Early Universe Physics Insensitive and Uncalibrated Cosmic Standards: Constraints on Ωm and Implications for the Hubble Tension

, , and

Published 2021 October 25 © 2021. The American Astronomical Society. All rights reserved.
, , Citation Weikang Lin et al 2021 ApJ 920 159 DOI 10.3847/1538-4357/ac12cf

Download Article PDF
DownloadArticle ePub

You need an eReader or compatible software to experience the benefits of the ePub3 file format.

0004-637X/920/2/159

Abstract

To further gain insight into whether pre-recombination models can resolve the Hubble tension, we explore constraints on the evolution of the cosmic background that are insensitive to early universe physics. The analysis of the CMB anisotropy has been thought to highly rely on early universe physics. However, we show that the fact that the sound horizon at recombination being close to that at the end of the drag epoch is insensitive to early universe physics. This allows us to link the absolute sizes of the two horizons and treat them as free parameters. Jointly, the CMB peak angular size, baryon acoustic oscillations, and Type Ia supernovae can be used as early universe physics insensitive and uncalibrated cosmic standards, which measure the cosmic history from recombination to today. They can set strong and robust constraints on the post-recombination cosmic background, especially the matter density parameter with Ωm = 0.302 ± 0.008 (68% C.L.), assuming a flat Λ cold dark matter universe after recombination. When we combine these with other nonlocal observations, we obtain several constraints on H0 with significantly reduced sensitivity to early universe physics. These are all more consistent with the Planck 2018 result than the local measurement results such as those based on Cepheids. This suggests a tension between the post-recombination, but nonlocal, observations, and the local measurements that cannot be resolved by modifying pre-recombination early universe physics.

Export citation and abstract BibTeX RIS

1. Introduction

The standard cosmological model—a spatially flat universe with a cosmological constant and cold dark matter (ΛCDM)—has been successful overall in explaining and predicting cosmological observations (Cooke et al. 2018; Scolnic et al. 2018; DES Collaboration et al. 2019a; Planck Collaboration et al. 2020; eBOSS Collaboration et al. 2021). However, recently several tensions on cosmological parameters measured via different observations have been reported to various extents; see, e.g., Riess et al. (2021), Asgari et al. (2021), and Joudaki et al. (2020). The currently most hotly debated tension is the Hubble tension: the difference in the Hubble parameter between the results from methods based on ΛCDM (such as the Planck cosmic microwave background (CMB) anisotropy and/or baryon acoustic oscillation (BAO)) and the results from the method of the distance ladder (especially the Cepheid-based local measurement) (Planck Collaboration et al. 2020; Di Valentino et al. 2021; Riess et al. 2021).

It is important to explore different methods to measure or infer H0 (Abbott et al. 2017; Moresco & Marulli 2017; Birrer et al. 2019; Domínguez et al. 2019) because this helps us conclude whether the Hubble tension has an origin beyond the Standard Model or which systematic effects need more investigation. Although most methods of determining H0 depend on an underlying cosmological model, the reliance varies and the constraints exhibit different degeneracy directions, e.g., in the H0–Ωm plane. Based on the standard ΛCDM model, it was shown that most constraints from different types of observations consistently overlap on some common parameter region in the H0–Ωm plane (Lin et al. 2020), which to some extent disfavors a beyond the Standard Model explanation.

Nonetheless, the Hubble tension is often described as a discrepancy between the early and the late 4 universe (Verde et al. 2019). This is partially because the two powerful methods with CMB and BAO assume standard pre-recombination physics and partially because observations such as Type Ia supernovae (SNe Ia) and BAO (even without assuming the standard pre-recombination physics) strongly constrain late-time deviations from ΛCDM (Evslin et al. 2018; Aylor et al. 2019; Krishnan et al. 2021). It is suggested that models that shorten the early expansion history and the sound horizon scale are promising (or the least unlikely) (Knox & Millea 2020). A number of examples of this class of models have been proposed, such as a form of dark energy that appeared between matter-radiation equality and recombination (Agrawal et al. 2019; Poulin et al. 2019; Niedermann & Sloth 2020; Sakstein & Trodden 2020; Smith et al. 2020), dark matter that partially interacts with dark radiation (Chacko et al. 2016; Choi et al. 2021), neutrinos with self-interaction (Das & Ghosh 2020; Kreisch et al. 2020) or interaction with dark matter (Ghosh et al. 2020), early modified gravity (Braglia et al. 2021), and the consideration of primordial magnetic field (Jedamzik & Pogosian 2020). Works are being undertaken to confront those proposals with different observations, and currently it still remains controversial whether or not they are viable solutions to the Hubble tension (Raveri et al. 2017; Archidiacono et al. 2020; Brinckmann et al. 2021; Hill et al. 2020; Ivanov et al. 2020; Philcox et al. 2020; Smith et al. 2020; Choudhury et al. 2021; D'Amico et al. 2021; Haridasu et al. 2021; Jedamzik et al. 2021).

Given the effort devoted to the early universe resolutions to the Hubble tension, it is especially important to explore methods that determine H0 independent of or insensitive to the early universe physics. There are already several methods that only involve post-recombination physics such as the cosmic chronometers (Jimenez & Loeb 2002), measuring the redshift dependence of γ-ray optical depth (Domínguez & Prada 2013), the strong-lensing time-delay technique (Refsdal 1964), inferring cosmic age from old stars (Jimenez et al. 2019), and the gravitational wave multi-messenger method (Abbott et al. 2017). But currently the precision of H0 from those methods is relatively low because of their own astrophysical uncertainties and their background degeneracy in the H0–Ωm plane. To mitigate the latter factor, some of those methods (especially the cosmic chronometers and the γ-ray optical depth) are combined with other observations to break the background degeneracy (Moresco & Marulli 2017; Domínguez et al. 2019). However, in doing so, either the constraining power of the other observation is not strong enough (e.g., combining with the uncalibrated SNe Ia), or a dependence on early universe physics (e.g., combining with BAO or a CMB prior on Ωm from Planck) is introduced. An important aim in this work is to develop a method that provides a strong constraint on the evolution of the cosmic background to break the degeneracy with other observations and yet is insensitive to early universe physics (i.e., the dependence on early universe physics is significantly reduced). The dependence on early universe physics may be traded with that on other potentially unaccounted for astrophysics effect. Thus, we investigate a number of independent probes that are different in methodology.

It is well known that when the sound horizon scale is treated as a free parameter, it is degenerate with H0, but the Hubble-distance normalized evolution of the cosmic background can still be constrained (Aylor et al. 2019). Several recent works have applied uncalibrated 5 BAO and SNe Ia jointly with large-scale structure to obtain sound horizon-independent constraints on H0 (Baxter & Sherwin 2020; Pogosian et al. 2020; Philcox et al. 2021). But uncalibrated BAO does not provide a strong enough constraint on, e.g., Ωm compared to that from Planck.

On the other hand, information obtained from the CMB has been thought to depend strongly on early universe physics. However, as we will show, the sound horizon at the recombination epoch (r*), related to the angular sizes of the CMB acoustic peaks (Planck Collaboration et al. 2020), has a property that significantly reduces the model sensitivity: namely, its difference (normalized by the Hubble distance) from the sound horizon at the end of the baryonic-drag epoch 6 (rd) is small, which we denote as ΔrH0. Both horizons share a common expression before recombination, making their difference to nearly cancel out the dependence on early universe physics. Furthermore, while the redshifts of these two epochs differ by only Δzs ≈ 30, the redshifts at which these two epochs imprinted observable effects—peaks in CMB power spectra and BAO extracted from two-point correlation functions of tracers of the underlying matter density fluctuation at large scales—differ by Δzs ∼ 1000. The former narrow redshift separation makes ΔrH0 both very small and insensitive even to nonstandard physics happening between the two epochs, while the latter large separation provides a long lever arm to effectively constrain model parameters (such as Ωm) once the two angular scales are used jointly.

We will treat the sound horizon as a free parameter for CMB and jointly analyzes the CMB acoustic peak angular scale, θcmb, and BAO. We will show that the aforementioned tight relation between the two horizon scales allows us to jointly analyze θcmb and BAO in a robust way insensitive to early universe physics. Such an analysis will be shown to have a constraining power on Ωm comparable to that from Planck in a full ΛCDM-based analysis. We call such a joint analysis using θcmb, BAO, and SNe Ia "early universe physics insensitive and uncalibrated cosmic standards" (EUPIUCS, or UCS for simplicity).

Reanalyses of observations relaxing some strong assumptions are very important. After relaxing some assumptions, if the result is inconsistent with the original one, it would be an indication that those assumptions could be problematic. On the contrary, if the result is consistent with the original one, it would cause difficulties for nonstandard models/considerations trying to modify those assumptions to resolve a tension.

