Next Article in Journal
Asymmetry in Galaxy Spin Directions—Analysis of Data from DES and Comparison to Four Other Sky Surveys
Next Article in Special Issue
The Galactic Interstellar Medium Has a Preferred Handedness of Magnetic Misalignment
Previous Article in Journal
Thermodynamics of Hot Neutron Stars and Universal Relations
Previous Article in Special Issue
Measuring the Modified Gravitational Wave Propagation beyond General Relativity from CMB Observations
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Stringy Signals from Large-Angle Correlations in the Cosmic Microwave Background?

by
Miguel-Angel Sanchis-Lozano
1,2
1
Department of Theoretical Physics, University of Valencia, Doctor Moliner, 50, 46011 Burjassot, Spain
2
Instituto de Física Corpuscular (IFIC), CSIC-University of Valencia, 46100 Burjassot, Spain
Universe 2022, 8(8), 396; https://doi.org/10.3390/universe8080396
Submission received: 20 May 2022 / Revised: 22 July 2022 / Accepted: 24 July 2022 / Published: 28 July 2022
(This article belongs to the Special Issue Cosmic Microwave Background)

Abstract

:
We interpret the lack of large-angle temperature correlations and the even-odd parity imbalance observed in the cosmic microwave background (CMB) by COBE, WMAP and Planck satellite missions as a possible stringy signal ultimately stemming from a composite inflaton field (e.g., a fermionic condensate). Based on causality arguments and a Fourier analysis of the angular two-point correlation function, two infrared cutoffs k min even , odd (satisfying k min even 2 k min odd ) are introduced to the CMB power spectrum associated, respectively, with periodic and antiperiodic boundary conditions of the fermionic constituents (echoing the Neveu–Schwarz–Ramond model in superstring theory), without resorting to any particular model.

1. Introduction

Since its discovery in 1964 by Penzias and Wilson, the study of the cosmic microwave background (CMB) has provided a wealth of information on the early universe and its evolution, even going back in time closer to the Big Bang than the recombination phase when it was released. Indeed, to account for the outstanding homogeneity of the CMB temperature across the sky, an inflationary phase has been put forward to solve the so-called horizon problem, together with the flatness issue and “unwanted” relics, such as unseen magnetic monopoles [1].
Although there are alternative models, inflation has currently become the commonly accepted paradigm to explain cosmological evolution because of a series of convincing reasons, of which we just highlight one: observations of the CMB temperatures across the sky show that the universe is approximately uniform and homogeneous at scales much larger than the particle horizon at decoupling time, and a phase of exponential growth can account for this.
In principle, the basic requirement needed for a successful inflation is a scalar field satisfying the slow-roll conditions and a graceful exit. Actually, the inflaton needs not be a fundamental (nor single) scalar field but an effective, e.g., composite, field (see ref. [2] for a review) stemming from more fundamental interacting fields (such as a fermionic condensate1). Through the process of inflation, initial small quantum fluctuations of the underlying inflationary field when exiting the Hubble horizon would have formed the seeds for the matter distribution and subsequent growth to structures seen in today’s universe [5], showing an overall isotropy and homogeneity at large scales. Alternatively, in a linearly expanding cosmology, like the so-called R h = c t universe (for an exhaustive review of this model, see [6]), the quantum fluctuations of the primordial field driving the expansion turned into (semi-)classical fluctuations once exiting the Planck domain at about the Planck time [7].
On the other hand, in spite of the many successes of the Standard Cosmological Model ( Λ CDM), anomalies and tensions have emerged from a variety of astrophysical and cosmological observations (see, e.g., [8,9,10,11,12,13,14]). In particular, the unexpected lack of large-angle angular correlations (usually related to missing power at low multipoles) observed by all three major satellite missions, COBE [15], WMAP [16] and Planck [17,18], together with the odd-parity preference in the two-point correlation function of the CMB [19,20,21], were examined in detail in Ref. [22]. To this end, an infrared cutoff was set in the primordial CMB power spectrum following the original work of [23]. At the same time, odd-parity dominance was taken into account by weighting the different multipole contributions to angular correlations. Needless to say, such a parity imbalance questions the large-scale isotropy of the observable universe and hence the cosmological principle.
Here we shall perform a similar analysis but while introducing two infrared cutoffs (rather than one) to the CMB power spectrum ( k min even , odd ), affecting differently the even- and odd-parity multipole contributions in the best fit of the temperature two-point correlation function C ( θ ) to Planck 2018 data. Firstly, the aim of using two infrared cutoffs, rather than one, is to simultaneously achieve both goals (reproducing missing large-angle correlations and odd-parity dominance) with fewer fitting parameters, while suggesting an interesting and novel physical interpretation. Indeed, large-angle correlations should provide details on the very first stages of the universe, maybe hinting new physics at unexplored large scales, as addressed elsewhere [24,25]. This possibility stands as the main motivation of this paper.
In order to obtain the best fit of C ( θ ) to Planck data, we find that the ratio k min even / k min odd must be numerically close to two. Later, we will provide a tentative theoretical explanation for this, based on an assumed “stringy”2 behaviour of the fundamental primordial field(s) driving inflation in the early universe.

