Top quark

Source: Wikipedia, the free encyclopedia.
(Redirected from
Top Quark
)

Top quark
ħ
Topness1
Weak isospinLH: + 1 /2, RH: 0
Weak hyperchargeLH: +1 /3, RH: +4/3

The top quark, sometimes also referred to as the truth quark, (symbol: t) is the most massive of all observed

Higgs Boson
. This coupling is very close to unity; in the Standard Model of particle physics, it is the largest (strongest) coupling at the scale of the weak interactions and above. The top quark was discovered in 1995 by the CDF[2] and [3] experiments at Fermilab.

Like all other

equal magnitude but opposite sign
.

The top quark interacts with

W boson and either a bottom quark (most frequently), a strange quark, or, on the rarest of occasions, a down quark.[a]

The Standard Model determines the top quark's

mean lifetime to be roughly 5×10−25 s.[5] This is about a twentieth of the timescale for strong interactions,[b] and therefore it does not form hadrons, giving physicists a unique opportunity to study a "bare" quark (all other quarks hadronize, meaning that they combine with other quarks to form hadrons
and can only be observed as such).

Because the top quark is so massive, its properties allowed indirect determination of the mass of the Higgs boson (see § Mass and coupling to the Higgs boson below). As such, the top quark's properties are extensively studied as a means to discriminate between competing theories of new physics beyond the Standard Model. The top quark is the only quark that has been directly observed due to its decay time being shorter than the hadronization time.[b][6]

History

In 1973,

doublet.[9][10]

The proposal of Kobayashi and Maskawa heavily relied on the

leptons, breaking the new symmetry
between leptons and quarks introduced by the GIM mechanism. Restoration of the symmetry implied the existence of a fifth and sixth quark.

It was in fact not long until a fifth quark, the bottom, was discovered by the

Leon Lederman at Fermilab in 1977.[14][15][16] This strongly suggested that there must also be a sixth quark, the top, to complete the pair. It was known that this quark would be heavier than the bottom, requiring more energy to create in particle collisions, but the general expectation was that the sixth quark would soon be found. However, it took another 18 years before the existence of the top was confirmed.[17]

Early searches for the top quark at

Z boson, it was again felt that the discovery of the top was imminent. As the SPS gained competition from the Tevatron at Fermilab there was still no sign of the missing particle, and it was announced by the group at CERN that the top mass must be at least 41 GeV/c2. After a race between CERN and Fermilab to discover the top, the accelerator at CERN reached its limits without creating a single top, pushing the lower bound on its mass up to 77 GeV/c2.[17]

The Tevatron was (until the start of

DØ detector, was added to the complex (in addition to the Collider Detector at Fermilab (CDF) already present). In October 1992, the two groups found their first hint of the top, with a single creation event that appeared to contain the top. In the following years, more evidence was collected and on 22 April 1994, the CDF group submitted their article presenting tentative evidence for the existence of a top quark with a mass of about 175 GeV/c2. In the meantime, DØ had found no more evidence than the suggestive event in 1992. A year later, on 2 March 1995, after having gathered more evidence and reanalyzed the DØ data (which had been searched for a much lighter top), the two groups jointly reported the discovery of the top at a mass of 176±18 GeV/c2.[2][3][17]

In the years leading up to the top-quark discovery, it was realized that certain precision measurements of the electroweak vector boson masses and couplings are very sensitive to the value of the top-quark mass. These effects become much larger for higher values of the top mass and therefore could indirectly see the top quark even if it could not be directly detected in any experiment at the time. The largest effect from the top-quark mass was on the

Martinus Veltman winning the Nobel Prize in physics in 1999.[18][19]

Properties

Production

Because top quarks are very massive, large amounts of energy are needed to create one. The only way to achieve such high energies is through high-energy collisions. These occur naturally in the Earth's upper atmosphere as

center-of-mass energy
of 7 TeV. There are multiple processes that can lead to the production of top quarks, but they can be conceptually divided in two categories: top-pair production, and single-top production.

Top-quark pairs

gluon–gluon fusion
t-channel
quark–antiquark annihilation

The most common is

Z-boson. However, these processes are predicted to be much rarer and have a virtually identical experimental signature in a hadron collider
like Tevatron.

