15
Decay-assisted collinear resonance ionization spectroscopy: Application to neutron-deficient francium K. M. Lynch, 1, 2, 3, * J. Billowes, 1 M. L. Bissell, 3 I. Budin˘ cevi´ c, 3 T.E. Cocolios, 1, 2 R.P. De Groote, 3 S. De Schepper, 3 V.N. Fedosseev, 4 K.T. Flanagan, 1 S. Franchoo, 5 R.F. Garcia Ruiz, 3 H. Heylen, 3 B.A. Marsh, 4 G. Neyens, 3 T.J. Procter, 1, R.E. Rossel, 4, 6 S. Rothe, 4, 6 I. Strashnov, 1 H.H. Stroke, 7 and K.D.A. Wendt 6 1 School of Physics and Astronomy, The University of Manchester, Manchester, M13 9PL, United Kingdom 2 ISOLDE, PH Department, CERN, CH-1211 Geneva-23, Switzerland 3 Instituut voor Kern- en Stralingsfysica, KU Leuven, B-3001 Leuven, Belgium 4 EN Department, CERN, CH-1211 Geneva 23, Switzerland 5 Institut de Physique Nucl´ eaire d’Orsay, F-91406 Orsay, France 6 Institut f¨ ur Physik, Johannes Gutenberg-Universit¨ at Mainz, D-55128 Mainz, Germany 7 Department of Physics, New York University, NY, New York 10003, USA (Dated: January 23, 2018) This paper reports on the hyperfine-structure and radioactive-decay studies of the neutron- deficient francium isotopes 202-206 Fr performed with the Collinear Resonance Ionization Spec- troscopy (CRIS) experiment at the ISOLDE facility, CERN. The high resolution innate to collinear laser spectroscopy is combined with the high efficiency of ion detection to provide a highly-sensitive technique to probe the hyperfine structure of exotic isotopes. The technique of decay-assisted laser spectroscopy is presented, whereby the isomeric ion beam is deflected to a decay spectroscopy station for alpha-decay tagging of the hyperfine components. Here, we present the first hyperfine-structure measurements of the neutron-deficient francium isotopes 202-206 Fr, in addition to the identification of the low-lying states of 202,204 Fr performed at the CRIS experiment. I. INTRODUCTION Recent advances in high-resolution laser spectroscopy have resulted in the ability to measure short-lived iso- topes with yields of less than 100 atoms per second [1, 2]. The Collinear Resonance Ionization Spectroscopy (CRIS) experiment [3], located at the ISOLDE facility, CERN, aims to push the limits of laser spectroscopy further, per- forming hyperfine-structure measurements on isotopes at the edges of the nuclear landscape. It provides a com- bination of high-detection efficiency, high resolution and ultra-low background, allowing measurements to be per- formed on isotopes with yields below, in principle, one atom per second. The first optical measurements of francium were per- formed in 1978. Liberman identified the 7s 2 S 1/2 7p 2 P 3/2 atomic transition, performing hyperfine- structure and isotope-shift measurements first with low- resolution [4] and later with high-resolution laser spec- troscopy [5]. The wavelength of this transition λ(D2) = 717.97(1) nm was in excellent agreement with the predic- tion of Yagoda [6], made in 1932 before francium was dis- covered. Further measurements of francium followed in the next decade. High-resolution optical measurements were performed on both the 7s 2 S 1/2 7p 2 P 3/2 atomic transition [7–9], as well as the 7s 2 S 1/2 8p 2 P 3/2 tran- sition [10], along with transitions into high-lying Ryd- berg states [11]. The CRIS technique, a combination of * [email protected] Present address: TRIUMF, Vancouver, British Columbia, V6T 2A3, Canada collinear laser spectroscopy and resonance ionization was originally proposed by Kudriavtsev in 1982 [12], but the only experimental realization of the technique was not performed until 1991 on ytterbium atoms [13]. The ability to study the neutron-deficient francium (Z = 87) isotopes at the CRIS beam line offers the unique opportunity to answer questions arising from the study of the nuclear structure in this region of the nuclear chart. As the isotopes above the Z = 82 shell closure become more neutron deficient, a decrease in the excitation en- ergy of the (π1i 13/2 ) 13/2 +,(π2f 7/2 ) 7/2 +,(ν 1i 13/2 ) 13/2 + and (π3s -1 1/2 ) 1/2 + states is observed. In 185 Bi (Z = 83) and 195 At (Z = 85), the (π3s -1 1/2 ) 1/2 + deformed intruder state has been observed to be the ground state [14, 15]. Recent radioactive-decay measurements suggest the existence of a (π3s -1 1/2 ) 1/2 + proton intruder state for 203 Fr and, with a lower excitation energy, for 201 Fr, suggest- ing that this state may become the ground state in 199 Fr [16, 17]. The intruder configurations polarize the nucleus, creating significant deformation. From the study of the nuclear structure of the neutron-deficient francium isotopes towards 199 Fr (by measuring the magnetic dipole moments and change in mean-square charge radii of the ground and isomeric states), the quantum configuration of the states and the shape of the nuclei can be investi- gated. Radioactive-decay measurements on the neutron- deficient francium isotopes have aimed to determine the level structure of the low-lying nuclear states, but their exact nature is still unknown [16–19]. High-resolution collinear laser spectroscopy has allowed determination of the ground-state properties of 204,205,206 Fr [20], confirm- ing the tentative spin assignments. The spin of 205 Fr was arXiv:1402.4266v1 [nucl-ex] 18 Feb 2014

arXiv:1402.4266v1 [nucl-ex] 18 Feb 2014 · 2018. 1. 23. · originally proposed by Kudriavtsev in 1982 [12], but the only experimental realization of the technique was not performed

  • Upload
    others

  • View
    1

  • Download
    0

Embed Size (px)

Citation preview

Page 1: arXiv:1402.4266v1 [nucl-ex] 18 Feb 2014 · 2018. 1. 23. · originally proposed by Kudriavtsev in 1982 [12], but the only experimental realization of the technique was not performed

Decay-assisted collinear resonance ionization spectroscopy:Application to neutron-deficient francium

K. M. Lynch,1, 2, 3, ∗ J. Billowes,1 M. L. Bissell,3 I. Budincevic,3 T.E. Cocolios,1, 2 R.P. De Groote,3 S. De

Schepper,3 V.N. Fedosseev,4 K.T. Flanagan,1 S. Franchoo,5 R.F. Garcia Ruiz,3 H. Heylen,3 B.A. Marsh,4

G. Neyens,3 T.J. Procter,1, † R.E. Rossel,4, 6 S. Rothe,4, 6 I. Strashnov,1 H.H. Stroke,7 and K.D.A. Wendt6

1School of Physics and Astronomy, The University of Manchester, Manchester, M13 9PL, United Kingdom2ISOLDE, PH Department, CERN, CH-1211 Geneva-23, Switzerland

3Instituut voor Kern- en Stralingsfysica, KU Leuven, B-3001 Leuven, Belgium4EN Department, CERN, CH-1211 Geneva 23, Switzerland

5Institut de Physique Nucleaire d’Orsay, F-91406 Orsay, France6Institut fur Physik, Johannes Gutenberg-Universitat Mainz, D-55128 Mainz, Germany

7Department of Physics, New York University, NY, New York 10003, USA(Dated: January 23, 2018)

This paper reports on the hyperfine-structure and radioactive-decay studies of the neutron-deficient francium isotopes 202−206Fr performed with the Collinear Resonance Ionization Spec-troscopy (CRIS) experiment at the ISOLDE facility, CERN. The high resolution innate to collinearlaser spectroscopy is combined with the high efficiency of ion detection to provide a highly-sensitivetechnique to probe the hyperfine structure of exotic isotopes. The technique of decay-assisted laserspectroscopy is presented, whereby the isomeric ion beam is deflected to a decay spectroscopy stationfor alpha-decay tagging of the hyperfine components. Here, we present the first hyperfine-structuremeasurements of the neutron-deficient francium isotopes 202−206Fr, in addition to the identificationof the low-lying states of 202,204Fr performed at the CRIS experiment.

I. INTRODUCTION

Recent advances in high-resolution laser spectroscopyhave resulted in the ability to measure short-lived iso-topes with yields of less than 100 atoms per second [1, 2].The Collinear Resonance Ionization Spectroscopy (CRIS)experiment [3], located at the ISOLDE facility, CERN,aims to push the limits of laser spectroscopy further, per-forming hyperfine-structure measurements on isotopes atthe edges of the nuclear landscape. It provides a com-bination of high-detection efficiency, high resolution andultra-low background, allowing measurements to be per-formed on isotopes with yields below, in principle, oneatom per second.

The first optical measurements of francium were per-formed in 1978. Liberman identified the 7s 2S1/2 →7p 2P3/2 atomic transition, performing hyperfine-structure and isotope-shift measurements first with low-resolution [4] and later with high-resolution laser spec-troscopy [5]. The wavelength of this transition λ(D2) =717.97(1) nm was in excellent agreement with the predic-tion of Yagoda [6], made in 1932 before francium was dis-covered. Further measurements of francium followed inthe next decade. High-resolution optical measurementswere performed on both the 7s 2S1/2 → 7p 2P3/2 atomic

transition [7–9], as well as the 7s 2S1/2 → 8p 2P3/2 tran-sition [10], along with transitions into high-lying Ryd-berg states [11]. The CRIS technique, a combination of

[email protected]† Present address: TRIUMF, Vancouver, British Columbia, V6T

2A3, Canada

collinear laser spectroscopy and resonance ionization wasoriginally proposed by Kudriavtsev in 1982 [12], but theonly experimental realization of the technique was notperformed until 1991 on ytterbium atoms [13].

