03.07.2013 Views

TRIAC Progress Report - KEK

TRIAC Progress Report - KEK

TRIAC Progress Report - KEK

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

<strong>TRIAC</strong> <strong>Progress</strong> <strong>Report</strong><br />

<strong>TRIAC</strong> collaboration<br />

(Edited by S. C. Jeong)<br />

<strong>KEK</strong> <strong>Progress</strong> <strong>Report</strong> 2011-1<br />

June 2011<br />

A/H<br />

High Energy Accelerator Research Organization


© High Energy Accelerator Research Organization (<strong>KEK</strong>), 2011<br />

<strong>KEK</strong> <strong>Report</strong>s are available from:<br />

High Energy Accelerator Research Organization (<strong>KEK</strong>)<br />

1-1 Oho, Tsukuba-shi<br />

Ibaraki-ken, 305-0801<br />

JAPAN<br />

Phone: +81-29-864-5137<br />

Fax: +81-29-864-4604<br />

E-mail: irdpub@mail.kek.jp<br />

Internet: http://www.kek.jp


i<br />

<strong>KEK</strong> <strong>Progress</strong> <strong>Report</strong> 2011-1<br />

June 2011<br />

A/H<br />

<strong>TRIAC</strong> <strong>Progress</strong> <strong>Report</strong><br />

<strong>TRIAC</strong> collaboration<br />

(Edited by S.C. Jeong)<br />

High Energy Accelerator Research Organization


<strong>TRIAC</strong> collaboration<br />

S. Arai 1 , Y. Arakaki 1 , Y. Fuchi 1 , Y. Hirayama 1 , N. Imai 1 , H. Ishiyama 1 ,<br />

S.C. Jeong 1 , I. Katayama 1 , H. Kawakami 1 , H. Miyatake 1 , K. Niki 1 , T. Nomura 1 ,<br />

M. Okada 1 , M. Oyaizu 1 , M.H. Tanaka 1 , E. Tojyo 1 , M. Tomizawa 1 , N. Yoshikawa 1 ,<br />

Y.X. Watanabe 1 , S. Abe 2 , T. Asozu 2 , S. Hanashima 2 , K. Horie 2 , S. Ichikawa 2 ,<br />

H. Ikezoe 2 , T. Ishii 2 , N. Ishizaki 2 , A. Iwamoto 2 , H. Kabumoto 2 , S. Kanda 2 ,<br />

K. Kutsukake 2 , M. Matsuda 2 , S. Mitsuoka 2 , M. Nakamura 2 , T. Nakanoya 2 , I. Ohuchi 2 ,<br />

A. Osa 2 , Y. Otokawa 2 , M. Sataka 2 , T.K. Sato 2 , S. Takeuchi 2 , H. Tayama 2 ,<br />

Y. Tsukihashi 2 , and T. Yoshida 2<br />

K. Matsuta 3* , J. Murata 4* , T. Fukuda 5* , W. Sato 6* , S. Takai 7* , T. Teranishi 8* ,<br />

H. Sugai 2* , H. Makii 2* , M. Shibata 9* , and D. Nagae 2*<br />

1 High Energy Accelerator Research Organization (<strong>KEK</strong>), Ibaraki 305-0801, Japan<br />

2 Japan Atomic Energy Agency (JAEA), Ibaraki 319-1195, Japan<br />

3 Research Center for Nuclear Physics, Osaka University, Osaka 565-0871, Japan<br />

4 Graduate School of Science, Rikkyo University, Tokyo 171-8501, Japan<br />

5<br />

Graduate School of Engineering, Osaka Electro-communication University,<br />

Osaka 572-8530, Japan<br />

6<br />

Graduate School of Natural science and Technology, Kanazawa University,<br />

Ishikawa 920-1192, Japan<br />

7<br />

Graduate School of Engineering, Tottori University, Tottori 680-8550, Japan<br />

8 Faculty of Sciences, Kyushu University, Fukuoka 812-8581, Japan<br />

9 Graduate School of Engineering, Nagoya University, Nagoya 464-8603, Japan<br />

* Spokespersons of the experimental proposals, not involved in the development of the<br />

<strong>TRIAC</strong> facilities. They are included as representatives of respective experiments since<br />

all experiments were performed in the framework of the <strong>TRIAC</strong> collaboration.<br />

ii


PREFACE<br />

The Tokai Radioactive Ion Accelerator Complex (<strong>TRIAC</strong>) is a radioactive ion<br />

beam (RIB) facility of Institute of Particle and Nuclear Studies (IPNS), High Energy<br />

Accelerator Research Organization (<strong>KEK</strong>), being installed at the Tokai Research and<br />

Development Center, Japan Atomic Energy Agency (JAEA). The <strong>TRIAC</strong> is, in Japan,<br />

the only low-energy accelerator complex for re-accelerating short-lived radioactive<br />

isotope beams, operating as User Facility since the November of 2005.<br />

The main components of <strong>TRIAC</strong> were primarily constructed at Institute for<br />

Nuclear Study (INS), University of Tokyo, from 1992 to 1995 as a prototype for the<br />

Exotic nuclear (E-) Arena of Japanese Hadron Facility (former name of the J-PARC).<br />

The prototype facility was closed in 1999 and started to be re-installed at the JAEA<br />

tandem accelerator facility from 2001 under the collaboration between <strong>KEK</strong> and JAEA,<br />

which is now called <strong>TRIAC</strong>. The project for <strong>TRIAC</strong> has started with following<br />

considerations: (1) The RIBs with energies above Coulomb barrier can be available at<br />

low construction cost, by utilizing two existing linear accelerators (linacs), namely the<br />

<strong>KEK</strong> linacs (SCRFQ / IH) and the JAEA superconducting linac (SC-linac). (2) The<br />

uranium target in a form of UCx is available for RIB production at the tandem facility.<br />

Although the connection of the two linacs has not yet been realized, the <strong>TRIAC</strong><br />

can provide beams of Uranium Fission Fragments with the maximum energy of 1.1<br />

MeV/nucleon produced by protons of 30 MeV and 1 µA (30 W in beam power,<br />

actually deposited in the production target) from the JAEA Tandem Accelerator.<br />

In the present <strong>TRIAC</strong>, considering current trends of world-wide RIB facilities, we<br />

recognize critical limitations in views of the re-accelerated energy and intensity of<br />

available RIBs. The possibility for upgrading the present <strong>TRIAC</strong> has been seriously<br />

discussed so far. As a result of such discussions, the <strong>TRIAC</strong> was determined to be<br />

closed at the end of 2010. Hereby, we summarize the scientific activities mainly<br />

performed at the <strong>TRIAC</strong> during the period of the operation (2006-2010). This<br />

documentation includes the activities for the development of the production and<br />

acceleration of RIBs, and experimental results obtained by using the short-lived RIBs<br />

of the <strong>TRIAC</strong>.<br />

We would like to express our great thanks to all scientists participating in all the<br />

scientific activities (<strong>TRIAC</strong> developments and experiments) performed at the <strong>TRIAC</strong>.<br />

iii


Contents<br />

1. Introduction -1<br />

2. Developments of the beam generation and acceleration at the <strong>TRIAC</strong><br />

2-1. Overview of the <strong>TRIAC</strong> -2<br />

2-2. Ion source and target development -3<br />

2-3. Charge Breeder -10<br />

2-4. Accelerator -20<br />

2-5. Beam transport line -39<br />

2-6. Control system -43<br />

2-7. Detector development -44<br />

3. Experiments at the <strong>TRIAC</strong><br />

3-1. Nuclear astrophysics -52<br />

3-2. Nuclear and fundamental physics -64<br />

3-3. Materials Science -79<br />

3-4. Experiments with the ISOL beam -90<br />

iv


1. Introduction<br />

Following the successful re-acceleration of short-lived radioactive ion beams<br />

(RIBs) from the isotope separator on-line (ISOL) in the early 1990s, many ISOL-based<br />

RIB facilities have been constructed over the world, and now are operating for various<br />

research fields such as in nuclear physics, nuclear astrophysics, and materials science.<br />

The low-energy RIBs, energy variable and with a high purity and good energy<br />

resolution, are very effective not only for the detailed spectroscopic studies on nuclear<br />

structure using nuclear reactions but also for studies on the structural and dynamical<br />

properties of the bulk of materials.<br />

At the Tandem accelerator facility of JAEA-Tokai, an ISOL-based radioactive ion<br />

beam facility, <strong>TRIAC</strong>-Tokai Radioactive Ion Accelerator Complex - is operational since<br />

the November of 2005. In the facility, short-lived radioactive ions produced by proton<br />

or heavy ion induced nuclear reactions, after being mass-separated by the JAEA-ISOL,<br />

can be accelerated up to the energy necessary for experiments. The energy is variable in<br />

the range from 0.14 to 1.09 MeV/nucleon, which is the energy region available at the<br />

first stage of the complete version of the <strong>TRIAC</strong>.<br />

The present energy range of RIBs available at the <strong>TRIAC</strong> is suitable for direct<br />

measurements for light ion induced nuclear reaction rates involved in the<br />

nucleosynthesis at the explosive stage of the stellar evolution of massive stars as well as<br />

in the early universe. The low energy RIBs can be also incorporated by implantation<br />

into materials of interest whose properties can be probed with an unprecedented<br />

sensitivity by using various nuclear spectroscopy techniques.<br />

Since the first public operation, 50 days per year, in maximum, has been allowed<br />

for experiments, including 10-day operation for the development of the RIB generation<br />

and the acceleration at the <strong>TRIAC</strong>. The experimental proposals were carefully reviewed,<br />

once a year, by the program advisory committee (PAC).<br />

In the following, we first introduce the scientific activities for the developments of<br />

RIB at the <strong>TRIAC</strong>. The summary of the experiments performed so far will follow.<br />

2. Developments of the beam generation and acceleration at the <strong>TRIAC</strong><br />

The developments have been concentrated for the production and acceleration of<br />

heavy (medium-mass) neutron-rich radioactive ion beams with high efficiency. The<br />

followings are included in this section; (1) the development of target / ion source of the<br />

JAEA-ISOL for effective production and ionization of uranium fission fragments, (2)<br />

the development for charge breeding the singly charged ions from the ISOL, (3) the<br />

development of post-accelerators for CW operation, upgrading acceleration energy, and<br />

1


eam bunching, (4) the development of beam transport line and remote control system,<br />

and (5) the development of detectors for beam diagnostics and experiments .<br />

2-1. Overview of the <strong>TRIAC</strong><br />

The layout of the facility is given in Fig. 2-1. The <strong>TRIAC</strong> is based on an isotope<br />

separator on-line (ISOL) and the radioactive nuclei are produced via proton-induced<br />

fission of 238 U or heavy-ion reactions with the primary beams from the JAEA tandem<br />

accelerator, which is not shown in Fig. 2-1. The produced radioactive nuclei are singly<br />

charged and mass-separated by the JAEA-ISOL. They are fed to the 18-GHz electron<br />

cyclotron resonance ion-source for charge-breeding (<strong>KEK</strong>CB), where the singly<br />

charged ions are converted to multi-charged ions. The charge-bred radioactive ions,<br />

usually with a mass (A) to charge-state (q) ratio of around 7 (A/q ~ 7), are extracted<br />

again and fed to the post accelerator for re-acceleration. The post accelerator consisting<br />

of two linear accelerators (linacs), a split-coaxial radio-frequency quadrupole (SCRFQ)<br />

linac and an interdigital-H (IH) linac, can accelerate the RIB to the energy necessary for<br />

experiments. The acceleration of the RIBs charge-bred by an ECR ion source was the<br />

first time over the world and the overall efficiency of transmission from ISOL to the<br />

experimental hall is about 2%. The basic parameters of the <strong>TRIAC</strong> are summarized in<br />

Table 2-1.<br />

Fig. 2-1. Layout of the <strong>TRIAC</strong><br />

Several experiments have been successfully performed, especially by using the 8 Li<br />

beam with the intensity of several 10 5 particles per second (pps). The <strong>TRIAC</strong> now<br />

provides beams of uranium fission fragments with the maximum energy of 1.1<br />

MeV/nucleon, which are routinely produced by protons of 34 MeV and around 1 µA<br />

(with a beam power of about 30 W) from the JAEA Tandem Accelerator, though the<br />

2


maximum available beam power for the production is 90 W.<br />

The progress report on the RIB generation and acceleration at the <strong>TRIAC</strong> made so<br />

far from the early stage of the construction is summarized in the following.<br />

Table 2-1. Basic parameters of the <strong>TRIAC</strong><br />

Primary Beam (JAEA) Ion (energy / intensity) P (36 MeV / 3 µA)*<br />

Production target<br />

(JAEA)<br />

ISOL (JAEA) Ion source<br />

Mass resolution<br />

Charge Breeder (<strong>KEK</strong>) Ion source<br />

Frequency / power<br />

Linac Complex (<strong>KEK</strong>) Injection energy<br />

Output energy<br />

(variable)<br />

: SCRFQ linac Frequency<br />

output energy<br />

duty cycle<br />

: IH linac Frequency<br />

output energy<br />

duty cycle<br />

* maximum proton beam power<br />

3<br />

7 Li (68 MeV / 300 pnA)<br />

19 F (78 MeV / 100 pnA), etc.<br />

UCx, BN, Mo, etc.<br />

FEBIAD, SI type<br />

1200<br />

ECRIS<br />

18 GHz / 1 kW<br />

2.1 keV/nucleon<br />

0.14-1.09 MeV/nucleon<br />

25.96 MHz<br />

178.4 keV/nucleon (A/q ≤ 28)<br />

100 % (A/q ≤ 16), 30 % (A/q =<br />

28)<br />

51.92 MHz<br />

0.14-1.09 MeV/nucleon (A/q ≤ 9)<br />

100 % (A/q ≤ 9)<br />

2-2. Ion source and target development<br />

The JAEA ISOL is used for producing singly charged radioactive ion beams for the<br />

<strong>TRIAC</strong>. One of the characteristics of the <strong>TRIAC</strong> is the acceleration of neutron-rich<br />

fission fragments of 238 U induced by low-energy protons. The proton beam of 34 MeV<br />

with intensity of ~ 1 µA is provided from the JAEA tandem accelerator. Two types of<br />

integrated target-ion sources were developed; a forced electron beam induced<br />

arc-discharge (FEBIAD-B2) type [2-1] and a surface ionization (SI) type ion sources.