This work is organized as follows. In Section 2, we describe the methods of our joint analysis of UCS, in particular the connection between θcmb and BAO. In Section 3, we present the constraint on the evolution of the post-recombination background, especially the strong constraint on Ωm, and discuss the robustness of this analysis against any possible early universe nonstandard physics. Then in Section 4, we combine the UCS with other early universe physics independent or insensitive observations to break the background degeneracy and obtain several constraints of H0. In Section 5, we discuss how to generalize the analysis of UCS to test post-recombination cosmological models independent of early universe physics. Finally, we summarize and conclude in Section 6. The Appendix contains some of the details of the analyses.

Throughout this work, we adopt units in which the speed of light is c = 1. So, e.g., H0 has dimensions of inverse distance unless its unit is explicitly specified.

2. Methods

It is well known that based on the standard ΛCDM model, uncalibrated SNe Ia and (late-time) BAO can constrain Ωm (Scolnic et al. 2018; Aylor et al. 2019). For SNe Ia, it is the relative change of the apparent brightness at different redshifts that constrains the matter fraction in the standard ΛCDM model, which determines the relative change of the Hubble parameter. For uncalibrated BAO, it is the change (at different redshifts) of the angular size or the redshift span of the sound horizon scale at the end of the drag epoch that sets the constraint. Those constraints are independent of early universe physics. While detailed descriptions of those two analyses can be found in the literature (Scolnic et al. 2018; Aylor et al. 2019), we provide a discussion in the Appendix using our own notation to unify these two analyses as well as that of the CMB acoustic peak. 7 Since the absolute magnitude of SNe Ia (M0) and the absolute (comoving) scale of the sound horizon at the drag epoch rd are both degenerate with the Hubble constant, we treat the combinations ${ \mathcal M }\equiv {M}_{0}-5{\mathrm{log}}_{10}(10\,\mathrm{pc}\times {H}_{0})$ and rd H0 as free parameters.

The angular size (θcmb) of the CMB acoustic peaks has been measured precisely (Planck Collaboration et al. 2020). Earlier, it has been proposed to use θcmb (along with the drift parameter R that we will not use) to, e.g., test late-time dark energy models (Bond et al. 1997). But this has already assumed that the scale of the sound horizon (r*) at recombination is calibrated by early universe physics in the sense that r* is calculated assuming a standard cosmological model at early times. Like rd for BAO, r* for θcmb is also degenerate with H0. Therefore, we also treat the combination r* H0 as a free parameter. But unlike BAO, which has measurements at multiple redshifts, θcmb is only measured at one single redshift, z*. Therefore, when analyzed alone, θcmb is not able to simultaneously constrain r* H0 and Ωm.

The underlying principle of our analysis is the fact that rd and r* are tightly related to each other. As shown in Appendix A.4, by taking the difference between rd H0 and r* H0 we have

Equation (1)

where z* and zd are the redshifts at recombination and at the end of the drag epoch, respectively, and E(z) represents the normalized total energy evolution as a function of redshift (also see Appendix A.1 for the definition of E(z)). We can thus treat r* H0 as a free parameter and link rd H0 to r* H0 by ΔrH0. The value of ΔrH0 is much smaller than r* H0 and this fact is almost independent of early universe physics. Therefore, a substantial change to ΔrH0 is required in order to alter our analysis. But the value of ΔrH0 itself is quite insensitive to early universe physics. This is because by taking the difference between rd H0 and r* H0 the effects from early universe physics get largely canceled. We discuss the robustness of these properties in Section 3.2.

We will first perform our default analysis where we use some standard assumptions (described below) to model and calculate ΔrH0. Later in Section 3.2, we will quantify the sensitivity to changes of those assumptions and show that the results obtained are robust against any reasonable modification of our standard assumptions. We will also perform analyses separately using the Planck (Planck Collaboration et al. 2020) and the Atacama Cosmology Telescope and Wilkinson Microwave Anisotropy Probe (ACT+WMAP; Aiola et al. 2020) results of θcmb and the priors on z* and Δzs, in order to show that our results are robust against possible systematic errors in each CMB experiment. We assume a flat-ΛCDM universe after recombination.

2.1. Our Standard Analysis

The standard assumptions in our default analysis are as follows. First, to calculate the sound speed, we need the reduced baryon fraction Ωb h2, for which we use the big bang nucleosynthesis (BBN) constraint Ωb h2 = 0.0222 ± 0.0005 (Cooke et al. 2018). We also need to know z* and zd, which we separately obtain from Planck (Planck Collaboration et al. 2020) (z* = 1089.99 ± 0.28 and z*zd = 30.27 ± 0.57) and from ACT+WMAP (Aiola et al. 2020) (z* = 1089.91 ± 0.42 and z*zd = 29.95 ± 0.75). Next, as we are assuming the standard ΛCDM after recombination, the function E(z) and the sound speed cs take their standard forms; see Equations (A3) and (A12). In particular, for the function E(z) we consider a universe with a cosmological constant filled with pressureless matter (CDM and baryons), photons, and neutrinos (one massless and two massive with a normal mass hierarchy).

The joint analysis of the θcmb likelihood and the BAO likelihood will break the degeneracy in the Ωmrd H0 plane and significantly improve the constraint on Ωm compared to the uncalibrated BAO alone. We illustrate how this works as follows. Recall that the angular size of the sound horizon at the drag epoch is given by

Equation (2)

where fM defined in Equation (A1) is the Hubble-distance normalized comoving distance. BAO uses the angular size and the redshift span of the (drag-epoch) sound horizon, θd and ${\rm{\Delta }}{z}_{{r}_{{\rm{d}}}}$, at different effective redshifts to constrain Ωm, which determines the relative change of the Hubble parameter with redshift; see Appendix A.2 for a discussion. With our standard assumptions, the difference ΔrH0 is ∼6 × 10−4 according to Equation (1), given that BAO alone can already constrain Ωm with a certain precision. With ΔrH0 roughly known, the measurement of θcmb can be viewed as another measurement of θd at z* (with an offset of ${\rm{\Delta }}\theta \equiv {\theta }_{{\rm{d}}}({z}_{* })-{\theta }_{\mathrm{cmb}}=\tfrac{{\rm{\Delta }}{{rH}}_{0}}{{f}_{{\rm{M}}}({z}_{* })}$). In Figure 1, we plot the measured or inferred values of θd at different redshifts along with the theoretical prediction according to the best-fit model. 8 We can see that the prediction matches the measured/inferred θd at different redshifts quite well. The precise measurement of θcmb (and then the inferred θd) adds a powerfully constraining anchor at a large redshift, so that the constraint on the relative change of the Hubble parameter can be significantly tightened, and hence so can the Ωm in the standard ΛCDM model. For the same (comoving) rd, the angular size θd asymptotically approaches the minimum value at large redshift, as indicated in the upper panel of Figure 1. This gives the advantage that z* has to be significantly incorrect in order to affect the analysis.

Figure 1.

Figure 1. The observed/inferred angular sizes θd of the drag-epoch sound horizon at different redshifts. Top: the solid line is the prediction of θd from the best-fit model. The dotted line represents the predicted asymptotic θd at infinite redshift. The blue circles are the observed θd's from some late-time BAOs. The green triangles are the inferred θd's for late-time BAO that only provides $\tfrac{{r}_{{\rm{d}}}{H}_{0}}{{f}_{{\rm{V}}}(z)}$. The red square is the inferred θd(z*) from the CMB. To get θd(z*) for the CMB, we added to θcmb the inferred angular size difference (Δθ) between the sound horizon at the end of the drag epoch and at recombination according to the best-fit model. The uncertainty of the inferred θd(z*) includes the derived uncertainty of Δθ. The error bars are too small to show. Bottom: the ratios of the observed and predicted θd's. The thin lines represent the predictions of the model with a subset of parameter points sampled in the Markov chain Monte Carlo (MCMC) analysis.

Standard image High-resolution image

BAO measurements at different redshifts give different degeneracy directions in the Ωmrd H0 plane (Cuceu et al. 2019). This also applies to θcmb. In Figure 2, we plot different constraints in the Ωmrd H0 plane at different redshifts, including the one inferred from θcmb. At lower redshifts, BAO measurements are more sensitive to rd H0, because fM(z) (and also E(z)) is not sensitive to Ωm at low redshifts. This is shown in Figure 2 as the constraint from the six degree field galaxy survey and main galaxy sample (6df+MGS) (zeff < 0.2) in the Ωmrd H0 plane is almost horizontal. At larger redshifts, BAO measurements are sensitive to both rd H0 and Ωm; this is shown in Figure 2 as constraints from BAO at higher redshifts (zeff > 0.3). The CMB constraint is diagonal in the Ωmrd H0 plane. Therefore, the tight constraint from θcmb breaks the degeneracy in the Ωmrd H0 plane of the BAO measurements at lower redshifts, and significantly improves the constraint on Ωm.

Figure 2.