The ability to study the neutron-deficient francium(Z = 87) isotopes at the CRIS beam line offers the uniqueopportunity to answer questions arising from the study ofthe nuclear structure in this region of the nuclear chart.As the isotopes above the Z = 82 shell closure becomemore neutron deficient, a decrease in the excitation en-ergy of the (π1i13/2)13/2+ , (π2f7/2)7/2+ , (ν1i13/2)13/2+

and (π3s−11/2)1/2+ states is observed. In 185Bi (Z = 83)

and 195At (Z = 85), the (π3s−11/2)1/2+ deformed intruder

state has been observed to be the ground state [14, 15].Recent radioactive-decay measurements suggest the

existence of a (π3s−11/2)1/2+ proton intruder state for 203Fr

and, with a lower excitation energy, for 201Fr, suggest-ing that this state may become the ground state in199Fr [16, 17]. The intruder configurations polarize thenucleus, creating significant deformation. From the studyof the nuclear structure of the neutron-deficient franciumisotopes towards 199Fr (by measuring the magnetic dipolemoments and change in mean-square charge radii of theground and isomeric states), the quantum configurationof the states and the shape of the nuclei can be investi-gated.

Radioactive-decay measurements on the neutron-deficient francium isotopes have aimed to determine thelevel structure of the low-lying nuclear states, but theirexact nature is still unknown [16–19]. High-resolutioncollinear laser spectroscopy has allowed determination ofthe ground-state properties of 204,205,206Fr [20], confirm-ing the tentative spin assignments. The spin of 205Fr was

arX

iv:1

402.

4266

v1 [

nucl

-ex]

18

Feb

2014

Page 2: arXiv:1402.4266v1 [nucl-ex] 18 Feb 2014 · 2018. 1. 23. · originally proposed by Kudriavtsev in 1982 [12], but the only experimental realization of the technique was not performed

2

measured to be 9/2−, the ground-state spins of 204,206Frwere confirmed as 3(+), but the low-lying spin (7+) and(10−) isomers are still under investigation.

General methods of isomer identification have alreadybeen achieved with in-source laser spectroscopy [21] (andreferences therein). In the case of 68,70Cu [22], follow-ing the selection of isomeric beams, experiments suchas Coulomb excitation [23] and mass measurements [24]have been performed. However, these experiments suf-fered from isobaric contamination, as well as significantground-state contamination due to the Doppler broaden-ing of the hyperfine resonances of each isomer [1]. Oneway of addressing the difficulties of in-source laser spec-troscopy (isobaric contamination, Doppler broadening,pressure broadening) is selecting the ground or isomericstate of interest by resonance ionization in a collineargeometry.

In a sub-Doppler geometry, the process of isomer selec-tive resonance laser ionization [21] can result in a high-purity isomeric beam. Deflection of the pure-state ionbeam to the decay spectroscopy station allows identifica-tion of the hyperfine component with alpha-decay spec-troscopy.

II. EXPERIMENTAL TECHNIQUE

Radioactive ion beams of francium were produced atthe ISOLDE facility, CERN [25] by impinging 1.4 GeVprotons onto a thick UCx target (up to 2 µA integratedproton current). The radioisotopes were surface ion-ized through interaction with the rhenium coating on thehot (2400 K) tantalum transfer line and extracted fromthe target-ion source at 50 keV. The isotope of interestwas mass selected using the high-resolution HRS separa-tor and bunched (at 31.25 Hz) with the radio-frequencycooler-buncher ISCOOL [26, 27]. The bunched-ion beamwas deflected into the CRIS beam line and transportedthrough a potassium-vapour charge exchange cell (CEC)(420 K, ∼10−6 mbar chamber pressure, 6×10−4 mbarvapour pressure [28]) to be neutralized. In the 1.2 mlong interaction region, the arrival of the atomic bunchwas synchronized with two co-propagating pulsed laserbeams to excite the state of interest followed by ion-ization in a step-wise scheme. The temporal length ofthe atomic bunch was 2-3 µs, corresponding to a spa-tial length of 45-70 cm. To reduce the background signalresulting from non-resonant collisional ionization, the in-teraction region aims at ultra-high vacuum (UHV) con-ditions. A pressure of <10−8 mbar was achieved duringthis experiment. A schematic diagram of the CRIS beamline is shown in Fig. 1.

A. Collinear resonance ionization spectroscopy

The resonant excitation step from the 7s 2S1/2 elec-

tronic ground state to the 8p 2P3/2 state was probed

with 422.7-nm light. The laser light of this resonantstep was provided by a narrow-band titanium:sapphire(Ti:Sa) laser of the ISOLDE RILIS installation [29, 30],pumped by the second harmonic output of a Nd:YAGlaser (Model: Photonics Industries DM-60-532, 10 kHz).The fundamental output from the tuneable Ti:Sa laserwas frequency doubled using a BBO crystal to pro-duce the required 422.7-nm laser light. The light wasfibre-coupled into the CRIS beam line through 35 m ofmultimode optical fibre (∼100 mW output). The laserlinewidth of 1.5 GHz limited the resolution achieved inthe present experiment, allowing only the lower-state(7s 2S1/2) splitting to be fully resolved. The second

(non-resonant) transition from the 8p 2P3/2 state to thecontinuum was driven using 1064-nm light. This lightwas produced by a fundamental Nd:YAG laser (Model:Quanta-Ray LAB 130, operated at 31.25 Hz) next to theCRIS setup, temporally overlapped with the 422.7-nmlaser beam and aligned through the laser/atom interac-tion region. The standard repetition rate of the RILISlasers (10 kHz) limited the repetition rate of the 1064-nm laser light to 31.25 Hz (one out of every 320 pulses of422.7-nm laser light was utilized). The bunching of theion beam with ISCOOL was matched to the lower repe-tition rate of 31.25 Hz to overlap the atom bunch withthe two laser pulses every 32 ms.

The synchronization of the first- and second-step laserpulses and the release of the ion bunch from ISCOOLwas controlled by a Quantum Composers digital delaygenerator (Model: QC9258). The 10 kHz pulse genera-tor of the Ti:Sa pump laser acted as the master clock,triggering the delay generator to output a sequence ofTTL pulses to synchronize the 1064-nm laser light andthe ion bunch with the 422.7-nm light, allowing resonanceionization of the francium atoms to occur. The laser ionswere detected by a micro-channel plate (MCP) housedin the decay spectroscopy station (DSS). The electronicsignal from the MCP was digitized by a LeCroy oscillo-scope (Model: WavePro 725Zi, 2 GHz bandwidth, 8 bitADC, 20 GS/s), triggered by the digital delay generator.The data were transferred from the oscilloscope using aLabVIEWTM program.

The frequency of the resonant excitation step, the422.7-nm laser light, was scanned to study the 7s 2S1/2 →8p 2P3/2 atomic transition. The scanning and stabi-lization of the frequency was controlled by the RILISEquipment Acquisition and Control Tool (REACT), aLabVIEW control program package that allows for re-mote control, equipment monitoring and data acquisi-tion [31]. This was achieved by controlling the etalontilt angle inside the Ti:Sa laser resonator to adjust thelaser wavelength, which was measured with a HighFi-nesse wavemeter (Model: WS7), calibrated with a fre-quency stabilised HeNe laser. The francium experimen-tal campaign at CRIS marked the first implementationof the REACT framework for external users. The remotecontrol LabVIEW interface for the Ti:Sa laser ran locallyat the CRIS setup, allowing independent laser scanning

Page 3: arXiv:1402.4266v1 [nucl-ex] 18 Feb 2014 · 2018. 1. 23. · originally proposed by Kudriavtsev in 1982 [12], but the only experimental realization of the technique was not performed

3

Pulsed laser beams

Bunched radioactive beam from ISOLDE

50 keV 30° bend

Charge exchange cell ~10-6 mbar

420 K Differential

pumping region

Deflector plates Dumped ions

Interaction region (UHV) <10-8 mbar

Deflector plates Dumped

atoms Quadrupole

doublet

Horizontal and vertical deflection plates

Copper plate

MCP The DSS

See inset

The ‘windmill’ system

APIPS PIPS

Collimator

Deflected ions

Carbon foil

1.93 m

1.2 m

FIG. 1. Schematic diagram of the CRIS beam line. Laser ions can be deflected to a copper plate and the correspondingsecondary electrons detected by the MCP, or implanted into a carbon foil for alpha-decay spectroscopy. (Inset) The decayspectroscopy station (DSS) ‘windmill’ system for alpha-decay tagging.

and control.

B. Decay-assisted laser spectroscopy

The technique of decay-assisted collinear laser spec-troscopy was further developed at the CRIS beam lineto take advantage of the ultra-pure ion beams producedby resonance ionization in a collinear geometry [32]. Theselectivity from resonance ionization of an isotope is a re-sult of the selectivity of the Lorentzian profile of the nat-ural linewidth (∼12.5 MHz) of the state and the Gaussianprofile of the laser linewidth (∼1.5 GHz). At a frequencyseparation of 4 GHz, the Gaussian component falls to 1%of its peak intensity and the selectivity is dominated bythe natural linewidth of the state. Thus, the maximumselectivity from resonance ionization is given by Eq. (1),

S =

N∏n=1

(∆ωAB,n

Γn

)2=

N∏n=1

Sn, (1)

where ∆ωAB is the separation in frequency of the twostates (A and B), Γn is the FWHM of the naturallinewidth of the state, Sn is the selectivity of the transi-tion and N is the number of transitions used. The totalselectivity of a resonance ionization process is given bythe product of the individual selectivities. In the caseof the two states being the ground state and isomer, theselectivity can be calculated from Eq. (1). When thetwo states are the isotope of interest and contaminationfrom a neighbouring isotope, additional selectivity canbe gained from the kinematic shift since the laser is over-lapped with an accelerated beam.

In addition to hyperfine-structure studies with iondetection, the decay spectroscopy station can be usedto identify the hyperfine components of overlappingstructures. This allows the hyperfine structure of twostates to be separated by exploiting their characteristicradioactive-decay mechanisms. This results in a smallererror associated with the hyperfine parameters, and a

better determination of the extracted nuclear observ-ables.

The decay spectroscopy station (DSS) consists of a ro-tatable wheel implantation system [33]. It is based on thedesign from KU Leuven [34] (Fig. 1 of Ref. [35]), whichhas provided results in a number of successful experi-ments [36] (and references therein). The wheel holds 9carbon foils, produced at the GSI target laboratory [37],with a thickness of 20(1) µg cm−2 (∼90 nm) into whichthe ion beam is implanted (at a depth of ∼25 nm).

