Brought to you by:

CHARGE SPECTRUM OF HEAVY AND SUPERHEAVY COMPONENTS OF GALACTIC COSMIC RAYS: RESULTS OF THE OLIMPIYA EXPERIMENT

, , , , , , , , , , , , , , , , , , , , and

Published 2016 September 28 © 2016. The American Astronomical Society. All rights reserved.
, , Citation Victor Alexeev et al 2016 ApJ 829 120 DOI 10.3847/0004-637X/829/2/120

0004-637X/829/2/120

ABSTRACT

The aim of the OLIMPIYA experiment is to search for and identify traces of heavy and superheavy nuclei of galactic cosmic rays (GCR) in olivine crystals from stony–iron meteorites serving as nuclear track detectors. The method is based on layer-by-layer grinding and etching of particle tracks in these crystals. Unlike the techniques of other authors, this annealing-free method uses two parameters: the etching rate along the track (Vetch) and the total track length (L), to identify charge Z of a projectile. A series of irradiations with different swift heavy ions at the accelerator facilities of GSI (Darmstadt) and IMP (Lanzhou) were performed in order to determine and calibrate the dependence of projectile charge on Vetch and L. To date, one of the most essential results of the experiment is the obtained charge spectrum of GCR nuclei within the range of Z > 40, based on about 11.6 thousand processed tracks. As the result of data processing, 384 nuclei with charges Z ≥ 75 have been identified, including 10 nuclei identified as actinides (90 < Z < 103). Three tracks were identified to be produced by nuclei with charges 113 < Z < 129. Such nuclei may be part of the Island of Stability of transfermium elements.

Export citation and abstract BibTeX RIS

1. INTRODUCTION

The existence of nuclei with atomic numbers larger than 170 is admissible within the framework of standard quantum electrodynamics (Lieb 2003). The microscopic theory of the atomic nucleus developed in the 1960s describes a stabilizing effect of nuclear shells in the formation of very heavy (VH) nuclei not yet observed in nature (Goeppert Mayer 1964). According to this theory, superheavy nuclei can have, depending on the ratio of the proton and neutron numbers, extended lifetimes and form so-called Islands of Stability (Nilsson et al. 1969).

The search for transfermium nuclei with charges Z > 100 of natural origin is one of the most significant and topical problems closely related to the approbation of the hypothesis of the transfermium elements' Island of Stability, as well as astrophysical models describing the extreme states of matter in the universe (Ginzburg 1999; Ginzburg et al. 2005; Panov et al. 2009; Ter-Akopian & Dmitriev 2015).

Due to peculiarities of nuclear instability, natural elements with nucleic charges around Z ≈ 100 have not been observed under conditions existing on Earth (Oganessian 2001). All currently observed nuclei with charges larger than the charge of uranium (Z > 92) have been synthesized artificially. The first transuranium elements were produced by the reactor method, consisting of the irradiation of 238U nuclei with neutrons. As the result of neutron capture and subsequent β decay, the charge of the initial nucleus increases by unity $({n}^{0}\to {\rm{p}}+{{\rm{e}}}^{-}+{\nu }_{{\rm{e}}})$, yielding nuclei with larger Z.

The development of high-powered nuclear reactors provided for the accumulation of transuranium elements in amounts sufficient to be used as targets for producing nuclei with large Z by irradiation with light charged particles (Ghiorso et al. 1950). This technique enabled the production of artificial elements up to fermium-100, whose isotope 257Fm has a lifetime of 94 days. To date, however, the potential of this method of heavy nuclei production has been exhausted, as excessively large neutron energies and flux densities are required to synthesize novel elements. Estimates have demonstrated that neutron densities on the order of 1019 cm−3 are required for the synthesis of elements heavier than fermium (the largest neutron flux in a nuclear reactor is about 1016 neutron cm−2 s−1) (Sarkisov & Puchkov 2011).

The nuclei of chemical elements heavier than Fm have been produced under laboratory conditions using high-energy ion beams available at accelerator facilities. Transuranium elements have been synthesized at the world's largest laboratories, including the Heavy Ion Linear Accelerator (SuperHILAC), Lawrence Berkeley Laboratory, California, USA (257Rf, 263Sg); GSI Helmholtzzentrum, Darmstadt, Germany (262Bh, 265Hs, 266Mt, 269Ds, 272Rg, 277Cn); the Institute of Physical and Chemical Research (RIKEN), Japan (278Uut) (Schädel & Shaughnessy 2014); and the Laboratory of Nuclear Reactions, Joint Institute for Nuclear Research (LNR JINR), Dubna, Russia (259No, 262Lr, 262Db, 288Uup, 289Fl, 292Lv, 293Uuo, 294Uus) (Mikheev et al. 1967; Oganessian et al. 1984, 2010, 2012, 2013a, 2013b).

The lifetimes of some of the produced nuclei last up to several seconds, or even minutes (see Figure 1).

Figure 1.

Figure 1. Lifetimes of the most stable superheavy isotopes vs. atomic numbers of elements (92 < Z < 118) (Emsley 2001; Shaviv 2012; Khuyagbaatar et al. 2014).

Standard image High-resolution image

The direct synthesis of nuclei with both atomic numbers above 100 and sufficiently large lifetimes under laboratory conditions confirms the possibility of the existence of stable superheavy nuclei. This has stimulated significant interest in searching for them in nature, first and foremost in galactic cosmic rays (GCR).

Specific conditions for the synthesis of superheavy elements can be realized in those areas of the universe where processes of matter evolution, completed within the boundaries of the solar system, are still running. According to the existing concepts, elements of natural origin, from carbon and up in atomic weight, are formed in stellar interiors and at supernova outbursts (Burbidge et al. 1957; Cameron 1957). Nucleosynthesis of superheavy elements heavier than bismuth takes place at a neutron concentration exceeding 1020 cm−3, as a result of the r-process (r for rapid) (Schramm & Fiset 1973; Fryer et al. 2006; Langanke et al. 2011) occurring near the boundary of neutron stability when the rate of neutron capture by nuclei is much larger than the rate of β decay (Kapitonov et al. 2009). It is assumed that superheavy nuclei with neutron numbers up to N = 184 can be generated this way.

In addition, within nonequilibrium shells of neutron stars, the formation of VH nuclei (with mass numbers up to 500) is considered possible at a neutron density on the order of 1030 cm−3 and at temperatures T < 108 K. Ejecta from these stars can lead to the emergence of ultraheavy elements (superheavy—SH group, 50 ≤ Z ≤ 80, ultraheavy—UH group, 80 ≤ Z ≤ 92, and transuranium Z > 92) in interstellar media, stars, and planets (Bisnovatyi-Kogan & Chechetkin 1979; Kramarovskiy & Chechev 1987). Generated in these processes, such heavy nuclei can propagate to considerable distances in intergalactic space, if their lifetimes are long enough, as part of cosmic rays.

The majority of experiments in the search for GCR superheavy nuclei have been performed on balloons (Fowler et al. 1970; Blanford et al. 1971; Price 1973; Cecchini et al. 2003) or in space (Shirk & Price 1978; Fowler et al. 1987; Binns et al. 1989; Font & Domingo 1998; Donnelly et al. 2001; Weaver & Westphal 2002). Despite the large number of these experiments, only several dozen events with Z > 86 were registered, and quite a small number of isolated fragmentary events were attributed by the authors to nuclei with Z > 92.

Although the experiments on satellites can last for several years, exposures in most experiments become inadequate for efficient, statistically significant registration of superheavy nuclei of GCR due to negligibly small values of heavy nuclei fluxes in near-Earth space (∼1–2 heavy nuclei/m2 per year). According to element abundance data obtained in studies with the use of balloons, satellites, and meteorites, the flux of superheavy elements of the lead–bismuth and thorium–uranium groups is 10 orders of magnitude weaker than the flux of hydrogen nuclei.

