This is the html version of the file https://arxiv.org/abs/2406.05050.
Google automatically generates html versions of documents as we crawl the web.
Page 1
DES-2024-0831
FERMILAB-PUB-24-0293-PPD
MNRAS 000, 1โ€“13 (2024)
Preprint 10 June 2024
Compiled using MNRAS LATEX style file v3.0
The Dark Energy Survey Supernova Program: Slow supernovae show
cosmological time dilation out to ๐‘ง โˆผ 1.
Ryan M. T. White1โ˜… , Tamara M. Davis1 , Geraint F. Lewis2 , Christopher Lidman3,4 , Paul Shah5 ,
T. M. C. Abbott6, M. Aguena7, S. Allam8 , F. Andrade-Oliveira9, J. Asorey10, D. Bacon11, S. Bocquet12 ,
D. Brooks5 , D. Brout13, E. Buckley-Geer14,8 , D. L. Burke15,16, A. Carnero Rosell17,7 , D. Carollo18,
J. Carretero19 , L. N. da Costa7, M. E. S. Pereira20, J. De Vicente21 , S. Desai22 , H. T. Diehl8 ,
S. Everett23, I. Ferrero24, B. Flaugher8 , J. Frieman8,25 , J. Garc๏ฟฝa-Bellido26 , E. Gaztanaga27,11,28 ,
G. Giannini19,25 , K. Glazebrook29, R. A. Gruendl30,31, S. R. Hinton1, D. L. Hollowood32,
K. Honscheid33,34 , D. J. James13 , R. Kessler14,25 , K. Kuehn35,36 , O. Lahav5 , J. Lee37 , S. Lee23,
M. Lima38,7, J. L. Marshall39 , J. Mena-Fern๏ฟฝndez40 , R. Miquel41,19 , J. Myles42, A. M๏ฟฝller29,
R. C. Nichol11, R. L. C. Ogando43 , A. Palmese44 , A. Pieres7,43 , A. A. Plazas Malag๏ฟฝn15,16 ,
A. K. Romer45 , M. Sako37, E. Sanchez21 , D. Sanchez Cid21 , M. Schubnell9 , M. Smith46 ,
E. Suchyta47 , M. Sullivan46 , B. O. S๏ฟฝnchez48,49 , G. Tarle9 , B. E. Tucker4, A. R. Walker6 ,
N. Weaverdyck50,51, and P. Wiseman46,
(DES Collaboration)
Affiliations are listed at the end of the paper.
Accepted XXX. Received YYY; in original form ZZZ
ABSTRACT
We present a precise measurement of cosmological time dilation using the light curves of 1504 type Ia supernovae from the
Dark Energy Survey spanning a redshift range 0.1 โ‰ฒ ๐‘ง โ‰ฒ 1.2. We find that the width of supernova light curves is proportional
to (1 + ๐‘ง), as expected for time dilation due to the expansion of the Universe. Assuming type Ia supernovae light curves are
emitted with a consistent duration ฮ”๐‘กem, and parameterising the observed duration as ฮ”๐‘กobs = ฮ”๐‘กem(1 + ๐‘ง)๐‘, we fit for the form of
time dilation using two methods. Firstly, we find that a power of ๐‘ โ‰ˆ 1 minimises the flux scatter in stacked subsamples of light
curves across different redshifts. Secondly, we fit each target supernova to a stacked light curve (stacking all supernovae with
observed bandpasses matching that of the target light curve) and find ๐‘ = 1.003 ๏ฟฝ 0.005 (stat) ๏ฟฝ 0.010 (sys). Thanks to the large
number of supernovae and large redshift-range of the sample, this analysis gives the most precise measurement of cosmological
time dilation to date, ruling out any non-time-dilating cosmological models at very high significance.
Key words: supernovae: general โ€“ cosmology: observations
1 INTRODUCTION
Time dilation is a fundamental implication of Einsteinโ€™s theory of
relativity in an expanding Universe โ€” the observed duration of an
event, ฮ”๐‘กobs, should be longer than the intrinsic emitted (or rest-
frame) duration, ฮ”๐‘กem, by a factor of one plus the observed redshift,
๐‘ง,
ฮ”๐‘กobs = ฮ”๐‘กem(1 + ๐‘ง).
(1)
The idea of using time dilation to test the hypothesis that the Universe
is expanding dates back as far as Wilson (1939) and was revisited
โ˜… E-mail:ryan.white@uq.edu.au
by Rust (1974). One of the first observational hints of time dilation
was the observation by Piran (1992) and Norris et al. (1994) that the
duration of gamma-ray bursts (GRBs) was inversely proportional to
their brightness โ€“ they used this to argue that at least some GRBs
must be cosmological. The first measurements of cosmological time-
dilation using supernovae were made by Leibundgut et al. (1996) for
a single Type Ia supernova (SN Ia) at ๐‘ง = 0.479 and Goldhaber et al.
(1997) for seven supernovae at 0.3 <๐‘ง< 0.5. Most relevant to this
work, we take Goldhaber et al. (2001) as the current state of the art
in identifying cosmological time dilation in SN Ia photometry. They
used 35 supernovae in the redshift range 0.30 โ‰ค ๐‘ง โ‰ค 0.70, to test a
model with a factor (1+๐‘ง)๐‘ time dilation and found ๐‘ โˆผ 1.07๏ฟฝ0.06.
To avoid degeneracy between the natural variation of light-curve
๏ฟฝ 2024 The Authors
arXiv:2406.05050v1 [astro-ph.CO] 7 Jun 2024

Page 2
2
White et al.
width and time dilation, Foley et al. (2005) and Blondin et al. (2008)
observed time dilation in the evolution of spectral features of high-๐‘ง
Type Ia supernovae (SNe Ia). The former found inconsistency with
no time dilation at the 99% confidence level and the latter finding
๐‘ = 0.97 ๏ฟฝ 0.10. Most recently, Lewis & Brewer (2023) inferred
๐‘ = 1.28
+0.28
โˆ’0.29
using the variability of 190 quasars out to ๐‘ง โˆผ 4.
Despite these successes, there remains continued discussion of hybrid
or static-universe models such as Tired Light (Zwicky 1929; Gupta
2023) that do not predict expansion-induced time dilation.
In this study, we measure cosmological time dilation using SNe Ia
from the full 5-year sample released by the Dark Energy Survey
(DES) (DES Collaboration et al. 2024), which contains โˆผ 1500 SN Ia
spanning the redshift range 0.1 โ‰ฒ ๐‘ง โ‰ฒ 1.2 โ€” significantly larger and
higher-redshift than any sample of supernovae previously used for a
time-dilation measurement. Such a large sample of SNe is important
in reducing statistical uncertainty and such a high-redshift sample
is the ideal regime to robustly identify time dilation. Over half the
DES-SN5YR sample are at ๐‘ง > 0.5, compared to only 12% in the
previous gold-standard Pantheon+ sample (Brout 2019) and 37% in
the Goldhaber analysis (Goldhaber et al. 2001). Therefore the DES
sample should have observed light-curve durations more than 1.5
times longer than their rest frame durations (up to 2.2 times longer
for those at ๐‘ง โˆผ 1.2). This means their time dilation signal should be
significantly larger than the intrinsic width variation expected due to
SNe Ia diversity in their subtypes.
We test the model that time dilation occurs according to,
ฮ”๐‘กobs = ฮ”๐‘กem(1 + ๐‘ง)๐‘
.
(2)
If standard time dilation is true we should find ๐‘ = 1. If no time
dilation occurs we should find ๐‘ = 0.
For this work we aim to keep supernova modelling assumptions to
a minimum to avoid circularity in our arguments (because most mod-
els of supernova light curves are generated assuming time-dilation
occurs). We therefore take two data-driven approaches to measuring
time dilation:
(i) Firstly, we simply take all the light curves, divide their time
axis by (1 + ๐‘ง)๐‘ relative to the time at peak brightness, and find the
value of ๐‘ that minimises the flux scatter.
(ii) Secondly, we โ€˜de-redshiftโ€™ all the supernova light curves and
stack them to define a data-driven SN Ia โ€˜reference light curveโ€™. Then
for each individual SN Ia we measure the observed light curve width
(๐‘ค) relative to the appropriate reference, ฮ”๐‘กobs = ๐‘คฮ”๐‘กreference. This
allows us to see if the time-dilation occurs smoothly with redshift
and find the best-fit value of ๐‘, where the expected result would be
๐‘ค = (1 + ๐‘ง) corresponding to ๐‘ = 1.
The first method is entirely data-driven and has no time-dilation
assumptions. In the second method for ease of computation we create
the stacked reference by dividing the time axis of the light curves by
(1+๐‘ง). This method therefore includes an assumption of time-dilation
in the generation of the reference. Even though this assumption is
justified by the result of the first method, the second method should
strictly be considered a consistency check. We note it is possible
to remove any circularity by keeping the reference light curves in
their rest frame and fitting to ๐‘ค = ((1 + ๐‘งtarget)/(1 + ๐‘งreference))๐‘,
which requires multiple different reference light curves per target
supernova; mathematically this is very similar to our approach but
requires even more data. To further check that this method rules out
no time dilation we re-test method two without de-redshifting the
reference light curves; it dramatically fails the consistency check,
see Appendix C.
This paper is arranged as follows. In Section 2 we discuss the use
of type Ia supernovae as standard clocks, and the challenges that
need to be taken into account when comparing SNe Ia light curves
observed in different bands across different redshifts. In Section 3, we
present the data used in this study, while Section 4 we describe our
approach of defining a reference light curve and the determination of
the redshift dependence of the time dilation signal. We discuss our
results in Section 5 and conclude in Section 6 that the null hypothesis
of no time dilation is inconsistent with the data.