Figure 2. Degeneracy direction of constraints on the Ωmrd H0 plane from BAO at different redshifts and θcmb. The θcmb and BAO constraints break the degeneracy to provide a constraint on, e.g., Ωm. See Appendix A.2 for a detailed descriptions of BAOs.

Standard image High-resolution image

To constrain Ωm, besides θcmb and uncalibrated BAO, we will also consider the Pantheon catalog of SNe Ia (without calibration). We denote the joint likelihood of all these UCS data as ${{ \mathcal L }}_{\mathrm{UCS}}$. In our standard analysis, the free parameters are Ωm, r* H0, ${ \mathcal M }$, and h (a very weak dependence; see Appendix A.1), with some assumed prior on z*, Δzs, and Ωb h2.

3. Results

3.1. A Tight Constraint on Matter Density Parameter Insensitive to Early Universe Physics

As mentioned earlier, even uncalibrated, standard rulers and candles can put a constraint on Ωm (Scolnic et al. 2018; Aylor et al. 2019). With θcmb added according to the method described above, the constraint is now much stronger. In Figure 3, we show the constraints on Ωm set by different combinations of those UCS. We first reproduced the results from Pantheon SNe Ia and uncalibrated BAO, respectively, shown by the green error bars (1σ) in the left panel of Figure 3. In the middle panel, we show the constraints from the combination of θcmb and three selected late-time BAO. We adopt two different θcmb results, one from Planck temperature and polarization (TTTEEE+lowE) with 9

Equation (3)

shown in blue (Planck Collaboration et al. 2020); and the other from ACT+WMAP with

Equation (4)

shown in red (Aiola et al. 2020). Among these three constraints, the one from the combination of θcmb and the BAO from the Sloan Digital Sky Survey Data Release 12 galaxy clustering (SDSS DR12 GC; Alam et al. 2017) is the tightest, which is due to both the degeneracy breaking and the strong constraining power of this late-time BAO. Nonetheless, the constraints on Ωm via those different combinations are consistent with each other and with the SN Ia-only and uncalibrated-BAO-only constraints.

Figure 3.

Figure 3. Constraints on Ωm obtained from different combinations of uncalibrated SN, BAO, and θcmb. The two green error bars are the constraints on Ωm from Pantheon SNe Ia and all late-time BAO uncalibrated. The middle three are the results from uncalibrated θcmb jointly with some individual late-time BAO. Different colors represent results using θcmb from Planck and ACT+WMAP, respectively. The error bar on the right is the final result with all UCS and the second one from the right excludes the SDSS DR12 GC BAOs (zeff = 0.38, 0.51, 0.68). The shaded horizontal band represents the final result (Ωm = 0.302 ± 0.008) with Planck θcmb. All results are consistent with each other.

Standard image High-resolution image

We then produce the joint constraint with all UCS. The final result is shown in the right panel of Figure 3 with

Equation (5)

using the Planck constraint on θcmb. This is a strong constraint on Ωm. In fact, with the same data the full analysis in ΛCDM (assuming the standard modeling of the sound horizon) gives a constraint on Ωm that is only a little tighter, Ωm = 0.310 ± 0.006. Although the average is somewhat lower, our constraint on Ωm is consistent with the full analysis that assumes the standard ΛCDM model for the entire cosmic history.

Compared to the full standard analysis of CMB anisotropy, our method using UCS is only based on the geometrical information from CMB anisotropy and relies on many fewer assumptions. First, as explained, our method is insensitive to early universe physics. Second, our method is less vulnerable to systematic bias in the CMB observations. Indeed, in the full analysis of the CMB, the matter fraction is constrained by the scale dependence of the CMB power spectra (Planck Collaboration et al. 2020). This renders the constraint of Ωm correlated with, e.g., the spectral index ns. Incidentally, there have been discussions on the possibility of some small inconsistency between the large and small scale spectra in the Planck CMB power spectra, which might be due to unaccounted for systematic errors or more radically some beyond the Standard Model physics (Addison et al. 2016); but also see Planck Collaboration et al. (2016) for a discussion. Our method is clearly free of those potential biases.

As we mentioned earlier, the combination of θcmb and the SDSS DR12 GC BAO provides the strongest constraint on Ωm. If there is some systematic error in the SDSS DR12 GC BAO, the result would be biased. Therefore, we also produce a joint constraint on Ωm using all UCS except for the SDSS DR12 GC, which is shown in the second to the right set of error bars in Figure 3. As we can see, even without the SDSS DR12 GC BAO, the result is fully consistent with the one using all UCS.

3.2. Robustness against Early Universe Physics

In the last subsection, we treat r* as a free parameter, which have already relaxed the most of the strong assumptions of early universe physics. In our standard analysis, we have only made an assumption that the difference between r* H0 and rd H0 can be calculated according to the standard ΛCDM model. We have assumed some priors on z* and Δzs from CMB observations and Ωb h2 from BBN, so that we can calculate ΔrH0. One may worry that these may render our results dependent upon early universe physics. We now show that results obtained based on this standard analysis are robust against changes caused by possible nonstandard early universe physics. We quantify how significantly these assumptions need to be modified to substantially affect our results. We also discuss several possibilities that may be contemplated in future model building to circumvent the arguments and constraints derived in this paper.

Recall that the underlying reason we can jointly analyze the θcmb and uncalibrated BAO likelihoods is that the two horizon scales (r* and rd) are very close to each other but the measurements are so separated in redshift that together they can provide a stringent constraint on the evolution of the post-recombination background. We discuss this in detail in the following.

First, the fact that the difference between r* H0 and rd H0 is very small is important and quite robust. Because it is small, the difference between the two angles θd (z*) and θcmb is comparable to or even smaller than the error on θd (z*) extrapolated from the BAO measurements alone. Therefore, the CMB measurement on θcmb can be viewed as another statistically significant measurement of the BAO sound horizon at a very high redshift with an uncertainty mainly caused by the uncertainty of Δθ.

Second, because the value of Ωm continuously and monotonically impacts the evolution of the background from z ∼ 1000 to z ∼ 1, with the same relative errors, the constraining power on Ωm from the data-derived θd values grows if the redshifts of these data are further separated. Consequently, since z* is so high, the effect on the constraint on Ωm due to a change in ΔrH0 from nonstandard physics is further reduced. The above two points, ΔrH0 being small and z* being high, argue that our analysis is robust against the change in ΔrH0 due to possibly unknown nonstandard physics.

Indeed, even if we artificially multiply Equation (1) by a factor of 110% mimicking some possible unknown modifications to our standard assumptions, the joint constraint on Ωm only changes to 0.303 ± 0.008, compared to 0.302 ± 0.008 in our standard analysis, or

Equation (6)

The small coefficient above verifies the conclusion that the constraint on Ωm is insensitive to changes to ΔrH0.

Third, even the small value of ΔrH0 itself is quite robust and it is not easy for nonstandard models to change it substantially. While the form of Equation (1) is generally true, one may change ΔrH0 by changing the sound speed, cs, the function E(z), and the redshift span between zd and z*. For example, changing the Ωb h2 can change the sound speed. Regarding this, due to the consistency of the Ωb h2 constraint between Planck and BBN (Cooke et al. 2018; Fields et al. 2020; Planck Collaboration et al. 2020), and between different uses of information from CMB anisotropy (Motloch 2020), it is difficult for the reduced baryon fraction to be significantly different from Ωb h2 ∼ 0.0222. Since the Ωb h2 term only has a relatively small effect on cs between zd and z*, it is even more difficult to change ΔrH0 via the Ωb h2 term. Quantitatively, even if we artificially change Ωb h2 by 10% (∼ 4σ away from the mean value of our adopted prior), which we use to mimic effects from possible changes in the sound speed, ΔrH0 only changes by ≲2%. For the E(z) function, if nonstandard physics occurred before recombination as in a number of currently proposed early universe nonstandard models, its functional form will not change. 10 Due to the narrow redshift span between zd and z*, it is difficult to substantially change ΔrH0 from its standard value by changing E(z) with modified energy compositions. Also, although a change to the redshift span itself can directly affect ΔrH0, to get a ∼ 10% change to ΔrH0, we would need a ∼10% change of Δzs, which is ∼ 5.5σ away from the mean value of Δzs inferred from Planck. It is difficult to achieve this because the decoupling between photons and baryons are based on well known physics. The redshift span between recombination and the drag epoch cannot be substantially different from the standard value unless there are some other significant interactions between, e.g., baryon and dark matter, which would have led to observable features in CMB anisotropy (Dvorkin et al. 2014; Boddy & Gluscevic 2018; Xu et al. 2018; de Putter et al. 2019). In support, Jedamzik et al. (2021) finds that r* ∼ 1.018rd persists in a number of pre-recombination nonstandard models. In summary, it is not easy to make a noticeable change (like 10%) to ΔrH0.

