You are on page 1of 10

Nuclear Physics A 805 (2008) 260c–269c

www.elsevier.com/locate/nuclphysa

Mass and Half-life Measurements of Stored Exotic


Nuclei at the FRS-ESR Facility
Yuri A. Litvinov
Gesellschaft für Schwerionenforschung GSI, 64291 Darmstadt, Germany
Justus-Liebig Universität, 35392 Gießen, Germany
St. Petersburg State University, 198504 St. Petersburg, Russia

Abstract

Isochronous and time-resolved Schottky mass spectrometry have been developed at the FRS-
ESR facility of GSI for precise mass and half-life measurements of exotic nuclei. Both techniques
are sensitive to single stored ions which makes them highly efficient. More than 1100 atomic
masses have been measured meanwhile. Half-life measurements are performed with bare and
few-electron ions whose decay modes can dramatically differ from the ones known in neutral
atoms. After a brief description of the experimental methods, a short review of the obtained
results and future perspectives will be presented.

Key words: Mass and half-life measurements, Exotic nuclei, Highly-charged ions, Storage rings
PACS: 21.10.Dr, 21.10.Tg, 23.40.-s, 29.20.Dh

1. Introduction

Three types of fundamental interactions, namely strong, weak, and electromagnetic,


play–by acting between the nucleons–a major role in atomic nuclei. The sum effect of
these interactions is reflected in the binding energy of the nucleus, which is directly
connected with its mass [1]. Experimental nuclear masses are often used as a tool to
reveal new nuclear structure effects. Indeed, the shell structure and pairing correlations
have been discovered via nuclear masses [1].
The actual pathways of the nucleosynthesis in stars are governed by the nuclear binding
energies and lifetimes. These processes happen at high temperatures and the nuclides,

Email address: y.litvinov@gsi.de (Yuri A. Litvinov).


URL: http://www-linux.gsi.de/∼litvinov/frs-esr/ (Yuri A. Litvinov).

0375-9474/$ – see front matter © 2008 Elsevier B.V. All rights reserved.
doi:10.1016/j.nuclphysa.2008.02.254
Yu.A. Litvinov / Nuclear Physics A 805 (2008) 260c–269c 261c

Production target Cocktail or


Monoisotopic
Br or Br-DE-Br beams
separation Electron cooler

FRS ESR
Gas-Jet
SIS target

Schottky Time-of-Fight
pick-ups detector

Injection from
UNILAC
Fig. 1. SIS-FRS-ESR facility at GSI. High-intensity primary beams are accelerated in the SIS to several
hundred MeV/u and impinged on the production target in front of the FRS. The reaction products
are separated in-flight in the FRS and injected into the ESR, where their masses and/or lifetimes are
measured. The main components used in our experiments are indicated.

therefore, appear in high ionic charge states. The decay properties of such highly-ionized
nuclides can be altered dramatically: decay modes known in neutral atoms can become
forbidden and the new ones can be opened up, which can play an essential role in the
nucleosynthesis processes [2,3].
Today, the challenge is to measure the masses and lifetimes of exotic nuclei close to
the borders of their existence. These nuclides can reveal new nuclear properties due to
the strong asymmetry of their proton-to-neutron ratio. However, they are difficult to
investigate due to their small production cross-sections and short lifetimes. Therefore,
very efficient and fast experimental techniques are required.
In this contribution we first give an overview on the experimental program on mass
and lifetime measurements with the FRS-ESR facility at GSI and then outline future
perspectives offered by the planned NUSTAR facility at FAIR [4].