2 TYPE IA SUPERNOVAE AS STANDARD CLOCKS
Any investigation into the physics at large length scales in the uni-
verse relies on known quantities, be they standard candles, rulers,
sirens, or clocks. SNe Ia have long fit the bill of a standardisable can-
dle on the basis of their extreme brightness and consistency (Tripp
1998; M๏ฟฝller-Bravo et al. 2022; Scolnic et al. 2023), allowing their
observation over cosmic distances with only little uncertainty in their
intrinsic properties. As SNe Ia are the explosions of a white dwarf
approaching the Chandrasekhar limit (Hoyle & Fowler 1960; Ruiter
2019), their properties are reasonably uniform across their popula-
tion compared to other SN types; not only are they standardisable in
brightness, but also in time (Phillips 1993; Leibundgut et al. 1996).
Hence, the observed duration of SN Ia explosions are well suited
to investigating time dilation as a result of an expanding universe
(Wilson 1939; Rust 1974).
The presence of a time dilation signal in SNe Ia data tests the gen-
eral relativistic prediction of an expanding universe having a factor
of (1 + ๐‘ง) time dilation (Wilson 1939; Blondin et al. 2008). This
signal needs to be corrected for in supernova cosmology analyses
(Leibundgut & Sullivan 2018; Carr et al. 2022) and so conclusively
quantifying the effect of time dilation is foundational to our cos-
mological model, especially considering the continued discussion of
hybrid or static-universe models such as Tired Light (Zwicky 1929;
Gupta 2023) that do not predict expansion-induced time dilation.
2.1 The importance of colour
SNe Ia are known to spectrally evolve over the duration of their
โˆผ 70 day bright period. The early-time spectrum is relatively blue
with spectral features dominated by transitions from intermediate
mass elements. The spectrum then reddens on the order of days from
heavier element emission lines and the cooling of the continuum
(Filippenko 1997). Previous papers have described the redward evo-
lution of SNe Ia spectra over time (Takanashi et al. 2008; Blondin
et al. 2012; Branch & Wheeler 2017), while photometric evidence
of this phenomenon is seen in the light curve peaking later in redder
bandpasses than in bluer ones (as in Figure 1) for a light curve in
the rest-frame optical. As such, the photometric behaviour of a light
curve is dependent on the rest-frame wavelength range observed.
This means that, to a good approximation, a typical high-๐‘ง SN Ia
observed in a redder band should have the same photometric and
spectral characteristics as a medium-๐‘ง SN Ia observed in a bluer band
(Figure 2). Since our photometric bands are fixed, they sample dif-
ferent rest-frame wavelength ranges as the supernovae are redshifted.
Therefore, it is critical to design a method that ensures time dilation
measurements compare light curves measured at similar rest-frame
wavelengths.1
1 A note on language: The phrase โ€˜rest-frameโ€™ wavelengths arises from the
usual assumption that redshifts are due to recession velocities. The fact red-
MNRAS 000, 1โ€“13 (2024)

Page 3
Time Dilation with DES SNe Ia
3
-20
0
20
40
60
Time Since Peak Brightness, t - tpeak (days)
0.0
0.2
0.4
0.6
0.8
1.0
Normalised
Flux
r-band data
i-band data
z-band data
Figure 1. The normalised (in flux) light curve of a SN Ia at ๐‘ง = 0.4754
shows intrinsically broader light curves at redder wavelengths that tend to
peak later (at least in the optical regime observed across our dataset). The
๐‘ฅ-axis represents time in the observer frame, and SALT3 model fits (solid
lines) are overlaid onto the data in each band.
400
500
600
700
800
Wavelength (nm)
0.0
0.2
0.4
0.6
0.8
1.0
Relativ
e
T
ransmittance/In
tensity
DES g
DES r
DES i
z = z1
z = z2
Figure 2. For a supernova at redshift ๐‘ง1 observed in a given filter, there exists
a higher redshift ๐‘ง2 > ๐‘ง1 such that another supernova at ๐‘ง2 observed in a
redder filter will have a similar rest frame effective wavelength as the nearer
supernova. We show here the spectrum of SN2001V (Matheson et al. 2008)
at ๐‘ง1 โˆผ 0.01 and again artificially redshifted to ๐‘ง2 โˆผ 0.35. Here we can
clearly see the Si II absorption line (615nm) redshifted from the ๐‘Ÿ band to
roughly the same position along the bandwidth in the ๐‘– band. Overlaid are the
transmission curves of the DES filters from Abbott et al. (2018) Fig. 1.
2.2 Stretch-Luminosity relation
Our fitting methods are independent of individual supernova light
curve models and are based only on the assumption that supernova
shifts occur is not in question here (so it is fine to use (1 + ๐‘ง) to calculate
matching rest-frame wavelengths, and this contains no time-dilation assump-
tion). The question is whether that redshift arises due to a recession velocity,
which would also cause time-dilation.
light curves are all similar. One dissimilarity between SNe Ia is the
โ€˜stretchโ€™ in their light curve โ€“ an intrinsic width variation of up to
โˆผ 20% that is separate from time dilation and strongly correlated
with the peak brightness (Phillips et al. 1999). With this amount of
data at such high redshifts we can simply treat this intrinsic variation
as noise. The stretch variation between SNe Ia essentially acts as
random (but intrinsic) scatter in the obtained widths of light curves.
This does not affect the overall trend of light curve width against
redshift. We therefore make no correction for the stretch-luminosity
relation in this work, to maintain maximal model-independence.
As long as there is a representative sample of the entire popula-
tion of SNe Ia at every redshift, this simple analysis should measure
time-dilation without bias. However, Malmquist bias can influence
the result since brighter supernovae have wider light curves. If faint
supernovae are under-represented at high-redshifts one might ex-
pect a slight bias toward a higher inferred time dilation at high-
๐‘ง. Thankfully, the DES data are well-sampled to such high-๐‘ง that
Malmquist bias has minimal impact on our results. M๏ฟฝller et al.
(2022) showed that the full stretch distribution is well represented in
the DES-SN5YR sample out to ๐‘ง โˆผ 1.1 (see also Fig. A2), meaning
that the stretch-luminosity relation should have a negligible effect on
our results.
Previous studies (e.g. Nicolas et al. 2021) have also found that
the stretch distribution of the SN Ia population drifts slightly with
redshift (getting wider by โˆผ 3% between 0.0 <๐‘ง< 1.4). Even though
we do not see this drift in the DES-SN5YR sample (see Fig. A2),
we quantify the possible impact of this effect on our time dilation
measurement in Appendix A and find it to be small.
3 DATA
We exclusively use the data of the 1635 type Ia supernovae measured
by the Dark Energy Survey Supernova Program (DES Collaboration
et al. 2024). The DECam instrument on the 4m Blanco telescope at
the Cerro Tololo Inter-American Observatory (Flaugher et al. 2015)
observed most of the photometrically classified SN Ia candidates in
the ๐‘”, ๐‘Ÿ, ๐‘–, and ๐‘ง bands according to the criteria in Smith et al. (2020).
The flux is determined by difference imaging (Kessler et al. 2015).
High-redshift SNe Ia typically show negligible flux in the ultraviolet
wavelength region, so ๐‘” and ๐‘Ÿ band light curves are only useful for
SNe at ๐‘ง โ‰ฒ 0.4 and ๐‘ง โ‰ฒ 0.85 respectively (see DES Collaboration
et al. 2024, Fig. 2). The SALT3 (Kenworthy et al. 2021) template fits
with a cadence of 2-day time sampling in each band were available for
each SN candidate. We use these fits to estimate the peak flux of each
supernova so we can normalize the observed flux values; the time
relative to the ๐ต-band maximum in the SALT3 fit for each light curve
is also used to define the time since peak brightness measurement in
our analysis. We otherwise discard the SALT supernova information.
We performed an initial quality cut on the sample of 1635 SNe Ia,
requiring the probability of being a type Ia PROBIa > 0.5 as classified
with SuperNNova (M๏ฟฝller & de Boissi๏ฟฝre 2020; M๏ฟฝller et al. 2022;
Vincenzi et al. 2024). This kept the sample of usable light curves
high while removing possible type II supernova contaminants. We
removed individual data points from each light curve that had an
error in their flux (FLUXCALERR2 for the flux value FLUXCAL) greater
than 20; this was done to restrict our fitting to the highest quality ob-
servations, particularly cutting those with very low signal-to-noise at
2 FLUXCALERR is the Poisson error on FLUXCAL, which is the variable used
for flux in SNANA corresponding to mag = 27.5 โˆ’ 2.5 log10 (FLUXCAL).
MNRAS 000, 1โ€“13 (2024)

Page 4
4
White et al.
2โˆ’1
2โˆ’2
2โˆ’3
2โˆ’4
2โˆ’5
Width Factor, ฮด
0.2
0.4
0.6
0.8
1.0
T
arget
SNe
Redshift
z
102
103
104
Reference
P
opulation
Figure 3. For each of the SNe Ia in our sample, we constructed a reference
light curve with a ๐›ฟ = 2โˆ’๐‘ฅ parameter according to Equation 3, with ฮ”๐œ†๐‘“
band FWHM. We counted how many data points populated the reference
curves (i.e. the number of points in Fig. 4 for example) changing ๐›ฟ in integer
steps of powers of 2. This plot shows the reference populations for a target
SN measured in the observer frame ๐‘– band, and is largely similar for the
different bands differing mainly by a linear shift on the vertical axis. That is,
an analogous plot in a bluer band would see the colour distribution shifted
downwards in target redshift and a redder band upwards.
high redshift (whose observations had comparatively low FLUXCAL).