In the mid-1960s, a method was proposed for studying the fluxes and spectra of heavy and superheavy nuclei in cosmic rays with natural detectors (Maurette et al. 1964), such as the stony–iron meteorites from the pallasite class (Yang et al. 2010) that consist of a porous iron–nickel matrix with inclusions of olivine crystals (Boesenberg et al. 2012; Lavrentjeva et al. 2012).

Olivine is semi-transparent and one of the most abundant silicate minerals in pallasites (up to 65% volume). Over hundreds of million years in cosmic space, olivine crystals are exposed to cosmic radiation to high fluencies. In olivine crystals, projectiles with charges Z > 24 create tracks due to induced structure transformations and broken bonds and preserve them for millions of years. It has been estimated that 1 cm3 of meteoric olivine, hundreds of million years in age, is equivalent (in terms of the number of tracks, i.e., registered events) to an exposure of 1–2 tons of photoemulsion in cosmic space for one year (Flerov & Il'inov 1982).

Investigations of tracks of heavy and superheavy nuclei of cosmic origin in pallasites started in the 1960s (Fleischer et al. 1967; Flerov & Ter-Akopian 1981). In the course of those studies, the method of etching meteorite olivine specimens in combination with annealing was developed (Otgonsuren et al. 1976; Perelygin & Stetsenko 1980; Bondar et al. 1997). Since 2005, tracks in olivine crystals from meteorites have been investigated within the framework of the OLIMPIYA project (Olivines from Meteorites: Search for Heavy and Superheavy Nuclei) (Ginzburg et al. 2005) performed by the collaboration between the Lebedev Physical Institute and the Vernadsky Institute of Geochemistry and Analytical Chemistry of the Russian Academy of Sciences. Small fragments from the Marjalahti meteorite found in 1902 in Finland (size, ∼30 cm; weight, ∼45 kg; age, ∼200 million years) and from the Eagle Station meteorite found in 1880 in the USA (size, ∼25 cm; weight, ∼38 kg; age, ∼70 million years) have been used in this project.

Annealing is sometimes used to avoid effects of temperature fluctuations during the etching and/or to remove a background consisting of tracks of light ions. We did not anneal samples before etching. The reason is that nanometric structure transformations of olivine along the heavy projectile trajectory provide enhanced etching of this region. Figure 3 and simulations made in Gorbunov et al. (2015) demonstrate that the diameter of an emerging amorphized track core is up to about 10 nm in the trajectory sector where the Bragg peak of the electronic stopping of heavy ions is realized. The chemical activity of this track core may be reduced due to recrystallization during annealing. To stimulate such recrystallization within etching, the etching temperature must reach thresholds activating (a) fast diffusion of atoms/structure defects supplying structure modifications at times much shorter than the etching time, or (b) melting of the track core followed by its rapid solidification. Because of olivines' high melting temperatures (1800°C–1850°C) the second scenario cannot be realized at the etching temperatures used (110°C) or during the hand-polishing of samples before etching. Such a temperature increase arising during treatments of samples cannot stimulate a fast diffusion of atoms, either, due to their high migration barriers (e.g., migration barriers of vacancies and interstitials in oxides with covalent binding exceeding 1–2 eV). This is well-illustrated in some experiments (Perelygin et al. 1985) when the procedure of track annealing is applied to study ancient tracks from GCR in olivine crystals from meteorites. Dissipation of "background" tracks of light iron group nuclei from GCR (initial density 1010–1011 cm−3) was detected in Perelygin & Stetsenko (1989) after annealing these crystals at higher temperatures (430 ± 1)°C for 32 hr before etching, and in a 6–8-fold decrease of track lengths for nuclei with Z ≥ 54. This correlates with the analysis (Perron et al. 1988), which demonstrated that the preliminary track annealing led to unpredictable changes in track lengths, resulting in a lower accuracy of nuclear charge determination. For example, path length variations of accelerated Kr and Xe nuclei (with energies of 12.5 and 10.0 MeV per nucleon, respectively), decelerated in olivine crystals from Marjalahti pallasite, depend on annealing time (Lal et al. 1969). The etched lengths of tracks of these nuclei are reduced by 2–3 times for the first 10–20 hr of annealing (382°C). A further increase of annealing time (up to 240 hr) is not followed by any significant decrease in track length, but these final lengths of tracks of Kr ions vary from 18 ± 3 μm to 11 ± 3 μm (40% difference), i.e., the dispersion of the measured lengths is too high. Annealing of tracks of U, Au, and Xe decelerated in olivine crystals from Marjalahti pallasite at temperatures of 430°C, 435°C, and 450°C resulted in a similar distribution of etched track lengths (Perron et al. 1988). Dispersion of L values measured in individual olivine crystals from Marjalahti pallasite sometimes reaches a 3–4-fold value. This effect has been observed, in particular, for tracks from U ions, annealed for 5 hr at a temperature of 450°C, when the L value measured in the same crystals varied within the range of Lmin = (217 ± 52) μm up to Lmax = (762 ± 77) μm. The annealing of tracks from U ions held for 5 hr at T = 435° gave the L values within the range of Lmin = (440 ± 100) μm to Lmax = (869 ± 53) μm (Perelygin & Stetsenko 1989). Similarly, almost twofold intervals of L variation were obtained for Xe and Au ion tracks. Taking into account these causes, the technique without preliminary annealing at a higher temperature is used in the presented work, i.e., we did not apply annealing of samples before their etching at a temperature of 110°C. The search for the heavy component in GCR within the framework of the OLIMPIYA project is based on the registration and measurement of the dynamic and geometric parameters of chemically etched tracks generated by nuclei with Z > 40 in combination with calibration experiments at heavy ion accelerator facilities. The detection method is an annealing-free technique based on layer-by-layer grinding and chemical etching. This technique provides for the geometrical parameters of tracks and the lengthwise track etching rate along the ion trace, as an additional parameter for identification of charges Z of the particle producing tracks.

To date, 11,647 tracks of nuclei with charges Z > 40 have been registered in specimens of the Marjalahti and Eagle Station meteorites. Among the processed tracks, 384 nuclei have been identified as having Z ≥ 75. Three of these tracks were identified as 113 < Z < 129, i.e., they may be attributed to long-lived superheavy elements. According to modern concepts, such nuclei may refer to the Island of Stability—their successful identification in nature justifies efforts to synthesize superheavy elements under Earth conditions.

The content of this article is organized as follows.

Section 2 contains a description of the experimental technique.

Section 2.1 discusses olivine crystals from meteorites as natural track detectors.

Detected structure changes and the possible mechanism of chemical activation of olivine during the relaxation of its excited electron subsystem in tracks are presented in Section 2.2, including the results of Monte Carlo modeling of the electron kinetics.

Section 2.3 contains the description of the mechanical and chemical treatment of olivine crystals intended to reveal tracks of charged particles.

Section 2.4 refers to measurements of nuclear track parameters: track identification (2.4.1), PAVICOM image processing (2.4.2), etching rate measurements (2.4.3). Section 2.5 refers to the results of the calibration experiments associating the detected track parameters with the charge of a projectile.

Section 3 presents an extensive database of the abundance of GCR nuclei with Z > 48 obtained by the described techniques (3.1) with special attention to registered transfermium nuclei (3.2) and nuclear lifetimes' estimates (3.3).

Section 4 sums up the results, with N emphasis on those that testify to the existence of transfermium elements in nature.

