Hypernucleus

From Justapedia, unleashing the power of collective wisdom
(Redirected from Hypernuclei)
Jump to navigation Jump to search

A hypernucleus is similar to a conventional atomic nucleus, but contains at least one hyperon in addition to the normal protons and neutrons. Hyperons are a category of baryon particles that carry non-zero strangeness quantum number, which is conserved by the strong and electromagnetic interactions.

A variety of reactions give access to depositing one or more units of strangeness in a nucleus. Hypernuclei containing the lightest hyperon, the lambda (Λ), have sharp nuclear energy levels on account of their long lifetimes regulated by their decays through the weak interaction. Because hyperons are particles distinguishable from protons and neutrons in the nuclear environment, there is a branch of experimental and theoretical nuclear spectroscopy that investigates the nuclear physics of these states, finding both similarities and differences with conventional nuclei.

Nomenclature

Diagram of the eight possible baryons with spin 1/2
Diagram of the ten possible baryons with spin 1/2
The combinations of three up, down, and strange quarks with total spin 12 (left) and with total spin 32 (right).

Hypernuclei are named in terms of their atomic number and baryon number, as in normal nuclei, plus the hyperon(s) which are listed in a left subscript of the symbol, with the caveat that atomic number is interpreted as the total charge of the hypernucleus, including charged hyperons such as the xi minus) as well as protons. For example, the hypernucleus 16
Λ
O
contains 8 protons, 7 neutrons, and one lambda (which carries no charge).[1]

History

The first was discovered by Marian Danysz and Jerzy Pniewski in 1952 using a nuclear emulsion plate exposed to cosmic rays, based on their energetic but delayed decay. This event was inferred to be due to a nuclear fragment containing a Λ baryon.[2] Experiments until the 1970s would continue to study hypernuclei produced in emulsions using cosmic rays, and later using pion (π) and kaon (K) beams from particle accelerators.[1]

Since the 1980s, more efficient production methods using pion and kaon beams have allowed further investigation at various accelerator facilities, including CERN, Brookhaven National Laboratory, KEK, DAφNE, and JPARC.[3][4] In the 2010s, heavy ion experiments such as ALICE and STAR first allowed the production and measurement of light hypernuclei formed through hadronization from quark–gluon plasma.[5]

Types

Λ hypernuclei

The simplest, and most well understood, type of hypernucleus includes only the lightest hyperon, the Λ.[6]

While two nucleons can interact through the nuclear force mediated by a virtual pion, the Λ becomes a Σ baryon upon emitting a pion,[a] so the Λ–nucleon interaction is mediated solely by more massive mesons such as the η and ω mesons, or through the simultaneous exchange of two or more mesons.[8] This means that the Λ–nucleon interaction is weaker and has a shorter range than the standard nuclear force, and the potential well of a Λ in the nucleus is shallower than that of a nucleon;[9] in hypernuclei, the depth of the Λ potential is approximately 30 MeV.[10] However, one-pion exchange in the Λ–nucleon interaction does cause quantum-mechanical mixing of the Λ and Σ baryons in hypernuclei (which does not happen in free space), especially in neutron-rich hypernuclei.[11][12][13] Additionally, the three-body force between a Λ and two nucleons is expected to be more important than the three-body interaction in nuclei, since the Λ can exchange two pions with a virtual Σ intermediate, while the equivalent process in nucleons requires a relatively heavy delta baryon (Δ) intermediate.[8]

Like all hyperons, Λ hypernuclei can decay through the weak interaction, which changes it to a lighter baryon and emits a meson or a lepton–antilepton pair. In free space, the Λ usually decays via the weak force to a proton and a π meson, or a neutron and a π0, with a total half-life of 263±ps.[14] A nucleon in the hypernucleus can cause the Λ to decay via the weak force without emitting a pion; this process becomes dominant in heavy hypernuclei, due to suppression of the pion-emitting decay mode.[15] The half-life of the Λ in a hypernucleus is considerably shorter, plateauing to about 215±14 ps near 56
Λ
Fe
,[16] but some empirical measurements substantially disagree with each other or with theoretical predictions.[17] In particular, the hypertriton lifetime averaged across all experiments (about 206+15
−13
 ps
) is substantially shorter than predicted by theory, as the non-mesonic decay mode is expected to be relatively minor, and some experimental results are substantially shorter or longer than this average.[18][19]