Besides the change of ΔrH0, one may worry about the prior on z* that we adopted. But as discussed in Appendix A.3, our result is not sensitive to the prior on z*, because z* is sufficiently large and fM is approaching a constant (though dependent of Ωm). To numerically verify this, we artificially shift the mean value of Planck result of z* from 1090–1110, which can be achieved, e.g., in the presence of primordial magnetic fields (Jedamzik & Pogosian 2020). This only leads to an ∼ 0.2% change in the mean value of Ωm (from 0.3021–0.3015).

Another possible concern is that we have adopted the constraint on θcmb based on the standard ΛCDM model. The uncertainty of θcmb from Planck is about 0.03%. Such a small uncertainty of θcmb may not represent the possible uncertainty due to some nonstandard early universe physics. For example, releasing the effective relativistic particle number, the Planck result becomes 100θcmb = 1.04136 ± 0.00060. And releasing the mass of neutrinos gives 100θcmb = 1.04105 ± 0.00032. So, the uncertainty of θcmb due to these potential nonstandard physics at early times is comparable to or larger than 0.03%. However, these uncertainties in θcmb will not substantially change our result in Ωm for the following reason. Take the combination of θcmb and the SDSS DR12 GC BAO, for example. As shown in Figure 2, the uncertainty of Ωm is mainly due to the width of the projection of the overlap between these two constraints in the Ωm direction. As we can visualize in that figure, even if we increase the width of the black contour or shift it a bit horizontally, the projection in the Ωm direction would not be substantially changed. Quantitatively, this insensitivity can be shown by alternatively using the result of θcmb from ACT+WMAP. We can see from Equations (3) and (4) that there is an ∼2σ (in terms of Planck's uncertainty) difference in the mean value of θcmb between Planck and ACT+WMAP, and also the uncertainty of θcmb for ACT+WMAP is more than twice that for Planck. But Figure 3 shows the resultant constraints on Ωm are nearly identical.

To evade the constraint on the evolution of the post-recombination CMB by UCS, one would need (1) some mechanism that makes the expressions for the two sound horizon scales different prior to the recombination; or (2) some nonstandard physics occurring during the narrow gap between z* and zd that significantly changes the function E(z) or the sound speed cs; or (3) a substantial change of the redshift span between recombination and the drag epoch. None of those possibilities are easy to achieve, but nonetheless could serve as guidelines if we would like to build models of counter examples.

Our UCS constraint on Ωm is strong and insensitive to early universe physics. This can be used as a strong prior or be combined with other observations to break degeneracy in the background. We will do that in Section 4 to obtain several early universe physics insensitive constraints on H0.

3.3. Constraint on the H0–Ωm Plane

In our standard analysis the constraint in the H0–Ωm plane is actually not exactly along the Ωm direction. That is because the normalized comoving distance fM has a weak dependence on the Hubble constant via the radiation and the massive neutrino terms; see Appendix A.1 for a discussion. A principal component analysis shows that

Equation (7)

Therefore, the constraint exhibits a small positive correlation in the H0–Ωm plane. In situations where joint analyses using MCMC with other data sets are unavailable (such as that with γ-ray optical depth, which shall be discussed), we will use Equation (7) as a prior to set weights in parameter chains obtained from other analyses in order to approximate the joint result. In practice, we find that using either Equations (5) or (7) as a prior only makes a negligible difference.

4. Joint Constraints with Post-recombination and Nonlocal Observations

Several post-recombination observations are independent of or insensitive to early universe physics; these are summarized in Lin et al. (2020). When combined with UCS, they can break the degeneracy in the evolution of the cosmic background and provide early universe physics insensitive constraints on H0. Specifically, we will consider the cosmic chronometers, γ-ray optical depth, cosmic age, and large-scale structure (LSS), respectively. Note that the joint constraints are not local measurements; they depend on a post-recombination cosmological model. The constraints on H0 using UCS+nonlocal observations trade the dependence on early universe physics with that on other astrophysical effects. Studies of how to mitigate and/or better account for astrophysical uncertainties in each observation are warranted. In this work, we include estimates of systematic uncertainties from the latest works; see the discussions in each entry below. Potentially unaccounted for systematic errors would cause biases in their H0 inferences, but the agreement between results from the different observations we consider suggest the consistency with current published estimates of systematic errors. LSS to some extent brings back the dependence on early universe physics but still provides a sound horizon insensitive constraint on H0; see a later discussion.

Cosmic chronometer s: The cosmic chronometer is a technique that directly measures the Hubble parameter H(z) by measuring the differential ages of passively evolving galaxies at two nearby redshifts (Moresco et al. 2016, 2020; Moresco & Marulli 2017). This technique does not depend on early universe physics, but may suffer from various astrophysical systematic errors (Moresco & Marulli 2017; Moresco et al. 2020). Therefore, we provide various results taking into account additional systematic errors considered in Moresco et al. (2020). The results include analyses of the data from (1) the current public release (current); (2) with uncorrelated systematic errors in the odd-one-out scenario (extra systematic); (3) with systematic errors in the odd-one-out scenario but conservatively include the full correlation of those systematic errors at different redshifts using Equation (9) in Moresco et al. (2020) (extra systematic, conservative). Note that the last case is a conservative estimation of extra systematic errors because systematic errors for data at different redshift bins are actually not quite correlated (Moresco et al. 2020).

The γ-ray optical depth: High energy γ-ray photons interact with the diffuse extragalactic background lights and pair-create electrons and positrons (Gould & Schréder 1966). The rate of γ-ray photons being attenuated depends on the proper density of the extragalactic background light as well as the proper distance along the line of sight, and hence depends on the evolution of the cosmic background. Measuring the γ-ray optical depth has thus provided an independent constraint on the evolution of the background (Ackermann et al. 2012; H.E.S.S. Collaboration et al. 2013; Biteau & Williams 2015; Domínguez et al. 2019). The extragalactic background light is mainly produced by star formation; Domínguez et al. (2019) studied the systematic errors due to different models in the calculations of the density of extragalactic light. We obtain the MCMC of the analysis carried in Domínguez et al. (2019) and estimate the joint constraint of UCS+γ-ray by adding weights to each sampled (H0, Ωm) of the chain according to Equation (7).

Cosmic age: Since most of the cosmic time is after recombination (in fact, the cosmic time after z = 100 contributes about 99.9% of the entire cosmic age), measuring ages of individual old stars, globular clusters, and galaxies is a promising way to test early universe resolutions to the Hubble tension (Lin et al. 2020; Bernal et al. 2021; Boylan-Kolchin & Weisz 2021). With some prior on the time or redshift when the old stars formed, the age of the universe tU can be inferred (Jimenez et al. 2019). The age determination of stars and globular clusters is systematic-error dominated (O'Malley et al. 2017; Wagner-Kaiser et al. 2017; Jimenez et al. 2019). Recently, Valcin et al. (2021) updated the age estimate of the oldest galactic globular clusters by reducing the systematic uncertainty due to the depth of the convection envelope, which is the most important known systematic error. They infer tU = 13.5 ± 0.27 Gyr based on a spectroscopic metallicity determination and tU = 13.5 ± 0.33 Gyr relying only on the clusters' color–magnitude diagrams. 11 We adopt these cosmic age estimations and combine them with UCS to provide early universe insensitive constraints on H0.

The LSS+BBN: Earlier, it was pointed out that the combination of several current LSS observations can provide an independent constraint on H0 (Lin & Ishak 2017). But some observations used, such as the redshift space distortion (with BAO included), mix the dependence on the sound horizon scale and thus are not independent of early universe physics. Later, Baxter & Sherwin (2020), Philcox et al. (2021), and Pogosian et al. (2020) showed that subsets of LSS observations with uncalibrated SN Ia and/or uncalibrated BAO can provide a sound horizon-independent constraint on H0. The idea is that the shape of the matter power spectrum is sensitive to the horizon scale at matter-radiation equality, which depends on the matter-to-radiation energy ratio today parameterized by Ωm h2 (Philcox et al. 2021). Since only the distance normalized by 1/H0 can be measured, LSS can put a constraint on Ωm h, and with our UCS constraint (mainly) on Ωm, can break the degeneracy in the H0–Ωm plane to provide a constraint on H0. In this work, we use the CMB lensing from Planck and the 3 × 2 correlation functions from the Dark Energy Survey (DES; also used in Pogosian et al. 2020), as the two together can probe the matter power spectrum on a wide range of scales, including resolving the peak of the matter power spectrum (Baxter & Sherwin 2020; Pogosian et al. 2020). The matter power spectrum has a weak dependence on Ωb h2 and we adopt a prior of Ωb h2 = 0.0222 ± 0.005 obtained from BBN (Cooke et al. 2018). Note that, unlike the other nonlocal observations, combining LSS to some extent raises the sensitivity of UCS to early universe physics. However, LSS depends on early universe physics in a very different way from that of the sound horizon scale, which is the most important quantity for the CMB inference of H0. As mentioned earlier, Baxter & Sherwin (2020), Philcox et al. (2021), and Pogosian et al. (2020) have shown the insensitivity of LSS (as well as the BBN constraint on Ωb h2) to sound horizon physics. Relating to this, the introduction of early universe physics that decreases the sound horizon tends to worsen the σ8 tension (Jedamzik et al. 2021).