2. Angular Correlations in the CMB

All three COBE, WMAP and Planck satellite missions have observed that the temperature angular distribution of the CMB is remarkably homogeneous across the sky, with anisotropies of order 1 part in 10 5 . A powerful test of these fluctuations relies on the two-point (and higher [28,29]) angular correlation function C ( θ ) , defined as the ensemble product of the temperature differences with respect to the average temperature T 0 from all pairs of directions in the sky defined by unitary vectors n 1 and n 2 :
C ( θ ) = δ T ( n 1 ) T 0 δ T ( n 2 ) T 0 ,
where θ [ 0 , π ] is the angle defined by the scalar product n 1 · n 2 .
Assuming azimuthal symmetry, C ( θ ) can be expanded using the Legendre polynomials as:
C ( θ ) = 1 4 π = 2 ( 2 + 1 ) C P ( cos θ ) ,
where the C multipole coefficients encode the information with astrophysical/cosmological significance. In practice, the sum starts at = 2 and ends at a given max , dictated by the resolution of the data (in our case we set = 400 as an upper limit). The first two terms are excluded because (i) the monopole ( = 0 ) is simply the average temperature over the whole sky and plays no role in the correlations, other than a global scale shift; and (ii) the dipole ( = 1 ) is greatly affected by Earth’s motion, creating an anisotropy dominating over the intrinsic cosmological dipole signal, being usually removed from the multipole analysis.
Under some simplifying assumptions [5], the coefficients C in Equation (2) can be evaluated according to the following expression:
C = N 0 d k k n s 2 j 2 ( k r d ) ,
where N is a normalization constant, n s 1 is the spectral index and j ( k r d ) denotes the spherical Bessel function of order , whose argument involves the comoving distance r d from the last scattering surface (LSS) to us, and the wavenumber k for each mode in the CMB power spectrum which varies, in principle, from zero to infinity.

3. Single Infrared Cutoff in the CMB Power Spectrum

As already commented, large-angle temperature correlations in the CMB provide information on the earliest stages of the primitive universe, well before recombination and the subsequent formation of the cosmic structure [5]. In this regard, the C ( θ ) function defined in Equation (2) was found to be close to zero above 60 –70 from all three COBE, WMAP and Planck data, constituting one of the anomalies between standard cosmology and observations [8]. Indeed, the suppression of large-angle correlations together with the existence of a downward tail at large angles (≳150 ) were unexpected in standard cosmology, given that inflation is supposed to provide the required number of e-folds (≥60) to solve both the horizon and flatness problems, thereby bringing angular coverage across the full sky.
In order to mitigate this tension, the authors of [23] introduced an infrared cutoff in the CMB power spectrum, leading to a lower limit k min in the integral of Equation (3)
C = N k min d k k n s 2 j 2 ( k r d ) ,
yielding a suppression of large-angle correlations. The normalization constant N and k min are obtained from a global fit to the whole angular correlation function.
A theoretical motivation to this truncation of k modes entering the integral of Equation (4) relies on a maximum correlation length obtained from a causality requirement at decoupling time t d [23,30]:
λ max = 2 π R h ,
where R h stands for the Hubble radius at that time. Actually, the Hubble radius R h is not necessarily a true horizon, but should provide an order of magnitude estimate of the maximum correlation length based on causality. In fact, λ max = α 2 π R h , where α 1 depends on the cosmological model, e.g., α 0.5 for the Λ CDM [31].
On the other hand, the comoving wavenumber k min associated with λ max reads
k min = 2 π a ( t ) λ max ,
which can be interpreted as the first quantum fluctuation in the underlying primordial field driving the early universe expansion either (i) having crossed the Hubble horizon once inflation started, or (ii) having emerged out of the Planck domain [7] should inflation have never happened, as in a R h = c t universe.
Next, changing the integration variable in Equation (3) from k to the dimensionless variable u k r d and setting n s = 1 for simplicity, one gets
C = N u min d u j 2 ( u ) u .
Let us point out that only those C coefficients with 10 are actually affected by the lower cutoff u min = k min r d 0 in the above integral.
In ref. [22], the numerical interval u min = 4.5 ± 0.5 was obtained from a best fit to the Planck 2018 dataset. In the upper panel of Figure 1, we plot C ( θ ) corresponding to the Λ CDM prediction ( u min = 0 ), together with the best fit to Planck data [17,18], as done in [23] with u min = 4.5 but keeping even-odd parity balance. One can see at once that the standard cosmological model fails to fit the data, while the latter curve indeed yields almost zero correlations at 70 , but fails to reproduce the observed downward tail at 180 . Finally, in the lower panel of Figure 1, we reproduce the excellent fit ( χ 2 / d . o . f . 1 ) performed in [22] to the same dataset requiring odd-parity dominance.
On the other hand, the maximum correlation angle θ max can be roughly estimated as
θ max λ max R d θ max 2 π u min ,
where R d = a ( t d ) r d stands for the proper distance from the LSS to us, and u min = k min r d .
For later use, let us regroup the even and odd -multipoles in Equation (2), namely
C ( θ ) = C even ( θ ) + C odd ( θ ) = 1 4 π even ( 2 + 1 ) C P ( cos θ ) + 1 4 π odd ( 2 + 1 ) C P ( cos θ ) .
To better understand the behaviour of the C ( θ ) curve, we plot separately in Figure 2 the even and odd parity pieces C even ( θ ) and C odd ( θ ) as a function of θ for [ 2 , 401 ] . Because of the oscillatory behaviour of the Legendre polynomials, both contributions add positively at small and middle angles, while a delicate balance is needed in order to obtain zero correlation at larger angles. This balance is indeed achieved [22] when using a single cutoff u min in the evaluation of the C coefficients in Equation (7).
However, if two distinct infrared cutoffs apply differently to even and odd -modes, such a delicate balance will be likely broken. Then, odd or even parity dominance can be obtained quite naturally. In the next section, we examine in depth this issue of paramount importance in this work.