Single top quarks

s-channel
t-channel
tW channel

The production of single top quarks via

CKM matrix
.

Decay

ttbar-decay final states
All possible final states of the decay of a top-quark pair

The only known way the top quark can decay is through the weak interaction, producing a W boson and a bottom quark.[a] Because of its enormous mass, the top quark is extremely short-lived, with a predicted lifetime of only 5×10−25 s.[5] As a result, top quarks do not have time before they decay to form hadrons as other quarks do.[b] The absence of a hadron surrounding the top quark provides physicists with the unique opportunity to study the behavior of a "bare" quark.

In particular, it is possible to directly determine the

branching ratio
:

The best current determination of this ratio is 0.957±0.034.[25] Since this ratio is equal to |Vtb|2 according to the Standard Model, this gives another way of determining the CKM element |Vtb|, or in combination with the determination of |Vtb| from single top production provides tests for the assumption that the CKM matrix is unitary.[26]

The Standard Model also allows more exotic decays, but only at one loop level, meaning that they are extremely rare. In particular, it is conceivable that a top quark might decay into another up-type quark (an up or a charm) by emitting a photon or a Z-boson.

confidence.[25]

Mass and coupling to the Higgs boson

The Standard Model generates fermion masses through their couplings to the Higgs boson. This Higgs boson acts as a field filling space. Fermions interact with this field in proportion to their individual coupling constants which generates mass. A low-mass particle, such as the electron has a minuscule coupling while the top quark has the largest coupling to the Higgs,

In the Standard Model, all of the quark and lepton Higgs–Yukawa couplings are small compared to the top-quark Yukawa coupling. This hierarchy in the fermion masses remains a profound and open problem in theoretical physics. Higgs–Yukawa couplings are not fixed constants of nature, as their values vary slowly as the energy scale (distance scale) at which they are measured. These dynamics of Higgs–Yukawa couplings, called "running coupling constants", are due to a quantum effect called the renormalization group.

The Higgs–Yukawa couplings of the up, down, charm, strange and bottom quarks are hypothesized to have small values at the extremely high energy scale of grand unification, 1015 GeV. They increase in value at lower energy scales, at which the quark masses are generated by the Higgs. The slight growth is due to corrections from the QCD coupling. The corrections from the Yukawa couplings are negligible for the lower-mass quarks.

One of the prevailing views in particle physics is that the size of the top-quark Higgs–Yukawa coupling is determined by a unique nonlinear property of the renormalization group equation that describes the running of the large Higgs–Yukawa coupling of the top quark. If a quark Higgs–Yukawa coupling has a large value at very high energies, its Yukawa corrections will evolve downward in mass scale and cancel against the QCD corrections. This is known as a (quasi-) infrared fixed point, which was first predicted by B. Pendleton and G.G. Ross,[28] and by Christopher T. Hill,[29] No matter what the initial starting value of the coupling is, if sufficiently large, it will reach this fixed-point value. The corresponding quark mass is then predicted.

The top-quark Yukawa coupling lies very near the infrared fixed point of the Standard Model. The renormalization group equation is:

where g3 is the color gauge coupling, g2 is the weak isospin gauge coupling, and g1 is the weak hypercharge gauge coupling. This equation describes how the Yukawa coupling changes with energy scale μ. Solutions to this equation for large initial values yt cause the right-hand side of the equation to quickly approach zero, locking yt to the QCD coupling g3.

The value of the top quark fixed point is fairly precisely determined in the Standard Model, leading to a top-quark mass of 220 GeV. This is about 25% larger than the observed top mass and may be hinting at new physics at higher energy scales.

The quasi-infrared fixed point subsequently became the basis of

top quark condensation and topcolor theories of electroweak symmetry breaking, in which the Higgs boson is composed of a pair of top and antitop quarks. The predicted top-quark mass comes into improved agreement with the fixed point if there are additional Higgs scalars beyond the standard model and therefore it may be hinting at a rich spectroscopy of new Higgs fields at energy scales that can be probed with the LHC and its upgrades.[30][31]

See also

Footnotes

  1. ^ a b The overwhelming majority of top quark decays produce a bottom quark, whose mass is closest to the top's. On very rare occasions it may decay into a strange quark; almost never a down quark.
  2. ^ a b c Top quark decay is an exceptional example of a weak process that is faster than a strong interaction.

