Next Article in Journal
X-ray Properties of Two Complementary Samples of Intermediate Seyfert Galaxies
Previous Article in Journal
Probing Dark Sectors with Neutron Stars
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

New Indication from Quantum Chromodynamics Calling for beyond the Standard Model

1
Center for Theoretical Physics and College of Physics, Jilin University, Changchun 130012, China
2
Center for Liberal Arts and Sciences, Toyama Prefectural University, Toyama 939-0398, Japan
3
School of Nuclear Science and Technology, University of Chinese Academy of Sciences, Beijing 100049, China
4
RIKEN BNL Research Center, Brookhaven National Laboratory, Upton, NY 11973, USA
5
Department of Information Technology, International Professional University of Technology, Osaka, 3-3-1, Umeda, Kita-Ku, Osaka 530-0001, Japan
*
Author to whom correspondence should be addressed.
Universe 2024, 10(2), 68; https://doi.org/10.3390/universe10020068
Submission received: 8 December 2023 / Revised: 22 January 2024 / Accepted: 22 January 2024 / Published: 1 February 2024
(This article belongs to the Section High Energy Nuclear and Particle Physics)

Abstract

:
We find that a big gap between indicators for the breaking strengths of the global chiral SU(2) and U(1) axial symmetries in the QCD of the standard model (SM) can be interpreted as a new fine-tuning problem. This may thus imply calling for a class beyond the SM, which turns out to favor having a new chiral symmetry, and the associated massless new quark is insensitive to the chiral SU(2) symmetry for the lightest up and down quarks so that the fine-tuning is relaxed. Our statistical estimate shows that QCD of the SM is by more than 300 standard deviations off the parameter space free from fine-tuning, and the significance will be greater as the lattice measurements on the QCD hadron observables become more accurate. We briefly address a dark QCD model with massless new quarks as one viable candidate.

1. Introduction

Several types of fine-tuning problems have so far been pointed out that cannot be resolved by the SM alone. All of these involve an unsatisfactory big cancellation, e.g., the gauge hierarchy problem [1,2,3,4] and the strong CP problem [5,6,7,8,9]. The associated fine-tuned small observables have been confirmed: the size of the Higgs mass is much smaller than the Planck scale and the yet-unobserved electromagnetic dipole moment of neutron, respectively.
Fine-tunings can generically be related to existence of a hidden new symmetry that relaxes the big cancellation so that the fine-tuning becomes absent in the symmetric limit, which includes the ‘t Hooft naturalness argument [10]1. Taking this seriously into account has so far motivated people to refine or go beyond standard theories with such a new hidden symmetry and opened numerous frontiers in research directions along the lines of theoretical particle and cosmological physics.
In this paper, we argue that a big gap between indicators for the breaking strengths of the global chiral S U ( 2 ) and U ( 1 ) axial symmetries in the QCD of the SM can be interpreted as a new fine-tuning problem.
QCD has been well explored and confirmed, but actually, we know less precisely how low-energy QCD and the vacuum depend on quark flavors (described like the Columbia plot); in particular, there is little understanding of how the relatively heavy strange quark contributes there. This important open issue is thought of as an analogy to the top quark contribution to the electroweak-symmetry broken vacuum in the Higgs potential of the SM, called the electroweak-vacuum stability problem. What the present work focuses on is such a still nontrivial quark–flavor structure of the QCD vacuum, in particular, the essential gap between the breaking strengths of the chiral S U ( 2 ) symmetry for the light up and down quarks and U ( 1 ) axial symmetry.
The proposed new fine-tuning issue arises in an anomalous chiral Ward–Takahashi identity, which dictates the chiral S U ( 2 ) breaking strength as the subtraction of the U ( 1 ) axial breaking and the rate of the fluctuation of the QCD topological charge (Equation (1)). There, a drastic cancellation between two independent infrared singularities is observed, at around the QCD scale, which is responsible for the existence of the soft pions and required to yield the finite quark condensate. This thus potentially causes fine-tuning and yields a gigantic gap between the chiral S U ( 2 ) and U ( 1 ) axial breaking strengths.
We find that the gap is big enough to require fine-tuning, which corresponds to more than 300 standard deviations for the QCD in the SM away from the parameter space free from fine-tuning. The prospected statistical significance is subject to the higher accuracy of the lattice QCD measurements on the hadronic observables. Thus, the fine-tuned big gap between the chiral S U ( 2 ) and U ( 1 ) axial breaking strengths can be interpreted as a new indication from QCD calling for beyond the SM.
As one viable candidate, a dark QCD model with massless new quarks is briefly addressed.

2. The Posed Fine-Tuning Problem and Hidden Symmetry

We begin by introducing a key equation, the anomalous-chiral Ward identity, showing a relation between the indicators for the breaking strengths of chiral S U ( 2 ) L × S U ( 2 ) R symmetry ( χ chiral ) and U ( 1 ) A axial symmetry ( χ axial ), together with the topological susceptibility ( χ top ). In light of the QCD of the SM, we consider massive three-flavor QCD only including the lightest up, down, and strange quarks relevant below the QCD scale of O ( 1 ) GeV. We perform the three-flavor chiral S U ( 3 ) rotations to derive the anomalous Ward identities. Detailed definitions of susceptibilities, as well as the explicit derivation for the Ward identities, are supplied in Appendix A. Combining relevant identities, we thus find
χ chiral = χ axial + 4 m l 2 χ top ,
where m l denotes the mass for the up and down quarks. χ chiral is given as the difference of the (volume-averaged) propagators of mesons forming the chiral partner, which becomes zero when the partner masses are identical. χ axial is the similar indicator for the breaking strength of the U ( 1 ) axial symmetry. χ top measures the rate of fluctuation of the topological charge2.
In the case with small-enough m l and finite strange quark mass m s ( m s m l 0 ), as in the QCD of the SM at the physical point, the topological susceptibility χ top can approximately be evaluated as (see also Appendix A)3
χ top | m l m s ( u ¯ u m l + d ¯ d m l + s ¯ s m s ) m ¯ 2 ,
where m ¯ = ( 2 / m l + 1 / m s ) 1 . Noting that u ¯ u , d ¯ d , and s ¯ s remain nonzero even when m l = 0 , because of the dynamical generation of quark condensates in QCD at the scale of O ( 1 ) GeV, we find that for small m l m s , the u ¯ u and d ¯ d terms are dominant in Equation (2) so that the χ top term in Equation (1) is well approximated as
χ top m l 2 | m l m s u ¯ u m l + d ¯ d m l m ¯ 2 m l 2 [ O ( 1 ) GeV ] 3 4 m l ,
with the minus sign of the quark-condensate value taken into account. Thus, the size of the χ top term becomes larger than [ O ( 1 ) GeV ] 2 (with minus sign). Note also that χ chiral > 0 , χ axial > 0 , and χ chiral < χ axial due to the measured meson spectroscopy (for more details, see Appendix A). Therefore, in the case with small m l and finite m s , we meet a big destructive cancellation in Equation (1) between χ axial and the χ top term, both of which are on the order bigger than [ O ( 1 ) GeV ] 2 , to have a highly suppressed χ chiral : Equation (1) looks like [ O ( 10 GeV ) ] χ axial 2 [ O ( 10 GeV ) ] χ top 2 = χ chiral [ O ( 1 GeV ) ] 2 for m l MeV. This can be interpreted as a fine-tuning, unless some symmetry is present to explain the extraordinarily small χ chiral , which can relax the big subtraction, as elaborated in the Introduction. Note, however, that even the conventional chiral S U ( 2 ) symmetry ( m l 0 ) makes the accidental big cancellation more serious. As it will turn out later, the QCD of the SM with light up and down quarks and relatively heavier strange quarks actually suffers from this kind of big subtraction.
The original form of the anomalous Ward identity Equation (1), as given in Appendix A, is constructed from several susceptibilities. The anomalous Ward identity Equation (1) is derived by making a couple of the chiral S U ( 3 ) transformations, corresponding to the S U ( 3 ) adjoint indices a = 1 , 2 , 3 , 8 , on the vacuum expectation values of the pseudoscalar operators with a = 0 , 8 (see also Appendix E), which gives some correspondence between the vacuum expectation values of scalar and pseudoscalar operators, leading to the relationship among susceptibilities. Once χ top is identified in terms of a set of the susceptibilities, other terms unambiguously correspond to χ chiral and χ axial , which can be independently observed in the lattice simulation. Therefore, a big subtraction between χ axial and χ top to yield a small χ chiral is physical. In fact, such big destructive cancellation has been observed even at high temperatures [11].
The existence of the fine-tuning is due to the accidental cancellation between two individual infrared singularities responsible for the soft pions in QCD, as discussed in Ref. [12], χ axial 1 / m π 2 1 / m l , χ chiral constant, and χ top / m l 2 1 / m l (see also Equation (3)) for m l m s and m l 0 ; hence, in this limit, Equation (1) looks like finite = . This observation may imply that the chiral limit, on the basis of which the QCD can be expanded in the way of the chiral perturbation and hence widely accepted and well established, is faced with an accidental fine-tuning. Thus, the proposed fine-tuning is completely separated from the already existing fine-tuning, e.g., on the tiny mass difference between proton and neutron.
Going away from the QCD of the SM, we consider a counter limit where m s 0 , keeping m l finite. In Equation (1), the χ top term then vanishes as m s 0 (see also Equation (2)), reflecting the flavor-singlet nature [12], so that the indicators for the breaking strengths of the chiral and axial symmetries become identically equal each other:
( χ chiral χ axial ) = 4 χ top m l 2 0 , as m s 0 .
In this case, the χ top term ( χ top / m l 2 ) is adjusted to zero by a big destructive subtraction, i.e., a fine-tuning between χ chiral and χ axial . However, this fine-tuning can be gone in the limit m s 0 , which makes ( χ top / m l 2 ) become zero, in contrast to the QCD of the SM argued above, though the case with m l m s 0 , where χ chiral χ axial , is unrealistic.
Note that the strange quark currently acts as a spectator for the chiral S U ( 2 ) symmetry, being a singlet. Hence, the introduction of a new massless quark, protected by its own chiral symmetry, i.e., hidden new symmetry, should play the same role as the strange quark to solve the fine-tuning problem, keeping the massive-enough strange quark in accordance with the observation. We will later introduce an explicit and phenomenologically viable model having massless new quarks ( χ ) with a new chiral symmetry, which makes the real-life QCD free from fine-tuning:
( χ chiral χ axial ) 0 , as m χ 0 ,
with m l and m s at the physical point.