4. Two Infrared Cutoffs in the CMB Power Spectrum

In ref. [22], the fit to Planck 2018 data was optimized using a k min together with the requirement of parity imbalance by weighting adequately the odd and even terms in Equation (2). We shall now carry out a similar analysis but introducing two infrared cutoffs in the CMB power spectrum, namely, k min even and k min odd , affecting the even- and odd-parity coefficients differently.
The goals of this new analysis are to provide: (i) a common explanation of both missing large-angle correlations and odd-parity dominance with fewer fitting parameters; and (ii) a physical interpretation behind. First, the numerical values of k min even and k min odd were heuristically determined from a fit of C ( θ ) to data, checking its consistency with odd-parity preference by means of a parity statistic to be discussed in Section 4.2. Depending on which k min even or k min odd is larger, odd or even dominance can be achieved.
On the one hand, the introduction of two infrared cutoffs in the angular power spectrum can be merely seen as a phenomenological approach to jointly account for missing large-angle correlations and the downwards tail in the C ( θ ) curve. Seen in this way, just an additional fitting parameter is required with respect to the analysis done in [23]. On the other hand, as we claim in this work, the ratio of both infrared cutoffs would become “fixed by theory” as a consequence of an extra (fermionic) spin degree of freedom due to the assumed fermionic nature of the underlying inflating field(s). Hence, no additional parameter is actually introduced into the fits.
Therefore, we compute separately the C coefficients of the even and odd -multipoles according to
C even / odd = N u min even / odd d u j 2 ( u ) u ,
equivalent to Equation (7) for a single k min without a distinction of parity. Now, the integral lower limits are: u min even , odd = k min even , odd r d , respectively, where r d denotes again the distance from the LSS to us.
In Figure 3, we plot the fit of of C ( θ ) to Planck data under the assumption of two lower cutoffs u min even , odd in the integral of Equation (7). From a best χ 2 / d . o . f . fit, we obtain u min odd 2.67 and, consequently, u min even 5.34 . Admittedly, the new fit as seen in Figure 3 looks a bit worse than in the lower panel of Figure 1. However, notice that the number of free parameters is now smaller and, above all, one can simultaneously accommodate the lack of large-angle correlations and the apparent odd-parity dominance seen in the downward tail 150 , without resorting to any further fine-tuning of the C coefficients.
Next, we estimate the statistical error of u min odd (and u min even = 2 u min odd ). As is well known, the C ( θ ) points are highly correlated, so that a calculation based on a simple χ d . o . f . 2 procedure is not the most suitable. We circumvented this problem by using a Monte Carlo procedure similar to what was done in [22,23] to sample the variation of C ( θ ) within the measurement errors. To this end, we generated 100 mock CMB catalogs with bin size of 0 . 5 each, corresponding to an ideal full sky case. Then, we computed the two-point correlation function C i ( θ ) ( i = 1 , 100 ) in each realization. Next, we obtained Δ C i ( θ ) = C i ( θ ) C 0 ( θ ) , where C 0 ( θ ) denotes the angular correlation function in standard cosmology. Defining C ( θ ) = C P l a n c k ( θ ) + Δ C i ( θ ) , we computed u min , i odd from a best χ d . o . f . 2 independent fit in each case, recalculating the C coefficients and the respective normalization constant N of Equation (4). Let us mention that the latter shows an almost linear dependence (of negative slope) on the lower cutoff, namely, N [ 340 , 180 ] versus u min odd [ 2.36 , 2.98 ] .
In Figure 4, we plot as a histogram the resulting u min , i odd / even distribution(s) for all 100 realizations, both showing approximate Gaussian shapes. From these resulting roughly Gaussian distribution(s), we determined their average values and r.m.s., yielding
u min odd = 2.67 ± 0.31 ; u min even = 5.34 ± 0.62 .
The above values imply non-zero values for u min odd / even corresponding to a 8.6 σ statistical significance in our analysis. Such a high confidence level reflects the need for two different lower cutoffs in the integral of Equation (10) for even and odd parity modes, in contrast to Λ CDM, to reproduce the angular correlation anomalies addressed in this work. As an aside, notice that the mean value of u min even , odd lies within the interval u min = 4.5 ± 0.5 found in [22] for a single lower cutoff, as naively expected.
Had we tried the opposite situation, k min even < k min odd , the C ( θ ) curve would have shown an upward tail, contrary to observational data. The mathematical reason why the condition k min even > k min odd goes in the “right” direction is that the C coefficients (for 10 ) decrease for odd to a lower extent than for even (as compared to no lower cutoff applied at all), thereby favoring odd-parity dominance, in accordance with observations. Let us note the greatest effect of the lower cutoffs affects C 2 and C 3 multipole coefficients (thereby decreasing the ratio C 2 / C 3 ), while other low- modes also become modified (reduced), but to a lesser extent. This makes a difference with respect to, e.g., the ellipsoidal universe where only the quadrupole term is modified as compared to standard cosmology [32].
Next, notice the following relations between Legendre polynomials and the square of cosine functions with entire and half-entire (2 π ) periods (these relations can also be formulated in terms of Chebyshev polynomials [33]), called to play a fundamental role in our later development:
P 1 ( cos θ ) = 1 + 2 cos 2 ( θ / 2 ) P 2 ( cos θ ) = 0.5 + 1.5 cos 2 ( θ ) P 3 ( cos θ ) = 1 + 0.75 cos 2 ( θ / 2 ) + 1.25 cos 2 ( 3 θ / 2 ) P 4 ( cos θ ) = 0.7184 + 0.6249 cos 2 ( θ ) + 1.0937 cos 2 ( 2 θ ) P 5 ( cos θ ) = 1 + 0.4687 cos 2 ( θ / 2 ) + 0.5469 cos 2 ( 3 θ / 2 ) + 0.9844 cos 2 ( 5 θ / 2 )
Higher order Legendre polynomials replicate the same pattern: besides a constant, even and odd Legendre polynomials either contain cos 2 [ n θ ] or cos 2 [ ( n + 1 / 2 ) θ ] terms, respectively. This pattern is just mathematical for the moment, below to be put in correspondence with the Fourier expansion of the constituents of a composite inflaton field, using a one-dimensional toy-model.
Inspired on very general grounds by superstring theory (obviously without entering at all in its complexity and intricacies [26,34]) and invoking a similar causality argument as employed for the single k min case, we address now the possibility that any constituent (fermionic) field ψ ( φ ) (for simplicity, no more variables or indices are explicitly written) of a composite inflaton 3 satisfies either the periodic or antiperiodic boundary condition:
ψ ( φ + 2 π ) = ψ ( φ )
ψ ( φ + 2 π ) = ψ ( φ ) ψ ( φ + 4 π ) = ψ ( φ ) .
The angular Fourier expansion of ψ ( φ ) for the periodic condition reads:
ψ ( φ ) = n Z α n e i n φ
and for a real function,
ψ ( φ ) = 2 n Z + Re α n e i n φ ,
so that only cos ( n φ ) terms appear in the Fourier expansion.
Let us define now a correlation function as in [36]
0 2 π ψ ( φ ) ψ ( φ + Δ φ ) d φ 2 π .
For random Gaussian Fourier coefficients, if we define θ = Δ φ / 2 , one finds
C ( Δ φ ) = 2 n Z + C n cos ( n Δ φ ) C ( θ ) = 4 n Z + C n cos 2 ( n θ ) 2 ,
with C n = α n α n * and θ [ 0 , π ] to be identified with the angle appearing as the argument of the two-point correlation function C ( θ ) .
For the antiperiodic conditions, the Fourier expansion reads
ψ ( φ ) = n Z + + 1 / 2 α n e i n φ ,
which guarantees that it changes sign when φ φ + 2 π . Therefore,
C ( θ ) = 4 n Z + + 1 / 2 C n cos 2 ( n θ ) 2 .
The above results obtained from this toy-model suggest the assignment of even- and odd-parity C even ( θ ) and C odd ( θ ) pieces in Equation (9), to fulfill the periodicity and antiperiodicity conditions given in Equations (12) and (13), respectively, somewhat recalling the well-known Ramond and Neveu–Schwarz sectors in superstring theory [26]. Hereafter, we shall assume that the above conclusions apply to the fundamental field(s) driving the expansion of the early universe.