The FEBIAD-B2 ion source is utilized for ionization of gaseous and volatile elements,<br />

while the SI ion source is used for ionization of alkali, alkaline earth and rare earth<br />

elements. With both ion sources, 105 neutron-rich radioisotopes of 19 elements were<br />

successfully ionized and mass-separated up to now. In the following, the performance<br />

of the ion sources is summarized.<br />

2-2-1. FEBIAD type ion source<br />

The FEBIAD-B2 ion source is shown in Fig. 2-2 (a) and compared to the<br />

FEBIAD-E source (Fig. 2-2 (b)) recently developed for higher operating temperature. A<br />

target container containing about 1 g/cm 2 of 238 UCx is directly attached to the hot<br />

discharge chamber of the source (‘anode with grid’ in the figure).<br />

Fig. 2-2. FEBIAD type ion sources with target container shown without the ion source housing. (1)<br />

Proton beam, (2) capsule cathode, (3) anode with grid, (4) anti-cathode, (5) gas-inlet for the<br />

operating gas from calibrated leaks. FEBIAD-B2 (a) with operating temperature of 1600–1900 K<br />

and FEBIAD-E (b) with operation temperature 2000–2300 K are compared [2-2].<br />

The present configuration of the target container with a heater, in addition to the<br />

heater for cathode (capsule cathode in Fig. 2-2) and the connection pipe ensures to<br />

effectively transport the fission products into the discharge chamber from the uranium<br />

target via diffusion-effusion process.<br />

UCx target preparation: A graphite fiber was chosen as starting material for<br />

synthesizing uranium carbide target for the integrated target–ion source system.<br />

Graphite fibers (φ = 11 µm, GC-20, Tokai Carbon Co.) were filled in the target<br />

4


container of ion source and uranyl nitrate ( nat UO2(NO3)2 / 3M HNO3) solution was<br />

impregnated. After drying in air atmosphere for the formation of UO2(NO3)2-6H2O, the<br />

nitrates were out-gassed and converted to oxides such as U3O8 at about 900 K for about<br />

1.5 hours in argon atmosphere. Then the target container of the oxides is loaded in the<br />

target-ion source system, where the oxides are converted into UCx by in-situ sintering at<br />

about 1800 K for about a half day. The target container (9 mm in diameter and 11-mm<br />

long in case of FEBIAD-B2) is heated by electron bombardment (EB-heating). The<br />

electrons were generated from the filament of the target heater. The potential difference<br />

between the filament and the target container is about 200 V. A certain variation of the<br />

target temperature does not affect the discharge parameters since the target chamber is<br />

equipotential with the anode. The electric power of about 120 W (160 V / 0.75 A) was<br />

supplied for the anode, about 180 W (220 V / 0.8 A) for EB-heating the target container,<br />

when operating at the temperature of 1850 ± 50 K and 1800 ± 50 K, respectively. The<br />

temperatures at the anode and the target were measured with an optical pyrometer and<br />

the corresponding input powers were calibrated with the temperatures. Along this way,<br />

in-situ preparation of UCx target was performed and the typical thickness of the carbide<br />

target corresponds to the thickness of 600-800 mg/cm 2 of nat U. The process for target<br />

preparation is summarized in Fig. 2-3.<br />

Fig. 2-3. Target processing and photograph of the target base material.<br />

The target/ion source was operated at temperature of about 1800 K and at the<br />

pressure approximately 2 x10 -4 Pa. During the operation, ionization efficiencies for<br />

stable xenon and krypton isotopes were periodically checked by using calibrated leaks<br />

of 2.3 x 10 -7 and 4.6 x10 -7 atm cm 3 /s, respectively. The separation yields and<br />

efficiencies for typical uranium fission fragments were summarized in Table 2-2. The<br />

separation efficiency for an isotope is defined as the ratio of the yield mass-separated to<br />

the calculated yield of production.<br />

Although high separation efficiency was observed in the FEBIAD-B2 type ion<br />

source, short-lived isotopes with half-lives shorter than 0.5 s such as 130, 131 In were<br />

5


failed to be mass-separated, which is possibly due to the decay loss caused by a long<br />

release time from the bulk of target material [2-3]. To reduce such decay losses for<br />

short-lived isotopes in the target-ion source system, the higher operating temperatures,<br />

e.g. around 2300 K just as the case of SI ion sources in Table 2-2, are desirable.<br />

Table 2-2. Separation yields and efficiencies for Kr, Rb, Sr, Ag, In, Sn, Xe, Cs and Ba<br />

Surface ionization type ion source operated at 2400 K, while FEBIAD-B2 operated at 1800 K. For<br />

comparison, the measured separation yield is transformed to that to be obtained with the uranium<br />

target thickness of 1 g/cm 2 and the proton beam current of 1 µA.<br />

For this reason, a FEBIAD-E type ion source, which can be operated at the temperature<br />

of 2000 – 2300 K [2-2], has been newly developed. In this type, as shown in Fig.2-2 (b),<br />

the target container is heated by electron bombardment from a couple of tungsten<br />

filaments surrounding the target container. The target container has a volume of 0.56<br />

cm 3 (φ = 6 mm, L = 20 mm), and is welded to a connection pipe so as to avoid leaks for<br />

vaporized fission products. The length of the target container is twice longer than that of<br />

B2 type, while the volume is smaller by 20%. The FEBIAD-E type integrated target–ion<br />

source, which was newly fabricated and off-line tested, has been recently on-line tested.<br />

In the on-line test, the release properties of the short-lived radioisotopes of 93 Kr (T1/2 =<br />

1.286 s), 129 In (T1/2 = 0.61 s) and 141 Xe (T1/2 = 1.73 s) were investigated [2-4]. The<br />

uranium target container with a reduced size in the FEBIAD-E type was heated by<br />

electron bombardment from a couple of surrounding tungsten filaments, and<br />

successfully operated at 2000 °C. Accordingly, the release time for gaseous and volatile<br />

elements was shortened as compared to the previous system (i.e. FEBIAD-B2); 2.6 s for<br />

Kr, 1.8 s for In, and 4.6 s for Xe. Using this ion source system, short-lived isotopes of<br />

6


gaseous and volatile elements (e.g. with lifetimes shorter than 1 s) could be provided for<br />

decay and in-beam spectroscopy with a reasonable intensity.<br />

2-2-2. SI type ion source<br />

For ionization of alkali, alkaline earth and rare earth elements, a surface ionization<br />

type ion source was used and shown in Fig. 2-4. A target container of tantalum with a<br />

thickness of 0.5 mm was directly attached to the ionizer. With this configuration, target<br />

materials can be loaded inside the ion source, which would be very effective to reduce<br />

the release time. The inner surface of ionizer was covered by 50 µm-thick rhenium foil,<br />

and the target container was loaded by about 650 mg/cm 2 of UCx. The integrated<br />

target-ion source is heated by electron bombardment, i.e. electrons emitted from<br />

filaments at a potential of typically 450 V relative to the ionizer / target container. The<br />

ion source operates continuously for over one week at a temperature of 2400 K (total<br />

power input of about 850 W), and can be used repeatedly. With the ion source, the<br />

intensities of the mass-separated alkali, alkaline earth isotopes produced by<br />

proton-induced fission of uranium were measured and summarized in Table 2-2.<br />

2-2-3. Separation yields of neutron-rich isotopes (uranium fission fragments)<br />

Using the FEBIAD-B2 type and SI type ion sources, we have extracted and<br />

mass-separated 105 isotopes of 19 elements as ion beams. Figure 2-5 shows the<br />

intensities of those isotopes available at the focal plane of the ISOL. Those intensities<br />

can be increased by a factor of around 10 with the maximum proton beam power (36<br />

MeV, 3 µA) and the uranium target thickness of 2.6 g/cm 2 .<br />

Fig. 2-4. Surface ionization type ion source (U-SIS) and its power supplying scheme for heating. (1)<br />

Proton beam, (2) thin foil, (3) uranium target, (4) target container (anode), (5) ionizer (the inner<br />

surface is covered by a thin Re foil), (6) filament-1, (7) filament-2, (8) heat shields.<br />

7


Fig. 2-5. Measured RIB intensities at the JAEA-ISOL. For some elements, more detailed<br />

comparison is given.<br />

Fig. 2-6. Separation efficiencies for Ba isotopes as a function of their half-lives. The solid line is<br />

fitted by ε = ε0exp(τsln(2)/T1/2), where T1/2 is the decay half-life, τs is release time and ε0 is the<br />

separation efficiency of 142 Ba. The dotted line is fitted to the data of 142,143,147 Ba. The real<br />

characteristic time (release time) should be between 3 and 9.3 s, since the observed dispersion is<br />

considered to be due to the experimental uncertainties.<br />

2-2-4. Half-life dependence of separation efficiency<br />

Release times and separation efficiencies for radioactive Kr, Rb, Sr, In, Xe, Cs, Ba<br />

and Eu isotopes produced by the proton-induced fission of 238 UCx target were observed<br />

to be strongly dependent on their life-times [2-3], which is shown for Ba isotopes as a<br />

typical example in Fig. 2-6. For example, while the separation efficiency for 142 Ba (T1/2<br />

8


= 10.6 m) were 4.5 %, that for 146 Ba (T1/2 = 2.2 s) was reduced to 0.4 %, i.e. by one<br />

order of magnitude. Considering the decay loss, a characteristic time was extracted from<br />

the correlation between separation efficiencies for isotopes and their half life-times as<br />

shown in Fig. 2-6. The characteristic time for Ba was about 9 s, in good agreement with<br />

release time measured for Ba in the present UCx target module. The release times were<br />

considered to be governed mostly by diffusion process in the bulk of the present<br />

uranium target; the diffusion speed from the uranium target used in this study changes<br />

in the following order, i.e. Rb ~ Cs > Ba > Eu, showing the element dependence as<br />

observed for release time.<br />

Fig. 2-7. Release profile of 7 2-2-5. Production of<br />

Li in the<br />

graphite and BN target<br />

8,9 Li<br />

For producing 8 Li (T1/2 = 838ms), we have used a neutron transfer reaction of 13 C<br />

( 7 Li, 8 Li), using a sintered target of 99%-enriched 13 C. The enriched 13 C graphite disk<br />

(diameter of ~10mm, ≤1mm thick) was mounted to the catcher position of the SI type<br />

ion source with a beam window of 3-µm thick<br />

tungsten [2-1]. The target was bombarded with a<br />

67-MeV 7 Li 3+ beam with intensity of about 100<br />

pnA. In this condition, the separation yield of 8 Li<br />

was evaluated to be 10 6 pps at the focal plane of<br />

the JAEA-ISOL. However, the separation yield<br />

of 9 Li (T1/2= 178ms) was observed as few as 10 2<br />

pps.<br />

A release profile of Li from the target ion<br />

source system was measured using the heavy ion<br />

implantation technique. As shown Fig. 2-7, the<br />

fast component of release time (τ) for Li ions from the 13 C sintered target was 3.2 s.<br />

Decay losses (defined as what remained on the release after decays) of 8 Li and 9 Li were<br />

calculated by exp(-ln(2)/T1/2 x τ) where T1/2 is their half-life and τ is the release time,<br />

giving 7.1x10 -2 for 8 Li and 3.9x10 -6 for 9 Li. Therefore, we could conclude that the<br />

significantly reduced yield of 9 Li is caused by the rather long release time; most 9 Li<br />

decayed out in the extraction processes (i.e. diffusion, effusion in the target materials)<br />

when the graphite target was used.<br />

In search for high-temperature-resistant target material for the production of 9 Li,<br />

we found that boron nitride (BN) has shorter release time for Li; as shown in Fig. 2-7,<br />

the fast component in the release profile from the 0.25mm-thick hot pressed BN sheet<br />

was 120 ms. The decay losses calculated by above function were improved to 9.1x10 -1<br />

9


for 8 Li and 6.3x10 -1 for 9 Li. With a hot pressed BN sheet target, we obtained the beam<br />

of 9 Li with an intensity of 10 4 pps after separation by the JAEA-ISOL.<br />

We observed that the release time became faster with decreasing thickness of the<br />

BN sheet; the value of τfast was 1.9 s for 0.6-mm-thick BN, and 120 ms for<br />

0.25-mm-thick BN. If the enhancement of the yield of 9 Li depends only on the fast<br />

release time, we should also observe any thickness dependence in the ratio of yields<br />

between 9 Li and 8 Li. However, the yield ratio of 9 Li / 8 Li was almost 1/10, irrespective<br />

of the thickness of the BN sheet. In order to understand the observation, further<br />

investigation would be necessary. Nevertheless, using the thick BN target, 9 Li of 3x10 4<br />

pps (one order of magnitude less than the yields of 8 Li) at the target position of the<br />

<strong>TRIAC</strong> were successfully supplied for experiments.<br />

2-3. Charge breeder: <strong>KEK</strong>CB<br />

The <strong>KEK</strong>CB is an electron cyclotron resonance (ECR) ion source operating at the<br />

microwave frequency of 18 GHz, and has been operated for converting singly charged<br />

radioactive ions from the JAEA-ISOL to highly charged ions with A/q < 7 for further<br />

acceleration at the <strong>TRIAC</strong>.<br />

2-3-1. Specification<br />

The cross sectional view of the <strong>KEK</strong>CB is shown in Fig. 2-8. Some of the<br />

important components are discussed as follows.<br />

Energetic electrons, high plasma density and a good ionic confinement are<br />

important ingredients for high charge breeding efficiency, leading to the effective<br />

production of highly charged ions. For producing ions with A/q = 7, for example,<br />

considerable fraction of electrons should have energies exceeding about 600 eV.<br />

Creating this kind of hot electrons with high density in a minimum-B structure is not<br />

difficult whenever a magnetic field configuration for the efficient confinement of hot<br />

electrons and a high microwave frequency are employed. Following the recent scaling<br />

law concerning the magnetic field configuration [2-5], we determined the axial and<br />

radial magnetic mirror ratios possible with conventional solenoid coils and permanent<br />

magnets; Raxial = 2.3 and Rradial = 1.7, where BECR = 0.64 T for 18 GHz. The mirror<br />

ratios are very close to the optimum for an ECRIS operating with this frequency.<br />

Ionic confinement time is closely related to the volume of the plasma chamber of<br />

the source; the ions can reside in plasma for a long time, so that the ionic confinement<br />

time becomes longer and accordingly highly charged ions can be created. With the help<br />

of the systematic data which compare the charge state distributions of Ar from sources<br />

10


with different plasma volumes [2-6], a plasma chamber around 1 l appears to be enough<br />

for our purpose. Furthermore, the volume determined in this way can accommodate a<br />

hot ECR zone with an axial length of 100 mm and a radial diameter of 50 mm. It would<br />

be sufficiently large so that the ECR plasma could efficiently capture injected ions for<br />

charge breeding.<br />

The deceleration system in Fig. 2-8 is a simplified version of that used in the pilot<br />

breeder [2-7]. Instead of the multi-step deceleration, two-step deceleration is adopted by<br />

using two concentric cylindrical electrodes. The outer cylinder, 40 mm in diameter, is<br />

operated as a final stage of deceleration, whereas the inner cylinder, 20 mm in diameter,<br />

stays on the ground potential. Therefore, the main deceleration happens between two<br />

electrodes and further smooth potential drop exists between the outer cylinder and the<br />

plasma chamber. Both of the electrodes are altogether movable axially around the<br />

position of the maximum axial magnetic field for the optimization of the injection under<br />

the influence of the field. The outer cylinder is now removed because of the aging<br />

problem of the insulator on which the electrode was attached. Instead, a similar shape of<br />

electrode is directly attached to the inner tube (plasma chamber). The aging effect was<br />

found due to the irradiation effect of RF wave unintentionally leaking from the ECR<br />

chamber. In the present system, only the ground electrode (i.e. inner cylinder) is<br />

movable and the outer electrode is always on the same potential as the plasma chamber.<br />

Fig. 2-8. Cross-sectional view and specifications of the 18-GHz <strong>KEK</strong>CB<br />

11


2-3-2. Charge breeding efficiency<br />

The charge breeding efficiency εCB was found to depend strongly on the properties<br />

of the beam injection and the ECR plasma [2-8]. In this section, we present two<br />

conditions for optimizing the charge breeding efficiency and the resultant εCB. Two<br />

stable beams of 129 Xe 1+ and 138 Ba 1+ were employed as pilot beams to search for the<br />

optimum conditions. The beams were provided by the JAEA-ISOL; as previously<br />

described in the section 2-2, the FEBIAD type ion source was used for gaseous and<br />

volatile elements such as Xe ions, while the SI type ion source for alkali and alkaline<br />

earth elements such as Ba ions.<br />

Beam injection: We installed two collimators for defining the beam axis when<br />

injecting singly charged ions to the <strong>KEK</strong>CB. One collimator functions as the entrance<br />

hole for injection and the other as the exit hole for beam extraction of the <strong>KEK</strong>CB; their<br />

diameters were 6 and 7 mm, respectively. The distance between them was about 300<br />

mm. An additional collimator (the third one) with a diameter of 6 mm was used to<br />

define the object point of the Einzel lens doublet placed upstream of the <strong>KEK</strong>CB as<br />

shown Fig. 2-8. The third collimator was located 1000 mm upstream from the entrance<br />

hole. At the first stage of beam tuning, we transported the stable beams from the<br />