2. DETAILS OF EXPERIMENTS AND METHODS APPLIED

2.1. Olivine Crystals from Meteorites as Natural Track Detectors

Olivine is distributed through pallasites in the form of crystals ranging from several millimeters up to 1–2 cm in size (Figure 2). According to current concepts, pallasites are either a product of unfinished differentiation of matter into the silicate phase and metal in the gravitation field, or a result of agglomeration of a silicate substance with iron. There are several models describing the formation of pallasites, including crystallization near the surface of an externally heated asteroid, crystallization of an impact melt, and nebular condensation (Goldstein et al. 2014).

Figure 2.

Figure 2. Marjalahti (large, 1) and Eagle Station (small, 2) pallasite fragments used in the OLIMPIYA project.

Standard image High-resolution image

By its crystal structure, olivine ((Mgx Fe1-x)2SiO4) is attributed to the silicates with an isolated island arrangement (nesosilicates) of silicon–oxygen tetrahedra (SiO4) bonded by means of Mg or Fe cations (Birle et al. 1968).

When passing through matter, heavy ions (M > 20 amu) with specific energies E > 1 MeV nucleon−1 lose the majority of their energy (>95%) due to the excitation of the electron subsystem of a material. Subsequent relaxation of the extremely excited electron subsystem, accompanied by the transfer of part of its excess energy to the lattice, leads to formation of ion tracks. The sizes of these tracks are a few nanometers in diameter and many tens or more micrometers in length. The tracks consist of materials that have been highly damaged (structural changes, broken bonds, etc.), resulting in an elevated level of chemical activity as compared with the undamaged surrounding matrix (Durrani & Bull 1987). The selectivity of the etching process (the etching rate of a track being significantly higher than that of the bulk) converts each individual track into an open channel (Fleischer et al. 1975; Kashkarov et al. 2009) that can be easily identified by optical microscopy.

It has been shown (Egorov et al. 2008, 2011; Aleksandrov et al. 2009b) that the etching efficiency of long path-length tracks in olivine crystals is the same for the polycrystalline highly oriented regular texture and the single-crystal texture, i.e., the etched track length and the etching rate do not depend on the orientation of tracks relative to the olivine crystallographic axes. This feature is of great significance for processing the experimental data because it enables precise measurements of etched GCR nuclear track parameters, as well as calibration measurements of tracks produced by high-energy nuclei at the accelerator facilities, without any additional crystallographic analysis.

The meteoritic olivine crystals used in this work were located relatively close (up to 5–6 cm depth) to the preatmospheric meteoroid surface. Thus, it is assumed that these crystals could register nuclei of superheavy GCR elements with energies up to ∼1.5–5 GeV/nucl (Aleksandrov et al. 2009a). According to estimates based on phenomenological models (Pellas et al. 1983; Alexandrov et al. 2013a), 1 cm3 of olivine crystals at a depth of about 5 cm from the preatmospheric meteoroid surface can contain 102–103 tracks of nuclei with Z > 90. Moreover, within a surface depth of about 1 cm, up to 104 of ion tracks of the VH (20 < Z < 30) group may have been accumulated over 108 years of exposure in cosmic space.

Unlike the other types of track detectors (nuclear emulsion, plastic), olivine is free from tracks of light nuclei with Z ≤ 24, because the energy-loss threshold for creating etchable tracks is sufficiently high in this material (about 18 MeV cm−2 mg−1 or 5 keV/nm) (Horn et al. 1967). Taking this threshold into account, only etched tracks from nuclei heavier than the iron group (24 ≤ Z ≤ 28) can be detected in olivine. This high threshold of the projectile charge for etched track developing in olivine enables studies of the nuclear charges of very very heavy (VVH group, 30 ≤ Z ≤ 50), superheavy (SH group, 50 ≤ Z ≤ 80), ultraheavy (UH group, 80 ≤ Z ≤ 92), and transuranium (Z > 92) elements in GCR. Glass track etch detectors can also be insensitive to light nuclei, but only meteorites are irradiated with cosmic rays for millions of years.

To date, supernovae are considered to be the most probable sources of heavy elements in GCRs (Ginzburg 1999). The evaluation of minimal possible lifetimes of those nuclei which have produced tracks in meteorites is discussed in Section 3.3.

2.2. Structure Changes and the Possible Mechanism of the Chemical Activation of Olivine in Ion Tracks

Structure transformations generated in the vicinity of ion trajectories were investigated using transmission electron microscopy (TEM). A TITAN 80-300TEM/STEM (FEI, US) operating at 300 kV was applied for this purpose.

For these investigations, thin electron-transparent foils of olivine along the trajectories of ion tracks were prepared using the focused ion beam (FIB) technique in a HELIOS dual-beam SEM/FIB system (FEI, US) equipped with C and Pt gas injectors and a micromanipulator (Omniprobe, US). A 2 μm Pt layer was deposited on the surface of the specimen prior to the preparation of cross-sections. Sections approximately 8 × 5 μm2 in size and 2 μm thick were cut by 30 kV Ga+ ions, removed from the sample, and then attached to an Omniprobe semi-ring (Omniprobe, US). Final thinning was performed with 5 kV Ga+ ions followed by cleaning by 2 keV Ga+ ions to electron transparency.

High-resolution images as well as conventional bright-field TEM images were recorded.

The diameters of structurally changed regions were measured from the recorded bright-field images and high-resolution TEM (HRTEM) images. HRTEM images of 2 GeV Au ion-induced tracks in plain view were obtained by carefully tilting the specimen to align the tracks parallel to the incident electron beam.

Figure 3(a) presents a typical HRTEM image of an ion track. The olivine crystal lattice was observed in [110] zone axis, and a fast Fourier transform (FFT) obtained from that image is shown in Figure 3(b). Close inspection of the track core shown in Figure 3(c), along with FFT analysis, indicated that the core is mostly amorphous, which is in good agreement with some other works (Szenes et al. 2010; Rodriguez et al. 2014; Park et al. 2015). However, in some areas of the track core, periodic features could be revealed in orientations that differed from the olivine crystal lattice. Because the scale bar can be precisely calibrated by the known olivine d-spacings (e.g., d001 = 5.99(7) Å; Heuer 2001), the radii of individual tracks can be measured reliably.

Figure 3.

Figure 3. HRTEM image of ion tracks in olivine irradiated with 11.1 MeV nucleon−1 Au ions. (a) Track cross-section, (b) FFT pattern of the olivine matrix in [110] zone axis, (c) enlarged image of the track core with FFT pattern from marked area (inset).

Standard image High-resolution image

These results demonstrate that considerable structural transformations occur only in the close (<10 nm) vicinity of the projectile trajectory. Therefore, these structure changes cannot stimulate an enhancement of the etching rate of olivine at distances ∼1 μm from the trajectory, which results in the detected syringe-like shape of GCR nuclear tracks (see Figure 7(a), (b) below).

We suggest that the chemical activation of olivine at such large distances from the ion trajectory is caused by spreading and specific relaxation of electronic excitations generated in the track (Gorbunov et al. 2013, 2014, 2015). It is assumed that this activation results from the neutralization of metal atoms linking SiO4 atomic groups that break up bonds providing this linking. In particular, we think that the change of the oxidation degree of polyvalent Fe cations by fast electrons appearing due to the excitation of the electronic subsystem of a target by a projectile is the most probable mechanism of the chemical activation of investigated iron-bearing olivine.

Indeed, Monte Carlo (MC) simulations based on the well-tested original TREKIS code describing the initial kinetics of excitation of the electronic subsystem of solids in swift heavy ion tracks (Medvedev et al. 2015) demonstrate that fast electrons generated in ion tracks in olivine may spread up to micrometers from the ion trajectory. Figure 4 illustrates this spreading at 100 fs after an impact of 11.1 MeV nucleon−1 Au ion realizing the electronic stopping around the Bragg peak.