Two Canberra Passivated Implanted Planar Silicon(PIPS) detectors for charged-particle detection (e.g. al-pha, electron, fission fragments) are situated on eitherside of the implantation carbon foil, as shown in Fig. 1.One PIPS detector (Model: BKA 300-17 AM, thickness300 µm) sits behind the carbon foil and another annularPIPS (APIPS) (Model: BKANPD 300-18 RM, thickness300 µm, with an aperture of 4 mm) is placed in front ofthe carbon foil. The detectors are connected to chargesensitive Canberra preamplifiers (Model: 2003BT) via aUHV type-C sub-miniature electrical feed-through.

Laser-produced ions from the interaction region in theCRIS beam line are deflected to the DSS by applyinga potential difference between a pair of vertical electro-static plates, see Fig. 1. The deflected ion beam is im-planted into the carbon foil, after passing though a col-limator with a 4 mm aperture and the APIPS detector.The collimator shields the APIPS detector from directimplantation of radioactive ions into the silicon wafer,see Fig. 1. Decay products from the carbon foil can bemeasured by either the APIPS or PIPS detector, with atotal solid-angle coverage of 63% (simulated, assuming auniform distribution of implanted activity). Operation ofthe single APIPS detector during the experiment gave analpha-detection efficiency of 25%. An electrical contactis made to the collimator, allowing the current gener-ated by the ion beam when it strikes the collimator to bemeasured and the plate to be used as a beam monitoringdevice. When it is not in use, it is electrically groundedto avoid charge build-up. A Faraday cup is installed in

Page 4: arXiv:1402.4266v1 [nucl-ex] 18 Feb 2014 · 2018. 1. 23. · originally proposed by Kudriavtsev in 1982 [12], but the only experimental realization of the technique was not performed

4

the location of one of the carbon foils. This copper plate(thickness 0.5 mm, diameter 10 mm) is electrically iso-lated from the steel wheel by PEEK rings and connectedby a spot-welded Kapton cable attached to a rotatableBNC connection in the centre of the wheel [33].

The alpha-decay spectroscopy data is acquired with adigital data acquisition system (DAQ), consisting of XIAdigital gamma finder (DGF) revision D modules [38].Each module has four input channels with a 40 MHzsampling rate. Signals fed into the digital DAQ are self-triggered with no implementation of master triggers.

Due to the reflective surface of the inside of the vacuumchambers, a significant fraction of 1064-nm laser lightwas able to scatter into the silicon detectors. Despitethe collimator in front of the APIPS detector to protectit from ion implantation (and laser light), the infra-redlight caused a shift in the baseline of the signal from thesilicon detector. This required the parameters for theDGF modules to be adjusted to account for this effectonline, since the reflections were due to the particularsetup of the experiment (power and laser-beam path).

The low energy resolution of the APIPS detector wasassociated with the necessity of optimizing the DGFparameters online with the radioactive 221Fr (t1/2 =4.9(2) min). In addition, a fluctuating baseline result-ing from the changing power of the 1064-nm laser lightmeant that only a resolution of 30 keV at 6.341 MeVwas achieved. This however was sufficient to identify thecharacteristic alpha decays of the neutron-deficient fran-cium isotopes under investigation.

III. RESULTS

The hyperfine structures of the neutron-deficient fran-cium isotopes 202−206Fr were measured with collinearresonance ionization spectroscopy, with respect to thereference isotope 221Fr. This paper follows the recentpublication reporting the hyperfine-structure studies of202,203,205Fr [39]. During the experimental campaign,the neutron-rich francium isotopes 218m,219,229,231Fr werealso studied. A detailed description of the nature of theseisotopes will be the topic of a future publication [40].

The resonance spectrum of the 7s 2S1/2 → 8p 2P3/2

transition was fit with a χ2-minimization routine. Thehyperfine AP3/2

factor was fixed to the ratio of the

7s 2S1/2 → 8p 2P3/2 transition of AP3/2/AS1/2

=

+22.4/6209.9, given in literature [10]. For the 8p 2P3/2

state, the hyperfine BP3/2factor is small enough to have

no impact on the fit to the data, and was consequentlyset to zero [41].

The intensities of the hyperfine transitions SFF ′ be-tween hyperfine levels F and F ′ (with angular momen-tum J and J ′ respectively) are related to the intensityof the underlying fine structure transition SJJ ′ [2]. Therelative intensities of the hyperfine transitions are given

0 10000 20000 30000 40000 50000 60000 70000Relative frequency (MHz)

0

50

100

150

200

250

300

350

400

Counts

A

B

C

D

E

204gFr204m1 Fr204m2 FrData

FIG. 2. Collinear resonance ionization spectroscopy of 204Frrelative to 221Fr. The hyperfine structure of the 3(+) groundstate of 204gFr is shown in blue, the 7+ state of 204m1Fr isshown in green and the (10−) state of 204m2Fr is shown inred.

200At 43(1) s

0 keV(3+)

47(1) ms

113(1) keV(7+)

112(5) keV(2,4+)

7.3 s

344(1) keV(10�)

204Fr 1.9(5) s

0 keV(3+)

1.6(5) ms

41(7) keV(7+)

0.8(2) s

316(7) keV(10�)

E↵=7031(5) keVI=99.4(2) %

E↵=6969(5) keVI=99.3(2) %

E↵=7013(5) keVI=100%

E�=230.9(2) keV

E�=113(1) keV

E↵=6916(8) keVI=0.6(2) %

E↵=7077(8) keVI=0.7(2) %

E�=275 keV

1

FIG. 3. The radioactive decay of 204Fr and its isomers [16,18, 19].

by

SFF ′

SJJ ′= (2F + 1)(2F ′ + 1)

{F F ′ 1J ′ J I

}2

, (2)

where {. . .} denotes the Wigner 6-j coefficient. Althoughthese theoretical intensities are only strictly valid forclosed two-level systems, and there was jitter on the tem-poral overlap of the two laser pulses in the interactionregion, they were used as currently the most reliable es-timate.

The AS1/2factor and the centroid frequency of the

hyperfine structure were determined for each scan indi-vidually. For isotopes with multiple scans, a weightedmean for the AS1/2

factor and the centroid frequency

Page 5: arXiv:1402.4266v1 [nucl-ex] 18 Feb 2014 · 2018. 1. 23. · originally proposed by Kudriavtsev in 1982 [12], but the only experimental realization of the technique was not performed

5

6200 6400 6600 6800 7000 7200Energy (keV)

0

50

100

150

200C

ounts

/ 4

keV

70

31

keV

69

69

keV

70

13

keV

64

64

keV

64

12

keV

65

37

keV

204gFr

204m1 Fr

204m2 Fr

200gAt

200m1 At

200m2 At

204gFr204m1 Fr204m2 Fr

FIG. 4. Alpha-particle spectroscopy of (blue) 204gFr,(green) 204m1Fr and (red) 204m2Fr allowed the hyperfinepeaks in Fig. 2 to be identified. The laser was detuned by11.503 GHz (peak A, 204gFr), 8.508 GHz (peak B, 204m1Fr)and 18.693 GHz (peak C, 204m2Fr) relative to the centroidfrequency of 221Fr.

were calculated based on the error of the fits. The uncer-tainty attributed to the AS1/2

factor was calculated as theweighted standard deviation of the values. The isotopeshifts were determined relative to 221Fr, with the uncer-tainty propagated from the error of the fits, the scatterand the drift in centroid frequency of the 221Fr referencescans [41].

A. Spectroscopic studies of 204Fr

The hyperfine structure of 204Fr is shown in Fig. 2,measured by detecting the laser ions with the MCP detec-tor as a function of the scanned first-step laser frequency.Five peaks are observed in the spectrum. Consideringthat only the lower-state splitting is resolved (associatedwith the 1.5 GHz linewidth of the scanning laser), twohyperfine resonances are expected per nuclear (groundor isomeric) state. Consequently, Fig. 2 contains the hy-perfine structure of three long-lived states in 204Fr, withone of the resonances unresolved (labeled E). In orderto identify the states of the hyperfine resonances, laserassisted alpha-decay spectroscopy was used.

The radioactive decay of the low-lying states in 204Fris presented in Fig. 3. The characteristic alpha decay ofeach nuclear state in 204Fr was utilized to identify thehyperfine-structure resonances of Fig. 2. The laser wastuned on resonance with each of the first three hyperfineresonances (labeled A, B and C) and alpha-decay spec-troscopy was performed on each. The alpha-particle en-ergy spectrum of these three states is illustrated in Fig. 4.The energy of the alpha particles emitted when the laserwas on resonance with an atomic transition of the hyper-

6200 6400 6600 6800 7000 7200Energy (keV)

0

20

40

60

80

100

120

Counts

(1

keV

per

bin

)

70

13

keV

204m2 Fr

64

12

keV

200m1 At

65

37

keV

204m2 At

69

69

keV

204m1 Fr

FIG. 5. Alpha-particle spectroscopy of the (10−) stateof 204m2Fr. The decay of 204m2Fr to 204m1Fr via an E3IT is observed through the presence of 204m1Fr alpha par-ticles of 6969 keV (denoted by ?). The laser was detuned by43.258 GHz relative to the centroid frequency of 221Fr.

fine spectrum characteristic of 204gFr is shown in blue.This transition occurred at 11.503 GHz (peak A of Fig. 2)relative to the centroid frequency of 221Fr. Similarly, thealpha-particle energy spectra for 204m1Fr and 204m2Fr areshown in green and red, when the laser was detuned by8.508 GHz and 18.693 GHz (peak B and C) from thereference frequency, respectively. Present in the alpha-particle energy spectrum are the alpha particles emittedfrom the decay of the 204Fr states (6950-7050 keV) inaddition to those emitted from the nuclear states in thedaughter isotope 200At (6400-6500 keV). Each state in204Fr has a characteristic alpha-particle emission energy:7031 keV for 204gFr, 6969 keV for 204m1Fr and 7013 keVfor 204m2Fr. This was confirmed by the presence ofthe corresponding daughter decays of 200gAt (6464 keV),200m1Fr (6412 keV), and 200m2At (6537 keV) in the alpha-particle energy spectrum.