In the analysis, we did not attempt to fit SNe light curve widths if
their light curve had fewer than 5 data points and if their reference
curve had fewer than 100 data points (discussed in Section 4). This
was done on a per-band basis; we estimated the width of each SN
light curve in each band where it satisfied these criteria. Individual
light curves were also omitted from the analysis if the ๐œ’2 width fitting
did not converge. All together, after these quality cuts we were left
with width measurements of 1504 unique SN Ia across the dataset.
4 FITTING SUPERNOVA PHOTOMETRY TO A
REFERENCE LIGHT CURVE
The photometric analysis in Goldhaber et al. (2001) relied funda-
mentally on template fitting to measure the expansion-induced time
dilation signal in SNe Ia. With the wealth of data now available we
can in principle obtain the same dilation signal using the data alone,
independent of a light-curve template. Herein we describe such a
template-independent method involving scaling the time axis of each
of the 1504 individual light curves to fit their own unique reference
curve composed of all the matching photometry in the DES dataset.
The time dilation signal in each light curve is the multiplicative in-
verse of said scaling factor, and this can be found for each of the 1504
SN Ia light curves in the data. This differs from a traditional template-
fitting method in that we do not assume the shape of a light curve
but instead let the data from other SNe compose the template-like
reference curve.
4.1 Reference curve construction
The main functionality of this method is to use the photometric data
of many SNe to automatically create a reference light curve unique
to any one target SN. The target SN is the SN whose width we are
trying to measure.
Since the shape of a SN Ia light curve is dependent on the rest-
frame effective wavelength at which it is observed (Fig. 1; see also
Takanashi et al. 2008; Blondin et al. 2012), the reference photometry
must be composed only of light curves that have the same (or very
similar) intrinsic shape as the target SN. Hence, we must choose ref-
erence photometry that samples the same rest-frame effective wave-
length as the target light curve. This effect is shown in Fig. 2, where,
for example, we might compare a low-๐‘ง supernova in some band
to a higher-๐‘ง supernova observed in a redder band with the same
(or similar) rest frame effective wavelength. We can compare the
photometry between the two events provided that their rest-frame
effective wavelength (and hence their light curve shape/evolution) is
alike.
If we chose instead to fit each target light curve in some band
against all of the photometry from that band, we would expect a non-
linear change in slope as a function of redshift on a width-vs-redshift
plot. The explanation for this lies in the fact that SNe Ia spectra get
redder over time; the light curves measured in a redder band are
intrinsically wider than those measured in a bluer band as shown
in Fig. 1. Hence, with this hypothetical method (comparing to all
photometry), we would observe a bluer than average rest-frame curve
for a high redshift SNe which would bias the obtained width to an
intrinsically thinner value. Conversely, we would be biased towards
a wider (redder) than average width for low redshift supernovae. To
avoid this bias, we use the aforementioned method of only using
reference photometry with a similar rest-frame wavelength as our
target light curve.
To find relevant light curves to populate the reference curve, we
pick all light curves out of a calculated redshift range. To fit a single
(target) SN light curve at redshift ๐‘ง imaged in a band of central
wavelength ๐œ† ๐‘“ , we can populate the reference curve with SNe within
the redshift range
๐œ†๐‘Ÿ (1 + ๐‘ง)
๐œ† ๐‘“
โˆ’ ๐›ฟ
ฮ”๐œ† ๐‘“
๐œ† ๐‘“
โ‰ค 1 + ๐‘ง๐‘Ÿ โ‰ค
๐œ†๐‘Ÿ (1 + ๐‘ง)
๐œ† ๐‘“
+ ๐›ฟ
ฮ”๐œ† ๐‘“
๐œ† ๐‘“
(3)
whose photometry is measured in a band of central wavelength ๐œ†๐‘Ÿ .
Here ๐›ฟ is a free parameter which, together with the band full width at
half maximum (FWHM) ฮ”๐œ† ๐‘“ , describe the acceptable wiggle room
in the relative band overlap. A derivation of this formula is given
in Appendix B and a graphical representation of this construction is
shown on the left-side plots in Fig. 6.
We show in Fig. 3 the number of points in the reference curves
(hereafter referred to as reference population) for all of the DES SNe
with a variable ๐›ฟ parameter. Ideally, this ๐›ฟ parameter should be as
small as practical to ensure that the reference curve is consistent in
shape (i.e. the spread of rest frame effective wavelengths is small). In
practice, we find a value of ๐›ฟ = 2โˆ’4 is the minimal value that provides
a large enough reference population for high/low redshift target SNe
(on the order of โˆผ 102 needed to satisfy the Section 3 criteria at
๐‘ง โˆผ 1.1). For medium range target SNe redshifts (in the context of
the DES-SN sample), we note that the reference population is large
(โˆผ 103) even for ๐›ฟ โ‰ฒ 2โˆ’4.
After populating the reference curve with data points we then
normalise the photometry in flux; as the curve is populated with
the data of several SNe at different redshifts, the curve must be
homogeneous in flux. To do this, we utilised the peak flux in the
SALT3 model light curves provided for each SNe. The data in each
constituent curve is normalised by this value before being added to
the reference. For convenience we also use the time of peak brightness
given by SALT3 as the reference point about which to stretch the light
MNRAS 000, 1โ€“13 (2024)

Page 5
Time Dilation with DES SNe Ia
5
โˆ’0.2
0.0
0.2
0.4
0.6
0.8
1.0
1.2
1.4
Normalised
Flux
(Pre-Correction)
โˆ’20
0
20
40
60
80
Time Since Peak Brightness, t โˆ’ tpeak (days)
โˆ’0.2
0.0
0.2
0.4
0.6
0.8
1.0
1.2
1.4
Normalised
Flux
(Corrected)
z โˆผ 0.21 (r band)
z โˆผ 0.48 (i band)
z โˆผ 0.74 (z band)
Figure 4. For a target SN Ia at ๐‘ง โ‰ƒ 0.48, the ๐‘–-band reference curve consists of
data from the ๐‘Ÿ, ๐‘–, and ๐‘ง bands. These data are chosen at redshifts according
to equation (3) (and visually shown in the left plots of Fig. 6). Top: the data
in different bands are not in phase (in the observer frame). It is visually
obvious that light curves appear wider in time at higher redshift. Bottom:
after a (1 + ๐‘ง) correction to the SN Ia light curves, we see a consistent trend
across all bandpasses and time. This alone is evidence for some degree of
time-dilation.
curves (see equation 5). These are the only uses of SALT3 information
and we expect that the same time dilation signal would be obtained
in the data with any other consistent normalisation measures.
4.2 First measure of time dilation: minimising scatter in the
reference curve
After the flux of the reference curve is normalised, we see that the
different bandpass data in the curve are temporally stretched (see the
colour gradient of the top plot in Fig. 4). As the redder bandpasses
are sampled at higher redshift, this is an immediate indication of time
dilation. Without assuming our expected cosmological time dilation
of (1 + ๐‘ง), we can scale the data in all of the reference curves by a
factor of (1 + ๐‘ง๐‘—)๐‘, where ๐‘ง๐‘— is the redshift of each constituent curve
in a reference and ๐‘ is a free parameter. We posit that minimising
the flux dispersion in the reference curve is analogous to finding the
optimal temporal scaling, simultaneously minimising the dispersion
in time. Hence, finding the value of ๐‘ that minimises the flux scatter
gives us our correction factor.
To investigate this, we generated reference curves for each of the
-1.0
-0.5
0.0
0.5
1.0
1.5
2.0
b
0.130
0.135
0.140
0.145
0.150
Reference
Flux
Disp
ersion
Figure 5. By scaling the reference photometry in time according to (1+๐‘ง)๐‘ for
some free parameter ๐‘, we find ๐‘ โˆผ 1 minimises the reference flux dispersion
across the entire SNe sample. The reference flux dispersion represents the
median dispersion of flux across the entire sample of normalised reference
light curves in each band (here averaged for the๐‘Ÿ๐‘–๐‘ง bands), where the errorbars
indicate one standard deviation in these values. We note that this figure yields
a signal of (1 + ๐‘ง) time dilation in the DES dataset, independent of the rest
of the analysis.
target SNe per observing band and scaled the data in time according
to the aforementioned relation in terms of ๐‘ (Fig. 4 shows this scaling
for ๐‘ = 1 as an example). Then, we binned the timeseries data into
30 equal-width time bins and found the standard deviation of the
flux within each bin. We calculated the median of these 30 standard
deviations as a representative estimate of the total flux scatter for that
reference curve with that tested ๐‘ value. We then took the median of
those results across all the SN reference curves as our estimate of the
dispersion for that ๐‘, which is shown in Fig. 5. That is, our reference
flux dispersion is
๐œŽrf(๐‘) = Med ({{Med(๐œŽ๐‘– ๐‘— (๐‘))|โˆ€๐‘— โˆˆ (1, ..., 30)}|โˆ€๐‘– โˆˆ (1, ..., ๐‘sn)})
(4)
for ๐‘sn supernova light curves in that band, and ๐œŽ๐‘– ๐‘— (๐‘) being the
flux standard deviation of the ๐‘–th light curve in the ๐‘—th timeseries
bin. This process was repeated for each of the ๐‘Ÿ๐‘–๐‘ง observing bands,
omitting the ๐‘” band due to the smaller number of SNe. We crudely
estimate the error (for each ๐‘) in this method as being the standard
deviation of the median dispersions across all light curves. We find
an optimal scaling corresponding to ๐‘ โˆผ 1 (Fig. 5) across the entire
dataset, which is the expected dilation factor of โˆผ (1 + ๐‘ง).