4.1. Physical Interpretation

In the following, we will consider, without entering into details, the inflaton as an effective scalar field Φ made out of two constituent fermionic fields ( Φ = Ψ ¯ Ψ ) driving the expansion of the early universe. The causality condition yielding λ max used for a single infrared cutoff is now assumed to split into two conditions applied independently to the periodic and the antiperiodic conditions in virtue of the spin degrees of freedom of the constituent fermionic fields of the effective inflaton.
Hence, in analogy to Equation (5), we shall now consider two maximum correlation lengths for even and odd parity multipoles, denoted as λ max even and λ max odd , satisfying
λ max even = 2 π R h , λ max odd = 4 π R h .
Then, two comoving wavenumbers k min even and k min odd can be defined from those length scales:
k min odd = 2 π a ( t d ) / λ max odd , k min even = 2 π a ( t d ) / λ max even
corresponding to odd and even parity modes, respectively. Let us remark that physical momenta are given by k / a ( t ) , but the ratio k min odd / k min even remains frozen after exiting the horizon. Note also that the numerical values of the lower cutoffs are obtained from a fit to correlation data and not from Equation (21). Indeed, what really matters in this work is that their ratio is equal (or close) to 2.
Moreover, two relevant angles can be defined following Equation (8) for a single k min :
θ even 2 π u min even 65 , θ odd 2 π u min odd 130 ,
defining three different angular intervals along the full 0 –180 range, where the C ( θ ) function exhibits a different behavior, as can be easily appreciated in Figure 1 and Figure 3.
Now, from our discussion in the previous section on the Fourier expansion of C ( θ ) , we will take into account that the C even ( θ ) and C odd ( θ ) pieces of Equation (9) can be associated with terms of the type cos 2 [ n θ ] and cos 2 [ ( n + 1 / 2 ) θ ] , respectively, as shown in Equation (12). Therefore, the two distinct lower limits u min even , odd for even and odd multipoles in the integral of Equation (4) can be related to the infrared cutoffs k min even , odd , respectively.
Moreover, from the fit to Planck data shown in Figure 3, the ratio u min even / u min odd = k min even / k min odd turns out to be quite close to 2. This phenomenological fact may be further understood by considering the periodicity and antiperiodicity conditions applied to the maximum correlation lengths:
u min even u min odd = k min even k min odd = λ max odd λ max even = 2 ,
as implied by Equation (21). In the framework of the R h = c t universe, this relation corresponds to two different exit times of the Planck regime: first the k min odd mode and later the k min even mode. In the usual inflationary scenario, it implies again two different times in similar order but exiting the Hubble horizon. Needless to say, the above ratio 2 is expected to be valid only approximately, but it has been exactly to this value in our numerical analysis.

