ia as cosmological distance indicators (Riess et al., 1998; Perlmutter et al., 1999; Scolnic et al., 2018), they are standardised using empirical relations between light curve shape, brightness, and colour around maximum light in the optical (Pskovskii, 1977; Phillips et al., 1993; Hamuy et al., 1996; Tripp, 1998). Around 2–3 weeks post maximum light, SN Ia light curves begin to rise again to a secondary bump in the optical i𝑖iitalic_i band and at near-infrared (NIR) wavelengths (Elias et al., 1981). A shoulder is seen in the optical r𝑟ritalic_r band at similar phases. This phenomenon is called the secondary maximum, and is significantly less well studied than the primary maximum due to the fainter magnitudes at these phases combined with the difficulties associated with observing in the NIR. Although less well understood, the secondary maximum encapsulates important information about the physical parameters of the explosion. Dhawan et al. (2016) has shown that the secondary maximum can be used to estimate the amount of 56Ni produced in the explosion, and Papadogiannakis et al. (2019a) has shown the properties of the secondary maximum to be linked to the transparency timescale and total ejected mass in the explosion. Moreover, it has been shown that the secondary maximum in the i𝑖iitalic_i and NIR bands can be used to standardise SN Ia light curves in a similar manner as the primary maximum (Nobili et al., 2005; Shariff et al., 2016), although this does not significantly reduce the Hubble residual scatter (Shariff et al., 2016).

The secondary maximum is thought to be caused by the ionisation transition of iron-group elements (IGEs) in the ejecta from doubly- to singly-ionised once the ejecta temperature drops below similar-to\sim7000K. The increased NIR emissivity of singly, compared to doubly ionised IGEs causes the NIR flux to increase (Pinto & Eastman, 2000; Kasen, 2006; Kasen & Woosley, 2007; Dhawan et al., 2015; Hoeflich, 2017). This explanation was corroborated by the identification of a strengthening Fe ii spectral feature in the i𝑖iitalic_i band of SN 2014J (Jack et al., 2015), coinciding with the onset of the secondary maximum. Slower evolving light curves with higher peak absolute magnitudes tend to have hotter ejecta and consequently, later secondary maxima (Kasen, 2006; Blondin et al., 2015). Fainter and faster evolving SNe Ia have earlier secondary maxima (Hamuy et al., 1996; Kasen, 2006; Taubenberger, 2017). The existence of the secondary maximum relies on at least some level of abundance stratification, meaning that explosions with large scale mixing are not expected to show a secondary maximum (Kasen, 2006). Moreover, asymmetric ejecta could lead to a viewing angle dependence in the properties of the secondary maximum (Kasen, 2006). The progenitor metallicity also affects the secondary maximum through its impact on the amount of stable iron produced, with increased progenitor metallicity leading to an earlier secondary maximum (Kasen, 2006).

Some of the more peculiar sub-types of SNe Ia lack a secondary maximum entirely (Turatto et al., 1996; Hoeflich et al., 2002; Gonzálezgonz´gonzález-Gaitán et al., 2014; Dhawan et al., 2017; Ashall et al., 2020; Lu et al., 2021; Hoogendam et al., 2022; Dimitriadis et al., 2023). In the case of 91bg-like SNe Ia or SNe Iax, the lack of secondary maximum is explained by a very low ejecta temperature, which causes the secondary maximum to merge with the primary peak (Kasen, 2006; Blondin et al., 2015; Taubenberger, 2017; Galbany et al., 2019). A similar effect has been seen in the late-time light curve of 91bg-like SN 2021qvv (Graur et al., 2023), which lacked a NIR plateau. The NIR plateau occurs between 70–500 d for normal SNe Ia and is the result of a shift in the dominant ionisation state of IGEs, resembling the secondary maximum (Deckers et al., 2023). On the other extreme end of the SN Ia absolute luminosity scale, the very bright 03fg-like SNe Ia are thought to lack a secondary maximum because the secondary re-brightening is obscured by the interaction with a dense circumstellar material (CSM) or envelope (Taubenberger et al., 2011; Taubenberger, 2017; Dimitriadis et al., 2021, 2023). The sub-class of 02es-like SNe Ia similarly lack a secondary maximum, and Hoogendam et al. (2023) suggest that this sub-class originates from the same progenitor scenario as 03fg-like SNe Ia, and that they similarly lack a secondary maximum due to interaction with a dense envelope.

Papadogiannakis et al. (2019a) study the secondary maximum of 422 SNe Ia provided by Carnegie Supernova Project (CSP-I, Contreras et al., 2010), CfA supernova programme (Hicken et al., 2009), the Palomar Transient Factory (PTF), and the intermediate Palomar Transient Factory (iPTF, Rau et al., 2009). Out of these surveys, only PTF/iPTF are untargeted, which makes up about half their sample. They find correlations between the properties of the secondary maximum in the r𝑟ritalic_r band and sBVsubscript𝑠𝐵𝑉s_{BV}italic_s start_POSTSUBSCRIPT italic_B italic_V end_POSTSUBSCRIPT (a proxy for light curve stretch measured by SNooPy, Burns et al., 2014) as well as the transparency timescale (t0subscript𝑡0t_{0}italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT), which can be directly compared to predictions from explosion models, highlighting the power of the secondary maximum.

In this paper, we build upon the analysis by Papadogiannakis et al. (2019a) by performing a large scale study of the secondary maximum in the r𝑟ritalic_r and i𝑖iitalic_i bands for 893 SNe Ia provided by the Zwicky Transient Facility’s (ZTF) second data release (DR2; Rigault et al., 2024). With our large spectroscopic data set, we aim to confirm which specific spectral features are producing the secondary maximum in the r𝑟ritalic_r band. We also investigate if the correlations found by Papadogiannakis et al. (2019a) between the secondary maximum and other light curve parameters hold for our larger sample. Finally, with 893 SNe Ia, DR2 contains many other sub-types beyond just normal SNe Ia, we investigate the secondary maximum properties of the more peculiar SNe Ia to test whether their properties can be connected to the normal SN Ia population through the secondary maximum as was shown by Li et al. (2022) for the transitional SN 2012ij.

In Sect. 2 we introduce the DR2 data set. We describe how we fit and analyse our light curves, introduce average spectral templates, and summarise the final samples analysed in this paper in Sect. 3. The results of our light-curve fits are presented in Sect. 4, and we provide a detailed discussion in Sect. 5. Lastly, we summarise our results in Sect. 6.

2 Data

In Section 2.1 we describe the ZTF DR2 dataset. In Section 2.2, we describe the coverage cuts we applied to ensure we only fit well sampled light curves.

2.1 The sample

For this investigation, we use data from the ZTF survey (ZTF, Bellm et al., 2019; Graham et al., 2019; Masci et al., 2019; Dekany et al., 2020). ZTF is a large field-of-view, galaxy-untargeted transient survey that has discovered thousands of transients since first light in 2018. In particular, we will be investigating the second SN Ia data release (DR2; Rigault et al., 2024), which contains SN Ia data of 3628 spectroscopically confirmed events from the first three years of operations (spanning 2018–2020). An overview of the sample statistics and technical details are described in Smith & ZTF Collaboration (2024).

We place a redshift cut at z0.06𝑧0.06z\leq 0.06italic_z ≤ 0.06 because the sample is considered spectroscopically complete for normal SNe Ia up to z𝑧absentz\leqitalic_z ≤0.06 (Amenouche et al., 2024). We note that the fainter sub-types in the sample will not all be discovered to this redshift (see Dimitriadis & ZTF Collaboration, 2024, for further discussion). Placing a redshift cut at z𝑧absentz\leqitalic_z ≤0.06 reduces the initial sample size from 3628 to 1584 SNe Ia.

Additional cuts on the sample are placed in Sects. 2.1, 2.2, 3.1, 3.2, and 3.3. We summarise the final samples in Sect. 3.6 and in Table 1.

Table 1: A summary of cuts applied to produce our final sample.
Criterion No. of SNe Removed
DR2 3628 -
z0.06𝑧0.06z\leq 0.06italic_z ≤ 0.06 1584 2044
Light curve coverage conditionsa 976 608
GP length scale <<< 44 d 973 3
σしぐまt2(r)<5.9subscript𝜎subscript𝑡2𝑟5.9\sigma_{t_{2}(r)}<5.9italic_σしぐま start_POSTSUBSCRIPT italic_t start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_r ) end_POSTSUBSCRIPT < 5.9 d 928 45
σしぐまr2<0.05subscript𝜎subscriptsubscript𝑟20.05\sigma_{\mathcal{F}_{r_{2}}}<0.05italic_σしぐま start_POSTSUBSCRIPT caligraphic_F start_POSTSUBSCRIPT italic_r start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT end_POSTSUBSCRIPT < 0.05 898 30
Manual inspection of outliers 893 5
Cosmological sub-sampleb 783 110
  • a

    Light curve coverage conditions require detections in at least two filters, in at least one filter pre-max and at least one filter post-max. At least two detections pre-peak and at least two detections post peak. A minimum of seven detections across all filters, and a minimum of four detections between +10 and +40 d in the r𝑟ritalic_r band.

  • b

    Sub-sample of SNe Ia that pass the following cuts: classified as normal or 91T-like, |x1|<3subscript𝑥13|x_{1}|<3| italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT | < 3, |σしぐまx1|1subscript𝜎subscript𝑥11|\sigma_{x_{1}}|\leq 1| italic_σしぐま start_POSTSUBSCRIPT italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_POSTSUBSCRIPT | ≤ 1, 0.2c0.80.2𝑐0.8-0.2\leq c\leq 0.8- 0.2 ≤ italic_c ≤ 0.8, |σしぐまc|0.1subscript𝜎𝑐0.1|\sigma_{c}|\leq 0.1| italic_σしぐま start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT | ≤ 0.1, |σしぐまt0|1subscript𝜎subscript𝑡01|\sigma_{t_{0}}|\leq 1| italic_σしぐま start_POSTSUBSCRIPT italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_POSTSUBSCRIPT | ≤ 1, and fit-probability > 1e7absent1superscript𝑒7>\ 1e^{-7}> 1 italic_e start_POSTSUPERSCRIPT - 7 end_POSTSUPERSCRIPT (see Sect. 3.6).

2.2 Light curve data

In this study we use light curves which were obtained by the ZTF survey in three optical filters (gri). We will perform Gaussian Process (GP) fits to characterise these light curves, which work best when the photometry is well sampled. Therefore, we apply quality cuts to the light curve data. We require each light curve to have detections in at least two filters, and in at least one of the same filter pre- and one post-maximum light. We also require a minimum of two detections pre-peak and two post-peak in any filter, and in total a minimum of seven detections across all filters. These criteria are applied to the stacked light curves, meaning that two data points on one epoch only count as a single detection. These light curve quality criteria ensure that we are able to fit the light curves around maximum light to estimate t0xsubscript𝑡subscript0𝑥t_{0_{x}}italic_t start_POSTSUBSCRIPT 0 start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT end_POSTSUBSCRIPT (the time of maximum in a specific filter, x). We also require a minimum of four detections between +10 and +40 d in the r𝑟ritalic_r band. This criterion is adopted from Papadogiannakis et al. (2019a) and ensures we are able to adequately constrain the secondary maximum. If there are less than four detections between +10 and +40 d in the i𝑖iitalic_i band, we keep the light curve in the sample but we do not attempt to measure the parameters of the secondary maximum in the i𝑖iitalic_i band. The aforementioned light curve coverage cuts reduce the size of the sample from 1584 to 976 SNe Ia.

3 Analysis

In Sect. 3.1, we describe the GP fits to our light curves and explain our motivation for the choice of parameters (see e.g., Rasmussen & Williams, 2006, for details on GP processes). In Sections 3.2 and 3.3, we describe the parametrisation of the timing and strength of the secondary maximum, respectively, while in Sect. 3.4, we discuss the impact of K-corrections on the parameters we measure. We describe why a number of SNe Ia were cut from the sample after we inspected outliers by eye, and then summarise all the cuts we applied to arrive at the final two samples used in this analysis. Finally, we introduce a number of spectral templates which we use to constrain the spectral evolution during the secondary maximum.

3.1 Fitting light curves using Gaussian Processes

To characterise the evolution of the r-band light curves, we perform GP fits to the data. GPs are a data-driven, non-parametric method of estimating an underlying function behind data. The same method was implemented by Papadogiannakis et al. (2019b, a) and Pessi et al. (2021) to characterise the r- and i𝑖iitalic_i-band light curves of SN Ia, respectively. GPs were also used for piscola (Müller-Bravo et al., 2022), a light curve fitter, and avocado (Boone, 2019), a photometric classifier.

We use the python package scikit-learn to implement our GP fits, which we perform in 2D in order to fit data across different filters simultaneously. We fit the light curves in 2D to maximise the amount of data available, which helps to constrain the timing of the primary maximum (Dimitriadis & ZTF Collaboration, 2024). Our analysis focuses primarily on the r𝑟ritalic_r band but we also analyse the i𝑖iitalic_i band, which generally has less available data than the r𝑟ritalic_r band. By fitting the different filters simultaneously we increase the quality of the i𝑖iitalic_i band light-curve fits. We fit our nightly stacked ZTF light curves in flux space so that we are able to include non-detections, which could not be used if we were fitting in magnitude space. An uncertainty floor of 2.5, 3.5, and 6 per cent were applied to the flux values in the g𝑔gitalic_g, r𝑟ritalic_r, and i𝑖iitalic_i band, respectively (Smith & ZTF Collaboration, 2024). We fit a phase range of --15 – 50 d with respect to maximum light. The fluxes are normalised to peak, as is standard for GP fits to ease maximum likelihood estimation (Rasmussen & Williams, 2006). An example of an average GP fit of a SN Ia in our sample is shown in the top panel of Fig. 1.

Refer to caption
Figure 1: Top: Example of GP fit to a r𝑟ritalic_r-band light curve. The GP fit is shown as a red line and the 1σしぐま1𝜎1\sigma1 italic_σしぐま uncertainty is shown as the red shaded region. The area over which the integrated flux under the secondary maximum, r2subscriptsubscript𝑟2\mathcal{F}_{r_{2}}caligraphic_F start_POSTSUBSCRIPT italic_r start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT, is calculated is indicated by the grey shaded region, and t2(r)subscript𝑡2𝑟t_{2}(r)italic_t start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_r ) by the black dashed line. Middle: The first derivative of the GP light curve. This light curve has a secondary shoulder rather than a secondary maximum because the first derivative is not equal to zero. Bottom: The second derivative of the GP light curve, showing that the time of the onset of the secondary maximum is the time at which the second derivative is equal to zero, also known as an inflection point.
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure 2: Impact of various kernel and length scale choices on a GP fit of a ZTF SN Ia light curve. Top left: If the uncertainties in the data are underestimated, not including a white noise kernel will result in unrealistic wiggles in an attempt to fit all the points. Top right: Even if the uncertainties are not underestimated in the data, a lack of a white noise kernel can result in an underestimation of the uncertainty of the GP fit, as is visible between 10–30 d in this light curve. Bottom left: With limited data at the edge of the fitting regions, not including a constant kernel can result in deviating behaviour at the edge regions. Bottom right: We optimise the length scale during the fit, which in this case resulted in \ellroman_ℓ = 12.3 d. By setting a length scale, the likelihood is higher to either over-fit if the length scale is too short (\ellroman_ℓ = 3 d), or to under-fit and miss important features if the length scale is too long (\ellroman_ℓ = 30 d). The ideal value for the length scale will depend on the cadence of the data, which varies across our sample.