Σ hypernuclei

The existence of hypernuclei containing a sigma baryon (Σ) is less clear. Several experiments in the early 1980s reported bound hypernuclear states above the Λ separation energy and presumed to contain one of the slightly heavier Σ baryons, but experiments later in the decade ruled out the existence of such states.[6] Results from exotic atoms containing a Σ bound to a nucleus by the electromagnetic force have found a net repulsive Σ–nucleon interaction in medium-sized and large hypernuclei, which means that no Σ hypernuclei exist in such mass range.[6] However, an experiment in 1998 definitively observed the light Σ hypernucleus 4
Σ
He
.[6]

ΛΛ and Ξ hypernuclei

Hypernuclei containing two Λ baryons have been made. However, such hypernuclei are much harder to produce due to containing two strange quarks, and As of 2016, only seven candidate ΛΛ hypernuclei have been observed.[20] Like the Λ–nucleon interaction, empirical and theoretical models predict that the Λ–Λ interaction is mildly attractive.[21][22]

Hypernuclei containing a Ξ baryon are known. Empirical studies and theoretical models indicate that the Ξ–proton interaction is attractive, but weaker than the Λ–nucleon interaction.[21] Like the Σ and other negatively charged particles, the Ξ can also form an exotic atom. When a Ξ is bound in an exotic atom or a hypernucleus, it quickly decays to a ΛΛ hypernucleus or to two Λ hypernuclei by exchanging a strange quark with a proton, which releases about 29 MeV of energy in free space:

Ξ + p → Λ + Λ[23][24][25]

Ω hypernuclei

Hypernuclei containing the Ω baryon were predicted using lattice QCD in 2018; in particular, the proton–Ω and Ω–Ω dibaryons (bound systems containing two baryons) are expected to be stable.[26][27] As of 2022, no such hypernuclei have been observed under any conditions, but the lightest such species could be produced in heavy-ion collisions,[28] and measurements by the STAR experiment are consistent with the existence of the proton–Ω dibaryon.[29]

Production

Several modes of production have been devised to make hypernuclei through bombardment of normal nuclei.

Strangeness exchange and production

One method of producing a K meson exchanges a strange quark with a nucleon and changes it to a Λ:[30]

p + K → Λ + π0
n + K → Λ + π

The cross section for the formation of a hypernucleus is maximized when the momentum of the kaon beam is approximately 500 MeV/c.[31] Several variants of this setup exist, including ones where the incident kaons are either brought to rest before colliding with a nucleus.[30]

In rare cases, the incoming K can instead produce a Ξ hypernucleus via the reaction:

p + K → Ξ + K+[32]

The equivalent strangeness production reaction involves a π+ meson reacts with a neutron to change it to a Λ:[33]

n + π+ → Λ + K+

This reaction has a maximum cross section at a beam momentum of 1.05 GeV/c, and is the most efficient production route for Λ hypernuclei, but requires larger targets than strangeness exchange methods.[33]

Elastic scattering

Electron scattering off of a proton can change it to a Λ and produce a K+:[34]

p + e → Λ + e′ + K+

where the prime symbol denotes a scattered electron. The energy of an electron beam can be more easily tuned than pion or kaon beams, making it easier to measure and calibrate hypernuclear energy levels.[34] Initially theoretically predicted in the 1980s, this method was first used experimentally in the early 2000s.[35]

Hyperon capture

The capture of a Ξ baryon by a nucleus can make a Ξ exotic atom or hypernucleus.[23] Upon capture, it changes to a ΛΛ hypernucleus or two Λ hypernuclei.[36] The disadvantage is that the Ξ baryon is harder to make into a beam than singly strange hadrons.[37] However, an experiment at J-PARC begun in 2020 will compile data on Ξ and ΛΛ hypernuclei using a similar, non-beam setup where scattered Ξ baryons rain onto an emulsion target.[23]

Properties

Hypernuclear physics differs from that of normal nuclei because a hyperon, having a non-zero strangeness quantum number, can share space and momentum coordinates with the usual four nucleon states that can differ from each other in spin and isospin. That is, they are not restricted by the Pauli exclusion principle from any single-particle state in the nucleus. The ground state of helium-5-Lambda, for example, must resemble helium-4 more than it does helium-5 or lithium-5 and must be stable, apart from the eventual weak decay of the lambda with a mean lifetime of 263±ps.[38] Sigma hypernuclei have been sought,[39] as have doubly-strange nuclei containing cascade baryons.

