Joel M. Moss

Last updated

Joel Marshall Moss (born November 29, 1942) is an American experimental nuclear physicist.

Education and career

Moss received his bachelor's degree from Fort Hays State University in 1964 [1] and his doctorate in physics from the University of California, Berkeley in 1969. [2] [3] As a postdoc he was from 1969 to 1971 a research associate at the Saclay Nuclear Research Centre and from 1971 to 1973 an instructor in physics at the University of Minnesota. He was from 1973 to 1978 an assistant professor and from 1978 to 1980 associate professor at Texas A&M University. [1] There he studied giant resonances of atomic nuclei with Texas A&M's cyclotron [4] and introduced a new technique of focal plane polarimetry [5] using a high-efficiency, high-resolution polarimeter in conjunction with an Enge [6] split-pole spectrograph. [7]

In 1979 he became a researcher in the physics division of Los Alamos National Laboratory (LANL), [1] where he developed and applied his technique of focal plane polarimetry at Los Alamos Meson Physics Facility (LAMPF) and also at the cyclotron of Indiana University. [5] For example, he used his polarimetry technique to search (unsuccessfully) for collective pion excitations in nuclei in spin-sensitive experiments. At LANL he was steadily promoted: Leader of the Nuclear Physics Group from 1982 to 1984, Leader from of the Medium Energy Physics Group from 1984 to 1987, Deputy Division Leader of the Medium Energy Physics Group from 1987 to 1993, and Program Director of Nuclear and Particle Physics from 1988 to 1990. [1] [8]

In 1986 Moss became the spokesperson for the E772 experiment at Fermilab, which involved dimuon production (i.e. of muon pairs via a Drell–Yan process and from charmonium decays) in high-energy proton-nucleus collisions with 800 GeV protons at the Tevatron. [5] [9] [10] [11] In particular, they obtained information about the antiquark distribution of sea quarks in the nucleons in the nuclei of hydrogen and deuterium targets and were able to study their dependence on the mass number of the nucleus. [12] There was no mass-number-dependent modification (i.e. a different behavior of nucleons in nuclei than in free nucleons), as was observed in 1983 in the EMC effect found in the European Muon Collaboration experiments on deep inelastic lepton scattering on nuclear targets. This was contrary to what was expected from the explanation of the EMC effect from pion effects (increased occurrence of antiquarks) in nuclei. E772 could not detect any such antiquark enhancement. In addition, the E772 science team gained evidence of charmonium and charm formation in nuclei from dimuon generation. [13]

He was also involved in experiments on deep inelastic scattering from nuclei and nucleons at Fermilab. He participated in experiments at the PHENIX detector of the RHIC heavy ion accelerator to study high-energy nuclear collisions and the spin structure of the nucleon. [5] [14] [15]

In 1983 Moss was elected a Fellow of the American Physical Society. [16] In 1998 he received the Tom W. Bonner Prize in Nuclear Physics with citation:

"For his pioneering experiments using dimuon production in proton-nucleus interactions which demonstrate that there is no antiquark enhancement in nuclei, and which delineate the characteristics of charmonium and open charm production in nuclear systems. [5]

Related Research Articles

<span class="mw-page-title-main">Hadron</span> Composite subatomic particle

In particle physics, a hadron is a composite subatomic particle made of two or more quarks held together by the strong interaction. They are analogous to molecules that are held together by the electric force. Most of the mass of ordinary matter comes from two hadrons: the proton and the neutron, while most of the mass of the protons and neutrons is in turn due to the binding energy of their constituent quarks, due to the strong force.

<span class="mw-page-title-main">Subatomic particle</span> Particle smaller than an atom

In physics, a subatomic particle is a particle smaller than an atom. According to the Standard Model of particle physics, a subatomic particle can be either a composite particle, which is composed of other particles, or an elementary particle, which is not composed of other particles. Particle physics and nuclear physics study these particles and how they interact. Most force carrying particles like photons or gluons are called bosons and, although they have discrete quanta of energy, do not have rest mass or discrete diameters and are unlike the former particles that have rest mass and cannot overlap or combine which are called fermions.