JAEA-ISOL to the exit of the <strong>KEK</strong>CB by using the collimator system without<br />

deceleration, where the injection axis of beam optics could be defined. The typical<br />

transport efficiency was around 60 %. We then began charge breeding by decelerating<br />

the beam at the entrance. The beam injection was optimized by using the Einzel lens<br />

doublet and electric quadrupole triplet located just upstream of the Einzel lens in the<br />

beam transport line at the <strong>TRIAC</strong>. Tuning the injection was found to be very critical,<br />

especially for charge breeding non-gaseous elements. The deceleration of the incident<br />

beams was further optimized by adjusting ∆V, where ∆V denotes the electric potential<br />

for adjusting the beam energy at injection into the ECR plasma. Figure 2-9 shows the<br />

dependence of charge breeding efficiencies (εCB) for the 129 Xe 1+ and 138 Ba 1+ beams on<br />

∆V. For Ba ions, a strong dependence was observed with a narrow width of about 4 V,<br />

in contrast to the weaker dependence observed for Xe ions. The element dependence of<br />

∆V is discussed below in detail. The results demonstrate that gaseous elements could be<br />

more energetic at injection for higher εCB. For charge breeding RIBs whose masses<br />

differed from that of the pilot stable beam, we only varied the parameters of the<br />

analyzing magnet of the JAEA-ISOL since the other optical components in the beam<br />

transport line are electric elements.<br />

12


ECR plasma: We used natural helium as the support gas and a microwave power of<br />

about 200 W to generate ECR plasma suitable for charge breeding. When we employed<br />

oxygen gas, a lower microwave power of 150 W was sufficient to obtain the same εCB.<br />

However, we adopted helium gas to avoid the long tail of the 16 O 2+ ions in the A/q < 8<br />

region. The plasma that we used was optimized for the charge states of A/q ~ 7. The<br />

charge state distributions for 129 Xe 1+ and 138 Ba 1+ are shown in Fig. 2-10, which show<br />

that the highest values for εCB were obtained at A/q ~ 7. Both distributions were found<br />

to have almost the same widths. The entrance hole for injection was found to play an<br />

important role in reducing the microwave power loss and stabilizing the plasma. Before<br />

installing the hole, a power as large as 400 W was required to achieve the same εCB for<br />

the charge states of A/q ~ 7. The small hole is very effective not only to reduce the<br />

conductance, but also to greatly reduce the microwave power loss, enabling us to<br />

produce good plasma even with a low microwave power. The low power resulted in a<br />

small amount of the total beam intensity being extracted, giving rise to a plasma stable<br />

against the electrical discharge of the high voltage supply for beam extraction.<br />

Fig. 2-9. ∆V versus charge breeding efficiency (εCB) for 129 Xe 1+ 129 Xe 19+ (dashed, but scaled down<br />

by a factor of 3) and 138 Ba 1+ 138 Ba 20+ (solid). The charge breeding efficiency was defined by εCB =<br />

(Iq+/q)/I1+ x100 (%), Iq+ and I1+ are the electric currents of charge-bred ions with a charge state of q+<br />

and injected singly charged ions, respectively. ∆V is the electric potential difference between the 1 +<br />

ion generator (SI for 138 Ba 1+ and FEBIAD for 129 Xe 1+ ) and the charge breeder. With increasing<br />

values of ∆V, the 1 + ions become more energetic at the entrance of the ECR plasma. For indium<br />

(metallic ions generated by FEBIAD as for Xe), the ∆V distribution was centered around 30 V in the<br />

middle of the upward slope of εCB values for Xe with a width as narrow as for Ba (not shown here).<br />

13


Fig. 2-10. Charge sate distributions of 129 Xe (lower panel) and 138 Ba (upper panel). Indices next to<br />

the peaks indicate the charge states.<br />

Table 2-3. Charge breeding efficiencies. Upper two rows indicate the gaseous elements, while the<br />

lower rows present the non-gaseous elements. See the text for details.<br />

Charge breeding efficiencies: The εCB for radioactive ion beams, 92 Kr 1+ and 126 In 1+ , as<br />

well as the stable ion beams are summarized in Table 2-3 [2-8]. The efficiencies for<br />

gaseous elements are higher by a factor of 3 than those for nongaseous elements. The<br />

14


higher efficiency for gaseous elements is considered partly due to desorption of the<br />

injected ions on the surface of the plasma chamber. The desorption of the gaseous ions<br />

may lead to the weak dependence of εCB on the ∆V over a rather wider range as shown<br />

in Fig. 2-8. On the other hand, the lower value of εCB for the nongaseous elements<br />

suggests that the adsorption to the surface is not negligible. The improvement of the<br />

beam injection efficiency to the ECR plasma may help increase εCB for the nongaseous<br />

elements, if the ionization efficiency of ECRIS is element independent. The εCB was<br />

found to be independent of the lifetime for radioactive nuclei with lifetimes of a few<br />

seconds, as shown in Table 2-3. This result suggests that the charge breeding time is<br />

shorter than the lifetimes of presently studied isotopes, which is consistent with the<br />

charge breeding time of about 100 ms measured at the <strong>KEK</strong>CB [2-9]. We suppose that<br />

εCB could be dependent on lifetime for RIBs with lifetimes shorter than a few 100 ms.<br />

2-3-3. Element-dependent charge breeding efficiency<br />

The charge breeding efficiencies are 7.4 % for Xe 20+ and 2.3 % for In 16+ , which are<br />

typical efficiencies for radioactive gaseous and metallic elements with a half-life time of<br />

~ 1 s., as well as for stable ions already shown in Fig. 2-8. The efficiencies for gaseous<br />

elements are observed to be higher than those for metallic ones, usually by a factor of<br />

three as long as the similar masses of the elements are concerned. Because of the<br />

significantly different behavior in the injection, which just depends on if the element is<br />

volatile or not at room temperature as discussed above, one can conclude that the lower<br />

charge breeding efficiencies for metallic ions could be improved by more careful<br />

optimization of the injection. The difference could be attributed to ad- and de-sorption<br />

of the ions on the surface of the plasma chamber during the injection; the gaseous<br />

elements adsorbed on the surface could be more easily desorbed and recycled for further<br />

ionization, as compared to the case of metallic, non-volatile elements that stick to the<br />

surface once adsorbed. On the other hand, the element-dependent ionization efficiency<br />

cannot be completely excluded. Therefore, in order to identify the direction of the<br />

development to obtain higher breeding efficiency, especially for metallic elements, it is<br />

of importance to study how the injected ions, but failed to be re-extracted, are<br />

distributed on the surface of the plasma chamber. The distribution of the injected ions<br />

accumulated on the surface of the plasma chamber would help us to understand when<br />

the injected ions are adsorbed, i.e. at the time of injection or in the course of the<br />

step-by-step ionization in the ECR plasma after being well captured.<br />

We have injected radioactive ions of 111 In into the <strong>KEK</strong>CB and, after charge<br />

breeding, measured the residual activity to investigate how ions externally injected<br />

15


should be lost in the course of charge breeding [2-10]. A typical azimuthal distribution<br />

around the Bmin is given in Fig. 2-11. We assumed that the distribution consisted of three<br />

components, azimuthally isotropic, 120 o -symmetric and asymmetric ones. At each<br />

longitudinal (axial) position, we decomposed the distribution according to the azimuthal<br />

symmetry; asymmetric and symmetric components. The symmetric includes isotropic<br />

and 120 o -symmetric components. Integrating over azimuthal angle, the longitudinal<br />

distributions were extracted and compared in Fig. 2-12, where axial magnetic field<br />

configuration is also given. The asymmetric component is localized around Bmin, while<br />

the symmetric component is concentrated around the Bmin as well as at the extraction<br />

side.<br />

Fig. 2-11. Azimuthal distribution of 111<br />

In around Bmin. Decomposed into 3 components; azimuthally<br />

isotropic, 120 o -symmetric, asymmetric ones.<br />

Fig. 2-12. Longitudinal (axial) distribution of the symmetric and asymmetric components. The<br />

injection and extraction positions are indicated by arrows, respectively. The symmetric includes both<br />

the isotropic and 120 o -symmetric components. Axial magnetic filed configuration is also given by a<br />

dotted line for comparison.<br />

16


The asymmetric component representing 13.5 % of the total activity residual on the<br />

surface might be associated with ions injected in a rather asymmetric manner, since the<br />

(axial and radial) magnetic field configuration for the confinement of electrons in ECR<br />

plasma cannot produce such asymmetry. More careful optimization of the beam optics<br />

for the injection is necessary for the elimination of the asymmetric component in the<br />

ion-losses in the course of charge breeding.<br />

The 120 o -symmetric distribution of ions reminds us of the electron losses due to<br />

the combined field configuration of the axial and the radial magnetic field for electron<br />

confinement [2-10]. Although the behavior of ions in ECR plasma has not yet been<br />

studied in detail both experimentally and theoretically, we can expect that both electrons<br />

and ions should show a quite similar behavior. Therefore, the 120 o -symmetric<br />

distribution can be considered as a characteristic behavior of ions in ECR plasma. When<br />

considering the charge-breeding process as two successive processes, i.e. stopping and<br />

ionizing processes [2-11], the ions with the 120 o azimuthal symmetry would correspond<br />

to those lost during the second, ionizing process. In the ionization process, those ions,<br />

though successfully captured by ECR plasma in the stopping process, are supposed to<br />

be lost to the wall of plasma container along the hexagonal radial magnetic field. The<br />

fraction, i.e. how many ions are lost in the ionization process, is nothing but the<br />

ionization efficiency for the ions in ECRIS.<br />

As shown in Fig. 2-11, the 120 o -symmetric component is on the top of the isotropic<br />

one. The isotropic one is actually the main component of the symmetric distribution.<br />

The isotropic component was observed all over the wall, while the 120 o -symmetric<br />

component was observed only around Bmin, even where the isotropic component<br />

amounts to a large fraction (refer to Fig. 2-11). Along the discussion given above for the<br />

120 o symmetry in the distribution, the isotropic component could be also associated<br />

with the incomplete confinement of ions, i.e. the losses simply by the ambipolar<br />

diffusion of plasma constituents.<br />

The symmetric component, which includes isotropic and 120 o -asymmetric<br />

components, represents 86.5 % of the total. The 38 % of the symmetric component were<br />

observed at the extraction side. The large fraction of ion losses around the extraction<br />

side could be understood qualitatively by a weak confinement field. The weak radial<br />

field caused by the geometrical limit of permanent magnets at the extraction side, as<br />

well as the weak axial field for efficient ion extraction, could lead to rather poor radial<br />

confinement of electrons and consequently large loss of ions along the radial direction.<br />

Further measurements using the gaseous element 140 Xe have been performed. The<br />

17


distribution of the residual activities of 140 Xe was compared with those of 111 In as shown<br />

in Fig. 2-13. We observed well-localized azimuthal distribution with 120 o periodicity for<br />

both elements, especially prominent around the minimum (Bmin at Z~200 mm) of the<br />

axial field configuration for electron confinement. However, for the 111 In, an<br />

asymmetric distribution with a strongly localized around φ~300 o and Z~200mm was<br />

observed around the Bmin, while the 140 Xe distribution had almost same peak heights<br />

(symmetric). In order to indentify the origin of the asymmetric components, further<br />

experimental investigation is in progress.<br />

Fig. 2-13. Distributions of residual activities of 111 In (left) and 140 Xe (right) on the surface of<br />

plasma chamber as function of azimuthal angle (φ) and longitudinal position (z). See the text for<br />

detail.<br />

Analysis of the residual activity distribution on the ECR plasma chamber surface<br />

revealed that most ions, especially gaseous 140 Xe ions, were lost on the side wall near<br />

the central region where there is a minimum in the axial magnetic field for electron<br />

confinement in the ECR plasma. While the 120° azimuthal periodicity is superimposed<br />

on the isotropic distribution around Bmin, the lost ion patterns are almost azimuthally<br />

isotropic on the extraction and injection sides. The azimuthal periodicity in the ion loss<br />

pattern, which was observed for both volatile and non-volatile elements, differs from the<br />

60° azimuthal periodicity around Bmin expected on the basis of the pattern of lost<br />

electrons, which is often observed as an imprinted pattern on the inner surface of ECR<br />

plasma chambers. The ion loss pattern, which was first observed by our present study,<br />

18


ut differs from the electron-loss pattern, could be due to the direct injection of 1 + ions<br />

for charge breeding; the injected ions may be preferentially stopped and captured on the<br />

extraction side as a result of ion-plasma collisions<br />

Additionally, a strong azimuthal asymmetric component localized near the central<br />

region and a symmetric component localized on the extraction side, were observed in<br />

the wall-loss distribution of the metallic 111 In ions. They account for 13.5 and 38.5 % of<br />

the total side-wall loss of 111 In, respectively. They were not observed for xenon.<br />

Therefore, the presence of such components in the wall-loss pattern of indium could be<br />

why non-volatile metallic elements have lower charge breeding efficiency than gaseous<br />

elements. We suggested that the charge breeding efficiencies for metallic ions at the<br />

<strong>KEK</strong>CB could be improved by further optimization of the injection conditions for<br />

charge breeding. However, it is not straightforward to further optimize our present<br />

injection system than we have already performed for charge breeding non-volatile<br />

metallic ions. Actually, the injection conditions can be varied by the parameters related<br />

to the conditions of ECR plasma, which are not usually controllable in injection.<br />

Therefore, more experimental works are required in order to identify the origin of the<br />

loss patterns of metallic ions during the injection for charge breeding. Detailed<br />

theoretical simulations are also highly desirable for gaining better understanding of the<br />

present observations.<br />

2-3-4. Beam impurities from <strong>KEK</strong>CB<br />

The ECR plasma inherently produces a variety of ions from residual gases, the<br />

plasma chamber, and other sources. Such ions are considered as impurities for the beam<br />

of interest. To evaluate the impurity quantitatively, we accelerated background ions up<br />

to 178 keV / nucleon by using the post-accelerator at the <strong>TRIAC</strong> [2-8]. As an example,<br />

the background ions of A/q = 7.68 were selected by using an analyzing magnet and slits<br />

with a full width of 10 mm along the horizontal direction (covering both sides of the<br />

central trajectory) located at exit of the magnet. The magnet had a resolving power of<br />

about 100 when the slits were open to 2-mm width. It is possible to identify the mass<br />

number of impurity ions by measuring the total energy of the beams, since the<br />

acceleration of the impurities has the same energy per nucleon. In the first trial, we<br />

observed various heavy elements with an intensity of 10 7 ~ 10 8 pps. After this test, all<br />

the materials exposed to the ECR plasma were converted to aluminum, since some of<br />

them were made of stainless steal and chromium. Before installation, we ground their<br />

surface using a sand blaster and high-pressure purified water, and cleaned them by<br />

ultrasonic cleaning in order to remove any elements that had attached to their surfaces in<br />