4.2. Parity Statistic Study

Let us now address the deviation from the even-odd parity balance employing the parity statistic [19,20], as done in [22]
P ( max ) = P + ( max ) P ( max ) ; P ± ( max ) = = 2 max γ ± ( + 1 ) 2 π C ,
with the projectors defined as γ + = cos 2 ( π / 2 ) and γ = sin 2 ( π / 2 ) . Assuming that ( + 1 ) C constant is satisfied at low , P ± can clearly be considered as a measurement of the degree of parity asymmetry. Any deviation from unity of this statistic points to an even-odd parity imbalance: below unity, it implies odd-parity dominance and vice versa. Below, we apply this statistic to compare different fits with one or two infrared cutoffs.
In Figure 5, we plot P ( max ) as a function of max corresponding to (a) upper panel: u min = 0 (expected from Λ CDM); (b) our previous result for u min = 4.5 and odd-parity dominance [22]; (c) the new analysis performed in this paper setting u min even = 2 u min odd 5.34 , as obtained before. From inspection of these plots it becomes apparent that the fit using a k min odd , even pair somewhat worsens with respect to the single k min 0 case: the reduced χ 2 / d . o . f . increases from almost unity to about twice. Nonetheless, as already emphasized, the main interest of providing a common explanation to both missing large-angle correlations and odd-parity dominance, remains.

5. Discussion and Conclusions

As claimed in ref. [23], the lack of large-angle correlations observed in the CMB from WMAP and Planck 2018 data can be reproduced, introducing a lower cutoff k min to the CMB power spectrum, thereby implying a lower u min in the computation of all (even and odd) C coefficients of the multipole expansion in Equation (4). Furthermore, the apparent odd-parity dominance manifesting as a downward tail of C ( θ ) at large angles ( θ 150 ), but keeping the good behaviour at lower angles, can also be accommodated thanks to ad hoc extra weights of the multipole coefficients C ( 10 ), leading at the same time to an excellent fit of the parity statistic P ( max ) [22].
No obvious theoretical connection between such an infrared cutoff k min (i.e., lack of large-angle correlations) in the power spectrum and parity imbalance has been found so far [21,22,37,38]. In this paper, however, we put forward a possible relationship between both by introducing two infrared cutoffs k min even and k min odd , whose ratio is close to 2 from a χ d . o . f . 2 best fit to Planck 2018 data. In this way we are able: (i) to reduce the number of fitting parameters with respect to [22], while reproducing the downward tail at large angle; (ii) to provide a theoretical connection (without resorting to any particular model) between both observations, setting the ratio k min even / k min odd = u min even / u min odd exactly equal to 2.
Thus, new fits of C ( θ ) and the parity statistic P ( max ) have been performed using the same Planck 2018 data set as for the single infrared cutoff case [22]. Since the number of fitting parameters is now considerably smaller, the goodness of the fits (determined by their χ 2 / d . o . f . ) somewhat worsens. However, let us emphasize the added value of our approach due to a common and suggestive explanation of both anomalies: missing correlations above 60 –70 and a downward tail at ≳150 , based on a suggestive physical interpretation.
For the numerical computations of the lower cutoffs and their associated errors, a Monte Carlo procedure was employed generating 100 mock CMB maps to sample the variation of C ( θ ) within the measurement uncertainties, yielding
u min odd = 2.67 ± 0.31 ; u min even = 5.34 ± 0.62 ,
corresponding to a 8.6 σ statistical significance of our working hypothesis. The need for non-zero lower cutoffs u min odd / even (in contrast to Λ CDM) to reproduce the anomalies of angular correlations examined in this paper has to be interpreted with care within our phenomenological approach.
On the other hand, based on a Fourier analysis and using a toy-model, we have associated even and odd multipoles of C ( θ ) with periodic and antiperiodic boundary conditions satisfied by the underlying (fermionic fields) of a composite inflaton, or a fermionic inflaton itself. Thus, the phenomenologically motivated choice for the ratio k min even / k min odd becomes supported by theoretical arguments based on spin, beyond the initial heuristic approach. This constitutes the main result of this work.
As is well known, spin degrees of freedom play a crucial role in elementary particle physics, e.g., predicting the gyromagnetic ratio g for the self-spinning electron to be about two times bigger than the value for an orbiting electron [39]. Needless to say, the famous muon g 2 deviation from zero stands as an important test of new physics, currently under close scrutiny [40]. The idea of searching for new physics and phenomena either at particle colliders on Earth and/or looking at the sky, in a complementary way, is certainly not new, with dark matter being a good example of it.
Of course, explanations alternative to ours for missing large-angle correlations together with odd-parity dominance are possible: a statistical fluke of data at large angles, contamination or non-cosmological effects [41,42], cosmic variance [43], or even other theoretical underlying reasons.
Finally, let us point out that the future very-high precision measurements of CMB polarization [44,45,46], might follow (or not) an angular pattern similar to the temperature fluctuations seen in the sky, thereby shedding light on this tantalizing hypothesis.