An additional alpha-decay measurement was per-formed on peak D in the hyperfine spectrum of 204Fr (seeFig. 2) at 43.258 GHz relative to the centroid frequencyof 221Fr. The observation of 7013 keV alpha particles al-lowed this state to be identified as 204m2Fr. This meantthe identity of all five hyperfine-structure peaks could beallocated to a state in 204Fr (hence the hyperfine struc-ture peak E is the overlapping structure of 204gFr and204m1Fr), allowing analysis of the hyperfine structure ofeach state.

In addition to the 7031 keV alpha particles of 204m2Fr,alpha particles of 6969 keV from the decay of 204m1Frwere also observed when the laser was on resonance withthe 204m2Fr state. The decay of the (10−) state to204m1Fr via an E3 internal transition (IT) has been pre-dicted [18] but only recently observed [17], see Fig. 3.This was achieved by tagging the conversion electronfrom the internal conversion of 204m2Fr with the emitted

Page 6: arXiv:1402.4266v1 [nucl-ex] 18 Feb 2014 · 2018. 1. 23. · originally proposed by Kudriavtsev in 1982 [12], but the only experimental realization of the technique was not performed

6

0 10000 20000 30000 40000 50000 60000 700000

20406080

100120140160180

Counts

0 10000 20000 30000 40000 50000 60000 70000Relative frequency (MHz)

6900

6950

7000

7050

7100

Energ

y (

4 k

eV

per

bin

)

1.5

3.0

4.5

6.0

7.5

9.0

10.5

12.0

FIG. 6. A two-dimensional histogram of the alpha-particleenergy as the hyperfine structure of 204Fr is probed. (Top)Projection of the frequency axis. The total number of alphaparticles detected at each laser frequency reveals the hyperfinestructure of 204Fr.

6969 keV alpha particles of 204m1Fr that followed (witha 5 s correlation time). This allowed the predicted en-ergy of the 275 keV isomeric transition to be confirmed.During the CRIS experiment, an additional alpha-decaymeasurement was performed on 204m2Fr, with the laserdetuned by 43.258 GHz relative to the centroid frequencyof 221Fr, see Fig. 5. This spectrum confirms the presenceof the 6969 keV alpha particle (denoted by ?), emittedfrom the decay of 204m1Fr. The ultra-pure conditionsof this measurement allowed the first unambiguous ex-traction of the branching ratios in the decay of 204m2Fr:Bα = 53(10)% and BIT = 47(10)% [42].

Decay-assisted laser spectroscopy was also performedon the hyperfine structure of the low-lying states of 204Fr.Just as the laser frequency of the resonant 422.7-nmionization step was scanned and resonant ions were de-tected in the collinear resonance ionization spectroscopyof 204Fr, the same technique was repeated with the mea-surement of alpha particles. At each laser frequency, aradioactive-decay measurement of 60 s was made at theDSS, measuring the alpha particles emitted from the im-planted ions. Fig. 6 (Top) shows the hyperfine peaksassociated with each state in 204Fr. Measurement of thealpha decay as a function of laser frequency allowed pro-duction of a matrix of alpha-particle energy versus laserfrequency, see Fig. 6.

In order to separate hyperfine structures for eachstates, an alpha-energy gating was used to maximize thesignal-to-noise ratio for the alpha particle of interest. The

0

30

60

90

120

Alpha-energy gate:7031 - 7200 keV

(a) 204gFr

0

20

40

606959 - 6979 keV(b) 204m1 Fr

0 10000 20000 30000 40000 50000 60000 70000Relative frequency (MHz)

0

20

40

607003 - 7023 keV(c) 204m2 Fr

FIG. 7. Alpha-tagged hyperfine structure of (a) 204gFr

spin 3(+) ground state, (b) 204m1Fr spin (7+) isomer, and (c)204m2Fr spin (10−) isomer. Additional peaks are discussed inthe text.

alpha-energy gates were chosen to be 7031-7200 keV for204gFr, 6959-6979 keV for 204m1Fr and 7003-7023 keV for204m2Fr.

By gating on the characteristic alpha-particle energiesof the three states in 204Fr, the hyperfine structures of in-dividual isomers become enhanced in the hyperfine spec-trum. Fig. 7(a) shows the hyperfine structure of 204gFr,(b) 204m1Fr, and (c) 204m2Fr. The presence of 204m2Fr canbe observed in the spectra of 204gFr due to the overlap-ping peaks of the alpha energies: the tail of the 7013 keValpha peak is present in the gate of the 204gFr alphapeak. The presence of 204m2Fr in the hyperfine structurespectrum of 204m1Fr is attributed to the E3 IT decay of204m2Fr to 204m1Fr: alpha particles of energy 6969 keVare observed when on resonance with 204m2Fr. Addi-tionally, 204gFr is present in the 204m2Fr spectra due tothe similar energies of the 7031 keV and 7013 keV al-pha particles. However, despite the contamination in thehyperfine spectra, each peak is separated sufficiently infrequency to be analysed independently.

From the resulting hyperfine structures of Fig. 7 pro-duced by the alpha-tagging process (in comparison to theoverlapping ion data of Fig. 2), each state of 204Fr can beanalysed individually and the hyperfine factors extractedwith better accuracy and reliability. The estimated errorof the AS1/2

factors was 30 MHz on account of the scatter

of AS1/2values for 221Fr. Likewise, an error of 100 MHz

was assigned to the isotope shifts.

B. Identification of the hyperfine structure of 202Fr

The hyperfine structure of 202Fr obtained withcollinear resonance ionization spectroscopy is presentedin Fig. 8. The four hyperfine resonances illustrate the

Page 7: arXiv:1402.4266v1 [nucl-ex] 18 Feb 2014 · 2018. 1. 23. · originally proposed by Kudriavtsev in 1982 [12], but the only experimental realization of the technique was not performed

7

0 10000 20000 30000 40000 50000 60000 70000 80000Relative frequency (MHz)

0

5

10

15

20

25

30C

ounts

A

B

202gFr202mFrData

FIG. 8. Collinear resonance ionization spectroscopy of 202Frrelative to 221Fr. The hyperfine structure of the (3+) groundstate of 202gFr is shown in blue and the (10−) state of 202mFris shown in red.

198At 3.8 s

0 keV(3+)

1.04 s

102+� keV(10�)

202Fr 0.30(5) s

0 keV(3+)

0.29(5) s

102+� keV(10�)

E↵=7241(8) keVI=100 %

E↵=7235(8) keVI=100%

1

FIG. 9. The radioactive decay of 202Fr and its isomer [43].

presence of the ground (202gFr) and isomeric (202mFr)states. Identification of these two states was performedwith laser-assisted alpha-decay spectroscopy. Accord-ing to literature, the radioactive decay of 202gFr (t1/2 =0.30(5) s) emits an alpha particle of energy 7241(8) keV,whereas 202mFr (t1/2 = 0.29(5) s) emits an alpha particleof energy 7235(8) keV [44]. The radioactive decay of theground and isomeric state of 202Fr is presented in Fig. 9.

The laser was tuned onto resonance with peak A(202gFr, 13.760 GHz relative to the centroid frequencyof 221Fr) and peak B (202mFr, 20.950 GHz relative tothe centroid frequency of 221Fr) of Fig. 8 obtained fromion detection. For each position, an alpha-decay mea-surement was performed, shown in Fig. 10. The alphaparticles emitted when the laser was on resonance withan atomic transition characteristic to 202gFr are shownin blue, and 202mFr in red. Due to the limited statis-tics of our measurement, and the similarity in energies ofthe alpha particles (within error), it is impossible to say

0

2

4

Counts

/ 4

keV

67

55

keV

198gAt

72

41

keV

202gFr(a) 202gFr

6600 6700 6800 6900 7000 7100 7200 7300 7400Energy (keV)

0

2

4

Counts

/ 4

keV

68

56

keV

198mAt

72

35

keV

202mFr(b) 202mFr

FIG. 10. Alpha-particle spectroscopy of (blue) 202gFr and(red) 202mFr allowed the hyperfine peaks in Fig. 8 to be iden-tified. The laser was detuned by (a) 13.760 GHz (peak A,202gFr) and (b) 20.950 GHz (peak B, 202mFr) relative to thecentroid frequency of 221Fr.

that alpha particles of different energies are observed inFig. 10.

Firm identification of the hyperfine components can beachieved however by studying the alpha particles emit-ted by the daughter isotopes 198g,mAt. Evident in thespectrum of 202gFr are the alpha particles emitted fromthe decay of the daughter nucleus 198gAt with an energyof 6755 keV. Similarly, present in the 202mFr spectrumare the alpha particles from the decay of 198mAt with anenergy of 6856 keV. The difference in energy of these twoalpha peaks illustrates the ability of the CRIS techniqueto separate the two states and provide pure ground stateand isomeric beams for decay spectroscopy.

C. Isomer identification of the resonance spectrumof 206Fr

Two sets of data were used in the determination of thenuclear observables from the hyperfine structure scans.The data for the francium isotopes 202−206,221Fr weretaken in Run I and the data for 202−205,221Fr were takenin Run II. Consistency checks were carried out, allowing206Fr to be evaluated with respect to the rest of the dataset from Run II. A detailed description of this analysiscan be found in Ref. [42]. In Run I, no alpha-tagging wasavailable and consequently the peaks in the ion-detectedhyperfine spectrum needed to be identified in a differentway. Recent measurements of the ground-state hyper-fine structure of 206gFr provided the AS1/2

factor for the

splitting of the 7s 2S1/2 state [20]. One peak of the (7+)isomeric state was also identified in this experiment (seeFig. 1(c) of Ref. [20]), allowing the positions of the over-lapping resonances to be determined. This left only the

Page 8: arXiv:1402.4266v1 [nucl-ex] 18 Feb 2014 · 2018. 1. 23. · originally proposed by Kudriavtsev in 1982 [12], but the only experimental realization of the technique was not performed

8

0

150

300

450

600

A B

(a) 206gFr206m1 Fr206m2 FrData

10000 0 10000 20000 30000 40000 50000 60000 70000Relative frequency (MHz)

0

150

300

450

600

A B

(b)

FIG. 11. Collinear resonance ionization spectroscopy of 206Frrelative to 221Fr. (a) Option 1: Peak A is assigned to 206m2Frand peak B to 206m1Fr. (b) Option 2: Peak A is assigned to206m1Fr and peak B to 206m2Fr.

identity of peaks A and B (shown in Fig. 11) unknown.Fig. 11(a) presents the hyperfine structures when peak Ais assigned to 206m2Fr and peak B to 206m1Fr. Fig. 11(b)shows the fit when peak A is 206m1Fr and peak B is206m2Fr. The suggested identity of the two resonances(based on mean-square charge radii and g-factor system-atics) is discussed in Sec. IV.