2. Experimental Setup

The combination of the high-energy heavy-ion synchrotron SIS [5], the in-flight frag-
ment separator FRS [6], and the cooler-storage ring ESR [7] at GSI, provides unique
experimental conditions for studying exotic nuclei.
Relativistic fragments of several hundred MeV/u are produced via fragmentation or
fission of primary beams in thick production targets. Typically 9 Be-targets with thick-
nesses of 2-8 g/cm2 are used. Secondary beams are separated in flight in the FRS within
about 150 nanoseconds and are then injected into the ESR [8]. Cocktail or clean mono-
isotopic beams can be prepared with the FRS by employing the magnetic rigidity (Bρ)
analysis and atomic energy loss (ΔE) in specially shaped matter. A review on the pro-
duction and separation of radioactive beams at the FRS can be found in Ref. [9]. The
262c Yu.A. Litvinov / Nuclear Physics A 805 (2008) 260c–269c

facility is schematically illustrated in Fig. 1.


The frequencies f of the circulating ions in the storage ring can be related in first-order
approximation to their mass-to-charge ratios (m/q) by the following expression:

Δf 1 Δ(m/q) γ 2 Δv
=− 2 + (1 − 2 ) , (1)
f γt (m/q) γt v
where γ is the relativistic Lorentz factor and γt is the transition point of a storage ring
(for more details see Ref. [10] and references cited therein). From this equation follows,
that in order to measure mass-to-charge ratios of stored ions one has to measure their
revolution frequencies and get rid of the second term on the right hand side. Inevitably,
the fragments have a velocity spread Δv/v of the order of a few percent due to their
stochastic creation process.
Two complementary experimental methods, namely Isochronous (IMS) and Schottky
(SMS) Mass Spectrometry, have been proposed [11,12] for accurate mass measurements.
The time-resolved SMS enables the possibility of radioactive decay studies [10].
In the SMS, the velocity spread can be reduced by stochastic [13] and electron cooling
[14], which force all stored ions towards the same mean velocity and thereby reduce the
velocity spread to roughly 5 · 10−7 for beam intensities below about 2000 ions. Thus the
second term in Eq. 1 becomes negligible and the measured revolution frequencies reflect
directly the mass-to-charge ratios of the stored ions [15,16]. The revolution frequencies
are measured by the Schottky-noise spectroscopy, which is described in details in Ref.
[17] and references cited therein. The disadvantage of SMS is that the cooling process
lasts a few seconds. Thus, only the nuclides with half-lives longer than about 1 second
can be investigated.
The area of a Schottky-frequency peak is proportional to the number of stored ions
[10,18,19]. Thus, by tracing in time the intensities of the stored ions, we can measure
their half-lives. Furthermore, the ESR can often be tuned such that daughter ions from
radioactive decays are also stored. In these cases the decay branchings can be accurately
determined.
Exotic nuclei with half-lives shorter than the cooling time can be investigated with a
time-of-flight technique by operating the ESR in the isochronous mode. In this case the
ions of interest are injected into the ESR with γ = γt ≈ 1.41. Thus, the term containing
the velocity spread in Eq. 1 is (in first order) equal to zero, i.e. the revolution frequency of
the circulating fragments does not depend on their velocity spread [20,21]. A time-of-flight
detector is used to measure the revolution frequencies. This detector records time stamps
of each ion passing through it at each revolution. The detector and its performance are
described in details in Refs. [20–22]. The fragments do not require cooling and the nuclei
with half-lives as short as 10 μs–about 20 revolutions in the ESR–can be measured with
this method.

3. Schottky Mass Spectrometry

Electron cooled ion beams occupy the entire acceptance of the ESR, which corresponds
to about Δ(m/q)/(m/q) ≈ 2.5%. Several ten different nuclides in up to three ionic charge
states can be measured simultaneously, see, e.g., Fig. 3 in Ref. [23]. It is essential that
the nuclides which are used for the calibration are stored together with the nuclides of
Yu.A. Litvinov / Nuclear Physics A 805 (2008) 260c–269c 263c

Er Tm Os Pt Hg
16 Lu W Pb
Au Tl
Yb Re Ir
Hf
Ta
14 Bi Ac Th Pa U
Po
Rn Ra
At
12
Fr
Fig. 2. Experimental values for two-proton
Np separation energies (S2p ). Contribution
10
S 2p [MeV]

of the measured mass surface at the


8 FRS-ESR is indicated with grey and black
filled symbols. The latter represent the val-
6
ues which could be obtained for the first
4 time. The error bars are mostly within
the symbol size. Linear extrapolations to
2
S2p = 0 and interpolations are shown with
0 dotted lines.