Funding

This work was partially funded by Spanish Agencia Estatal de Investigación under grant PID2020-113334GB-I00/AEI/10.13039/501100011033, by Generalitat Valenciana under grant PROMETEO/2019/113 (EXPEDITE).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

I acknowledge inspiring discussions with V.Sanz at the beginning of this work. I also thank F. Melia and M. Lopez-Corredoira for comments and suggestions.

Conflicts of Interest

The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript, or in the decision to publish the results.

Abbreviations

The following abbreviations are used in this manuscript:
CMBCosmic microwave background
Λ CDMStandard Λ cold dark matter cosmological model
LSSLast scattering surface

Notes

1
Historically, the interpretation of scalars as fermionic bound states, e.g., in the Nambu–Jona–Lasinio model, dates back to the 1960s in the past century, to address strong interaction issues [3,4]). Nowadays, the possibility of a composite Higgs boson is also seriously contemplated.
2
In this paper stringy refers generically to properties associated with periodic and antiperiodic boundary conditions, such as those found in superstring theory [26], and not, in principle, to cosmic strings or superstrings [27], understood as one-dimensional structures of false vacuum [1].
3
The possibility that a fermionic field, minimally or non-minimally coupled with the gravitational field, drives the inflation, without resorting to a composite inflaton, has also been studied (see e.g., [35] and references therein).