If there was no time dilation we would expect the minimum dis-
persion in the reference curve to be at ๐‘ โˆผ 0 (i.e. no time scaling) in
Fig. 5. The fact that we find ๐‘ โ‰ˆ 1 is evidence for time dilation of
the expected form. This a rough but completely model independent
measure of time dilation and it is the paperโ€™s first main result.
4.3 Second measure of time dilation: Finding each light curve
width
After constructing the reference curves for a target SN, we are ready
to fit for the width, ๐‘ค, of each individual target light curve and look
for a trend with redshift. This method enables a more precise measure
of ๐‘.
MNRAS 000, 1โ€“13 (2024)

Page 6
6
White et al.
0.00
0.25
0.50
0.75
1.00
Target SNe z
0.0
0.2
0.4
0.6
0.8
1.0
1.2
1.4
Reference
z
g Band
r Band
i Band
z Band
-10
0
10
20
30
40
Time Since Peak Brightness (days)
-0.2
0.0
0.2
0.4
0.6
0.8
1.0
1.2
1.4
Normalised
Flux
Reference Photometry
Target Photometry (Obs. Frame)
Target Photometry (Scaled)
0.00
0.25
0.50
0.75
1.00
Target SNe z
0.0
0.2
0.4
0.6
0.8
1.0
1.2
1.4
Reference
z
g Band
r Band
i Band
z Band
-20
-10
0
10
20
30
40
Time Since Peak Brightness (days)
-0.2
0.0
0.2
0.4
0.6
0.8
1.0
1.2
1.4
Normalised
Flux
Reference Photometry
Target Photometry (Obs. Frame)
Target Photometry (Scaled)
0.00
0.25
0.50
0.75
1.00
Target SNe z
0.0
0.2
0.4
0.6
0.8
1.0
1.2
1.4
Reference
z
g Band
r Band
i Band
z Band
-30
-20
-10
0
10
20
30
40
Time Since Peak Brightness (days)
-0.2
0.0
0.2
0.4
0.6
0.8
1.0
1.2
1.4
Normalised
Flux
Reference Photometry
Target Photometry (Obs. Frame)
Target Photometry (Scaled)
Figure 6. We show the reference curve construction and subsequent target SN fit for 3 SNe at redshifts ๐‘ง โ‰ƒ 0.22, ๐‘ง โ‰ƒ 0.43, and ๐‘ง โ‰ƒ 1.02 and in fitting bands ๐‘Ÿ,
๐‘–, and ๐‘ง respectively (in descending order). The left plots show the allowed ranges for reference curve SN sampling given the target redshift (and ๐›ฟ = 2
โˆ’4). The
vertical line of dots is plotted at the target SN redshift, with each dot representing the redshift of a DES supernova (vertical axis). The dots that fall in the narrow
coloured bands are the SNe that make up the reference population, as those data all share approximately the same rest-frame wavelength in their respective
bands. The right plots show the constructed (1 + ๐‘ง) time-scaled reference curve (small coloured points) with respect to the target SN photometry (blue points)
and subsequent target photometry scaled on the time axis to fit the reference (best-fit widths of 1.42, 1.49, and 2.17 respectively). Due to the statistics associated
with such large reference curve populations, the contribution of any individual reference point uncertainty to the overall reference curve uncertainty is negligible
and not plotted; the uncertainty in the target data has a much higher contribution to the uncertainty in the fitting.
MNRAS 000, 1โ€“13 (2024)

Page 7
Time Dilation with DES SNe Ia
7
0.5
1.0
1.5
2.0
2.5
3.0
3.5
Ligh
tcurv
e
width
w
g
1 + z
(1 + z)1.021๏ฟฝ0.027
Data
Binned Data
ะณ
(1 + z)1.004๏ฟฝ0.008
1.0
1.2
1.4
1.6
1.8
2.0
1 + z
0.5
1.0
1.5
2.0
2.5
3.0
3.5
Ligh
tcurv
e
width
w
i
(1 + z)0.988๏ฟฝ0.008
1.0
1.2
1.4
1.6
1.8
2.0
1 + z
ฮถ
(1 + z)0.997๏ฟฝ0.008
Figure 7. Using the reference-scaling method described in Section 4.3, we plot the fitted SNe widths of light curves observed in the ๐‘”, ๐‘Ÿ, ๐‘–, and ๐‘ง bands (left to
right, top down respectively). The lines of best fit (blue dashed) are in excellent agreement with the expected (1 + ๐‘ง) time dilation (black dotted). The binned
data are purely to visualise rough trends in 50 data point bins. 361 SNe in the ๐‘” band passed the quality cuts described in Section 3, while the ๐‘Ÿ band has 1380
SNe, the ๐‘– band 1465, and the ๐‘ง band 1381. The reduced chi-square values, ๐œ’2
๐œˆ, of each fit (left to right, top down) are 0.537, 0.729, 0.788 and 0.896 respectively.
We first normalise the target data to the peak flux using the SALT3
fit (as with the reference curves). The free parameter in the fit is the
scaling parameter 1/๐‘ค, whereby changing this value would stretch
and squash the data relative to ๐‘กpeak (the time since peak flux) until
the ๐œ’2 is minimised. That is, we assume the SN Ia light curve of the
๐‘–th supernova is of a mathematical form similar to that described in
Goldhaber et al. (2001),
๐น๐‘– (๐‘ก) โ‰ƒ ๐‘“๐‘–
(๐‘ก โˆ’ ๐‘กpeak
๐‘ค
)
(5)
and change ๐‘ค until the data most closely matches the reference. Here,
๐‘“๐‘– (๐‘ก) corresponds to the ๐‘–th target light curve; ๐น๐‘– (๐‘ก) corresponds to
the ๐‘–th reference curve where each point is now scaled in time by
(1 + ๐‘ง) relative to ๐‘กpeak as per the results of Section 4.2.
To fit the target light curve width using its reference curve, we
minimised the ๐œ’2 value of the differences in the target flux compared
to the median reference flux in a narrow bin around time values of the
target photometry. That is, for each target light curve we minimised
๐œ’2
๐‘– =
๐‘๐‘
โˆ‘
๐‘—
( ๐‘“๐‘– ๐‘— โˆ’ Med {๐น๐‘– (๐‘ก)|โˆ€๐‘ก โˆˆ [๐‘ก๐‘– ๐‘—/๐‘ค โˆ’ ๐œ, ๐‘ก๐‘– ๐‘—/๐‘ค + ๐œ]})
2
๐œŽ2
๐‘– ๐‘—
(6)
for ๐‘๐‘ number of points in the ๐‘–th target SN light curve ( ๐‘“๐‘–). The
points in the reference curve (๐น๐‘–) bin that are averaged and compared
to each target SN flux value ( ๐‘“๐‘– ๐‘— โ€“ with error ๐œŽ๐‘– ๐‘—) are selected within
the time range [๐‘ก๐‘– ๐‘—/๐‘ค โˆ’ ๐œ, ๐‘ก๐‘– ๐‘—/๐‘ค + ๐œ]; here ๐‘ก๐‘– ๐‘— is the time since peak
brightness of each target data point scaled by the fitted width ๐‘ค, and
๐œ is the half bin width either side of the central time value ๐‘ก๐‘– ๐‘—.
During the fitting process, the bounds of this narrow bin around
each time value changes as the target data is scaled in time but remains
the same width. We chose a bin width, 2๐œ, of 4 rest-frame days (i.e.
๏ฟฝ๐œ = ๏ฟฝ2 of a central value); ideally this would be as low as practical
to maximise intrinsic similarity between the target data point position
and the reference curve slice, but needs to be large enough to provide
a sufficiently populated sample of the reference to compare to. We
find that a width of 4 days (just under the width of a minor tick span in
Fig. 4) is low enough that the reference curve does not significantly
change in flux but still contains enough points even for high/low
redshift target SNe with small reference populations. With this ๐œ = 2
value we find โ‰ณ 50 data points per time slice at the highest and lowest
redshifts, where a ๐œ = 1 yields a prohibitively small โ‰ฒ 20 data points
per slice even in the most well sampled photometric band (๐‘–-band).
In fitting the data, we did not include any target SN data points that
extended past the maximum time value in the reference curve; the
late-time light curves of SNe dwindle slowly and are less constraining
for width-measurements than those near the peak. We also omitted
any points that had observation times prior to the first reference curve
point from the fitting procedure.
MNRAS 000, 1โ€“13 (2024)

Page 8
8
White et al.
1.0
1.2
1.4
1.6
1.8
2.0
1 + z
0.5
1.0
1.5
2.0
2.5
3.0
Ligh
tcurv
e
width
w
1 + z
(1 + z)1.003๏ฟฝ0.005
Data
Binned Data
1
2
3
4
Num
b
er
of
Av
eraged
Bands
Figure 8. We show here the width value for each SNe averaged across all available bands. Since cosmological time dilation is independent of the observed band
of any SN, we are able to average the widths over observed bands to form a more robust estimate of the light curve width. This relationship of (1 + ๐‘ง)1.003๏ฟฝ0.005
time dilation (reduced chi-square ๐œ’2
๐œˆ โ‰ƒ 1.441) is comprised of the 1504 unique SNe across the 4 bandpasses, where the error bars here are the Gaussian
propagation of the errors in each band. Points are coloured according to how many bandpasses were used in computing the averaged width. A linear model fit to
the data recovers ๐‘ค = (0.988 ๏ฟฝ 0.016)(1 + ๐‘ง)+(0.020 ๏ฟฝ 0.024) (with the same ๐œ’2
๐œˆ to 4 significant figures), consistent with our power model fit above.