140 160 180 200 220 240


Mass Number

interest. In this way the instabilities of the magnetic fields in the ESR are experienced
by all stored ions and can, therefore, be corrected in the analysis [23,24].
It is possible to detect and measure the revolution frequencies of single stored ions with
atomic numbers Z≥12 [25]. Detecting single ions is of a great advantage for resolving low-
lying isomeric states since the ion can only be present in one or the other state [26]. Thus,
ground or isomeric states can be assigned even for very small excitation energies, which
is not possible under the condition when both states are simultaneously populated [26–
28]. In this way, a long-lived isomeric state in 125 Ce with excitation energy of E ∗ =
103(12) keV has been discovered recently [29].
The mass resolving power of the time-resolved SMS of about 2·106 (FWHM) has been
achieved [23,26]. 285 new and more than 300 improved mass values of neutron-deficient
nuclides in the range 36≤Z≤92 have been contributed (including mass values obtained
indirectly via Q-values of α, β, or proton decays) [23,30,31] to the latest Atomic Mass
Evaluation AME’03 [32]. Recently, masses of 373 neutron-deficient nuclides have been
measured. The masses for 18 nuclides (84 Zr, 92 Ru, 94 Rh, 107,108,110 Sb, 111,112,114 I, 118 Ba,
122,123
La, 124 Ce, 127 Pr, 129 Nd, 132 Pm, 134 Sm, 137 Eu) have been determined for the first
time [25]. In addition, the masses for 111,112 I and 113 Xe have been obtained via known
α-decay energies [25]. The data on the masses of neutron-rich 238 U-projectile fragments
[33] are still in the analysis. In this experiment we could identify and measure masses
and half-lives of hitherto unknown isotopes, e.g. 235 Ac [33,34].
The mass surface covered in SMS experiments is illustrated in Fig. 7. The relative
mass accuracy of the SMS amounts presently to Δ(m)/m ≈ 1.5 · 10−7 [23–26]. Somewhat
better mass accuracy of about Δ(m)/m ≈ 4 · 10−8 can be reached in special cases by
using narrow frequency ranges and well known masses for calibration. For example, the
mass excess value M E = −84376(4) keV has been obtained for the isomeric state 93m M o
[25], which is in good agreement with the literature value M E = −84378(4) keV [32].
The measured mass surface can be used for systematic investigations of nuclear struc-
ture effects. For example, the location of the one-proton and two-proton drip-lines became
possible for heavy nuclei around lead [24,31,35]. The two-proton separation energies S2p
are plotted in Fig. 3. From this figure one can see that Hg, Tl, Pb, and Bi isotopes have
known two-proton unbound isotopes.
Our new data can be used for testing the predictive powers of various mass models.
Benchmark tests have been performed in Refs. [27,31,36–38]
264c Yu.A. Litvinov / Nuclear Physics A 805 (2008) 260c–269c

protons Er Hf neutrons
Gd Yb W Hg Pb
Sm Dy Os Pt Pt
1.6 1.6 Os
Xe BaCe Nd Te
Ba Dy
Hf W
Gd
Xe Nd Sm Yb
1.4 1.4 Ce Er
OES [MeV]

OES [MeV]
1.2 1.2

1.0 1.0

0.8 0.8
N=82

N=82
0.6 0.6

0.4 0.4

0.2 0.2
4 4
6 6
8 80 8 80
10 75 10 75

N-Z
12
14 65
70 Z N-Z
12
14 65
70 Z
16 16
18 60 18 60
20 55 20 55

Fig. 3. Odd-even staggering of nuclear masses for even-even isotopes above Sn. The experimental values
are obtained with the 5-mass formula [1]. Note, that the OES values increase towards the proton drip-line
for all presented isotopic chains. The neutron shell-closure at N=82 hampers the extraction of the OES
values.