References

  1. Kolb, E.W.; Turner, M. The Early Universe, Fontiers in Physics; Westview Press: Boulder, CO, USA, 1994. [Google Scholar]
  2. Samart, D.; Pongkitivanichkul, C.; Channuie, P. Composite dynamics and cosmology: Inflation. Eur. Phys. J. Spec. Top. 2022, 231, 1325–1344. [Google Scholar] [CrossRef]
  3. Nambu, Y.; Jona-Lasinio, G. Dynamical Model of Elementary Particles Based on an Analogy with Superconductivity. Phys. Rev. 1961, 122, 345. [Google Scholar] [CrossRef]
  4. Gross, D.J.; Neveu, A. Dynamical symmetry breaking in asymptotically free field theories. Phys. Rev. D 1974, 10, 3235. [Google Scholar] [CrossRef]
  5. Mukhanov, V.F. Physical Foundations of Cosmology; Cambridge University Press: Cambridge, UK, 2005. [Google Scholar]
  6. Melia, F. The Cosmic Spacetime; CRC Press: Boca Raton, FL, USA, 2020. [Google Scholar]
  7. Melia, F. Quantum Fluctuations at the Planck Scale. EPJ-C 2019, 79, 455. [Google Scholar] [CrossRef]
  8. Abdalla, E.; Abellan, G.F.; Aboubrahim, A.; Agnello, A.; Akarsu, O.; Akrami, Y.; Pettorino, V. Cosmology Intertwined: A Review of the Particle Physics, Astrophysics, and Cosmology Associated with the Cosmological Tensions and Anomalies. J. High Energy Astrophys. 2022, 34, 49–211. [Google Scholar] [CrossRef]
  9. Perivolaropoulos, L.; Skara, F. Challenges for ΛCDM: An update. arXiv, arXiv:2105.05208.Unpublished.
  10. Di Valentino, E.; Mena, O.; Pan, S.; Visinelli, L.; Yang, W.; Melchiorri, A.; Mota, D.F.; Riess, A.G.; Silk, J. In the Realm of the Hubble tension: A Review of Solutions. Class. Quantum Grav. 2021, 38, 153001. [Google Scholar] [CrossRef]
  11. Vagnozzi, S.; Loeb, A.; Moresco, M. Eppur e piatto? The cosmic chronometer take on spatial curvature and cosmic concordance. Astrophys. J. 2021, 908, 84. [Google Scholar] [CrossRef]
  12. Shaikh, S.; Mukherjee, S.; Das, S.; Wandelt, B.; Souradeep, T. Joint Bayesian analysis of large angular scale CMB temperature anomalies. JCAP08 2019, 2019, 007. [Google Scholar] [CrossRef]
  13. Muir, J.; Adhikari, S.; Huterer, D. Covariance of CMB anomalies. Phys. Rev. D 2018, 98, 023521. [Google Scholar] [CrossRef]
  14. Rassat, A.; Starck, J.-L.; Paykari, P.; Sureau, F.; Bobin, J. Planck CMB Anomalies: Astrophysical and Cosmological Secondary Effects and the Curse of Masking. JCAP 2014, 8, 6. [Google Scholar] [CrossRef]
  15. Hinshaw, G.; Banday, A.J.; Bennett, C.L.; Górski, K.M.; Kogut, A.; Lineweaver, C.H.; Wright, E.L. Two-Point Correlations in the COBE DMR Four-Year Anisotropy Maps. Astrophys. J. 1996, 464, L25. [Google Scholar] [CrossRef]
  16. Bennett, C.L.; Larson, D.; Weiland, J.L.; Jarosik, N.; Hinshaw, G.; Odegard, N.; Wright, E.L. First-Year Wilkinson Microwave Anisotropy Probe (WMAP) Observations: Preliminary Maps and Basic Results. Astrophys. J. Suppl. Ser. 2003, 148, 97. [Google Scholar] [CrossRef]
  17. Aghanim, N.; Akrami, Y.; Ashdown, M.; Aumont, J.; Baccigalupi, C.; Ballardini, M.; Roudier, G. Planck 2018 results. VI. Cosmological parameters. Astron. Astrophys. 2018, 641, A86. [Google Scholar]
  18. Akrami, Y.; Ashdown, M.; Aumont, J.; Baccigalupi, C.; Ballardini, M.; Banday, A.J.; Zonca, A. Planck 2018 results. VII. Isotropy and Statistics of the CMB. Astron. Astrophys. 2019, 641, A7. [Google Scholar]
  19. Aluri, P.K.; Jain, P. Parity Asymmetry in the CMBR Temperature Power Spectrum. MNRAS 2012, 419, 3378. [Google Scholar] [CrossRef]
  20. Panda, S.; Aluri, P.K.; Samal, P.K.; Rath, P.K. Parity in Planck full-mission CMB temperature maps. Astropart. Phys. 2021, 125, 102493. [Google Scholar] [CrossRef]
  21. Land, K.; Magueijo, J. Is the Universe odd? Phys. Rev. D 2005, 72, 101302. [Google Scholar] [CrossRef]
  22. Sanchis-Lozano, M.A.; Lopez-Corredoira, M.; Sanchis-Gual, N. Missing large-angle correlations versus odd-parity dominance in the cosmic microwave background. Astron. Astrophys. 2022, 10, 142–149. [Google Scholar]
  23. Melia, F.; López-Corredoira, M. Evidence of a truncated spectrum in the angular correlation function of the cosmic microwave background. Astron. Astrophys. 2018, 610, A87. [Google Scholar] [CrossRef]
  24. Gruppuso, A.; Kitazawa, N.; Lattanzi, M.; Mandolesi, N.; Natoli, P.; Sagnotti, A. The Evens and Odds of CMB Anomalies. Phys. Dark Universe 2018, 20, 49–64. [Google Scholar] [CrossRef]
  25. Ashtekar, A.; Gupt, B.; Jeong, D.; Sreenath, V. Alleviating the Tension in the Cosmic Microwave Background Using Planck-Scale Physics. Phys. Rev. Let. 2020, 125, 051302. [Google Scholar] [CrossRef] [PubMed]
  26. Zwiebach, B. A First Course in String Theory; Cambridge University Press: Cambridge, UK, 2004; pp. 262–268. [Google Scholar]
  27. Ringeval, C.; Yamauchi, D.; Yokoyama, J.; Bouchet, F. Large scale CMB anomalies from thawing cosmic strings. JCAP 2016, 1602, 033. [Google Scholar] [CrossRef]
  28. Gangui, A.; Lucchin, F.; Matarrese, S.; Mollerach, S. The Three–Point Correlation Function of the Cosmic Microwave Background in Inflationary Models. Astrophys. J. 1994, 430, 447. [Google Scholar] [CrossRef]
  29. Kamionkowski, M.; Souradeep, T. The Odd-Parity CMB Bispectrum. Phys. Rev. D 2011, 83, 027301. [Google Scholar] [CrossRef]
  30. Melia, F. Angular Correlation of the CMB in the Rh=ct Universe. A&A 2012, 561, A80. [Google Scholar]
  31. Melia, F. Proper Size of the Visible Universe in FRW Metrics with Constant Spacetime Curvature. Class. Quant. Grav. 2013, 30, 155007. [Google Scholar] [CrossRef]
  32. Cea, P. CMB two-point angular correlation function in the Ellipsoidal Universe. arXiv 2022, arXiv:2203.14229. [Google Scholar]
  33. Abramowitz, M.; Stegun, I.A. Handbook of Mathematical Functions: With Formulas, Graphs and Mathematical Tables; Dover Publications, Inc.: New York, NY, USA, 1970. [Google Scholar]
  34. Liu, Z.G.; Guo, Z.K.; Piao, Y.S. CMB anomalies from an inflationary model in string theory. Eur. Phys. J. C 2014, 74, 3006. [Google Scholar] [CrossRef]
  35. Grams, G.; de Souza, R.; Kremer, G. Fermion field as inflaton, dark energy and dark matter. Class. Quant. Grav. 2014, 31, 185008. [Google Scholar] [CrossRef]
  36. Copi, C.J.; Gurian, J.; Kosowsky, A.; Starkman, G.D.; Zhang, H. Exploring suppressed long-distance correlations as the cause of suppressed large-angle correlations. Mon. Not. R. Astron. Soc. 2018, 490, 1–8. [Google Scholar]