We note that this method of fitting is not fundamentally limited to
target SN data with pre-peak brightness observations in each band.
Given enough target data (on the order of several well space points in
time), the mapping of this data to the corresponding reference curve
phases is unique regardless of whether pre-peak data is available.
The uncertainty in each estimated width was found via Monte
Carlo uncertainty propagation, where the target data was resampled
200 times according to its Gaussian error; for each iteration we fit
the width and the final error is the standard deviation in these widths.
To provide some measure of the error in the reference curve, we
imposed an error floor of ๐œŽ๐‘š(1 + ๐‘ง) on the Monte Carlo uncertainty,
for ๐œŽ๐‘š representing the median (normalised) flux dispersion in the
reference curve. The ๐œŽ๐‘š dependence is from our rationale that the
fit width is only as good as the quality of the reference curve, and the
(1 + ๐‘ง) dependence arises from our scaling of the reference curve.
An example of how a reference curve is created is shown in Fig. 4.
We note the difference in the reference curve timescale before and
after (1 + ๐‘ง) correction, and the larger dispersion in flux between the
points at any one time prior to correction. Several examples of width
fitting and reference curve construction for low, medium, and high
redshift target SNe (in the context of the DES-SN sample) are shown
in Fig. 6.
We note that while the width fitting for the whole dataset was calcu-
lated in all four DECam bands, only the ๐‘– band data encompasses the
entire redshift range of the DES-SN sample. Due to the spectral shift-
ing inherent in redshifted data, the ๐‘” and ๐‘Ÿ filters are unable to detect
SNe at sufficiently high redshift (๐‘ง โ‰ณ 0.4 and ๐‘ง โ‰ณ 0.85 respectively)
as the observed wavelengths shift to lower emitted wavelengths (see
Fig. 2 of DES Collaboration et al. 2024) and become fainter as a re-
sult. Fig. 6 shows that fitting ๐‘ง โ‰ฒ 0.2 SNe in the ๐‘ง-band would require
negative redshifted SNe in the other bands to populate the reference;
hence there is an inherent redshift floor for ๐‘ง-band fits leaving the
๐‘–-band as the only suitable bandpass for the entire redshift range.
The widths obtained in all four bands separately are shown in
Fig. 7. We see the truncated ๐‘”, ๐‘Ÿ and ๐‘ง band data, and fit widths
consistent with the expected (1 + ๐‘ง) relationship across all bands.
The averaged widths of all the bands are shown in Fig. 8, again
showing excellent agreement with the (1 + ๐‘ง) expected theory.
As mentioned in the introduction, this method has an element of
circularity because we de-time-dilated the observed light curves to
generate each reference light curve. As a cross-check to ensure that
we are not just getting the answer we put in we repeated the analysis
without de-redshifting the data. This effectively makes the reference
light curves more noisy and wider (like the top plot of Fig. 4). If
time dilation is absent we should get a consistent ๐‘ = 0 fit in this
case. However, if (1 + ๐‘ง) time dilation is present we should find a
slope inconsistent with ๐‘ = 0 and an intercept offset from ๐‘ค = 1.0
(because the reference light curve will itself be time-dilated). We
find, as expected, that the ๐‘ = 0 result is excluded strongly by this
test (see Appendix C and Fig. C1).
5 DISCUSSION
As we see in Fig. 7, there is a clear and significant non-zero time
dilation signature in the DES SN Ia dataset, conclusively ruling out
any static universe models. Our method described in Section 4 detects
a time dilation signature in all of the ๐‘”, ๐‘Ÿ, ๐‘–, and ๐‘ง DECam bandpasses
MNRAS 000, 1โ€“13 (2024)

Page 9
Time Dilation with DES SNe Ia
9
1.0
1.5
2.0
2.5
3.0
g
Band
Width
Raw Widths
Binned Widths
1.0
1.5
2.0
2.5
3.0
r
Band
Width
1.0
1.5
2.0
2.5
3.0
i Band Width
1.0
1.5
2.0
2.5
3.0
z
Band
Width
Figure 9. Since most SNe were observed in multiple bands, the fit widths in
each band for each SNe should be intrinsically correlated as they arise from
the same event. Hence, the widths for the same target SN should show some
agreement between bands. We plot their agreement relative to the ๐‘– band
which had the most SNe pass the quality cuts. Each of the data points here
corresponds to a width in Fig. 7 of bands ๐‘”, ๐‘Ÿ, or ๐‘ง against the ๐‘– band widths.
A 1:1 dashed line is shown to represent perfect agreement and binned points
are plotted to represent the trends in the agreement.
as expected. The power-law fits to the data in each bandpass are all
consistent with the expected (1 + ๐‘ง) law to within 2๐œŽ.
Since there is a well documented stretch-luminosity relationship in
Ia light curves (Phillips 1993; Phillips et al. 1999; Kasen & Woosley
2007), it is possible that Malmquist bias could skew the data to larger
widths at high redshift where we may not see the less-luminous
SNe. Regardless, this does not greatly influence the quality of our
fits since the DES SN data extend to such high redshifts that the
intrinsic dispersion in widths is significantly smaller than the time
dilation signal. In fact, we find that the standard deviation in the
width residuals (i.e. of ๐‘ค๐‘– โˆ’ (1 + ๐‘ง๐‘–)1.003 for all SNe) within Fig. 8
is only โˆผ 0.15 (which includes observational uncertainty as well as
intrinsic stretch dispersion). At ๐‘ง โˆผ 1 we would expect a time dilation
factor of โˆผ 2 and so the contribution of time dilation far outweighs
the intrinsic light curve stretch in the supernova population.
Nicolas et al. (2021) (see also Howell et al. 2007) showed that
there is a 5๐œŽ โ€˜redshift driftโ€™ in the stretch of their unbiased multi-
survey sample of SNe Ia; that is, higher redshift SNe Ia tend to be
intrinsically wider. Although we do not see this trend in the DES-
SN5YR sample, we nevertheless quantify how such a drift could
affect a time dilation measurement (Appendix A). Given that the
drift is so small, we find its impact would be minimal even if it is
hidden in the data (|ฮ”๐‘| โ‰ฒ 0.02).
As a test of the robustness of our method, we reran all the width-
fitting code with a requirement of pre-peak observations in each light
curve. At most, this changed the power law fit by ฮ”๐‘ = โˆ’0.004 for
the ๐‘” band. The calculated ๐‘ values in the other bands were increased
by one or two thousandths (including the averaged fit of Fig. 8), or
not at all. Interestingly, including this pre-peak restriction reduced
our number of unique SNe widths by only 24 in total. This reduction
does not let us conclusively say if this method is robust at fitting
light curves without pre-peak data, and future analyses may look
at purposefully degrading the dataset (e.g. by manually removing
pre-peak data) to investigate this.
Our method creates a unique reference light curve for each SN as
a function of bandpass and redshift, and so we are able to infer a time
dilation signature no matter the photometric band. This is in contrast
to Goldhaber et al. (2001) who showed time dilation in the ๐ต-band
and suggested it would hold in the other bands (see also Wang et al.
2003). Fig. 9 compares the calculated widths in each band relative
to the ๐‘–-band sample (which has the most SNe of any DECam band).
The apparent discrepancies toward high widths in the ๐‘– band might
be explained by dust effects or noise domination in the observation of
high redshift supernovae (M๏ฟฝller et al. 2022, 2024). With that said,
we see generally broad agreement between the widths in the different
bands as expected of our source wavelength-flexible model.
To avoid de-redshifting reference light curves we devised another
method, similar in concept to method two above, that would yield a
posterior distribution of ๐‘. This entailed generating a reference curve
without first scaling it in time (as in the second method), and fitting
target data with a free ๐‘ (as opposed to ๐‘ค) to the reference data of
each constituent band. That is,
ฮ”๐‘กtarget = ฮ”๐‘กreference
( 1 + ๐‘งtarget
1 + ๐‘งreference
)๐‘
(7)
and we attempted this with the same ๐œ’2 minimisation procedure as
in equation (6). With this method we tried 1) separating the reference
data into the constituent bands, fitting the ๐œ’2 to each band reference
and minimising the weighted sum of these ๐œ’2. The sum was weighted
according to how many points made up each constituent band ref-
erence curve, and this method is preferred as it assesses the fit to
all bands fairly. We also tried 2) fitting ๐‘ using the ๐œ’2 on the entire
scaled reference (i.e. all bands together), but this is not preferred as
the reference is not necessarily composed of points equally sampled
from all bands and so can prefer fitting ๐‘ towards a single bands ref-
erence (i.e. a particular redshift). This method was abandoned overall
as we did not have enough data (even with the DES dataset) to ac-
curately fit ๐œ’2 values to each band. We expect that this procedure
would be viable in the future with an even larger SNe Ia dataset of
a comparable redshift distribution, or by using a Bayesian approach
MNRAS 000, 1โ€“13 (2024)

Page 10
10
White et al.
(which will be the topic of future work). We instead performed the
analysis with flux-scatter-minimisation and width-fitting methods as
the former does not rely on fitting target SN data and the latter fits to
a unified (in phase) reference curve composed of all available bands.