The simplest hypernucleus, for example, is the hypertriton, which consists of one proton, one neutron, and one Lambda hyperon. This system of particles is bound by only about 200 keV, and has a lifetime similar to that of the free ground state Lambda hyperon of about 263 picoseconds. Exact values for these quantities are an area of active research because they are related to the nature of the hyperon-nucleon interaction potential.

Samanta formula

A generalized mass formula has been developed for both the non-strange normal nuclei and strange hypernuclei can estimate masses of hypernuclei containing Lambda, Lambda-Lambda, Sigma, and Cascade hyperon(s).[40][41] The neutron and proton driplines for hypernuclei are predicted and existence of some exotic hypernuclei beyond the normal neutron and proton driplines are suggested.[42] This generalized mass formula was named the "Samanta formula" by Botvina and Pochodzalla and used to predict relative yields of hypernuclei in multifragmentation of nuclear spectator matter.[43][jargon]

Similar species

Kaonic nuclei

The K meson can orbit a nucleus in an exotic atom, such as in kaonic hydrogen.[44] Although the K-proton strong interaction in kaonic hydrogen is repulsive,[45] the K–nucleus interaction is attractive for larger systems, so this meson can enter a strongly bound state closely related to a hypernucleus;[6] in particular, the K–proton–proton system is experimentally known and more tightly bound than a normal nucleus.[46]

Charmed hypernuclei

Nuclei containing a charm quark have been predicted theoretically since 1977.[47] and are described as charmed hypernuclei despite the possible absence of strange quarks.[48] In particular, the lightest charmed baryons, the Λc and Σc baryons,[b] are predicted to exist in bound states in charmed hypernuclei, and could be created in processes analogous to those used to make hypernuclei.[48] The depth of the Λc potential in nuclear matter is predicted to be 58 MeV,[48] but unlike Λ hypernuclei, larger hypernuclei containing the positively charged Λc would be less stable than the corresponding Λ hypernuclei due to Coulomb repulsion.[49] The mass difference between the Λc and the
Σ+
c
is too large for appreciable mixing of these baryons to occur in hypernuclei.[50] Weak decays of charmed hypernuclei have strong relativistic corrections compared to those in ordinary hypernuclei, as the energy released in the decay process is comparable to the mass of the Λ baryon.[51]

Notes

  1. ^ Isospin (I), a number describing the up and down quark content of the system, is preserved in the strong interaction. Since the isospin of a pion is 1, the Λ baryon (I = 0) must become a Σ (I = 1) upon emitting a pion.[7]
  2. ^ The subscript c in the symbols for charmed baryons indicate that a strange quark in a hyperon is replaced with a charm quark; the superscript, if present, still represents the total charge of the baryon.