We summarize the results of the joint analyses with each of the above-mentioned observations and UCS in Table 1. All constraints prefer a lower value of H0 that is more consistent with Planck and the tip of the red giant branch (TRGB)-based local measurement 12 from Freedman et al. (2020) than with the Cepheid-based local measurement. In particular, the fact that our result from UCS+CMBlens+DES+BBN prefers a lower value of H0 is consistent with the finding that combining Planck and LSS disfavors early dark energy (Hill et al. 2020; Ivanov et al. 2020).

Table 1. Summary of the Constraints on H0

Methods H0 (km s−1 Mpc−1) n-σ from R21
UCS and individual nonlocal observationWithout θcmb With θcmb Without θcmb With θcmb
Cosmic chronometers    
Current public data69.1 ± 1.7 68.8 ± 1.6 1.9σ 2.1σ
Extra systematic69.4 ± 2.3 69.2 ± 2.1 1.4σ 1.6σ
Extra systematic, conservative69.3 ± 3.4 68.9 ± 3.3 1.1σ 1.2σ
γ-ray optical depth66.2 ± 3.566.1 ± 3.41.9σ 2.0σ
Cosmic age    
tU = 13.5 ± 0.27 Gyr70.2 ± 1.7 69.8 ± 1.5 1.4σ 1.7σ
tU = 13.5 ± 0.33 Gyr70.3 ± 2.1 69.8 ± 1.9 1.2σ 1.5σ
CMBlens+DES+BBN68.8 ± 2.4 68.6 ± 2.0 1.6σ 1.9σ
UCS and joint nonlocal observations a     
All nonlocal observations69.1 ± 1.5 68.8 ± 1.3 2.0σ 2.4σ
Nonlocal observations without cosmic age68.3 ± 1.9 68.1 ± 1.6 2.1σ 2.5σ
Nonlocal observations without LSS69.1 ± 1.6 68.8 ± 1.5 2.0σ 2.2σ
Time-delay strong lensing b     
TDCOSMO (Millon et al. 2020)74.2 ± 1.6  
TDCOSMO+SLACS (Birrer et al. 2020) ${67.4}_{-3.2}^{+4.1}$   
Local measurements c (distance ladder)    
Cepheid+SN Ia (Riess et al. 2021)73.2 ± 1.3  
TRGB+SN Ia (a) (Freedman et al. 2020)69.8 ± 1.9  
TRGB+SN Ia (b) (Yuan et al. 2019)72.4 ± 2.0  
TRGB+SN Ia (c) (Soltis et al. 2021)72.1 ± 2.0  
Mira+SN Ia (Huang et al. 2020)73.3 ± 3.9  
Cepheid+SBF+SN Ia (Khetan et al. 2021)70.5 ± 4.1  
Cepheid/TRGB+SBF (Blakeslee et al. 2021)73.3 ± 2.5  
Cepheid/TRGB+TFR (Kourkchi et al. 2020; Schombert et al. 2020)76.0 ± 2.5  
Cepheid/TRGB+SN II (de Jaeger et al. 2020) ${75.8}_{-4.9}^{+5.2}$   
Local measurements (non-distance ladder)    
Megamaser Cosmology Project (Pesce et al. 2020)73.9 ± 3.0  
Standard siren multi-messenger (The LIGO Scientific Collaboration et al. 2021) ${69}_{-8}^{+16}$   

Notes. The first subgroup contains the improved results using the method of this paper, namely, combining UCS with each of the listed early universe physics independent or insensitive observations. (See the text for detailed descriptions for each case.) They are all more consistent with Planck (the full standard analysis) or the TRGB-based local measurement from Freedman et al. (2020) than with the Cepheid-based local measurement. For nonlocal determinations, in addition to the results with the full UCS indicated by the bold values, we also include the results without using θcmb in UCS (i.e., only using UnCalibrated (U.C.) BAO and Pantheon as UCS). Comparing the results with and without θcmb, we can see the various extents to which nonlocal results improve their uncertainties and aggravate their tensions with the the Cepheid+SN Ia determination (R21). Strictly speaking, CMBlens+DES+BBN is not independent of early universe physics, but depends on early universe physics in a very different way from the methods that rely on the modeling of the sound horizon; see the text for a discussion. Since all nonlocal results are consistent and independent of each other, for completeness we also provide the constraint on H0 from jointly analyzing UCS+all late-time but nonlocal observations using a likelihood analysis. The UCS+all nonlocal result is in 2.4σ tension with R21. For comparison, in other entries we cite the constraints on H0 from other measurements that are independent of early universe physics. (Most updated results from independent groups are cited.)

a Conservative settings in cosmic chronometers and cosmic age are used. b The time-delay strong-lensing technique is not a method of local measurement, but is insensitive to the underlying cosmological model. When the standard-ΛCDM is assumed, its determination of H0 is insensitive to Ωm. Compared to the TDCOSMO entry, the result in TDCOSMO+SLACS relaxes the strong assumption of the lens mass density profile, considers correlations of the modeling between lenses, and adds imaging and spectroscopic data of other lenses. c Local measurements based on the distance ladder technique are represented as primary+secondary distance indicators. SBF stands for surface brightness fluctuation, Mira for Mira variables, TFR for Tully–Fisher relation, and SN II for Type II Supernova. Note that not all local measurements are independent, since some of them use the same type of primary or secondary distance indicator. It is also worth pointing out that the zero-point of the TRGB used in Cepheid/TRGB+TFR (Kourkchi et al. 2020; Schombert et al. 2020), Cepheid/TRGB+SN II (de Jaeger et al. 2020) and Cepheid/TRGB+SBF (Blakeslee et al. 2021) are somewhat different from that in TRGB+SN Ia (a) (Freedman et al. 2020).

Download table as:  ASCIITypeset image

Since all nonlocal results are consistent with and independent of each other, for completeness we jointly analyze UCS+all late-time but nonlocal observations. In doing so, we adopt conservative error estimates on cosmic chronometers and cosmic age. The joint constraint on H0 is in a 2.4σ tension with the Cepheid+SN Ia determination. In addition, we provide one joint result excluding the cosmic age due to its prior on the formation time of the globular cluster and some other potentially unaccounted for systematic errors, and one excluding LSS so that the remaining constraints are highly insensitive to early universe physics. These two weaker joint results are also in tension with the Cepheid+SN Ia determination with >2σ.

For comparison, we also list in Table 1 other measurements of H0, including the results from the time-delay strong-lensing technique (TDCOSMO; Birrer et al. 2020) and local measurements. Although with larger uncertainties than that of the Planck 2018 inference of H0, the post-recombination but nonlocal constraints on H0 appear to have some tension with (and are all consistently lower than) most of the local measurements, except for the TRGB+SN Ia result reported in Freedman et al. (2020). Incidentally, it is worth noting that not all local measurements are totally independent from each other because some of them share the same type of primary or secondary distance indicator; see Table 1.

The tension we find here may be due to either post-recombination new physics, unknown systematic errors in some local measurements, or unknown systematic errors in other nonlocal observations, but cannot be resolved by introducing early universe physics. Regarding this, the fact that nonlocal results are independent in terms of statistics and methodology but still consistent with each other disfavors the explanation of systematic errors in nonlocal observations. Our results challenge early universe physics resolutions to the Hubble tension, which is in agreement with Jedamzik et al. (2021) who came to a similar conclusion in a very different way.

5. Discussion

In our analyses, we have assumed a flat-ΛCDM universe after recombination since we are interested in constraining the cosmic background without a strong assumption of the standard cosmological model at early times. But our method can also be generalized to allow a nonstandard cosmological model after recombination. This can serve as a strong and early universe physics insensitive test to late-time models, such as voids (Shanks et al. 2019a, 2019b; Ding et al. 2020) and modified gravity/dark energy models (Belgacem et al. 2018; Pan et al. 2019; Park & Ratra 2019; Gonzalez et al. 2020; Shimon 2020; Yang et al. 2021). The advantages of this test are (1) the constraint on late-time models will be free from biases due to any possible early universe nonstandard physics or systematic errors in the CMB data that affect the amplitude or the shape of the power spectrum; (2) while releasing the assumption of the standard cosmological model before recombination, the constraining power on the late-time background has been shown to be largely maintained.

To use UCS as an early universe physics insensitive test on late-time cosmological models, all that is needed is to modify the normalized comoving distance accordingly (i.e., Equation (A2)). Note that one is able to constrain late-time cosmological models without constraining the value of H0 because late-time models predict the relative change of H(z) as a function of redshift instead of predicting its absolute value. UCS is a clear and strong probe of the relative change of H(z). Therefore, when constraining late-time nonstandard cosmological models, it is neither necessary nor recommended to adopt a prior of H0 obtained from the Cepheid-based local measurement. It is also unnecessary to combine with the observations listed in Section 4, as the test uses a minimal set of observations and still remains strong, although doing so allows us to gain some additional constraining power.