The choice of kernels is important for GP fitting. We test both the radial basis function (RBF) kernel and the Matern kernel. The RBF kernel is an infinitely differentiable, stationary kernel, meaning that it produces smooth output functions (Rasmussen & Williams, 2006; Duvenaud, 2014), and depends only on the distance of two data-points, which is parameterised by a length scale parameter (\ellroman_ℓ), and not on their absolute values. The Matern kernel is also stationary, but it has an additional parameter νにゅー𝜈\nuitalic_νにゅー, which controls the smoothness of the function. The Matern kernel only produces smooth, infinitely differentiable functions at particular values of νにゅー𝜈\nuitalic_νにゅー (Rasmussen & Williams, 2006; Duvenaud, 2014). We find that the Matern kernel overfits the ZTF light curves, resulting in spurious wiggles in the model. We therefore implement the RBF kernel for our GP fits. Pessi et al. (2019, 2021) also opted for a RBF kernel, whereas Papadogiannakis et al. (2019a) opted for a Matern kernel.

We test how adding additional kernels (white noise, constant) to our GP model impacts the fits (see Fig. 2). In the top left panel of Fig. 2 we highlight that if a light curve has large gaps in the data, a GP without a white noise kernel will over-fit the data resulting in unphysical fluctuations in the flux predictions. The top right panel of Fig. 2 shows that a GP without a white noise kernel is likely to under-predict the uncertainties, as is shown between 10–30 d. The bottom left panel of Fig. 2 shows that excluding the constant kernel can result in deviating behaviour at the edges of the function, particularly if there is limited data in these regions. As a result of these tests, we opted to include a constant and a white noise kernel.

We also test the impact of the length scale of the time axis of the RBF kernel on the GP fits. In the bottom right panel of Fig. 2 we show the fit for a typical event in terms of data coverage, by fixing the length scale to either a small value (\ellroman_ℓ = 3 d) or a large value (\ellroman_ℓ = 30 d). The small value results in over-fitting of the data, while the large value results in an overly smoothed function that does not capture the crucial variation of SN Ia light curves around the secondary maximum. We also show a fit where the length scale is optimised without bounds in the bottom right panel of Fig. 2. For this example event the length scale after optimisation is 12.3 d, and the fit captures the evolution of the light well without over-fitting.

Motivated by our tests, we choose to use the RBF kernel, combined with a constant and a noise kernel. We optimise the length scale of the RBF kernel for each SN, since the optimum values will depend on the cadence of each individual light curve. We find a range of length scales between 4–70 d with a median of 10 d. Upon inspection by eye, light curves with very long length scales are not well fit, so we remove any SNe Ia from our sample that are fit with a length scale >44absent44>44> 44 d (rejecting measurements which are further than 3σしぐま𝜎\sigmaitalic_σしぐま away from the median of the distribution). This cut removes three out of 973 SNe Ia from our sample. We implement non-informative priors on the hyper parameters of the other kernels, and run an optimisation for each SN fit.

We estimate the uncertainties on the timing and strength of the secondary maximum (Sect. 3.2 and 3.3) by resampling. We perform 1000 iterations where we perturb the stacked flux data points within their Gaussian uncertainties and rerun the GP fits. We repeat our measurements and take the standard deviation across the 1000 iterations as the uncertainty.

Refer to caption
Figure 3: The distribution of t2(r)subscript𝑡2𝑟t_{2}(r)italic_t start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_r ) for the 63 SNe Ia with a detected true bump (i.e. dF/dt= 0𝑑𝐹𝑑𝑡 0dF/dt\ =\ 0italic_d italic_F / italic_d italic_t = 0 exists), measured either as the local maximum (dF/dt= 0𝑑𝐹𝑑𝑡 0dF/dt\ =\ 0italic_d italic_F / italic_d italic_t = 0, x-axis) or the inflection point (d2F/dt2= 0superscript𝑑2𝐹𝑑superscript𝑡2 0d^{2}F/dt^{2}\ =\ 0italic_d start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_F / italic_d italic_t start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT = 0, y-axis). We systematically find a lower t2(r)subscript𝑡2𝑟t_{2}(r)italic_t start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_r ) if measured as the point of inflection vs. the local maximum in the light curve.
Refer to caption
Figure 4: Impact of K-corrections on t2(r)subscript𝑡2𝑟t_{2}(r)italic_t start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_r ) (top panel) and r2subscriptsubscript𝑟2\mathcal{F}_{r_{2}}caligraphic_F start_POSTSUBSCRIPT italic_r start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT (bottom panel), as a function of redshift and x1subscript𝑥1x_{1}italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT (left) and c𝑐citalic_c (right). The colour of the data points represents the residual between the measured parameter in the SALT3 model at that point and the same model at z𝑧zitalic_z = 0. Overall the impact of K-corrections is small, with a maximum residual on r2subscriptsubscript𝑟2\mathcal{F}_{r_{2}}caligraphic_F start_POSTSUBSCRIPT italic_r start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT of 0.014 and on t2(r)subscript𝑡2𝑟t_{2}(r)italic_t start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_r ) of 0.66 d.

3.2 Determining the time of the secondary maximum

To determine the time of the secondary maximum, we trace the GP light curves with a univariate spline and calculate the first and second derivatives. The secondary maximum is expected to occur in the range of 10–40 d with respect to maximum light (Papadogiannakis et al., 2019a). Papadogiannakis et al. (2019a) search this range for where the first derivative is equal to zero and the second derivative is negative, which would correspond to a local maximum in the light curve. Although this would work in the i𝑖iitalic_i band, the secondary maximum is less prominent in the r band and it is usually better characterised as a shoulder rather than a local maximum. In order to constrain a shoulder in the r-band light curve, we locate the first inflection point (where the second derivative is equal to zero, see Fig. 1) in the range 13–35 d, which we define as t2(r)subscript𝑡2𝑟t_{2}(r)italic_t start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_r ). We reduced the upper bound of the range from 40 d to 35 d because Papadogiannakis et al. (2019a) found no secondary maxima later than 35 d, and we find that there is often limited data at this range which can result in an unexpected behaviour of the GP function. We reject any measurements of t2(r)subscript𝑡2𝑟t_{2}(r)italic_t start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_r ) with an uncertainty more than 3σしぐま𝜎\sigmaitalic_σしぐま away from the median of the uncertainty distribution (σしぐまt2> 5.9subscript𝜎subscript𝑡25.9\sigma_{t_{2}}\ >\ 5.9italic_σしぐま start_POSTSUBSCRIPT italic_t start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT > 5.9 d), to eliminate objects with poor GP fits, which removes 45 SNe Ia.

We find 20 91bg-, seven 02cx-, and three 02es-like SNe Ia (including SN 2019yvq), with t2(r)subscript𝑡2𝑟t_{2}(r)italic_t start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_r ) ¿20 d. These sub-types are not expected to show strong, late secondary maxima, making them stand out significantly from the rest of the sample. We have checked all their light curves by eye and find that their light curves flatten as they approach zero flux around 20 d. Because the absolute flux is very low at these epochs, small fluctuations are picked up in the derivatives and can be classified as inflection points, but these are not a real secondary shoulders/maxima. This is not a problem for most of the normal SNe Ia in the sample because they decline less rapidly, and therefore do not have such low flux values in the window where we search for the secondary maximum. Because the light curves of these objects are well fit by the GP function, we do not remove these SNe Ia from our sample, but instead we only remove their measured t2(r)subscript𝑡2𝑟t_{2}(r)italic_t start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_r ) values. The only exceptions are SN 2018jpk (an 02cx-like SN Ia) and SN 2019ner (an 02es-like SN Ia) which do appear to have a real secondary maximum at 20.9 and 20.6 d, respectively. We have kept both of their t2(r)subscript𝑡2𝑟t_{2}(r)italic_t start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_r ) values in the sample.

Our t2(r)subscript𝑡2𝑟t_{2}(r)italic_t start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_r ) values are better described as the onset of the shoulder, rather than the peak of the secondary maximum (see Fig. 1). Papadogiannakis et al. (2019a) also measured the inflection point, but only for light curves where they could not locate a true secondary maximum (dF/dt𝑑𝐹𝑑𝑡dF/dtitalic_d italic_F / italic_d italic_t = 0). We opt to use the point where d2F/dt2= 0superscript𝑑2𝐹𝑑superscript𝑡2 0d^{2}F/dt^{2}\ =\ 0italic_d start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_F / italic_d italic_t start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT = 0 (inflection point) for all light curves to ensure we are consistent across the sample because we are interested in relative t2(r)subscript𝑡2𝑟t_{2}(r)italic_t start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_r ) values. We are able to find a secondary maximum for 808 SNe Ia, and 60 of these could be characterised as having a true bump (dF/dt𝑑𝐹𝑑𝑡dF/dtitalic_d italic_F / italic_d italic_t = 0)111In this paper we refer to the secondary shoulder in the r𝑟ritalic_r band, t2(r)subscript𝑡2𝑟t_{2}(r)italic_t start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_r ), as a secondary maximum, even if dF/dt𝑑𝐹𝑑𝑡dF/dtitalic_d italic_F / italic_d italic_t = 0 does not exist in the light curve.. A direct comparison of our t2(r)subscript𝑡2𝑟t_{2}(r)italic_t start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_r ) values to those presented in Papadogiannakis et al. (2019a) is hindered by the fact that they do not specify which t2(r)subscript𝑡2𝑟t_{2}(r)italic_t start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_r ) values correspond to dF/dt𝑑𝐹𝑑𝑡dF/dtitalic_d italic_F / italic_d italic_t = 0 vs. d2F/dt2= 0superscript𝑑2𝐹𝑑superscript𝑡2 0d^{2}F/dt^{2}\ =\ 0italic_d start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_F / italic_d italic_t start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT = 0.

In Fig. 3 we highlight the importance of using a consistent definition of t2(r)subscript𝑡2𝑟t_{2}(r)italic_t start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_r ). We measure t2(r)subscript𝑡2𝑟t_{2}(r)italic_t start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_r ) only for SNe Ia with an identified bump (dF/dt= 0𝑑𝐹𝑑𝑡 0dF/dt\ =\ 0italic_d italic_F / italic_d italic_t = 0), using both methods. We find that the median difference between the inflection point d2F/dt2= 0superscript𝑑2𝐹𝑑superscript𝑡2 0d^{2}F/dt^{2}\ =\ 0italic_d start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_F / italic_d italic_t start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT = 0 and the local maximum in the light curve (dF/dt= 0𝑑𝐹𝑑𝑡 0dF/dt\ =\ 0italic_d italic_F / italic_d italic_t = 0) is 1.9 d, with the inflection point generally occurring earlier.

We also measure t2subscript𝑡2t_{2}italic_t start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT using the inflection point (d2F/dt2= 0superscript𝑑2𝐹𝑑superscript𝑡2 0d^{2}F/dt^{2}\ =\ 0italic_d start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_F / italic_d italic_t start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT = 0) for the i𝑖iitalic_i band (t2(i)subscript𝑡2𝑖t_{2}(i)italic_t start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_i )). We find that 103 of the i𝑖iitalic_i band light curves could be characterised as having a true bump (dF/dt= 0𝑑𝐹𝑑𝑡 0dF/dt\ =\ 0italic_d italic_F / italic_d italic_t = 0). This is more than in the r𝑟ritalic_r band, which is not surprising because the secondary maximum is known to be stronger in the i𝑖iitalic_i band. The median offset between the point at which dF/dt= 0𝑑𝐹𝑑𝑡 0dF/dt\ =\ 0italic_d italic_F / italic_d italic_t = 0 and d2F/dt2= 0superscript𝑑2𝐹𝑑superscript𝑡2 0d^{2}F/dt^{2}\ =\ 0italic_d start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_F / italic_d italic_t start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT = 0 is 4.6 d, which is larger than in the r𝑟ritalic_r band. This is because the secondary maximum is more prominent and lasts longer in the i𝑖iitalic_i band (see fig. 10, Kasen, 2006), and as a result the time between the onset and the peak of the secondary maximum is greater.

3.3 Constraining the strength of the secondary maximum

We use the integrated flux under the secondary maximum, r2subscriptsubscript𝑟2\mathcal{F}_{r_{2}}caligraphic_F start_POSTSUBSCRIPT italic_r start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT, to quantify the strength of the secondary maximum (see shaded region in the top panel of Fig. 1). The same metric was used in Krisciunas et al. (2001) and Papadogiannakis et al. (2019a) to quantify the strength of the secondary maximum in the i and r bands, respectively. r2subscriptsubscript𝑟2\mathcal{F}_{r_{2}}caligraphic_F start_POSTSUBSCRIPT italic_r start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT is calculated by normalising the flux to peak in the r𝑟ritalic_r band, integrating the flux between +15 and +40 d (where rest-frame phase is defined relative to peak in the r𝑟ritalic_r band) and dividing by the length of the interval (25 d). We reject any measurements of r2subscriptsubscript𝑟2\mathcal{F}_{r_{2}}caligraphic_F start_POSTSUBSCRIPT italic_r start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT with an uncertainty more than 3σしぐま𝜎\sigmaitalic_σしぐま away from the median of the uncertainty distribution (σしぐまr2> 0.05subscript𝜎subscriptsubscript𝑟20.05\sigma_{\mathcal{F}_{r_{2}}}\ >\ 0.05italic_σしぐま start_POSTSUBSCRIPT caligraphic_F start_POSTSUBSCRIPT italic_r start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT end_POSTSUBSCRIPT > 0.05) in order to eliminate objects with poor GP fits, which removes 30 SNe Ia.

3.4 Uncertainties due to K-corrections

At high redshifts, it is important to apply K-corrections because the observed r𝑟ritalic_r-band will begin to resemble rest-frame g𝑔gitalic_g-band, where no secondary maximum is expected. Unfortunately, this is not straightforward due to a lack of reliable templates at the phases we are investigating (+15 to +40 d post peak). Moreover, some of the more peculiar sub-types present in our sample do not have sufficient template coverage to calculate reliable K-corrections at any phases.

We estimate the impact of K-corrections using the SALT3 (Spectral Adaptive Light-curve Template; Kenworthy et al., 2021) template in the sncosmo library (Barbary et al., 2022). We produce a grid of light curves which resemble the normal SNe Ia in our sample, with 3x1 33subscript𝑥13-3\ \leq\ x_{1}\ \leq\ 3- 3 ≤ italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ≤ 3 in steps of 0.5 and 0.2c 0.80.2𝑐0.8-0.2\ \leq\ c\ \leq\ 0.8- 0.2 ≤ italic_c ≤ 0.8 in steps of 0.1. We place these light curves at a range of redshifts (0z 0.060𝑧0.060\ \leq\ z\ \leq\ 0.060 ≤ italic_z ≤ 0.06). We next measure t2(r)subscript𝑡2𝑟t_{2}(r)italic_t start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_r ) and r2subscriptsubscript𝑟2\mathcal{F}_{r_{2}}caligraphic_F start_POSTSUBSCRIPT italic_r start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT for all the simulated light curves and find how much the values differ compared to the simulated light curve with the same x1subscript𝑥1x_{1}italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT and c𝑐citalic_c at z= 0𝑧 0z\ =\ 0italic_z = 0 (see Fig. 4). We adopt the residuals as additional uncertainties due to unaccounted for K-corrections in our measured parameters (t2(r/i)subscript𝑡2𝑟𝑖t_{2}(r/i)italic_t start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_r / italic_i ), r2/i2subscriptsubscript𝑟2subscript𝑖2\mathcal{F}_{r_{2}/i_{2}}caligraphic_F start_POSTSUBSCRIPT italic_r start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT / italic_i start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT). We select which grid residual point to adopt for each SN based on its redshift, x1subscript𝑥1x_{1}italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT, and c𝑐citalic_c in bins of 0.01, 0.5, and 0.1, respectively. The measured x1subscript𝑥1x_{1}italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT- and c𝑐citalic_c-values are not always reliable for the SNe Ia that are not in our cosmological sample (see Sect. 3.6), but we nonetheless use them here to obtain a rough estimate of the additional uncertainty due to unaccounted for K-corrections.