3. Quantifying the Fine-Tuning

We define the ratio [12]
R χ chiral χ axial ,
which also reads R = 1 4 χ top / m l 2 χ axial = 1 1 + 4 χ top / m l 2 χ chiral via the Ward identity in Equation (1). Thus, the deviation from R = 1 dictates a fine-tuning, and hence R serves as the estimator of the fine-tuning.
To compute the estimator R, one needs to work on the QCD in the deep-infrared region, which is highly nonperturbative because of the strong coupling nature in the low-energy scale. The best method to compute such nonperturbative dynamics is the numerical simulations of QCD on the lattice. However, the lattice simulations have never measured the susceptibilities at vacuum with varying m s 4.
Instead of the lattice simulation, in the spirit of Weinberg [15], we can invoke effective models of low-energy QCD, which realize the same breaking structure of the chiral and axial symmetries, and so forth, as that in low-energy QCD. In this paper, as the low-energy QCD description, we thus adapt a class of the Nambu–Jona–Lasinio (NJL) model made of only quarks with several quarkonic interactions. The NJL model has extensively been utilized in the field of hadron physics and so far provided us with lots of qualitative interpretations for low-energy QCD features, associated with chiral and axial symmetry breaking, together with successful phenomenological predictions [16].
We evaluate R as a function of m s based on a best-fit NJL model, which exhibits good fitness with lattice data on 2 + 1 flavor-QCD at the physical point, such as the observed meson masses. Details are provided in Appendix B, Appendix C and Appendix D. Since the NJL model currently does not incorporate the isospin breaking as well as radiative electromagnetic and weak interactions, it would not be suitable to input experimental values of QCD observables that implicitly include all those corrections. We therefore have used as inputs observables in lattice QCD with 2 + 1 flavors in the isospin-symmetric limit at the physical point available from the literature [17,18], which are exclusive for the gauge interactions external to QCD, and applied the least- χ 2 test to fix the parameters by using five representative observables5.
We thus compute the estimator R at the best-fit point, including the errors associated with the lattice data, and find
R = 0.0469 ± 0.0028 .
This clarifies that the QCD at the physical point is by about 340 standard deviation off the theory free from fine-tuning with R = 1 ! This is due to a too-large m s , as noted above.
We may take into account a possible theoretical uncertainty of about 30%, which could arise from the leading order approximation in the 1 / N c expansion, on which the present NJL model prediction is based. Currently disregarded corrections, associated with the isospin breaking, electromagnetic, and electroweak interactions, would also be small enough to be covered by the 30% uncertainty. Therefore, the estimated value of R in Equation (7) with the theoretical uncertainty of 30% would be the one corresponding to the prediction of the SM. Combining this 30% (“theor.”) with the error in Equation (7) associated with the uncertainties of inputs from the lattice data (“lat.”), we would then have R = 0.0469 ± ( 0.0028 ) lat . ± ( 0.0141 ) theor . . It is still about 66 standard deviations.
To make this disfavor visualized, varying the value of m s and m l with other model parameters fixed at the best-fit values, we plot contours of the estimator R on the ( m l , m s ) plane, which is displayed in Figure 1.
The value of R tends to saturate to be ≃0.02, even in the massive two-flavor limit with m s and m l = 5.75 MeV. This trend is exactly what we have suspected from the m s scaling of the χ top term in Equation (3). With m s fixed, say, to the physical point, R tends to get close to 1 as m l becomes larger and actually reaches 1 before the decoupling limit of the up and down quarks, as clarified in Ref. [12].
The deviation in Equation (7) can be interpreted as an indication of a violation of a new symmetry, as conjectured in Equation (5), which the SM does not possess. The significance of this new symmetry is subject to the accuracy in the current lattice simulation reflected in the size of the error of R, which is as low as 10%. This significance may therefore be compared to the significance for the discovery of the small isospin breaking in the W and Z boson masses observed at UA1 and UA2 experiments [21,22,23,24], which was the same 10% level in accuracy at the final stage of the discovery era (with data taking til 1985) [25,26]. The estimator R can be defined also for the W and Z masses as R W Z m W m Z such that the mass difference is written as Δ m W Z = m Z m W = m Z ( 1 R W Z ) , which actually forms a big cancellation structure and hence could be thought of as a fine-tuning in the same way as in χ axial and χ top / m l 2 with R in Equation (6). The final UA2 result reads [26] R WZ 1982 1985 = 0.876 ± 0.026 , so it is about a 4.8 σ deviation from the isospin-symmetric limit R WZ = 1 . This is, however, of course, trivial and can be explained by the isospin breaking (related to the so-called custodial symmetry) in the SM due to the hypercharge gauge interaction and the presence of isospin breaking in the quark masses. Compared to this, the new indication from QCD in Equation (7) is by about one order of magnitude more significant.
Current precision measurements on m W and m Z give R W Z 2022 = 0.88147 ± 0.00013 [27], which corresponds to 890 σ deviations from the isospin-symmetric limit. Similarly, the error of R in Equation (7) is also expected to become smaller as the precision in the lattice simulations becomes higher in the future; hence, the significance of the violation of a new symmetry will be enlarged to be as big as the current one for the isospin breaking.
This prospected significance might also become comparable with the current significance of the isospin breaking in the proton and neutron mass difference with R p n m p m n = 1 Δ m p n m n = 1 1 Δ m p n m p , which is read as R p n 2022 = 0.998623477 ( 316 ) [27], leading to 4 × 10 4 σ .
The gauge hierarchy problem caused by the big destructive loop correction to the Higgs mass m h yields R H = m h 2 ( m h ) m h 2 ( M p ) 1 3 8 π 2 M p 2 m h 2 m h 2 ( M p ) = 1 1 + 3 8 π 2 M p 2 m h 2 m h 2 , where M p denotes the Planck scale 10 18 GeV, m h ( m h ) 125 GeV, and we have simply taken into account only the top loop with the top Yukawa coupling 1 . This R H is estimated to be ∼ 6.3 × 10 32 . A similar estimate can be made also for the strong CP problem, which would yield R θ = θ QCD θ EW < 10 10 , where θ QCD and θ EW , respectively, denote the QCD and electroweak origins for the CP phase of the quark mass term. Thus, the statistical significance of R H will remain the biggest unless the accuracy in measuring the Higgs mass and top Yukawa coupling becomes better than the level of 10 32 , the size of R H .
The fine-tuning problem that we presently address is nothing sensory, but is essentially related to the existence of a hidden symmetry which makes R = 1 or χ top / m l 2 = 0 or χ chiral = χ axial in terms of our estimator. This is in contrast to the conventional argument: Something delicately fine or not is controlled merely by whether the tuning, to some extent, is necessary to make a big cancellation. The quantification of the present fine-tuning is unambiguously made on the basis of the statistical significance, and the standard deviations are subject to the accuracy in the measurement of Rs, which should therefore be compared with those having the same level of accuracy, as seen above.
Note that the strong CP problem in the SM is trivially solved in the limit m l 0 , whereas the new fine-tuning problem becomes more serious ( R 0 ). This discrepancy in the two problems can be understood via the Ward identity Equation (1), where χ top itself can be sent to zero when m l 0 , which is the massless up quark solution to the strong CP problem [5,28]; however, χ top / m l 2 then blows up, leaving the new fine-tuning problem. Thus, the two problems are generically separated within the framework of the SM. This fact also proves that the chiral S U ( 2 ) symmetry (with m l 0 ) does not make QCD free from fine-tuning, in sharp contrast to the naïve folklore. We will discuss a more definite discrimination from the strong CP problem in a later section (Section 5).
We could start with the definition of R, R = χ chiral / χ axial , instead of the Ward identity Equation (1), and discuss the difference Δ axial chiral = χ axial χ chiral , so that R = 1 Δ axial chiral / χ axial . Then, the form of Equation (1) is unambiguously fixed as it stands, telling us that the symmetric limit R = 1 is realized when χ top / m l 2 0 , which cannot be made when m l 0 (because χ top m l for small m l ), but can be achieved when m s 0 , which is based on the symmetry argument. This alternative view would also help readers to more definitely see that the m l = 0 limit separates the new fine-tuning problem from the strong CP problem, where R 0 and R θ = θ QCD / θ EW 1 , respectively.