We also note that BOSS Collaboration et al. (2015) has used the idea of treating the measurement of the CMB angular scale as an extra data point for BAO. 13 In their method, they assume the Standard Model expression of rd as a function of Ωm h2 and Ωb h2 but drastically increase the prior ranges for these two parameters. Effectively, rd becomes a free parameter because Ωb h2 is unconstrained. In our work, we directly leave rd as a free parameter and only model its difference from r*. For the purpose of our paper, namely, to disentangle the interference of the early universe physics, the latter approach is more advantageous because, e.g., it removes the dependence on BBN, which (when applied) can determine the parameter Ωb h2 and thus constrain rd (and h) by the Standard Model in the former approach.

6. Summary and Conclusion

Constraining the Hubble constant in an early universe physics independent or insensitive way is important because this can check whether the current tension between Planck and the Cepheid-based local measurement is due to some undiscovered early universe physics. In this work, we have introduced a method that straightforwardly analyzes θcmb in an early universe physics insensitive way. We jointly analyzed the angular size of CMB acoustic peaks, BAO, and SNe Ia by treating their absolute scales or luminosity as free parameters.

These uncalibrated standard rulers and candles, measured across cosmic time from recombination to today, provide a strong constraint on the evolution of the cosmic background insensitive to early universe physics. When a flat ΛCDM is assumed after recombination, UCS mainly constrains the matter density parameter today, but the exact constraint exhibits a small positive degeneracy in the H0–Ωm plane.

The method of UCS can also be generalized and used to test post-recombination nonstandard cosmological models in a way that is insensitive to modifications of early universe physics or systematic errors that affect the amplitude of the CMB power spectra at different scales, yet can still be powerful.

To analyze the problem of the Hubble tension in a way that is insensitive to pre-recombination physics, we combined UCS with other early universe physics independent or insensitive, but nonlocal, observations to break the degeneracy in the H0–Ωm plane and obtained several constraints on H0. The combination with LSS mildly raises the dependence to early universe physics, but still insensitive to the sound-horizon physics as discussed in Baxter & Sherwin (2020), Philcox et al. (2021), Pogosian et al. (2020), and Jedamzik et al. (2021), and its result on H0 is consistent with other nonlocal determinations. All nonlocal constraints are more consistent with the Planck and a TRGB-based local measurement than with the Cepheid-based local measurements (see Table 1).

In the literature, the Hubble tension is usually attributed to some discrepancy between the early (pre-recombination) and late (post-recombination) universe. Our analyses provide some new insights into this statement. Hubble constant measurements of the late universe include both nonlocal and local measurements. Our results suggest that the nonlocal ones are still consistent with the CMB result, and the tension has more to do with a tension between the nonlocal (including both pre- and post-recombination) measurements and most of the local measurements.

This discrepancy may be due to new physics in the post-recombination epoch, or may disappear after future improvement of systematic errors in some local measurements or all nonlocal observations. Indeed, since the H0 determinations using UCS+nonlocal observations significantly reduce the dependence on early universe physics, it suggests an increased focus on astrophysical effects. While we have adopted the latest treatments in each nonlocal observation, more studies are warranted to reduce and better treat the systematic uncertainties in each observation. Fortunately, the fact that all nonlocal probes are independent in methodologies and that their constraints on H0 (jointly with UCS) are consistent with each other suggests an absence of major unaccounted for systematic errors in those nonlocal observations. In any case, this tension between the local determination and UCS+nonlocal results will not be resolved by introducing nonstandard physics in the pre-recombination early universe.

Identifying the cause of the reported inconsistencies—being either beyond the Standard Model physics or unaccounted for/underestimated systematic errors—has become and will continue to be an important topic in cosmology; see, e.g., Lin et al. (2020), Efstathiou (2014), Yao et al. (2020), Garcia-Quintero et al. (2019), Yu et al. (2018), Lin & Ishak (2017), Huang & Wang (2016), Gómez-Valent & Solà Peracaula (2018), Vagnozzi (2020), and Yang et al. (2019). Reanalyzing data by relaxing strong assumptions is a useful way to check those assumptions. This philosophy has been used to check the validity of other astrophysical assumptions as well, such as in the strong-lensing time-delay technique of determining the Hubble constant (Birrer et al. 2020; Blum et al. 2020; Gomer & Williams 2020). In the future, galaxy surveys such as the Dark Energy Spectroscopic Instrument (DESI) will provide measurement of BAO with unprecedented precision (DESI Collaboration et al. 2016). The Vera C. Rubin Observatory will discover many more SNe Ia (LSST Dark Energy Science Collaboration 2012). Measuring BAO from 21 cm anisotropy will fill in standard rulers during cosmic dawn (Muñoz 2019). In the meantime, the techniques using cosmic chronometers, γ-ray optical depth, and large-scale structure will be improved, which will help provide more precise and robust measurement of H0 without strongly assuming early universe physics. All these data will further reduce the errors and make the method of UCS more powerful. With these, we shall soon be able to narrow down the cause of the Hubble tension.

We thank Gongjun Choi for helpful discussions, and Eoin Colgain, Michele Moresco, Levon Pogosian, Matias Zaldarriaga, and an anonymous referee for helpful feedback and comments.

Appendix: Data and Likelihoods

A.1. Notations and Background Evolution

Here, we discuss in detail the framework and the UCS data we used in this work. We normalize the (transverse) comoving distance dM by the Hubble distance (1/H0),

Equation (A1)

with fM(z; Ωm, ⋯) dependent on a cosmological model. We assume the standard spatially flat ΛCDM model after recombination and

Equation (A2)

Equation (A3)

with the constraint $1={{\rm{\Omega }}}_{{\rm{m}}}+{{\rm{\Omega }}}_{{\rm{\Lambda }}}+{{\rm{\Omega }}}_{{\rm{r}}}+{\sum }_{i}{{\rm{\Omega }}}_{{\nu }_{i}}$. The subscript "m" stands for matter, "Λ" for the cosmological constant, "r" for radiation (photon and massless neutrino), and "νi " for massive neutrinos where i runs through the neutrino species. The function ei (z) represents the redshift dependence of the evolution of the density fraction of the massive neutrinos and can be well approximated by (W. Lin et al. 2021, in preparation)

Equation (A4)

where $a=\tfrac{1}{1+z}$ is the scale factor, n = 1.8367, ${a}_{{\rm{T}},i}=3.1515\tfrac{{T}_{\nu }^{0}}{{m}_{{\nu }_{i}}}$, ${T}_{\nu }^{0}=1.676\times {10}^{-4}$ eV, and ${m}_{{\nu }_{i}}$ is the ith neutrino mass. The radiation and neutrino terms only have some small contribution to fM at redshifts close to recombination, but we include them for completeness. In our standard analysis, we consider one massless and two massive neutrinos in the normal hierarchy with constraints on the difference of the mass-squared ${\rm{\Delta }}{m}_{12}^{2}=7.9\,\times {10}^{-5}\,{\mathrm{eV}}^{2}$ and ${\rm{\Delta }}{m}_{13}^{2}=2.2\times {10}^{-3}\,{\mathrm{eV}}^{2}$ (Maltoni et al. 2004). With the measured CMB temperature today and the assumption of thermal neutrino relics, we have ${{\rm{\Omega }}}_{{\rm{r}}}={{\rm{\Omega }}}_{\gamma }+{{\rm{\Omega }}}_{\nu }^{\mathrm{massless}}\,=3.0337{h}^{-2}\times {10}^{-5}$ and ${{\rm{\Omega }}}_{{\nu }_{i}}=\tfrac{{m}_{{\nu }_{i}}}{93.14\,{h}^{2}\,\mathrm{eV}}$, where $h\,\equiv \tfrac{{H}_{0}}{100\,\mathrm{km}\,{{\rm{s}}}^{-1}\,{\mathrm{Mpc}}^{-1}}$ (Lesgourgues & Pastor 2012). While we have ignored the uncertainties of the mass-squared difference and the mass of the lightest neutrino, we found that our results are insensitive to them for a reasonable range of the lowest neutrino mass for both hierarchy schemes. This is because the radiation and neutrino terms only have small effects on fM at high redshifts close to recombination. For a reasonable range of the mass of the lightest neutrino, all neutrinos were relativistic before z ∼ 100 when they have (small) effects on fM. In fact the effects of the radiation and the neutrinos terms are so small that our constraints on Ωm are essentially unchanged even if we completely ignore these two terms.

In our standard treatment, fM mainly depends on Ωm. From our adopted normalization (i.e., Equation (A1)) we shall see more clearly the degeneracy of H0 with the scale of the sound horizon (r* or rd) and with the absolution magnitude of SNe Ia, M0. However, it is worth pointing out that fM still weakly depends on h via the radiation and neutrino terms, especially at higher redshifts. 14 As a consequence, our constraints will not be totally uninformative in H0; see Section 3.3.