References

  1. ^ a b Zyla, P.A.; et al. (Particle Data Group) (2020). "2020 Review of Particle Physics". Progress of Theoretical and Experimental Physics: 083C01.
  2. ^ a b Abe, F.; et al. (
    S2CID 119451328
    .
  3. ^ a b Abachi, S.; et al. (
    S2CID 42826202
    .
  4. ^ Elert, Glenn. "Quantum Chromodynamics". The Physics Hypertextbook. Retrieved 23 March 2019.
  5. ^ a b Quadt, A. (2006). "Top quark physics at hadron colliders".
    S2CID 121887478
    .
  6. . Retrieved 11 June 2020.
  7. ^ Harari, H. (1975). "A new quark model for hadrons". .
  8. ^ Staley, K.W. (2004). The Evidence for the Top Quark. .
  9. ^ Perkins, D.H. (2000). Introduction to High Energy Physics. .
  10. ^ Close, F. (2006). The New Cosmic Onion. .
  11. ^ Glashow, S.L.; Iliopoulous, J.; Maiani, L. (1970). "Weak interactions with lepton–hadron symmetry". .
  12. ^ Pickering, A. (1999). Constructing Quarks: A sociological history of particle physics. .
  13. ^ Perl, M.L.; et al. (1975). "Evidence for anomalous lepton production in
    e+

    e
    annihilation". .
  14. ^ "Discovery of the bottom quark" (Press release). Discoveries at Fermilab. Fermilab. 7 August 1977. Retrieved 24 July 2009.
  15. ^ Lederman, L.M. (2005). "Logbook: Bottom quark". Symmetry Magazine. Vol. 2, no. 8. Archived from the original on 4 October 2006.
  16. ^ Herb, S.W.; et al. (1977). "Observation of a di‑muon resonance at 9.5 GeV in 400 GeV proton–nucleus collisions".
    OSTI 1155396
    .
  17. ^ a b c d Liss, T.M.; Tipton, P.L. (1997). "The discovery of the top quark" (PDF). .
  18. ^ "The Nobel Prize in Physics 1999" (Press release).
    The Nobel Foundation
    . Retrieved 10 September 2009.
  19. ^ "The Nobel Prize in Physics 1999" (Press release).
    The Nobel Foundation
    . 12 October 1999. Retrieved 10 September 2009.
  20. ^ Chakraborty, D.; et al. (.
  21. ^ Abazov, V.M.; et al. (
    S2CID 1147194
    .
  22. ^ a b Abazov, V.M.; et al. (
    S2CID 14919683
    .
  23. ^ Abazov, V.M.; et al. (
    S2CID 14937909
    .
  24. ^ Aaltonen, T.; et al. (
    S2CID 8029289
    .
  25. ^ a b Zyla, P.A.; et al. (Particle Data Group) (2020). "QUARKS" (PDF). Progress of Theoretical and Experimental Physics: 083C01. Retrieved 22 May 2022.
  26. ^ Abazov, V.M.; et al. (
    S2CID 2638258
    .
  27. ^ Chekanov, S.; et al. (
    S2CID 119494760
    .
  28. ^ Pendleton, Brian; Ross, Graham (1981). "Mass and mixing angle predictions from infrared fixed points". .
  29. ^ Hill, Christopher T. (1981). "Quark and lepton masses from renormalization group fixed points". .
  30. ^ Hill, Christopher T.; Machado, Pedro; Thomsen, Anders; Turner, Jessica (2019). "Where are the next Higgs bosons?". Physical Review D. 100 (1): 015051.
    S2CID 104291827
    .
  31. ^ Hill, Christopher T.; Machado, Pedro; Thomsen, Anders; Turner, Jessica (2019). "Scalar democracy". Physical Review D. 100 (1): 015015.
    S2CID 119193325
    .

Further reading

External links