Positron annihilation spectroscopy

Last updated
Condensed matter
experiments
Levitation of a magnet on top of a superconductor 2.jpg
ARPES
ACAR
Neutron scattering
X-ray spectroscopy
Quantum oscillations
Scanning tunneling microscopy

Positron annihilation spectroscopy (PAS) [1] or sometimes specifically referred to as positron annihilation lifetime spectroscopy (PALS) is a non-destructive spectroscopy technique to study voids and defects in solids. [2] [3]

Contents

Theory

A Feynman diagram of an electron and positron annihilating into a photon. Electron-positron-annihilation.svg
A Feynman diagram of an electron and positron annihilating into a photon.

The technique operates on the principle that a positron or positronium will annihilate through interaction with electrons. This annihilation releases gamma rays that can be detected; the time between emission of positrons from a radioactive source and detection of gamma rays due to annihilation corresponds to the lifetime of positron or positronium.

When positrons are injected into a solid body, they interact in some manner with the electrons in that species. For solids containing free electrons (such as metals or semiconductors), the implanted positrons annihilate rapidly unless voids such as vacancy defects are present. If voids are available, positrons will reside in them and annihilate less rapidly than in the bulk of the material, on time scales up to ~1 ns. For insulators such as polymers or zeolites, implanted positrons interact with electrons in the material to form positronium.

Positronium is a metastable hydrogen-like bound state of an electron and a positron which can exist in two spin states. Para-positronium, p-Ps, is a singlet state (the positron and electron spins are anti-parallel) with a characteristic self-annihilation lifetime of 125 ps in vacuum. [4] Ortho-positronium, o-Ps, is a triplet state (the positron and electron spins are parallel) with a characteristic self-annihilation lifetime of 142 ns in vacuum. [4] In molecular materials, the lifetime of o-Ps is environment dependent and it delivers information pertaining to the size of the void in which it resides. Ps can pick up a molecular electron with an opposite spin to that of the positron, leading to a reduction of the o-Ps lifetime from 142 ns to 1-4 ns (depending on the size of the free volume in which it resides). [4] The size of the molecular free volume can be derived from the o-Ps lifetime via the semi-empirical Tao-Eldrup model. [5]

While the PALS is successful in examining local free volumes, it still needs to employ data from combined methods in order to yield free volume fractions. Even approaches to obtain fractional free volume from the PALS data that claim to be independent on other experiments, such as PVT measurements, they still do employ theoretical considerations, such as iso-free-volume amount from Simha-Boyer theory. A convenient emerging method for obtaining free volume amounts in an independent manner are computer simulations; these can be combined with the PALS measurements and help to interpret the PALS measurements. [6]

Pore structure in insulators can be determined using the quantum mechanical Tao-Eldrup model [7] [8] and extensions thereof. By changing the temperature at which a sample is analyzed, the pore structure can be fit to a model where positronium is confined in one, two, or three dimensions. However, interconnected pores result in averaged lifetimes that cannot distinguish between smooth channels or channels having smaller, open, peripheral pores due to energetically favored positronium diffusion from small to larger pores.

The behavior of positrons in molecules or condensed matter is nontrivial due to the strong correlation between electrons and positrons. Even the simplest case, that of a single positron immersed in a homogeneous gas of electrons, has proved to be a significant challenge for theory. The positron attracts electrons to it, increasing the contact density and hence enhancing the annihilation rate. Furthermore, the momentum density of annihilating electron-positron pairs is enhanced near the Fermi surface. [9] Theoretical approaches used to study this problem have included the Tamm-Dancoff approximation, [10] Fermi [11] and perturbed [12] hypernetted chain approximations, density functional theory methods [13] [14] and quantum Monte Carlo. [15] [16]

Implementation

The experiment itself involves having a radioactive positron source (often 22Na) situated near the analyte. Positrons are emitted near-simultaneously with gamma rays. These gamma rays are detected by a nearby scintillator.[ citation needed ]

Related Research Articles

<span class="mw-page-title-main">Positron</span> Subatomic particle

The positron or antielectron is the particle with an electric charge of +1e, a spin of 1/2, and the same mass as an electron. It is the antiparticle of the electron. When a positron collides with an electron, annihilation occurs. If this collision occurs at low energies, it results in the production of two or more photons.

<span class="mw-page-title-main">Quantum electrodynamics</span> Quantum field theory of electromagnetism

In particle physics, quantum electrodynamics (QED) is the relativistic quantum field theory of electrodynamics. In essence, it describes how light and matter interact and is the first theory where full agreement between quantum mechanics and special relativity is achieved. QED mathematically describes all phenomena involving electrically charged particles interacting by means of exchange of photons and represents the quantum counterpart of classical electromagnetism giving a complete account of matter and light interaction.