The variation over the redshift range covered by our sample is <<< 0.66 d for t2(r)subscript𝑡2𝑟t_{2}(r)italic_t start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_r ), <<< 0.20 d for t2(i)subscript𝑡2𝑖t_{2}(i)italic_t start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_i ), <<< 0.014 for r2subscriptsubscript𝑟2\mathcal{F}_{r_{2}}caligraphic_F start_POSTSUBSCRIPT italic_r start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT, and <<< 0.087 for i2subscriptsubscript𝑖2\mathcal{F}_{i_{2}}caligraphic_F start_POSTSUBSCRIPT italic_i start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT, suggesting that the impact of K-corrections on our results is minimal. As a sanity check, Fig. 4 also check how the measured parameters vary for the Hsiao template (Hsiao et al., 2007) compared to the SALT3 template with x1subscript𝑥1x_{1}italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = 0 and c𝑐citalic_c = 0. Although the absolute values of the measured parameters differ from those measured from SALT3, the deviation with redshift is comparable.

3.5 Outlier rejection

Although we try to ensure only good quality GP fits are included in our sample by applying a number of cuts detailed in the previous sections, a few poor fits are still present. Due to the large sample size we only inspect the light curves and GP fits for SNe Ia with extreme t2(r)subscript𝑡2𝑟t_{2}(r)italic_t start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_r ) or r2subscriptsubscript𝑟2\mathcal{F}_{r_{2}}caligraphic_F start_POSTSUBSCRIPT italic_r start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT values by eye. As a result of these manual inspections, we have removed a few objects from our sample.

SN 2018but (a normal SN Ia) was removed because the GP function showed unrealistic extrapolation due to a lack of data at > 30absent30>\ 30> 30 d. SN 2019sua (91bg-like) was removed because the data between 20–50 d showed spurious fluctuation with underestimated uncertainties, which resulted in unrealistic wiggles in the GP fit. Three additional SNe Ia were removed (SNe 2018ggw, 2020bxp, and 2020acol, all three normal) because their GP fits produce a dip in the light curve around 20 d even though there is no data in any band. After removing these additional five SNe Ia, the final sample contains 893 SNe Ia.

3.6 Summary of final samples

We provide a summary of the various cuts applied in Sects. 2.1, 2.2, 3.1, 3.2, and 3.3, and the outliers removed in this section that were used to arrive at our main sample and cosmological sample that was presented in Table 1.

Our final sample consists of 893 SNe Ia. Although the sample is dominated by normal SNe Ia (706), there are many sub-types of SNe Ia present. The most ubiquitous are the over-luminous 91T-like SNe Ia (56), followed by sub-luminous 91bg-like SNe Ia (36). A full break-down of the various sub-types is shown in Table 2, and a more detailed analysis of the various sub-types and their rates in DR2 will be presented in Dimitriadis & ZTF Collaboration (2024).

In this study we will use the decrease in magnitude between peak and +15 d in the g𝑔gitalic_g band, Δでるたm15(g)Δでるたsubscript𝑚15𝑔\Delta m_{15}(g)roman_Δでるた italic_m start_POSTSUBSCRIPT 15 end_POSTSUBSCRIPT ( italic_g ), as a measure of the stretch of a light curve. We adopt the Δでるたm15(g)Δでるたsubscript𝑚15𝑔\Delta m_{15}(g)roman_Δでるた italic_m start_POSTSUBSCRIPT 15 end_POSTSUBSCRIPT ( italic_g )-values from Dimitriadis & ZTF Collaboration (2024), which are estimated after K-correction using GP fits to the light curves. SALT2 based K-corrections are applied to normal and 91T-like SNe Ia, template K-corrections are applied to all other sub-types, apart from Ia-CSM, which have no applied K-corrections (see Dimitriadis & ZTF Collaboration, 2024, for further discussion). We also aim to perform comparisons between our parameters and those presented in other DR2 cosmology papers, which implement x1subscript𝑥1x_{1}italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT and c𝑐citalic_c measured using SALT2.4 (Guy et al., 2007; Taylor et al., 2021). SALT2.4 is trained predominantly on templates of normal SNe Ia and can therefore not accurately measure the light curve parameters for more peculiar SNe Ia. We therefore define a ‘cosmology’ sub-sample, which is defined using the same cuts as those implemented by Rigault et al. (2024) for the cosmological analysis of the DR2 sample (i.e.x1subscript𝑥1x_{1}italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT and c𝑐citalic_c: |x1|<3subscript𝑥13|x_{1}|<3| italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT | < 3, 0.2c0.80.2𝑐0.8-0.2\leq c\leq 0.8- 0.2 ≤ italic_c ≤ 0.8, cuts on the uncertainties of |σしぐまx1|1subscript𝜎subscript𝑥11|\sigma_{x_{1}}|\leq 1| italic_σしぐま start_POSTSUBSCRIPT italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_POSTSUBSCRIPT | ≤ 1, |σしぐまc|0.1subscript𝜎𝑐0.1|\sigma_{c}|\leq 0.1| italic_σしぐま start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT | ≤ 0.1, |σしぐまt0|1subscript𝜎subscript𝑡01|\sigma_{t_{0}}|\leq 1| italic_σしぐま start_POSTSUBSCRIPT italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_POSTSUBSCRIPT | ≤ 1, a fit-probability > 1e7absent1superscript𝑒7>\ 1e^{-7}> 1 italic_e start_POSTSUPERSCRIPT - 7 end_POSTSUPERSCRIPT, and either no sub-classification, or a sub-classification of normal, 91T- or 99aa-like). These cuts result in a sub-sample of 783 SNe Ia. We will distinguish this sub-sample from our main sample by referring to it as our cosmological sub-sample because it adheres to the cuts normally applied in order to perform cosmological analysis. However, we note that the DR2 is currently not suitable for inferring cosmological parameters (see Rigault et al., 2024, for further discussion).

Table 2: A summary of the SN Ia sub-types present in our final sample.
Sub-type Number
Normal 706
91T-like 56
Not sub-typed 50
91bg-like 36
99aa-like 19
02cx-like 13
03fg-like 7
02es-like∗∗ 4
Ia-CSM 2

*SNe Ia which were spectroscopically classified as SNe Ia but either due to low quality spectra or the phases of the spectra being too late, a sub-classification could not be determined. **We include SN 2019yvq (Miller et al., 2020; Siebert et al., 2020; Burke et al., 2021) in the 02es-like sub-class because it was tentatively classified as a 02es-like SN Ia by Burke et al. (2021).

3.7 Averaged spectral templates

In this paper, we are interested in constraining the spectral properties of SNe Ia from maximum light with the aim of understanding the spectral features contributing before, during, and after the time of secondary maximum (e.g. from peak to +30 d). To investigate this, we have compiled average spectral templates as a function of x1subscript𝑥1x_{1}italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT and phase from a sample of SN Ia spectra from the ZTF DR2 cosmological sub-sample (Johansson & ZTF Collaboration, 2024; Burgaz et al., 2024), supplemented with additional spectra taken by ZTF post-DR2 (obtained between February 2021 – September 2023) along with literature spectra obtained from WISEREP. The additional spectra were used to increase the sample size at the upper and lower ends of our x1subscript𝑥1x_{1}italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT range at later phases (+15 to +30 d).

We excluded spectra with a signal-to-noise ratio (S/N) in the continuum of <<<5, and we binned the spectra to 10 Å resolution to account for varying resolution across the sample, in particular the low resolution spectra (R 100similar-to𝑅100R\ \sim\ 100italic_R ∼ 100) taken by the SEDMachine (Rigault et al., 2019; Blagorodnova et al., 2018; Kim et al., 2022). In total, we include 265 spectra of SNe Ia in our cosmological sub-sample, where the phase is defined relative to the time of maximum in the g𝑔gitalic_g band from the SALT2.4 fit. We also included 9 SN Ia spectra taken by ZTF post-DR2 at z0.06𝑧0.06z\leq 0.06italic_z ≤ 0.06 at phases between +10 and +40 d. We fit the light curves of SNe Ia obtained post-DR2 with SALT2.4 using the same procedure as for the rest of the sample to obtain x1subscript𝑥1x_{1}italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT and t0subscript𝑡0t_{0}italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT values. Finally, we searched WISEREP for spectra taken between +10 and +40 d for normal SNe at z𝑧absentz\leqitalic_z ≤0.06 and with published x1subscript𝑥1x_{1}italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT and t0subscript𝑡0t_{0}italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT values, adding a further 17 spectra. These additional spectra were pre-processed in the same way as the ZTF DR2 spectra.

In Fig. 5 we show the r𝑟ritalic_r- and i𝑖iitalic_i-band region of the averaged spectra of SNe Ia from our cosmological sub-sample, sorted into two x1subscript𝑥1x_{1}italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT bins (3x1<03subscript𝑥10-3\leq x_{1}<0- 3 ≤ italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT < 0 and 0x1<30subscript𝑥130\leq x_{1}<30 ≤ italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT < 3), and six phase bins (0–10, 10–20, 20–30, 30–40, and 40–50 d). The spectra in each bin were first sigma-clipped (5σしぐま𝜎\sigmaitalic_σしぐま), then averaged, and finally smoothed with a Gaussian convolution with a standard deviation of 5. All averaged spectra are then normalised to the continuum region at 7000 Å. We also include spectra of SN 1999by (a 91bg-like SN Ia, Matheson et al., 2008) taken from WISEREP in the bottom panel of Fig. 5. The spectra of SN 1999by are also normalised to the continuum region around 7000 Å.

Refer to caption
Figure 5: Averaged spectra between 0–10, 10–20, 20–30, 30–40, and 40–50 d in the r𝑟ritalic_r- and i𝑖iitalic_i-band wavelength ranges. The phases are relative to g𝑔gitalic_g-band maximum. Spectra are binned into 3x1<03subscript𝑥10-3\leq x_{1}<0- 3 ≤ italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT < 0 (dashed) and 0x1<30subscript𝑥130\leq x_{1}<30 ≤ italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT < 3 (solid). The grey shaded regions indicate the spectral regions where a feature appears during the onset of the secondary maximum. The numbers in the brackets represent the number of spectra going into each bin (first number represents the number of spectra in the 3x1<03subscript𝑥10-3\leq x_{1}<0- 3 ≤ italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT < 0 bin, and the second number are the number of spectra in the 0x1<30subscript𝑥130\leq x_{1}<30 ≤ italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT < 3 bin). We also show the filter response functions of the ZTF r𝑟ritalic_r and i𝑖iitalic_i bands in black. In the bottom panel we show a spectral timeseries of SN 1999by (a 91bg-like SN Ia).

4 Results

After the additional cuts on the length scale and uncertainties on t2(r)subscript𝑡2𝑟t_{2}(r)italic_t start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_r ) and r2subscriptsubscript𝑟2\mathcal{F}_{r_{2}}caligraphic_F start_POSTSUBSCRIPT italic_r start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT, as well as the removal of outliers, the main sample remaining consists of 893 SNe Ia. In this section we test whether the correlations found by Papadogiannakis et al. (2019a) for the CfA and CSP samples persist for our sample. We then analyse the available spectra in order to constrain the spectral feature responsible for the secondary maximum in the r𝑟ritalic_r band. We next use the integrated flux under the secondary maximum to estimate the transparency timescales for the SNe Ia in our main sample, and compare it to model predictions. We also provide estimated total ejected masses derived from the transparency timescale, and compare these to masses derived from the SALT2.4 light curve width parameter, x1subscript𝑥1x_{1}italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT, for our cosmological sub-sample.

4.1 Timing and strength of secondary maximum

We have measured the timing and the strength of the secondary maximum for the r𝑟ritalic_r and i𝑖iitalic_i band. All four parameters, t2(r)subscript𝑡2𝑟t_{2}(r)italic_t start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_r ), t2(i)subscript𝑡2𝑖t_{2}(i)italic_t start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_i ), r2subscriptsubscript𝑟2\mathcal{F}_{r_{2}}caligraphic_F start_POSTSUBSCRIPT italic_r start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT, and i2subscriptsubscript𝑖2\mathcal{F}_{i_{2}}caligraphic_F start_POSTSUBSCRIPT italic_i start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT are shown as a function of Δでるたm15(g)Δでるたsubscript𝑚15𝑔\Delta m_{15}(g)roman_Δでるた italic_m start_POSTSUBSCRIPT 15 end_POSTSUBSCRIPT ( italic_g ) in the four panels of Fig. 6.

Refer to caption
Figure 6: The timing of secondary maximum (t2(x)subscript𝑡2𝑥t_{2}(x)italic_t start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_x ), top panel) and normalised integrated flux in the range 15–40 d (x2subscriptsubscript𝑥2\mathcal{F}_{x_{2}}caligraphic_F start_POSTSUBSCRIPT italic_x start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT, bottom panel) in the r𝑟ritalic_r band (left) and i𝑖iitalic_i band (right) as a function of Δでるたm15(g)Δでるたsubscript𝑚15𝑔\Delta m_{15}(g)roman_Δでるた italic_m start_POSTSUBSCRIPT 15 end_POSTSUBSCRIPT ( italic_g ) for our sample. We also show simple linear regression fits to the whole sample (black dashed lines). We find a non-zero slope at the >5σしぐまabsent5𝜎>5\sigma> 5 italic_σしぐま confidence level for t2(r)subscript𝑡2𝑟t_{2}(r)italic_t start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_r ), t2(i)subscript𝑡2𝑖t_{2}(i)italic_t start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_i ), r2subscriptsubscript𝑟2\mathcal{F}_{r_{2}}caligraphic_F start_POSTSUBSCRIPT italic_r start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT, but not for i2subscriptsubscript𝑖2\mathcal{F}_{i_{2}}caligraphic_F start_POSTSUBSCRIPT italic_i start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT with Δでるたm15(g)Δでるたsubscript𝑚15𝑔\Delta m_{15}(g)roman_Δでるた italic_m start_POSTSUBSCRIPT 15 end_POSTSUBSCRIPT ( italic_g ).

4.1.1 r band

We have measured the timing of the secondary maximum/shoulder in the r band, t2(r)subscript𝑡2𝑟t_{2}(r)italic_t start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_r ), for our ZTF sample and find values ranging between similar-to\sim13 – 34 d with respect to maximum light. This is similar to the range seen in Papadogiannakis et al. (2019a) but our values are slightly smaller due to the different definition of t2(r)subscript𝑡2𝑟t_{2}(r)italic_t start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_r ) (see Sect. 3.2).

In the top panel of Fig. 6 we show t2(r)subscript𝑡2𝑟t_{2}(r)italic_t start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_r ) as a function of Δでるたm15(g)Δでるたsubscript𝑚15𝑔\Delta m_{15}(g)roman_Δでるた italic_m start_POSTSUBSCRIPT 15 end_POSTSUBSCRIPT ( italic_g ). We find a strong correlation for the normal SN Ia population, with a Pearson correlation coefficient r𝑟ritalic_r of 0.690.69-0.69- 0.69 at a >5σしぐまabsent5𝜎>5\sigma> 5 italic_σしぐま confidence level. Papadogiannakis et al. (2019a) found the timing of the secondary maximum to correlate with SNooPy light-curve width parameter, sBVsubscript𝑠𝐵𝑉s_{BV}italic_s start_POSTSUBSCRIPT italic_B italic_V end_POSTSUBSCRIPT, which itself is strongly correlated with Δでるたm15(B)Δでるたsubscript𝑚15𝐵\Delta m_{15}(B)roman_Δでるた italic_m start_POSTSUBSCRIPT 15 end_POSTSUBSCRIPT ( italic_B ) (Burns et al., 2018). We were able to measure t2(r)subscript𝑡2𝑟t_{2}(r)italic_t start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_r ) for a number of unusual classes of SNe Ia (also shown in Fig. 6), and find that they broadly follow the normal sample relation. The correlation between t2(r)subscript𝑡2𝑟t_{2}(r)italic_t start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_r ) and Δでるたm15(g)Δでるたsubscript𝑚15𝑔\Delta m_{15}(g)roman_Δでるた italic_m start_POSTSUBSCRIPT 15 end_POSTSUBSCRIPT ( italic_g ) becomes weaker but remains statistically significant when tested for the whole sample rather than just the normal SNe Ia (Pearson correlation coefficient r𝑟ritalic_r of 0.600.60-0.60- 0.60 at a >5σしぐまabsent5𝜎>5\sigma> 5 italic_σしぐま confidence level).