19


the machining processes. As a result, several peaks corresponding transition metallic<br />

elements disappeared, and the intensity of the impurity peaks decreased dramatically to<br />

about 600 pps in total, making it feasible to perform experiments with the RIBs. The<br />

main component of the remaining impurities is considered to be krypton, which is<br />

naturally abundant in the residual gases in the source. When the beam was defined more<br />

rigorously by the slits, from 10 to 2 mm in width, a reduction of more than a factor of 5<br />

was expected for the background. In reality, however, the actual desired beam reduction<br />

was a factor of 2. We also checked the region of A/q between 6 and 7. As an example,<br />

we studied the background ions of A/q = 6.45, which is very close to 6.4 ( 32 S 5+ ) and 6.5<br />

( 13 C 2+ ), with the slits of 2 mm in width. The background intensity was found to be only<br />

several tens of pps.<br />

It should be noted that the intensity of background components measured at the<br />

focal plane of the analyzing magnet of the <strong>KEK</strong>CB was about 1 nA, not depending on<br />

the surface-cleaning: Most of them can be removed at the <strong>TRIAC</strong>, not only by<br />

additional beam optics fine-tuned to A/q of 6.45 in the beam transport line but also<br />

during acceleration.<br />

2-4. Accelerator<br />

The post accelerator at the <strong>TRIAC</strong> comprises a 25.96-MHz split-coaxial RFQ with<br />

modulated vanes and a 51.92-MHz interdigital–H (IH) linac, as shown in Fig. 2-1. Main<br />

specifications of the RFQ / IH linac are listed in Table 2-4.<br />

2-4-1. Specification: SCRFQ / IH linacs<br />

SCRFQ linac: The SCRFQ accelerates ions with a charge-to-mass ratio (q/A) greater<br />

than 1/28 from 2 to 178 keV/nucleon. The duty factor can be 30 % for ions with q/A =<br />

1/28 and 100 % for q/A > 1/16. The cavity, 0.9 m in inner diameter and 8.6 m in length,<br />

comprises four unit cavities, each of which is composed of three modules. Obtained<br />

flatness of the longitudinal field distribution is within ±1 %. Unloaded Q-value is 5800.<br />

The resonance resistance (= V 2 /2P) is 24.55 ± 0.44 kΩ, which is obtained from a<br />

relation between the input RF-power and the intervane voltage derived from the<br />

endpoint energy of X-rays generated from the cavity. The input power for accelerating<br />

q/A ~ 1/28 ions is nearly 225 kW.<br />

Transport line between RFQ and IH linac: A transport system between the RFQ and<br />

IH linac comprises a rebuncher and two sets of quadrupole doublet. The rebuncher is a<br />

25.96-MHz double coaxial quarter wave resonator with six gaps. The power<br />

20


consumption in the cavity is less than 1.5 kW. The rebuncher operates at a 100-% duty<br />

factor.<br />

Spacing Rod<br />

Beam<br />

Cooling Channel<br />

Back Plate<br />

Stem Flange<br />

Table 2-4. Main specifications of the <strong>TRIAC</strong> linac<br />

Vertical Vane<br />

Horizontal Vane<br />

Window<br />

Stem<br />

Fig. 2-14. SCRFQ view (one of the four unit cavities).<br />

IH linac: The IH linac accelerates ions with a q/A greater than 1/9 up to 1.09<br />

MeV/nucleon. The IH is a separated function drift-tube linac (SDTL), which comprises<br />

four tanks and three magnetic quadrupole triplets between tanks. Output beam energy<br />

can be continuously varied by adjusting independently the phase and amplitude of the<br />

RF field in each tank. The accelerating mode is π-π, and no transverse focusing element<br />

21


is installed in the drift tubes. The achieved effective shunt impedances of the 1st through<br />

4th tanks are 274, 268, 249 and 228 MΩ/m, respectively. From the measured shunt<br />

impedances, we figured out the RF powers required for accelerating ions with q/A = 1/9,<br />

i.e. 12, 22, 30 and 50 kW for the 1st through 4th tanks, respectively.<br />

Magnetic Q-triplet Rf-Coupler<br />

L-Tuner Ridge<br />

Drift Tube<br />

Fig. 2-15. IH view<br />

Modifications of cavities: The SCRFQ [2-12] and IH linacs [2-13] primarily have<br />

resonant frequencies of 25.5 and 51 MHz, respectively, while the JAEA Super<br />

Conducting (SC-) linac has a resonant frequency of 129.8 MHz. We changed the<br />

resonant frequency of the RFQ from 25.5 to 25.96 MHz and that of IH from 51 to 51.92<br />

MHz for future’s connection to SC-linac. The resonant frequency of the RFQ was tuned<br />

by changing locally interelectrode capacitance and stem (used for supporting the<br />

electrodes) inductance. Values of the capacitance and inductance to be changed were<br />

well estimated by using an equivalent circuit analysis; with the changed frequency, the<br />

accelerating performance of the SCRFQ was shown to be kept except that the beam<br />

energy changes from 172 to 178 keV/nucleon. The resonant frequency of the IH linac<br />

was tuned by increasing the gap lengths between drift-tubes, that is, by decreasing the<br />

capacitance. All drift-tubes were replaced to the modified ones. The gap lengths were<br />

determined by model tests after rough estimations by MAFIA calculations.<br />

2-4-2. Performances (obtained before the cavity modifications)<br />

Transmission efficiency: The transmission efficiency of the RFQ was measured as a<br />

function of the intervane voltage by using N2 + , N + and H2 + beams. Beam currents were<br />

22<br />

Stem


measured with Faraday cups, FC1, FC2 and FC3. FC1 is installed at the entrance, and<br />

FC2 at the exit of RFQ. A quadrupole doublet is installed between FC2 and FC3. Under<br />

the condition when the magnetic field strength of the quadrupole doublet is sufficient to<br />

focus the accelerated ions into FC3, the transmission for the accelerated ions is obtained<br />

from I(FC3) / I(FC1). Here, I(FCi) is the beam current from the i-th Faraday cup. The<br />

transmission for the drift-through ions (containing unaccelerated ions) is given by<br />

I(FC2) / I(FC1).<br />

2-16. Transmission of the RFQ vs. intervane voltage<br />

Fig. 2-17. Transmission along the accelerator.<br />

The measured transmission is about 90 % at design voltage and agrees well with the<br />

simulation by PARMTEQ [2-12]. From Fig. 2-16, it was verified that the intervane<br />

voltage is applicable in a very wide range from 109 kV down to 1/14 of that voltage.<br />

The transmission efficiencies of the RF/ IH linac were measured along the accelerator<br />

23


y using N 2+ beam. For the measurement, FC4 and FC5 installed at entrance and exit of<br />

the IH linac respectively were used as well as FC1 and FC2. The transmission of the<br />

RFQ is nearly 95 % and that of the IH nearly 100 %, as shown in Fig. 2-17.<br />

Energy and energy spread: We measured the output energy T and energy spread ∆T.<br />

Figure 2-18 shows the output energy spectra measured at six operating modes, which<br />

are described later. The IH-tank4 energy agrees well with a designed one, 1.053<br />

MeV/nucleon. In Table 2-5, the measured T and ∆T values are compared with the<br />

calculated values in parentheses. ∆T is defined by 2-rms of the spectrum containing<br />

90% ions. The measured energy spreads are smaller than calculated one except for the<br />

RFQ mode and variable modes.<br />

Fig. 2-18. Energy spectra for six operating modes<br />

Table 2-5. Energy and its spread for six operating modes. Numbers in parentheses are from<br />

calculations. The results summarized here were obtained before frequency modifications<br />

Different accelerating modes: The linac is operated at different accelerating modes to<br />

change the output energy. For example, the IH-tank2 energy in Table 2-5 was obtained<br />

24


y turning off the supply of RF power for the tank3 and tank4, and a variable energy by<br />

adjusting the RF power voltage and phase in tank4. Deceleration test of the RFQ beam<br />

was conducted by using the IH-tank1, in order to extend the variable-energy range. In<br />

the experiment, He + beam was injected at a decelerating phase of 135 o of the IH-tank1,<br />

and the output energy and its spread were measured as a function of the IH-tank1 RF<br />

voltage. Figure 2-19 shows that the minimum energy is 112 keV/nucleon at a<br />

normalized voltage of about 1.8.<br />

Fig. 2-19. Energy and its spread vs. normalized voltage.<br />

RF stability: The RF power sources have an automatic gain and phase control system,<br />

in which a feedback signal is picked up from the forward power to the cavity. As for the<br />

forward power, voltage variation is within 0.25 %, the phase one within 0.2 o . However,<br />

we cannot ignore the voltage and phase variations in the cavity due to the temperature<br />

change. We added three feedback systems to compensate the temperature change effect<br />

on the RF voltage and phase in the cavity. One is a piston-turner control system for<br />

tuning the cavity. The piston tuners are moved automatically so as to minimize the<br />

reflected power from the cavity. The control system comprises the circuits for reading<br />

the reflected power levels, the tuner driver circuits, and a personal computer. The<br />

second is for keeping the cavity RF level constant in long-term by adjusting a control<br />

signal (DC voltage) for RF output at a ten-second interval. The third is for constantly<br />

keeping the phase difference between different cavities, the test result shows the phase<br />

stability is within 0.2 o .<br />

We have found the instability of the RF power supplied for SCRFQ. The instability<br />

is associated with two components of voltage fluctuation: one is caused randomly in a<br />

few minutes (slow component) and another is by electronic noises synchronized with<br />

AC line (50 Hz component). The slow and 50 Hz components of the cavity voltage<br />

fluctuation are shown by two red curves (one is in the inset) in Fig. 2-20, and the<br />

magnitude of them corresponds to 20 % and 8 % of the nominal cavity voltage,<br />

25


espectively. By modulating the control voltage signal for RF output so as to keep the<br />

cavity voltage constant, both components were reduced to 2 % in magnitude. The RF<br />

power stabilization system is now well operational for the stable beam delivery at the<br />

<strong>TRIAC</strong>.<br />

Fig. 2-20. Time variations of the cavity voltage are compared before and after introducing feedback<br />

control for RF output. The short-term variations are given in the inset.<br />

2-4-3. Simulation for upgrading output energy [2-14]<br />

The present SCRFQ / IH complex can accelerate RIBs with a charge-to-mass ratio<br />

(q/A) greater than 1/9 from 2.1 keV/nucleon to 1.09 MeV/nucleon. In order to increase<br />

the beam energy up to 5-8 MeV/nucleon, we have made a plan to connect the IH linac<br />

to the 129.8-MHz JAEA SC-linac existing at the Tandem facility. Although the plan has<br />

not yet been realized, the frequencies of the SCRFQ and IH linacs have been already<br />

changed from 25.5 to 25.96 MHz and 51 to 51.92 MHz, respectively, so as that their<br />

frequencies become equal to 1/5 and 2/5 of the SC-linac frequency. The performance of<br />

the 51.92-MHz IH linac was confirmed by the beam simulation using TRACEP code<br />

[2-15]. After the beam simulation for normal operation based on the design, our<br />

attention was paid on the special operation to bring the output beam energy to the<br />

maximum restricted by the accelerating voltage limit. As a result of the simulation for<br />

normal operation, the relation between output beam energy and synchronous phase of<br />

the first accelerating cell is obtained for each tank. This relation is efficiently used as a<br />

standard for adjusting the accelerating voltage and phase in the real operation of the<br />

linac. Most importantly, by a subsequent simulation for the special operation, we show<br />

that the IH linac could accelerate ions with a q/A = 1/4 up to 1.4 MeV/nucleon higher<br />

than designed energy of 1.09 MeV/nucleon.<br />

26


Encouraged by the new finding, as the most realistic solution for output energy<br />

upgrade at <strong>TRIAC</strong>, we have examined the possibility for directly connecting the present<br />

IH to the SC-linac. A most suitable energy for injection into the SC-linac is about 2<br />

MeV/nucleon. However, we investigated the possibility of 1.4 MeV/nucleon injection<br />

from the following reasons: 1) It is found that the IH linac is able to accelerate ions with<br />

a q/A = 1/4 up to 1.4 MeV/nucleon with a tolerable reduction of beam quality for the<br />

beam energy higher than designed, 2) Connection of the <strong>TRIAC</strong> to the SC-linac can be<br />

realized with a significantly reduced cost and time, if it is found to be unnecessary to<br />

construct a new linac for boosting the IH energy to 2 MeV/nucleon.<br />

Fig. 2-21. Relation between input and output energies of SC-linac.<br />

Fig. 2-22. Beam transport from IH to SC-linac<br />

27


The JAEA SC-linac comprises 10 cryostats and 9 magnetic quadrupole doublets<br />

set between cryostats. Four super-conducting cavities made of niobium are installed in<br />

each cryostat. The cavities are 129.8-MHz two-gap λ/4 resonators. The optimum ion<br />

velocity is 10 % of light velocity for all cavities. This linac has large velocity<br />

acceptance, since the number of accelerating gap per cavity is only two and RF phase of<br />

each cavity can be independently adjusted. The relation between input and output<br />

energies were calculated for the cases of q/A = 1/4 and 1/7 under the condition that the<br />

gap voltage is 375 kV and the accelerating phase -20 o at first gap. Figure 2-21 shows<br />

that the SC-linac can accelerate ions with the injection energies higher than 1.1<br />

MeV/nucloen. When the injecting energy is 1.4 MeV/nucleon, the maximum energy is<br />

~ 7 MeV/nucleon for ions with q/A = 1/4.<br />

A designed beam-transport line between IH and SC linacs is 37.71 m in total<br />

length, and comprises 19 quadrupole magnets, two 45 o deflection magnets and two<br />

25.96-MHz rebunchers [2-16]. Performances of the beam transport line were examined<br />

by means of the beam simulations using MAGIC code [2-17]. The results are shown in<br />

Fig. 2-22. The obtained important results are as follows: Beam transmission efficiency<br />

from IH entrance to SC linac exit is 67 %. If 129.8-MHz buncher and 259.6-MHz<br />

sub-harmonic buncher for injecting the tandem beam into SC linac are used instead of<br />

two rebunchers, the efficiency is 40 %. From above results, it is concluded to be a<br />

realistic plan to connect the <strong>TRIAC</strong> to the SC linac by using only existing devices<br />

without constructing newly a booster linac and two rebunchers.<br />

Test experiments: As mentioned above, a slight modification of acceleration mode of<br />

the IH linac, especially final two acceleration stages of the IH linac, has been simulated<br />

for higher maximum energy than designed: The higher energy could allow us to further<br />

accelerate RI beams by SC-linac without additional equipments related to acceleration,<br />

such as re-buncher and pre-booster between the IH and the SC linacs. The simulation<br />

was confirmed by a test operation; indeed the maximum energy can be increased from<br />

1.1 to 1.4 MeV/nucleon at the present facility. For the SC-linac where at least 2<br />

MeV/nucleon at the injection is required for the effective acceleration, on the other hand,<br />

a 40 Ar 10+ beam injected from the JAEA tandem accelerator was successfully accelerated<br />

from 1.4 to 4.53 MeV/nucleon. It should be noted, however, that in this scheme the<br />

beam transport and acceleration efficiency of the SC-linac is lower than what is<br />

expected in the ideal case. Therefore, the present conclusion resulting from the test<br />

experiments and simulations should be taken as a first step for step-by-step realization<br />

of the ideal plan.<br />

28


2-4-4. Development of superconducting low-β resonators [2-18]<br />

As mentioned earlier for the specification of the post-accelerators at the <strong>TRIAC</strong><br />

(see subsection 2-4-1), the cavities were modified to match the resonant frequency of<br />

the superconducting resonators (129.8-MHz two-gap λ/4 resonators) of the JAEA<br />