<span class="mw-page-title-main">Positronium</span> Bound state of an electron and positron

Positronium (Ps) is a system consisting of an electron and its anti-particle, a positron, bound together into an exotic atom, specifically an onium. Unlike hydrogen, the system has no protons. The system is unstable: the two particles annihilate each other to predominantly produce two or three gamma-rays, depending on the relative spin states. The energy levels of the two particles are similar to that of the hydrogen atom. However, because of the reduced mass, the frequencies of the spectral lines are less than half of those for the corresponding hydrogen lines.

<span class="mw-page-title-main">Antihydrogen</span> Exotic particle made of an antiproton and positron

Antihydrogen is the antimatter counterpart of hydrogen. Whereas the common hydrogen atom is composed of an electron and proton, the antihydrogen atom is made up of a positron and antiproton. Scientists hope that studying antihydrogen may shed light on the question of why there is more matter than antimatter in the observable universe, known as the baryon asymmetry problem. Antihydrogen is produced artificially in particle accelerators.

<span class="mw-page-title-main">Ionization</span> Process by which atoms or molecules acquire charge by gaining or losing electrons

Ionization is the process by which an atom or a molecule acquires a negative or positive charge by gaining or losing electrons, often in conjunction with other chemical changes. The resulting electrically charged atom or molecule is called an ion. Ionization can result from the loss of an electron after collisions with subatomic particles, collisions with other atoms, molecules and ions, or through the interaction with electromagnetic radiation. Heterolytic bond cleavage and heterolytic substitution reactions can result in the formation of ion pairs. Ionization can occur through radioactive decay by the internal conversion process, in which an excited nucleus transfers its energy to one of the inner-shell electrons causing it to be ejected.

<span class="mw-page-title-main">Electron–positron annihilation</span> Collision causing gamma ray emission

Electron–positron annihilation occurs when an electron and a positron collide. At low energies, the result of the collision is the annihilation of the electron and positron, and the creation of energetic photons:

In solid-state physics, heavy fermion materials are a specific type of intermetallic compound, containing elements with 4f or 5f electrons in unfilled electron bands. Electrons are one type of fermion, and when they are found in such materials, they are sometimes referred to as heavy electrons. Heavy fermion materials have a low-temperature specific heat whose linear term is up to 1000 times larger than the value expected from the free electron model. The properties of the heavy fermion compounds often derive from the partly filled f-orbitals of rare-earth or actinide ions, which behave like localized magnetic moments. The name "heavy fermion" comes from the fact that the fermion behaves as if it has an effective mass greater than its rest mass. In the case of electrons, below a characteristic temperature (typically 10 K), the conduction electrons in these metallic compounds behave as if they had an effective mass up to 1000 times the free particle mass. This large effective mass is also reflected in a large contribution to the resistivity from electron-electron scattering via the Kadowaki–Woods ratio. Heavy fermion behavior has been found in a broad variety of states including metallic, superconducting, insulating and magnetic states. Characteristic examples are CeCu6, CeAl3, CeCu2Si2, YbAl3, UBe13 and UPt3.

Di-positronium, or dipositronium, is an exotic molecule consisting of two atoms of positronium. It was predicted to exist in 1946 by John Archibald Wheeler, and subsequently studied theoretically, but was not observed until 2007 in an experiment performed by David Cassidy and Allen Mills at the University of California, Riverside. The researchers made the positronium molecules by firing intense bursts of positrons into a thin film of porous silicon dioxide. Upon slowing down in the silica, the positrons captured ordinary electrons to form positronium atoms. Within the silica, these were long lived enough to interact, forming molecular di-positronium. Advances in trapping and manipulating positrons, and spectroscopy techniques have enabled studies of Ps–Ps interactions. In 2012, Cassidy et al. were able to produce the excited molecular positronium angular momentum state.

<span class="mw-page-title-main">Positronium hydride</span> Exotic molecule consisting of a hydrogen atom bound to a positronium atom

Positronium hydride, or hydrogen positride is an exotic molecule consisting of a hydrogen atom bound to an exotic atom of positronium. Its formula is PsH. It was predicted to exist in 1951 by A Ore, and subsequently studied theoretically, but was not observed until 1990. R. Pareja, R. Gonzalez from Madrid trapped positronium in hydrogen laden magnesia crystals. The trap was prepared by Yok Chen from the Oak Ridge National Laboratory. In this experiment the positrons were thermalized so that they were not traveling at high speed, and they then reacted with H ions in the crystal. In 1992 it was created in an experiment done by David M. Schrader and F.M. Jacobsen and others at the Aarhus University in Denmark. The researchers made the positronium hydride molecules by firing intense bursts of positrons into methane, which has the highest density of hydrogen atoms. Upon slowing down, the positrons were captured by ordinary electrons to form positronium atoms which then reacted with hydrogen atoms from the methane.