Figure 4.

Figure 4. Radial distribution of fast electrons at 100 fs after the impact of 11.1 MeV nucleon−1 Au ion in olivine (Gorbunov et al. 2015).

Standard image High-resolution image

Within the framework of the activated complex theory (Connors 1990; Berg et al. 2002) the change of the specific Gibbs energy determines the level of chemical activation of a modified material.

Figure 5 (Gorbunov et al. 2015) illustrates the increase of chemical activity (the relative reaction rates) of olivine due to changes of the Gibbs energy in the vicinity of the trajectory of a gold ion having the energy E = 11.1 MeV/nucl. At distances less than 5 nm from the trajectory, the Gibbs energy of the material increases due to structure transformations stimulated by relaxation of the energy and momentum transferred into the olivine lattice during the relaxation of the excited electron subsystem. These transformations are confirmed by TEM images (Figure 3), as well as by simulations of the coupled kinetics of the relaxing electronic subsystem and lattice in swift heavy ion tracks in olivine (Gorbunov et al. 2013, 2014). Chemical activation of olivine at distances r > 5 nm from the ion trajectory is caused by neutralization of Fe ions binding SiO4 tetrahedra in an olivine lattice by fast electrons generated in an ion track.

Figure 5.

Figure 5. The relative reaction rate caused by strong structure transformations in the nanometric track core and Fe atoms reduction by fast electrons generated in a 11.1 MeV nucleon−1 Au ion track in olivine. The Gibbs energy of an Fe-reduced atom is taken to be equal to the free energy of Fe vacancies in Fe-bearing olivine (Gorbunov et al. 2015); kT is the reaction rate constant of the track, kB is that of the undamaged material (bulk).

Standard image High-resolution image

2.3. Primary Treatment of Olivines

Due to numerous microcracks, only about a quarter of olivine crystals up to 2 mm in size extracted from the (Fe, Ni) pallasite matrix are sufficiently transparent for track analysis by optical microscopy.

Within the framework of the OLIMPIYA project, a special method of multi-stage grinding and etching of a certain thickness of crystal layer was developed yielding the most precise detection of the parameters of etched tracks and efficient use of each chosen crystal volume.

First, polishing is applied because the quality of the crystal surface determines the accuracy of geometrical track parameter measurements. After the first cut, the resulting surface of the crystal has a roughness of 20 ± 10 μm. Grinding and polishing are performed up to a roughness no larger than 3–5 μm and 1–2 μm, respectively. Such quality is achieved by polishing with diamond pastes using powder grains decreasing successively from 14 μm down to 1 μm.

Sets of polished crystals, several pieces per set, are embedded into epoxy tablets of a size corresponding to the slot on the microscope table.

The presented work makes use of the common standard WN technique for chemical track etching in olivine crystals from pallasites. This technique was first described in Krishnaswami et al. (1971) and refined at the Flerov Laboratory of Nuclear Reactions, Joint Institute for Nuclear Research (Petrova et al. 1995). The WN solution is prepared in the proportion of 40 g of trilon B (disodium salt of ethylenediaminetetraacetic acid), 1 ml of orthophosphoric acid, and 1 g of oxalic acid for 100 ml of water. For optimum pH = 8.0 ± 0.1, the obtained original solution is gradually supplemented with aqueous NaOH (4:3), the amount of which is determined empirically for each solution preparation. The solution is usually prepared in a large volume (1–2 liters), and all the subsequent etching operations occur in almost identical conditions. The temperature of the etching solution is maintained with an accuracy of about 1% and is +(110 ± 1)°C at positive solution vapor pressure (up to ∼1.5 atm). It has been established empirically that (a) this mode provides constant concentration of the solution and (b) there is no need to bring the solution up to its boiling point. As a result, the etching process occurs under almost constant conditions without any variations over a very long (up to ∼100 hr) etching time. The etching time is determined by the necessity to detect, with utmost accuracy, the value of the track etching rate changing along the path of a charged particle. The minimum etching time is 8 hr; after checking whether it is possible to search for high-velocity etching tracks, the procedure is continued, with checks at 12, 24, and 36 hr.

Olivines from meteorites are able to conserve heavy GCR nuclear tracks for hundreds of millions of years. Experiments show that, at low temperatures in outer space, tracks in olivine can persist much longer than 1010 years (Maurette et al. 1964; Fleischer et al. 1967). Such resistance of olivine allows it to conserve tracks at a relatively low temperature of (110 ± 1)°C throughout the etching process in a boiling solution for several tens of hours (Goswami et al. 1984). Moreover, the tracks survive polishing of the crystals by hand at room temperature. In this regard, determination of the nucleus charge does not require any contributions from track temperature characteristics at space exposure, as well as at olivine processing in the laboratory.

2.4. Measurements of Nuclear Track Parameters

2.4.1. Identification of Tracks

At high energies, the energy loss of a projectile is not high enough to produce sufficiently severe damage for tracks to be etchable. During the slowing down of a particle, its energy loss along the trajectory gradually increases, finally surpassing the threshold (about 18 MeV cm−2 mg−1 or 5 keV/nm in olivine) for the creation of etchable tracks. At higher energy losses, the track etching rates are higher. Within the region of a maximum energy loss (Bragg peak) the high etching rate leads to a maximum diameter of the etched channel. The cylindrical section of the etched track demonstrates a fivefold increase of the etching rate (Perron & Maury 1986; Perron & Bourot-Denise 1986) in this region. Close to the stopping end of the particle trajectory, the electronic energy loss drops below the etching threshold, and the channel ends with a short, sharp tip. This form of an etched track is illustrated in Figure 6.

Figure 6.

Figure 6. Top: electronic energy loss of a projectile. Bottom: geometry of various parts of etched track. The relation between the lengths of the wide and narrow parts of the etched track differ for different ions, according to the different electronic energy losses of these ions.

Standard image High-resolution image

Micrographs of two types of etched tracks in olivine confirming the scheme of Figure 6 are presented in Figure 7. Figure 7(a), (b) (magnification 40×) refers to final segments of the trajectory of an incident particle. Therefore, we assume further that an etched track consists of two regions: syringe-like, with a cylindrical part at the end of the path (where the energy losses are maximum); and needle-like, corresponding to the high-energy segment of the trajectory. About 2% of the tracks in the total data set are syringe-like. On the other hand, needle-like tracks are predominant (98% in the total data set) which, it seems, can be described statistically by considering the small sizes of olivine crystals compared to the long paths of high-energy particles. Figure 7(c) (magnification 20×) and Figure 7(d) (magnification 40×) demonstrate both types (syringe and needle-like tracks) in the same crystal; Figure 7(e) (magnification 40×) demonstrates a needle-like track; Figure 7(f), presents the general view of an olivine crystal with a few heavy projectile tracks (magnification 8×).

Figure 7.

Figure 7. Micrographs of etched tracks of GCR superheavy nuclei in olivine crystals from pallasites. Fields of view: (a) 220 × 280 μm2; (b) 220 × 280 μm2; (c) 444 × 555 μm2 and (d) 220 × 280 μm2 (the same crystal); (e) 220 × 280 μm2; (f) 1360 × 1090 μm2.

Standard image High-resolution image

Being exposed to cosmic radiation, olivine crystals accumulate tracks from nuclei with trajectories coming in from different directions. During etching, only tracks whose trajectories intersect the specimen surface are affected by the etchant.

Successive step-by-step etching has demonstrated that the etching rate Vetch (that is, the increment of the length of a track divided by the time of etching) considerably decreases when the tracks achieve a length of 100–200 μm (Bagulya et al. 2009). At very long etching times (up to 96 hr), tracks of 300–400 μm either continue to slowly lengthen or remain invariable.

The uncertainty associated with this effect of an uncontrolled decrease of Vetch can be ruled out with the technique of multi-stage slicing of layers at the layer-by-layer etching of the crystal (Kashkarov et al. 2008; Aleksandrov et al. 2010). After etching and analysis of the crystal surface, a thin layer of 50–100 μm is removed from the crystal by grinding (to an accuracy of several microns), and the procedure is repeated. The thickness of the layer removed by precision grinding and polishing is established for each particular set of observed tracks. After each etching step, the geometrical parameters (length and diameter) of the developed etched tracks at a given processing stage are measured. The successive stages of the treatment are shown schematically in Figure 8.

Figure 8.

Figure 8. Scheme of stepwise cutting and etching of a given track-containing olivine crystal. The thickness of the cut layer is d = 30–100 μm.

Standard image High-resolution image

Recording the data of etched track segments in different layers allows us to reconstruct the full track length throughout all processed layers (Figure 8, tracks 1 and 3). Such reconstruction requires an accurate determination of the coordinates of the track ends and angles of inclination relative to the etched surface. In order to produce this reconstruction, the PAVICOM tool is applied.

2.4.2. PAVICOM Image Processing

The search for tracks of heavy and superheavy charged particles of GCRs crucially depends on the ability to identify individual particle tracks and separate them from the background, as well as on the accuracy of the track geometry data in the (x, y ) plane and along the depth (z) of the crystal.

The (x, y , z) coordinates of the etched tracks were analyzed by PAVICOM (Completely Automated Measuring Complex) (Aleksandrov et al. 2004, 2012; Polukhina 2012)—a multi-functional programmable measuring facility developed for processing data obtained in various types of solid-state track detectors.

The software identifies the etched tracks because they appear as dark objects in the transparent olivine crystal (Aleksandrov et al. 2009b). In a fully automated process, PAVICOM determines the coordinates of track segments and the angle of inclination of a track with respect to the crystal surface, thus reconstructing the complete track length within all processed layers. Automatic scanning of the track elements is performed by means of optical microscopy combined with a movable sample stage (MiCos, Germany): MS-8 (working range, 205 × 205 mm; lateral accuracy, 0.5 μm) and LS-110 (working range, 305 mm; depth accuracy, 0.2 μm). A CMOS camera is fixed on the z-axis stage, allowing variation of the focus at different depths with steps of about 3 μm. The images are recorded by a 1.3 MP camera with a frame rate of 376 fps while the xy stage is at rest and the z stage moves at constant velocity (the so-called Stop-and-Go motion algorithm). The scanning speed of this system is about 20–80 cm2/h−1. Mathematical processing of the digitized images is performed by using a library of C++ programs.

A problem associated with automated processing is the change of the optical properties of the etched tracks at different depths of view. Near the crystal surface, the etched track is rather dark and has clear-cut and sharp edges. However, when going deeper, the tracks become increasingly pale and blurred, and it becomes difficult to discriminate them from background features. In order to reliably determine the track parameters deep in the crystal, several parameters in the analysis program are adapted to the depth of the crystal by automatically changing, e.g., filters, or geometrical parameters of etched tracks.

For data processing, it is also taken into consideration that etched tracks represent open channels or cavities with uneven walls. The image of an etched track often looks like a series of dark, light, and very light spots. In particular, the image of the tracks under a small angle of incidence of a projectile with respect to the crystal surface is subjected to greatest distortions due to optical effects. As a result, these images, scattering into separate clusters, have segments of elongated shape with the direction of the axis close to the direction of track's axis. When reconstructing the track, these features are taken into account by extending the axis of an elongated cluster and considering other clusters close to this axis (Figure 9).

Figure 9.

Figure 9. Optical microscope image (300 μ× 400 μm) of etched olivine crystal before (a) and after (b) processing by the PAVICOM image recognition program. The lines designate the track axes.

Standard image High-resolution image

The track analysis is difficult in some olivine crystals because of numerous extraneous spots from various sorts of inhomogeneities, microcracks, and optical effects (reflection, refraction, focusing etc.). To reduce this background and to increase the efficiency of track identification, the software is equipped with the ability to isolate an area of the image where the signal-to-noise ratio is most favorable. Spot-like features are separated from true tracks by a complete analysis through the entire crystal depth: the image of a real track shifts with depth in the direction of the particle trajectory, whereas the positions of extraneous objects remain fixed at a certain depth in the crystal. This additional test enables a more efficient selection and identification of real tracks.

2.4.3. Etching Rate as an Important Track Parameter

The main and only parameter measured after chemical track etching of olivine crystals, used in previous experiments on the search for heavy nuclei in meteorites (Otgonsuren et al. 1976; Flerov & Ter-Akopian 1981), was the length of etched tracks, i.e., the segment of charged particle deceleration path along which a region with increased chemical activity is formed. Within the approach developed in the OLIMPIYA experiment, we record the etching rate Vetch along the projectile trajectory as the second main track characteristic used for determination of the nuclear charges (Aleksandrov et al. 2008). This additional parameter provides quite essential information, because the crystals accessible for the experiments are sometimes too small to observe the complete length of tracks. Particularly for VH nuclei, the track length may exceed the size of a crystal and even multiple step-wise etching does not allow the measurement of the complete track length by the available technique (Figure 10). For example, the full length of an etched uranium track reaches 10 mm (at U energy of about 400 MeV nucleon−1), whereas the average thickness of a crystal is at least 10 times smaller.

Figure 10.

Figure 10. Scheme of a long etched track (e.g., like that where evaluation gives Z = 119 for a projectile, see Section 3.2; two other tracks with Z = 119 are of a needle-like form) in an olivine crystal from pallasite. L(B) denotes the needle-like length of the etched track, L(A–B) corresponds to a transition zone between B and A, and L(A) is the cylindrical section of the track observed as a syringe (Alexandrov et al. 2013b).

Standard image High-resolution image

Experiments with ion beams of various elements of fixed energy available at accelerator facilities were carried out within the framework of this project. They demonstrated (Alexandrov et al. 2013a) that the etching rate Vetch along the projectile trajectory changes noticeably following changes in the energy loss of a projectile, as well as in the spectrum of electrons generated in a track. For instance, for U ions, the track etching rate Vetch = (7 ± 1) μm h−1 at the trajectory fragment corresponding to the residual path length R ≅ 1000 μm increases up to (25 ± 2) μm h−1 at R ≅ 200 μm.

In general, because the projectiles are not always stopped in a crystal, the measured segment L of the nucleus deceleration path does not coincide with the residual path length R, so it might be either LR or L < R. The former case is attributed to tracks etched up to the stopping point. In the latter case, the projectile passes through the crystal without being stopped and the analysis of the registered segment of the track must be based on accurate measurements of the etching rates. In this case, use of the developed method of multi-stage etching process and the automated track parameter measurements make it possible to determine Vetch with good accuracy.

It should also be noted that most tracks exit from olivine before the particle comes to a halt. Due to this, track lengths L registered in our experiments are, in most cases, less than the values of R, and the measured parameters L and Vetch give the lower estimates of nuclear charges.

In summary, in combination with the results of the calibration experiments (see the next section) these two parameters of etched tracks in olivine crystals from meteorites, i.e., the measured length of the etched track L and the etching rates Vetch, form a database which can be used for determination of GCR nuclear charges.

2.5. Calibration Experiments and Nuclear Charge Determination

Using the above-mentioned parameters (L and Vetch, see Section 2.4) charges of nuclei Z can be uniquely determined only when (ZVetchL) dependence is determined from the calibration experiments.

A series of such calibration irradiations were performed at the UNILAC and SIS accelerators (GSI Helmholtz Centre for Heavy Ion Research, Darmstadt, Germany) with Kr, Xe, Au, Bi, and U ions, as well as at the accelerator complex of the Institute of Modern Physics (Lanzhou, China) with Bi ions.

The calibration campaign included the irradiation of olivine crystals from Marjalahti and Eagle Station pallasites, as well as Earth olivine samples, with Kr, Xe, Au, Bi, and U ions of 11.1 MeV nucleon−1 and 150 MeV nucleon−1 (for U ions), and Bi with energies of 2.5, 4.5, and 9.4 MeV nucleon−1. The irradiations were made at incident angles of 90° and 45° to the normal of the sample surfaces, with fluences varied between 105 and 1012 ions cm−2.

For uranium, the data obtained by Perron & Maury (1986) were also used.

Figure 11 presents optical micrographs of olivine crystals containing various almost-parallel etched ion tracks. The detected density of registered tracks varies from 30 to 80 tracks per crystal.

Figure 11.

Figure 11. Optical micrograph (∼500 × 700 μm2) of olivine with etched tracks from (a) 11.1 MeV nucleon−1 Xe ions (measured lengths, 67 ± 6 μm; in accordance with SRIM-2008, 75 ± 5 μm); (b) 150 MeV nucleon−1 U ions (measured lengths, 91 ± 5 μm; in accordance with SRIM-2008, 89 ± 5 μm).

Standard image High-resolution image

As an example, some of the calibration results are presented in Table 1 and Figure 12. The majority of our results are in good agreement within the errors with the calculations by SRIM-2008. The agreement of the results at energies of 2.5 and 4.3 MeV nucleon−1 is worse, which may be associated with a wider spectrum shape after particle deceleration to the desired energy.

Figure 12.

Figure 12. Distribution of track lengths for 605 etched tracks of 2.5 MeV nucleon−1 Bi ions in olivine. The arrow marks the calculated length.

Standard image High-resolution image

Table 1.  Results of the Calibration Experiments

Projectile Beam Energy (MeV nucleon−1) Etching Rate (μm h−1) Measured Track Length (μm) Calculated Track Length (μm) (SRIM-2008)
Kr 11.1a 0.5–1 71 ± 5 75
Xe 11.1a 3–4 67 ± 6 ≈75
Au 11.1a 18 ± 3 69 ± 6 ≈67
Bi 2.5b 19.4 ± 4 26.6 ± 2.9 ≈25
  4.3b 19.2 ± 4 47.5 ± 1.1 ≈40
  9.5b 19.0 ± 4 56 ± 1.8 ≈60
  11.1a 18.8 ± 4 81 ± 4.8 ≈80

Notes.

aIrradiations in GSI, Darmstadt. bIrradiations in IMP, Lanzhou.

Download table as:  ASCIITypeset image

The experimentally measured lengths of tracks in the irradiated olivine crystals show a good agreement with the corresponding values calculated using the SRIM-2008 code (Ziegler et al. 1985, 2008).

Based on the results of the calibration experiments, the charge of the projectiles can be plotted as a function of the track lengths and track etching rates measured in the experiments (Figure 13). The accuracy of charge determination by this technique ranges in ±1 to ±2 charge units.

Figure 13.

Figure 13. The charge Z of an incident ion as a function of the measured track length L and track etching rate Vetch. This dependence was obtained by fitting the data from the calibration experiments.

Standard image High-resolution image

Within the charge range of 67 < Z < 93, this calibration surface can be fitted by a five-parameter function given by Equation (1),

Equation (1)

The fitting parameters A, B, C, D, and E of the function depend smoothly on charge Z and were approximated by a linear function with a small quadratic term. This allows a good extrapolation of charge values to several units beyond the limits of the calibration experiments (Z > 92).

Figures 14 and 15 present the results of the calibration experiments and their fitting by Equation (1).

Figure 14.

Figure 14. The experimental dependence Vetch on the track length L for U nuclei with maximum energy 150 MeV/nucleon−1 in olivine fitted by Equation (1). Each experimental data point corresponds to a specific ion energy, which can be deduced from the dE/dx vs. range dependence, e.g., by using SRIM. The inserted figure presents the experimental track-etching rate Vetch as a function of track length for Fe (1, triangles), Kr (2, squares), and Xe (3, circles) ions fitted by Equation (1).

Standard image High-resolution image
Figure 15.

Figure 15. The family of Vetch(Z, L) curves obtained by fitting the experimental data from calibration experiments by Equation (1) and extrapolating the fitted values to all charges from Z = 54 up to Z = 92 (filled circles) and for larger charges from Z = 94 up to Z = 100 (empty circles) (the intervals between curves are two charge units). Insufficient experimental data prevent extending this dependence to the entire V(L, Z) surface. For this reason, we restrict ourselves to illustrating the region of Z ≤ 100.

Standard image High-resolution image

3. RESULTS AND DISCUSSION

3.1. Estimates of Charge Abundance

The application of our layer-by-layer grinding and etching method, in combination with the system of automatic detection of etched tracks (PAVICOM), has allowed us to analyze a large set of experimental track data. Within the OLIMPIYA project, a total of 442 crystals of meteoritic olivine crystals (264 crystals from the Marjalahti meteorite and 178 crystals from the Eagle Station meteorite) were investigated. Using the technique of projectile-charge determination described in Sections 2.4 and 2.5, and based on the data on the track length and the track-etching rate, in combination with calibration results, 11,647 tracks were assigned to GCR nuclei having charges Z > 20 (in accordance with Otgonsuren (1973), uranium content in terrestrial olivines is 10−11–10−12 g/g, whereas the concentration of uranium in the surrounding rock minerals is about 10−7–10−9 g/g, and occasional spontaneous or induced fission tracks in meteorite olivine was evaluated at the same level of uranium content, 10−11–10−12 g/g). Out of them, 10,283 tracks were produced by nuclei with Z ≥ 50, 1233 tracks by nuclei with Z ≥ 70, 384 tracks with charges assessed as Z ≥ 75, and 9 tracks by nuclei with Z ≥ 88. Compared to other studies, the OLIMPIYA experiment has yielded a huge number of GCR tracks, providing one of the best-sampled data sets (see Table 2).

Table 2.  Registered Events of Heavy and Superheavy Nuclei in Various Experiments

  Z Interval Ariel 6 (1) HEAO-3 (2) UHCRE (3) OLYMPIYA
1 Z ≥ 50 412 362 10283
2 50 ≤ Z ≤ 58 240 204 5612
3 60 ≤ Z ≤ 68 84 34 2814
4 Z ≥ 70 88 62 2567 1233
5 70 ≤ Z ≤ 73 29 10 715
6 74 ≤ Z ≤ 80 29 42 449
7 81 ≤ Z ≤ 86 27 10 59
8 88 ≤ Z ≤ 103 3 0 35 9
9 Z ≥ 92 2 0 4

Note. Others experiments: (1) Fowler et al. (1987), (2) Binns et al. (1985), (3) Donnelly et al. (2012). In this table, a dash indicates that no information is available.

Download table as:  ASCIITypeset image

Currently, there are three sets of results obtained in the largest satellite experiments registering cosmic heavy ions: ARIEL-6 (Fowler et al. 1987), HEAO-3 (Binns et al. 1985), and UHCRE (Donnelly et al. 2012). These experimental data sets available to date and our results are presented in Figure 16, showing the relative abundances of nuclei with Z ≥ 48; the data were normalized by the content of iron nuclei A(26Fe) = 106. The nuclei charge determination accuracy in OLYMPIYA is determined by the geometric parameters of tracks, measured on the automated microscope with precision German mechanics (it is 0.5 μm for all coordinates). ARIEL-6, HEAO-3m and UHCRE data show a peak at Z = 79, 77, and 78, respectively, and the height of the UHCRE peak is two times higher than the others. The existing disagreement between the available results requires further research. The results of this work within the errors have the best agreement with HEAO data, although the decline of the distribution is steeper in the region of Z > 80.

Figure 16.

Figure 16. Relative abundance (the data were normalized by the content of iron nuclei A(26Fe) = 106) of GCR heavy nuclei registered in the OLIMPIYA experiment (crosses, with statistical errors) as compared with the results of other experiments: ARIEL-6 (triangles) (Fowler et al. 1987), HEAO-3 (squares) (Binns et al. 1989), and UHCRE (diamonds) (Donnelly et al. 2012). The inset shows three transfermium nuclei registered in this work. The abundance evaluation of these three nuclei (with charge 113 < Z < 129) for inset figure is A = 0.015+0.042−0.003 with level of confidence 95% based on the rare-event processing method (Gehrels 1986). In the region of Z < 56, the data of the OLIMPIYA experiment are statistically less significant because the inadequate length of the corresponding tracks in olivine makes them difficult to measure.

Standard image High-resolution image

Registrations of transuranium nuclei with charges of 92 < Z < 100 were mentioned in experiments by Shirk & Price (1978), Donnelly et al. (2001), and Perelygin et al. (2003a, 2003b). Unfortunately, these papers do not comment on the nature of these events.

It should also be noted that transuranium nuclei with charges of 92 < Z < 100 registered in ARIEL-6 and UHCRE experiments have lifetimes that would be too short to reach our solar system, starting from the nearest supernova.

We think that events with Z > 92 appear, not due to inaccuracies of the applied method or instrumental failures, but as the result of the fragmentation of heavier, more stable nuclei from the region of the Island of Stability.

Analyzing the superlong particle tracks in meteoritic olivines, we also found several events that can be attributed to particles with Z > 92.

3.2. Registration of Transfermium Nuclei

Three superlong tracks with lengths of more than 500 μm and etching rates exceeding 35 μm h−1 were detected when analyzing the data accumulating during the OLIMPIYA experiment, see Figures 17(a)–(d) below.

Figure 17.

Figure 17. (a) Syringe superheavy nuclei track image, 444 × 555 μm2; (b) scheme of this track in olivine; (c) Needle-like long nuclei track image (starting in the bottom right corner, finishing in the top left corner), 444 × 555 μm2; in the middle part of the image, a crystal defect; this olivine crystal was irradiated by U with energy of 150 MeV/n at 45°; after that, the crystal was etched for 36 hr; the mean U track lengths are 132 ± 26 μm. All accelerated projectiles are in one direction, right to left; (d) Needle-like long nuclei track image (starting at the top left corner, finishing in the bottom right corner), 444 × 555 μm2; this olivine crystal was irradiated by U with energy of 150 MeV/n at 45°; after that, the crystal was etched for 36 hr, the mean U track lengths are 132 ± 26 μm. All accelerated projectiles are in one direction, right to left.

Standard image High-resolution image

Taking into account that the maximum value of the experimentally measured etching rate for tracks of uranium nuclei in olivine before their stoppage point is Vetch,U = 26 ± 1 μm h−1, this indicates that the charges of these registered nuclei significantly exceed the charge of uranium.

The charge assessment for these nuclei is based on the dependence of the etching rate V near the stoppage point of the charge value. For nuclei with Z ≤ 92, we used the empirical function V(L, Z) to determine, by interpolation, the charges of the nuclei, Parameters of this function were determined by comparison with experimental data on nuclei up to Z = 92. At present, we cannot pronounce upon how well this function works for Z > 92 because insufficient experimental data make it impossible to extrapolate the function to this region. In any event, the curves obtained for higher Z, up to 100, are shown in Figure 15, as illustrated. The parameter V was used for charge determination near the stoppage point to evaluate our data in the larger charge region. The etching rate V is almost independent from the residual length L in this area. The character features of our tracks demonstrate that the nuclei are only located near the stoppage. Figure 18 shows that, in this case, the V dependence from the charge is well-approximated by a straight line (there are five experimental points well-approximated by a straight line at our disposal). Thus, charge determination by extrapolation in the region up to Z = 110–120 is more reliable than using empirical function V(L, Z). The obtained approximation is quite accurate and possesses a great level of confidence (95%). Further extrapolation of the straight line and the corridor of errors to an etching rate of 35 μm h−1 gives the required evaluation. The performed regression analysis of the etching rate near the stoppage point (Alexandrov et al. 2013a) made it possible to refine the assessment of charge of one of these nuclei to ${119}_{-6}^{+10}$ with a 95% probability (Figure 18). The obtained nucleus charge error estimation is carried out by extrapolation of the characteristics of a complex chemical process (etching) with the use of the results of previously performed calibration measurements. The error estimation is made with consideration of the calibration errors by means of approximation of the experimental points by a straight line, and building an error corridor with a specified confidence level.

Figure 18.

Figure 18. Regression analysis of the experimental results. The fit curve is shown as a solid line; the dashed lines indicate the error corridor at a significance level of 95%. The vertical lines identify a possible interval of the charge at a 95% significance level at an etching rate of 35 μm h−1 near the stopping point of projectiles in olivine.

Standard image High-resolution image

Thus, it seems that the data obtained in the OLYMPIA experiment can supply arguments supporting the existence of the theoretically predicted Island of Stability for long-lived transfermium nuclei of natural origin. Estimates of their minimum lifetimes are presented in Section 3.3.

3.3. Estimates of Nuclear Lifetimes

In order to reach our solar system and form tracks in meteorite olivine, nuclei registered in pallasites should exist for a sufficiently long time. Their average lifetimes should, as a minimum, be equal to the time necessary for their flight from their point of origin to our solar system's asteroid belt, from where a predominant number of meteorites come. The minimum estimate of the lifetimes of the nuclei registered in pallasites can be determined as the time of their flight from the closest supernova (SN) to our solar system, which depends on the velocity V to which a nucleus could accelerate after being formed and on the distance to our solar system.

Generally, the trajectory of a nucleus in space is not a straight line due to the existence of the galactic magnetic field. In a homogeneous magnetic field, the trajectory of a charged particle is a spiral with a Larmor radius, which in the relativistic case appears (in the SI system) as shown in Landau & Lifshitz (1975):

where m is the particle mass, e is its charge, ${v}_{\perp }$ is the initial part of the velocity in perpendicular to the field direction, B is the magnitude of a homogeneous magnetic field, and E is the particle relativistic energy.

This radius will depend on (1) the magnetic field strength, (2) charge of projectile (a cosmic nucleus may bear some electrons, and we cannot say how many electrons it bears), (3) direction of the magnetic field, which may change during the meteorite exposition (exposition time of a meteorite is comparable to the period of rotation of our solar system in the galaxy), (4) the kinetic energy of the nucleus and (5) ratio between initial components of nucleus velocity in perpendicular and parallel directions to the magnetic field's.

Taking into account all of these factors requires special calculations. We would like to estimate the minimum life time;  so, in the limit case, when ${v}_{\perp }$ in the initial timepoint is zero, and the particle velocity is directed along the magnetic field, ${v}_{\perp }$ will be zero at every other timepoint. In this case, the trajectory coincides with a straight line, which gives an assessment of a minimum possible flight time for a nucleus. If ${v}_{\perp }\gt 0$, the time of flight will only be greater.

As for nucleus collisions with particles of the galactic interstellar medium, it should be considered that up to 70% of the volume of galaxy is a hot ionized gas with particle density n = 65·10−4 atoms cm−3 (meaning dimensionless probability [n] = 65·10−4 to find an interstellar atom in one cubic centimeter) (Ferriere 2001). Half of these particles are fully ionized hydrogen and helium atoms, and the other half consists of electrons of these atoms. An estimation of the collision probability can be obtained using the experimental data on the scattering of relativistic nuclei on hydrogen and helium targets. Without going into the details of the dependence of the total cross-section of the charge and mass of the projectile nucleus, one can say that, for nuclei with energies above 500 MeV per nucleon, the collision cross-section is limited by the value of 1 barn (Webber et al. 1990). Therefore, the likelihood for the incident nucleus to scatter on a gas particle in each cubic centimeter is equal to w = (σ·[n])/(1 cm2) = 65·10−28. For example, the probability for a nucleus to scatter on a distance of 20 pc = 6·1019 cm equals W  = w·6·1019 = 3.6·10−7, and thus can obviously be neglected.

To assess the velocity of a nucleus hitting an olivine crystal from a meteorite, a specific feature of olivines' etching mechanism can be used, namely the existence of the threshold value of the ionization losses (about 18 MeV cm−2 mg−1) of projectiles, below which no etched channel emerges. For a uranium nucleus, losses above this threshold value occur at ion energies E < 500 MeV per nucleon; for a lead nucleus, at E < 300 MeV per nucleon. Herewith, the residual path length of a uranium nucleus with energy of 500 MeV per nucleon in olivine is 14 mm; it is 7 mm for a lead nucleus with energy of 300 MeV per nucleon.

The meteorite represents a nickel–iron matrix with olivine cells up to 1 cm in size. The matrix wall thickness ranges from 2 to 5 mm. Using the SRIM-2008 code, we calculate the specific losses and path lengths of uranium nucleus for a solid material, including olivine and nickel. Table 3(a) presents such values of initial kinetic energy Ekin of uranium nuclei, prior to their entry into the meteorite at which, at a corresponding preatmospheric depth PD, their energy will equal 500 MeV per nucleon—this represents the upper boundary where energy Emax will result in an etchable track. The lower boundary Emin corresponds to nuclei that approach the crystal at a corresponding depth with almost zero energy. Figure 19 shows the intervals Emin and Emax as functions of depth PD for the uranium nucleus. Only nuclei with energy E within the range Emax < E < Emin can produce an etched channel in the crystal at a preatmospheric depth PD. The width of this range is essentially constant and equals approximately 300 MeV/nucleon. Thus, there is a relation between PD and energy and, therefore, the velocity of a penetrating nucleus.

Figure 19.

Figure 19. The values Ekin (upper) and Emin (lower) for uranium nucleus as a function of preatmospheric depth PD.

Standard image High-resolution image

Table 3.  The Time-of-flight Values for Uranium (a) and Lead (b) Assessed Using This Technique, for Example, for DD = 100 pc

(a) Uranium, for DD = 100 pc
DP (mm) 10 20 30 40
Ekin, GeV/nucleon 0.84 1.15 1.47 1.79
v/c 0.84 0.88 0.91 0.93
v, pc/yr 0.258 0.2710 0.281 0.287
T, yr 217 175 148 128
(b) Lead, for DD = 100 pc
Ekin, GeV/nucleon 0.79 1.02 1.25 1.5
v/c 0.83 0.87 0.9 0.92
v, pc/yr 0.255 0.267 0.275 0.28
T, yr 219 186 162 142

Download table as:  ASCIITypeset image

Table 3 shows the velocities v for various values of preatmospheric depth PD for preselected kinetic energies Ekin = Emax (PD). By setting the distance DD from a particle's point of origin to our solar system, we can find the time of flight and recalculate it with account for the effect of deceleration into the particle's time of flight T (c is the speed of light):

There are different experimental estimations of the distance to the closest supernovae, based on direct observations of supernova remnants in space. They range from ∼200 pc (Aschenbach 1998) up to ∼1 Kpc (Badenes et al. 2006). However, it is known that supernovae have appeared within every 30 years, and it may be some are closer to Earth, but not all of them are registered. For example, there is an estimation based on measurements of 60Fe abundance in the samples from the deep ocean estimating the distance to be ∼30pc (Fields & Ellis 1999). Table 3 presents the time-of-flight values for uranium (a) and lead (b) assessed using this technique, for DD = 100 pc.

OLIMPIYA experiments detected a transfermium nucleus with Z = 119. Assessment of its minimal lifetime Tmin by the considered method (time-of-flight evaluation) requires extrapolation into the region of large charges. To carry it out, we performed necessary calculations for three heavy nuclei: Ra, Pb, and W. Figure 20 shows the summary results of extrapolation for four values of charge and various depths of PD, at a distance DD = 100 pc to a supernova. Figure 20 demonstrates that Tmin for Z = 119 falls into the interval of 50 years < Tmin < 100 years. Note that, for uranium nuclei, the corresponding value of T ranges from 150 years < Tmin < 300 years. It should be mentioned that the estimated time-of-flight for a uranium nucleus is many orders of magnitude shorter than the lifetime of uranium nuclei. Thus, the real lifetime of a nucleus detected in the experiments with Z = 119 can also be larger than its assessed Tmin. It is important that these estimated minimal lifetimes taken as the times-of-flight are many orders of magnitude larger than the lifetimes of transfermium nuclei produced artificially at accelerators.

Figure 20.

Figure 20. Dependence of the times of flights of different nuclei for various olivine crystal depths at DD = 100 pc, extrapolated by regression analysis into the region of large charges. PD  = 10 mm (1), 20 mm (2), 30 mm (3), and 40 mm (4).

Standard image High-resolution image

4. CONCLUSIONS

In the OLIMPIYA project, olivine crystals from meteorites were successfully used as detectors for identification of heavy and superheavy nuclei in the spectra of galactic cosmic rays.

The applied method is based on layer-by-layer grinding and etching of particle tracks in these crystals. Unlike other authors' techniques, this annealing-free technique makes use of two parameters for identification of charge Z of a projectile: the etching rate along the track (Vetch) and the total detected track length (L). A series of irradiations with different swift heavy ions at the accelerator facilities of GSI (Darmstadt) and IMP (Lanzhou) were performed in order to determine and calibrate the dependence of projectile charge Z on the measured Vetch and L.

The applied approach made it possible not only to determine the charges of the detected GCR nuclei with an accuracy of ±2 charge units within the region of Z < 92, but also to increase significantly the accuracy of charge determination for nuclei with charges Z > 92.

A considerable amount of experimental data on superheavy nuclei of natural origin was obtained, and accumulation of data continues. The statistics of 11,647 GCR heavy nuclei (Z > 40) enables evaluation of this data bank as one of the largest to date–representing a sizable contribution to the world's collection of experimental data on superheavy nuclei of natural origin.

Three detected tracks were identified as having been produced by nuclei with charges 113 < Z < 129. Their estimated lifetimes are many orders of magnitude larger than those of transfermium nuclei produced artificially at accelerators. This gives reason to believe that the data obtained in the OLYMPIA experiment supply arguments supporting the existence of the theoretically predicted Island of Stability for long-lived transfermium nuclei of natural origin.

The authors are very grateful to Professor Yu. Oganesyan for strong support of this work and useful discussions.

S.A. Gorbunov and A.E. Volkov acknowledge financial support from the Russian Foundation for Basic Research, Russia (Grants 16-38-00877 and 15-02-02875, respectively).

Please wait… references are loading.
10.3847/0004-637X/829/2/120