References

  1. ^ a b Gal et al. 2016, p. 2.
  2. ^ Danysz, M.; Pniewski, J. (March 1953). "Delayed disintegration of a heavy nuclear fragment: I". The London, Edinburgh, and Dublin Philosophical Magazine and Journal of Science. 44 (350): 348–350. doi:10.1080/14786440308520318.
  3. ^ Gal et al. 2016, p. 4.
  4. ^ Tolos & Fabbietti 2020, p. 29.
  5. ^ Tolos & Fabbietti 2020, pp. 53–54.
  6. ^ a b c d e Feliciello, A; Nagae, T (1 September 2015). "Experimental review of hypernuclear physics: recent achievements and future perspectives". Reports on Progress in Physics. 78 (9): 096301. doi:10.1088/0034-4885/78/9/096301.
  7. ^ Gal et al. 2016, pp. 20.
  8. ^ a b Gal et al. 2016, pp. 2, 20–21.
  9. ^ Gal et al. 2016, p. 6.
  10. ^ Tolos & Fabbietti 2020, p. 50.
  11. ^ Gal et al. 2016, pp. 20–21.
  12. ^ Tolos & Fabbietti 2020, p. 52.
  13. ^ Umeya, A.; Harada, T. (20 February 2009). "Λ–Σ coupling effect in the neutron-rich Λ hypernucleus in a microscopic shell-model calculation". Physical Review C. 79 (2): 024315. arXiv:0810.4591. doi:10.1103/PhysRevC.79.024315.
  14. ^ Amsler, C.; et al. (Particle Data Group) (2008). "
    Λ
    "
    (PDF). Particle listings. Lawrence Berkeley Laboratory.
  15. ^ Tolos & Fabbietti 2020, p. 50–51.
  16. ^ Sato, Y.; Ajimura, S.; Aoki, K.; Bhang, H.; Hasegawa, T.; Hashimoto, O.; Hotchi, H.; Kim, Y. D.; Kishimoto, T.; Maeda, K.; Noumi, H.; Ohta, Y.; Omata, K.; Outa, H.; Park, H.; Sekimoto, M.; Shibata, T.; Takahashi, T.; Youn, M. (9 February 2005). "Mesonic and nonmesonic weak decay widths of medium-heavy Λ hypernuclei". Physical Review C. 71 (2): 025203. arXiv:nucl-ex/0409007v2. doi:10.1103/PhysRevC.71.025203.
  17. ^ Gal et al., 2016 & 17–18.
  18. ^ Tolos & Fabbietti 2020, p. 52–53.
  19. ^ ALICE Collaboration (October 2019). " and lifetime measurement in Pb–Pb collisions at s NN = 5.02 TeV via two-body decay". Physics Letters B. 797: 134905. doi:10.1016/j.physletb.2019.134905.
  20. ^ Gal et al. 2016, p. 41.
  21. ^ a b Tolos & Fabbietti 2020, pp. 43–45, 59.
  22. ^ ALICE Collaboration (10 October 2019). "Study of the Λ–Λ interaction with femtoscopy correlations in pp and p–Pb collisions at the LHC". Physics Letters B. 797: 134822. doi:10.1016/j.physletb.2019.134822. ISSN 0370-2693.
  23. ^ a b c Yoshida, J.; et al. (The J-PARC 07 Collaboration) (25 March 2021). "J-PARC E07: Systematic Study of Double Strangeness System with Hybrid Emulsion Method". Proceedings of the 3rd J-PARC Symposium (J-PARC2019). Journal of the Physical Society of Japan. 33. doi:10.7566/jpscp.33.011112.
  24. ^ Gal et al. 2016, pp. 16, 43.
  25. ^ Tolos & Fabbietti 2020, p. 53.
  26. ^ Iritani, Takumi; et al. (HALQCD Collaboration) (May 2019). "NΩ dibaryon from lattice QCD near the physical point". Physics Letters B. 792: 284–289. arXiv:1810.03416. doi:10.1016/j.physletb.2019.03.050.
  27. ^ Gongyo, Shinya; et al. (HALQCD Collaboration) (23 May 2018). "Most Strange Dibaryon from Lattice QCD". Physical Review Letters. 120 (21): 212001. doi:10.1103/PhysRevLett.120.212001.
  28. ^ Zhang, Liang; Zhang, Song; Ma, Yu-Gang (May 2022). "Production of $$\Omega NN$$ and $$\Omega \Omega N$$ in ultra-relativistic heavy-ion collisions". The European Physical Journal C. 82 (5): 416. doi:10.1140/epjc/s10052-022-10336-7.
  29. ^ STAR Collaboration (March 2019). "The proton–Ω correlation function in Au + Au collisions at s NN = 200 GeV". Physics Letters B. 790: 490–497. doi:10.1016/j.physletb.2019.01.055.
  30. ^ a b Gal et al. 2016, pp. 6–10.
  31. ^ Tolos & Fabbietti 2020, p. 49.
  32. ^ Gal et al. 2016, p. 16.
  33. ^ a b Gal et al. 2016, pp. 10–12.
  34. ^ a b Gal et al. 2016, p. 12.