SC-linac (i.e. from 25.5 to 25.96 MHz, from 51 to 52.92 MHz for the SCRFQ and the<br />

IH, respectively), where the injection energy of at least 2 MeV/nucleon is required for<br />

the effective acceleration. Therefore, we need a pre-booster between the IH and SC<br />

linacs for boosting the output energy of the IH (1.1 MeV/nucleon) to 2 MeV/nucleon,<br />

although we have discussed in the previous subsection the possibility to connect the<br />

present <strong>TRIAC</strong> to the SC linac by using only existing devices.<br />

We have designed and fabricated a superconducting twin-quarter-wave resonator<br />

(Twin-QWR) made of niobium and copper as a pre-booster at the <strong>TRIAC</strong>. It has 2 drift<br />

tubes and 3 acceleration gaps, as shown in Fig. 2-23. The resonant frequency is 129.8<br />

MHz, which is the same for the existing QWRs in the SC linac. The optimum beam<br />

velocity βopt (= vopt/c) is 0.06.<br />

Fig. 2-23. Cross sectional view of Twin-QWR.<br />

29


mode, and this is the mode for acceleration. The measured frequency separation<br />

between two modes is about 0.86 MHz, which is enough for stable operation. Table 2-6<br />

presents the main parameters of the present Twin-QWR obtained from the measurement<br />

and calculation. Figure 2-24 shows the transit time factors of Twin-QWR and QWR as a<br />

function of beam velocity. The optimum beam velocity βopt of Twin-QWR is 0.06, and<br />

βopt of QWR is 0.10. The beams accelerated by the present <strong>TRIAC</strong> have the velocity of<br />

about 0.048×c (1.1 MeV/nucleon), which can be efficiently accelerated by the presently<br />

developed Twin-QWR.<br />

Transit Time Factor<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0.0<br />

-0.2<br />

Table 2-6. Main parameters of Twin-QWR<br />

Beam from <strong>TRIAC</strong>(1.1MeV/u)<br />

0 0.05 0.1 0.15 0.2<br />

Beam velocity β(= v/c)<br />

31<br />

3gap Twin-QWR<br />

2gap QWR<br />

Fig. 2-24. Transit time factors of presently developed Twin-QWR and exiting QWR as a function of<br />

beam velocity.<br />

In the design of the inner conductor part, it is important to hinder frequency<br />

instability. The frequency instability is mainly caused by a deformation of the top end


accelerating the radioactive ion beams from the present <strong>TRIAC</strong> (i.e. from 1.1<br />

MeV/nucleon to 2 MeV/nucleon as an input energy for the SC linac).<br />

2-4-5. Development of re-bunchers for energy up-grading of the present <strong>TRIAC</strong><br />

As mentioned in the previous subsection, the original project of the <strong>TRIAC</strong> was<br />

made to connect the present <strong>TRIAC</strong> with the JAEA superconducting linac, increasing<br />

the energy of the RIB to 5~8 MeV/nucleon. The beam transport line between them has<br />

been designed, and Fig. 2-26 shows the calculated beam envelopes.<br />

Fig. 2-26. Calculated beam envelopes between the <strong>TRIAC</strong> facility and the JAEA superconducting<br />

linac. The charge-to-mass ratio of the beam is assumed to be 1/7. The horizontal and longitudinal<br />

beam profiles are shown in the upper half (blue and green lines, respectively), and the vertical profile<br />

(red line) is shown in the lower half. The symbol “G” indicates a re-buncher. The narrow and wide<br />

boxes indicate a quadrupole and dipole magnet, respectively.<br />

Fig. 2-27 (Left) Cold model of the re-bunchers to transport RIBs from the present <strong>TRIAC</strong> to the<br />

JAEA SC-linac (Right) Correlation between the resonance frequency and the partition plate position<br />

of the re-buncher model.<br />

34


Two re-bunchers need to be newly constructed, and a cold model of them has been<br />

made and tested. Figure 2-27 (Left) shows the model of the bunchers. The cavity has a<br />

double coaxial structure with a movable partition plate to tune the resonance frequency.<br />

Figure 2-27 (Right) shows the correlation between the resonance frequency and the<br />

partition plate position. The resonance frequency of 51.814 MHz, which is required to<br />

match the RIBs from the <strong>TRIAC</strong> to the JAEA SC-linac, has been achieved by adjusting<br />

the partition plate position. The electric field in the model has also been measured and<br />

compared with the calculation.<br />

2-4-5. Beam pre-bunching system [2-22, -23]<br />

A beam pre-bunching system has been recently installed for producing a pulsed<br />

beam with a width of less than 10 ns and an interval between 250 ns and 500 ns. The<br />

use of the pulsed α beam is essential for the experiment (RNB-KJ04) to directly<br />

measure the 12 C(α, γ) reaction cross sections at a highly reduced γ-rays background<br />

expected in the proposed experimental set-up.<br />

The system consists of a multi-layer beam chopper and a two-gap pre-buncher,<br />

respectively installed at the F2-chamber and just upstream of the SCRFQ linac, as<br />

shown in Fig. 2-28.<br />

F2 Chamber<br />

SCRFQ<br />

Fig. 2-28. Layout of the beam line upstream of the SCRFQ, where the beam pre-bunching system is<br />

installed<br />

35


Fig. 2-29. Photos of the pre-buncher; outside view (Left panel) and inside view (Right panel).<br />

The pre-buncher, as shown in Fig. 2-29, has two gaps at both ends of a single drift<br />

tube respectively applied by a RF-wave synthesized with three harmonics (from the first<br />

to the third), which are combined together into a pseudo saw-tooth wave-form for beam<br />

bunching. By adjusting the length of the drift tube to be 140/360 in units of βλ<br />

according to the beam-bunching frequency, where β is the velocity of the beam (2 keV/<br />

nucleon at the <strong>TRIAC</strong>) normalized by the light velocity and λ is the wavelength of the<br />

applied RF-waves for beam bunching, the pseudo saw-tooth wave could be properly<br />

applied to the beam passing through the two gaps. In this way, the frequency of the<br />

bunched beam can be varied between 2 and 4 MHz (i.e. variable time interval from 250<br />

ns to 500 ns).<br />

The beam test of the pre-buncher at a 2-MHz operation was performed using 16 O 4+<br />

and 12 C 3+ beams supplied by <strong>KEK</strong>CB. The time structures of the beam observed with<br />

buncher-on and off are compared in Fig. 2-30. At the buncher-off operation was<br />

indentified 13 bunches of the beam by the acceleration frequency of 26-MHz SCRFQ<br />

during the time interval of the frequency of 2-MHz (500 ns) beam-bunching (blue line<br />

in Fig. 2-30). When the pre-buncher was on, a clear bunch structure (bunched into the<br />

middle of 13 bunches with an interval of 500 ns) was observed at a target position 10 m<br />

away from the exit of the <strong>TRIAC</strong> (red line in Fig. 2-30), although there exist a small<br />

fraction of the beam failed to be bunched (i.e. observed as a side bunch instead of the<br />

main, central bunch).<br />

The bunching efficiency, defined as the intensity ratio of a bunch of beam (central<br />

bunch in Fig. 2-30 when the pre-buncher is on) to the sum of 13 bunches with<br />

buncher-off, was found to be 42 %. Incomplete beam bunching, partly observed as a<br />

remainder of the 26-MHz operation of SCRFQ in Fig. 2-30 with buncher-on, is due to<br />

36


the imperfect saw-tooth wave-form for beam-bunching. A fraction of the beam, 47 % in<br />

the present observation, was lost during acceleration and not observed in Fig. 2-30. The<br />

remaining fraction of 11% was properly accelerated but observed in the side bunches as<br />

shown in Fig. 2-30. The present observations are well described by the beam simulation<br />

code based on the TRACEP [2-15].<br />

Fig. 2-30. Beam bunch structure with buncher-on (red line) and buncher-off (SCRFQ time structure)<br />

(blue line).<br />

Fig. 2-31. Front view of the multi-layer beam chopper on the beam direction<br />

37


For the proposed experiment, the portion of beam in the side bunches, especially<br />

during the time of 200 ns ahead of the main bunch of beam, should be less than 10 -4<br />

relative to the intensity of the main bunch. Such contribution can be removed by the<br />

operation of the chopper installed upstream of the pre-buncher, avoiding beam injection<br />

into the buncher during the shaping time of the saw-tooth wave-form for beam bunching.<br />

The chopper consists of 19 electrodes of thin plates, which are vertically piled on top of<br />

each other against the beam injection at regular intervals of 1.9 mm, as shown in Fig.<br />

2-31. The plate is made of phosphor bronze, which is 10-mm long along the beam<br />

direction, 40-mm wide in the horizontal direction and 0.1-mm thick in the vertical<br />

direction. The beam can be deflected upwards and downwards by applying an electric<br />

potential to every two plates; every plate on the ground potential is sandwiched between<br />

two plates on an electric potential. Time profile of the applied voltage is shown by red<br />

line in Fig. 2-32. It should be noted that the fast-rising and -falling wave-form of the<br />

applied potential (around 50 ns) is essential to keep the bunching efficiency high while<br />

significantly reducing the portion of beam arriving prior to the main bunch of beam.<br />

Thanks to the multi-layer structure, the applied voltage required for proper<br />

deflection of beam becomes as low as about 120 V and the leakage of the electric field<br />

along the beam direction is minimized, while about 14% of beam is lost by hitting the<br />

electrodes. The beam chopped by the applied potential was injected into the SCRFQ and<br />

turned out to be well accepted for acceleration; only during the time duration of<br />

potential-off, the accelerated beam was observed, as shown in Fig. 2-32.<br />

Fig. 2-32. Time profile of the potential applied to the chopper electrodes (red line by left vertical<br />

axis), compared with the time structure of chopped beam after SCRFQ acceleration without beam<br />

bunching beam (blue line by right vertical axis).<br />

38


Fig. 2-33. Time structure of the bunched beam with the operation of the chopper (black) and without<br />

the operation (grey)<br />

The time structures of the bunched beam with chopper-off and -on are compared in<br />

Fig. 2-33. There are not observed the beam for 250 ns prior to the central bunch,<br />

indicating that the suppression of beam in the time range is better than 5x10 -5 . The<br />

bunching system has been recently applied for the experiment.<br />

2-5. Beam transport line<br />

The beam transport line at the <strong>TRIAC</strong> can be divided into three sections, a<br />

low-energy beam transport line (LBT: before acceleration), a beam transport line for the<br />

SCRFQ / IH linac (ACC: during acceleration) and a high-energy beam transport line<br />

(FBT: after acceleration).<br />

2-5-1. Low energy beam line (LBT)<br />

Ion optical configuration of the JAEA-ISOL, located upstream of the LBT, is<br />

EQ-EQ-MD-ED, where EQ, MD, and ED stand for electric quadrupole, magnetic dipole,<br />

and electric dipole, respectively. The mass resolving power of the JAEA-ISOL is M/∆M<br />

= 1200.<br />

After mass-separated by the ISOL, the singly charged RIB is transported to the<br />

<strong>KEK</strong>CB by the first part of LBT whose optical configuration is shown in Fig. 2-34. The<br />

designed acceptance is 200π mm•mrad. Two beam profile monitors, BPM-LBT01 and<br />

BPM-LBT02 as shown in Fig. 2-34, are installed for stable ion beams. Formed in a<br />

crossed wire and driven by a stepping motor, the monitor can probe simultaneously the<br />

vertical and horizontal components of the beam profiles, when crossing the beam along<br />

39


a diagonal direction. The monitors are followed by Faraday cups, FC-LBT01 and<br />

FC-LBT02, respectively. For monitoring the intensity of RIB just upstream of the<br />

<strong>KEK</strong>CB, a plastic scintillator (2 x 2 x 2 cm 3 ), SCI-LBT01 in Fig. 2-34, is installed in<br />

the injection side of the <strong>KEK</strong>CB. It was covered by an aluminum plate of 100-µm thick<br />

where the RIB was implanted. The detection efficiency for β-rays from the implanted<br />

RIBs is about 50 %.<br />

After being charge-bred, the RIB with a higher charge state is transported through<br />

the second part of LBT to ACC. The ion-optical configuration between the <strong>KEK</strong>CB and<br />

ACC is also shown in Fig. 2-34. The first component of 2MQ-MD-2MQ is dedicated to<br />

charge state analysis of ions extracted from the <strong>KEK</strong>CB. The resolving power of<br />

mass-to-charge state ratio (A/q) is designed to be (A/q)/δ(A/q) = 128 for the beams with<br />

emittance of 100π mm•mrad. Three Faraday cups (FC-LBTFi, i = 1, 2, 3) and three<br />

plastic scintillators (SCI-LBTFi, i = 1, 2, 3) are installed at F1, F2 and F3 chambers. A<br />

beam profile monitor, BPM-LBT04, is installed between F1 and F2 chambers. The<br />

designed acceptance of the second part of LBT is also 200π mm•mrad. Figure 2-35<br />

shows the result of ion optical simulation with an initial emittance of 200π mm•mrad by<br />

using the GIOS code.<br />

Fig. 2-34 .Ion transporting optical components and beam monitors for low energy beam transport.<br />

40


upstream of the target position not only for picking up timing information of beam<br />

particles, one by one, but also for monitoring beam intensity during experiments.<br />

Table 2-7 shows the present transport efficiencies of the respective transport lines<br />

for 16 O 2+ and 129 Xe 19+ beams. The transport efficiency from ISOL to CB is almost<br />

100 % for various ion beams.<br />

2-5-3. Transport for charge-bred ions<br />

In transporting charge-bred ions at the <strong>TRIAC</strong>, we have found that the transport<br />

efficiencies for highly charged ions are significantly reduced. This is mostly due to the<br />

charge-exchange collisions of highly charged ions during the beam transportation<br />

through the 15-m long low-energy beam line. Such collisions with residual gases in<br />

flight tend to transform the highly charged ions to lower charge states by electron<br />

captures.<br />

Fig. 2-37. Fraction of charge states measured at F3 for Xe 19+<br />

In order to reduce the electron capture rates of highly charged ion beams in flight as<br />

much as possible, higher vacuum in the low-energy beam line is necessary. To obtain<br />

better vacuum, we installed eight cryopumps along the beam lines, 4 for LBT and<br />

another 4 for SCRFQ. The averaged vacuum pressure in LBT was improved from<br />

2×10 -4 Pa to 1.5×10 -5 Pa after the installation. The charge state distributions of Xe 19+<br />

after its flight with 2 keV/nucleon from F1 to F3 (see Fig. 2-34) are compared under<br />

different conditions of vacuum in Fig. 2-37. In this measurement, we selected the initial<br />

42


charge state of xenon ions by using the analyzing magnet for the <strong>KEK</strong>CB. As shown in<br />

Fig. 2-37, the escape fraction to lower charge states reduced from 70% to 20% in the<br />

case of the 129 Xe 19+ .<br />

Table 2-7. Beam transport efficiencies on each transport line.<br />

LBT (CB -> ACC) ACC FBT<br />

16 O 2+ 95 % 89 % 100 %<br />

129 Xe 19+ 85 % 73 % 65 %<br />

2-6. Control system<br />

In order to operate the <strong>TRIAC</strong>, we have three control rooms which are dedicated to<br />

the control of the devices for low-energy beam, accelerated beam and experimental<br />

measurement, respectively. The equipments for low-energy beam consist of ISOL,<br />

<strong>KEK</strong>CB, low-energy beam transportation line (LBT). Those for accelerated beam<br />

consist of SCRFQ and IH linear accelerators (ACC.), and fast beam transportation line<br />