Fig. 3 illustrates the experimental data on odd-even staggering of nuclear masses (OES)
obtained with the so-called 5-mass formula [1]. The OES values are associated with the
nucleon-nucleon pairing gap energies. The OES values increase towards the proton drip-
line for both protons and neutrons [39] which has been first noticed in Ref. [40]. This
isospin-dependent trend is not reproduced by the modern mass models, which is illus-
trated in Fig. 4 for the case of even-even Hf-isotopes [41]. One can reproduce this trend by
tuning the parameters of the isospin-dependent pairing force in macroscopic-microscopic

1.8 1.8
neutrons Parameterizations
protons Bohr-Mottelson
1.6 1.6 Vogel et al.
Mass models
OES (neutrons) [MeV]
OES (protons) [MeV]

HF-BCS (MSk7)
1.4 1.4
HFB (BSk2)
HFB (BSk14)
1.2 1.2 FRLDM
Experiment
1.0 1.0

0.8 0.8

0.6 0.6

10 20 30 40 50 10 20 30 40 50
(N - Z) (N - Z)

Fig. 4. Comparison of the experimental and theoretical OES values for even-even Hf-isotopes. The data
are normalized to the OES value of the stable 182 Hf. The parameterizations of Bohr-Mottelson and Vogel
et al. are taken from Ref. [1] and Ref. [42], respectively. The description of the Hartree-Fock models can
be found in Ref. [43]. The macroscopic-microscopic FRLDM model is described in Ref. [44].
Yu.A. Litvinov / Nuclear Physics A 805 (2008) 260c–269c 265c

models [41]. However, the most recent self-consistent HFB calculations employing BSk14
force [45] fail to describe this effect (see Fig. 4). We note that this model has been fitted
to the masses from the AME’03 [32] which contains our values.

4. Isochronous Mass Spectrometry

In the IMS the revolution frequency spectrum which can be covered in a single ion-
optical setting of the FRS-ESR corresponds to Δ(m/q)/(m/q) ≈ 13% [46], which is about
6 times broader than a single Schottky frequency spectrum. The spectrum is accumulated
over several hundred injections into the ESR where each injection contains only a few
(typically below ten) ions. After about a hundred revolutions the revolution time of a
stored ion can be obtained with a relative accuracy of ΔT /T ≈ 2 · 10−6 [47].
In the first IMS experiments we used restricted regions of the revolution frequency
spectra. By doing this, the masses of several short lived nuclei relevant for the astro-
physical rp-process [20,48] (see Fig. 7) have been measured. The achieved mass resolving
power was about 1.1 · 105 (FWHM) and the mass accuracy was about 100-500 keV.
However, the experiments have demonstrated that the isochronous condition is strictly
fulfilled only in a small m/q-range [26,49]. This is illustrated in Fig. 4, where one can
see how the isochronous conditions change for three different m/q-values. The ions with
m/q = 2.62 are well isochronous which is reflected by a small spread of revolution times.
The ions with m/q = 2.52 and even more with m/q = 2.45 show significantly different
revolution times for different magnetic rigidities (velocities).
In the last experiment, masses of about 130 238 U-fission fragments have been measured,
about 50 of them for the first time [49,50]. In this experiment the range of magnetic
rigidities was restricted to about Δ(Bρ)/Bρ ≈ 1.5 · 10−4 at the dispersive mid-focal
plane of the FRS (see Fig. 4). Though, the data analysis is still underway, preliminary
results show that the mass resolving power is improved by a factor of about 2 for the
isochronous part and that this condition is preserved nearly over the entire time-of-
flight spectrum. This is an important achievement, although at the expense of strongly
reduced transmission. However, only several particles in total are now needed in order
to determine their mass value with a relative accuracy of about (preliminary) Δm/m ≈
2 · 10−6 . This compensates the loss in transmission.