<span class="mw-page-title-main">J/psi meson</span> Subatomic particle made of a charm quark and antiquark

The
J/ψ
(J/psi) meson is a subatomic particle, a flavor-neutral meson consisting of a charm quark and a charm antiquark. Mesons formed by a bound state of a charm quark and a charm anti-quark are generally known as "charmonium" or psions. The
J/ψ
is the most common form of charmonium, due to its spin of 1 and its low rest mass. The
J/ψ
has a rest mass of 3.0969 GeV/c2, just above that of the
η
c
, and a mean lifetime of 7.2×10−21 s. This lifetime was about a thousand times longer than expected.

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.

<span class="mw-page-title-main">Exotic hadron</span> Subatomic particles consisting of quarks and gluons

Exotic hadrons are subatomic particles composed of quarks and gluons, but which – unlike "well-known" hadrons such as protons, neutrons and mesons – consist of more than three valence quarks. By contrast, "ordinary" hadrons contain just two or three quarks. Hadrons with explicit valence gluon content would also be considered exotic. In theory, there is no limit on the number of quarks in a hadron, as long as the hadron's color charge is white, or color-neutral.

<span class="mw-page-title-main">Deep inelastic scattering</span> Type of collision between subatomic particles

In particle physics, deep inelastic scattering is the name given to a process used to probe the insides of hadrons, using electrons, muons and neutrinos. It was first attempted in the 1960s and 1970s and provided the first convincing evidence of the reality of quarks, which up until that point had been considered by many to be a purely mathematical phenomenon. It is an extension of Rutherford scattering to much higher energies of the scattering particle and thus to much finer resolution of the components of the nuclei.

<span class="mw-page-title-main">Drell–Yan process</span> Process in high-energy hadron–hadron scattering

The Drell–Yan process occurs in high energy hadron–hadron scattering. It takes place when a quark of one hadron and an antiquark of another hadron annihilate, creating a virtual photon or Z boson which then decays into a pair of oppositely-charged leptons. Importantly, the energy of the colliding quark-antiquark pair can be almost entirely transformed into the mass of new particles. This process was first suggested by Sidney Drell and Tung-Mow Yan in 1970 to describe the production of lepton–antilepton pairs in high-energy hadron collisions. Experimentally, this process was first observed by J. H. Christenson et al. in proton–uranium collisions at the Alternating Gradient Synchrotron.

<span class="mw-page-title-main">CDHS experiment</span>

CDHS was a neutrino experiment at CERN taking data from 1976 until 1984. The experiment was officially referred to as WA1. CDHS was a collaboration of groups from CERN, Dortmund, Heidelberg, Saclay and later Warsaw. The collaboration was led by Jack Steinberger. The experiment was designed to study deep inelastic neutrino interactions in iron.

<span class="mw-page-title-main">MINERνA</span> Neutrino scattering experiment at Fermilab in Illinois, USA

Main Injector Experiment for ν-A, or MINERνA, is a neutrino scattering experiment which uses the NuMI beamline at Fermilab. MINERνA seeks to measure low energy neutrino interactions both in support of neutrino oscillation experiments and also to study the strong dynamics of the nucleon and nucleus that affect these interactions.

Kurt Gottfried was an Austrian-born American physicist who was professor emeritus of physics at Cornell University. He was known for his work in the areas of quantum mechanics and particle physics and was also a co-founder with Henry Way Kendall of the Union of Concerned Scientists. He wrote extensively in the areas of physics and arms control.