D. Yield measurements

TABLE I. Yields of the neutron-deficient francium isotopesat the ISOLDE facility (1.4 GeV protons on a UCx target).The nuclear-state composition of the radioactive beams for202,204,206Fr are presented.

A Yield Proportion of beam

(ions/s) Spin 3(+) Spin (7+) Spin (10−)

202 1× 102 76(14)% 24(6)%

203 1× 103

204 1× 104a 63(3)% 27(3)% 10(1)%

205 2× 105

206 3× 106 63(7)% 27(5)% 9(1)%

a Estimate based on yield systematics.

The yields of the neutron-deficient francium isotopes202−206Fr are presented in Table I. The quoted yields,scaled ISOLDE-database yields based on an independentyield measurement of 202Fr, can be expected to vary bya factor of two due to different targets. The quoted valuefor 204Fr is estimated based on francium yields systemat-ics. The composition of the beam for 202,204,206Fr wascalculated from the ratio of hyperfine-peak intensities

1.5 1.4 1.3 1.2 1.1 1.0 0.9 0.8 0.7µA,A

′δνA,A

718 (MHz amu) 1e8

1.5

1.4

1.3

1.2

1.1

1.0

0.9

0.8

0.7

0.6

µA,A

′ δνA,A

422

(M

Hz

am

u)

1e8

[221,207]

[221,211]

[221,212]

[221,220]

[212, 213]µA,A

′δνA,A

422 = F422

F718µA,A

′δνA,A

718 +M422−F422

F718M718

King plot

FIG. 12. A King plot for the extraction of atomic factors Fand M for the 422.7-nm transition. See text for details.

(based on the strongest hyperfine-structure resonance)from the CRIS ion data. The composition of the beamfor 202Fr was confirmed with the alpha-decay data mea-sured with the DSS.

E. King-plot analysis

The atomic factors F and M were evaluated by theKing-plot method [52]. This combines the previouslymeasured isotope shifts by Coc [7] of the 7s 2S1/2 →7p 2P3/2 transition with 718-nm laser light, with those

made by Duong [10] of the 7s 2S1/2 → 8p 2P3/2 tran-

sition (422.7 nm). The isotope shifts of δν207,221 andδν211,221 from this work were combined with δν220,221

and δν213,212 from Duong. These values were plottedagainst the corresponding isotope shifts from Coc [7],shown in Fig. 12. From the linear fit of the data, andusing

µA,A′δνA,A

422 =F422

F718µA,A

′δνA,A

718 +M422−F422

F718M718, (3)

where µA,A′

= AA′/(A′ − A), enabled the evaluation ofF422/F718 = +0.995(3) and M422 − (F422/F718)M718 =+837(308) GHz amu respectively. From these values, theatomic factors for the 422.7-nm transition were calcu-lated to be

F422 = −20.67(21) GHz/fm2,

M422 = +750(330) GHz amu.

For comparison, the atomic factors evaluated forthe 718-nm transition were determined by Dzuba tobe F718 = −20.766(208) GHz/fm2 and M718 =−85(113) GHz amu [51].

Page 9: arXiv:1402.4266v1 [nucl-ex] 18 Feb 2014 · 2018. 1. 23. · originally proposed by Kudriavtsev in 1982 [12], but the only experimental realization of the technique was not performed

9

TABLE II. Spins, half-lives, AS1/2factors, isotope shifts, magnetic moments and change in mean-square charge radii of the

neutron-deficient francium isotopes 202−206Fr with reference to 221Fr for the 7s 2S1/2 → 8p 2P3/2 atomic transition. AllAS1/2

-factor and magnetic-moment values were deduced using the nuclear spins presented. The half-life values are taken from

Refs. [16, 18, 45–49].

A I t1/2 (s) AS1/2(GHz) µ (µN ) δνA,221 (GHz) δ〈r2〉 (fm2)

Lit. Exp. Lit. Exp. Lit. Exp. Exp. Lit.

202g (3+) 0.30(5) +12.80(5) +3.90(5) 32.68(10) -1.596(18)

202m (10−) 0.29(5) +2.30(3) +2.34(4) 32.57(13) -1.591(19)

203 (9/2−) 0.53(2) +8.18(3) +3.73(4) 31.32(10) -1.530(18)

204ga 3(+) 1.9(5) +12.99(3) +13.1499(43)b +3.95(5) +4.00(5)b,c 32.19(10) -1.571(18) -1.5542(4)b

204m1a (7+) 1.6(5) +6.44(3) +4.57(6) 32.32(10) -1.577(18)

204m2a (10−) 0.8(2) +2.31(3) +2.35(4) 30.99(10) -1.513(17)

205 9/2− 3.96(4) +8.40(3) +8.3550(11)b +3.83(5) +3.81(4)b,c 30.21(10) -1.475(17) -1.4745(4)b

206g 3(+) 15.9(3) +13.12(3) +13.0522(20)b +3.99(5) +3.97(5)b,c 30.04(12) -1.465(17) -1.4768(4)b

206m1d (7+) 15.9(3) +6.61(3) +4.69(6) 30.23(16) -1.475(18)

206m2d (10−) 0.7(1) +3.50(3) +3.55(5) 23.57(12) -1.153(14)

206m1e (7+) 15.9(3) +6.74(4) +4.79(6) 29.69(15) -1.449(17)

206m2e (10−) 0.7(1) +3.40(3) +3.45(5) 24.13(12) -1.180(14)

207 9/2− 14.8(1) +8.48(3) +8.484(1)f +3.87(5) +3.87(4)c,f 28.42(10) -1.386(16) -1.386(3)g

211 9/2− 186(1) +8.70(6) +8.7139(8)f +3.97(5) +3.98(5)c,f 24.04(10) -1.171(13) -1.1779(4)g

220 1+ 27.4(3) -6.50(4) -6.5494(9)f -0.66(1) -0.66(1)c,f 2.75(10) -0.134(5) -0.133(10)g

221 5/2− 294(12) +6.20(3) +6.2046(8)f +1.57(2) +1.57(2)c,f 0 0

a Calculated from the alpha-decay gated hyperfine structure scan of 204Fr. See text for details.b Literature value taken from Ref. [20].c Literature magnetic-moment values re-calculated in reference to µ(210Fr) [50]d Based on the isomeric identity of the hyperfine resonances of Option 1. See text for details.e Based on the isomeric identity of the hyperfine resonances of Option 2. See text for details.f Literature value taken from Ref. [7].g Literature value taken from Ref. [51].

The mass factor is the linear combination of two com-ponents: the normal mass shift, KNMS, and the specificmass shift, KSMS,

M422 = KNMS422 +KSMS

422 , (4)

and is dependent on the frequency of the transitionprobed. Subtraction of the normal mass shift of the422.7-nm transition (KNMS

422 = +389 GHz amu) from themass factor M422 allows for calculation of the specificmass shift, giving KSMS

422 = +360(330) GHz amu. Thespecific mass shift for the 718-nm line was determined byDzuba to be KSMS

718 = −314(113) GHz amu [51].

F. Hyperfine structure observables

Table II presents the hyperfine AS1/2factor, isotope

shift, change in mean-square charge radius and magneticmoment values extracted from the CRIS data for thefrancium isotopes 202−206Fr with reference to 221Fr. Ad-ditional data for 207,211,220Fr (used in the creation of theKing plot of Fig. 12) in included for completeness. Allvalues were deduced using the nuclear spins presented.

The hyperfine AS1/2factor is defined as

A =µIBeI · J

, (5)

with µI the magnetic dipole moment of the nucleus andBe the magnetic field of the electrons at the nucleus.For each isotope, it was calculated from the weightedmean of AS1/2

values for isotopes where more than onehyperfine structure scan is present. A minimum error of30 MHz was attributed to the AS1/2

factor values dueto the scatter of the measured AS1/2

for the reference

isotope 221Fr [41, 53].

The isotope shift, δνA,A′, between isotopes A and A′

is expressed as

δνA,A′

= MA′ −AAA′

+ Fδ〈r2〉A,A′. (6)

As with the AS1/2values, the isotope shifts were cal-

culated as the weighted mean of all isotope shifts for agiven nucleus. The error on the isotope shift was deter-mined to be 100 MHz due to the long-term drift of thecentroid frequency of 221Fr as the experiment progressed,and the scan-to-scan scatter in centroid frequency. Whenthe calculated weighted standard deviation of the isotope

Page 10: arXiv:1402.4266v1 [nucl-ex] 18 Feb 2014 · 2018. 1. 23. · originally proposed by Kudriavtsev in 1982 [12], but the only experimental realization of the technique was not performed

10

shift was higher than 100 MHz, this error is quoted in-stead. Combining the extracted F and M atomic factorsfrom the King-plot analysis with the measured isotopeshifts, evaluation of the change in mean-square chargeradii, δ〈r2〉A,A′ , between francium isotopes can be per-formed, see Eq. (6).

The magnetic moment of the isotopes under investiga-tion can be extracted from the known moment of anotherisotope of the element, using the ratio

µ = µrefIA

IrefAref. (7)

In this work, calculation of the magnetic moments wasevaluated in reference to the magnetic moment of 210Fr,measured by Gomez (µ = +4.38(5) µN , Iπ = 6+,AS1/2

= +7195.1(4) MHz [7, 50]). This represents themost accurate measurement of the magnetic moment ofa francium isotope to date, due to probing the 9s 2S1/2

hyperfine splitting which has reduced electron-correlationeffects than that of the ground state. The current eval-uated magnetic moments of the francium isotopes aremade in reference to the magnetic moment of 211Fr ofEkstrom [54].