A composite fermion is the topological bound state of an electron and an even number of quantized vortices, sometimes visually pictured as the bound state of an electron and, attached, an even number of magnetic flux quanta. Composite fermions were originally envisioned in the context of the fractional quantum Hall effect, but subsequently took on a life of their own, exhibiting many other consequences and phenomena.

Bose–Einstein condensation can occur in quasiparticles, particles that are effective descriptions of collective excitations in materials. Some have integer spins and can be expected to obey Bose–Einstein statistics like traditional particles. Conditions for condensation of various quasiparticles have been predicted and observed. The topic continues to be an active field of study.

<span class="mw-page-title-main">Breit–Wheeler process</span> Electron-positron production from two photons

The Breit–Wheeler process or Breit–Wheeler pair production is a proposed physical process in which a positron–electron pair is created from the collision of two photons. It is the simplest mechanism by which pure light can be potentially transformed into matter. The process can take the form γ γ′ → e+ e where γ and γ′ are two light quanta.

<span class="mw-page-title-main">True muonium</span> Predicted exotic atom

In particle physics, true muonium is a theoretically predicted exotic atom representing a bound state of an muon and an antimuon (μ+μ). The existence of true muonium is well established theoretically within the Standard Model. Its properties within the Standard Model are determined by quantum electrodynamics, and may be modified by physics beyond the Standard Model.

AEgIS, AD-6, is an experiment at the Antiproton Decelerator facility at CERN. Its primary goal is to measure directly the effect of Earth's gravitational field on antihydrogen atoms with significant precision. Indirect bounds that assume the validity of, for example, the universality of free fall, the Weak Equivalence Principle or CPT symmetry also in the case of antimatter constrain an anomalous gravitational behavior to a level where only precision measurements can provide answers. Vice versa, antimatter experiments with sufficient precision are essential to validate these fundamental assumptions. AEgIS was originally proposed in 2007. Construction of the main apparatus was completed in 2012. Since 2014, two laser systems with tunable wavelengths and synchronized to the nanosecond for specific atomic excitation have been successfully commissioned.

<span class="mw-page-title-main">Angular Correlation of Electron Positron Annihilation Radiation</span> Experimental techniques of solid-state physics

Angular Correlation of Electron Positron Annihilation Radiation (ACAR or ACPAR) is a technique of solid state physics to investigate the electronic structure of metals. It uses positrons which are implanted into a sample and annihilate with the electrons. In the majority of annihilation events, two gamma quanta are created that are, in the reference frame of the electron-positron pair, emitted in exactly opposite directions. In the laboratory frame, there is a small angular deviation from collinearity, which is caused by the momentum of the electron. Hence, measuring the angular correlation of the annihilation radiation yields information about the momentum distribution of the electrons in the solid.

<span class="mw-page-title-main">Buffer-gas trap</span> Device used to accumulate positrons

The buffer-gas trap (BGT) is a device used to accumulate positrons efficiently while minimizing positron loss due to annihilation, which occurs when an electron and positron collide and the energy is converted to gamma rays. The BGT is used for a variety of research applications, particularly those that benefit from specially tailored positron gases, plasmas and/or pulsed beams. Examples include use of the BGT to create antihydrogen and the positronium molecule.

The rotating wall technique is a method used to compress a single-component plasma confined in an electromagnetic trap. It is one of many scientific and technological applications that rely on storing charged particles in vacuum. This technique has found extensive use in improving the quality of these traps and in tailoring of both positron and antiproton plasmas for a variety of end uses.

<span class="mw-page-title-main">John H. Malmberg</span> American physicist

John Holmes Malmberg was an American plasma physicist and a professor at the University of California, San Diego. He was known for making the first experimental measurements of Landau damping of plasma waves in 1964, as well as for his research on non-neutral plasmas and the development of the Penning–Malmberg trap.

<span class="mw-page-title-main">Penning–Malmberg trap</span> Electromagnetic device used to confine particles of a single sign of charge

The Penning–Malmberg trap, named after Frans Penning and John Malmberg, is an electromagnetic device used to confine large numbers of charged particles of a single sign of charge. Much interest in Penning–Malmberg (PM) traps arises from the fact that if the density of particles is large and the temperature is low, the gas will become a single-component plasma. While confinement of electrically neutral plasmas is generally difficult, single-species plasmas can be confined for long times in PM traps. They are the method of choice to study a variety of plasma phenomena. They are also widely used to confine antiparticles such as positrons and antiprotons for use in studies of the properties of antimatter and interactions of antiparticles with matter.

<span class="mw-page-title-main">Vladimir Kocharovsky</span>

Vladimir Kocharovsky is a Russian physicist, academic and researcher. He is a Head of the Astrophysics and Space Plasma Physics Department at the Institute of Applied Physics of the Russian Academy of Sciences and a professor at N.I. Lobachevsky State University of Nizhny Novgorod.