Finally, in the spirit of the results of Goldhaber et al. (2001),
we similarly state that our method with the DES dataset would dis-
favour the null hypothesis of no cosmological time dilation to a
1.003/0.005 โ‰ƒ 200๐œŽ significance. (If ๐œŽ values were still meaningful
at that extreme!) Our uncertainty estimate is a statistical uncertainty
only; in Appendix A we look at a possible systematic effect due to
evolution of the stretch of supernova light curves as a function of red-
shift. We find it to be small, with a likely upper bound of ๐œŽ
sys
๐‘
โ‰ƒ 0.01.
Even with an upper limit to the uncertainty of ๐œŽ
sys
๐‘ +๐œŽstat
๐‘
โ‰ƒ 0.015this
remains the most precise constraint on cosmological time dilation.
6 CONCLUSIONS
Using two distinct methods, we have conclusively identified (1 + ๐‘ง)
cosmological time dilation using the multi-band photometry of 1504
SNe Ia from the Dark Energy Survey that met our quality cuts. We
make this detection with the most model-independent methods yet in
the literature and with the largest survey of high redshift supernovae.
For both methods, we create a โ€˜reference curveโ€™ unique to each
supernova (and each bandpass) which describes the expected light
curve shape without accounting for the stretch variation associated
with SN Ia subtypes. Doing this relied on the scale of the available
DES data (the number of SNe, the frequency of imaging, and the
redshift range) and would not be possible with a significantly smaller
survey. Creating this reference curve only relies on the assumption
that SNe Ia should be standard candles/clocks.
Using this reference curve we show an inherent preference of
โˆผ (1 + ๐‘ง)1 time dilation in the data, first by minimising the flux
scatter in the data via a redshift-dependent temporal scaling, and
then with a more traditional light curve width estimation. The latter
allows for numerical estimates with uncertainty with which we obtain
a factor of (1+๐‘ง)1.003๏ฟฝ0.005 time dilation signature โ€“ the most precise
constraint on cosmological time dilation yet.
We discuss factors and choices that affect our fits and notably see
no indication that Malmquist bias or light-curve stretch significantly
impacts our results. Our results infer a cosmological time dilation
signature aligning strongly with the expected theory, corroborating
past findings (Leibundgut et al. 1996; Goldhaber et al. 2001; Blondin
et al. 2008; Lewis & Brewer 2023) with more SNe and at a higher
redshift than ever before.
ACKNOWLEDGEMENTS
RMTW, TMD, RCa, SH, acknowledge the support of an Australian
Research Council Australian Laureate Fellowship (FL180100168)
funded by the Australian Government. AM is supported by the ARC
Discovery Early Career Researcher Award (DECRA) project number
DE230100055.
Funding for the DES Projects has been provided by the U.S. De-
partment of Energy, the U.S. National Science Foundation, the Min-
istry of Science and Education of Spain, the Science and Technology
Facilities Council of the United Kingdom, the Higher Education
Funding Council for England, the National Center for Supercomput-
ing Applications at the University of Illinois at Urbana-Champaign,
the Kavli Institute of Cosmological Physics at the University of
Chicago, the Center for Cosmology and Astro-Particle Physics at
the Ohio State University, the Mitchell Institute for Fundamental
Physics and Astronomy at Texas A&M University, Financiadora de
Estudos e Projetos, Funda๏ฟฝ๏ฟฝo Carlos Chagas Filho de Amparo ๏ฟฝ
Pesquisa do Estado do Rio de Janeiro, Conselho Nacional de Desen-
volvimento Cient๏ฟฝfico e Tecnol๏ฟฝgico and the Minist๏ฟฝrio da Ci๏ฟฝncia,
Tecnologia e Inova๏ฟฝ๏ฟฝo, the Deutsche Forschungsgemeinschaft and
the Collaborating Institutions in the Dark Energy Survey.
The Collaborating Institutions are Argonne National Laboratory,
the University of California at Santa Cruz, the University of Cam-
bridge, Centro de Investigaciones Energ๏ฟฝticas, Medioambientales y
Tecnol๏ฟฝgicas-Madrid, the University of Chicago, University Col-
lege London, the DES-Brazil Consortium, the University of Edin-
burgh, the Eidgen๏ฟฝssische Technische Hochschule (ETH) Z๏ฟฝrich,
Fermi National Accelerator Laboratory, the University of Illinois at
Urbana-Champaign, the Institut de Ci๏ฟฝncies de lโ€™Espai (IEEC/CSIC),
the Institut de F๏ฟฝsica dโ€™Altes Energies, Lawrence Berkeley National
Laboratory, the Ludwig-Maximilians Universit๏ฟฝt M๏ฟฝnchen and the
associated Excellence Cluster Universe, the University of Michigan,
NSFโ€™s NOIRLab, the University of Nottingham, The Ohio State Uni-
versity, the University of Pennsylvania, the University of Portsmouth,
SLAC National Accelerator Laboratory, Stanford University, the Uni-
versity of Sussex, Texas A&M University, and the OzDES Member-
ship Consortium.
Based in part on observations at Cerro Tololo Inter-American
Observatory at NSFโ€™s NOIRLab (NOIRLab Prop. ID 2012B-0001;
PI: J. Frieman), which is managed by the Association of Universities
for Research in Astronomy (AURA) under a cooperative agreement
with the National Science Foundation.
The DES data management system is supported by the Na-
tional Science Foundation under Grant Numbers AST-1138766
and AST-1536171. The DES participants from Spanish institutions
are partially supported by MICINN under grants ESP2017-89838,
PGC2018-094773, PGC2018-102021, SEV-2016-0588, SEV-2016-
0597, and MDM-2015-0509, some of which include ERDF funds
from the European Union. IFAE is partially funded by the CERCA
program of the Generalitat de Catalunya. Research leading to these re-
sults has received funding from the European Research Council under
the European Unionโ€™s Seventh Framework Program (FP7/2007-2013)
including ERC grant agreements 240672, 291329, and 306478. We
acknowledge support from the Brazilian Instituto Nacional de Ci๏ฟฝn-
cia e Tecnologia (INCT) do e-Universo (CNPq grant 465376/2014-
2).
This manuscript has been authored by Fermi Research Alliance,
LLC under Contract No. DE-AC02-07CH11359 with the U.S. De-
partment of Energy, Office of Science, Office of High Energy Physics.
DATA AVAILABILITY
The data are available on Zenodo and GitHub as described in the
DES supernova cosmology paper (DES Collaboration et al. 2024)
and DES-SN5YR data release paper (Sanchez et al. 2024). The gen-
erated width fits and associated uncertainties for all 1504 SNe are
included in the analysis GitHub (see Code Availability section), as
are supplementary plots not included in the paper.
CODE AVAILABILITY
Our code in all analysis and plotting relied on the open source Python
packages NumPy (Harris et al. 2020), Matplotlib (Hunter 2007),
Pandas (pandas development team 2023), and SciPy (Virtanen et al.
MNRAS 000, 1โ€“13 (2024)

Page 11
Time Dilation with DES SNe Ia
11
2020) โ€“ specifically the Nelder-Mead algorithm described in Nelder
& Mead (1965).
The code used to generate the width fits/reference curves and all
associated figures is available at github.com/ryanwhite1/DES-Time-
Dilation.
REFERENCES
Abbott T. M. C., et al., 2018, ApJS, 239, 18
Blondin S., et al., 2008, ApJ, 682, 724
Blondin S., et al., 2012, AJ, 143, 126
Branch D., Wheeler J. C., 2017, Observational Properties. Springer Berlin
Heidelberg, Berlin, Heidelberg, pp 483โ€“517, doi:10.1007/978-3-662-
55054-0_20, https://doi.org/10.1007/978-3-662-55054-0_20
Brout D., 2019, in KITP Conference: Tensions between the Early and the Late
Universe. p. 26
Carr A., Davis T. M., Scolnic D., Said K., Brout D., Peterson E. R., Kessler
R., 2022, Publ. Astron. Soc. Australia, 39, e046
DES Collaboration et al., 2024, arXiv e-prints, p. arXiv:2401.02929
Filippenko A. V., 1997, Annual Review of Astronomy and Astrophysics, 35,
309
Flaugher B., et al., 2015, AJ, 150, 150
Foley R. J., Filippenko A. V., Leonard D. C., Riess A. G., Nugent P., Perlmutter
S., 2005, ApJ, 626, L11
Goldhaber G., et al., 1997, Observation of Cosmological Time Dilation Us-
ing Type Ia Supernovae as Clocks. Springer, Dordrecht, pp 777โ€”-784,
doi:10.1007/978-94-011-5710-0_48
Goldhaber G., et al., 2001, ApJ, 558, 359
Gupta R. P., 2023, Monthly Notices of the Royal Astronomical Society, 524,
3385
Guy J., et al., 2007, A&A, 466, 11
Harris C. R., et al., 2020, Nature, 585, 357
Howell D. A., Sullivan M., Conley A., Carlberg R., 2007, ApJ, 667, L37
Hoyle F., Fowler W. A., 1960, ApJ, 132, 565
Hunter J. D., 2007, Computing in Science & Engineering, 9, 90
Kasen D., Woosley S. E., 2007, ApJ, 656, 661
Kenworthy W. D., et al., 2021, ApJ, 923, 265
Kessler R., et al., 2015, AJ, 150, 172
Leibundgut B., Sullivan M., 2018, Space Sci. Rev., 214, 57
Leibundgut B., et al., 1996, ApJ, 466, L21
Lewis G. F., Brewer B. J., 2023, Nature Astronomy,
Matheson T., et al., 2008, AJ, 135, 1598
M๏ฟฝller A., de Boissi๏ฟฝre T., 2020, MNRAS, 491, 4277
M๏ฟฝller A., et al., 2022, MNRAS, 514, 5159
M๏ฟฝller A., et al., 2024, arXiv e-prints, p. arXiv:2402.18690
M๏ฟฝller-Bravo T. E., et al., 2022, A&A, 665, A123
Nelder J. A., Mead R., 1965, The Computer Journal, 7, 308
Nicolas N., et al., 2021, A&A, 649, A74
Norris J. P., Nemiroff R. J., Scargle J. D., Kouveliotou C., Fishman G. J.,
Meegan C. A., Paciesas W. S., Bonnell J. T., 1994, ApJ, 424, 540
Phillips M. M., 1993, ApJ, 413, L105
Phillips M. M., Lira P., Suntzeff N. B., Schommer R. A., Hamuy M., Maza
J., 1999, AJ, 118, 1766
Piran T., 1992, ApJ, 389, L45
Ruiter A. J., 2019, Proceedings of the International Astronomical Union, 15,
1โ€“15
Rust B. W., 1974, PhD thesis, Oak Ridge National Laboratory, Tennessee
Sanchez et al. B., 2024, in prep
Scolnic D., et al., 2023, ApJ, 954, L31
Smith M., et al., 2020, AJ, 160, 267
Takanashi N., Doi M., Yasuda N., 2008, MNRAS, 389, 1577
Tripp R., 1998, A&A, 331, 815
Vincenzi M., et al., 2024, arXiv e-prints, p. arXiv:2401.02945
Virtanen P., et al., 2020, Nature Methods, 17, 261
Wang L., Goldhaber G., Aldering G., Perlmutter S., 2003, ApJ, 590, 944
Wilson O. C., 1939, ApJ, 90, 634
Zwicky F., 1929, Proceedings of the National Academy of Sciences, 15, 773
pandas development team T., 2023, pandas-dev/pandas: Pandas,
doi:10.5281/zenodo.8364959, https://doi.org/10.5281/zenodo.