m/q = 2.45
6
DBr/Br ~ 1.5 . 10
-4
Fig. 5. Dependence of the revolution time
spread on the magnetic rigidity spread for
4
three ion species with different m/q-values.
The curve for m/q = 2.62 has been
2
DT/T [10-5]

measured using cooled 238 U91+ primary


0 beam at different velocities. The curves for
m/q = 2.52 and m/q = 2.45 have been
-2 m/q = 2.52 m/q = 2.62 calculated from the measured one assum-
ing that the ions with the same magnetic
-4 rigidities have the same orbits in the ESR.
The grey band represents the magnetic
-6
rigidity acceptance of ΔBρ/Bρ ≈ 1.5·10−4
-2 -1 0 1
which has been selected in the FRS in the
2
-3 last experiments (see text).
DBr/Br [10 ]
266c Yu.A. Litvinov / Nuclear Physics A 805 (2008) 260c–269c

140
100
Pr 58+
NPr (t) = NPr (0) . e-lt Fig. 6. Decay and growth curves of the
+
mother (hydrogen-like 140 Pr58 ) ions and
Intensity [arb. units]

electron-capture daughter (fully ionized


140 Ce58+ ) ions. The lines represent the fits

according to the given equations, where


λ-is the decay constant of radioactive de-
cay of 140 Pr and λloss reflects the particle
10
140 losses due to collisions with residual gas
Ce58+ and atomic electron capture in the electron
NCe (t) = NPr (0) l-ll loss [e -lloss t- e-lt] + NCe (0) . e-lloss t cooler of the ESR. The decay constant λ is
a sum of the electron capture decay con-
stant λEC and the three-body β + decay
constant λβ + (λ = λEC + λβ + ).
0 200 400 600 800
Time after injection [sec]

5. Decay studies with highly-charged ions

The unique feature of the FRS-ESR facility is that highly-charged radioactive ions can
be produced, separated and stored over extensive periods of time. The decay rates can be
measured in the ESR by applying the time-resolved SMS or by using particle detectors
behind a dipole bending magnet for intercepting the daughter ions [10].
The electron capture and electron conversion decays do not occur in the absence of
orbital electrons, which is fulfilled in fully-ionized ions. Thus, the pure β + -decay branch
has been measured in 52 Fe26+ ions [18]. The half-lives of isomeric states in 149 Dy and
151
Er are prolonged by factors 22(2) and 33(5), respectively [51].
Stripping bound electrons can also enable new decay channels. One of such decay
modes is the bound-state β − -decay in which the emitted electron occupies a free atomic
orbital. This decay mode can dramatically modify nuclear lifetimes [52]. Thus, the fully-
ionized 187 Re75+ ions decay by nine orders of magnitude faster than neutral 187 Re atoms
with a half-life of 42 Gyr [53]. Please note that the pair 187 Re/187 Os serves as a cosmic
clock. Bare 163 Dy66+ nuclei, being stable as neutral atoms, become radioactive, thus
allowing the s-process, the astrophysical slow-neutron capture process of nucleosynthesis,
to branch [54]. A simultaneous measurement of β-decay to the continuum and bound
states in 207 Tl81+ ions has been performed recently [55]. The branching ratio of bound
to continuum β − decay in 207 Tl81+ has been found to be 0.188(18) [55].
Orbital electron-capture (EC) decay rates in hydrogen-like and helium-like 140 Pr ions
have been measured [56]. An example of the decay curve of stored hydrogen-like 140 Pr ions
and the growth of the EC-decay daughter–fully ionized 140 Ce–ions is depicted in Fig. 5.
The striking observation is that the measured EC decay rate in 140 Pr ions with one bound
electron is 1.49(8) times larger than in 140 Pr with two bound electrons. Moreover, the half-
life of 140 Pr58+ with a single orbital electron, T1/2 = ln(2)/λ = 3.04(9) min, is even shorter
than the half-life T1/2 = 3.39(1) min [32] of the neutral 140 Pr0+ atoms with 59 orbital
electrons. This result can be explained by taking into account the conservation of total
angular momentum of the nucleus-lepton system since only particular spin orientations
of the nucleus and of the captured electron can contribute to the decay [57]. In the initial
state (i), the total angular momentum Fi of a 140 Pr nucleus with spin Ii = 1 and a single
Yu.A. Litvinov / Nuclear Physics A 805 (2008) 260c–269c 267c