References

  1. Dupasquier, Alfredo E.; Dupasquier, A.; Hautojarvi, Pekka; Hautojärvi, Pekka (1979). Positrons in solids. Berlin: Springer-Verlag. ISBN   0-387-09271-4.
  2. Siegel, R W (1980). "Positron Annihilation Spectroscopy". Annual Review of Materials Science . 10: 393–425. Bibcode:1980AnRMS..10..393S. doi:10.1146/annurev.ms.10.080180.002141.
  3. F. Tuomisto and I. Makkonen (2013). "Defect identification in semiconductors with positron annihilation: Experiment and theory" (PDF). Reviews of Modern Physics . 85 (4): 1583–1631. Bibcode:2013RvMP...85.1583T. doi:10.1103/RevModPhys.85.1583. hdl:10138/306582. S2CID   41119818.
  4. 1 2 3 Jean, Y. C.; Schrader, D. M.; Mallon, P. E. (2002). Principles and Applications of Positron and Positronium Chemistry. World Scientific Publishing Co Pte Ltd.
  5. Eldrup, M.; Lightbody, D.; Sherwood, J. N. (1981). "The temperature dependence of positron lifetimes in solid pivalic acid". Chemical Physics. 63 (1–2): 51. Bibcode:1981CP.....63...51E. doi:10.1016/0301-0104(81)80307-2. S2CID   93631779.
  6. Capponi, S.; Alvarez, F.; Racko, D. (2020), "Free Volume in a PVME Polymer–Water Solution", Macromolecules, 53 (12): 4770–4782, Bibcode:2020MaMol..53.4770C, doi:10.1021/acs.macromol.0c00472, hdl: 10261/218380 , S2CID   219911779
  7. Eldrup, M.; Lightbody, D.; Sherwood, J.N. (1981). "The temperature dependence of positron lifetimes in solid pivalic acid". Chemical Physics. 63 (1–2): 51–58. Bibcode:1981CP.....63...51E. doi:10.1016/0301-0104(81)80307-2. S2CID   93631779.
  8. Tao, S. J. (1972). "Positronium Annihilation in Molecular Substances". The Journal of Chemical Physics. 56 (11): 5499–5510. Bibcode:1972JChPh..56.5499T. doi:10.1063/1.1677067.
  9. S. Kahana (1963). "Positron Annihilation in Metals". Physical Review. 129 (4): 1622–1628. Bibcode:1963PhRv..129.1622K. doi:10.1103/PhysRev.129.1622.
  10. J. Arponen; E. Pajanne (1979). "Electron liquid in collective description. III. Positron annihilation". Annals of Physics. 121 (1–2): 343–389. Bibcode:1979AnPhy.121..343A. doi:10.1016/0003-4916(79)90101-5.
  11. L. J. Lantto (1987). "Variational theory of multicomponent quantum fluids: An application to positron-electron plasmas at T=0". Physical Review B. 36 (10): 5160–5170. Bibcode:1987PhRvB..36.5160L. doi:10.1103/PhysRevB.36.5160. PMID   9942150.
  12. E. Boronski; H. Stachowiak (1998). "Positron-electron correlation energy in an electron gas according to the perturbed-hypernetted-chain approximation". Physical Review B. 57 (11): 6215–6218. Bibcode:1998PhRvB..57.6215B. doi:10.1103/PhysRevB.57.6215.
  13. N. D. Drummond; P. Lopez Rios; C. J. Pickard & R. J. Needs (2010). "First-principles method for impurities in quantum fluids: Positron in an electron gas". Physical Review B. 82 (3): 035107. arXiv: 1002.4748 . Bibcode:2010PhRvB..82c5107D. doi:10.1103/PhysRevB.82.035107. S2CID   118673347.
  14. B. Barbiellini & J. Kuriplach (2015). "Proposed Parameter-Free Model for Interpreting the Measured Positron Annihilation Spectra of Materials Using a Generalized Gradient Approximation". Physical Review Letters. 114 (14): 147401. arXiv: 1504.03359 . Bibcode:2015PhRvL.114n7401B. doi:10.1103/PhysRevLett.114.147401. PMID   25910161. S2CID   9425785.
  15. E. Boronski (2006). "Positron-electron annihilation rates in an electron gas studied by variational Monte Carlo simulation". Europhysics Letters. 75 (3): 475–481. Bibcode:2006EL.....75..475B. doi:10.1209/epl/i2006-10134-5. S2CID   250844357.
  16. N. D. Drummond; P. Lopez Rios; R. J. Needs & C. J. Pickard (2011). "Quantum Monte Carlo Study of a Positron in an Electron Gas". Physical Review Letters. 107 (20): 207402. arXiv: 1104.5441 . Bibcode:2011PhRvL.107t7402D. doi:10.1103/PhysRevLett.107.207402. PMID   22181773. S2CID   14125414.