The hyperfine anomaly for the francium isotopes isgenerally considered to be of the order of 1% and is in-cluded as a contribution to the error of the hyperfineAS1/2

factors and magnetic moments [55].Table II presents the experimental results alongside

comparison to literature of the hyperfine AS1/2factor,

change in mean-square charge radius and magnetic mo-ment values. The literature values for 204−206Fr havebeen taken from Ref. [20] and 207,211,220,221Fr fromRef. [7]. The magnetic-moment values from literaturehave been re-calculated in reference to µ(210Fr) [50],the most accurate measurement to date. The changein mean-square charge radii values for 207,211,220Fr havebeen taken from Ref. [51]. All experimental results arein broad agreement with those of literature.

IV. DISCUSSION

A. Charge radii of the neutron-deficient francium

Located between radon and radium, francium (Z = 87)has 5 valence protons occupying the π1h9/2 orbital, ac-cording to the shell model of spherical nuclei. Below theN = 126 shell closure, the neutron-deficient francium iso-topes were studied down to 202Fr (N = 115). The changein mean-square charge radii for the francium and lead iso-topes are presented in Fig. 13. The data of francium showthe charge radii of 207−213Fr re-evaluated by Dzuba [51]alongside the CRIS values which extends the data set to202Fr. The blue data points show the francium groundstates, while the (7+) isomeric states are in green andthe (10−) states in red. The error bars attributed tothe CRIS values are propagated from the experimental

114 116 118 120 122 124 126Neutron number

0.6

0.5

0.4

0.3

0.2

0.1

0.0

δ⟨ r2⟩ (

fm2

)

00.10.15β2 =0.2

Option 1Option 2

Option 1Option 2

Pb: Ground stateFr: Ground state

Fr: Spin (7+ ) stateFr: Spin (10− ) state

FIG. 13. Mean-square charge radii of the francium (cir-cle) isotopes [54] presented alongside the lead (diamond) iso-topes [56]. The dashed lines represent the prediction of thedroplet model for given iso-deformation [57]. The data werecalibrated by using β2(213Fr)= 0.062, evaluated from the en-ergy of the 2+

1 state in 212Rn [58]. Option 1 and 2 for the(7+) and (10−) states in 206Fr are based on the isomeric iden-tification given in Fig. 11.

error of the isotope shift and the systematic error as-sociated with the atomic factors F422 and M422. Thesystematic error is the most significant contribution tothe uncertainty associated with the mean-square chargeradii, and not that arising from the isotope shift. Thefrancium data is presented with the lead data of Ansel-ment [56] to illustrate the departure from the sphericalnucleus. The change in mean-square charge radii of thefrancium isotopes have been overlapped with the chargeradii of the lead isotopes, by using 213Fr (N = 126) and208Pb (N = 126) as reference points. The dashed iso-deformation lines represent the prediction of the dropletmodel for the francium isotopes [57]. The data were cali-brated using β2(213Fr)= 0.062, evaluated from the energyof the 2+1 state in 212Rn [58].

The doubly-magic 208Pb represents a model spheri-cal nucleus, with the shape of the nucleus remainingspherical with the removal of neutrons from the closedN = 126 shell. This trend is observed until N = 114,where a small deviation from the spherical droplet model(isodeformation line β2 = 0.0) is interpreted as enhancedcollectivity due to the influence of particle-hole excita-tions across the Z = 82 shell closure [59]. The change inmean-square charge radii for the francium isotopes showsagreement with the lead data as the ν3p3/2, ν2f5/2 andν3p1/2 orbitals are progressively depleted. The deviation

from sphericity at N = 116 with 203Fr marks the onsetof collective behaviour. The spectroscopic quadrupolemoments were not measured in this work, since they re-quire a laser linewidth of <100 MHz. Measurement ofthe quadrupole moment will provide information on thestatic deformation component of the change in mean-

Page 11: arXiv:1402.4266v1 [nucl-ex] 18 Feb 2014 · 2018. 1. 23. · originally proposed by Kudriavtsev in 1982 [12], but the only experimental realization of the technique was not performed

11

114 116 118 120 122 124 126Neutron number

0.6

0.5

0.4

0.3

0.2

0.1

0.0

δ⟨ r2⟩ (

fm2

)

00.10.15β2 =0.2

Option 1Option 2

Option 1Option 2

Rn: Ground stateFr: Ground state

Fr: Spin (7+ ) stateFr: Spin (10− ) state

FIG. 14. Mean-square charge radii of the francium (circle)isotopes [54] presented alongside the radon (diamond) iso-topes [60]. The dashed lines represent the prediction of thedroplet model for given iso-deformation [57]. The data werecalibrated using β2(213Fr)= 0.062, evaluated from the energyof the 2+

1 state in 212Rn [58]. Option 1 and 2 for the (7+) and(10−) states in 206Fr are based on the isomeric identificationgiven in Fig. 11.

square charge radii, allowing better understanding of thistransition region.

Recent laser spectroscopy measurements on theground-state properties of 204,205,206Fr suggest this devi-ation occurs earlier, at 206Fr (N = 119) [20]. In Ref. [20],a more pronounced odd-even staggering is observed in re-lation to the lead isotopes, where the mean-square chargeradius of 205Fr is larger than that of 206Fr. The CRIS ex-periment observed a smaller mean-square charge radiusof 205Fr in comparison to 206Fr, the deviation from thelead isotopes occurring at 203Fr instead. However, bothexperiments are in broad agreement within errors downto N = 117.

Figure 13 presents the two options of the mean-squarecharge radii of 206m1Fr and 206m2Fr (as defined by theirhyperfine peak identity in Fig. 11). Option 1 is favouredover option 2 due to the smaller mean-square charge radiiof 206m1Fr (compared to 206gFr) agreeing with the sys-tematics of the states in 204Fr. As seen in Fig. 13, 206gFr(N = 119) overlaps with the lead data within errors. Thelarge change in the mean-square charge radius of 206m2Frsuggests a highly deformed state for the (10−) isomer.

The mean-square charge radii of francium are overlaidwith the radon (Z = 86) charge-radii of Borchers (down toN = 116, with the exception of N = 117) [60] in Fig. 14.The mean-square charge-radii of radon have been cali-brated to the francium pair δ〈r2〉211,213 to account forthe uncertainty in F and M for the optical transitionprobed (the original isotope shifts are presented graph-ically). Despite this, the agreement between the mean-square charge radii of the francium and radon data isclear. The addition of a single π1h9/2 proton outside the

TABLE III. Extracted β2 values. (Exp.) The dropletmodel [57] was used to extract the rms values for β2 fromthe change in mean-square charge radii. The charge-radii val-ues were calibrated using β2(213Fr)= 0.062, as before. (Lit.)β2 values were extracted from the quadrupole moments ofRef. [20]. See text for details.

A I 〈β22〉1/2 β2

Exp. Lit.

202g (3+) 0.11

202m (10−) 0.11

203 (9/2−) 0.11

204g 3(+) 0.06a -0.0140(14)

204m1 (7+) 0.06a

204m2 (10−) 0.09a

205 9/2− 0.08 -0.0204(2)

206g 3(+) 0.05 -0.0269(8)

206m1 (7+) 0.04b

206m2 (10−) 0.17b

202m1 (7+) 0.07c

202m2 (10−) 0.17c

a Calculated from the alpha-decay gated hyperfine structure scanof 204Fr. See text for details.

b Based on the isomeric identity of the hyperfine resonances ofOption 1. See text for details.

c Based on the isomeric identity of the hyperfine resonances ofOption 2. See text for details.

radon even-Z core does not affect the charge-radii trend,suggesting the valence proton acts as a spectator particle.

Table III presents a comparison of β2 values with lit-erature. The droplet model [57] was used to extract therms values for β2 (column 3) from the change in mean-square charge radii (calibrated using β2(213Fr)= 0.062, asbefore). Column 4 presents β2 values extracted from thequadrupole moments of Ref. [20]. The larger β2 valuesextracted from the mean-square charge radii, comparedto those extracted from the quadrupole moments, sug-gest that the enhanced collectivity observed in Figs. 13and 14 is due to a large dynamic component of the nu-clear deformation.

B. Interpretation of the nuclear g-factors

Figures 15 and 17 show the experimental g-factors forodd-A and even-A francium isotopes, respectively. Theseplots present the CRIS data alongside the data from Ek-strom [54]. The Ekstrom data has been re-evaluated withrespect to the µ(210Fr) measurement of Gomez [50].

In Fig. 15, the blue line represents the empirical g-factor (gemp) of the odd-A isotopes for the single-particleoccupation of the valence proton in the π1h9/2 orbital.gemp(π1h9/2) was determined from the magnetic mo-

ment of the single-particle state in 209Bi [61]. Simi-larly, gemp(π3s1/2) was estimated from the magnetic mo-

Page 12: arXiv:1402.4266v1 [nucl-ex] 18 Feb 2014 · 2018. 1. 23. · originally proposed by Kudriavtsev in 1982 [12], but the only experimental realization of the technique was not performed

12

114 116 118 120 122 124 126Neutron number

0.5

1.0

1.5

2.0

2.5

3.0

3.5

4.0

4.5g-

fact

or

gemp(π1h9/2)

gemp(π3s1/2)

Fr: Ground stateFr: Ekstrom (1986)Tl: Stone (2011)

FIG. 15. g-factors for francium (blue) [50, 54] and thallium(red) isotopes with odd-A [63]. The g-factors for the π3s1/2and π1h9/2 proton orbitals have been calculated empirically.See text for details.

ment of the single-hole ground-state in 207Tl [62]. FromN = 126 to 116, every isotope has a g-factor consistentwith the proton occupying the π1h9/2 orbital. This in-

dicates that the 9/2− state remains the ground state,and the (π3s−11/2)1/2+ proton intruder state has not yet

inverted. This lowering in energy of the π3s1/2 state tobecome the ground state would be apparent in the sud-den increase in g-factor of the ground state, as illustratedby the black gemp(π3s1/2) line.

Figure 15 highlights the robustness of the Z = 82 andN = 126 shell closure with a shell-model description validover a range of isotopes. A close-up of gemp(1π1h9/2) inFig. 16 illustrates that the g-factor is sensitive to bulknuclear effects. The departure from the gemp(1π1h9/2)line shows the sensitivity of the g-factor to second-ordercore polarization in the odd-A thallium, bismuth andfrancium isotopes. The systematic decrease in g-factorof francium is attributed to second-order core polariza-tion associated with the presence of five valence particles,compared to one-particle (hole) in the bismuth (thallium)isotopes, enough to significantly weaken the shell closure.The linear trend observed in bismuth, thallium and fran-cium (until N = 118) is suggested to be related to theopening of the neutron shell, yet allowing for more neu-tron and proton-neutron correlations.