(FBT). It is desirable to control almost all devices at experimental measurement room,<br />

in order to transport stable and/or radioactive beams to user's experimental setup. Thus<br />

we constructed a remote control system which connects the three control rooms through<br />

a network.<br />

Figure 2-38 shows the diagram of the control system. The interface programs of<br />

control system are loaded in LabView (National Instruments Co. Ltd.) which is installed<br />

in the personal computers (PC, Windows XP) in each control room. These PCs connect<br />

via Ethernet to six board PCs, seven Field Point (FP) network modules, four GPIB<br />

E-NET100 modules (National Instruments Co. Ltd.) and one programmable logic<br />

controller (PLC) locally placed in the experimental hall and the power room. These<br />

board PCs and FPs control some digital to analog converters (DAC) and motion-driving<br />

modules which access to HV power and current suppliers for electromagnetic deflectors<br />

and lenses, and motor drivers for some kind of beam monitors and RF tuners for<br />

accelerators. The GPIB E-NET100 controls the devices which read the status of vacuum<br />

gauges, magnetic field of deflectors and beam current of Faraday cups. The interlock<br />

systems of beam and RF power generation are also controlled by FP modules. The<br />

<strong>KEK</strong>CB control system belongs to another network group. A PC connects to the<br />

sequencer which access to the devices for <strong>KEK</strong>CB operation. The ion source,<br />

mass-separator and electric lenses for the JAEA-ISOL are operated independently at the<br />

console. These elements are also controllable in any rooms by an additional PC having<br />

DAC modules.<br />

43


a thin film. Figure 2-39 shows a schematic view of the detector. It is composed of a thin<br />

metalized Formvar film (~0.2 µm), an acceleration electrode, a drift space, a mirror<br />

electrode, a two-layered MCP and a two-dimensional position sensitive board. When a<br />

beam particle passes through the entrance film, secondary electrons are ejected, and<br />

they are accelerated and reflected by high voltages applied to mirror electrodes. After<br />

being multiplied by the MCP, they are captured on a readout board (2-dimensional<br />

position sensitive board in Fig. 2-39) where the signals for position and timing<br />

information are induced. The right panel of Fig. 2-39 shows the tracks of the secondary<br />

electrons calculated under the electric potential generated by a high voltage<br />

configuration, HV1 = -5500 V, HV2 = -4000 V, HV3 = -2300 V and HV4 = -100 V. It is<br />

shown that positions of electrons ejected at the thin film are projected on the readout<br />

board, well keeping the primary information on their emission positions. The<br />

acceleration electrode and mirror electrode are composed of gold-tungsten wires of 20-<br />

µm diameter stretched at intervals of 1 mm.<br />

2-dimensional position<br />

sensitive board<br />

MCP<br />

thin film<br />

beam<br />

acceleration<br />

electrode<br />

e −<br />

mirror electrode<br />

signal<br />

HV4<br />

HV3<br />

HV2<br />

HV1<br />

Fig. 2-39. Schematic view of the beam diagnostic detector (left) and secondary electron tracks<br />

under the electric potential distribution configured by high voltages that are respectively applied<br />

to the electrodes (e.g. HV1 = -5500 V, HV2 = -4000 V, HV3 = -2300 V and HV4 = -100 V).<br />

The two-dimensional position sensitive board has an area of 5 x 5 cm 2 . Figure 2-40<br />

shows a part of the pattern of the readout electrodes on the board. For position<br />

measurements, there are two kinds of electrodes, x-cells and y-cells. All x-cells are<br />

connected to each other by delay lines in a row, and the time difference between two<br />

signals from both ends of the row gives the position hitted by secondary electrons along<br />

the x-axis as defined by dotted lines in Fig. 2-40. In the same way, another position (y)<br />

can be read by the y-cells. Since the total delay time in the row is 100 ns, the sum of the<br />

delay times of two signals from both ends of the row (relative to the timing by a striped<br />

45<br />

5 cm


electrode as shown in Fig. 2-40) should be that value. Applying such criterion on the<br />

detected events, we can identify real events among them, which include some<br />

unintentional events induced by noises. Rejecting those unlike events can be applied up<br />

to the beam intensity of a few tens of Mpps, i.e. the intensity of ~10 7 pps is the<br />

maximum beam rates countable by the present MCP. For higher beam rates, such event<br />

selection is not applicable because of multiple hitting. The timing-readout strip<br />

electrodes are connected directly to each other, and give a reference time signal for the<br />

starting point of delay-line signals.<br />

timing-readout<br />

x-readout<br />

Fig. 2-40. Pattern of readout electrodes on the two-dimensional position sensitive board.<br />

46<br />

y-readout<br />

(a) (b)<br />

1 mm<br />

x-cell<br />

y-cell<br />

Fig. 2-41. Position distributions for alpha particles measured with two different kinds of<br />

masks: (a) a hole of the diameter 1.8 mm and (b) a diagonal slit of width 1 mm.


Using the alpha particles from a 241 Am source, position resolutions of the detector<br />

were measured. Figure 2-41 shows the measured position distributions of alpha particles<br />

with two different kinds of masks put in front of the thin film; (a) a hole of 1.8 mm in<br />

diameter and (b) a diagonal slit of 1 mm in width. Fitting the distributions with<br />

Gaussian ones, the position resolution was deduced to be 6.5 mm in full width half<br />

maximum (FWHM). The present resolution is considered to be largely limited by the<br />

small number of electrons primarily produced on the thin foil by the impact of alpha<br />

particles. For heavier ions, the resolution could be improved. A simplified version of the<br />

presently developed detector, where is used a collector of the secondary electrons<br />

multiplied by MCP instead of the position sensitive board, has been used as a beam<br />

counter (i.e. intensity monitor and time pick-up detector) providing beam trigger signal<br />

at <strong>TRIAC</strong> experiments to be described in the section 3.<br />

2-7-2. Development of an active target<br />

For the experiments using RIBs whose intensities are at most about 10 5 pps at the<br />

present <strong>TRIAC</strong>, it is very efficient to employ a gas counter (e.g. Multiple Sampling<br />

Tracking Proportional Chamber: MSTPC) as an active target [2-24], where the<br />

operating gas for the counter played simultaneously the role of gas target for the nuclear<br />

reaction of interest. Using such a detector, the detection efficiency for reaction products<br />

of charged particles could be in principle100 % and the excitation function could be<br />

measured at a fixed incident-energy of RIB. However, the gas-gain instability in the<br />

MSTPC was observed under an injection rate higher than ~10 3 pps due to space-charge<br />

limitation around anode wires of the MSTPC. We have developed a new type of the<br />

active target working at a beam injection rate as high as 10 5-6 pps.<br />

For higher injection-rate capability, a thick foil of gas electron multiplier<br />

(THGEM) with a thickness of 400 µm [2-25] was used to overcome the gain instability<br />

of a multi-wire proportional counter. We examined the performance of various kinds of<br />

THGEMs by using low-energy heavy-ion beam from the <strong>TRIAC</strong>. For the investigation,<br />

we made a new counter called as GEM-MSTPC, which consists of a drift space (active<br />

volume of 100 3 mm 3 ), one (or more) THGEM, and an anode plate. The anode plate (Pad<br />

in Fig. 2-42) is composed of backgammon-type electrodes (24 segmented pads) to<br />

measure trajectories of charged particles in the horizontal plane against the vertical-drift<br />

space. The drift times of electrons primarily produced from the position of a particle’s<br />

track are measured, giving the position of the particle trajectory along the drift direction<br />

of electrons (i.e. vertical position of the particle track). Figure 2-42 presents a<br />

photograph of the GEM-MSTPC. The typical operating gas and its pressure are He +<br />

47


CO2 (10 %) and 16 kPa, respectively, for the measurement of the 8 Li(α, n) 11 B reaction<br />

cross sections. The gas gain on the THGEM was measured under the condition of the<br />

operating gas. The measured gas gain was high enough (~10 3 ) in the single THGEM<br />

configuration [2-26, -27].<br />

Fig. 2-42. (Left) The photograph of the GEM-MSTPC. The drift space has two electrodes and is<br />

enclosed by field-guiding wires (100 x 100 x 100 mm 3 ); ‘Top’ is a plate electrode and ‘Bottom’ is an<br />

electrode which supports insulating pillars. ‘Pad’ indicates an anode plate, segmented for having<br />

position sensitivity. (Right) The (vertical) configuration of the electrodes in the GEM-MSTPC is<br />

shown, where the beam comes into the middle of the drift space (across the position of 50mm among<br />

the vertical length of the 100 mm.)<br />

When the GEM is incorporated in a 3-dimensional tracking detector, it is known<br />

that the vertical position of the particle track is distorted by ion feedback [2-27]; the<br />

electron drift time is disturbed due to the electric field induced by numerous ions<br />

drifting away from the proportional region in gas amplification around the GEM into<br />

the drift space. In order to suppress the ion feedback from the GEM foil to the drift<br />

space, the double GEM configuration was finally adopted as shown in the right panel of<br />

Fig. 2-42. Such a configuration helps reduce the operational gas-gain on the one GEM,<br />

subsequently reducing the number of ions that back to the drift space.<br />

We measured the time dependence of pulse height of anode output by injecting 4 He<br />

from a radioactive source of 241 Am or the 12 C beam of 13.2 MeV provided by the<br />

<strong>TRIAC</strong>. We observed shifts in the pulse-height with time, as reported in the previously<br />

developed THGEM [2-25]. This phenomenon may come from the gas-gain instability<br />

due to the electrical charge-up on the insulator between the GEM electrodes. It strongly<br />

depends on the structure of holes punching through the electrodes. In order to avoid the<br />

48


charge-up, we modified the structure of the holes and the surface of the electrode.<br />

Table 2-8 shows a summary of such modifications together with observed<br />

pulse-height shifts under various injection rates of particles with the energy of about 1<br />

MeV/nucleon. THGEM #1 indicates the THGEM previously developed in Ref. 2-25,<br />

which has an insulator rim of 100 µm in width around a hole. THGEM #2 has no rim<br />

and the gain shift was suppressed to be less than 12 % under the 4 He injection rate of<br />

10 4 pps. However, a discharge often occurred in several hours from the beginning of the<br />

injection, and the gain became unstable once the discharge occurred. In order to avoid<br />

the discharge, the diameter of the hole of THGEM#3 was reduced to be smaller than of<br />

THGEM#2.<br />

Table 2-8. Summary of modified THGEMs together with pulse height shifts at various injection<br />

rates. THGEM#1 indicates the previously developed THGEM [2-25]. THGEM#2 has no rim on the<br />

hole of 500 µm in diameter. The diameter of the hole of THGEM#3 was smaller than for<br />

THGEM#2. The surface of copper electrodes was coated with gold for THGEM#4. ‘n/a’ indicates<br />

not-measured.<br />

Pulse<br />

height<br />

THGEM# #1 #2 #3 #4<br />

Hole diameter [µm] 300 500 300 300<br />

Rim yes no no no<br />

Hole pitch [µm] 700 700 700 700<br />

Electrode surface Au Cu Cu Au<br />

4 He (E = 5.4 MeV, 10 0 pps ) 100 a few n/a n/a<br />

4 He (E = 5.4MeV, 10 4 pps ) n/a 12 1~5 n/a<br />

shift (%) 12 C (E = 13.2 MeV, 10 5 pps) n/a n/a 8 < 3<br />

Fig. 2-43. Pulse-height shifts under the 12 C-beam injection rate from 400 pps to 10 5 pps using<br />

THGEM #3 (left panel) and THGEM #4 (right panel). The horizontal axis indicates time from the<br />

beginning of the beam injection.<br />

49


Although the gain shift of THGEM#3 was suppressed to be within 5 % under the<br />

same 4 He injection rate, it became larger (about 8 %) in the case of the 12 C injection rate<br />

of 10 5 pps as shown in the left panel of Fig. 2-43. Moreover, the gain shift of<br />

THGEM#3 gradually increased in several consecutive measurements with the 4 He<br />

injection of 10 4 pps. Actually, the shift was about 1 % at the first measurement, whereas<br />

it changed to be 5 % at the third measurement. This phenomenon could be explained as<br />

the ‘ageing’ of the electrodes, such as oxidation of the surface of the copper electrode.<br />

Hence, we coated the electrode of THGEM#4 with gold. THGEM #4 showed good gain<br />

stability as shown in the right panel of Fig. 2-43. The pulse height became stable within<br />

3 % under the 12 C injection rate from 400 to 1.2 x 10 5 pps.<br />

The energy resolution under the 12 C injection rate of 10 5 pps was also measured. It<br />

was obtained from the distribution of energy-loss signals from the segmented anode<br />

pads. The energy resolutions under the injection rate of 400 pps of the two THGEMs of<br />

#3 and #4 were almost similar (7% in σ). However, with the higher injection rate, e.g.<br />

10 5 pps, THGEM #4 gave better energy resolution than THGEM #3; 8 % vs. 13 %.<br />

The observed gain stability and energy resolution of THGEM #4 under the<br />

injection rate of 10 5 pps satisfy experimental requirements. The GEM-MSTPC using<br />

THGEM #4 (400-µm thick, 300-µm hole-diameter, Cu electrode coated by Au without<br />

rim) was successfully employed to measure the 8 Li(α, n) 11 B reaction cross sections<br />

(RNB-K05).<br />

References<br />

[2-1] S. Ichikawa et al., Nucl. Instr. and Meth. B204 (2003) 372.<br />

[2-2] R. Kirchner et al., Nucl. Instr. and Meth. 186 (1981) 295.<br />

[2-3] A. Osa et al., Nucl. Instr. and Meth. B266 (2008) 4373.<br />

[2-4] Y. Otokawa et al., Rev. Sci. Instrum. 81 (2010) 02A902.<br />

[2-5] D. Hitz et al., Rev. Sci. Instrum. 73 (2002) 509.<br />

[2-6] N. Angert et al., Proc. of the 14th International Workshop on ECR sources,<br />

CERN, Geneva, Switzerland, 1999, p220.<br />

[2-7] M. Oyaizu et al., Rev. Sci. Instrum. 73 (2002) 806.<br />

[2-8] N. Imai et al., Rev. Sci. Instrum. 79 (2008) 02A906.<br />

[2-9] S.C. Jeong et al., Rev. Sci. Instrum. 75 (2004) 1631.<br />

[2-10] M. Oyaizu et al., AIP CP1120 (2009) 308.<br />

[2-11] S.C. Jeong et al., Rev. Sci. Instrum. 73 (2002) 803.<br />

[2-12] S. Arai et al., Nucl. Instr. and Meth. A390 (1997) 9.<br />

[2-13] S. Arai et al., <strong>KEK</strong> preprint 98-99, 1998.<br />

50


[2-14] S. Arai et al., <strong>KEK</strong> <strong>Report</strong> 2008-8, 2008.<br />

[2-15] S. Yamada, ‘TRACEP’, NIRS HIMAC program manual No. 1, May 1990,<br />

Private communication.<br />

[2-16] S. Arai et al., <strong>KEK</strong> <strong>Report</strong> 2008-1, 2008.<br />

[2-17] M. J. Lee et al., ‘MAGNET INSERTION CODE’, NAL TM-447, Oct. 1973.<br />

[2-18] H. Kabumoto et al., Nucl. Instr. and Meth. A612 (2010) 221.<br />

[2-19] K.W. Shepard, C.H. Scheibelhut, R. Benaroya and L.M. Bollinger, IEEE Trans.<br />