<span class="mw-page-title-main">Quark–gluon plasma</span> Phase of quantum chromodynamics (QCD)

Quark–gluon plasma is an interacting localized assembly of quarks and gluons at thermal and chemical (abundance) equilibrium. The word plasma signals that free color charges are allowed. In a 1987 summary, Léon van Hove pointed out the equivalence of the three terms: quark gluon plasma, quark matter and a new state of matter. Since the temperature is above the Hagedorn temperature—and thus above the scale of light u,d-quark mass—the pressure exhibits the relativistic Stefan-Boltzmann format governed by temperature to the fourth power and many practically massless quark and gluon constituents. It can be said that QGP emerges to be the new phase of strongly interacting matter which manifests its physical properties in terms of nearly free dynamics of practically massless gluons and quarks. Both quarks and gluons must be present in conditions near chemical (yield) equilibrium with their colour charge open for a new state of matter to be referred to as QGP.

The EMC effect is the surprising observation that the cross section for deep inelastic scattering from an atomic nucleus is different from that of the same number of free protons and neutrons. From this observation, it can be inferred that the quark momentum distributions in nucleons bound inside nuclei are different from those of free nucleons. This effect was first observed in 1983 at CERN by the European Muon Collaboration, hence the name "EMC effect". It was unexpected, since the average binding energy of protons and neutrons inside nuclei is insignificant when compared to the energy transferred in deep inelastic scattering reactions that probe quark distributions. While over 1000 scientific papers have been written on the topic and numerous hypotheses have been proposed, no definitive explanation for the cause of the effect has been confirmed. Determining the origin of the EMC effect is one of the major unsolved problems in the field of nuclear physics.

<span class="mw-page-title-main">Fermilab E-906/SeaQuest</span>

Fermilab E-906/SeaQuest is a particle physics experiment which will use Drell–Yan process to measure the contributions of antiquarks to the structure of the proton or neutron and how this structure is modified when the proton or neutron is included within an atomic nucleus.

<span class="mw-page-title-main">Nuclear drip line</span> Atomic nuclei decay delimiter

The nuclear drip line is the boundary beyond which atomic nuclei are unbound with respect to the emission of a proton or neutron.

The rms charge radius is a measure of the size of an atomic nucleus, particularly the proton distribution. The proton radius is about one femtometre = 10−15 metre. It can be measured by the scattering of electrons by the nucleus. Relative changes in the mean squared nuclear charge distribution can be precisely measured with atomic spectroscopy.

The polarized targets are used as fixed targets in scattering experiments. In high energy physics they are used to study the nucleon spin structure of simple nucleons like protons, neutrons or deuterons. In deep inelastic scattering the hadron structure is probed with electrons, muons or neutrinos. Using a polarized high energy muon beam, for example, on a fixed target with polarized nucleons it is possible to probe the spin dependent part of the structure functions.

The proton spin crisis is a theoretical crisis precipitated by a 1987 experiment by the European Muon Collaboration (EMC), which tried to determine the distribution of spin within the proton.

The nucleon magnetic moments are the intrinsic magnetic dipole moments of the proton and neutron, symbols μp and μn. The nucleus of an atom comprises protons and neutrons, both nucleons that behave as small magnets. Their magnetic strengths are measured by their magnetic moments. The nucleons interact with normal matter through either the nuclear force or their magnetic moments, with the charged proton also interacting by the Coulomb force.

<span class="mw-page-title-main">The shape of the atomic nucleus</span>

The shape of the atomic nucleus been depicted as a compact bundle of the two types of nucleons that look like little balls stuck together, protons (red) and neutrons (blue). This depiction of the atomic nucleus approximates the empirical evidence for the size and shape of nucleons and nuclei as outlined in the article below, beginning with the discovery of the quadrapole moment in 1935 and its role in shape. Factors affecting nuclear shape include the prolate spheroid shape of the nucleon, the distance between nucleons, and the radial charge density distribution. The unusual cosmic abundance of alpha nuclides has inspired geometric arrangements of alpha particles as a solution to nuclear shapes, although the atomic nucleus generally assumes a prolate spheroid shape. Nuclides can also be discus-shaped, triaxial or pear-shaped.