Further measurements towards the limit ./of stabil-ity are needed to better understand the prediction ofthe inversion of the π3s1/2 intruder orbital with the

π1h9/2 ground state. A re-measurement of 203Fr could

determine the presence of the spin 1/2+ isomer (t1/2 =43(4) ms [19]), which was not observed during this ex-periment.

The g-factors for the odd-odd francium isotopes arepresented in Fig. 17. With the coupling of the single va-lence proton in the π1h9/2 orbital with a valence neutron,

106 108 110 112 114 116 118 120 122 124 126Neutron number

0.82

0.84

0.86

0.88

0.90

0.92

g-fa

ctor

gemp(π1h9/2)

Fr: This work (2013)Fr: Ekstrom (1986)Bi: Stone (2011)Tl: Barzakh (2012)

FIG. 16. Close up of the g-factors for francium (blue) [50, 54],bismuth (green) [63] and thallium (red) isotopes [64] withodd-A. The g-factor for the π1h9/2 proton orbital has beencalculated empirically. See text for details.

a large shell model space is available. The empiricallycalculated g-factors for the coupling of the π1h9/2 pro-ton with the valence neutrons are denoted by the coloredlines. These g-factors were calculated from the additivityrelation

g =1

2

[gp + gn + (gp − gn)

jp(jp + 1)− jn(jn + 1)

I(I + 1)

], (8)

as outlined by Neyens [65]. The empirical g-factors ofthe odd valence neutrons were calculated from the mag-netic moments of neighbouring nuclei: 201Po for the bluegemp(π1h9/2 ⊗ ν3p3/2) and red gemp(π1h9/2 ⊗ ν1i13/2)

line [66]; 213Ra for the black gemp(π1h9/2 ⊗ ν3p1/2) line;

and 211Ra for the green gemp(π1h9/2 ⊗ ν2f5/2) line [67].The empirical g-factors for the valence proton in theπ1h9/2 orbital were calculated from the magnetic mo-

ment of the closest odd-A francium isotope (203Fr and213Fr respectively) from the CRIS data.

The ground state of 202,204,206Fr display similar g-factors, with the valence proton and neutron coupling togive a spin 3(+) state. The tentative configuration in lit-erature of (π1h9/2 ⊗ ν2f5/2)3+ for 202gFr is based on the

configuration of the (3+) state in 194Bi (from favouredFr-At-Bi alpha-decay chain systematics) [44]. Similarly,the assignment of the same configuration for 204gFr and206gFr is based on the alpha-decay systematics of neigh-bouring nuclei 196,198Bi. However, the initial assignmentof 194gBi was declared to be either (π1h9/2 ⊗ ν2f5/2)3+or (π1h9/2 ⊗ ν3p3/2)3+ [68]. From the g-factors of the

ground states of 202,204,206Fr, it is clear that the configu-ration of these states is indeed (π1h9/2 ⊗ ν3p3/2)3+ .

Figure 17 also presents the g-factors of 206m1Fr and206m2Fr for option 1 and 2 (see Fig. 11). The first iso-meric states of 204,206Fr (7+) have a valence neutron thatoccupies the ν2f5/2 state. This coupling of the proton-

Page 13: arXiv:1402.4266v1 [nucl-ex] 18 Feb 2014 · 2018. 1. 23. · originally proposed by Kudriavtsev in 1982 [12], but the only experimental realization of the technique was not performed

13

114 116 118 120 122 124 126Neutron number

0.5

1.0

1.5

2.0g-

fact

or

gemp(π1h9/2⊗ν3p1/2)

gemp(π1h9/2⊗ν3p3/2)

gemp(π1h9/2⊗ν2f5/2)

gemp(π1h9/2⊗ν1i13/2)

Option 1

Option 2

Option 1Option 2

Ekstrom (1986)Ground state

Spin (7+ ) stateSpin (10− ) state

FIG. 17. g-factors for francium isotopes with even-A [54]:ground state (blue), spin (7+) state (green) and spin (10−)state (red). The g-factor for the coupling of the proton andneutron orbitals have been calculated empirically. See textfor details.

particle neutron-hole results in a (π1h9/2⊗ν2f5/2)7+ con-

figuration [69]. For 202mFr, 204m2Fr and 206m2Fr, theparticle proton-neutron hole coupling result in a tenta-tive (π1h9/2 ⊗ ν1i13/2)10− configuration assignment foreach isomer [70]. However, while the agreement of theg-factors of the spin (10−) state in 202,204Fr point toa ν1i13/2 occupancy, the observed value for 206m2Fr is

in disagreement with the g-factor of such a (10−) state.The charge radius of 206m2Fr indicates a highly deformedconfiguration, where the single-particle description of thenucleus is no longer valid. This is consistent with the g-factor of this state: it is no longer obeying a simple shell-model description. This leads to the conclusion, thatwhile a (π1h9/2⊗ν1i13/2)10− configuration for 206m2Fr issuggested, the charge radii and magnetic moment pointto a drastic change in the structure of this isomeric state.

For completeness, the configurations of the odd-oddfrancium isotopes 208,210,212Fr are presented. The cou-pling of the valence proton and neutron in the π1h9/2

and ν2f5/2 orbital in 208Fr and 210Fr leads to a (π1h9/2⊗ν2f5/2)7+ and (π1h9/2 ⊗ ν2f5/2)6+ configuration respec-tively [71]. With the ν2f5/2 neutron orbital fully occu-

pied, the valence neutron in 212Fr occupies the ν3p1/2

orbital, resulting in a (π1h9/2 ⊗ ν3p1/2)5+ configura-tion [72].

The agreement of the experimental and empirical g-factors, as shown in Figs. 15-17, illustrates the suitabilityof the single-particle description of the neutron-deficientfrancium isotopes, with the exception of the (10−) statein 206m2Fr. A model-independent spin and spectroscopicquadrupole moment determination is needed to clarifythe nature of this isomeric state. The neutron-deficientfrancium isotopes display a single-particle nature wherethe additivity relation is still reliable.

V. CONCLUSION AND OUTLOOK

The hyperfine structures and isotope shifts of theneutron-deficient francium isotopes 202−206Fr with ref-erence to 221Fr were measured with collinear resonanceionisation spectroscopy, and the change in mean-squarecharge radii and magnetic moments extracted. The se-lectivity of the alpha-decay patterns allowed the unam-biguous identification of the hyperfine components of thelow-lying isomers of 202,204Fr for the first time.

The resonant atomic transition of 7s 2S1/2 → 8p 2P3/2

was probed, and the hyperfine AS1/2factor mea-

sured. A King plot analysis of the 422.7-nm tran-sition in francium allowed the atomic factors to becalibrated. The field and mass factors were deter-mined to be F422 = −20.670(210) GHz/fm2 and M422

= +750(330) GHz amu, respectively.The novel technique of decay-assisted laser spec-

troscopy in a collinear geometry was performed on theisotopes 202,204Fr. The decay spectroscopy station wasutilized to identify the peaks in the hyperfine spectraof 202,204Fr. Alpha-tagging the hyperfine structure scanof 204Fr allowed the accurate determination of the nu-clear observables of the three low-lying isomeric statesand the determination of the branching ratios in the de-cay of 204m2Fr.

Analysis of the change in mean-square charge radiisuggests an onset of collectivity that occurs at 203Fr(N = 116). However, measurement of the spectroscopicquadrupole moment is required to determine the natureof the deformation (static or dynamic). The magneticmoments suggest that the single-particle description ofthe neutron-deficient francium isotopes still holds, withthe exception of the (10−) isomeric state of 206m2Fr.Based on the systematics of the region, the tentative as-signment of the hyperfine structure peaks in 206Fr resultin magnetic moments and mean-square charge radii thatsuggest a highly deformed state. Laser assisted nucleardecay spectroscopy of 206Fr would unambiguously deter-mine their identity.

The occupation of the valence proton in the π1h9/2 or-bital has been suggested for all measured isotopes downto 202Fr, indicating the (π1s−11/2)1/2+ intruder state does

not yet invert with the π1h9/2 orbital as the ground state.Further measurements of the very neutron-deficient fran-cium isotopes towards 199Fr are required to fully deter-mine the nature of the proton-intruder state. A laserlinewidth of 1.5 GHz was enough to resolve the lower-state (7s 2S1/2) splitting of the hyperfine structure andmeasure the AS1/2 factor. In the future, the inclusion of anarrow-band laser system for the resonant-excitation stepwill enable the resolution of the upper-state (8p 2P3/2)splitting, providing the hyperfine BP3/2

factor. This willallow extraction of the spectroscopic quadrupole momentand determination of the nature of the deformation.

Successful measurement of 202Fr was performed dur-ing this experiment, with a yield of 100 atoms per sec-ond. By pushing the limits of laser spectroscopy, further

Page 14: arXiv:1402.4266v1 [nucl-ex] 18 Feb 2014 · 2018. 1. 23. · originally proposed by Kudriavtsev in 1982 [12], but the only experimental realization of the technique was not performed

14

measurements of 201Fr (with a yield of 1 atom per sec-ond) and 200Fr (less than 1 atom per second) are thoughtto be possible. The ground state (9/2−) of 201Fr has ahalf life of 53 ms and its isomer (1/2+) a half life of19 ms. By increasing the sensitivity of the CRIS tech-nique, the presence of the 1/2+ isomers in 201,203Fr can beconfirmed. A positive identification will lead to nuclear-structure measurements that will determine (along withthe verification of nuclear spin) the magnetic momentswhich are sensitive to the single-particle structure andthus to the (π3s1/2)1/2+ proton intruder nature of thesestates. With sufficient resolution (<100 MHz), the spec-troscopic quadrupole moment of these neutron-deficientstates (with I ≥ 1) will be directly measurable and thetime-averaged static deformation can be determined.