Nucl. Sci. NS-24(3) (1977) 1147.<br />

[2-20] K. Saito and P. Kneisel, Proc. of the 3rd EPAC, Berlin, Germany, (1992) 1231.<br />

[2-21] P. Kneisel, B.Lewis and L. Turlington., Proc. of the 6th Workshop on RF<br />

Superconductivity, CEBAF, (1993) 628.<br />

[2-22] K. Niki et al., Proc. of the 6th Part. Accel. Soc. Meeting, 2009, Tokai, Japan.<br />

[2-23] M. Okada et al., Proc. of the 6th Part. Accel. Soc. Meeting, 2009, Tokai, Japan.<br />

[2-24] T. Hashimoto et al., Nucl. Instrum. Methods A 556 (2006) 239.<br />

[2-25] C. Shalem et al., Nucl. Instrum. Methods A 558 (2006) 475.<br />

[2-26] K. Yamaguchi et al., Nucl. Instrum. Methods A 623 (2010) 135.<br />

[2-27] S.K. Das et al., Nucl. Instrum. Methods A 625 (2010) 39.<br />

51


3. Experiments at the <strong>TRIAC</strong><br />

Since the first public operation of the <strong>TRIAC</strong>, 50 days of operation per year, in<br />

maximum, has been allowed for experiments, including the operation for 10 days for the<br />

development of the RIB generation and acceleration at the <strong>TRIAC</strong>. The experimental<br />

proposals were carefully reviewed, once a year, by the program advisory committee<br />

(PAC). As listed in Table 3-1, 15 proposals have been approved and, among them, 13<br />

proposals have been already performed. Two remaining proposals (KJ02 and KJ03)<br />

have been already scheduled to be performed soon. There were 3 additional proposals<br />

only using ISOL beams without reacceleration by post-accelerators.<br />

The approved proposals are grouped according to relevant research field and<br />

respectively summarized as followings: nuclear astrophysics (RNB-K01, K05, J01,<br />

KJ03 and KJ04), nuclear and fundamental physics (RNB-K02/03, K04, and K08), and<br />

materials science (RNB-K06, K07, J02/03, KJ01/02). The experiments using RIBs<br />

without re-acceleration are also briefly described at the end.<br />

3-1. Nuclear astrophysics<br />

In recent decades, nuclear reactions involving the unstable nucleus 8 Li have<br />

attracted considerable attention among nuclear physicists since they are considered to<br />

influence largely the synthesis of heavy elements in neutron-rich high-temperature<br />

environments. Such environments in the temperature of T9 = 1-3 (in unit of 10 9 K) could<br />

happen, for example, in the early stage of supernova explosion and/or in the early<br />

universe of inhomogeneous Big-Bang model. Continuous efforts have been made to<br />

experimentally determine the relevant reaction rates at energies below the Coulomb<br />

barrier. The low-energy 8 Li beam available at the <strong>TRIAC</strong> provides unique opportunities<br />

to measure directly the relevant cross sections in the energy regime of astrophysical<br />

interest.<br />

The RNB-K01 (Spokesperson: H. Ishiyama (<strong>KEK</strong>) and T. Hashimoto (JAEA),<br />

Collaboration (Institution (number of collaborators)): <strong>KEK</strong> (9), JAEA (8), Osaka<br />

Electro-comm. Univ. (3), Osaka Univ. (2), GSI (1), NAO (1)) was the first attempt to<br />

measure the cross sections of 8 Li(d, t) 7 Li reaction, which has been thought to play an<br />

important role in destroying 8 Li nuclei and therefore limiting the synthesis of elements<br />

over the mass of 8. By using the 8 Li beam from the <strong>TRIAC</strong>, the measurements were<br />

made at seven different energies of 0.18, 0.24, 0.30, 0.40, 0.46, 0.60, and 0.75<br />

MeV/nucleon, respectively corresponding to center-of-mass energies (Ecm) of 0.3, 0.4,<br />

0.5, 0.7, 0.8, 1.0, and 1.1 MeV. Those Ecm cover the Gamow peaks of T9 = 1-3.<br />

52


Table 3-1. List of experiments performed at the <strong>TRIAC</strong>. Approved experiments are identified by the<br />

subject numbers assigned; RNB-Kxx and RNB-Jyy indicate the subject number approved by the<br />

<strong>KEK</strong>-PAC and the JAEA-PAC, respectively. Additional experiments numbered by RNB-KJxx are<br />

approved recently by a joint PAC.<br />

Subject No. Title of subject Spokesperson(Affiliation)<br />

RNB-K01 Direct measurement of astrophysical reaction rates through 8 Li H. Ishiyama (<strong>KEK</strong>)<br />

RNB-K02/K03 Nuclear spectroscopy through β-delayed decay of spin-polarized nuclei<br />

around doubly magic nucleus 132 Sn<br />

The typical intensity was 1x10 5 particles per second (pps), energy resolution 0.85 % in<br />

standard deviation (σ), and diameter of beam spot at the target position 3.4 mm in σ.<br />

A deuteron target (deuterated polyethylene (CD2)) with a thickness of 130 µg/cm 2<br />

was used. The detection system employed in the measurements is shown in Fig. 3-1.<br />

Figure 3-2 shows the measured cross sections together with previous results [3-1].<br />

The present experiment provides the first data in the low-energy region below Ecm = 1.5<br />

MeV and covers the Gamow peaks in the temperature range of T9 = 1-3. An<br />

enhancement was observed at around Ecm = 0.9 MeV in the excitation function, where<br />

the cross sections are significantly larger than at the higher Ecm.<br />

Figure 3-3 shows the resultant reaction rate, which is compared with the rate<br />

estimated by the extrapolation of data at higher energies [3-1].<br />

53<br />

Y. Hirayama (<strong>KEK</strong>)<br />

K. Matsuta (Osaka Univ.)<br />

RNB-K04 R&D for a test of time reversal symmetry using polarized nuclei J. Murata (Rikkyo Univ.)<br />

RNB-K05 Measurement of the 8 Li(α, n) 11 B reaction cross section for astrophysical<br />

interest<br />

T. Fukuda (Osaka<br />

Electro-Com.Univ.)<br />

RNB-K06 Local fields at 111 Cd implanted in Highly Oriented Pyrolytic Graphite W. Sato (Osaka Univ.)<br />

RNB-K07 Diffusion Coefficient Measurements on Perovskite-type Structured<br />

Lithium Ion Conductor<br />

S. Takai<br />

(Tottori Univ.)<br />

RNB-K08 Search for highly excited states of 11 Be via 9 Li+d reaction T. Teranishi (Kyushyu Univ.)<br />

RNB-J01 Search for highly excited states in 10 Be using deuteron elastic reaction to<br />

RNB-J02/J03<br />

RNB-KJ01<br />

8 Li<br />

Diffusion studies in Li ionic conductors by using short-lived radioactive<br />

beam of 8 Li<br />

RNB-KJ02 In-situ measurements of Li diffusion in Li micro-batteries & Nano-scale<br />

diffusion experiments in Li ionic conductors by using 8 Li<br />

H. Miyatake<br />

(JAEA/<strong>KEK</strong>)<br />

H. Sugai (JAEA)<br />

S.C. Jeong (<strong>KEK</strong>)<br />

H. Ishiyama (<strong>KEK</strong>)<br />

S.C. Jeong (<strong>KEK</strong>)<br />

RNB-KJ03 “Extremely” safe Coulomb excitation of 142 Ba N. Imai (<strong>KEK</strong>)<br />

RNB-KJ04 Measurement of the cross section of 12 C (a,γ) reaction H. Makii (JAEA)


Fig. 3-1. Schematic view of the detection system for studying the 8 Li(d, t) 7 Li reaction. The<br />

recoiled tritons were measured by using the ∆E-E telescope (43-µm-thick double-sided Si detector<br />

and 375-µm-thick Si detector with a size of 50 x 50 mm 2 ), SSD1 and SSD2 (375-µm-thick Si<br />

detector with a size of 18 x 18 mm 2 ). The scattered 7 Li particles were measured by SSSD1 and<br />

SSSD2 (68-µm-thick single-sided Si detector a size of 50 x 50 mm 2 ) in coincidence with SSD1 and<br />

SSD2. The beam intensity was monitored and its amount was measured by the plastic scintillator<br />

installed just downstream of the Al plate, on which the beam particles were implanted after passing<br />

through the target.<br />

Fig. 3-2. Excitation function of cross section of the 8 Li(d, t) 7 Li reaction. The red circles show the<br />

results of the present study. The open triangles show th results of the previous study [3-1]. Energy<br />

regions of astrophysical interest are indicated by temperatures corresponding to the Gamow peaks<br />

(red arrows).<br />

54


The reaction rate deduced from the data shown in Fig. 3-2 is faster by about one order<br />

of magnitude at about T9 = 1 than the previously reported values. The fast reaction rate<br />

is due to the resonance like structure in the excitation function of the reaction observed<br />

at around Ecm = 0.9 MeV. The reaction rate of the 8 Li(d, t) 7 Li reaction is 3x10 3 cm 3 /s,<br />

which is 45 times larger than those of the 8 Li(n, γ) 9 Li [3-2] and 8 Li(α, n) 11 B [3-3]<br />

reactions at around T9 = 1.<br />

Fig. 3-3. Reaction rate of the 8 Li(d, t) 7 Li reaction as a function of temperature . The red line<br />

indicates the present result and the blue line represents the result of Balbes et al. [3-1].<br />

The present result indicates that the 8 Li(d, t) 7 Li reaction could suppress the<br />

synthesis of heavy elements through 8 Li when the number density of deuteron (Yd) is<br />

comparable to those of neutron (Yn) and/or α particle (Yα). However, in the process of<br />

nucleosynthesis in type II supernovae presently considered as a candidate of<br />

astronomical sites for synthesizing heavy elements by rapid neutron captures (r-process),<br />

the suppression would be negligible since the estimated Yd/Yn and Yd/Yα values are<br />

very small at T9 =1 (about 10 -6 and 10 -8 , respectively). Figure 3-4 shows an example of<br />

such considerations [3-5], where the relative reaction rates (YxY8Li) were<br />

calculated and compared for several 8 Li-involved reactions. In the calculations, the<br />

fraction of light elements, Yx (i.e. proton (Yp), deuteron (Yd), neutron (Yn) and alpha<br />

particle (Yα)) were deduced by a network calculation for the r-process using the<br />

neutrino-driven wind model in the Type II supernovae [3-6]. The environmental initial<br />

parameters (i.e. electron fraction, entropy, etc.) in the model were selected to reproduce<br />

55


-element abundances up to the third (A ~ 195) peak. For easy comparison, the fraction<br />

of 8 Li, Y8Li was assumed to be unity. As can be seen in Fig. 3-4, at T9 > 3.7, the 8 Li(p,<br />

α)αn reaction is the fastest reaction, which destroys the 8 Li. In T9 = 0.7 – 3.7, the 8 Li(α,<br />

n) 11 B reaction becomes the fastest one, which leads to the synthesis of heavier elements<br />

than 8 Li. However, The relative rate of 8 Li(d, t) 7 Li is negligibly small as compared to<br />

that of 8 Li(α, n) 11 B, and therefore the 8 Li(d, t) 7 Li gives little effect to the r-element<br />

abundances in this model, although the reaction rate is relatively high. Nevertheless, our<br />

present reaction rate of the 8 Li(d, t) 7 Li reaction should be included in the network<br />

calculation for more accurate estimate of r-element abundances especially when using<br />

other competitive astrophysical models for the calculations, since the astrophysical site<br />

for the r-process is still controversial.<br />

Fig. 3-4. Relative reaction rates (YxY8Li) of several 8 Li-involved nuclear reactions. The Yx is<br />

the fraction of light element, x (x = n, p, d, α), and the Y8Li is the fraction of 8 Li. The reaction rates<br />

of the 8 Li(d, t) 7 Li reaction estimated by using our present results were adopted, as shown in Fig. 3-4.<br />

The rates of 8 Li(α, n) 11 B, 8 Li(p, α)αn, 8 Li(n, γ) 9 Li are from Ref. [3-3], [3-7], and [3-2], respectively.<br />

The RNB-J01 (Spokesperson: H. Miyatake (JAEA/<strong>KEK</strong>), Collaboration: <strong>KEK</strong> (6),<br />

JAEA (5), Kyushu Univ. (1), Tsukuba Univ.(1), Ibaraki Univ. (1), Osaka Univ. (2),<br />

Osaka Electro-Comm. Univ. (2), CNS (1)) was dedicated to study the nuclear structure<br />

around the resonant-like peak in the excitation function of 8 Li(d, t) reactions around Ecm<br />

= 1.0 MeV previously observed at the RNB-K01 (as shown in Fig. 3-2). The energy<br />

56


corresponds to an excited state of 22.4 MeV in 10 Be, which was also reported as a<br />

proton/triton decaying state both in the neutron capture experiment of 9 Be and in the<br />

resonant particle decay spectroscopy of 7 Li + 7 Li. It has been recently suggested that the<br />

state might be a candidate of a coexisting cluster state with cluster structures of (α+t+t)<br />

and (α+α+n+n) [3-8], which are of great importance not only for understanding the<br />

clustering phenomena in neutron rich nuclei but also for an accurate estimate of the<br />

reaction rate linked with such a state in the process of nucleosynthesis. The aim of this<br />

experiment is to determine the spin and parity of this state, by using the thick-target<br />

resonance elastic scattering of deuteron by 8 Li as shown in Fig. 3-5.<br />

Fig. 3-5. Experimental set-up and principle of the thick-target elastic scattering method for<br />

resonance-searching.<br />

Following the PAC recommendation, a feasibility test for this measurement has<br />

been performed for 3 days. A schematic figure for the experimental set-up and the<br />

measurement principle is shown in Fig. 3-5. An elastic scattering with thick target was<br />

employed in the inverse kinematics to search for the resonance states. The 8 Li beam of<br />

0.94 MeV/nucleon with an intensity of 4.5 kpps passed through a time pick-up detector<br />

(C-foil + Micro-Channel Plate (MCP) in Fig. 3-5) bombarded to the CD2 target of 2.6<br />

mg/cm 2 in thickness. The energy of the scattered particles was measured by a set of<br />

Solid State Detectors (SSD), and their time-of flight (TOF) between the MCP and SSD<br />

was also measured. As shown in Fig. 3-6, where the correlations between the energy<br />

and the TOF for detected particles were plotted, scattered deuterons were successfully<br />

identified. In the figure were also observed tritons, 8 Li particles passed through<br />

pin-holes on the target, and α-particles from decaying 8 Li, in addition to deuterons.<br />

These particles were well resolved. Based on the present result, we can conclude that<br />

the thick-target elastic scattering method for resonance-searching is feasible with the<br />

beam intensity of 8 Li presently available at <strong>TRIAC</strong> (i.e. higher than 100 kpps). For<br />

57


higher reliability of the experimental method, it would be desirable that the present<br />

detection system should be extended to cover larger angles; it allows better statistics and<br />

more experimental information at larger angles, if any.<br />

Fig. 3-6. Correlations between the energy and the time-of-flight (TOF) for detected particles by<br />

MCP and SSD (θLab.= 0). Different particles are localized in their own way and well identified, as<br />

indicated by groups in the figure.<br />

The RNB-K05 (Spokesperson: T. Fukuda (Osaka Electro-Comm. Univ.), H.<br />

Ishiyama (<strong>KEK</strong>), Collaboration: Osaka Electro-Comm. Univ. (3), Tsukuba Univ. (4),<br />