Mass surface measured with the


time-resolved Schottky Mass
82
Spectrometry and determined via
known energies of a, b, and proton
decays.
126
r-process
path

50 Mass surface which can be


rp-process
measured at the FRS-ESR
path

82
Mass surface which will
be accessible at FAIR
28

20 50
Mass surface measured
with Isochronous Mass
Spectrometry
8 28
20
8

Fig. 7. The chart of nuclides. The mass surface covered at the FRS-ESR facility with SMS is indicated
in grey color. The nuclides which masses have been measured with IMS are shown in black color. Dotted
line shows the nuclei which masses and half-lives can still be measured with the present FRS-ESR
facility. Dashed line shows the nuclei which will be investigated at the future NUSTAR facility at FAIR.
Astrophysical r- and rp-processes are indicated (see Ref. [33].)

bound electron with spin s = 1/2 can have two values: Fi = Ii −s = 1/2 and Fi = Ii +s =
3/2. In the final state (f ), however, we can have only one value, Ff = 1/2, which results
from the sum of the zero angular momentum of the 140 Ce nucleus If = 0 and spin s = 1/2
of the emitted electron-neutrino. The only transitions between Fi = Ff = 1/2 conserve
the total angular momentum and, therefore, allowed. The decay from the Fi = 3/2 state
requires that the emitted neutrino carries away two units of orbital angular momentum,
which corresponds to a much slower (twice forbidden) β-decay. The magnetic moment
of 140 Pr has been estimated to be μ ≈ 2.5 μN . Thus, 140 Pr58+ ions are stored–due
to short relaxation time–in Fi = 1/2 hyperfine ground state. The detailed theoretical
calculations are given in Ref. [57]. In our case that only one Fi -state contributes to the
decay, the EC-decay rate depends on the ratio of the statistical weights of the transition,
i. e. (2Ii + 1)/(2Fi + 1).
A measurement of the EC decay rate in hydrogen-like 64 Cu has been proposed in
Ref. [58]. This nucleus has μ = −0.217(2) μN and Ii = 1 [59] and the hyperfine ground
state has, therefore, Fi = 3/2. This state cannot decay by an allowed EC-decay to the
ground state of 64 Ni (If = 0).

6. Conclusion and Outlook

Our mass measurement program at the FRS-ESR facility has provided a significant
contribution to the knowledge of atomic masses. The mass surface presently covered
268c Yu.A. Litvinov / Nuclear Physics A 805 (2008) 260c–269c

in the FRS-ESR experiments is illustrated in Fig. 7. Masses of more than 1100 nuclides
were measured with SMS and IMS, and determined via known energies of α, β, or proton
decay. Masses of about 350 nuclides have been measured for the first time. These new
data have contributed to the nuclear structure and nuclear astrophysics research. Several
examples have been discussed here.
Decay studies at the FRS-ESR have shown that the half-lives of highly-ionized ions
can dramatically differ from the ones measured in neutral atoms. Such measurements
have obvious impact on the astrophysical calculations.
The ultimate sensitivity of both SMS and IMS techniques to single stored ions gives
access to nuclides with very small production rates. The region of the nuclides which
masses and half-lives can still be investigated at the FRS-ESR facility is indicated in
Fig. 7 by the dotted line.
The ILIMA-project (Isomeric beams, LIfetimes and MAsses) [60] aims for mass and
lifetime measurements at the NUSTAR facility at the future international Facility for
Antiprotons and Ion Research FAIR [4]. Production of pure isomeric beams is discussed,
e.g., in Ref. [61]. A new double-ring synchrotron system (100/300 Tm) will accelerate
ions up to uranium with intensities of 1012 particles/second. Radioactive isotopes will
be separated with a large-acceptance super-conducting fragment separator (Super-FRS)
[62] which will efficiently handle also the large phase space of the fission fragments. Two
storage rings CR and NESR will be used for half-life and mass measurements [63]. The
layout of the planned storage ring complex can be found, e.g., in Refs. [49,50,63].
The Collector Ring (CR) [64] will be operated in the isochronous mode to investigate
short-lived isotopes [65,66]. Two time-of-flight detectors on a straight section of the CR
will be used for the in-ring measurement of the velocities of stored ions. Together with
measured revolution frequencies, this will allow us to overcome the problem with changing
isochronous conditions illustrated in Fig. 4 without restricting the transmission.
In the New Experimental Storage Ring (NESR) [67] a system of several Schottky
pickups will be installed which will help us to improve the signal-to-noise ratio. For
decay studies of exotic nuclei particle counters will be mounted in the arcs of the ring to
record those decay products which cannot stay inside the ring acceptance. The range of
nuclides which will be accessible at FAIR is indicated with dashed line in Fig. 7.
This paper is entirely based on the work and on previous publications of my colleagues
from the FRS-ESR collaboration 1 . To all of them I am deeply obliged.