There are seven, out of a total of 46, 91bg-like SNe Ia with an identified t2(r)subscript𝑡2𝑟t_{2}(r)italic_t start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_r ), which is unexpected because this sub-class is characterised by a lack of secondary maxima (e.g. Taubenberger, 2017). We have inspected all their light curves by eye and for six out of seven, their secondary maxima look real (i.e. there is sufficient data to constrain a secondary maximum and all data points around t2(r)subscript𝑡2𝑟t_{2}(r)italic_t start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_r ) are consistent with the GP fit, and if i𝑖iitalic_i-band data is available t2(r)subscript𝑡2𝑟t_{2}(r)italic_t start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_r ) agrees with t2(i)subscript𝑡2𝑖t_{2}(i)italic_t start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_i ) within 3σしぐま𝜎\sigmaitalic_σしぐま). We remove the t2subscript𝑡2t_{2}italic_t start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT measurements of SN 2020acbx from the sample because t2(i)subscript𝑡2𝑖t_{2}(i)italic_t start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_i ) is not coincident with t2(r)subscript𝑡2𝑟t_{2}(r)italic_t start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_r ) within 3σしぐま𝜎\sigmaitalic_σしぐま. We further find a secondary maximum for SN 2018jpk, which is a 02cx-like SNe Ia. The 02cx-like sub-class are not expected to have secondary maxima due to their well-mixed ejecta composition (Jha, 2017). We find one SN Ia-CSM (SN 2020uem Sharma et al., 2023) with a late secondary maximum (t2(r)=subscript𝑡2𝑟absentt_{2}(r)\ =\ italic_t start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_r ) =31.6 d). Although the secondary maximum looks real, it is possible that this increase in flux in the light curve is fuelled by the interaction with the CSM rather than recombination of IGE because the light curve remains approximately flat for similar-to\sim50 d.

We measured the normalised integrated flux under the secondary maximum (r2subscriptsubscript𝑟2\mathcal{F}_{r_{2}}caligraphic_F start_POSTSUBSCRIPT italic_r start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT), which we use as a measure of the strength of the secondary maximum. The typical range of r2subscriptsubscript𝑟2\mathcal{F}_{r_{2}}caligraphic_F start_POSTSUBSCRIPT italic_r start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT for the normal SNe Ia in our sample is 0.20 – 0.55, which deviates slightly from what was found by Papadogiannakis et al. (2019a) of 0.25 – 0.60. In Fig. 7 we compare the distributions of r2subscriptsubscript𝑟2\mathcal{F}_{r_{2}}caligraphic_F start_POSTSUBSCRIPT italic_r start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT-values between the two samples and find that the ZTF DR2 population peaks at lower values, with the medians of the two distributions differing by 0.1. We do not expect any intrinsic difference between the SN Ia samples. However, the integrated flux under the secondary maximum was originally defined as a parameter by Krisciunas et al. (2001) to quantify the strength of the secondary maximum in the i𝑖iitalic_i band. Krisciunas et al. (2001) integrated over the range 20–40 d, and normalised by dividing the integral by 20 d. Although Papadogiannakis et al. (2019a) described their integration as over the same range as this paper (15–40 d), if we divide our integrals by 20 d instead of 25 d, the r2subscriptsubscript𝑟2\mathcal{F}_{r_{2}}caligraphic_F start_POSTSUBSCRIPT italic_r start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT distribution from ZTF lines up very well with that of PTF and iPTF (the shifted distribution is also shown in Fig. 7). As a sanity check, we compare our i2subscriptsubscript𝑖2\mathcal{F}_{i_{2}}caligraphic_F start_POSTSUBSCRIPT italic_i start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT-values to those presented in Krisciunas et al. (2001) and they cover the same range, whereas those presented by Papadogiannakis et al. (2019a) are offset to higher i2subscriptsubscript𝑖2\mathcal{F}_{i_{2}}caligraphic_F start_POSTSUBSCRIPT italic_i start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT-values, just as their r2subscriptsubscript𝑟2\mathcal{F}_{r_{2}}caligraphic_F start_POSTSUBSCRIPT italic_r start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT-values are.

The bottom panel of Fig. 6 shows r2subscriptsubscript𝑟2\mathcal{F}_{r_{2}}caligraphic_F start_POSTSUBSCRIPT italic_r start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT as a function of Δでるたm15(g)Δでるたsubscript𝑚15𝑔\Delta m_{15}(g)roman_Δでるた italic_m start_POSTSUBSCRIPT 15 end_POSTSUBSCRIPT ( italic_g ), and we also find a very strong correlation between these parameters for the normal SN Ia population, with a Pearson correlation coefficient r𝑟ritalic_r of 0.720.72-0.72- 0.72 at a >5σしぐまabsent5𝜎>5\sigma> 5 italic_σしぐま confidence level. r2subscriptsubscript𝑟2\mathcal{F}_{r_{2}}caligraphic_F start_POSTSUBSCRIPT italic_r start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT is a useful metric because we are able to measure it for all the light curves in our sample, not just those with a measured secondary maxima. This means we are able to include sub-types of SNe Ia which typically do not show a secondary maximum. When measured for the full sample, the strong correlation remains (a Pearson correlation coefficient r𝑟ritalic_r of 0.710.71-0.71- 0.71 at a >5σしぐまabsent5𝜎>5\sigma> 5 italic_σしぐま confidence level).

The most distinct outlier visible in the bottom panel of Fig. 6 is a Ia-CSM, which sits 21σしぐま𝜎\sigmaitalic_σしぐま above the trend. In general, the 03fg-like sub-class also have larger r2subscriptsubscript𝑟2\mathcal{F}_{r_{2}}caligraphic_F start_POSTSUBSCRIPT italic_r start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT than expected for their Δでるたm15(g)Δでるたsubscript𝑚15𝑔\Delta m_{15}(g)roman_Δでるた italic_m start_POSTSUBSCRIPT 15 end_POSTSUBSCRIPT ( italic_g ) value. Since both these sub-classes of SNe Ia are thought to be powered by some interaction with a dense envelope/CSM (Taubenberger et al., 2011; Taubenberger, 2017; Dimitriadis et al., 2021, 2023; Sharma et al., 2023), a higher r2subscriptsubscript𝑟2\mathcal{F}_{r_{2}}caligraphic_F start_POSTSUBSCRIPT italic_r start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT is expected. Some of the 02es-like SNe Ia also sit above the trend, implying a higher than expected r2subscriptsubscript𝑟2\mathcal{F}_{r_{2}}caligraphic_F start_POSTSUBSCRIPT italic_r start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT. Hoogendam et al. (2023) suggested that this sub-class is linked to the 03fg-like sub-class because the two share similar photometric and spectroscopic properties. Interaction with a dense envelope could also explain the higher r2subscriptsubscript𝑟2\mathcal{F}_{r_{2}}caligraphic_F start_POSTSUBSCRIPT italic_r start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT for the 02es-like sub-class. We note that all three of these sub-classes show a range of r2subscriptsubscript𝑟2\mathcal{F}_{r_{2}}caligraphic_F start_POSTSUBSCRIPT italic_r start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT-values, with the majority having larger r2subscriptsubscript𝑟2\mathcal{F}_{r_{2}}caligraphic_F start_POSTSUBSCRIPT italic_r start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT-values than normal SNe Ia, but some are consistent with the normal population.

There is a spread in the r2subscriptsubscript𝑟2\mathcal{F}_{r_{2}}caligraphic_F start_POSTSUBSCRIPT italic_r start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT values of the 02cx-like sub-class, with most being consistent with the general trend, but several sitting slightly above the trend. 02cx-like SNe Ia are thought to leave behind a bound remnant which could explain the additional flux at these phases (Kromer et al., 2013a; Jha, 2017).

The mean r2subscriptsubscript𝑟2\mathcal{F}_{r_{2}}caligraphic_F start_POSTSUBSCRIPT italic_r start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT for the 91T-like sub-class is 0.41, whereas the mean r2subscriptsubscript𝑟2\mathcal{F}_{r_{2}}caligraphic_F start_POSTSUBSCRIPT italic_r start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT for normal SNe Ia in the same Δでるたm15(g)Δでるたsubscript𝑚15𝑔\Delta m_{15}(g)roman_Δでるた italic_m start_POSTSUBSCRIPT 15 end_POSTSUBSCRIPT ( italic_g ) range (0.3 Δでるたm15(g)absentΔでるたsubscript𝑚15𝑔absent\leq\Delta m_{15}(g)\ \leq≤ roman_Δでるた italic_m start_POSTSUBSCRIPT 15 end_POSTSUBSCRIPT ( italic_g ) ≤ 1.05) is 0.37. Leloudas et al. (2015) suggested the brighter peak absolute magnitudes of the 91T-like sub-class can be attributed to interaction with a dense CSM originating from a non-degenerate companion. The r2subscriptsubscript𝑟2\mathcal{F}_{r_{2}}caligraphic_F start_POSTSUBSCRIPT italic_r start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT-values we find for this sub-class can be interpreted to support this proposition if we assume that the additional flux from the interaction contributes more between 15-40 d than at peak, since r2subscriptsubscript𝑟2\mathcal{F}_{r_{2}}caligraphic_F start_POSTSUBSCRIPT italic_r start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT is calculated from the peak normalised light curves. SN 2020uoo has a weak secondary maximum (r2=subscriptsubscript𝑟2absent\mathcal{F}_{r_{2}}\ =\ caligraphic_F start_POSTSUBSCRIPT italic_r start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT =0.30), but we note that although the GP fit quality is good, this SN passed our light curve coverage criteria due to having sufficient i𝑖iitalic_i-band data on the rise, but it has limited g𝑔gitalic_g- and r𝑟ritalic_r-band data, which could result in a biased estimate of the time of maximum light and therefore Δでるたm15(g)Δでるたsubscript𝑚15𝑔\Delta m_{15}(g)roman_Δでるた italic_m start_POSTSUBSCRIPT 15 end_POSTSUBSCRIPT ( italic_g ). This is unusual since in most cases, ZTF light curves have significantly more data in g𝑔gitalic_g and r𝑟ritalic_r than i𝑖iitalic_i.

It is notable that the 91bg-like SNe Ia appear like a continuous extension of the normal SN Ia population in the bottom left panel of Fig. 6, since there is ongoing debate as to whether 91bg-like SNe Ia have a different physical origin from normal SNe Ia (Taubenberger et al., 2008; Li et al., 2022; Graur, 2023; Harvey et al., 2023). Based solely on the behaviour of r2subscriptsubscript𝑟2\mathcal{F}_{r_{2}}caligraphic_F start_POSTSUBSCRIPT italic_r start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT as a function of Δでるたm15(g)Δでるたsubscript𝑚15𝑔\Delta m_{15}(g)roman_Δでるた italic_m start_POSTSUBSCRIPT 15 end_POSTSUBSCRIPT ( italic_g ), it appears that 91bg-like and normal SNe Ia are members of the same population.

4.1.2 i band

We find that the timing of the secondary maximum is broadly consistent between the r𝑟ritalic_r and i𝑖iitalic_i band (see Fig. 8), suggesting that the secondary maximum in the two filters share a common cause. The correlation we found between Δでるたm15(g)Δでるたsubscript𝑚15𝑔\Delta m_{15}(g)roman_Δでるた italic_m start_POSTSUBSCRIPT 15 end_POSTSUBSCRIPT ( italic_g ) and t2(r)subscript𝑡2𝑟t_{2}(r)italic_t start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_r ) persists for the i𝑖iitalic_i band (see top right panel of Fig. 6), with a Pearson correlation coefficient r𝑟ritalic_r of 0.450.45-0.45- 0.45 at a >5σしぐまabsent5𝜎>5\sigma> 5 italic_σしぐま confidence level for the normal population, and a Pearson correlation coefficient r𝑟ritalic_r of 0.340.34-0.34- 0.34 at a >5σしぐまabsent5𝜎>5\sigma> 5 italic_σしぐま confidence level for the full sample. This is slightly weaker than the correlation in the r𝑟ritalic_r band, but there are many fewer SNe with sufficient i𝑖iitalic_i-band data to fit the secondary maximum.

We also compare i2subscriptsubscript𝑖2\mathcal{F}_{i_{2}}caligraphic_F start_POSTSUBSCRIPT italic_i start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT to r2subscriptsubscript𝑟2\mathcal{F}_{r_{2}}caligraphic_F start_POSTSUBSCRIPT italic_r start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT, and find that i2subscriptsubscript𝑖2\mathcal{F}_{i_{2}}caligraphic_F start_POSTSUBSCRIPT italic_i start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT is higher than r2subscriptsubscript𝑟2\mathcal{F}_{r_{2}}caligraphic_F start_POSTSUBSCRIPT italic_r start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT for the majority of the sample (174 out of 182 that could be fitted in the i𝑖iitalic_i band), as was found by Papadogiannakis et al. (2019a). Since the secondary maximum is known to be stronger in the i𝑖iitalic_i band than in the r𝑟ritalic_r band, a higher integrated flux in this phase range is expected.

The strong correlation between i2subscriptsubscript𝑖2\mathcal{F}_{i_{2}}caligraphic_F start_POSTSUBSCRIPT italic_i start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT and Δでるたm15(g)Δでるたsubscript𝑚15𝑔\Delta m_{15}(g)roman_Δでるた italic_m start_POSTSUBSCRIPT 15 end_POSTSUBSCRIPT ( italic_g ) does not persist for the i𝑖iitalic_i band, shown in the bottom right panel of Fig. 6, where we find r=0.27𝑟0.27r=-0.27italic_r = - 0.27 at a >5σしぐまabsent5𝜎>5\sigma> 5 italic_σしぐま confidence level for the full sample and no significant correlation at all for the normal population alone. The 91T-like SNe Ia in particular show a wide spread of i2subscriptsubscript𝑖2\mathcal{F}_{i_{2}}caligraphic_F start_POSTSUBSCRIPT italic_i start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT values (0.18 i2absentsubscriptsubscript𝑖2absent\leq\mathcal{F}_{i_{2}}\leq≤ caligraphic_F start_POSTSUBSCRIPT italic_i start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT ≤ 0.85 vs. 0.28 r2absentsubscriptsubscript𝑟2absent\leq\mathcal{F}_{r_{2}}\leq≤ caligraphic_F start_POSTSUBSCRIPT italic_r start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT ≤ 0.47, although we note that the uncertainties are larger in the i𝑖iitalic_i band due to the poorer data coverage). Further discussion of the potential causes of the differences in the r𝑟ritalic_r and i𝑖iitalic_i bands can be found in Sect. 5.1.