Note that in this work, we have assumed the standard ΛCDM for cosmic time after recombination, but our analyses can be generalized and used to test post-recombination nonstandard models where Equation (A2) is replaced by the corresponding normalized comoving distance. Such a test of post-recombination nonstandard models will have the advantage of being robust against any potentially unknown early universe physics while retaining a strong constraining power.

A.2. The Late-time BAO

The late-time BAO measurements are compressed information obtained from two-point correlation functions of tracers of the underlying matter fluctuation. With rd the sound horizon scale at the end of the drag epoch (at redshift zd), measurements can be given by a pair of quantities that are the angular size ${\theta }_{{\rm{d}}}=\tfrac{{r}_{{\rm{d}}}{H}_{0}}{{f}_{{\rm{M}}}(z)}$ and the redshift span ${\rm{\Delta }}{z}_{{r}_{{\rm{d}}}}={r}_{{\rm{d}}}{H}_{0}E(z)$ of the baryon acoustic sound horizon when placed perpendicular to and along the light of sight. Alternatively, the set $\left[\tfrac{{r}_{{\rm{d}}}{H}_{0}}{{f}_{{\rm{V}}}(z)},\,{F}_{\mathrm{AP}}(z)=\tfrac{{\rm{\Delta }}{z}_{{r}_{{\rm{d}}}}}{{\theta }_{{\rm{d}}}}=E(z){f}_{{\rm{M}}}(z)\right]$ will be given or can be constructed from $({\theta }_{{\rm{d}}},\,{\rm{\Delta }}{z}_{{r}_{{\rm{d}}}})$. Here,

Equation (A5)

which can be interpreted as the cubic root of the differential comoving volume (w.r.t. the solid angle and logarithmic redshift) normalized by ${(1/{H}_{0})}^{3}$. The quantity $\tfrac{{r}_{{\rm{d}}}{H}_{0}}{{f}_{{\rm{V}}}(z)}$ can be interpreted in the following way. Imagine a simple cubic lattice formed in the universe with the comoving length of each dimension equal to the size of the drag sound horizon. The comoving volume element of each lattice cell is ${r}_{{\rm{d}}}^{\,3}$. Then, ${\left(\tfrac{{r}_{{\rm{d}}}{H}_{0}}{{f}_{{\rm{V}}}(z)}\right)}^{3}$ is (the inverse of) the number of those volume elements seen per solid angle per logarithmic redshift. The quantity FAP(z) is the ratio between ${\rm{\Delta }}{z}_{{r}_{{\rm{d}}}}$ and θd, which is useful as it contains no dependence of ΔrH0.

It is worth pointing out that rd always goes with H0 in those BAO measurements, while fM and fV are nearly independent of H0. So, when rd is treated as a free parameter, it is degenerate with H0; also see Aylor et al. (2019). We therefore treat rd H0 as one free parameter when analyzing the BAO data.

In Table 2, we summarize some BAO measurements obtained or inferred from (in ascending order with effective redshift) the 6dF galaxy survey (Beutler et al. 2011), SDSS DR7 MGS (Ross et al. 2015), SDSS DR12 GC (Alam et al. 2017), SDSS DR16 luminous red galaxy, and emission line galaxy samples (Luminous Red Galaxy (LRG) and Emission Line Galaxy)) (Wang et al. 2020), and eBOSS DR16 quasars (QSO) and Lyα forest (Lyα) auto and cross correlations (Blomqvist et al. 2019). There are correlations between θd and ${\rm{\Delta }}{z}_{{r}_{{\rm{d}}}}$, which are shown in the last column of Table 2. In particular, there are correlations among the three SDSS DR12 GC BAO measurements with

Equation (A6)

Table 2. BAO Measurements Obtained or Inferred from Wang et al. (2020), Blomqvist et al. (2019), Alam et al. (2017), Ross et al. (2015), and Beutler et al. (2011)

LabelsReferences zeff $\tfrac{{r}_{{\rm{d}}}{H}_{0}}{{f}_{{\rm{V}}}(z)}$ FAP ${\theta }_{{\rm{d}}}\equiv \tfrac{{r}_{{\rm{d}}}{H}_{0}}{{f}_{{\rm{M}}}(z)}$ Δzrd H0 × E(z) $\rho ({\theta }_{{\rm{d}}},{\rm{\Delta }}{z}_{{r}_{{\rm{d}}}})$
6dfBeutler et al. (2011)0.1060.336 ± 0.015N/AN/AN/AN/A
MGSRoss et al. (2015)0.150.224 ± 0.0084N/AN/AN/AN/A
SDSS DR12 GCAlam et al. (2017)0.380.1002 ± 0.00110.410 ± 0.0160.0977 ± 0.00160.0400 ± 0.0012Equation (A6)
  0.510.07893 ± 0.000800.599 ± 0.0210.0748 ± 0.00110.0448 ± 0.0011Equation (A6)
  0.680.06898 ± 0.000710.761 ± 0.0270.0641 ± 0.00100.0488 ± 0.0012Equation (A6)
LRG and LEGWang et al. (2020)0.770.05707 ± 0.000900.960 ± 0.0350.0530 ± 0.00110.0509 ± 0.00158.02 × 10−4
QuasarBlomqvist et al. (2019)1.520.0383 ± 0.0017N/AN/AN/AN/A
Lyα-quasarBlomqvist et al. (2019)2.350.03256 ± 0.000654.13 ± 0.200.02698 ± 0.00090.1115 ± 0.00280.645

Download table as:  ASCIITypeset image

The uncalibrated BAO likelihood reads as

Equation (A7)

where Θi m, rd H0) and Q i are predictions and measurements of $({\theta }_{{\rm{d}}},\,{\rm{\Delta }}{z}_{{r}_{{\rm{d}}}})$ or $\left(\tfrac{{r}_{{\rm{d}}}{H}_{0}}{{f}_{{\rm{V}}}(z)},\,{F}_{\mathrm{AP}}(z)\right)$, respectively, and C i is the covariance matrix and the index i denotes different BAO measurements.

A.3. The Angular Size of the Acoustic Scale from the CMB

The photon-baryon plasma perturbation also manifests as acoustic oscillations in the CMB temperature and polarization angular power spectra. These oscillations correspond to a sharply defined angular scale (θcmb), which has been robustly and precisely measured and is almost independent of the underlying cosmological model (Planck Collaboration et al. 2020). It reads as ${\theta }_{\mathrm{cmb}}={r}_{* }/{d}_{{\rm{M}}}^{\mathrm{rec}}$, where r* is the comoving sound horizon at recombination and ${d}_{{\rm{M}}}^{\mathrm{rec}}$ is the comoving distance at recombination. Plugging in Equation (A1) we have ${\theta }_{\mathrm{cmb}}=\tfrac{{r}_{* }{H}_{0}}{{f}_{M}({z}_{* })}$. So, θcmb mainly depends on r* H0 and Ωm. It weakly depends on h via the radiation and neutrino terms in E(z) within fM(z); see Appendix A.1. It also depends on z*. We first assume a prior on z* obtained from the Planck baseline constraint based on the standard ΛCDM model (Planck Collaboration et al. 2020) or from ACT+WMAP (Aiola et al. 2020). But since ${f}_{{\rm{M}}}(z\to \infty )\to \mathrm{constant}$, corresponding to the particle horizon normalized by the Hubble distance and z* is sufficiently large, the dependence of θcmb on z* is very weak. Therefore, our results would be barely changed for a reasonable range of z* and we verify this in Section 3.2.

The θcmb likelihood reads as

Equation (A8)

where θ = (θcmb, z*) and the subscripts "o" and "p" denote "observation" and "prediction," respectively. Here, z* only serves as a nuisance parameter, and, since the derived constraint on θcmb correlates with z* for both Planck and ACT+WMAP, we have included their correlation in the θcmb likelihood Equation (A8). Thus, the parameters contained in the θcmb likelihood are r* H0, Ωm, and h (a very weak dependence).

A.4. Linking the Two Baryon Acoustic Sound Horizons

The two sound horizon scales, r* at recombination and rd at the end of the drag epoch, are closely related to each other. They are given by

Equation (A9)

Equation (A10)

where cs is the sound speed. A number of early universe nonstandard models/considerations have been proposed to make both r* and rd smaller to mitigate the Hubble tension between Planck and the Cepheid-based local measurement, e.g., Poulin et al. (2019), Kreisch et al. (2020), and Jedamzik & Pogosian (2020). But those models (in fact, almost all early universe nonstandard models) shift the two horizons by approximately the same constant, so their difference remains intact and small. (See discussions in Section 3.2 for the details and possible exceptions.) This permits us to jointly analyze the uncalibrated BAO and the θcmb likelihoods by treating r* H0 as a free parameter and adopting some reasonable and model insensitive assumptions about the calculation of the difference between r* H0 and rd H0.