The successful measurements performed by the CRISexperiment demonstrates the high sensitivity of thecollinear resonance ionization technique. The decay spec-troscopy station provides the ability to identify overlap-ping hyperfine structure and eventually perform laser as-

sisted nuclear decay spectroscopy measurements on pureground and isomeric-state beams [32, 73].

ACKNOWLEDGEMENTS

The authors extend their thanks to the ISOLDE teamfor providing the beam, the GSI target lab for producingthe carbon foils, and IKS-KU Leuven and The Universityof Manchester machine shops for their work. This workwas supported by the IAP project P7/23 of the OSTCBelgium (BRIX network) and by the FWO-Vlaanderen(Belgium). The Manchester group was supported by theSTFC consolidated grant ST/F012071/1 and continua-tion grant ST/J000159/1. K.T. Flanagan was supportedby STFC Advanced Fellowship Scheme grant numberST/F012071/1. The authors would also like to thank EdSchneiderman for continued support through donationsto the Physics Department at NYU.

[1] B. Cheal and K. T. Flanagan, J. Phys. G 37, 113101(2010).

[2] K. Blaum, J. Dilling, and W. Nortershauser, PhysicaScripta 2013, 014017 (2013).

[3] T. J. Procter and K. T. Flanagan, Hyperfine Interact.216, 89 (2013).

[4] S. Liberman et al., C. R. Acad. Sci. Paris Ser. B 286,353 (1978).

[5] S. Liberman et al., Phys. Rev. A 22, 2732 (1980).[6] H. Yagoda, Phys. Rev. 40, 1017 (1932).[7] A. Coc et al., Phys. Lett. B 163, 66 (1985).[8] F. Touchard et al., Atomic Masses and Fundamental

Constants 7 , 353 (1984).[9] J. Bauche et al., J. Phys. B: At. Mol. Opt. Phys. 19,

L593 (1986).[10] H. T. Duong et al., Europhys. Lett. 3, 175 (1987).[11] S. V. Andreev, V. I. Mishin, and V. S. Letokhov, Phys.

Rev. Lett. 59, 1274 (1987).[12] Y. Kudriavtsev and V. Letokhov, Applied Physics B 29,

219 (1982).[13] C. Schulz et al., J. Phys. B: At. Mol. Opt. Phys. 24, 4831

(1991).[14] C. N. Davids et al., Phys. Rev. Lett. 76, 592 (1996).[15] H. Kettunen et al., Eur. Phys. J. A 16, 457 (2003).[16] J. Uusitalo et al., Phys. Rev. C 71, 024306 (2005).[17] U. Jakobsson et al., Phys. Rev. C 85, 014309 (2012).[18] M. Huyse, P. Decrock, P. Dendooven, G. Reusen,

P. Van Duppen, and J. Wauters, Phys. Rev. C 46, 1209(1992).

[19] U. Jakobsson et al., Phys. Rev. C 87, 054320 (2013).[20] A. Voss, M. R. Pearson, J. Billowes, F. Buchinger,

B. Cheal, J. E. Crawford, A. A. Kwiatkowski, C. D. P.Levy, and O. Shelbaya, Phys. Rev. Lett. 111, 122501(2013).

[21] V. N. Fedosseev, Y. Kudryavtsev, and V. I. Mishin,Phys. Scripta 85, 058104 (2012).

[22] L. Weissman et al., Phys. Rev. C 65, 024315 (2002).[23] I. Stefanescu et al., Phys. Rev. Lett. 98, 122701 (2007).

[24] J. Van Roosbroeck et al., Phys. Rev. Lett. 92, 112501(2004).

[25] E. Kugler, Hyperfine Interact. 129, 23 (2000).[26] A. Jokinen, M. Lindroos, E. Molin, and M. Petersson,

Nucl. Instrum. Methods Phys. Res. B 204, 86 (2003).[27] E. Mane et al., Eur. Phys. J. A 42, 503 (2009).[28] W. H. Haynes, Handbook of Chemistry and Physics, 94th

ed. (CRC, 2013).[29] V. N. Fedosseev et al., Rev. Sci. Instrum. 83, 02A903

(2012).[30] S. Rothe, V. N. Fedosseev, T. Kron, B. A. Marsh, R. E.

Rossel, and K. D. A. Wendt, Nucl. Instrum. MethodsPhys. Res. B 317, Part B, 561 (2013).

[31] R. E. Rossel, V. N. Fedosseev, B. A. Marsh, D. Richter,S. Rothe, and K. D. A. Wendt, Nucl. Instrum. MethodsPhys. Res. B 317, Part B, 557 (2013).

[32] K. M. Lynch, T. E. Cocolios, and M. M. Rajabali, Hy-perfine Interact. 216, 95 (2013).

[33] M. M. Rajabali et al., Nucl. Instrum. Methods Phys. Res.A 707, 35 (2013).

[34] P. Dendooven, Ph.D. thesis, IKS, KU Leuven (1992).[35] A. N. Andreyev et al., Phys. Rev. Lett. 105, 252502

(2010).[36] J. Elseviers et al., Phys. Rev. C 88, 044321 (2013).[37] B. Lommel, W. Hartmann, B. Kindler, J. Klemm, and

J. Steiner, Nucl. Instrum. Methods Phys. Res. A 480,199 (2002).

[38] W. Hennig, H. Tan, M. Walby, P. Grudberg, A. Fallu-Labruyere, W. K. Warburton, C. Vaman, K. Starosta,and D. Miller, Nucl. Instrum. Methods Phys. Res. B 261,1000 (2007).

[39] K. T. Flanagan et al., Phys. Rev. Lett. 111, 212501(2013).

[40] I. Budincevic et al., (unpublished).[41] T. E. Cocolios et al., Nucl. Instrum. Methods Phys. Res.

B 317, Part B, 565 (2013).[42] K. M. Lynch, Ph.D. thesis, The University of Manchester

(2013).

Page 15: arXiv:1402.4266v1 [nucl-ex] 18 Feb 2014 · 2018. 1. 23. · originally proposed by Kudriavtsev in 1982 [12], but the only experimental realization of the technique was not performed

15

[43] H. De Witte et al., Eur. Phys. J. A 23, 243 (2005).[44] S. Zhu and F. G. Kondev, Nuclear Data Sheets 109, 699

(2008).[45] W. Lourens, Ph.D. thesis, Technische Hogeschool Delft

(1967).[46] B. G. Ritchie, K. S. Toth, H. K. Carter, R. L. Mlekodaj,

and E. H. Spejewski, Phys. Rev. C 23, 2342 (1981).[47] F. Kondev and S. Lalkovski, Nuclear Data Sheets 112,

707 (2011).[48] B. Singh, D. Abriola, C. Baglin, V. Demetriou, T. John-

son, E. McCutchan, G. Mukherjee, S. Singh, A. Sonzogni,and J. Tuli, Nuclear Data Sheets 114, 661 (2013).

[49] E. Browne and J. Tuli, Nuclear Data Sheets 112, 1115(2011).

[50] E. Gomez, S. Aubin, L. A. Orozco, G. D. Sprouse,E. Iskrenova-Tchoukova, and M. S. Safronova, Phys.Rev. Lett. 100, 172502 (2008).

[51] V. A. Dzuba, W. R. Johnson, and M. S. Safronova, Phys.Rev. A 72, 022503 (2005).

[52] W. H. King, J. Opt. Soc. Am. 53, 638 (1963).[53] K. M. Lynch et al., EPJ Web of Conferences 63, 01007

(2013).[54] C. Ekstrom, L. Robertsson, and A. Rosen, Phys. Scripta

34, 624 (1986).[55] H. H. Stroke, R. J. Blin-Stoyle, and V. Jaccarino, Phys.

Rev. 123, 1326 (1961).[56] M. Anselment, W. Faubel, S. Goring, A. Hanser,

G. Meisel, H. Rebel, and G. Schatz, Nucl. Phys. A 451,471 (1986).

[57] W. D. Myers and K.-H. Schmidt, Nucl. Phys. A 410, 61(1983).

[58] S. Raman, C. N. Jr., and P. Tikkanen, Atomic Data andNuclear Data Tables 78, 1 (2001).

[59] H. De Witte et al., Phys. Rev. Lett. 98, 112502 (2007).[60] W. Borchers, R. Neugart, E. W. Otten, H. T. Duong,

G. Ulm, and K. Wendt, Hyperfine Interact. 34, 25(1987).

[61] T. Bastug, B. Fricke, M. Finkbeiner, and W. R. Johnson,Z. Phys. D 37, 281 (1996).

[62] R. Neugart, H. H. Stroke, S. A. Ahmad, H. T. Duong,H. L. Ravn, and K. Wendt, Phys. Rev. Lett. 55, 1559(1985).

[63] N. Stone, Table of Nuclear Magnetic Dipole and ElectricQuadrupole Moments (Nuclear Data Services, Interna-tional Atomic Energy Agency, Vienna, Austria, 2011).

[64] A. E. Barzakh, L. K. Batist, D. V. Fedorov, V. S. Ivanov,K. A. Mezilev, P. L. Molkanov, F. V. Moroz, S. Y. Orlov,V. N. Panteleev, and Y. M. Volkov, Phys. Rev. C 86,014311 (2012).

[65] G. Neyens, Rep. Prog. Phys. 66, 633 (2003).[66] J. Wouters, N. Severijns, J. Vanhaverbeke, and

L. Vanneste, J. Phys. G: Nucl. Part. Phys. 17, 1673(1991).

[67] S. A. Ahmad, W. Klempt, R. Neugart, E. W. Otten,K. Wendt, and C. Ekstrom, Physics Letters B 133, 47(1983).

[68] P. V. Duppen, P. Decrock, P. Dendooven, M. Huyse,G. Reusen, and J. Wauters, Nucl. Phys. A 529, 268(1991).

[69] F. G. Kondev, Nuclear Data Sheets 109, 1527 (2008).[70] C. J. Chiara and F. G. Kondev, Nuclear Data Sheets

111, 141 (2010).[71] M. J. Martin, Nuclear Data Sheets 108, 1583 (2007).[72] E. Browne, Nuclear Data Sheets 104, 427 (2005).[73] K. M. Lynch et al., J. Phys.: Conf. Ser. 381, 012128

(2012).