References

  1. 1 2 3 4 "Joel M. Moss, Biography". American Institute of Physics (AIP).
  2. "Joel Marshall Moss". Physics Tree.
  3. "Abstract # 5673". Nuclear Science Abstracts, Volume 25, Number 3, Abstracts 4008-8192. United States Atomic Energy Commission, Division of Technical Information. February 15, 1971. p. 546.
  4. Youngblood, D. H.; Bacher, A. D.; Brown, D. R.; Bronson, J. D.; Moss, J. M.; Rozsa, C. M. (1977). "Particle decay from the giant resonance region of 40Ca". Physical Review C. 15 (1): 246–259. Bibcode:1977PhRvC..15..246Y. doi:10.1103/PhysRevC.15.246. hdl: 1969.1/126734 .
  5. 1 2 3 4 5 "1998 Tom W. Bonner Prize in Nuclear Physics Recipient, Joel M. Moss". American Physical Society.
  6. Enge, Harald A. (1958). "Combined Magnetic Spectrograph and Spectrometer". Review of Scientific Instruments. 29 (10): 885–888. Bibcode:1958RScI...29..885E. doi:10.1063/1.1716028.
  7. Moss, J.M.; Brown, D.R.; Cornelius, W.D. (1976). "Proton polarimetry using an Enge split-pole spectrograph". Nuclear Instruments and Methods. 135 (1): 139–143. Bibcode:1976NucIM.135..139M. doi:10.1016/0029-554X(76)90837-5.
  8. information from American Men and Women of Science, Thomson Gale 2004
  9. McGaughey, P. L.; Moss, J. M.; Peng, J. C. (December 1999). "High-Energy Hadron-Induced Dilepton Production from Nucleons and Nuclei". Annual Review of Nuclear and Particle Science. 49: 217–253. arXiv: hep-ph/9905409 . Bibcode:1999ARNPS..49..217M. doi:10.1146/annurev.nucl.49.1.217. S2CID   16426595.
  10. Vasiliev, M. A.; et al. (1999). "Parton Energy Loss Limits and Shadowing in Drell-Yan Dimuon Production". Physical Review Letters. 83 (12): 2304–2307. arXiv: hep-ex/9906010 . Bibcode:1999PhRvL..83.2304V. doi:10.1103/PhysRevLett.83.2304. S2CID   119409844.
  11. preprints of the E772 collaboration at LANL
  12. Hawker, E. A.; et al. (1998). "Measurement of the Light Antiquark Flavor Asymmetry in the Nucleon Sea". Physical Review Letters. 80 (17): 3715–3718. arXiv: hep-ex/9803011 . Bibcode:1998PhRvL..80.3715H. doi:10.1103/PhysRevLett.80.3715. S2CID   54921026. (over 600 citations)
  13. Alde, D. M.; Baer, H. W.; Carey, T. A.; Garvey, G. T.; Klein, A.; Lee, C.; Leitch, M. J.; Lillberg, J. W.; McGaughey, P. L.; Mishra, C. S.; Moss, J. M.; Peng, J. C.; Brown, C. N.; Cooper, W. E.; Hsiung, Y. B.; Adams, M. R.; Guo, R.; Kaplan, D. M.; McCarthy, R. L.; Danner, G.; Wang, M. J.; Barlett, M. L.; Hoffmann, G. W. (1990). "Nuclear dependence of dimuon production at 800 GeV". Physical Review Letters. 64 (21): 2479–2482. Bibcode:1990PhRvL..64.2479A. doi:10.1103/PhysRevLett.64.2479. PMID   10041723. (over 550 citations)
  14. Brooks, M. L.; Moss, J. M. (1997). "Heavy-Flavor Production via Single Muon Detection in the PHENIX Detector at √s= 200 GeV" (PDF). PHENIX Technical Note 361.
  15. Akikawa, H.; et al. (2003). "PHENIX Muon Arms". Nuclear Instruments and Methods in Physics Research Section A: Accelerators, Spectrometers, Detectors and Associated Equipment. 499 (2–3): 537–548. Bibcode:2003NIMPA.499..537A. doi:10.1016/S0168-9002(02)01955-1.
  16. "APS Fellow Archive". American Physical Society. (search on year=1983 and institution=Los Alamos National Laboratory)