4. A Solution: New Quarks with Dark QCD Colors

The proposed new fine-tuning problem is present at the scale only around the order of 1 GeV. When the electroweak symmetry becomes manifest at higher scales, the fine-tuning problem will be obscure. This is because the global chiral S U ( 2 ) and U ( 1 ) axial symmetries are explicitly broken by the electroweak interactions as well as quark masses (or Yukawa couplings between the Higgs and quarks). In that case, the key Equation (1) will be modified involving the electroweak “topological” susceptibilities. Also at scales m π , the fine-tuning will be nontransparent due to the decoupling of pions, which is the most dominant source to generate the big gap between χ chiral and χ axial . Thus, the new fine-tuning problem needs to be solved by a new physics with the scale Λ NEW in a range of m π Λ NEW 1 GeV.
The hint for this avenue is seen in Equation (4), which indicates introducing massless new quarks. In fact, the topological susceptibility χ top goes to zero when a massless new quark couples to the other three quarks due to the flavor-singlet nature so that Equation (5) is realized. The detailed proof is given in Appendix E. This motivates one to consider an explicit model beyond the SM.
We consider a new chiral quark ( χ ) to be neutral under the electroweak charges, which, instead, carries a dark color of S U ( N d ) group under the fundamental representation. The group representation table for the χ quark thus goes like χ L , R ( N d , 3 , 1 ) 0 for S U ( N d ) × S U ( 3 ) c × S U ( 2 ) W × U ( 1 ) Y , where the latter three symmetries correspond to the SM ones (QCD color, weak, and hypercharge). The dark color symmetry, as well as the electroweak neutrality, forbids creating extra light hadrons composed of the ordinary light quarks and the χ quark, such as u ¯ χ and u u χ . For simplicity, we also assume that the dark QCD coupling g d becomes strong almost at the same scale as the ordinary QCD coupling does, ( Λ d Λ QCD = O ( 1 ) GeV ) , namely, g d g s .
Below the scale ∼ 1 GeV, the dark QCD dynamically breaks the dark chiral U ( 3 ) L × U ( 3 ) R symmetry for the χ quark, down to the vectorial part, where the extra factor of 3 in the number of flavors comes from the QCD color multiplicity. Only the hadrons singlet under both the dark and ordinary QCD colors survive in the vacuum. Then, there emerges only one composite Nambu–Goldstone boson, η d , which becomes pseudo due to the axial anomaly in the dark QCD sector and acquires the mass of O ( 1 ) GeV. Moreover, at almost the same scale, the ordinary QCD breaks the approximate chiral S U ( 3 + N d ) L × S U ( 3 + N d ) R symmetry involving the χ quark down to the vectorial one, where, again, only the color singlets are relevant. The spontaneous breaking of this extended chiral symmetry does not yield excessive meson spectra made of ordinary quarks because of the double color symmetries, as aforementioned. Thus, the new low-lying spectra consist only of the dark sector: η d χ ¯ i γ 5 χ and its dark chiral partner σ d χ ¯ χ , as well as spin-1 dark mesons ( χ ¯ γ μ χ and χ ¯ γ μ γ 5 χ ) and dark baryons ( χ χ χ χ ). All those low-lying dark hadrons have a mass on the order of 1 GeV by feeding the chiral breaking contributions from both the ordinary QCD and dark QCD sectors.
The most stringent phenomenological constraint on this type of model comes from the extra massless-quark contribution to the running evolution of the ordinary QCD coupling α s , where, in the present case, N d species of new quarks in the fundamental representation of S U ( 3 ) c group come into play. Collider experiments have confirmed the asymptotic freedom with high accuracy in a wide range of higher energy scales, in particular above 10 GeV, over 1 TeV [27]. When α s is evolved up to higher scales using α s ( M Z ) measured at the Z boson pole as input, the tail of the asymptotic freedom around O ( 1 ) TeV can thus have sensitivity to exclude new quarks. Current data on α s at the scale around O ( 1 ) TeV involve large theoretical uncertainties. This results in uncertainty of determination of α s ( M Z ) for various experiments (LHC-ATLAS, -CMS, Tevatron-CDF, D0, etc.), which yields α s ( M Z ) 0.110 0.130 , consistent with the world average α s ( M Z ) 0.118 within the uncertainties [27]. We have worked on the two-loop perturbative computation of α s 6, and find that as long as N d is moderately large ( N d 5 ) , the measured ultraviolet scaling (for the renormalization scale of μ = 10 GeV, a few TeV) can be consistent with the current data [27] within the range of α s ( M z ) above.
Precise measurements in lower scales 10 GeV have not well been explored so far due to the deep-infrared complexity of QCD. The low-energy running of α s is indeed still uncertain and can be variant as discussed in a recent review, e.g., [29]. The present dark QCD could dramatically alter the infrared running feature of α s , due to new quarks and the running of the dark QCD coupling α d . This will also supply a decisive answer to the possibility of the infrared-near conformality of the real-life QCD.
Thus, a few massless new quarks can still survive constraints on α s at the current status7. More precise measurements of α s in the future will clarify how many light or massless new quarks can be hidden in QCD, which will fix the value of N d in terms of the present dark QCD.
Regarding other phenomenological aspects, we have checked that the present dark QCD model can survive current constraints, such as the measurement of the ordinary QCD hadron physics, dark meson couplings to diphoton, and limits on the dark baryon as dark matter candidates, and so forth. Furthermore, the dark meson and dark baryon yield several smoking guns to be probed in the near future, respectively, through the triphoton ( 3 γ ) signal at the mass around 1 GeV, at the Belle II experiment with 20–50 fb 1 data [31], and a planned dark matter detection experiment aiming at the sub-GeV mass range [32] with the spin-independent dark matter–nucleon cross-section ∼ 10 43 cm 2 8.