  37. Kim, J.; Naselsky, P.; Hansen, M. Symmetry and Antisymmetry of the CMB Anisotropy Pattern. Adv. Astron. 2012, 2012, 960509. [Google Scholar] [CrossRef]
  38. Schwarz, D.J.; Copi, C.J.; Huterer, D.; Starkman, G.D. CMB anomalies after Planck. Class. Quantum Gravity 2016, 33, 184001. [Google Scholar] [CrossRef]
  39. Thomson, M. Modern Particle Physics; Cambridge University Press: Cambridge, UK, 2013; p. 517. [Google Scholar]
  40. Zyla, P.A. [Particle Data Group] Review of Particle Physics. Prog. Theor. Exp. Phys. 2022, 2020, 083C01. [Google Scholar]
  41. Creswell, J.; Naselsky, P. Asymmetry of the CMB map: Local and global anomalies. J. Cosmol. Astropart. Phys. 2021, 2021, 103. [Google Scholar] [CrossRef]
  42. Zhao, W.; Santos, L. Preferred axis in cosmology. Universe 2015, 3, 9. [Google Scholar]
  43. Copi, C.J.; Huterer, D.; Schwarz, D.J.; Starkman, G.D. Large angle anomalies in the CMB. Adv. Astron. 2010, 2010, 847541. [Google Scholar] [CrossRef]
  44. Armitage-Caplan, C.; Avillez, M.; Barbosa, D.; Banday, A.; Bartolo, N.; Battye, R.; Bernard, J.P.; de Bernardis, P.; Basak, S.; Bersanelli, M. et al. [COrE Collaboration] COrE (Cosmic Origins Explorer) A White Paper. arXiv 2011, arXiv:1102.2181. [Google Scholar]
  45. Young, K.; Alvarez, M.; Battaglia, N.; Bock, J.; Borrill, J.; Chuss, D.; Crill, B.; Delabrouille, J.; Devlin, M.; Fissel, L. et al. [NASA PICO] PICO: Probe of Inflation and Cosmic Origins. arXiv 2019, arXiv:1902.10541. [Google Scholar]
  46. Errard, J.; Feeney, S.M.; Peiris, H.V.; Jaffe, A.H. Robust forecasts on fundamental physics from the foreground-obscured, gravitationally-lensed CMB polarization. J. Cosmol. Astropart. Phys. 2016, 2016, 052. [Google Scholar] [CrossRef]
Figure 1. Two-point correlation function C ( θ ) ) (solid curves) obtained from the fit to Planck 2018 data. Upper panel: Λ CDM prediction (in green) clearly deviating from data in almost the whole angular range; blue curve: u min = 4.5 without imposing a odd-parity dominance requirement. Lower panel: Red curve for u min even = 4.5 with odd-parity dominance from [22].
Figure 1. Two-point correlation function C ( θ ) ) (solid curves) obtained from the fit to Planck 2018 data. Upper panel: Λ CDM prediction (in green) clearly deviating from data in almost the whole angular range; blue curve: u min = 4.5 without imposing a odd-parity dominance requirement. Lower panel: Red curve for u min even = 4.5 with odd-parity dominance from [22].
Universe 08 00396 g001
Figure 2. The C even ( θ ) (in blue) and C odd ( θ ) (in magenta) contributions to the two-point angular correlation function C ( θ ) Equation (9), using u min = 4.5 . Notice the delicate balance between odd and even -multipoles (even at high-) in order to yield zero correlations above ≈60 –70 . Any slight deviation from this balance would yield a downwards (upwards) tail at ≃180 for odd (even) dominance of Legendre polynomials.
Figure 2. The C even ( θ ) (in blue) and C odd ( θ ) (in magenta) contributions to the two-point angular correlation function C ( θ ) Equation (9), using u min = 4.5 . Notice the delicate balance between odd and even -multipoles (even at high-) in order to yield zero correlations above ≈60 –70 . Any slight deviation from this balance would yield a downwards (upwards) tail at ≃180 for odd (even) dominance of Legendre polynomials.
Universe 08 00396 g002
Figure 3. Two-point correlation function C ( θ ) (orange solid curve) obtained from the fit to Planck 2018 data for u min even = 5.34 and u min odd = u min even / 2 = 2.67 . The fit is not so good as in Figure 1 (lower panel), but the number of fitting parameters is smaller and, above all, provides a theoretical foundation that simultaneously explains the lack of large-angle correlations and the observed odd-parity dominance leading to a downward tail near 180 .
Figure 3. Two-point correlation function C ( θ ) (orange solid curve) obtained from the fit to Planck 2018 data for u min even = 5.34 and u min odd = u min even / 2 = 2.67 . The fit is not so good as in Figure 1 (lower panel), but the number of fitting parameters is smaller and, above all, provides a theoretical foundation that simultaneously explains the lack of large-angle correlations and the observed odd-parity dominance leading to a downward tail near 180 .
Universe 08 00396 g003
Figure 4. Histogram of u min , i odd (left) and u min , i even (right), both showing approximate Gaussian shapes. Let us point out that the condition u min , i even = 2 u min , i odd was required throughout the data analysis.
Figure 4. Histogram of u min , i odd (left) and u min , i even (right), both showing approximate Gaussian shapes. Let us point out that the condition u min , i even = 2 u min , i odd was required throughout the data analysis.
Universe 08 00396 g004
Figure 5. P ( max ) statistic as a function of max compared to the Planck 2018 data for upper left panel: u min = 0 and even-odd parity balance ( Λ CDM); upper right panel: u min = 4.5 and weighted C to get odd-parity dominance; lower panel: u min even = 5.34 and u max odd = 2.67 .
Figure 5. P ( max ) statistic as a function of max compared to the Planck 2018 data for upper left panel: u min = 0 and even-odd parity balance ( Λ CDM); upper right panel: u min = 4.5 and weighted C to get odd-parity dominance; lower panel: u min even = 5.34 and u max odd = 2.67 .
Universe 08 00396 g005
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Sanchis-Lozano, M.-A. Stringy Signals from Large-Angle Correlations in the Cosmic Microwave Background? Universe 2022, 8, 396. https://doi.org/10.3390/universe8080396

AMA Style

Sanchis-Lozano M-A. Stringy Signals from Large-Angle Correlations in the Cosmic Microwave Background? Universe. 2022; 8(8):396. https://doi.org/10.3390/universe8080396

Chicago/Turabian Style

Sanchis-Lozano, Miguel-Angel. 2022. "Stringy Signals from Large-Angle Correlations in the Cosmic Microwave Background?" Universe 8, no. 8: 396. https://doi.org/10.3390/universe8080396

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop