Nuclear isomers: stepping stones to the unknown

Similar documents
ISOMER BEAMS. P.M. WALKER Department of Physics, University of Surrey, Guildford GU2 7XH, UK

K-isomer decay rates. Phil Walker. - introduction - exploring the different degrees of freedom - exotic nuclei

Shape Effects in E2 Transition Rates from Z 76 High-Spin Isomers

and shape coexistence

High-spin isomers: structure and applications

Probing neutron-rich isotopes around doubly closed-shell 132 Sn and doubly mid-shell 170 Dy by combined β-γ and isomer spectroscopy.

arxiv:nucl-th/ v1 14 Nov 2005

Nuclear Isomerism. Phil Walker. University of Surrey. on the occasion of the 70th birthday of Geirr Sletten

arxiv:nucl-th/ v1 19 May 2004

Isomers in Heavy Deformed Nuclei

Shape Coexistence and Band Termination in Doubly Magic Nucleus 40 Ca

Evidence for K mixing in 178 Hf

Some (more) High(ish)-Spin Nuclear Structure. Lecture 2 Low-energy Collective Modes and Electromagnetic Decays in Nuclei

Quadrupole moment of the K π = 14 + isomer in 176 W

Review of Metastable States in Heavy Nuclei

Magnetic rotation past, present and future

and 14 + isomers in 175,176 W

arxiv: v1 [astro-ph.he] 28 Dec 2010

K-hindered decay of a six-quasiparticle isomer in 176 Hf

c E If photon Mass particle 8-1

Theoretical basics and modern status of radioactivity studies

Nuclear Spectroscopy I

Coexistence phenomena in neutron-rich A~100 nuclei within beyond-mean-field approach

Gamma-ray spectroscopy I

RFSS: Lecture 8 Nuclear Force, Structure and Models Part 1 Readings: Nuclear Force Nuclear and Radiochemistry:

Projected shell model analysis of tilted rotation

Conversion Electron Spectroscopy in Transfermium Nuclei

Mean field studies of odd mass nuclei and quasiparticle excitations. Luis M. Robledo Universidad Autónoma de Madrid Spain

Search for very long-lived isomers in the hafnium-tungsten region

A Comparison between Channel Selections in Heavy Ion Reactions

arxiv: v2 [nucl-th] 8 May 2014

New T=1 effective interactions for the f 5/2 p 3/2 p 1/2 g 9/2 model space: Implications for valence-mirror symmetry and seniority isomers

Some Aspects of Nuclear Isomers and Excited State Lifetimes

Identification of yrast high-k isomers in 177 Lu and characterisation of 177m Lu

Lisheng Geng. Ground state properties of finite nuclei in the relativistic mean field model

arxiv: v1 [nucl-th] 8 Sep 2011

Chapter 6. Summary and Conclusions

Structure of Sn isotopes beyond N =82

CHEM 312 Lecture 7: Fission

14. Structure of Nuclei

Highly deformed Bands in 175 Hf

Probing the shell model using nucleon knockout reactions

H.O. [202] 3 2 (2) (2) H.O. 4.0 [200] 1 2 [202] 5 2 (2) (4) (2) 3.5 [211] 1 2 (2) (6) [211] 3 2 (2) 3.0 (2) [220] ε

Nuclides with excess neutrons need to convert a neutron to a proton to move closer to the line of stability.

Basic science. Atomic structure. Electrons. The Rutherford-Bohr model of an atom. Electron shells. Types of Electrons. Describing an Atom

B(E2) value of even-even Pd isotopes by interacting boson model-1 *

Role of Hexadecupole Deformation in the Shape Evolution of Neutron-rich Nd Isotopes

Composite Nucleus (Activated Complex)

Chapter VIII: Nuclear fission

SPIN-PARITIES AND HALF LIVES OF 257 No AND ITS α-decay DAUGHTER 253 Fm

2007 Fall Nuc Med Physics Lectures

Nuclear Isomers: What we learn from Tin to Tin

Discovery of highly excited long-lived isomers in neutron-rich hafnium and tantalum isotopes through direct mass measurements

Chapter 44. Nuclear Structure

Stability of heavy elements against alpha and cluster radioactivity

Alpha decay. Introduction to Nuclear Science. Simon Fraser University Spring NUCS 342 February 21, 2011

Nuclear Physics for Applications

High-spin studies and nuclear structure in three semi-magic regions of the nuclide chart High-seniority states in Sn isotopes

Nuclear vibrations and rotations

Projected shell model for nuclear structure and weak interaction rates

SURROGATE REACTIONS. An overview of papers by Jason Burke from LLNL

Nuclear Fission Fission discovered by Otto Hahn and Fritz Strassman, Lisa Meitner in 1938

Fission fragment mass distributions via prompt γ -ray spectroscopy

arxiv:nucl-th/ v1 1 Mar 2001

Evolution Of Shell Structure, Shapes & Collective Modes. Dario Vretenar

Fast-Timing with LaBr 3 :Ce Detectors and the Half-life of the I π = 4 Intruder State in 34 P

Annax-I. Investigation of multi-nucleon transfer reactions in

Decay studies of 170,171 Au, Hg, and 176 Tl

Rotational motion in thermally excited nuclei. S. Leoni and A. Bracco

Nuclear Physics and Astrophysics

4. Rotational motion in thermally excited nuclei *

Allowed beta decay May 18, 2017

RFSS: Lecture 6 Gamma Decay

CHAPTER 12 The Atomic Nucleus

Lecture 11 Krane Enge Cohen Williams. Beta decay` Ch 9 Ch 11 Ch /4

Nuclear and Radiation Physics

The IC electrons are mono-energetic. Their kinetic energy is equal to the energy of the transition minus the binding energy of the electron.

Correction to Relativistic Mean Field binding energy and N p N n scheme

arxiv:nucl-th/ v2 16 Mar 2003


O WILEY- MODERN NUCLEAR CHEMISTRY. WALTER D. LOVELAND Oregon State University. DAVID J. MORRISSEY Michigan State University

STRUCTURE FEATURES REVEALED FROM THE TWO NEUTRON SEPARATION ENERGIES

Introduction to Nuclear Physics and Nuclear Decay

Introduction to Nuclear Science

Exploring the Structure of Cold and Warm Nuclei Using Particle Accelerators in India

Karlsruhe Nuclide Chart

High-spin states in 90 Ru and the projected shell model description

arxiv:nucl-th/ v4 30 Jun 2005

Structure of hot nuclear states

Sunday Monday Thursday. Friday

HEIDI WATKINS. Plunger Measurements of Shape Coexistence in the Neutron Deficient 174 Pt Nuclei

Collective rotation and vibration in neutron-rich 180,182 Hf nuclei

FACTS WHY? C. Alpha Decay Probability 1. Energetics: Q α positive for all A>140 nuclei

Collective Excitations in Exotic Nuclei

Nuclear and Particle Physics

INVESTIGATION OF THE EVEN-EVEN N=106 ISOTONIC CHAIN NUCLEI IN THE GEOMETRIC COLLECTIVE MODEL

Fission fragment angular distribution - Analysis and results

Theoretical Nuclear Physics

Shape coexistence and beta decay in proton-rich A~70 nuclei within beyond-mean-field approach

Probing the evolution of shell structure with in-beam spectroscopy

Transcription:

Nuclear isomers: stepping stones to the unknown P.M. Walker Department of Physics, University of Surrey, Guildford GU2 7XH, UK Abstract. The utility of isomers for exploring the nuclear landscape is discussed, including their role in superheavy-element research, and the possibility of observing neutron radioactivity. Emphasis is given to K isomers in deformed nuclei. Transition rates are examined in the N p N n scheme for 2- and 3-quasiparticle K-isomer decays, and in connection with level densities for higher quasiparticle numbers. Keywords: nuclear structure, isomers, limits of stability, high spin PACS: 21.10.-k, 27.70.+q INTRODUCTION Nuclear isomers continue to make key contributions to the development and understanding of nuclear structure physics [1, 2]. Isomers are widely distributed in neutron number, proton number, excitation energy and angular momentum. With half-lives ranging from nanoseconds to years, they offer a variety of opportunities to explore unusual and extreme states of nuclei. The understanding of their occurrence and degree of stability promises additional nuclear structure insights, with the potential for novel applications. Figure 1 shows a restricted set of isomers, with long half-lives and high excitation energies. It is clear that many such isomers are correlated with shell closures. These can be categorised as spin traps: they arise from special combinations of single-particle orbits in a spherical potential, with decay transitions that must carry high angular momentum (hence the long half-lives). Away from the closed shells are regions of deformed nuclei, where axial symmetry gives rise to the K quantum number. The associated K traps involve transitions with large changes in the orientation of the angular momentum (though not necessarily in the magnitude). In Figure 1, only 242 Am represents shape isomers, where it is a nuclear shape change, rather than an angular-momentum change, that hinders the isomer deexcitation. A convenient listing of isomers, with their angular momenta and half-lives, can be found in ref. [3]. However, the tabulation is incomplete for isomers with half-lives < 1 ms. The delay time for isomer decays provides an important experimental tool for separating them from the bulk of the radiation from nuclear reactions. This is all the more important when exploring the limits of nuclear stability, with non-selective reactions that lead to many final products. In the study of neutron-rich nuclei with deep-inelastic reactions, fragmentation reactions or fission, the additional selectivity provided by isomer decays can be especially valuable. 16

Z 80 60 isomers >1 MeV >1 ms 40 20 20 40 60 80 120 140 160 N FIGURE 1. Nuclear chart: squares naturally occurring nuclides; circles isomers with half-lives > 1 ms and excitation energies > 1 MeV; and lines closed shells (adapted from ref. [1]). HIGH-K ISOMERS The K value is the angular-momentum projection on the nuclear symmetry axis. This is not a directly measured quantity, but it is usually assumed to be equal to the total angular momentum (I) of the isomer (which is itself typically observed as the lowestenergy, lowest-spin member of a rotational band). To a first approximation, the K value is unaltered by collective rotation, which has it angular momentum perpendicular to the symmetry axis. However, inertial forces, principally the Coriolis force, cause K mixing and rotational alignment of individual nucleon orbits. It is these and other K-mixing effects that determine K-isomer decay rates. The wide range of K-isomer half-lives is evident in Figure 2. It is also clear that the majority of K isomers are found in the A 180 region. Indeed, it is only in that region that isomers involving > 3 quasiparticles are so far observed. Since the K values of isomers increase with quasiparticle number, it is through the properties of isomers in the A 180 region that angular-momentum (and pairing) effects can best be studied. Arguably the most famous K isomer is the 31-year, K π = 16 +, 2.45-MeV excitation of 178 Hf, originally discovered over 35 years ago following reactor-neutron bombardment of natural hafnium [4]. The vast majority of the 31-year isomer decay strength is by (unobserved) conversion electrons associated with a 13-keV E3 transition to the I = 13 17

250 mass (amu) 200 150-9 -8-7 -6-5 -4-3 -2-1 0 1 2 3 4 5 6 7 8 9 log T1/2 (s) FIGURE 2. K-isomer mass number versus half-life (> 5 ns). The dots represent 2-quasiparticle isomers (excluding odd-odd nuclei) and 3-quasiparticle isomers, and the circles increase in size with quasiparticle number, up to 9 quasiparticles. Isomers in nuclides within four mass units of closed shells are omitted. member of a K = 8 band. With decay by a λ = 3 transition, the isomer can be called a spin trap, and with a K change of 8 units the transition has a forbiddenness of ν = K λ = 5, classifying the isomer also as a K trap. This is an exceptional spin-trap/k-trap situation, which can explain the remarkable combination of high excitation energy and long halflife. Only recently have γ-ray emissions directly from the isomer been observed [5], in the form of low-intensity M4 and E5 branches, accounting for 0.015% and 0.006% of its decay, respectively. In comparison with Weisskopf single-particle estimates, these decays are hindered by factors, F W 10 6 10 9, while the reduced hindrances are much more uniform: f ν = F 1/ν W. This much is reasonably well understood, and the structure of the isomer can be explained in terms of the single-particle orbitals close to the Fermi surface: a broken neutron pair and a broken proton pair each contribute half of the 4-quasiparticle isomer s 16 h. A puzzle remains in the experimental data, however. Unplaced transitions, illustrated in Figure 3, are observed with similar intensity to the E5 decay of the isomer. Considering that the source was prepared from an irradiation performed in 1980, and the illustrated measurements date from 2003 [5], it would appear that the unplaced tran- 18

250 200 154-keV gate *912 Counts 150 50 868 *901 959 0 200 150 850 875 900 925 950 868-keV gate 154 *181 213 Counts 50 (91) *113 0 125 150 175 200 E γ (kev) FIGURE 3. Gamma-ray coincidence spectra from a 178 Hf isomer source measurement [5] (see text). Known contaminant transitons are indicated by asterisks. Other labeled transitions are of unknown origin. sitions, though of low intensity, are associated with a half-life of many years. Whether they might come from a higher-lying isomer in 178 Hf is speculation, but no trivial explanation has yet been found. A more complete understanding of the γ-ray spectrum associated with the 178 Hf 31- year isomer decay is not only of academic interest. There has been much comment, see refs [1, 2], concerning the possibility of storing energy in nuclear isomers, potentially with controlled release of that energy. The 178 Hf isomer has been the most sought-after candidate, but the ability to trigger the release of the isomer s energy with < kev photons has not yet been adequately demonstrated. Further knowledge of the nuclear structure of 178 Hf could prove valuable in this regard. It may be noted that, while isomer triggering in 178 Hf remains unproven, the triggering of the only naturally occurring isomer, 180m Ta, is well established [6], albeit with > 1 MeV photons; and the triggering pathway can be correlated in detail with the known level structure [7, 8]. This is an interesting example from an astrophysics perspective, as photo-activation could destroy the isomer in stellar s-process environments [6], if that is where the isomer is synthesised in the first place. LIMITS OF NUCLEAR STABILITY The drip lines, with associated proton and (potentially) neutron radioactivity, define the limits of nuclear stability in the N-Z plane, together with the fission and α decay of superheavy nuclei. Isomers can enhance our understanding of these limits, albeit 19

complicated by their angular-momentum and excitation-energy dimensions. In the superheavy domain, detailed measurements have now established high-k µs and ms isomers in 254 102No [9], highlighting the importance of conversion-electron detection. Going even heavier, a 6-ms isomer was recently found at 1 MeV in 270 110Ds [10], with a half-life that is longer than that of its ground state. Xu et al. [11] have calculated that comparable isomers should exist in a range of even-even superheavy nuclei, and in a real sense isomers can provide extra stability, which may be important for future experimental discoveries. Approaching the proton drip line for lighter masses, the isomer 53m Co provided the first example of proton radioactivity [12]. Peker et al. [13] subsequently pointed out that the angular-momentum barrier could lead to neutron radioactivity from high-spin isomers close to the neutron drip line. In this context, it is interesting that some high-spin isomers, such as the K π = 57/2, 22-ns isomer in 175 Hf [14, 15], are unbound to neutron decay. However, the high angular momentum makes the probability very small for neutron emission, and it would be necessary to find much longer-lived isomers on the neutron-rich side of stability, to have a realistic chance of detecting their neutron decay. Such a possibility could perhaps apply to 187 Hf, where favored multi-quasiparticle isomers are predicted, as shown in Figure 4. Nevertheless, 7-quasiparticle excitations are needed in order to exceed the neutron separation energy. Experimentally, the ground state of 187 Hf has already been observed [17] from projectile-fragmentation reactions, but whether the required high angular momenta are within reach is not yet proven. However, the population of a spin 43/2 isomer in 215 Ra following 238 U fragmentation [18], and the inference that there must be a large and unexpected collective contribution to the angular momentum generated in the final isomeric fragment, gives encouragement. While, therefore, the objective of finding neutron decay from isomers remains challenging, the issue is not one of too-rapid decay in the absence of the Coulomb barrier, it is one of too-hindered decay in the presence of the angular-momentum barrier. In the course of performing isomer calculations in the region of 187 Hf, it has been found that collective oblate rotation becomes an increasingly important mode, as the neutron richness increases [19]. In 190 Hf, the ground state itself is predicted to be oblate, and that shape remains favored up to high spin [20]. At the same time, coexisting multi-quasiparticle prolate states are predicted. Consequently, with sufficient primary beam intensity, it may be possible to identify the oblate collective states through their population following prolate isomer decays. K-ISOMER DECAY RATES Even though the configurations and excitation energies of K isomers can be calculated with some confidence, it remains problematical to predict their half-lives. The wavefunction overlaps are small, and even the relative importance of the different mechanisms for generating the overlaps is not well defined. The principal K-mixing mechanisms, necessary for non-zero K-forbidden transition rates, can be categorised as: (i) Coriolis (orientation) mixing; (ii) γ-tunneling (shape) mixing; (iii) statistical (thermal) mixing; and (iv) mixing due to chance near-degeneracies. These mechanisms have been discussed in many different contexts, see for example 20

10 187 Hf 8 Energy (MeV) 6 4 2 0 1qp ( π) 1qp (+π) 3qp ( π) 3qp (+π) 5qp ( π) 5qp (+π) 7qp ( π) 7qp (+π) 9qp ( π) 9qp (+π) 0 200 400 600 800 0 K(K+1) FIGURE 4. Multi-quasiparticle states in 187 Hf calculated with the Nilsson+BCS method of Jain et al. [16]. The neutron separation energy is approximately 4 MeV. ref. [21], and here the presentation is restricted to extensions of the N p N n scheme for 2- and 3-quasiparticle isomers [22], and statistical mixing considerations for higher quasiparticle numbers [23], in both cases concentrating on K-forbidden E2 transitions in the A 180 region. As described above, a simple measure of the goodness of the K quantum number is the hindrance per degree of K forbiddenness, often referred to as the reduced hindrance. For 2- and 3-quasiparticle E2 decays, there is found [22] to be a strong correlation of reduced hindrance with N p N n, the product of the valence nucleon numbers, which is itself a simple measure of collectivity. One may consider that for larger N p N n values 21

0.85 rigid rotor f ν 10 179Ta(49/2) 174Yb(14) 175Hf(57/2) 1 1.0 1.5 2.0 2.5 E K -E R (MeV) FIGURE 5. Variation of reduced hindrance with energy relative to a rotor with 85% of the rigid-body moment of inertia. Quasiparticle numbers are equal to 4 (circles even-even nuclei), 5 (squares), 7 (triangle up), and 9 (triangle down). For the odd-mass nuclei, a pairing energy of 0.9 MeV has been added. The data are for E2 and E3 decays with K > 5. The full line represents the predicted level-density dependence [23]. The figure is adapted from ref. [28]. The labeled open symbols are discussed in the text. the nuclei are more stably deformed and less susceptible to mixing effects, hence the correlation with reduced hindrance. Recent data for a 6-ns (31/2 + ) isomer in 191 Os [24] indicate an extension of the systematic reduced-hindrance behavior to a low value of N p N n = 66, with f ν = 1.9. This is in marked contrast to the value for the 6 + isomer in 174 Yb, having N p N n = 264, and f ν = 322, which represents the other extreme of our present knowledge in the A 180 region (and disagrees with the γ-tunneling model [25]). The remarkably high 174 Yb reduced hindrance has been discussed and confirmed recently by Dracoulis et al. [25], though not in the context of N p N n systematics. Data at even larger N p N n values can be anticipated in the near future, as 6 +, 2-quasineutron isomers are predicted in the isotones 172 Er and 170 Dy [26], with N p N n = 308 and 352, respectively. For higher quasiparticle numbers (> 3) the systematic behavior is distinctly different, perhaps as a consequence of blocked pairing correlations. Statistical mixing seems to be an important factor. This can be related to the level density, which may itself be determined from the excitation energy relative to a rigid-rotor reference [23]. If only a small range of angular momentum is considered (previously 4- and 5-quasiparticle isomers) then there is little sensitivity to the reference moment of inertia. However, including recent data from 7- and 9-quasiparticle isomers in 179 Ta [27] and 175 Hf [15], respectively, reduction of the reference moment of inertia by a factor of 0.85 is seen to be appropriate [28]. The resulting dependence of the reduced hindrance on the 22

relative excitation energy is illustrated in Figure 5, showing that the general behavior is indeed systematic, but there are considerable fluctuations that presumably result from other significant degrees of freedom (see also ref. [21]). The new data point for a 4- quasiparticle isomer in 174 Yb [25] is in good accord with expectations. SUMMARY Isomers provide many insights into the nature of nuclear excitations and the limits of stability. The present paper has reviewed some of these aspects and pointed out that neutron decay from high-spin isomers may be observable. Extensions of the systematic behavior of K-isomer reduced-hindrance values have been presented, covering the full range of observed multi-quasiparticle decays. ACKNOWLEDGMENTS This work involves many collaborations: special thanks to Rodi Herzberg, Filip Kondev, Zsolt Podolyák, and Furong Xu. The UK EPSRC and Royal Society are thanked for financial support. REFERENCES 1. P. M. Walker, and G. D. Dracoulis, Nature 399, 35 (1999). 2. P. M. Walker, and J. J. Carroll, Physics Today 58-6, 39 (2005). 3. G. Audi, O. Bersillon, J. Blachot, and A. H. Wapstra, Nucl. Phys. A 729, 3 (2003). 4. R. G. Helmer, and C. W. Reich, Nucl. Phys. A 114, 649 (1968). 5. M. B. Smith et al., Phys. Rev. C 68, 031302(R) (2003). 6. D. Belic et al.,phys. Rev. C 65, 035801 (2002). 7. P. M. Walker, G. D. Dracoulis, and J. J. Carroll, Phys. Rev. C 64, 061302(R) (2001). 8. P. M. Walker, Hyp. Int., 143, 143 (2002). 9. R.-D. Herzberg et al., Physica Scripta, in press. 10. S. Hofmann et al., Eur. Phys. J. A 10, 5 (2001). 11. F. R. Xu, E. G. Zhao, R. Wyss, P. M. Walker, Phys. Rev. Lett. 92, 252501 (2004). 12. K. P. Jackson et al., Phys. Lett. B 33, 281 (1970). 13. L. K. Peker, E. I. Volmyansky, V. E. Bunakov, and S. G. Ogloblin, Phys. Lett. B 36, 547 (1971). 14. N. L. Gjørup et al., Zeit. Phys. A 337, 353 (1990). 15. F. G. Kondev et al., to be published. 16. K. Jain et al., Nucl. Phys. A 591, 61 (1995). 17. J. Benlliure et al., Nucl. Phys. A 660, 87 (1999). 18. Zs. Podolyák et al., to be published. 19. F. R. Xu, P. M. Walker, and R. Wyss, Phys. Rev. C 62, 014301 (2000). 20. F. R. Xu et al., to be published. 21. P. M. Walker, and G. D. Dracoulis, Hyp. Int. 135, 83 (2001). 22. P. M. Walker, J. Phys. G 16, L233 (1990). 23. P. M. Walker et al., Phys. Lett. B 408, 42 (1997). 24. G. A. Jones et al., J. Phys. G, in press. 25. G. D. Dracoulis et al., Phys. Rev. C 71, 044326 (2005). 26. P. H. Regan et al., Phys. Rev. C 65, 037302 (2002). 27. F. G. Kondev et al., Eur. Phys. J. A 22, 23 (2004). 28. P. M. Walker, Acta Phys. Pol. B, 36, 1055 (2005). 23