8364959
APPENDIX A: STRETCH DRIFT WITH REDSHIFT
There is evidence that the stretch distribution of SNe evolves with
redshift, as the fraction of older and younger progenitors evolves.
Nicolas et al. (2021) give the following relation for the evolution of
the SN stretch distribution,
๐‘ƒ(๐‘ฅ1) = ๐›ฟ(๐‘ง)N(๐œ‡1, ๐œŽ2
1
)+(1โˆ’๐›ฟ(๐‘ง))
(
๐‘ŽN(๐œ‡1, ๐œŽ2
1
)+(1 โˆ’ ๐‘Ž)N(๐œ‡2, ๐œŽ2
2
)
)
,
(A1)
where N(๐œ‡, ๐œŽ2) is a normal distribution with mean ๐œ‡ and variance
๐œŽ2 and the values of the parameters were (๐‘Ž, ๐œ‡1, ๐œ‡2, ๐œŽ1, ๐œŽ2, ๐พ) =
(0.51, 0.37, โˆ’1.22, 0.61, 0.56, 0.87); and the fraction of young su-
pernovae in the population is given by,
๐›ฟ(๐‘ง) =
(
๐พโˆ’1(1 + ๐‘ง)โˆ’2.8 + 1
)โˆ’1
.
(A2)
The distribution given by equation (A1) is shown in the upper panel
of Fig. A1 for several redshifts, where the vertical dashed lines show
the resulting change in the mean ๐‘ฅ1. The relationship between ๐‘ฅ1 and
the stretch of the supernova is given by (Guy et al. 2007),
๐‘  = 0.98 + 0.091๐‘ฅ1 + 0.003๐‘ฅ2
1
โˆ’ 0.00075๐‘ฅ3
1
(A3)
and this is shown in the lower panel of Fig. A1.
Since light curve width is directly proportional to stretch, that
means that light curves at redshift ๐‘ง = 1 should be approximately
3% wider than those at ๐‘ง = 0. This is therefore substantially sub-
dominant to the factor of 2 widening expected from time-dilation
over the same range.
Compounding this effect, supernovae with wider light curves tend
to be brighter, so selection effects might also cause you to find a shift
to wider light curves at higher redshifts.
Nevertheless, in the DES-SN5YR data there is no indication that
the mean stretch parameter ๐‘ฅ1 changes with redshift, see Fig. A2.
Despite the consistency of ๐‘ฅ1 in the DES sample we want to
quantify how large the potential drift in the light curve widths could
be if equation (A1) holds. Thankfully this over-estimation can be
readily quantified. When you make a mock data set with this intrinsic
widening included you find you would actually get a line of ฮ”๐‘ก โ‰ˆ
1.03(1 + ๐‘ง) โˆ’ 0.05 (see Fig. A3). In other words, it would change the
slope by โˆผ3%. This is in contrast to the recovered linear model fit
in the Fig. 8 caption, hence indicating that this redshift-dependent
stretch is not evident in the DES-SN5YR data.
The impact of high-redshift supernovae tending to have a few per-
cent wider stretch than their low-redshift counterparts would cause us
to slightly overestimate ๐‘. The magnitude of the impact on ๐‘ depends
on your redshift distribution, we estimate a shift of |ฮ”๐‘| โ‰ฒ 0.01 for
the DES data, and we consider this a likely upper limit to the system-
atic uncertainty on our result. Since our aim in this paper is to fit the
light curves with the minimal modelling assumptions (and since we
do not see an ๐‘ฅ1 trend in our light curve fits) we have chosen not to
correct for this trend. Instead we note that any potential effect would
only be a small deviation around the slope of ๐‘ค/(1 + ๐‘ง) โˆผ 1 that we
see.
MNRAS 000, 1โ€“13 (2024)

Page 12
12
White et al.
-3
-2
-1
0
1
2
3
x1
0.0
0.2
0.4
0.6
0.8
1.0
x1
distribution
main
secondary
0.0
0.2
0.4
0.6
0.8
1.0
1.2
1.4
Redshift,
z
0.0
0.2
0.4
0.6
0.8
1.0
1.2
1.4
Redshift, z
0.98
0.99
1.00
Mean
Stretch
Figure A1. Upper panel: Distribution of ๐‘ฅ1 values predicted by Nicolas et al.
(2021). The grey and black dashed Gaussians show the two components of
the supernova population. The coloured lines show the total distribution for
several different redshifts. The vertical dashed lines show the mean of the
redshift distribution (in the same colours as the legend). One can see that the
mean drifts from low to high ๐‘ฅ1 as redshift increases. Lower panel: The black
line shows the evolution of the mean stretch (๐‘ ) of the supernova population
with redshift, where colours match the redshifts in the upper legend. The
intrinsic light curve width is proportional to ๐‘ , and therefore light curves are
expected to be about 3% wider at ๐‘ง = 1 than at ๐‘ง = 0. This is much less than
the factor of two widening due to time dilation.
0.2
0.4
0.6
0.8
1.0
Redshift, z
0.8
0.9
1.0
1.1
1.2
Stretch
Best fit: s = 0.0030z + 0.96
โˆ’3
โˆ’2
โˆ’1
0
1
2
-0.2
-0.1
0.0
0.1
0.2
SN
Colour
Figure A2. The distribution of stretch in the DES-SN5YR data as a function
of redshift (calculated from the SALT3 fitted ๐‘ฅ1 values using equation A2),
with ๐‘ฅ1 shown on the right axis. Fitting a straight line to this distribution
shows no significant trend in the stretch with redshift.
1.0
1.2
1.4
1.6
1.8
2.0
2.2
2.4
1 + z
1.0
1.2
1.4
1.6
1.8
2.0
2.2
2.4
Ligh
t
Curv
e
Width,
w
Time dilation only: โˆ†t = (1 + z)
with stretch drift: โˆ†t = 1.03(1 + z) - 0.05
no time dilation: โˆ†t = 0
Figure A3. The effect of adding the predicted stretch evolution of SNe Ia vs
redshift is to cause the width vs redshift plot to be slightly steeper. If this
result is present we therefore expect to slightly overestimate ๐‘, as we will
attribute that widening to time dilation.
APPENDIX B: REFERENCE CURVE SELECTION
DERIVATION
We begin with the definition of redshift,
1 + ๐‘ง =
๐œ†๐‘œ
๐œ†๐‘’
(B1)
where ๐‘ง is the source redshift, ๐œ†๐‘œ is the observed wavelength of
light and ๐œ†๐‘’ is the original emission wavelength of light. If we are
sampling from a band with central wavelength ๐œ†๐‘Ÿ during reference
curve construction, we need to find a central redshift ๐‘ง๐‘Ÿ for our
reference SNe selection, given that we are fitting the target light
curve in a band of central wavelength ๐œ† ๐‘“ . The idea is to match
the original emitted wavelengths and so we can divide by another
instance of equation (B1),
1 + ๐‘ง
1 + ๐‘ง๐‘Ÿ
=
๐œ† ๐‘“ /๐œ†๐‘’
๐œ†๐‘Ÿ /๐œ†๐‘’
=
๐œ† ๐‘“
๐œ†๐‘Ÿ
(B2)
Then, we can rearrange to find an expression for our target redshift
๐‘ง๐‘Ÿ ,
๐œ†๐‘Ÿ (1 + ๐‘ง) = ๐œ† ๐‘“ (1 + ๐‘ง๐‘Ÿ )
(B3)
๐‘ง๐‘Ÿ =
๐œ†๐‘Ÿ (1 + ๐‘ง)
๐œ† ๐‘“
โˆ’ 1
(B4)
We can then append a term ๏ฟฝฮ”๐‘ง on equation( B4) to give us a range of
applicable redshift values as in Section 4.1. Finally, it is useful in the
broader context of the paper (and Fig. 3) to show this redshift range
in terms of some fraction of the band FWHM of the band that the
target SN was observed in, ๐›ฟฮ”๐œ† ๐‘“ . To do this we set ฮ”๐‘ง = ๐›ฟฮ”๐œ† ๐‘“ /๐œ† ๐‘“
and shift the term into the fraction within equation (B4),
๐‘ง๐‘Ÿ =
๐œ†๐‘Ÿ (1 + ๐‘ง) ๏ฟฝ ๐›ฟฮ”๐œ† ๐‘“
๐œ† ๐‘“
โˆ’ 1
(B5)
which yields the redshift sampling range of equation (3) that we use
in the analysis.
MNRAS 000, 1โ€“13 (2024)

Page 13
Time Dilation with DES SNe Ia
13
APPENDIX C: NULL TEST โ€” NO DE-REDSHIFTING OF
REFERENCE LIGHT CURVES
To confirm that our method is able to rule out no time dilation we
repeated the analysis without de-redshifting the reference curves.
That means that the reference curves would look like the top panel of
Fig. 4. If the data were not time dilated, then we should fit consistently
๐‘ = 0 in this case. Fig. C1 clearly shows that this null test fails. Be-
cause time-dilation is present in the data, it means that the reference
curves are now wider than they should be โ€” making the measured
light curve widths narrower. Despite this, the trend for higher-redshift
light curves to be wider persists, in very strong contradiction to the
no-time-dilation hypothesis.
AFFILIATIONS
1 School of Mathematics and Physics, The University of Queensland,
QLD 4072, Australia
2 Sydney Institute for Astronomy, School of Physics, A28, The Uni-
versity of Sydney, NSW 2006, Australia
3 Centre for Gravitational Astrophysics, College of Science, The
Australian National University, ACT 2601, Australia
4 The Research School of Astronomy and Astrophysics, Australian
National University, ACT 2601, Australia
5 Department of Physics & Astronomy, University College London,
Gower Street, London, WC1E 6BT, UK
6 Cerro Tololo Inter-American Observatory, NSFโ€™s National Optical-
Infrared Astronomy Research Laboratory, Casilla 603, La Serena,
Chile
7 Laborat๏ฟฝrio Interinstitucional de e-Astronomia - LIneA, Rua Gal.
Jos๏ฟฝ Cristino 77, Rio de Janeiro, RJ - 20921-400, Brazil
8 Fermi National Accelerator Laboratory, P. O. Box 500, Batavia, IL
60510, USA
9 Department of Physics, University of Michigan, Ann Arbor, MI
48109, USA
10 Departamento de F๏ฟฝsica Te๏ฟฝrica and Instituto de F๏ฟฝsica de Part๏ฟฝcu-
las y del Cosmos (IPARCOS-UCM), Universidad Complutense de
Madrid, 28040 Madrid, Spain
11 Institute of Cosmology and Gravitation, University of Portsmouth,
Portsmouth, PO1 3FX, UK
12 University Observatory, Faculty of Physics, Ludwig-Maximilians-
Universit๏ฟฝt, Scheinerstr. 1, 81679 Munich, Germany
13 Center for Astrophysics | Harvard & Smithsonian, 60 Garden
Street, Cambridge, MA 02138, USA
14 Department of Astronomy and Astrophysics, University of
Chicago, Chicago, IL 60637, USA
15 Kavli Institute for Particle Astrophysics & Cosmology, P. O. Box
2450, Stanford University, Stanford, CA 94305, USA
16 SLAC National Accelerator Laboratory, Menlo Park, CA 94025,
USA
17 Instituto de Astrofisica de Canarias, E-38205 La Laguna, Tenerife,
Spain
18 INAF-Osservatorio Astronomico di Trieste, via G. B. Tiepolo 11,
I-34143 Trieste, Italy
19 Institut de F๏ฟฝsica dโ€™Altes Energies (IFAE), The Barcelona Insti-
tute of Science and Technology, Campus UAB, 08193 Bellaterra
(Barcelona) Spain
20 Hamburger Sternwarte, Universit๏ฟฝt Hamburg, Gojenbergsweg
112, 21029 Hamburg, Germany
21 Centro de Investigaciones Energ๏ฟฝticas, Medioambientales y Tec-
nol๏ฟฝgicas (CIEMAT), Madrid, Spain
22 Department of Physics, IIT Hyderabad, Kandi, Telangana 502285,
India
23 Jet Propulsion Laboratory, California Institute of Technology,
4800 Oak Grove Dr., Pasadena, CA 91109, USA
24 Institute of Theoretical Astrophysics, University of Oslo. P.O. Box
1029 Blindern, NO-0315 Oslo, Norway
25 Kavli Institute for Cosmological Physics, University of Chicago,
Chicago, IL 60637, USA
26 Instituto de Fisica Teorica UAM/CSIC, Universidad Autonoma de
Madrid, 28049 Madrid, Spain
27 Institut dโ€™Estudis Espacials de Catalunya (IEEC), 08034
Barcelona, Spain
28 Institute of Space Sciences (ICE, CSIC), Campus UAB, Carrer de
Can Magrans, s/n, 08193 Barcelona, Spain
29 Centre for Astrophysics & Supercomputing, Swinburne Univer-
sity of Technology, Victoria 3122, Australia
30 Center for Astrophysical Surveys, National Center for Supercom-
puting Applications, 1205 West Clark St., Urbana, IL 61801, USA
31 Department of Astronomy, University of Illinois at Urbana-
Champaign, 1002 W. Green Street, Urbana, IL 61801, USA
32 Santa Cruz Institute for Particle Physics, Santa Cruz, CA 95064,
USA
33 Center for Cosmology and Astro-Particle Physics, The Ohio State
University, Columbus, OH 43210, USA
34 Department of Physics, The Ohio State University, Columbus, OH
43210, USA
35 Australian Astronomical Optics, Macquarie University, North
Ryde, NSW 2113, Australia
36 Lowell Observatory, 1400 Mars Hill Rd, Flagstaff, AZ 86001,
USA
37 Department of Physics and Astronomy, University of Pennsylva-
nia, Philadelphia, PA 19104, USA
38 Departamento de F๏ฟฝsica Matem๏ฟฝtica, Instituto de F๏ฟฝsica, Univer-
sidade de S๏ฟฝo Paulo, CP 66318, S๏ฟฝo Paulo, SP, 05314-970, Brazil
39 George P. and Cynthia Woods Mitchell Institute for Fundamental
Physics and Astronomy, and Department of Physics and Astronomy,
Texas A&M University, College Station, TX 77843, USA
40 LPSC Grenoble - 53, Avenue des Martyrs 38026 Grenoble, France
41 Instituci๏ฟฝ Catalana de Recerca i Estudis Avan๏ฟฝats, E-08010
Barcelona, Spain
42 Department of Astrophysical Sciences, Princeton University, Pey-
ton Hall, Princeton, NJ 08544, USA
43 Observat๏ฟฝrio Nacional, Rua Gal. Jos๏ฟฝ Cristino 77, Rio de Janeiro,
RJ - 20921-400, Brazil
44 Department of Physics, Carnegie Mellon University, Pittsburgh,
Pennsylvania 15312, USA
45 Department of Physics and Astronomy, Pevensey Building, Uni-
versity of Sussex, Brighton, BN1 9QH, UK
46 School of Physics and Astronomy, University of Southampton,
Southampton, SO17 1BJ, UK
47 Computer Science and Mathematics Division, Oak Ridge National
Laboratory, Oak Ridge, TN 37831
48 Department of Physics, Duke University Durham, NC 27708, USA
49 Universit๏ฟฝ Grenoble Alpes, CNRS, LPSC-IN2P3, 38000 Greno-
ble, France
50 Department of Astronomy, University of California, Berkeley, 501
Campbell Hall, Berkeley, CA 94720, USA
51 Lawrence Berkeley National Laboratory, 1 Cyclotron Road, Berke-
ley, CA 94720, USA
This paper has been typeset from a TEX/LATEX file prepared by the author.
MNRAS 000, 1โ€“13 (2024)

Page 14
14
White et al.
0.4
0.6
0.8
1.0
1.2
1.4
1.6
1.8
2.0
Ligh
tcurv
e
width
w
g
1 + z
(1 + z)โˆ’0.346๏ฟฝ0.039
No Time Dil
Data
Binned Data
ะณ
(1 + z)โˆ’0.016๏ฟฝ0.013
1.0
1.2
1.4
1.6
1.8
2.0
1 + z
0.4
0.6
0.8
1.0
1.2
1.4
1.6
1.8
2.0
Ligh
tcurv
e
width
w
i
(1 + z)0.087๏ฟฝ0.012
1.0
1.2
1.4
1.6
1.8
2.0
1 + z
ฮถ
(1 + z)0.221๏ฟฝ0.011
Figure C1. Light curve widths measured with respect to a reference curve that has not been de-time-dilated. We nevertheless still see a persistent trend of
increasing light curve width with redshift. The vertical offset from the (1 + ๐‘ง) line arises because the non-de-time-dilated reference curves are wider than
rest-frame light curves, i.e. this offset is yet another indication of time dilation. The black horizontal dashed line indicates no time dilation and the blue dashed
lines are (poor) (1 + ๐‘ง)๐‘ model fits to the data. If there was no time dilation, these fits would be horizontal lines with ๐‘ = 0.
MNRAS 000, 1โ€“13 (2024)