5. Conclusions

Toward a deeper understanding of the flavor dependence of the QCD vacuum, in the present work, we have focused on a gap between the breaking strengths of the chiral S U ( 2 ) symmetry for light up and down quarks and U ( 1 ) axial symmetry. It might be too premature to conclude that this gap is related merely to the mystery of the quark-generation structure in the SM. We have found an alternative interpretation based on the symmetry argument: the gap can be relaxed by a chiral symmetry for the strange quark, and the role of the massless/light strange quark can be replaced by a new massless/light quark (called χ ).
The existence of the symmetric limit m s 0 , where the “chiral S U ( 2 ) = U ( 1 ) axial” is realized as above, is manifest in the flavor dependence of the QCD vacuum, and we have shown how the gap, i.e., the fine-tuning, is seriously large on the basis of the statistical analysis along with comparison with the existing fine-tuned quantities.
Thus, the QCD of the SM may be yet incomplete if the fine-tuning is considered to be serious and hence may call for more quarks to keep the equivalence of the strengths of the chiral S U ( 2 ) and U ( 1 ) axial breaking. Given these new criteria, new quarks need protecting to be (nearly) massless by a new chiral symmetry and can be introduced in QCD consistently with existing experiments; that is what we have demonstrated in the present work. This symmetry is independent of the existing chiral or isospin symmetry, which ensures the smallness of masses of light quarks and has so far played the role to make QCD free from fine-tuning, e.g., for the small proton–neutron mass difference compared to individual proton and neutron masses.
The new fine-tuning problem is triggered due to the large strange quark mass, and brings a big gap between the chiral and axial order parameters, detected as a small size of the estimator R in Equation (6). This trend actually persists even in a whole temperature, as can be understood by tracing the analysis in the recent literature [12]. The present work is sort of an extension from the reference, and what is more, it essentially paves the way to a new criterion on the necessity of beyond the SM, which is completely out of the scope of the literature. Still, as noted also in [12], the characteristic small R can be checked on the lattice QCD simulations in the future, which will be an indirect observation of the proposed new indication.
In this paper, a benchmark of the fine-tuned QCD was placed based on the well-known low-energy effective model, NJL. This is a pioneering step and should be motivated and confirmed by various approaches in the future, such as the lattice calculation and functional-renormalization group analysis. One can also work on the chiral perturbation theory to evaluate R. No work has so far been conducted properly taking into account the flavor-singlet condition for χ top ; hence, it would also be an interesting issue to be left in the future.
Relaxing the new fine-tuning is tied with the vanishing curvature of the QCD vacuum at around the QCD scale: χ top 0 . The new fine-tuning problem is present irrespective of the place of the QCD vacuum, i.e., the value of the (net) QCD θ ¯ parameter: even a shifted QCD vacuum with θ ¯ gone, say by assumption of a QCD axion, keeps nonzero χ top as the developed axion potential energy, including the axion mass, which takes precisely the same flavor-singlet form as in Equation (2). Thus, the new fine-tuning problem is definitely separated from the strong CP problem.
In general, the dark QCD solution instead implies a nontrivial relation between θ ¯ and θ d , the theta parameter in dark QCD, to realize χ top = 0 in the presence of the massless new χ quark: the anomalous axial rotation of the massless χ leaves the θ ¯ -dependence into the dark QCD topological sector, ( θ d + ( N c / N d ) θ ¯ ) Q top d , where Q top d denotes the dark QCD topological charge. Thereby, one obtains χ top = ( N c / N d ) 2 χ top d 0 , with χ top d being the topological susceptibility in dark QCD, unless θ d = ( N c / N d ) θ ¯ . This nontrivial relation is required no matter what size θ ¯ is, i.e., which is independent of the strong CP problem.
The required relation might be trivial when QCD itself relaxes θ ¯ to 0 at a deep-infrared fixed point to be consistent with realization of the confinement, as recently discussed in lattice QCD [34,35,36]. This self-relaxation is applicable also to dark QCD; hence, in that case, one has θ ¯ = θ d = 0 , and χ top = 0 in the presence of the massless χ quark. In this sense, the proposed fine-tuning might be linked also with a deeper understanding of the color confinement.
In contrast, solving the new fine-tuning problem disfavors QCD axion models as the solution of the strong CP: first of all, the QCD axion needs not to be present until the QCD pions are decoupled from R, otherwise the axion potential energy necessarily yields nonzero χ top even along with massless new quarks. In this sense, a composite axion [37,38] with the dynamical/composite scale scaled down to the QCD scale (or lower) might be the candidate, where the composite scale is set to m π 4 π f a with the axion decay constant f a . However, such a low-scale QCD axion model with both a QCD axion and its small decay constant on the QCD scale has already been ruled out by the LEP search for Z π γ [39] due to too-large axion coupling to diphoton. Thus, the QCD axion is incompatible with the solution for the new fine-tuning problem.
Other solutions alternative to the presently addressed dark QCD are worth investigating. In closing, we provide one general “recipe”. First of all, try to introduce new Dirac massless quarks, which act as a spectator of the global chiral S U ( 2 ) symmetry for the up and down quarks. They are generically allowed to feel the electroweak charge, whichever way it is ganged vectorlikely or chirally. The former case would be phenomenologically viable in light of the electroweak precision tests, and the limit on the number of quark generations placed from the Z boson decays. Those new quarks would be preferable not to form the Yukawa interaction with ordinary quarks and Higgs fields (doublets and triplets, and so on), which develop the vacuum expectation values at the weak scale, and yield the mass for the new quarks. If new quarks could be coupled to such Higgses, solving the strong CP problem (without introducing an axion), as well as keeping the light enough new quarks down until the QCD scale, would be hard and challenging.
Given this recipe, one might think that though it would sound somewhat ad hoc, the presumably most minimal setup would be to introduce an electroweak-singlet quark with a negative charge under a new parity while assigning a positive charge for ordinary quarks, so as to avoid spoiling the successful light hadron spectroscopy. In the dark QCD model introduced in the present paper, the role of such a new parity has been played by the dark QCD color charge. Such alternatives are to be addressed in detail elsewhere.

Author Contributions

Conceptualization and methodology, H.I., M.K., S.M. and A.T.; formal analysis, C.-X.C., J.-Y.L. and S.M.; writing—review and editing, C.-X.C. and S.M. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported in part by the National Science Foundation of China (NSFC) under Grant No. 11975108, 12047569, and 12147217, and the Seeds Funding of Jilin University (S.M.), the RIKEN Special Postdoctoral Researcher program and partially by JSPS KAKENHI Grant Number JP20K14479 (A.T.), and the Fundamental Research Funds for the Central Universities (M.K.).

Data Availability Statement

No new data were created or anlyzed in this study. Date sharing is not applicable to this article.

Acknowledgments

We thank Ryosuke Sato for useful comments and discussions.

Conflicts of Interest

The authors declare no conflicts of interest.

Appendix A. Chiral Ward Identities and Topological Susceptibility: Road to Equation (1) in the Main Text

A set of generic anomalous Ward identities for the three-flavor chiral S U ( 3 ) L × S U ( 3 ) R symmetry in QCD is derived by the corresponding chiral variations for the generating functional of QCD [12,40,41] (see also Appendix E):
u ¯ u + d ¯ d = m l χ π , u ¯ u + d ¯ d + 4 s ¯ s = [ m l χ P u u + χ P d d + 2 χ P u d 2 ( m s + m l ) χ P u s + χ P d s + 4 m s χ P s s ] , u ¯ u + d ¯ d 2 s ¯ s = [ m l ( χ P u u + χ P d d + 2 χ P u d ) + ( m l 2 m s ) ( χ P u s + χ P d s ) 2 m s χ P s s ] ,
where the isospin-symmetric mass m u = m d m l has been taken for up and down quarks; χ p u u , χ P d d , χ P u d , χ P s s , χ P u s , and χ P d s are the pseudoscalar susceptibilities; and χ π is the pion susceptibility, defined as
χ P f 1 f 2 = d 4 x ( q ¯ f 1 ( 0 ) i γ 5 q f 1 ( 0 ) ) ( q ¯ f 2 ( x ) i γ 5 q f 2 ( x ) ) , for q f 1 , 2 = u , d , s , χ π = d 4 x [ ( u ¯ ( 0 ) i γ 5 u ( 0 ) ) ( u ¯ ( x ) i γ 5 u ( x ) ) conn + ( d ¯ ( 0 ) i γ 5 d ( 0 ) ) ( d ¯ ( x ) i γ 5 d ( x ) ) conn ] ,
with · · · conn being the connected part of the correlation function.
The topological susceptibility χ top is related to the θ vacuum configuration of the QCD. It is defined as the curvature of the θ -dependent vacuum energy V ( θ ) in the QCD at θ = 0 :
χ top = d 4 x δ 2 V ( θ ) δ θ ( x ) δ θ ( 0 ) | θ = 0 .
Performing the U ( 1 ) A rotation for quark fields together with the flavor-singlet condition [42,43], one can transfer the θ dependence coupled to the topological gluon configurations, via the axial anomaly, into current quark mass terms. Thus, χ top is as follows [41]:
χ top = m ¯ 2 [ u ¯ u m l + d ¯ d m l + s ¯ s m s + χ P u u + χ P d d + χ P s s + 2 χ P u d + 2 χ P u s + 2 χ P d s ] = 1 4 m l u ¯ u + d ¯ d + m l 2 χ P u u + χ P d d + 2 χ P u d = m s s ¯ s + m s 2 χ P s s ,
where m ¯ 1 m u + 1 m d + 1 m s 1 . χ top 0 , when either of quarks becomes massless ( m l or m s 0 ), reflecting the flavor-singlet nature of the QCD vacuum.
By combining the Ward identities in Equation (A1), χ top in Equation (A3) is expressed as
χ top = 1 2 m l m s χ P u s + χ P d s = 1 4 m l 2 ( χ η χ π ) ,
where χ η is the η meson susceptibility, which is defined as
χ η = d 4 x [ ( u ¯ ( 0 ) i γ 5 u ( 0 ) ) ( u ¯ ( x ) i γ 5 u ( x ) ) + ( d ¯ ( 0 ) i γ 5 d ( 0 ) ) ( d ¯ ( x ) i γ 5 d ( x ) ) + 2 ( u ¯ ( 0 ) i γ 5 u ( 0 ) ) ( d ¯ ( x ) i γ 5 d ( x ) ) ] = χ P u u + χ P d d + 2 χ P u d .
Equation (A5) can be written as
( χ η χ δ ) = ( χ π χ δ ) + 4 m l 2 χ top ,
where χ δ is the susceptibility for the δ meson channel (which is a 0 meson in terms of the Particle Data Group identification), defined in the same way as χ π in Equation (A1) with the factors of ( i γ 5 ) replaced with identity 1. ( χ η χ δ ) and ( χ π χ δ ) play the roles of the chiral and axial breaking indicators, respectively, which signal the restorations when those (asymptotically) reach zero.
Renaming ( χ η χ δ ) and ( χ π χ δ ) as
χ chiral χ η χ δ , χ axial χ π χ δ ,
one reaches Equation (1) in the main text.
In deriving Equation (1), one could choose another scalar susceptibility χ σ , which makes the chiral S U ( 2 ) and axial partners for χ π and χ η , respectively. (The definition of χ σ is the same as χ η with the ( i γ 5 ) factors replaced by the identity ( 1 ) .) Then, Equation (1) would be replaced by χ chiral = χ axial + 4 χ top m l 2 , where χ chiral χ σ χ π and χ axial χ σ χ η . Even if this alternative identification is taken, the present proposal of the new fine-tuning problem still holds: a big gap between χ axial and the χ chiral is relaxed by a symmetry sending m s 0 to make χ top / m l 2 vanishing. Thus, the presently proposed new fine-tuning problem is robust and unambiguous.

Appendix B. Nambu–Jona–Lasinio Model Description

The Nambu–Jona–Lasinio (NJL) model with three flavors, which is adapted in the main text, takes the form (for a review, see [16]):
L = q ¯ ( i γ μ μ M ) q + L 4 f + L KMT , L 4 f = G S 2 a = 0 8 [ ( q ¯ λ a q ) 2 + ( q ¯ i γ 5 λ a q ) 2 ] , L KMT = G D [ det i , j q i ¯ ( 1 + γ 5 ) q j + h . c . ] ,
where the quark field q is represented as the triplet of S U ( 3 ) group in the flavor space, q = ( u , d , s ) T , and λ a ( a = 0 , 1 , , 8 ) are the Gell–Mann matrices with λ 0 = 2 / 3 · 1 3 × 3 . The determinant in L KMT acts on the flavor indices, and M = diag { m l , m l , m s } .
L 4 f is the standard-scalar four-Fermion interaction term with the coupling strength G S . This is the most minimal interaction term involving the smallest number of quark fields for Lorentz scalar and pseudoscalar channels, which could be generated at low-energy QCD via the gluon exchange. The L 4 f is U ( 3 ) L × U ( 3 ) R invariant under the chiral transformation: q U · q with U = exp [ i γ 5 a = 0 8 ( λ a / 2 ) θ a ] and the chiral phases θ a . The mass term in L explicitly breaks U ( 3 ) L × U ( 3 ) R symmetry. The determinant term L KMT is called the Kobayashi–Maskawa–‘t Hooft [44,45,46,47,48] term, which is a six-point interaction induced from the QCD instanton configuration coupled to quarks, with the effective coupling constant G D . This interaction gives rise to the mixing between different flavors and also uplifts the η mass to be no longer a Nambu–Goldstone boson. The KMT term preserves S U ( 3 ) L × S U ( 3 ) R invariance (associated with the chiral phases labeled as a = 1 , , 8 ), but breaks the U ( 1 ) A (corresponding to a = 0 ) symmetry.
The approximate chiral S U ( 3 ) L × S U ( 3 ) R symmetry is spontaneously broken down to the vectorial symmetry S U ( 3 ) V , when the couplings G s and/or G D become strong enough, by nonperturbatively developing nonzero quark condensates q ¯ q 0 , to be consistent with the underlying QCD feature. The present NJL model monitors the spontaneous breakdown by the large N c expansion, where N c stands for the number of QCD colors.
The NJL model itself is a (perturbatively) nonrenormalizable field theory because L 4 f and L KMT describe the higher dimensional interactions with mass dimensions greater than four. Therefore, a momentum cutoff Λ needs to be introduced to make the NJL model regularized.

Appendix C. Scalar and Pseudoscalar Susceptibilities in the NJL Model

In this section, the explicit formulae for scalar and pseudoscalar susceptibilities derived from the NJL model Equation (A9) are listed. We will leave the details of the calculations here and refer readers to a review paper [16], which contains all necessary information to reach the final formulas that we will present below. In what follows, we use the notation for quark condensates as α = u ¯ u , β = d ¯ d , and γ = s ¯ s , and will work on computations in Euclidean momentum space.

Appendix C.1. Pseudoscalar Meson Channel

In the η η -coupled channel, the pseudoscalar meson susceptibility on the generator basis of U ( 3 ) is defined as
χ P i j = d 4 x ( i q ¯ ( x ) γ 5 λ i q ( x ) ) ( i q ¯ ( 0 ) γ 5 λ j q ( 0 ) ) ,
where i , j = 0 , 8 .
In the NJL model, after prescribing the resummation technique, this χ P i j takes the form [16]
χ P = 1 1 + G P Π P · Π P ,
where G P is the coupling strength matrix, and Π P is the polarization tensor evaluated at zero momentum transfer:
G P = G P 00 G P 08 G P 80 G P 88 = G S 2 3 ( α + β + γ ) G D 2 6 ( 2 γ α β ) G D 2 6 ( 2 γ α β ) G D G S 1 3 ( γ 2 α 2 β ) G D ,
Π P = Π P 00 Π P 08 Π P 80 Π P 88 = 2 3 ( 2 I P u u + I P s s ) 2 2 3 ( I P u u I P s s ) 2 2 3 ( I P u u I P s s ) 2 3 ( I P u u + 2 I P s s ) ,
with I P i i ( ω , p ) being the pesudoscalar one-loop polarization functions, and the relevant quantities:
I P i i = N c π 2 0 Λ d p p 2 1 E i , for i = u , d , s , E i = M i 2 + p 2 , M u = m l 2 G S α 2 G D β γ , M d = m l 2 G S β 2 G D α γ , M s = m s 2 G S γ 2 G D α β , q ¯ i q i = 2 N c Λ d p ( 2 π ) 3 M i E i .
By performing the basis transformation, the pseudoscalar susceptibilities in the flavor basis are obtained as
1 2 χ P u u + 1 2 χ P u d χ P u s χ P s s = 1 6 2 6 1 12 1 6 2 12 1 6 1 6 2 3 1 3 χ P 00 χ P 08 χ P 88 ,
where we have taken the isospin-symmetric limit into account, i.e., χ P u u = χ P d d and χ P u s = χ P d s . The χ η is then given as in Equation (A6).
In a way similar to Equation (A11), the present NJL model gives the explicit formula of χ π defined in Equation (A2) as
χ π = 1 1 + G π Π π · Π π ,
where G π = G S + G D γ , which is the coupling strength in the pion channel, and Π π is the quark-loop polarization function for χ π , which is evaluated by using I P i i in Equation (A14) as
Π π = I P u u + I P d d = 2 I P u u .

Appendix C.2. Scalar Meson Channel

The definitions of scalar susceptibilities are similar to those for pseudoscalars’, which are given just by removing i γ 5 in the definition of pseudoscalar susceptibilities, and supplying the appropriate one-loop polarization functions and the corresponding coupling constants.
In the same way as in the pseudoscalar susceptibilities in the 0–8 coupled channel in Equation (A10), the scalar susceptibility matrix χ S is evaluated in the present NJL on the generator basis as
χ S = 1 1 + G S Π S · Π S ,
where G S is the coupling strength matrix,
G S = G S 00 G S 08 G S 80 G S 88 = G S + 2 3 ( α + β + γ ) G D 2 6 ( 2 γ α β ) G D 2 6 ( 2 γ α β ) G D G S + 1 3 ( γ 2 α 2 β ) G D .
The coupling constant matrices in the scalar and pseudoscalar channels (Equations (A12) and (A19)) are different in sign in front of G D , reflecting attractive and repulsive interactions, respectively. The scalar polarization tensor matrix Π S in Equation (A18) is given by
Π S = Π S 00 Π S 08 Π S 80 Π S 88 = 2 3 ( 2 I S u u + I S s s ) 2 2 3 ( I S u u I S s s ) 2 2 3 ( I S u u I S s s ) 2 3 ( I S u u + 2 I S s s ) ,
I S i i = N c π 2 0 Λ p 2 d p E i p 2 M i 2 E i 3 , for i = u , d , s .
By moving on to the flavor base via the base transformation, the scalar susceptibilities are cast into the form:
1 2 χ S u u + 1 2 χ S u d χ S u s χ S s s = 1 6 2 6 1 12 1 6 2 12 1 6 1 6 2 3 1 3 χ S 00 χ S 08 χ S 88 ,
in which we have read χ S u u = χ S d d and χ S u s = χ S d s .
Similarly to χ π in Equation (A16), in the NJL model the explicit formula for χ δ reads
χ δ = Π δ 1 + G δ Π δ ,
where G δ = G S G D γ , which is the coupling strength in the δ channel, and Π δ = I S u u + I S d d = 2 I S u u is the corresponding quark-loop polarization function.

Appendix D. The Best-Fit NJL Model

Table A1. The result of the least χ 2 statistical test of the present NJL model derived by fitting to the lattice QCD data with 2 + 1 flavors in the isospin-symmetric limit at the physical point [17,18]. Details of the global fit of the NJL model including other available lattice data can be found in another publication.
Table A1. The result of the least χ 2 statistical test of the present NJL model derived by fitting to the lattice QCD data with 2 + 1 flavors in the isospin-symmetric limit at the physical point [17,18]. Details of the global fit of the NJL model including other available lattice data can be found in another publication.
ObservableLattice DataBest-Fit Value
f π (MeV) 92.07 ± 0.99  [17] 91.93 ± 1.35
m π (MeV) 138 ± 0.3  [17] 137.95 ± 1.95
m K (MeV) 494.2 ± 0.4  [17] 493.50 ± 8.09
m η (MeV) 554.7 ± 9.2  [18] 503.82 ± 15.37
m η (MeV) 930 ± 21  [18] 978.77 ± 61.59
Table A2. The best-fit values of the model parameters.
Table A2. The best-fit values of the model parameters.
Model ParameterBest-Fit Value
m l (MeV) 5.75 ± 0.05
m s (MeV) 130.69 ± 0.98
G S ( MeV 2 ) ( 14.34 ± 0.41 ) × 10 6
G D ( MeV 5 ) ( 18.17 ± 1.35 ) × 10 14
Λ (MeV) 569.4 ± 16.5
There are five model parameters that need to be fixed: the light quark mass m l , the strange quark mass m s , the coupling constants G S and G D , and the (three-) momentum cutoff Λ . Since the present NJL model does not incorporate the isospin breaking as well as radiative electromagnetic and weak interactions, it would not be suitable to input experimental values of QCD observables that implicitly include all those corrections. We thus use as inputs observables in lattice QCD with 2 + 1 flavors in the isospin-symmetric limit at the physical point available from the literature [17,18], which are exclusive for the gauge interactions external to QCD. We apply the least- χ 2 test to fix the parameters by using five representative observables as in Table A1. The resultant values of the best-fit model parameters are given in Table A2. The least χ 2 test shows good agreement with the lattice data within the 1 σ uncertainties. The best-fit NJL model predicts the susceptibilities relevant to R in Equation (6) of the main text as χ chiral = ( 2.2784 ± 0.0026 ) × 10 5 , χ axial = ( 4.8597 ± 0.0077 ) × 10 6 , and χ top / m l 2 = ( 1.158 ± 0.064 ) × 10 6 , in unit of MeV 2 .

Appendix E. Anomalous Chiral Ward Identities and Topological Susceptibility, with One Extra Quark

This section provides derivation of anomalous chiral Ward identities, including one extra quark in three flavors of QCD.
The anomalous chiral Ward identities are directly read off from chiral variations of the generating functional of the QCD action with N quark flavors. The central formula then takes the form
δ a O b ( 0 ) = i d 4 x O b ( 0 ) · Q ¯ ( x ) i γ 5 { T a , M } Q ( x ) ,
where Q denotes a quark field forming N-plet of S U ( N ) ; T a ( a = 1 , , N 2 1 ) are generators of S U ( N ) ; δ a stands for the infinitesimal variation of the chiral S U ( N ) transformation associated with the generator T a , under which Q transforms as δ a Q = i γ 5 T a Q ; O b ( 0 ) ( b = 0 , , N 2 1 ) is an arbitrary operator; M is the Q-quark mass matrix, taken to be diagonal, like M = diag { m 1 , m 2 , , m N } .
In particular, for the pseudoscalar operators O b = Q ¯ i γ 5 T b Q , taking the case with N = 3 , with Q = ( q l , s ) T = ( ( u , d ) , s ) T , m 1 = m 2 = m l , and m 3 = m s , and choosing a = 1 , 2 , 3 , 8 and b = 0 , 8 , we get the first and second identities in Equation (A1) in the main text.
Including a new quark field χ with mass m 4 = m χ into the Q-field to form the S U ( 4 ) -quartet: Q = ( q l , s , χ ) T , we work on the chiral S U ( 4 ) transformations. Among the S U ( 4 ) generators, we focus only on the subgroup part of S U ( 3 ) embedded as T a = 1 , , 8 = 1 2 λ a 0 1 × 3 0 3 × 1 0 , and the Cartan generator T a = 15 = 1 2 6 1 3 × 3 0 1 × 3 0 3 × 1 3 , together with the unit matrix T a = 0 = 1 2 2 · 1 4 × 4 . We thus find a couple of additional Ward identities, except for the first two in Equation (A1):
( a , b ) = ( 8 , 0 ) : u ¯ u + d ¯ d 2 s ¯ s = [ m l χ P u u + χ P d d + 2 χ P u d + ( m l 2 m s ) χ P u s + χ P d s 2 m s χ P s s + m χ χ P u χ + χ P d χ 2 m s χ P s χ ] ; ( a , b ) = ( 15 , 8 ) : u ¯ u + d ¯ d 2 s ¯ s = [ m l χ P u u + χ P d d + 2 χ P u d + ( m s 2 m l ) χ P u s + χ P d s 2 m s χ P s s 3 m χ χ P u χ + χ P d χ + 6 m χ χ P s χ ] ; ( a , b ) = ( 8 , 15 ) : u ¯ u + d ¯ d 2 s ¯ s = [ m l χ P u u + χ P d d + 2 χ P u d + ( m l 2 m s ) χ P u s + χ P d s 2 m s χ P s s 3 m l χ P u χ + χ P d χ + 6 m s χ P s χ ] ;
( a , b ) = ( 15 , 0 ) : u ¯ u + d ¯ d + s ¯ s 3 χ ¯ χ = [ m l χ P u u + χ P d d + 2 χ P u d + ( m l + m s ) χ P u s + χ P d s + m s χ P s s + ( m l 3 m χ ) χ P u χ + χ P d χ + ( m s 3 m χ ) χ P s χ 3 m χ χ P χ χ ] ; ( a , b ) = ( 15 , 15 ) : u ¯ u + d ¯ d + s ¯ s + 9 χ ¯ χ = [ m l χ P u u + χ P d d + 2 χ P u d + ( m l + m s ) χ P u s + χ P d s + m s χ P s s 3 ( m l + m χ ) χ P u χ + χ P d χ 3 ( m s + m χ ) χ P s χ + 9 m χ χ P χ χ ] ,
where pseudoscalar susceptibilities including the χ quark are defined in the same way as those for other quarks in Equation (A2). From these with Equation (A1), we have
χ P u χ + χ P d χ = m s m χ χ P u s + χ P d s , χ P s χ = m l 2 m χ χ P u s + χ P d s , χ ¯ χ = m χ χ P χ χ + m l m s 2 m χ χ P u s + χ P d s .
Using the first two relations in Equation (A26), we see that the first Ward identity in Equation (A25) becomes the same as the third one in Equation (A1).
Including the new χ quark, the topological susceptibility defined in Equation (A3) now takes the form
χ top = m ¯ 2 [ u ¯ u m l + d ¯ d m l + s ¯ s m s + χ ¯ χ m χ + χ P u u + χ P d d + 2 χ P u d + 2 χ P u s + χ P d s + χ P s s + 2 χ P u χ + χ P d χ + 2 χ P s χ + χ P χ χ ] ,
with m ¯ 1 = 2 m l + 1 m s + 1 m χ . Using the relations in Equation (A26) together with those in Equation (A1), we find
χ top = 1 2 m l m s χ P u s + χ P d s = 1 2 m l m χ χ P u χ + χ P d χ = m s m χ χ P s χ = 1 4 m l u ¯ u + d ¯ d + m l 2 χ P u u + χ P d d + 2 χ P u d = m s s ¯ s + m s 2 χ P s s = m χ χ ¯ χ + m χ 2 χ P χ χ .
In the case of dark QCD modeling as in the main text, actually, the dark QCD coupling to the χ quark explicitly breaks the chiral S U ( 4 ) L × S U ( 4 ) R symmetry, as well as the mass terms. However, this breaking effect does not modify the anomalous identities associated with the chiral transformations for a = 1 , 2 , 3 , 8 , 15 , because the dark QCD coupling to the χ quark only breaks the vectorial S U ( 4 ) flavor symmetry down to S U ( 3 ) × U ( 1 ) , which still keeps those chiral symmetries: T a , 0 3 × 3 1 [ T a , T b = 15 ] = 0 .
The argument in this section can straightforwardly be extended to the case with more extra quarks, like 3 + N d quarks.
The presence of the dark QCD theta parameter and the QCD’s one modifies χ top in Equation (A28) as χ top χ top + ( N c m ¯ / m χ ) χ top d with the dark QCD topological susceptibility χ top d , N c = 3 , and m χ in the original χ top as well as m ¯ replaced by m χ / N d .

Notes

1
One example is the tiny mass difference between the proton and neutron. Individual masses, being mainly fed by the isospin-symmetric dynamical quark mass, are of O ( 1 ) GeV on the typical QCD scale, but the mass difference is of O ( 10 3 ) GeV. This is regarded as a big subtraction and can be explained due to a small violation of the isospin symmetry for up and down quarks, including the current quark masses m u and m d arising from the Higgs via the electroweak-symmetry breaking and their electromagnetic charge difference. The mass difference, in fact, goes to zero in the symmetric limit.
2
Susceptibilities are not direct observables in terrestrial experiments, nor astrophysical observatories, in contrast to the existing fine-tuning problems as aforementioned. Those can rather be observed in the lattice QCD, though they would not have a correlation with definite phenomenological observables.
3
Throughout the present paper, we take the signs of quark condensates and quark masses to be negative and positive, respectively, so that χ top < 0 .
4
This is mainly because, firstly, it has not been well motivated, and moreover, costs of lattice calculations for small mass are proportional to 1 / m , where m is the lightest quark mass. Simulations for light strange quarks can be performed using the same technology as in [13], and employing similar calculations in [11,14] with light m s .
5
The best-fit NJL model predicts χ top = ( 0.025 ± 0.002 ) / fm 4 . For this χ top , comparison with the results from the lattice QCD simulations with 2 + 1 flavors is available, which are χ top = 0.019 ( 9 ) / fm 4 [19], and χ top = 0.0245 ( 24 ) stat ( 03 ) flow ( 12 ) cont / fm 4 [20]. Here, for the latter, the first error is statistical; the second one comes from the systematic error; and the third one arises due to changing the upper limit of the lattice spacing range in the fit. Although their central values do not agree with each other, we may conservatively say that the difference between them is interpreted as a systematic error from the individual lattice QCD calculation. Thus, the present NJL model is, in that sense, in good agreement with the lattice QCD results also on χ top .
6
The dark QCD running coupling ( α d ) contributes to the running of α s at the two-loop level. This contribution is, however, safely negligible: when α s α d at low-energy, because α d α s at high energy due to the smaller number of dark QCD quarks (with the net number 3 coming in the beta function of α d ) than that of the ordinary QCD quarks (with the net number 5 or 6 + N d in the beta function of α s ). Taking into account only the additional N d quark-loop contributions to the running of α s , we are thus allowed to evaluate the two-loop beta function at the leading order.
7
Other stringent bounds on the extra light quarks or colored scalars come from the ALEPH search for gluino and squark pairs tagged with the multijets at the Large Electron Positron (LEP) collider experiment [30]. However, this limit has no sensitivity below the mass 2 GeV and hence is not applicable to the present benchmark model.
8
More details on this beyond the SM have been presented in a separate paper [33].