References

[1] A. Bohr and B.R. Mottelson, Nuclear Structure, World Scient. Publ., Singapore, 1998.
[2] J.N. Bahcall, Phys. Rev. 124, 495 (1961).
[3] K. Takahashi, K. Yokoi, Nucl. Phys. A 404, 578 (1983).
[4] An international accelerator facility for beams of ions and antiprotons, Conceptual Design Report,
GSI 2001, http://www.gsi.de/GSI-Future/CDR/
[5] K. Blasche and B. Franczak, in: Proc.: 3rd Eur. Part. Acc. Conf., Berlin, 1992, p. 9.
[6] H. Geissel, et al., Nucl. Instr. and Meth. B 70, 286 (1992).
[7] B. Franzke, Nucl. Instr. and Meth. B 24/25, 18 (1987).
[8] H. Geissel, et al., Phys. Rev. Lett. 68, 3412 (1992).
[9] H. Geissel, Progress in Part. and Nucl. Phys. 42, 3 (1999).

1 http://www-linux.gsi.de/∼litvinov/frs-esr/collaboration.shtml
Yu.A. Litvinov / Nuclear Physics A 805 (2008) 260c–269c 269c

[10] F. Bosch, Lect. Notes Phys. 651, 137 (2004).


[11] H. Wollnik, et al., Exp. Proposal for the SIS-FRS-ESR Facilities 1986.
[12] B. Franzke, H. Geissel, and G. Münzenberg, Exp. Proposal for the SIS-FRS-ESR Facilities 1987.
[13] F. Nolden, et al., Nucl. Instr. and Meth. A 532, 329 (2004).
[14] M. Steck, et al., Nucl. Instr. and Meth. A 532, 357 (2004).
[15] T. Radon, et al., Phys. Rev. Lett. 78, 4701 (1997).
[16] H. Geissel, et al., AIP Conf. Proc. Vol. 495, 327 (1999).
[17] F. Nolden, et al., AIP Conf. Proc. Vol. 821, 211 (2006).
[18] H. Irnich, et al., Phys. Rev. Lett. 75, 4182 (1995).
[19] F. Attallah, et al., Nucl. Phys. A 701, 561c (2002).
[20] M. Hausmann, et al., Hyperfine Interactions 132, 291 (2001).
[21] M. Hausmann, et al., Nucl. Instr. and Meth. A 446, 569 (2000).
[22] J. Trötscher, et al., Nucl. Instr. Meth. B 70, 455 (1992).
[23] Yu.A. Litvinov, et al., Nucl. Phys. A 756, 3 (2005).
[24] H. Geissel, et al., Nucl. Phys. A 685, 115c (2001).
[25] Yu.A. Litvinov, et al., Hyperfine Interactions 173, 55 (2006).
[26] H. Geissel and Yu.A. Litvinov, J. Phys. G 31, S1779 (2005).
[27] Yu.A. Litvinov, et al., Hyperfine Interactions 132, 283 (2001).
[28] Z. Liu, Yu.A. Litvinov, and L. Chen, Int. J. Modern Phys. E 15, 1645 (2006).
[29] B. Sun, et al., Eur. Phys. J. A 31, 393 (2007).
[30] T. Radon, et al., Nucl. Phys. A 677, 75 (2000).
[31] Yu.N. Novikov, et al., Nucl. Phys. A 697, 92 (2002).
[32] A. Wapstra, G. Audi, and C. Thibault, Nucl. Phys. A 729, 129 (2003).