To do this, we take the difference between Equations (A10) and (A9),

Equation (A11)

In our standard analysis, we first assume that ΔrH0 can be modeled in the same way as in the standard ΛCDM model, so that we have the sound speed given by

Equation (A12)

and E(z) given by Equation (A3). With Ωγ h2 = 2.47 × 10−5, cs(z) is determined if Ωb h2 is given. In the standard analysis, we assume a BBN prior on Ωb h2 = 0.0222 ± 0.0005. In the θcmb likelihood, we already assume a prior on z* obtained from Planck or ACT+WMAP, so the only thing needed to model ΔrH0 is the redshift difference between z* and zd. In the standard analysis, we also assume a prior on Δzsz*zd obtained from Planck (Δzs = 30.26 ± 0.57) or ACT+WMAP (Δzs = 29.95 ± 0.75). Then, the joint θcmb and uncalibrated BAO likelihood, i.e., $\mathrm{ln}{{ \mathcal L }}_{{\theta }_{\mathrm{cmb}}}+\mathrm{ln}{{ \mathcal L }}_{{\rm{u}}.{\rm{c}}.\ \mathrm{bao}}$, contain parameters r* H0, Ωm, and h (a very weak dependence), with some assumed priors on z*, Δzs, and Ωb h2.

A.5. Type Ia SNe and a Fast Algorithm to Compute the SN Ia Likelihood

In a Friedmann–Lemaître–Robertson–Walker universe, the luminosity distance (observed in the heliocentric frame) is related to the comoving distance by dL = (1 + zhel)dM(zcmb) (see, e.g., Equation (2) in DES Collaboration et al. 2019) and we have

Equation (A13)

where zhel and zcmb are the redshifts of SNe Ia observed in the heliocentric frame and in the CMB frame, m is the apparent magnitude of the SN Ia, and ${ \mathcal M }$ is defined as

Equation (A14)

with M0 the absolute magnitude of SNe Ia at some reference stretch and color. We treat ${ \mathcal M }$ as a free parameter in the SN Ia likelihood analysis. The SN Ia likelihood reads as

Equation (A15)

where m o and ${{\boldsymbol{m}}}_{{\rm{p}}}({{\rm{\Omega }}}_{{\rm{m}}},{ \mathcal M })$ are the observed and predicted apparent magnitudes, and C stat and C sys are the statistic and systematic covariance matrices. In our standard analyses, the SN Ia likelihood has two free parameters, Ωm and ${ \mathcal M }$. When analyzing the SN Ia data alone, we fixed h = 0.7 for the radiation and neutrino terms. Since the relevant redshifts are low and these terms are very small, ignoring them leads to almost the same constraint on Ωm and ${ \mathcal M }$. In this work, we use the Pantheon compilation of SNe Ia data (Scolnic et al. 2018).

Here, we present a fast algorithm 15 to compute the above SN Ia likelihood, suitable for data sufficiently compact in the redshift space like the Pantheon compilation. The calculation of m p involves many evaluations of fM(z); each needs an integration over redshift from 0 to z (see Equation (A2)). For the Pantheon compilation, there are 1048 SNe Ia so that there are 1048 integrals to evaluate. To reduce computational cost, usually the comoving distance is first evaluated for some redshift grids and then interpolation is used to get the comoving distance at each observed redshift. Here, we present an even faster method essentially without a loss of accuracy. This is to take the advantage of the fact that all SN Ia are quite evenly and compactly distributed within 0.01 ≲ z ≲ 2.3, using a Runge–Kutta-like algorithm as follows. We first sort the data in a redshift-ascending order and use i to denote the ith SN Ia. The redshift separation between every adjoining pair of data points is usually small. This allows us to calculate fM(zi+1) by adding a small increase to fM(zi ). For the first SN Ia with the smallest redshift, we calculate fM(z1) using the integral Equation (A2). Then beginning with the second SN Ia, we successively calculate fM(zi+1) as

Equation (A16)

where Δzi,i+1zi+1zi and E(z) is defined in Equation (A3). In other words, we are replacing each integral for i ≥ 2 by an addition. The number of kj terms needed, n, and the coefficients cj and bj are determined by the required order of accuracy. The more terms included in Equation (A16), the more accurate but slower the algorithm becomes. In Table 3, we list the values of c and b that allow the error of fM(zi+1) − fM(zi ) to be of the order ${ \mathcal O }({\rm{\Delta }}{z}_{i,i+1}^{3})$ for including one term, ${ \mathcal O }({\rm{\Delta }}{z}_{i,i+1}^{5})$ for two terms, and ${ \mathcal O }({\rm{\Delta }}{z}_{i,i+1}^{7})$ for three terms. If Δzi,i+1 > 0.1, we force the algorithm to use Equation (A2) to calculate fM(zi+1). We tested our algorithm on a python code, and found that it is about an order of magnitude faster than the algorithm that uses Equation (A2) with the function quad in the scipy package to calculate fM(z). While having such a high speed, our algorithm does not lose any accuracy. For the case including two k terms in Equation (A16), the numerical difference ($| {\rm{\Delta }}\mathrm{ln}{ \mathcal L }| $) between our approximation and the traditional method is well below 10−5. Considering both accuracy and speed, we use the two-term algorithm in this work. For the Pantheon SN Ia compilation, we reproduced Ωm = 0.298 ± 0.022 in the standard ΛCDM model (Scolnic et al. 2018).

Table 3. List of Coefficients for One, Two, and Three Terms to be Included in Equation (A16)

 One TermTwo TermsThree Terms
Coefficients c1 = 1, ${b}_{1}=\tfrac{1}{2}$ ${c}_{1}=\tfrac{1}{2}$, ${b}_{1}=\tfrac{1}{3-\sqrt{3}}$, ${c}_{2}=\tfrac{1}{2}$, ${b}_{2}=\tfrac{1}{3+\sqrt{3}}$ ${c}_{1}=\tfrac{4}{9}$, ${b}_{1}=\tfrac{1}{2}$, c2 = 5/18, ${b}_{2}=\tfrac{1}{5-\sqrt{15}}$, c3 = 5/18, ${b}_{3}=\tfrac{1}{5+\sqrt{15}}$
Orders [accumulated] ${ \mathcal O }({\rm{\Delta }}{z}_{i,i+1}^{3})$ $[{ \mathcal O }({\rm{\Delta }}{z}^{2})]$ ${ \mathcal O }({\rm{\Delta }}{z}_{i,i+1}^{5})$ $[{ \mathcal O }({\rm{\Delta }}{z}^{4})]$ ${ \mathcal O }({\rm{\Delta }}{z}_{i,i+1}^{7}$ $[{ \mathcal O }({\rm{\Delta }}{z}^{6})]$
Speeds up by∼14 times∼8 times∼6 times

Note. The two-term approximation is what we use with considerations of both numerical accuracy and speed.

Download table as:  ASCIITypeset image

Footnotes

  • 4  

    In this work, early refers to cosmic time before recombination and late refers to cosmic time later than that.

  • 5  

    Being uncalibrated means that the absolute magnitude or scale of the standard candles or rulers are not determined by calibration with another astrophysical observation or calculated with an assumed model.

  • 6  

    The end of the drag epoch refers to the time when photon pressure is no longer able to prevent baryons from falling into the potential wells of the cold dark matter (CDM).

  • 7  

    We also provide a fast algorithm to compute the SN Ia likelihood in Appendix A.5.

  • 8  

    To illustrate how uncalibrated θcmb and BAO constrain the evolution of the cosmic background, we only plot θd for BAO but not ${\rm{\Delta }}{z}_{{r}_{{\rm{d}}}}$. The bottom panel of our Figure 1 is similar to Figure 2 in eBOSS Collaboration et al. (2021) except that we are plotting θd in the y axis. But we note that we are adding θcmb in the analysis in an early universe physics insensitive way.

  • 9  

    There is some correlation between θcmb and z*, which has also been considered in our analyses.

  • 10  

    For a late-time nonstandard model, one needs to modify the standard E(z) using the corresponding nonstandard physics, to capture the effect on ΔrH0.

  • 11  

    While there is a limit on how early the actual formation time of the old stars can be, if they formed later, the inferred cosmic age would be higher and the inferred H0 would be lower.

  • 12  

    It is however worth pointing out that the TRGB-based result remains controversial; see Freedman et al. (2019), Yuan et al. (2019), Freedman et al. (2020), and Soltis et al. (2021).

  • 13  

    We thank Matias Zaldarriaga for bringing the reference (BOSS Collaboration et al. 2015) to our attention.

  • 14  

    In other words, if we can measure UCS extremely well, H0 can be inferred even without any calibration to cosmic standards!

  • 15  

    The speed comparison is based on a single-core machine.

Please wait… references are loading.
10.3847/1538-4357/ac12cf