References

  1. Shaposhnikov, M. Is there a new physics between electroweak and Planck scales? arXiv 2007, arXiv:0708.3550. [Google Scholar]
  2. Nielsen, H.B. PREdicted the Higgs Mass. Bled Workshops Phys. 2012, 13, 94–126. [Google Scholar]
  3. Masina, I. Higgs boson and top quark masses as tests of electroweak vacuum stability. Phys. Rev. D 2013, 87, 053001. [Google Scholar] [CrossRef]
  4. Degrassi, G.; Di Vita, S.; Elias-Miró, J.; Espinosa, J.R.; Giudice, G.F.; Isidori, G.; Strumia, A. Higgs mass and vacuum stability in the Standard Model at NNLO. J. High Energy Phys. 2012, 2012, 1–33. [Google Scholar] [CrossRef]
  5. Peccei, R.D.; Quinn, H.R. Constraints imposed by CP conservation in the presence of pseudoparticles. Phys. Rev. D 1977, 16, 1791–1797. [Google Scholar] [CrossRef]
  6. Weinberg, S. A New Light Boson? Phys. Rev. Lett. 1978, 40, 223–226. [Google Scholar] [CrossRef]
  7. Wilczek, F. Decays of Heavy Vector Mesons into Higgs Particles. Phys. Rev. Lett. 1977, 39, 1304–1306. [Google Scholar] [CrossRef]
  8. Shifman, M.; Vainshtein, A.; Zakharov, V. Can confinement ensure natural CP invariance of strong interactions? Nucl. Phys. B 1980, 166, 493–506. [Google Scholar] [CrossRef]
  9. Kim, J.E.; Carosi, G. Axions and the Strong CP Problem. Rev. Mod. Phys. 2010, 82, 557–602, Erratum in Rev. Mod. Phys. 2019, 91, 049902. [Google Scholar] [CrossRef]
  10. Hooft, G. Naturalness, chiral symmetry, and spontaneous chiral symmetry breaking. NATO Sci. Ser. B 1980, 59, 135–157. [Google Scholar] [CrossRef]
  11. Bhattacharya, T.; Buchoff, M.I.; Christ, N.H.; Ding, H.-T.; Gupta, R.; Jung, C.; Karsch, F.; Lin, Z.; Mawhinney, R.D.; McGlynn, G.; et al. QCD Phase Transition with Chiral Quarks and Physical Quark Masses. Phys. Rev. Lett. 2014, 113, 082001. [Google Scholar] [CrossRef]
  12. Cui, C.-X.; Li, J.-Y.; Matsuzaki, S.; Kawaguchi, M.; Tomiya, A. New interpretation of the chiral phase transition: Violation of the trilemma in QCD. Phys. Rev. D 2022, 105, 114031. [Google Scholar] [CrossRef]
  13. Dini, L.; Hegde, P.; Karsch, F.; Lahiri, A.; Schmidt, C.; Sharma, S. Chiral phase transition in three-flavor QCD from lattice QCD. Phys. Rev. D 2022, 105, 034510. [Google Scholar] [CrossRef]
  14. Bazavov, A.; Bhattacharya, T.; Buchoff, M.I.; Cheng, M.; Christ, N.H.; Ding, H.-T.; Gupta, R.; Hegde, P.; Jung, C.; Karsch, F.; et al. The chiral transition and U(1)A symmetry restoration from lattice QCD using domain wall fermions. Phys. Rev. D 2012, 86, 094503. [Google Scholar] [CrossRef]
  15. Weinberg, S. Phenomenological Lagrangians. Phys. A Stat. Mech. Its Appl. 1979, 96, 327–340. [Google Scholar] [CrossRef]
  16. Hatsuda, T.; Kunihiro, T. QCD phenomenology based on a chiral effective Lagrangian. Phys. Rep. 1994, 247, 221–367. [Google Scholar] [CrossRef]
  17. Aoki, S. et al. [FLAG Working Group] Review of lattice results concerning low-energy particle physics. Eur. Phys. J. C 2014, 74, 2890. [Google Scholar] [CrossRef]
  18. Bali, G.S. et al. [The RQCD Collaboration] Masses and decay constants of the η and η’ mesons from lattice QCD. J. High Energy Phys. 2021, 2021, 1–94. [Google Scholar] [CrossRef]
  19. Bonati, C.; D’elia, M.; Mariti, M.; Martinelli, G.; Mesiti, M.; Negro, F.; Sanfilippo, F.; Villadoro, G. Axion phenomenology and θ-dependence from Nf = 2 + 1 lattice QCD. J. High Energy Phys. 2016, 2016, 1–27. [Google Scholar] [CrossRef]
  20. Borsanyi, S.; Fodor, Z.; Guenther, J.; Kampert, K.-H.; Katz, S.D.; Kawanai, T.; Kovacs, T.G.; Mages, S.W.; Pasztor, A.; Pittler, F.; et al. Calculation of the axion mass based on high-temperature lattice quantum chromodynamics. Nature 2016, 539, 69–71. [Google Scholar] [CrossRef]
  21. Arnison, G. et al. [UA1 Collaboration] Experimental observation of isolated large transverse energy electrons with associated missing energy at s = 540 GeV. Phys. Lett. B 1983, 122, 103–116. [Google Scholar] [CrossRef]
  22. Banner, M. et al. [UA2 Collaboration] Observation of single isolated electrons of high transverse momentum in events with missing transverse energy at the CERN pp collider. Phys. Lett. B 1983, 122, 476–485. [Google Scholar] [CrossRef]
  23. Arnison, G. et al. [UA1 Collaboration] Experimental observation of lepton pairs of invariant mass around 95 GeV/c2 at the CERN SPS collider. Phys. Lett. B 1983, 126, 398–410. [Google Scholar] [CrossRef]
  24. Bagnaia, P. et al. [UA2 Collaboration] Evidence for Z0→e+e− at the CERN pp collider. Phys. Lett. B 1983, 129, 130–140. [Google Scholar] [CrossRef]
  25. Albajar, C. et al. [UA1 Collaboration] Studies of intermediate vector boson production and decay in UA1 at the CERN proton-antiproton collider. Eur. Phys. J. C 1989, 44, 15–61. [Google Scholar]
  26. Ansari, R. et al. [UA2 Collaboration] Measurement of the standard model parameters from a study of W and Z bosons. Phys. Lett. B 1987, 186, 440–451. [Google Scholar] [CrossRef]
  27. Zyla, P.A. et al. [Particle Data Group] Review of Particle Physics. Prog. Theor. Exp. Phys. 2020, 2020, 083C01. [Google Scholar] [CrossRef]
  28. Peccei, R.D.; Quinn, H.R. CP Conservation in the Presence of Pseudoparticles. Phys. Rev. Lett. 1977, 38, 1440–1443. [Google Scholar] [CrossRef]
  29. Deur, A.; Brodsky, S.J.; de Téramond, G.F. The QCD running coupling. Prog. Part. Nucl. Phys. 2016, 90, 1–74. [Google Scholar] [CrossRef]
  30. Barate, R. et al. [ALEPH Collaboration] A measurement of the QCD colour factors and a limit on the light gluino. Z. Phys. C Part. Fields 1997, 76, 1–14. [Google Scholar] [CrossRef]
  31. Dolan, M.J.; Ferber, T.; Hearty, C.; Kahlhoefer, F.; Schmidt-Hoberg, K. Revised constraints and Belle II sensitivity for visible and invisible axion-like particles. J. High Energy Phys. 2017, 12, 094, Erratum in J. High Energy Phys. 2021, 3, 190. [Google Scholar] [CrossRef]
  32. Liao, J.; Gao, Y.; Liang, Z.; Peng, Z.; Ouyang, Z.; Zhang, L.; Zhang, L.; Zhou, J. A low-mass dark matter project, ALETHEIA: A Liquid hElium Time projection cHambEr In dArk matter. arXiv 2021, arXiv:2103.02161. [Google Scholar]
  33. Cui, C.X.; Ishida, H.; Kawaguchi, M.; Li, J.Y.; Matsuzaki, S.; Tomiya, A. QCD knows new quarks. arXiv 2021, arXiv:2112.13533. [Google Scholar]
  34. Nakamura, Y.; Schierholz, G. The strong CP problem solved by itself due to long-distance vacuum effects. Nucl. Phys. B 2023, 986, 116063. [Google Scholar] [CrossRef]
  35. Nakamura, Y.; Schierholz, G. Does confinement imply CP invariance of the strong interactions? arXiv 2019, arXiv:1912.03941. [Google Scholar] [CrossRef]
  36. Schierholz, G. Dynamical solution of the strong CP problem within QCD? EPJ Web Conf. 2022, 274, 01009. [Google Scholar] [CrossRef]
  37. Kim, J.E. Composite invisible axion. Phys. Rev. D 1985, 31, 1733–1735. [Google Scholar] [CrossRef]
  38. Choi, K.; Kim, J.E. Dynamical axion. Phys. Rev. D 1985, 32, 1828–1834. [Google Scholar] [CrossRef]
  39. Jaeckel, J.; Spannowsky, M. Probing MeV to 90 GeV axion-like particles with LEP and LHC. Phys. Lett. B 2016, 753, 482–487. [Google Scholar] [CrossRef]
  40. Nicola, A.G.; de Elvira, J.R. Pseudoscalar susceptibilities and quark condensates: Chiral restoration and lattice screening masses. J. High Energy Phys. 2016, 2016, 1–28. [Google Scholar] [CrossRef]
  41. Kawaguchi, M.; Matsuzaki, S.; Tomiya, A. Analysis of nonperturbative flavor violation at chiral crossover criticality in QCD. Phys. Rev. D 2021, 103, 054034. [Google Scholar] [CrossRef]
  42. Baluni, V. CP-nonconserving effects in quantum chromodynamics. Phys. Rev. D 1979, 19, 2227–2230. [Google Scholar] [CrossRef]
  43. Kim, J.E. Light pseudoscalars, particle physics and cosmology. Phys. Rep. 1987, 150, 1–177. [Google Scholar] [CrossRef]
  44. Kobayashi, M.; Maskawa, T. Chiral Symmetry and η-XMixing. Prog. Theor. Phys. 1970, 44, 1422–1424. [Google Scholar] [CrossRef]
  45. Kobayashi, M.; Kondo, H.; Maskawa, T. Symmetry Breaking of the Chiral U(3) ⊗ U(3) and the Quark Model. Prog. Theor. Phys. 1971, 45, 1955–1959. [Google Scholar] [CrossRef]
  46. ‘t Hooft, G. Symmetry Breaking through Bell-Jackiw Anomalies. Phys. Rev. Lett. 1976, 37, 8–11. [Google Scholar] [CrossRef]
  47. ‘t Hooft, G. Computation of the quantum effects due to a four-dimensional pseudoparticle. Phys. Rev. D 1976, 14, 3432–3450, Erratum in Phys. Rev. D 1978, 18, 2199. [Google Scholar] [CrossRef]
  48. Kunihiro, T. Chiral restoration, flavor symmetry and the axial anomaly at finite temperature in an effective theory. Nucl. Phys. B 1991, 351, 593–622. [Google Scholar] [CrossRef]
Figure 1. This plot visualizes that the QCD of the SM is disfavored in terms of the presently addressed fine-tuning problem. The estimated numbers of the estimator R are displayed in the ( m l , m s ) plane, where the cross mark “×”, labeled as the “QCD of standard model”, has been plotted by using the best-fit model parameters in Table A2 of Appendix D. A possible theoretical uncertainty of 30% for the present NJL model to match the full QCD of the SM has also been reflected there (see also the text), which is drawn by light-blue arrows. The size of deviations is maximally about 340 σ , and will be at least over 66 σ even when the theoretical uncertainty of 30% is considered.
Figure 1. This plot visualizes that the QCD of the SM is disfavored in terms of the presently addressed fine-tuning problem. The estimated numbers of the estimator R are displayed in the ( m l , m s ) plane, where the cross mark “×”, labeled as the “QCD of standard model”, has been plotted by using the best-fit model parameters in Table A2 of Appendix D. A possible theoretical uncertainty of 30% for the present NJL model to match the full QCD of the SM has also been reflected there (see also the text), which is drawn by light-blue arrows. The size of deviations is maximally about 340 σ , and will be at least over 66 σ even when the theoretical uncertainty of 30% is considered.
Universe 10 00068 g001
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Cui, C.-X.; Li, J.-Y.; Ishida, H.; Kawaguchi, M.; Matsuzaki, S.; Tomiya, A. New Indication from Quantum Chromodynamics Calling for beyond the Standard Model. Universe 2024, 10, 68. https://doi.org/10.3390/universe10020068

AMA Style

Cui C-X, Li J-Y, Ishida H, Kawaguchi M, Matsuzaki S, Tomiya A. New Indication from Quantum Chromodynamics Calling for beyond the Standard Model. Universe. 2024; 10(2):68. https://doi.org/10.3390/universe10020068

Chicago/Turabian Style

Cui, Chuan-Xin, Jin-Yang Li, Hiroyuki Ishida, Mamiya Kawaguchi, Shinya Matsuzaki, and Akio Tomiya. 2024. "New Indication from Quantum Chromodynamics Calling for beyond the Standard Model" Universe 10, no. 2: 68. https://doi.org/10.3390/universe10020068

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