Refer to caption
Figure 7: Histogram showing the distribution of r2subscriptsubscript𝑟2\mathcal{F}_{r_{2}}caligraphic_F start_POSTSUBSCRIPT italic_r start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT for the ZTF DR2 sample with the flux integral divided by 25 d (shaded orange) or divided by 20 d (dashed orange), compared to the distribution found for PTF and iPTF by Papadogiannakis et al. (2019a) (black). The median of the PTF+iPTF population is 1.5σしぐま𝜎\sigmaitalic_σしぐま higher than the median of the ZTF DR2 population when the integral is divided by 25 d, but the ZTF distribution is consistent with the PTF+iPTF distribution if the integral is divided by 20 d instead.
Refer to caption
Figure 8: The measured values of t2(r)subscript𝑡2𝑟t_{2}(r)italic_t start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_r ) vs. t2(i)subscript𝑡2𝑖t_{2}(i)italic_t start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_i ). We also show a simple linear regression fit (black dashed line) through the full data set.

4.2 Spectral features causing the secondary maximum

Jack et al. (2015) identify Fe ii as the dominant spectral feature that causes the secondary maximum in the i𝑖iitalic_i band. Their models produce Co ii lines and blended Fe ii features (at 6900 Å and 7500 Å) in the r𝑟ritalic_r band, which strengthen around the time of the secondary maximum. However, these features alone are not sufficient to explain the increase in flux during the r𝑟ritalic_r-band shoulder. There is another feature around 6500 Å in the spectra of SN 2014J presented by (Jack et al., 2015, see their fig. 8), which their model was not able to reproduce.

We can see from Fig. 5, which shows the r𝑟ritalic_r- and i𝑖iitalic_i-band wavelength range of averaged spectra of SNe Ia from our cosmological sub-sample, that there is an emission feature in the r𝑟ritalic_r band, red-ward of Si ii 6355 Å  around 6500 Å, which become stronger in SNe Ia with narrower light curves (3x1<03subscript𝑥10-3\leq x_{1}<0- 3 ≤ italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT < 0) as early as 10–20 d after peak. In SNe Ia with broader light curves (0x1<30subscript𝑥130\leq x_{1}<30 ≤ italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT < 3) this occurs later, around 20–30 d after peak. The flux values are dependent on the region used to normalise the spectra, but the line profiles are similar in the r𝑟ritalic_r-band region from similar-to\sim20 d onwards. Since t2(r)subscript𝑡2𝑟t_{2}(r)italic_t start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_r ) scales with x1subscript𝑥1x_{1}italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT (see Sect. 4.1), we suggest that the appearance of these features play a dominant role in the secondary maximum in the r𝑟ritalic_r band, just like the Fe ii feature identified by Jack et al. (2015) causes the secondary maximum in the i𝑖iitalic_i band. There is an additional spectral feature blue-ward of Si ii 6355 Å  around 5850 Å, which appears to decrease in strength between 10–30 d, but then gets stronger and finally peaks in strength between 40–50 d. However, due to the limited number of spectra in these final bins, and because the spectra of SN 2014J presented by Jack et al. (2015) also showing a decreasing flux in this region, we do not trust that this feature plays a dominant role in the appearance of the secondary maximum. In the i𝑖iitalic_i band region of the top panel of Fig. 5 we also clearly see the broad Fe ii feature (between 7200–8000 Å) which was previously identified by Jack et al. (2015). This feature increases in strength between 10–40 d, and per each phase bin, it is stronger for spectra of SNe Ia with smaller x1subscript𝑥1x_{1}italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT values.

In Fig. 9 we show the breakdown of species contributing to each feature for a non local thermodynamic equilbrium (NLTE) radiative transfer model using artis of a W7 explosion model (Nomoto et al., 1984; Shingles et al., 2020, 2022) at 30, 40, and 50 d post explosion (approximately 12, 22, and 32 d past maximum light). From this model, it appears that the predominant IGE contributor to the feature at similar-to\sim6500 Å (which is similar to the unidentified feature in SN 2014J, Jack et al., 2015) is Fe ii. The feature that sits blue-ward of Si ii (around 5850 Å) has a strong contributions of Co iii and Fe iii  which fade away as recombination occurs, explaining why this feature initially fades. Although the region also shows a presence of Fe ii  which get stronger as time progresses, the relative strength of the Fe ii in this spectral feature compared to that at 6500 Å is weaker. The growing contribution of Fe ii in the i𝑖iitalic_i band region between 7200–8000 Åis also clearly visible in these models. We use the W7 model since it has been shown to be able to reproduce a secondary maximum in the i𝑖iitalic_i band and a secondary maximum in the r𝑟ritalic_r band (Jack et al., 2011). However, the choice of assumed explosion model can affect the relative contributions of each ion to the net spectrum and therefore, these line identifications are not entirely reliable.

One of the defining photometric characteristics of 91bg-like SNe Ia is their lack of secondary maxima (Hoeflich et al., 2002; Gonzálezgonz´gonzález-Gaitán et al., 2014; Dhawan et al., 2017; Ashall et al., 2020; Hoogendam et al., 2022). It has been argued that these faint events are faster evolving because they have cooler ejecta, enabling recombination to occur sooner, shifting their secondary maximum to earlier times, causing it to blend with the primary maximum (Kasen, 2006; Blondin et al., 2015; Taubenberger, 2017). We included spectra of SN 1999by (which is a 91bg-like SN Ia) in the bottom panel of Fig. 5. The feature we have tentatively classified as Fe ii, which coincides with the onset of the secondary maximum in normal SNe Ia, starts to come through in the spectra of SN 1999by as early as similar-to\sim3 d post maximum light. This is significantly earlier than for the normal SNe Ia with narrow light curves (3x1<03subscript𝑥10-3\leq x_{1}<0- 3 ≤ italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT < 0) shown in the top panel of Fig. 5. In the future, we aim to get more detailed spectroscopic time series of 91bg-like SNe Ia to determine the range of phases at which the recombination occurs for this sub-class.

Refer to caption
Refer to caption
Refer to caption
Figure 9: Spectra showing the breakdown of species contributing to each feature from a NLTE radiative transfer model of the W7 explosion model at 30 d (top), 40 d (middle) and 50 d (bottom) past explosion.

4.3 Transparency timescale

Refer to caption
Figure 10: Top: We show the distributions of transparency timescales for the whole cosmological sub-sample (grey), including a bimodal Gaussian fit (solid grey line) and its individual components (dashed grey). The arrows show the ranges of transparency timescales predicted by doubledet_2010 (Fink et al., 2010; Kromer et al., 2010), doubledet_2020, (Gronow et al., 2020), doubledet_2021(Gronow et al., 2021), ddt_2010 (Fink, 2010), ddt_2013 (Seitenzahl et al., 2013; Sim et al., 2013), ddt_2014 (Ohlmann et al., 2014), SCH, PDDEL, and DDC (Blondin et al., 2017) explosions models. Bottom: We separate the transparency timescale distribution into SNe Ia with blue local colours ((gz)local 1.0subscript𝑔𝑧𝑙𝑜𝑐𝑎𝑙1.0(g-z)_{local}\ \leq\ 1.0( italic_g - italic_z ) start_POSTSUBSCRIPT italic_l italic_o italic_c italic_a italic_l end_POSTSUBSCRIPT ≤ 1.0, blue) or red local colours ((gz)local> 1.0subscript𝑔𝑧𝑙𝑜𝑐𝑎𝑙1.0(g-z)_{local}\ >\ 1.0( italic_g - italic_z ) start_POSTSUBSCRIPT italic_l italic_o italic_c italic_a italic_l end_POSTSUBSCRIPT > 1.0, red). We show bimodal Gaussian the fits to both distributions, and the individual components making up the fit are shown as dotted lines. The Akaike Information Criterion (AIC) test prefers a bimodal fit for the SNe Ia in red local environments, but it does not show a preference for either the double component fit or single component fit for the SNe Ia population in blue environments, so we also show the single component fit as a dashed line to this population.

The transparency timescale is defined by Jeffery (1999) as the epoch at which the ejecta have an optical depth of unity to γがんま𝛾\gammaitalic_γがんま-rays. For normal SNe Ia, this value ranges between 30–50 d (Scalzo et al., 2014; Dhawan et al., 2018). The transparency timescale is a valuable metric because it has been shown to relate directly to the total ejected mass (Jeffery, 1999; Stritzinger et al., 2006a; Scalzo et al., 2014; Dhawan et al., 2018, 2017), which can be used to constrain the explosion models for SNe Ia.

Papadogiannakis et al. (2019a) were able to calculate pseudo-bolometric light curves for SNe Ia in their sample provided by CSP because these have multi-wavelength coverage. From the pseudo-bolometric light curves they calculate the transparency timescale (t0subscript𝑡0t_{0}italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT). They compare their values of t0subscript𝑡0t_{0}italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT for the CSP sub-sample to their measured t2(r)subscript𝑡2𝑟t_{2}(r)italic_t start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_r ) and r2subscriptsubscript𝑟2\mathcal{F}_{r_{2}}caligraphic_F start_POSTSUBSCRIPT italic_r start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT values, and find a strong correlation with the latter, providing the following linear relation:

t0=44.92(±5.86)×r2+15.00(±2.32),subscript𝑡044.92plus-or-minus5.86subscriptsubscript𝑟215.00plus-or-minus2.32t_{0}=44.92(\pm 5.86)\times\mathcal{F}_{r_{2}}+15.00(\pm 2.32),italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 44.92 ( ± 5.86 ) × caligraphic_F start_POSTSUBSCRIPT italic_r start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT + 15.00 ( ± 2.32 ) , (1)

relating r2subscriptsubscript𝑟2\mathcal{F}_{r_{2}}caligraphic_F start_POSTSUBSCRIPT italic_r start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT to t0subscript𝑡0t_{0}italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT. Due to the offset in our r2subscriptsubscript𝑟2\mathcal{F}_{r_{2}}caligraphic_F start_POSTSUBSCRIPT italic_r start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT distribution compared to that presented in Papadogiannakis et al. (2019a), see Sect. 4.1, we cannot directly use the above linear correlation to estimate t0subscript𝑡0t_{0}italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT for our sample. Instead, we recalculate r2subscriptsubscript𝑟2\mathcal{F}_{r_{2}}caligraphic_F start_POSTSUBSCRIPT italic_r start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT by dividing the integrated flux between 15–40 d by 20 d (which we will refer to as ¯r2subscript¯subscript𝑟2\mathcal{\bar{F}}_{r_{2}}over¯ start_ARG caligraphic_F end_ARG start_POSTSUBSCRIPT italic_r start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT), ensuring that the distribution matches that of Papadogiannakis et al. (2019a) as shown in Fig. 7.

Using Eq. 1 and ¯r2subscript¯subscript𝑟2\mathcal{\bar{F}}_{r_{2}}over¯ start_ARG caligraphic_F end_ARG start_POSTSUBSCRIPT italic_r start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT, we estimate the t0subscript𝑡0t_{0}italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT values for our sample. In the top panel of Fig. 10 we show the distributions of t0subscript𝑡0t_{0}italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT for the cosmological SN Ia sample (see definition in Sect 3.6). The t0subscript𝑡0t_{0}italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT distribution does not show a single Gaussian distribution but instead appears bimodal, with a stronger peak at longer t0subscript𝑡0t_{0}italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT than the weaker peak at shorter t0subscript𝑡0t_{0}italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT values. This is similar to the bimodal x1subscript𝑥1x_{1}italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT distribution identified in Nicolas et al. (2021) and Ginolin et al. (2024).

To estimate the properties of this bimodal distribution, we fit the t0subscript𝑡0t_{0}italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT values with a double Gaussian model. The fit is defined as P(t0)=r𝒩(μみゅーhigh,σしぐまhigh)+(1r)𝒩(μみゅーlow,σしぐまlow)𝑃subscript𝑡0𝑟𝒩subscript𝜇𝑖𝑔subscript𝜎𝑖𝑔1𝑟𝒩subscript𝜇𝑙𝑜𝑤subscript𝜎𝑙𝑜𝑤P(t_{0})=r\mathscr{N}(\mu_{high},\sigma_{high})+(1-r)\mathscr{N}(\mu_{low},% \sigma_{low})italic_P ( italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) = italic_r script_N ( italic_μみゅー start_POSTSUBSCRIPT italic_h italic_i italic_g italic_h end_POSTSUBSCRIPT , italic_σしぐま start_POSTSUBSCRIPT italic_h italic_i italic_g italic_h end_POSTSUBSCRIPT ) + ( 1 - italic_r ) script_N ( italic_μみゅー start_POSTSUBSCRIPT italic_l italic_o italic_w end_POSTSUBSCRIPT , italic_σしぐま start_POSTSUBSCRIPT italic_l italic_o italic_w end_POSTSUBSCRIPT ), where P(t0)𝑃subscript𝑡0P(t_{0})italic_P ( italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) is the probability distribution, 𝒩𝒩\mathscr{N}script_N is the Gaussian distribution, r𝑟ritalic_r is the amplitude of the high t0subscript𝑡0t_{0}italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT Gaussian, μみゅー𝜇\muitalic_μみゅー is the mean and σしぐま𝜎\sigmaitalic_σしぐま is the standard deviation of each Gaussian. We summarise the results of this fit in Table 3. We find r= 0.74± 0.05𝑟plus-or-minus0.740.05r\ =\ 0.74\ \pm\ 0.05italic_r = 0.74 ± 0.05, suggesting that the higher t0subscript𝑡0t_{0}italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT mode is the dominant mode over those SNe Ia with lower t0subscript𝑡0t_{0}italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT values. This result is consistent with what was found for the bimodal x1subscript𝑥1x_{1}italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT distribution in Nicolas et al. (2021) (r= 0.755± 0.05𝑟plus-or-minus0.7550.05r\ =\ 0.755\ \pm\ 0.05italic_r = 0.755 ± 0.05), but not with Ginolin et al. (2024), r= 0.60± 0.07𝑟plus-or-minus0.600.07r\ =\ 0.60\ \pm\ 0.07italic_r = 0.60 ± 0.07).

Table 3: Summary of bimodal fit to the t0subscript𝑡0t_{0}italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT distribution of the cosmological sample discussed in Section 4.3.
Parameter Result
r𝑟ritalic_r 0.74± 0.05plus-or-minus0.740.050.74\ \pm\ 0.050.74 ± 0.05
μみゅーlowsubscript𝜇𝑙𝑜𝑤\mu_{low}italic_μみゅー start_POSTSUBSCRIPT italic_l italic_o italic_w end_POSTSUBSCRIPT 30.9± 0.4plus-or-minus30.90.430.9\ \pm\ 0.430.9 ± 0.4
σしぐまlowsubscript𝜎𝑙𝑜𝑤\sigma_{low}italic_σしぐま start_POSTSUBSCRIPT italic_l italic_o italic_w end_POSTSUBSCRIPT 1.8± 0.4plus-or-minus1.80.41.8\ \pm\ 0.41.8 ± 0.4
μみゅーhighsubscript𝜇𝑖𝑔\mu_{high}italic_μみゅー start_POSTSUBSCRIPT italic_h italic_i italic_g italic_h end_POSTSUBSCRIPT 35.9± 0.1plus-or-minus35.90.135.9\ \pm\ 0.135.9 ± 0.1
σしぐまhighsubscript𝜎𝑖𝑔\sigma_{high}italic_σしぐま start_POSTSUBSCRIPT italic_h italic_i italic_g italic_h end_POSTSUBSCRIPT 1.7± 0.1plus-or-minus1.70.11.7\ \pm\ 0.11.7 ± 0.1

We also compare the distribution of t0subscript𝑡0t_{0}italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT in our sample to theoretical explosion models, akin to fig. 5 in Papadogiannakis et al. (2019a). We fit the following energy deposition function to the bolometric light curves:

Edepsubscript𝐸𝑑𝑒𝑝\displaystyle E_{dep}italic_E start_POSTSUBSCRIPT italic_d italic_e italic_p end_POSTSUBSCRIPT =λらむだNiNNiQNi,γがんまexp(λらむだNit)absentsubscript𝜆𝑁𝑖subscript𝑁𝑁𝑖subscript𝑄𝑁𝑖𝛾subscript𝜆𝑁𝑖𝑡\displaystyle=\lambda_{Ni}N_{Ni}Q_{Ni,\gamma}\exp{(-\lambda_{Ni}t})= italic_λらむだ start_POSTSUBSCRIPT italic_N italic_i end_POSTSUBSCRIPT italic_N start_POSTSUBSCRIPT italic_N italic_i end_POSTSUBSCRIPT italic_Q start_POSTSUBSCRIPT italic_N italic_i , italic_γがんま end_POSTSUBSCRIPT roman_exp ( - italic_λらむだ start_POSTSUBSCRIPT italic_N italic_i end_POSTSUBSCRIPT italic_t ) (2)
+λらむだCoλらむだNiλらむだNiλらむだCo[exp(λらむだCot)exp(λらむだNit)]subscript𝜆𝐶𝑜subscript𝜆𝑁𝑖subscript𝜆𝑁𝑖subscript𝜆𝐶𝑜delimited-[]subscript𝜆𝐶𝑜𝑡subscript𝜆𝑁𝑖𝑡\displaystyle+\ \lambda_{Co}\frac{\lambda_{Ni}}{\lambda_{Ni}-\lambda_{Co}}[% \exp{(-\lambda_{Co}t})-\exp{(-\lambda_{Ni}t})]+ italic_λらむだ start_POSTSUBSCRIPT italic_C italic_o end_POSTSUBSCRIPT divide start_ARG italic_λらむだ start_POSTSUBSCRIPT italic_N italic_i end_POSTSUBSCRIPT end_ARG start_ARG italic_λらむだ start_POSTSUBSCRIPT italic_N italic_i end_POSTSUBSCRIPT - italic_λらむだ start_POSTSUBSCRIPT italic_C italic_o end_POSTSUBSCRIPT end_ARG [ roman_exp ( - italic_λらむだ start_POSTSUBSCRIPT italic_C italic_o end_POSTSUBSCRIPT italic_t ) - roman_exp ( - italic_λらむだ start_POSTSUBSCRIPT italic_N italic_i end_POSTSUBSCRIPT italic_t ) ]
×[QCo,γがんま+QCo,e+(1exp((t/t0)2))],absentdelimited-[]subscript𝑄𝐶𝑜𝛾subscript𝑄𝐶𝑜superscript𝑒1superscript𝑡subscript𝑡02\displaystyle\times\ [Q_{Co,\gamma}+Q_{Co,e^{+}}(1-\exp{((-t/t_{0})^{2})})],× [ italic_Q start_POSTSUBSCRIPT italic_C italic_o , italic_γがんま end_POSTSUBSCRIPT + italic_Q start_POSTSUBSCRIPT italic_C italic_o , italic_e start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT end_POSTSUBSCRIPT ( 1 - roman_exp ( ( - italic_t / italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) ) ] ,

where λらむだNisubscript𝜆𝑁𝑖\lambda_{Ni}italic_λらむだ start_POSTSUBSCRIPT italic_N italic_i end_POSTSUBSCRIPT and λらむだCosubscript𝜆𝐶𝑜\lambda_{Co}italic_λらむだ start_POSTSUBSCRIPT italic_C italic_o end_POSTSUBSCRIPT are the inverse of the e-folding times of nickel and cobalt (1/8.8 and 1/111.3 d-1, respectively), NNisubscript𝑁𝑁𝑖N_{Ni}italic_N start_POSTSUBSCRIPT italic_N italic_i end_POSTSUBSCRIPT represents the number of nickel atoms present, which is calculated from the mass of 56Ni, which is derived from the peak of the bolometric light curve (Arnett, 1982). QNi,γがんまsubscript𝑄𝑁𝑖𝛾Q_{Ni,\gamma}italic_Q start_POSTSUBSCRIPT italic_N italic_i , italic_γがんま end_POSTSUBSCRIPT and QCo,γがんまsubscript𝑄𝐶𝑜𝛾Q_{Co,\gamma}italic_Q start_POSTSUBSCRIPT italic_C italic_o , italic_γがんま end_POSTSUBSCRIPT represent the energy released from γがんま𝛾\gammaitalic_γがんま-ray decays of 56Ni\rightarrow 56Co and 56Co\rightarrow 56Fe, respectively (1.75 and 3.61 MeV) and QCo,e+subscript𝑄𝐶𝑜superscript𝑒Q_{Co,e^{+}}italic_Q start_POSTSUBSCRIPT italic_C italic_o , italic_e start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT end_POSTSUBSCRIPT is the energy release of the positron decay of 56Co\rightarrow 56Fe (0.12 MeV).

We consider two types of explosion models from the Heidelberg Supernova Model Archive (HESMA; Kromer et al., 2017), double detonation and delayed detonation models. Specifically, we include the ranges predicted by doubledet_2010 (Fink et al., 2010; Kromer et al., 2010), doubledet_2020, (Gronow et al., 2020), doubledet_2021222The models presented in Gronow et al. (2021) extend the range of core and He shell masses and were produced to explain the variations seen in the bolometric magnitudes of SNe Ia. (Gronow et al., 2021), ddt_2010 (Fink, 2010), ddt_2013 (Seitenzahl et al., 2013; Sim et al., 2013), and ddt_2014 (Ohlmann et al., 2014). The t0subscript𝑡0t_{0}italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ranges predicted by these models are shown in the top panel of Fig. 10. We also include model predictions from Blondin et al. (2017) for pure central detonations of sub-Mch WDs (SCH), pulsational Mch delayed-detonations (PDDEL), and standard Mch delayed-detonations (DDC). We checked the t0subscript𝑡0t_{0}italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT distributions for gravitationally confined detonation (Seitenzahl et al., 2016; Lach et al., 2022) and merger models (Pakmor et al., 2010, 2012; Kromer et al., 2013b, 2016), but the shortest t0subscript𝑡0t_{0}italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT value produced by these models was 45 d and their t0subscript𝑡0t_{0}italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT values extended out to 70 d, which is not consistent with any of our observations.

We find that the t0subscript𝑡0t_{0}italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT predictions from the doubledet models are consistent with 97 per cent of our cosmological sub-sample, whereas only 28 per cent are consistent with the predictions from the ddt models. From the models presented by Blondin et al. (2017), the t0subscript𝑡0t_{0}italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT values predicted by the SCH models are consistent with the largest percentage of our sample (26 per cent). The other two explosion models, DDC and PDDEL, are consistent with equal percentages of the sample (8 per cent).

We also check which explosion model t0subscript𝑡0t_{0}italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT predictions are most consistent with the common SN Ia sub-types. The 91bg-like sub-class spans 22.2 t0absentsubscript𝑡0absent\leq\ t_{0}\leq≤ italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ≤ 35.3 d with a median of 26.1 d, and the 91T-like sub-class spans 31.0 t0absentsubscript𝑡0absent\leq\ t_{0}\leq≤ italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ≤ 41.7 d with a median of 38.1 d. None of the 91bg-like SNe Ia are consistent with the predictions from the ddt explosion models, 18 per cent are consistent with the doubledet explosion models, and only 3 per cent are consistent with the SCH explosion models. On the other hand, 100 per cent of the transparency timescales of the 91T-like SN Ia population are consistent with the doubledet, 76 percent are consistent with ddt, 40 per cent are consistent with both PDDEL and DDC, and only 3 per cent are consistent with the SCH explosion models.

In order to test whether the low t0subscript𝑡0t_{0}italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT component originates from SNe Ia in particular host galaxy environments, we bin the distribution into locally blue or red environments ((gz)local 1.0subscript𝑔𝑧𝑙𝑜𝑐𝑎𝑙1.0(g-z)_{local}\ \leq\ 1.0( italic_g - italic_z ) start_POSTSUBSCRIPT italic_l italic_o italic_c italic_a italic_l end_POSTSUBSCRIPT ≤ 1.0 or (gz)local> 1.0subscript𝑔𝑧𝑙𝑜𝑐𝑎𝑙1.0(g-z)_{local}\ >\ 1.0( italic_g - italic_z ) start_POSTSUBSCRIPT italic_l italic_o italic_c italic_a italic_l end_POSTSUBSCRIPT > 1.0, shown in the bottom panel of Fig. 10). We fit the resulting two distributions with either a single or bimodal Gaussian, the only priors being that the amplitudes of the Gaussians cannot be negative and they cannot share the same mean. We use the Akaike Information Criterion (AIC, Burnham & Anderson, 2004), which is a good test for comparing models because it penalises for additional parameters, to determine whether the population is better described by a single or double component Gaussian. We find that the SNe Ia in locally blue environments can be fit by either a single or double component, with the AIC values differing by <2absent2<2< 2, which is the standard cut-off point for determining whether one model is preferred over the other (Burnham & Anderson, 2004). The distribution of SNe Ia in locally red environments on the other hand shows a strong preference for two components, with an AIC value 39 units lower than for the single component fit.

4.4 Total ejected mass

One of the strongest constraints for different progenitor scenarios is the total ejected mass (Mejsubscript𝑀ejM_{\textnormal{ej}}italic_M start_POSTSUBSCRIPT ej end_POSTSUBSCRIPT), since most explosion scenarios can be placed in one of two categories: Mchsubscript𝑀𝑐M_{ch}italic_M start_POSTSUBSCRIPT italic_c italic_h end_POSTSUBSCRIPT or sub-Mchsubscript𝑀𝑐M_{ch}italic_M start_POSTSUBSCRIPT italic_c italic_h end_POSTSUBSCRIPT. Unfortunately, we cannot calculate the total ejected mass directly since this requires bolometric light curves (e.g. Arnett, 1982; Jeffery, 1999), which are not available for the ZTF sample where we only have g𝑔gitalic_g, r𝑟ritalic_r, and i𝑖iitalic_i coverage.

However, we can use our t0subscript𝑡0t_{0}italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT values (which were derived from ¯r2subscript¯subscript𝑟2\mathcal{\bar{F}}_{r_{2}}over¯ start_ARG caligraphic_F end_ARG start_POSTSUBSCRIPT italic_r start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT) to estimate Mejsubscript𝑀ejM_{\textnormal{ej}}italic_M start_POSTSUBSCRIPT ej end_POSTSUBSCRIPT using the equation;

Mej/M=1.38(1/3q)(ve3000km s1)2(t036.80d)2subscript𝑀ejsubscriptMdirect-product1.3813𝑞superscriptsubscript𝑣𝑒3000superscriptkm s12superscriptsubscript𝑡036.80d2M_{\textnormal{ej}}/\textnormal{M}_{\odot}=1.38\cdot\left(\frac{1/3}{q}\right)% \cdot\left(\frac{v_{e}}{3000\ \textnormal{km s}^{-1}}\right)^{2}\cdot\left(% \frac{t_{0}}{36.80\ \textnormal{d}}\right)^{2}italic_M start_POSTSUBSCRIPT ej end_POSTSUBSCRIPT / M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT = 1.38 ⋅ ( divide start_ARG 1 / 3 end_ARG start_ARG italic_q end_ARG ) ⋅ ( divide start_ARG italic_v start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT end_ARG start_ARG 3000 km s start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT end_ARG ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⋅ ( divide start_ARG italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG start_ARG 36.80 d end_ARG ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT (3)

which was presented in this form by Papadogiannakis et al. (2019a) but originally derived by Jeffery (1999). In Eq. 3, q𝑞qitalic_q is a qualitative description of the distribution of the material of the ejecta (a high value for q𝑞qitalic_q implies a more centrally concentrated ejecta) and vesubscript𝑣𝑒v_{e}italic_v start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT is the e-folding velocity assuming an exponential density profile. We adopt the same values as Papadogiannakis et al. (2019a): q= 1/3𝑞13q\ =\ 1/3italic_q = 1 / 3 and v2= 3000subscript𝑣23000v_{2}\ =\ 3000italic_v start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = 3000 km s-1, leaving t0subscript𝑡0t_{0}italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT as the only variable. These assumed values are approximately valid for normal SNe Ia, with more or less luminous SNe Ia having higher or lower e-folding velocities, respectively. We calculate Mejsubscript𝑀ejM_{\textnormal{ej}}italic_M start_POSTSUBSCRIPT ej end_POSTSUBSCRIPT for the SNe Ia in our cosmological sub-sample using the t0subscript𝑡0t_{0}italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT values calculated in Sect. 4.3 (see Fig. 11). Although they are a member of the cosmological sub-sample, we highlight the 91T-like SN Ia population in Fig. 11, which sit at the upper end of the Mejsubscript𝑀ejM_{\textnormal{ej}}italic_M start_POSTSUBSCRIPT ej end_POSTSUBSCRIPT distribution. We also present the calculated Mejsubscript𝑀ejM_{\textnormal{ej}}italic_M start_POSTSUBSCRIPT ej end_POSTSUBSCRIPT-values for the 91bg-like population in Fig. 11, but because the assumptions on q𝑞qitalic_q and vesubscript𝑣𝑒v_{e}italic_v start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT are not necessarily valid for this sub-populations, their measured ejecta masses may be less accurate.

Scalzo et al. (2014) studied SNe Ia with multi-wavelength coverage (enabling the construction of bolometric light curves) and measured the total ejected mass using a comparable method to Papadogiannakis et al. (2019a) by measuring t0subscript𝑡0t_{0}italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT from the bolometric light curve by fitting a radioactive decay energy deposition curve to the tail of the observations. However, unlike Papadogiannakis et al. (2019a), Scalzo et al. (2014) infer the ejected mass in a Bayesian context using a semi-analytic model of the ejecta. Scalzo et al. (2014) found a strong correlation between the Mejsubscript𝑀ejM_{\textnormal{ej}}italic_M start_POSTSUBSCRIPT ej end_POSTSUBSCRIPT and x1subscript𝑥1x_{1}italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT, providing the following equation;

Mej/M=(1.253±0.022)+(0.172±0.021)x1subscript𝑀ejsubscriptMdirect-productplus-or-minus1.2530.022plus-or-minus0.1720.021subscript𝑥1M_{\textnormal{ej}}/\textnormal{M}_{\odot}=(1.253\pm 0.022)+(0.172\pm 0.021)x_% {1}italic_M start_POSTSUBSCRIPT ej end_POSTSUBSCRIPT / M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT = ( 1.253 ± 0.022 ) + ( 0.172 ± 0.021 ) italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT (4)

which we use as a secondary method to derive Mejsubscript𝑀ejM_{\textnormal{ej}}italic_M start_POSTSUBSCRIPT ej end_POSTSUBSCRIPT. We note that the Mejsubscript𝑀ejM_{\textnormal{ej}}italic_M start_POSTSUBSCRIPT ej end_POSTSUBSCRIPT-values derived using this equation are only reliable for our cosmological sub-sample of the SNe Ia since they have reliable x1subscript𝑥1x_{1}italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT values (see Sect. 3.6), but we do also show the 91bg-like SN Ia population for comparison in Fig. 11.

Although in general we find a slightly higher ejected mass from ¯r2subscript¯subscript𝑟2\mathcal{\bar{F}}_{r_{2}}over¯ start_ARG caligraphic_F end_ARG start_POSTSUBSCRIPT italic_r start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT than x1subscript𝑥1x_{1}italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT (see Fig. 11), the median ejected masses derived from the two methods are consistent with each other (with a median calculated from ¯r2subscript¯subscript𝑟2\mathcal{\bar{F}}_{r_{2}}over¯ start_ARG caligraphic_F end_ARG start_POSTSUBSCRIPT italic_r start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT of 1.3± 0.3plus-or-minus1.30.31.3\ \pm\ 0.31.3 ± 0.3 Msubscript𝑀direct-productM_{\odot}italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT and a median calculated from x1subscript𝑥1x_{1}italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT of 1.2± 0.2plus-or-minus1.20.21.2\ \pm\ 0.21.2 ± 0.2 Msubscript𝑀direct-productM_{\odot}italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT). This is despite the fact that the correlation used to derive t0subscript𝑡0t_{0}italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT from ¯r2subscript¯subscript𝑟2\mathcal{\bar{F}}_{r_{2}}over¯ start_ARG caligraphic_F end_ARG start_POSTSUBSCRIPT italic_r start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT presented by Papadogiannakis et al. (2019a) is based on 17 SNe Ia and includes two faint 91bg-like SNe Ia, whereas the correlation presented by Scalzo et al. (2014) is based on 21 SNe Ia and includes normal and bright SNe Ia, such as 91T-like or 03fg-like SNe Ia. These two studies are sampling opposite extremes of the absolute magnitude distribution of SNe Ia, and it has not been established whether these extremes are members of the same population as normal SNe Ia. Although 91T-like SNe Ia have very similar properties to normal SNe Ia, studies have suggested that both 91bg- and 03fg-like SNe Ia could originate from different progenitor scenarios or explosion mechanisms than normal SNe Ia (Stritzinger et al., 2006b; Taubenberger et al., 2008, 2013; Scalzo et al., 2014; Taubenberger et al., 2019; Graur et al., 2023; Hoogendam et al., 2023). Moreover, the method implemented by Papadogiannakis et al. (2019a) to derive Mejsubscript𝑀ejM_{\textnormal{ej}}italic_M start_POSTSUBSCRIPT ej end_POSTSUBSCRIPT from t0subscript𝑡0t_{0}italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT differs from that implemented by Scalzo et al. (2014), with Scalzo et al. (2014) using a Bayesian framework. However, since both methods use the bolometric light curves to derive t0subscript𝑡0t_{0}italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT, it is not surprising that the resulting distributions are consistent with each other, but it does suggest that the correlation used to derive t0subscript𝑡0t_{0}italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT from r2subscriptsubscript𝑟2\mathcal{F}_{r_{2}}caligraphic_F start_POSTSUBSCRIPT italic_r start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT provides reliable results.

Refer to caption
Figure 11: A comparison between the total ejected mass derived from ¯r2subscript¯subscript𝑟2\mathcal{\bar{F}}_{r_{2}}over¯ start_ARG caligraphic_F end_ARG start_POSTSUBSCRIPT italic_r start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT and x1subscript𝑥1x_{1}italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT for the cosmology sub-sample (which contains the 91T-like SNe Ia) as well as the 91bg-like SN Ia population. The black line is the line of equality. The median total ejected mass calculated from ¯r2subscript¯subscript𝑟2\mathcal{\bar{F}}_{r_{2}}over¯ start_ARG caligraphic_F end_ARG start_POSTSUBSCRIPT italic_r start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT is 1.3± 0.3plus-or-minus1.30.31.3\ \pm\ 0.31.3 ± 0.3 Msubscript𝑀direct-productM_{\odot}italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT and the median when calculated from x1subscript𝑥1x_{1}italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT is 1.2± 0.2plus-or-minus1.20.21.2\ \pm\ 0.21.2 ± 0.2 Msubscript𝑀direct-productM_{\odot}italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT. We find a range of 0.64 – 1.91 Msubscript𝑀direct-productM_{\odot}italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT for Mejsubscript𝑀ejM_{\textnormal{ej}}italic_M start_POSTSUBSCRIPT ej end_POSTSUBSCRIPT for the normal SNe Ia in the cosmological sample, ignoring the single outlier at 2.24 Msubscript𝑀direct-productM_{\odot}italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT.

5 Discussion

In this section we discuss the differences between the r𝑟ritalic_r and i𝑖iitalic_i band secondary maxima. We also analyse a potential non-linearity we find in the correlation between x1subscript𝑥1x_{1}italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT and r2subscriptsubscript𝑟2\mathcal{F}_{r_{2}}caligraphic_F start_POSTSUBSCRIPT italic_r start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT, and compare this to recent literature results which find a similar trend with Hubble residuals. We briefly discuss the potential of using the secondary maximum in the r𝑟ritalic_r band for standardisation. We introduce TURTLS radiative transfer model results that enable us to estimate the total mass of 56Ni produced in the explosion, and use this in combination with the results presented in the previous section to try to constrain the dominant parameter affecting the secondary maximum in the r𝑟ritalic_r band.

5.1 Differences between the r𝑟ritalic_r and i𝑖iitalic_i band secondary maximum

In Sect. 4.1 we presented the strength and timing of the secondary maximum in the r𝑟ritalic_r and i𝑖iitalic_i bands. Whilst we found the timing of the secondary maximum to be consistent between the r𝑟ritalic_r and i𝑖iitalic_i bands, the integrated flux under the secondary maximum is larger in the i𝑖iitalic_i band. Furthermore, we find strong correlations between Δでるたm15(g)Δでるたsubscript𝑚15𝑔\Delta m_{15}(g)roman_Δでるた italic_m start_POSTSUBSCRIPT 15 end_POSTSUBSCRIPT ( italic_g ) and t2(r)subscript𝑡2𝑟t_{2}(r)italic_t start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_r ), t2(i)subscript𝑡2𝑖t_{2}(i)italic_t start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_i ), and r2subscriptsubscript𝑟2\mathcal{F}_{r_{2}}caligraphic_F start_POSTSUBSCRIPT italic_r start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT but not with i2subscriptsubscript𝑖2\mathcal{F}_{i_{2}}caligraphic_F start_POSTSUBSCRIPT italic_i start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT.

The secondary maximum is known to be more prominent in the i𝑖iitalic_i band than in the r𝑟ritalic_r band (Kasen, 2006). The integrated flux is normalised to peak, but because the secondary maximum is stronger in the i𝑖iitalic_i band relative to peak, it is not surprising that we find i2subscriptsubscript𝑖2\mathcal{F}_{i_{2}}caligraphic_F start_POSTSUBSCRIPT italic_i start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT to be higher than r2subscriptsubscript𝑟2\mathcal{F}_{r_{2}}caligraphic_F start_POSTSUBSCRIPT italic_r start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT in general.

It is notable that the strong correlation seen between r2subscriptsubscript𝑟2\mathcal{F}_{r_{2}}caligraphic_F start_POSTSUBSCRIPT italic_r start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT and Δでるたm15(g)Δでるたsubscript𝑚15𝑔\Delta m_{15}(g)roman_Δでるた italic_m start_POSTSUBSCRIPT 15 end_POSTSUBSCRIPT ( italic_g ) does not persist for the i𝑖iitalic_i band. The strong correlation in the r𝑟ritalic_r band can be attributed to the nickel mass being one of the main drivers of the strength of the secondary maximum (Kasen, 2006). Whilst the integrated flux under the secondary maximum is normalised to peak and should not show a trend with the peak magnitude (driven by the nickelmass, Arnett, 1982), a higher mass of nickel also leads to a higher temperature. This higher temperature can result in a higher opacity of the ejecta that increases the timescale of the evolution of the light curve (Mazzali et al., 2001; Hoflich & Khokhlov, 1996). Therefore, if more nickel is produced in the explosion, we expect a smaller Δでるたm15(g)Δでるたsubscript𝑚15𝑔\Delta m_{15}(g)roman_Δでるた italic_m start_POSTSUBSCRIPT 15 end_POSTSUBSCRIPT ( italic_g ) and a more prominent secondary maximum, which is what we see in the r𝑟ritalic_r band.

The additional scatter seen in the i𝑖iitalic_i band for i2subscriptsubscript𝑖2\mathcal{F}_{i_{2}}caligraphic_F start_POSTSUBSCRIPT italic_i start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT compared to in the r band could imply that there is another factor influencing the strength of the secondary maximum, which has a larger effect in the i𝑖iitalic_i band than in the r𝑟ritalic_r band, and does not correlate with the decline rate. Kasen (2006) found that changing the metallicity affects the secondary maximum whilst having little to no impact on the primary maximum. This could explain the spread seen in the strengths of the secondary maxima in the i𝑖iitalic_i band for a constant decline rate in the bottom right panel of Fig. 6. We speculate that the reason this effect is more visible in the i𝑖iitalic_i band than in the r𝑟ritalic_r band is because it impacts the strength of the secondary maximum but not the decline rate post-maximum. With the secondary maximum being weaker in the r𝑟ritalic_r band, the overall decline rate will make up a larger contribution of r2subscriptsubscript𝑟2\mathcal{F}_{r_{2}}caligraphic_F start_POSTSUBSCRIPT italic_r start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT than for i2subscriptsubscript𝑖2\mathcal{F}_{i_{2}}caligraphic_F start_POSTSUBSCRIPT italic_i start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT, where the flux in the secondary bump dominates.

5.2 Non-linear correlation between x1subscript𝑥1x_{1}italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT and r2subscriptsubscript𝑟2\mathcal{F}_{r_{2}}caligraphic_F start_POSTSUBSCRIPT italic_r start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT

In Sect. 4.1 we found a strong correlation between r2subscriptsubscript𝑟2\mathcal{F}_{r_{2}}caligraphic_F start_POSTSUBSCRIPT italic_r start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT and Δでるたm15(g)Δでるたsubscript𝑚15𝑔\Delta m_{15}(g)roman_Δでるた italic_m start_POSTSUBSCRIPT 15 end_POSTSUBSCRIPT ( italic_g ), supporting the findings of Papadogiannakis et al. (2019a). Here we test whether this correlation extends to r2subscriptsubscript𝑟2\mathcal{F}_{r_{2}}caligraphic_F start_POSTSUBSCRIPT italic_r start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT with x1subscript𝑥1x_{1}italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT. We opted to use Δでるたm15(g)Δでるたsubscript𝑚15𝑔\Delta m_{15}(g)roman_Δでるた italic_m start_POSTSUBSCRIPT 15 end_POSTSUBSCRIPT ( italic_g ) for the majority of this study because it can be measured from our GP fits, enabling us to measure it for all sub-types of SNe Ia, whereas x1subscript𝑥1x_{1}italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT relies on a reliable SALT2.4 fit, and since SALT2.4 templates are predominantly trained on normal SNe Ia it breaks down for more peculiar sub-types. This means that we can only reliably use x1subscript𝑥1x_{1}italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT values for our cosmological sub-sample.

We plot r2subscriptsubscript𝑟2\mathcal{F}_{r_{2}}caligraphic_F start_POSTSUBSCRIPT italic_r start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT as a function of x1subscript𝑥1x_{1}italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT for the cosmological sub-sample in Fig. 12 and as for r2subscriptsubscript𝑟2\mathcal{F}_{r_{2}}caligraphic_F start_POSTSUBSCRIPT italic_r start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT and Δでるたm15(g)Δでるたsubscript𝑚15𝑔\Delta m_{15}(g)roman_Δでるた italic_m start_POSTSUBSCRIPT 15 end_POSTSUBSCRIPT ( italic_g ), we find a >5σしぐまabsent5𝜎>5\sigma> 5 italic_σしぐま correlation between these parameters. Ginolin et al. (2024) find a strongly non-linear stretch-magnitude relationship (a change of slope at x10superscriptsubscript𝑥10x_{1}^{0}italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT, with x10=0.53± 0.09superscriptsubscript𝑥10plus-or-minus0.530.09x_{1}^{0}\ =\ -0.53\ \pm\ 0.09italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT = - 0.53 ± 0.09). Senzel et al. (2024) find a similar result with the local host colour and x1subscript𝑥1x_{1}italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT. Motivated by these studies, we test for similar non-linearity between r2subscriptsubscript𝑟2\mathcal{F}_{r_{2}}caligraphic_F start_POSTSUBSCRIPT italic_r start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT and x1subscript𝑥1x_{1}italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT by fitting the sample with a ‘broken-line’:

r2={m1x1+c1ifx1x10m2x1+c2ifx1>x10subscriptsubscript𝑟2casessubscript𝑚1subscript𝑥1subscript𝑐1ifsubscript𝑥1superscriptsubscript𝑥10subscript𝑚2subscript𝑥1subscript𝑐2ifsubscript𝑥1superscriptsubscript𝑥10\mathcal{F}_{r_{2}}=\begin{cases}m_{1}x_{1}+c_{1}\quad&\text{if}\,\ x_{1}\leq x% _{1}^{0}\\ m_{2}x_{1}+c_{2}\quad&\text{if}\,\ x_{1}>x_{1}^{0}\\ \end{cases}caligraphic_F start_POSTSUBSCRIPT italic_r start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT = { start_ROW start_CELL italic_m start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT + italic_c start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_CELL start_CELL if italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ≤ italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT end_CELL end_ROW start_ROW start_CELL italic_m start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT + italic_c start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_CELL start_CELL if italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT > italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT end_CELL end_ROW (5)

where x10superscriptsubscript𝑥10x_{1}^{0}italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT is the x1subscript𝑥1x_{1}italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT value at which the line breaks and this is a free parameter which is allowed the vary over the full x1subscript𝑥1x_{1}italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT range of the sample, m1subscript𝑚1m_{1}italic_m start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT and m2subscript𝑚2m_{2}italic_m start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT are the slopes and c1subscript𝑐1c_{1}italic_c start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT and c2subscript𝑐2c_{2}italic_c start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT are the y-intercepts before and after the line break (x10superscriptsubscript𝑥10x_{1}^{0}italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT), respectively. We find a split in the population atx10=0.5± 0.2superscriptsubscript𝑥10plus-or-minus0.50.2x_{1}^{0}\ =\ -0.5\ \pm\ 0.2italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT = - 0.5 ± 0.2, and the broken-line fit is also shown in Fig. 12. Using AIC as a test to compare the models, the ‘broken-line’ model is statistically preferred (difference of 52 AIC units). Interestingly, the best-fit value of the stretch split (x10=0.5± 0.2superscriptsubscript𝑥10plus-or-minus0.50.2x_{1}^{0}\ =\ -0.5\ \pm\ 0.2italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT = - 0.5 ± 0.2) is consistent with what was found by Ginolin et al. (2024), (x10=0.49± 0.06superscriptsubscript𝑥10plus-or-minus0.490.06x_{1}^{0}=-0.49\ \pm\ 0.06italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT = - 0.49 ± 0.06) where they argue that the αあるふぁ𝛼\alphaitalic_αあるふぁ term in SN Ia standardisation is non-linear, with the low stretch mode only originating from older populations and the high stretch mode containing SNe Ia from both old and young populations (Nicolas et al., 2021).

Refer to caption
Figure 12: r2subscriptsubscript𝑟2\mathcal{F}_{r_{2}}caligraphic_F start_POSTSUBSCRIPT italic_r start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT as a function of x1subscript𝑥1x_{1}italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT for SNe Ia in our cosmological sub-sample. We fit the population with a broken line (blue) and linear fit (black, dashed). We find the best-fit value of the stretch split to be at x10=0.5± 0.2superscriptsubscript𝑥10plus-or-minus0.50.2x_{1}^{0}\ =\ -0.5\ \pm\ 0.2italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT = - 0.5 ± 0.2. The blue points show the data binned by stretch, taking the mean and standard error of the mean in each bin.