  35. ^ Nakamura, Satoshi N.; Fujii, Yuu; Tsukada, Kyo (2013). "Precision spectroscopy of lambda hypernuclei using electron beams". Nippon Butsuri Gakkai-Shi. 68 (9): 584–592. ISSN 0029-0181.
  36. ^ Gal et al. 2016, p. 16,43.
  37. ^ Tolos & Fabbietti 2020, p. 43.
  38. ^ Gal et al. 2016, p. 30.
  39. ^ M. May (1994). "Recent results and directions in hypernuclear and kaon physics" (PDF). In A. Pascolini (ed.). PAN XIII: Particles and Nuclei. World Scientific. ISBN 978-981-02-1799-0. OSTI 10107402.
  40. ^ C. Samanta (2006). "Mass formula from normal to hypernuclei". In S. Stoica; L. Trache; R.E. Tribble (eds.). Proceedings of the Carpathian Summer School of Physics 2005. World Scientific. p. 29. ISBN 978-981-270-007-0.
  41. ^ C. Samanta, P. Roy Chowdhury, D.N.Basu (2006). "Generalized mass formula for non-strange and hyper nuclei with SU(6) symmetry breaking". Journal of Physics G. 32 (3): 363–373. arXiv:nucl-th/0504085. Bibcode:2006JPhG...32..363S. doi:10.1088/0954-3899/32/3/010.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  42. ^ C. Samanta, P. Roy Chowdhury and D.N.Basu (2008). "Lambda hyperonic effect on the normal driplines". Journal of Physics G. 35 (6): 065101–065110. arXiv:0802.3172. Bibcode:2008JPhG...35f5101S. doi:10.1088/0954-3899/35/6/065101.
  43. ^ A.S. Botvina; J. Pochodzalla (2007). "Production of hypernuclei in multifragmentation of nuclear spectator matter". Physical Review C. 76 (2): 024909–024912. arXiv:0705.2968. Bibcode:2007PhRvC..76b4909B. doi:10.1103/PhysRevC.76.024909.
  44. ^ Iwasaki, M.; Hayano, R. S.; Ito, T. M.; Nakamura, S. N.; Terada, T. P.; Gill, D. R.; Lee, L.; Olin, A.; Salomon, M.; Yen, S.; Bartlett, K.; Beer, G. A.; Mason, G.; Trayling, G.; Outa, H.; Taniguchi, T.; Yamashita, Y.; Seki, R. (21 April 1997). "Observation of Kaonic Hydrogen K α X Rays". Physical Review Letters. 78 (16): 3067–3069. doi:10.1103/PhysRevLett.78.3067.
  45. ^ Bazzi, M.; Beer, G.; Bombelli, L.; Bragadireanu, A.M.; Cargnelli, M.; Corradi, G.; Curceanu (Petrascu), C.; dʼUffizi, A.; Fiorini, C.; Frizzi, T.; Ghio, F.; Girolami, B.; Guaraldo, C.; Hayano, R.S.; Iliescu, M.; Ishiwatari, T.; Iwasaki, M.; Kienle, P.; Levi Sandri, P.; Longoni, A.; Lucherini, V.; Marton, J.; Okada, S.; Pietreanu, D.; Ponta, T.; Rizzo, A.; Romero Vidal, A.; Scordo, A.; Shi, H.; Sirghi, D.L.; Sirghi, F.; Tatsuno, H.; Tudorache, A.; Tudorache, V.; Vazquez Doce, O.; Widmann, E.; Zmeskal, J. (October 2011). "A new measurement of kaonic hydrogen X-rays". Physics Letters B. 704 (3): 113–117. arXiv:1105.3090. doi:10.1016/j.physletb.2011.09.011.
  46. ^ Sakuma, F.; et al. (December 2021). "Recent Results and Future Prospects of Kaonic Nuclei at J-PARC". Few-Body Systems. 62 (4): 103. doi:10.1007/s00601-021-01692-3.
  47. ^ Dover, C. B.; Kahana, S. H. (12 December 1977). "Possibility of Charmed Hypernuclei". Physical Review Letters. 39 (24): 1506–1509. doi:10.1103/PhysRevLett.39.1506.
  48. ^ a b c Krein, Gastão (2019). "Charmed hypernuclei and nuclear-bound charmonia": 020022. doi:10.1063/1.5118390. {{cite journal}}: Cite journal requires |journal= (help)
  49. ^ Güven, H.; Bozkurt, K.; Khan, E.; Margueron, J. (10 December 2021). "Ground state properties of charmed hypernuclei within a mean field approach". Physical Review C. 104 (6): 064306. doi:10.1103/PhysRevC.104.064306.
  50. ^ Vidaña, I.; Ramos, A.; Jiménez-Tejero, C. E. (23 April 2019). "Charmed nuclei within a microscopic many-body approach". Physical Review C. 99 (4): 045208. doi:10.1103/PhysRevC.99.045208.
  51. ^ Fontoura, C E; Krmpotić, F; Galeão, A P; Conti, C De; Krein, G (1 January 2018). "Nonmesonic weak decay of charmed hypernuclei". Journal of Physics G: Nuclear and Particle Physics. 45 (1): 015101. doi:10.1088/1361-6471/aa982a.