[33] H. Geissel, et al., AIP Conf. Proc. Vol. 831, 108 (2006).
[34] F. Bosch, et al., Int. J. Mass Spectrometry 251, 212 (2006).
[35] C. Scheidenberger, et al., Acta Phys. Hung. 10, 177 (1999).
[36] H. Geissel, et al., Nucl. Phys. A 746, 150c (2004).
[37] Yu.A. Litvinov, et al., Nucl. Phys. A 734, 473 (2004).
[38] Yu.A. Litvinov, et al., Nucl. Phys. A 787, 315c (2007).
[39] T. Radon, et al., Pramana 53, 609 (1999).
[40] G.D. Alkhazov, et al., Z. Phys. A 311, 245 (1983).
[41] Yu.A. Litvinov, et al., Phys. Rev. Lett. 95, 042501 (2005).
[42] P. Vogel, et al., Phys. Lett. B 139, 227 (1984).
[43] S. Goriely, et al., http://www-astro.ulb.ac.be.
[44] P. Möller, et al., At. Data Nucl. Data Tables 59, 185 (1995).
[45] S. Goriely, M. Samyn, and J.M. Pearson, Phys. Rev. C 75, 064312 (2007).
[46] C. Scheidenberger, et al., Acta Phys. Hung. 19, 165 (2004).
[47] M. Matoš, PhD thesis, JLU Gießen, 2004.
[48] J. Stadlmann, et al., Phys. Lett. B 586, 27 (2004).
[49] H. Geissel, et al., Hyperfine Interactions 173, 49 (2006).
[50] R. Knöbel, et al., AIP Conf. Proc. Vol. 891, 199 (2007).
[51] Yu.A. Litvinov, et al., Phys. Lett. B 573, 80 (2003).
[52] F. Bosch, Hyperfine Interactions 173, 1 (2006).
[53] F. Bosch, et al., Phys. Rev. Lett. 77, 5190 (1996).
[54] M. Jung, et al., Phys. Rev. Lett. 69, 2164 (1992).
[55] T. Ohtsubo, et al., Phys. Rev. Lett. 95, 052501 (2005).
[56] Yu.A. Litvinov, et al., submitted for publication (2007).
[57] Z. Patyk, et al., submitted for publication (2007).
[58] Yu.A. Litvinov, et al., Exp. Proposal for the SIS-FRS-ESR Facilities E078, 2007.
[59] B. Singh, Nucl. Data Sheets 108, 197 (2007).
[60] Yu.N. Novikov, et al., ILIMA, Letter of Intent, GSI, 2004.
[61] C. Scheidenberger, et al., Hyperfine Interactions 173, 61 (2006).
[62] H. Geissel et al., Nucl. Instr. and Meth. B 204, 71 (2003).
[63] H. Weick, et al., AIP Conf. Proc. Vol. 912, 95 (2007).
[64] F. Nolden, et al., in Proc.: 10th Eur. Part. Acc. Conf., Edinburgh, 2006, p. 208.
[65] A. Dolinskii, et al., Nucl. Instr. and Meth. A 574, 207 (2007).
[66] S.A. Ltvinov, et al., in Proc.: 10th Eur. Part. Acc. Conf., Edinburgh, 2006, p. 1618.
[67] M. Steck, et al., in Proc.: 10th Eur. Part. Acc. Conf., Edinburgh, 2006, p. 217.

You might also like