5.3 The secondary maximum in the r𝑟ritalic_r band for standardisation

We also test the secondary maximum in the r𝑟ritalic_r band as an alternative standardisation parameter in the absence of x1subscript𝑥1x_{1}italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT, as Shariff et al. (2016) did for the secondary maximum in the J𝐽Jitalic_J band. To this end we use colour-corrected Hubble residuals (not corrected for x1subscript𝑥1x_{1}italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT or the host mass step, values adopted from Ginolin et al., 2024), and compare the correlation with x1subscript𝑥1x_{1}italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT, t2(r)subscript𝑡2𝑟t_{2}(r)italic_t start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_r ), and r2subscriptsubscript𝑟2\mathcal{F}_{r_{2}}caligraphic_F start_POSTSUBSCRIPT italic_r start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT. We find that although the correlation with t2(r)subscript𝑡2𝑟t_{2}(r)italic_t start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_r ) is statistically significant (>5σしぐまabsent5𝜎>5\sigma> 5 italic_σしぐま), the correlation between the Hubble residual and t2(r)subscript𝑡2𝑟t_{2}(r)italic_t start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_r ) is weaker than with x1subscript𝑥1x_{1}italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT, with Pearson’s correlation coefficients of r=𝑟r=-italic_r = -0.34, --0.50, respectively. r2subscriptsubscript𝑟2\mathcal{F}_{r_{2}}caligraphic_F start_POSTSUBSCRIPT italic_r start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT on the other hand shows a slightly stronger correlation with the colour-corrected Hubble residuals, with r=0.52𝑟0.52r=-0.52italic_r = - 0.52 at the >5σしぐまabsent5𝜎>5\sigma> 5 italic_σしぐま confidence level. The root mean square (rms) dispersion of the Hubble residuals is also reduced when using r2subscriptsubscript𝑟2\mathcal{F}_{r_{2}}caligraphic_F start_POSTSUBSCRIPT italic_r start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT instead of x1subscript𝑥1x_{1}italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT (rms = 0.220 and 0.224 mag, respectively). We perform bootstrapping to estimate the robustness of these results (Efron, 1983; Efron & Tibshirani, 1997). We sample data points with replacement from the original sample 1000 times to generate of alternative sub-samples of the same size. We compute 95 per cent confidence intervals from our bootstrapping and find 0.58r0.460.58𝑟0.46-0.58\leq r\leq-0.46- 0.58 ≤ italic_r ≤ - 0.46 for the correlation between the colour-corrected Hubble residuals and r2subscriptsubscript𝑟2\mathcal{F}_{r_{2}}caligraphic_F start_POSTSUBSCRIPT italic_r start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT. The 95 per cent confidence interval for the correlation with x1subscript𝑥1x_{1}italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT is 0.55r0.420.55𝑟0.42-0.55\leq r\leq-0.42- 0.55 ≤ italic_r ≤ - 0.42, meaning these correlations are consistent. The correlations between the colour-corrected Hubble residuals and x1subscript𝑥1x_{1}italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT as well as r2subscriptsubscript𝑟2\mathcal{F}_{r_{2}}caligraphic_F start_POSTSUBSCRIPT italic_r start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT are shown in Fig. 13.

Hayden et al. (2019) found that x1subscript𝑥1x_{1}italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT measured from the rising part of the light curve showed a stronger correlation with the peak magnitude than x1subscript𝑥1x_{1}italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT calculated from the post-peak light curve. Our results are not aligned with this and suggest that the post-peak light curve contains more information. Although it is interesting that we find r2subscriptsubscript𝑟2\mathcal{F}_{r_{2}}caligraphic_F start_POSTSUBSCRIPT italic_r start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT to be a better standardisation parameter than x1subscript𝑥1x_{1}italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT, the measurement of r2subscriptsubscript𝑟2\mathcal{F}_{r_{2}}caligraphic_F start_POSTSUBSCRIPT italic_r start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT relies on the presence of data at phases >10absent10>10> 10 d, at which time the SN will have dimmed significantly compared to at maximum light. Moreover, the measurement of r2subscriptsubscript𝑟2\mathcal{F}_{r_{2}}caligraphic_F start_POSTSUBSCRIPT italic_r start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT also requires data around maximum light because the peak magnitude in the r𝑟ritalic_r band is required to normalise the light curve to peak. Consequently, standardisation using x1subscript𝑥1x_{1}italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT requires less data and would be preferable in most cases.

Refer to caption
Figure 13: Colour corrected Hubble residuals as a function of x1subscript𝑥1x_{1}italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT (left) and r2subscriptsubscript𝑟2\mathcal{F}_{r_{2}}caligraphic_F start_POSTSUBSCRIPT italic_r start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT (right) for the cosmological sub-sample. Pearson’s r𝑟ritalic_r correlation coefficients and the rms dispersion of the Hubble residuals are shown in the plots.

5.4 Which intrinsic parameters impact the secondary maximum?

The timing of the secondary maximum is thought to be a temperature effect: the lower the temperature, the earlier the shift in the dominant ionisation state of the IGEs. However, the models presented by Kasen (2006) showed that the amount of stable and radioactive IGEs produced in the explosion, the distribution of IGEs throughout the ejecta, and metallicity may also affect the secondary maximum. We will use our observational data to test the impact of these parameters on the secondary maximum in the r𝑟ritalic_r band.

We first investigate the impact of the temperature of the ejecta on the timing of the secondary maximum by checking if there is evolution of t2(r)subscript𝑡2𝑟t_{2}(r)italic_t start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_r ) with the pseudo-equivalent width (pEW) of the Si ii features, using the pEW values presented by Burgaz et al. (2024). Specifically, we use the pEW of the 5972 Å line as a proxy for the temperature of the ejecta. The ratio between the pEWs of the 5972 and 6355 Å lines is more commonly used as a proxy for temperature (Nugent et al., 1995; Hachinger et al., 2008), but Burgaz et al. (2024) find that the pEW of the 5972 Å line is a better tracer of peak magnitude/x1subscript𝑥1x_{1}italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT and therefore temperature, because the 6355 Å line becomes saturated for faint SNe Ia. Our findings agree with this, since we find a statistically significant linear trend between t2(r)subscript𝑡2𝑟t_{2}(r)italic_t start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_r ) and the pEW of the 5972 Å (Pearson’s correlation coefficient r=0.38𝑟0.38r=-0.38italic_r = - 0.38 at the >5σしぐまabsent5𝜎>5\sigma> 5 italic_σしぐま level), but no statistically significant trend between t2(r)subscript𝑡2𝑟t_{2}(r)italic_t start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_r ) and the ratio of the pEWS of the 5972 Å and 6355 Å lines (2.5σしぐま𝜎\sigmaitalic_σしぐま). Assuming the pEW of the 5972 Å Si ii line is a good proxy for temperature, this correlation indicates that t2(r)subscript𝑡2𝑟t_{2}(r)italic_t start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_r ) is impacted by temperature, as expected. However, the correlation is weak (Pearson’s correlation coefficient r=0.38𝑟0.38r=-0.38italic_r = - 0.38). This could be driven by the fact that SNe Ia with the largest t2(r)subscript𝑡2𝑟t_{2}(r)italic_t start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_r ) values are expected to have very weak Si ii 5972 Å pEW values, and in some cases may not be measurable. At the other extreme, SNe Ia with strong Si ii 5972 Å pEW values may not have a measurable secondary maximum. If both of these extremes were measurable, it could strengthen the correlation. Alternatively, there could be other factors affecting the timing of the secondary maximum, such as nickel-mixing, metallicity, or nickel mass, although we note that the nickel mass should also strongly correlate with temperature.

In order to test the impact of the mass of IGEs produced and their distribution throughout the ejecta, we fit the g𝑔gitalic_g- and r𝑟ritalic_r-band light curves in our sample with the TURTLS 1D radiative transfer code, which models the early light curves of thermonuclear SNe (Magee et al., 2018, 2020). We use the same suite of 355 Mchsubscript𝑀𝑐M_{ch}italic_M start_POSTSUBSCRIPT italic_c italic_h end_POSTSUBSCRIPT-models as presented in Deckers et al. (2022), which contains four ejected 56Ni-masses (0.4, 0.5, 0.6, and 0.8 Msubscript𝑀direct-productM_{\odot}italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT), five different 56Ni-distributions, two density profiles, and nine different kinetic energies of the ejecta. We follow the fitting method described in Deckers et al. (2022). In order to reliably constrain the best fitting model, the light curve needs to have good data coverage, especially at early times (Magee et al., 2020). We require the first data point to be no later than 14 d before maximum light in the g band, and a minimum cadence of 3 d. This reduces the number of light curves that we are able to fit with the TURTLS models to 499. We take the 56Ni-mass and 56Ni-distribution from the best matching TURTLS model to each SN Ia, and the 3σしぐま𝜎\sigmaitalic_σしぐま confidence interval is determined by the range of model that fall within 3σしぐま𝜎\sigmaitalic_σしぐま of the normalised reduced-χかい2=1superscript𝜒21\chi^{2}=1italic_χかい start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT = 1 (see Deckers et al., 2022, for further details). Similarly to the findings in Deckers et al. (2022), we find that the majority of the SN Ia light curves are fit with the two most extended nickel distributions and are not able to test for correlations with this parameter. We are able to analyse the effect of nickel mass and find that SNe Ia with a higher estimated nickel mass tend to have a later secondary maximum at the 3.6σしぐま𝜎\sigmaitalic_σしぐま level, as well as a stronger secondary maximum at the 2.5σしぐま𝜎\sigmaitalic_σしぐま level, which agrees with the prediction made by Kasen (2006). We note that the nickel mass parameter in the TURTLS models measures the amount of radioactive nickel produced in the explosion, whereas the study by Kasen (2006) considered the total mass of IGE, combining radioactive and stable material.

Kasen (2006) also showed that metallicity impacts the secondary maximum, but we are unable to measure metallicity for the SNe Ia in our sample. However, Kasen (2006) predicted that as metallicity increases, the secondary maximum is pushed to earlier phases and that earlier secondary maxima should be as bright, if not brighter than later ones. Our observations of normal SNe Ia in the r𝑟ritalic_r band do not support this because r2subscriptsubscript𝑟2\mathcal{F}_{r_{2}}caligraphic_F start_POSTSUBSCRIPT italic_r start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT shows a positive linear trend with t2(r)subscript𝑡2𝑟t_{2}(r)italic_t start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_r ) at a >5σしぐまabsent5𝜎>5\sigma> 5 italic_σしぐま confidence level, meaning that in our sample, later secondary maxima tend to be brighter. However, we do not find a significant positive trend between i2subscriptsubscript𝑖2\mathcal{F}_{i_{2}}caligraphic_F start_POSTSUBSCRIPT italic_i start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT and t2(i)subscript𝑡2𝑖t_{2}(i)italic_t start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_i ). As discussed in Sect. 5.1, one interpretation of these results could be that a change in metallicity has a stronger impact on the secondary maximum in the i𝑖iitalic_i band than in the r𝑟ritalic_r band.

The models presented in Kasen (2006) predict a third bump in SN Ia light curves at around similar-to\sim 80 d post maximum light in the B𝐵Bitalic_B band, due to the recombination of singly ionised iron lines to neutral iron. Unfortunately, the ZTF DR2 light curves do not have sufficiently high S/N at these phases to confirm or reject this prediction. Moreover, Kasen (2006) states that the LTE approximations made to produce the models are not applicable at these late epochs, and this may result in a further delay of the third maximum as well as a weakening of the bump. Kasen (2006) also predicted that the third bump should be most prominent in the J𝐽Jitalic_J band, but the model predictions presented by (Blondin et al., 2015, appendix C) show that the only strong Fe i features at 100 d post explosion exist in the U𝑈Uitalic_U band, and so it is unlikely that this recombination would leave any visible signature in the r𝑟ritalic_r/i𝑖iitalic_i-band photometry. However, we would encourage a future study to focus on getting high S/N data, in a range of wavelengths, at epochs between 80–150 d to search for signatures of the third maximum.

Refer to caption
Figure 14: The time of t2(r)subscript𝑡2𝑟t_{2}(r)italic_t start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_r ) in the r𝑟ritalic_r band as a function of the pEW of the 5972 Å Si ii line.

6 Conclusions

We measured the time and strength of the secondary maximum in the r band for a sample of 893 SNe Ia using GP fits. Our main results are as follows:

  1. 1.

    We find >5σしぐまabsent5𝜎>5\sigma> 5 italic_σしぐま correlations between Δでるたm15(g)Δでるたsubscript𝑚15𝑔\Delta m_{15}(g)roman_Δでるた italic_m start_POSTSUBSCRIPT 15 end_POSTSUBSCRIPT ( italic_g ) and the timing and strength of the secondary maximum, confirming that fainter and faster evolving SNe Ia have earlier and weaker secondary maxima, and brighter, slower evolving SNe Ia have later and stronger secondary maxima.

  2. 2.

    We find that sub-classes of SNe Ia which are speculated to experience interaction with a CSM/envelope (03fg- and 02es-like, as well as Ia-CSM) have higher values of r2subscriptsubscript𝑟2\mathcal{F}_{r_{2}}caligraphic_F start_POSTSUBSCRIPT italic_r start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT than expected from the general trend of the sample, which can be explained by additional flux contribution from said interaction. Interestingly, we also find the 91T-like SNe Ia in general show higher r2subscriptsubscript𝑟2\mathcal{F}_{r_{2}}caligraphic_F start_POSTSUBSCRIPT italic_r start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT-values than the trend, which could point to these SNe Ia experiencing some interaction. Several 02cx-like SNe Ia show higher r2subscriptsubscript𝑟2\mathcal{F}_{r_{2}}caligraphic_F start_POSTSUBSCRIPT italic_r start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT-values which could be attributed to a bound remnant left behind after an incomplete explosion.

  3. 3.

    We constrain the transparency timescales for our sample, and find that the distribution peaks at 35.9 ±plus-or-minus\pm± 0.1 d, but there is a tail to lower transparency timescales. We fit the distribution with a bimodal Gaussian and find that the higher transparency timescale mode dominates, but there is a second component which peaks at 30.9 ±plus-or-minus\pm± 0.4 d.

  4. 4.

    The lower transparency timescale component of the bimodal Gaussian, peaking at 30.9 ±plus-or-minus\pm± 0.4 d, is produced almost exclusively by SNe Ia residing in locally red environments. The main peak at 35.9 ±plus-or-minus\pm± 0.1 d is produced by SNe Ia in both locally red and blue environments.

  5. 5.

    The transparency timescales of most of the SNe Ia in our cosmological sample (97 per cent) are consistent with predictions from double detonation explosion models, whereas only 28 per cent are consistent with the predictions from delayed detonation models.

  6. 6.

    We measure the total ejected mass using two methods: from the transparency timescale, which was estimated from r2subscriptsubscript𝑟2\mathcal{F}_{r_{2}}caligraphic_F start_POSTSUBSCRIPT italic_r start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT, or from the correlation Mejsubscript𝑀ejM_{\textnormal{ej}}italic_M start_POSTSUBSCRIPT ej end_POSTSUBSCRIPT shows with x1subscript𝑥1x_{1}italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT. Both methods are derived from bolometric light curves, and the distributions of Mejsubscript𝑀ejM_{\textnormal{ej}}italic_M start_POSTSUBSCRIPT ej end_POSTSUBSCRIPT we find from two methods are consistent, with medians of 1.3 ±plus-or-minus\pm± 0.3 and 1.2 ±plus-or-minus\pm± 0.2 Msubscript𝑀direct-productM_{\odot}italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT, respectively.

  7. 7.

    We identified a spectral feature that sits in the r𝑟ritalic_r band around 6500 Å which begins to strengthen during the onset of the secondary maximum, and using a NLTE radiative transfer model we tentatively identify this feature as Fe ii.

  8. 8.

    From the spectral time series of SN 1999by (a 91bg-like SN Ia) we find that these same features appear in the spectrum as early as 3 d post maximum, supporting the suggestion that the secondary maximum is not visible in 91bg-like SNe Ia because it coincides with the primary peak.