<strong>KEK</strong> (8), JAEA (7), Osaka Univ. (1), CNS (4), GSI (1), NAO (1)) was proposed to<br />

measure the cross sections of 8 Li(α, n) 11 B reaction. This reaction is considered to be a<br />

key reaction for heavy-element nucleosynthesis, since it bridges the gap of atomic mass<br />

number of A = 8 in the stellar element abundance. We had already measured the cross<br />

sections in the energy range of Ecm = 0.7 – 2.6 MeV [3-3], which corresponds to the<br />

Gamow peaks of T9 = 1 – 3. In the measurement, we used a recoil mass separator<br />

(RMS) as an in-flight fragment separator for producing the 8 Li beam, and employed a<br />

gas counter (Multiple Sampling Tracking Proportional Chamber: MSTPC) as an active<br />

target, where the operating gas for the counter played simultaneously the role of gas<br />

target for the nuclear reaction of interest, and neutron detectors [3-9]. The cross sections<br />

were about two times smaller than those of previous inclusive measurements where<br />

only 11 B were measured, and a resonance-like structure in the excitation function was<br />

58


successfully employed to measure the 8 Li(α, n) 11 B reaction cross sections. The 8 Li beam<br />

with an intensity of 3x10 4 pps on average was injected into the GEM-MSTPC for about<br />

3 days at the <strong>TRIAC</strong>.<br />

Neutron counter<br />

GEM-MSTPC<br />

8 Li beam<br />

Fig. 3-8. Experimental set-up for the measurements of 8 Li(α,n) cross sections. The beam of 8 Li is<br />

counted one-by-one by the MCP and then injected into the GEM-MSTPC, where the occurrence<br />

of the reaction (sudden change of the energy loss by recoil 11 B) is identified and the<br />

simultaneously emitted neutron is measured by the neutron counter.<br />

Although the intensity was limited by the data acquisition rate allowed in the<br />

experiment because of unexpectedly low reduction rates of the veto counter installed<br />

downstream of the GEM in order to remove beam particles without inducing reactions<br />

of our interest (i.e. elastic scattering), the average injection rate is about 10 times larger<br />

than that in our previous measurement using the RMS. In the experiment, the width of<br />

the pad reduced by a factor of three as compared that in the precious measurement at the<br />

RMS (but with the same operating pressure of the active gas) allows narrower beam<br />

energy binning, accordingly higher energy resolution for Ecm in the measurement of the<br />

excitation function. The picture of the experimental set-up is shown in Fig. 3-8.<br />

The trigger for data acquisition was generated from the coincidence of signals of<br />

the MCP and the neutron counter wall. Since neutron detectors had significant room<br />

background, recorded data were mainly composed of accidental events. In order to<br />

identify true reaction events, first of all, it is necessary to search for events with a<br />

sudden change in the energy loss along the particle trajectory identified by the<br />

60<br />

MCP


GEM-MSTPC among the accumulated data. Figure 3-9 shows a typical example of true<br />

reaction. The data analysis for selecting true reaction events on event-by-event bases is<br />

in progress.<br />

Fig. 3-9. Typical event pattern of 8 Li(α, n) 11 B reaction. (Left): Energy losses measured by pad<br />

(dE/pad) are presented as a function of the number of anode pads, where the beam comes in from the<br />

low pad number. (Right): Horizontal (x) and vertical (y) positions of the three-dimensional particle<br />

trajectories, where the beam direction was defined as the z-axis (i.e. increasing pad numbers in the<br />

figures).<br />

Using a stable ion beam of 4 He provided by the <strong>TRIAC</strong>, the RNB-KJ04<br />

(Spokesperson: H. Makii (JAEA) and H. Miyatake (JAEA), Collaboration: <strong>KEK</strong> (9),<br />

JAEA (5), RIKEN (1), Osaka Univ. (1)) was proposed to directly measure the 12 C(α,<br />

γ) 16 O reaction cross sections and determine the strength of electric-dipole (E1) and/or<br />

-quadrupole (E2) components in the α-capture cross sections (decaying strength of<br />

excited 16 O by emitting γ-rays of corresponding multi-polarities) at stellar energies. The<br />

reaction plays an important role at the stage of helium-burning in stellar evolution; its<br />

reaction rate determines the mass fraction of 12 C and 16 O after the stellar helium-burning,<br />

the abundance of elements from oxygen to iron, and the iron-core mass before<br />

61


supernovae explosion [3-11]. Therefore, it is quite important to accurately determine the<br />

cross section at Ecm ≈ 0.3 MeV corresponding to the stellar temperature of interest. The<br />

direct measurement of the cross section at Ecm ≈ 0.3 MeV, however, is not possible with<br />

current experimental techniques, because the cross section is too small, around 10 -17<br />

barn. Hence one has to estimate the cross section by extrapolating the cross sections<br />

measured at energies higher than 1.0 MeV into the region of corresponding energies,<br />

and the extrapolations rely on theoretical calculations that gave scattered results<br />

according to their theoretical approaches. Consequently, the estimated cross section at<br />

Ecm ≈ 0.3 MeV has a large uncertainty, which turned out to be largely due to the poor<br />

experimental determination of the relative strengths of E2- to E1-radiative capture<br />

reactions (σE2 / σE1) at higher energies.<br />

Recently, it has been shown that the ratio of σE2 / σE1 could be determined with a<br />

good accuracy by using high-efficiency anti-Compton NaI(Tl) spectrometers to detect<br />

the γ-rays from the reaction, where a pulsed α-beam was employed to discriminate true<br />

events from background events induced by neutrons from 13 C(α,n) 16 O reaction with a<br />

time-of-flight (TOF) method and the target thickness and α-beam intensity were<br />

monitored during the measurement by the Rutherford backscattering spectrum of<br />

α-particles from the target [3-12, -13]. In order to provide a stringent constrain to the<br />

extrapolation down to Ecm ≈ 0.3 MeV, the present proposal aims at measuring the ratios<br />

at several energies, where large discrepancies have been observed in the values of the<br />

ratios predicted by various theoretical calculations, by using the experimental set-up in<br />

Ref. 3-13. The measurements would provide good insight into distinguishing between<br />

different calculations, and thereby help reduce the large uncertainty in the<br />

extrapolations.<br />

In the proposed experimental set-up, the use of the pulsed intense α beam is<br />

essential for reducing the background γ-rays induced by neutrons originating from the<br />

13<br />

C(α,n) reaction; the background γ-rays can be removed because they have a delayed<br />

detection time as compared to the prompt γ-rays of interest. The pulsed beam with a<br />

width of less than 10 ns and an interval between 250 ns and 500 ns was developed for<br />

the experiment at the <strong>TRIAC</strong>. We fabricated a beam-bunching system consisting of a<br />

pseudo saw-tooth wave pre-buncher and a multi-layer chopper, and tested the<br />

pre-buncher with a repetition frequency of 2 MHz using 16 O 4+ and 12 C 3+ beams with the<br />

same mass-to-charge ratio as 4 He 1+ from the <strong>TRIAC</strong>. A clear bunched structure with a<br />

width of 2 ns in standard deviation was observed at a target position 10 m away from<br />

the exit of <strong>TRIAC</strong> [3-14]. Incomplete beam bunching, observed as a remainder of the<br />

26-MHz operation of SCRFQ was removed by the operation of the chopper installed<br />

62


upstream of the pre-buncher, avoiding beam injection into the buncher during the<br />

shaping time of the saw-tooth wave-form for beam bunching [3-15].<br />

Fig. 3-10. Experimental set-up for the measurements of 12 C(α,γ) cross sections. Three NaI(Tl)<br />

detectors are placed at 40 o , 90 o , 130 o relative to the beam direction. The scattering chamber was<br />

installed at the center of the detectors and includes a water cooled target holder where several<br />

12 C-enriched targets on the gold foil backing were set. The vacuum in the chamber was about 10 -7<br />

Pa.<br />

Fig. 3-11. Correlation in the energy and the time of γ-rays measured by the NaI(Tl) detector at 40 o<br />

using bunched α-beams of 0.701 MeV/nucleon. Vertical and horizontal axes correspond to the<br />

observed γ-ray energies in unit of MeV and the times-of-flight from the RF signals in unit of<br />

nanosecond, respectively. Intensities of correlated events are presented by different colors as<br />

indicated in the right-hand side of the figure. Events in the red circle indicate the characteristic γ<br />

transitions from excited state populated by the reaction to the ground state in 16 O.<br />

Using the beam-bunching system and experimental set-up shown in Fig. 3-10, the<br />

experiment has been recently performed at four incident energies ((0.701, 0.774, 1.12<br />

and 1.13 MeV/nucleon) of the α-beam. Figure 3-10 shows the experimental set-up,<br />

which is similar to the one in Ref. 3-13 except for a water cooled target holder for the<br />

63


pure (99.9 %) 12 C target. A monitor SSD was set at 77 cm upstream of the target to<br />

measure elastically backward scattered α particles, whose energies reflect the effective<br />

thickness and uniformity of the irradiated target. Typical beam intensity was 5 µA.<br />

Irradiation times at 0.7 and 1.1 MeV/nucleon are 5 and 2 hours, respectively.<br />

Figure 3-11 shows one of the measured correlation spectra for the energies and the<br />

times-of-flight of detected events by NaI(Tl) detectors. The start trigger of the<br />

time-of-flight measurement was the RF signal (26 MHz) from the <strong>TRIAC</strong>. The<br />

characteristic γ-rays of about 9 MeV decaying from excited states in 16 O can be found at<br />

around 200 ns as a part of prompt events, which are strongly correlated to the arrival of<br />

the pulsed α-beam. Such prompt events are clearly separated from delayed γ-ray events<br />

starting from around 210 ns; the present observation well demonstrates the successful<br />

operation of the beam pre-bunching system (no beam bunches prior to the main bunch<br />

as shown in Fig. 2-33), otherwise the prompt γ-rays around 9 MeV would be hindered<br />

the delayed γ-rays. Some low-energy, randomly existing γ-rays were also observed,<br />

which indicates that their origin are related to the thermal neutron capture reactions by<br />

the contaminants around target or in detectors (e.g. γ-rays of 6.8 MeV originate from<br />

127<br />

I(n, γ) reactions induced by thermal neutrons in the NaI(Tl) detectors). The reaction<br />

cross section will be deduced for those prompt events with γ-ray energy of 9 MeV.<br />

Further analysis is in progress.<br />

3-2. Nuclear and fundamental physics<br />

The following experiments to study nuclear structure and fundamental physics<br />

were carried out using the radioactive ion beams (RIBs) provided by the <strong>TRIAC</strong>: (1)<br />

RNB-K02/K03: Nuclear spectroscopy through β-delayed decay of spin-polarized nuclei<br />

around doubly magic nucleus 132 Sn (Spokesperson: Y. Hirayama (<strong>KEK</strong>) and K.<br />

Matsuta (Osaka Univ.), Collaboration: <strong>KEK</strong> (9), JAEA (6), Osaka Univ. (5), Niigata<br />

Univ. (2), Kochi Univ. of Tech. (1), RIKEN (1)). (2) RNB-K04: Test of time reversal<br />

symmetry using polarized nuclei (Spokesperson: J. Murata (Rikkyo Univ.) and Y.<br />

Hirayama (<strong>KEK</strong>), Collaboration: Rikkyo Univ. (9), <strong>KEK</strong> (3), JAEA (2), and RIKEN<br />

(1)). (3) RNB-K08: Search for the highly excited state in 11 Be via 9 Li + d reaction<br />

(Spokesperson: T. Teranishi (Kyushu Univ.) and H. Ishiyama (<strong>KEK</strong>), Collaboration:<br />

<strong>KEK</strong> (6), JAEA (7), CNS/Univ. of Tokyo (1), Osaka Univ. (2), and Tsukuba Univ. (1)).<br />

These experiments are respectively summarized in the following subsections. The status<br />

of the RNB-KJ03 (Spokesperson: N. Imai (<strong>KEK</strong>)), which is to be carried out soon,<br />

will be presented at the end of this section.<br />

64


The process of nuclear polarization by the TF technique is schematically shown in<br />

Fig. 3-13: The tilted-foil method uses the anisotropic atomic collision of incident ions<br />

with the conduction (constituent) electrons at the exit surface of the foil. When the ion<br />

beams passing through a thin foil, tilted against the incident beam direction at an<br />

oblique angle (θ in Fig. 3-13), the electronic states of the outgoing ions are polarized.<br />

Polarization is initially introduced in the orbital motion of the electrons by the surface<br />

interaction at the instant of exit of ions from the foil. During flight in free space, some<br />

of this electron spin (J) polarization is transferred to the nuclear spin (I) through<br />

hyperfine interaction. The direction of the polarization is well defined; n x v where n is<br />

the unit vector normal to the surface of the foil at the exit and v is the ion’s velocity. By<br />

a successive passage of several such foils, interspersed with regions of free flight to<br />

allow a significant nuclear precession around the total angular momentum F=I+J in<br />

flight, the polarization effect is enhanced. In this way, rather sizable nuclear polarization,<br />

~10 %, for a wide variety of elements can be achieved [3-18].<br />

The features of this method can be summarized as follows: The method is<br />

applicable to any elements and the RIBs of a few hundred keV/nucleon are considered<br />

to be most suitable for the effective capture of electrons, which is essential for the<br />

primary atomic polarization. The degree of polarization increases with the number of<br />

foils [3-19] and the tilted angle between the beam axis and the axis normal to the foil<br />

surface [3-20]. The direction of polarization is easily reversed by reversing the normal<br />

axis of the foil surface. The degree of nuclear polarization PI induced by the multi-foil<br />

can be approximated as follows [3-21],<br />

N { Q } PJ<br />

I + 1<br />

PI ( N)<br />

~ 1−<br />

.<br />

J + 1<br />

Here, N is the number of foils, I and J are respectively nuclear and atomic spin, PJ is the<br />

degree of atomic polarization and Q is expressed approximately for small PJ by:<br />

2 2<br />

1− 2/<br />

3⋅<br />

J / I if I / J > 1 , 1 / 2 if I / J = 1 , 1 / 3 if I / J < 1 . From the nuclear<br />

polarization measured as a function of the number of foils, dominant atomic state for the<br />

atomic and nuclear polarization could be estimated by using the equation for PI.<br />

Experimental setup: Figure 3-14 shows a schematic side view of the experimental<br />

setup for the production and measurement of nuclear polarization. Reaccelerated RIBs<br />

pass through a stack of tilted-foils in vacuum chamber. After being spin-polarized by the<br />

TF method, they are implanted into a catcher (often called as catcher or stopper foil),<br />

where a magnetic field B0 is applied to preserve the nuclear polarization. The catcher<br />

can be cooled by a refrigerator (down to about 10 K, if necessary) in order to achieve<br />

66


much longer spin-relaxation time than the lifetimes of RIBs (necessary condition for the<br />

measurement of nuclear polarization by β-NMR method). It is the most critical to<br />

choose the catcher that is suitable for each RIB to hold the nuclear polarization during<br />

their lifetimes, otherwise the nuclear spin would be depolarized before measurement. A<br />

radio-frequency (RF) oscillating field B1 for the NMR measurement is generated by the<br />

RF coils fixed around the catcher. The emitted β-rays are detected by the double layered<br />

plastic scintillator telescopes in the atmosphere, which are placed above and below the<br />

catcher. Nuclear polarization PI is evaluated from the up/down ratio W(0 o )/W(180 o ),<br />

where W(θ) is the angular distribution W(θ) ~ 1 + APcosθ of emitted β-rays from<br />

polarized nuclei. Here A and θ are the asymmetry parameter of β-transition and the<br />

emission angle with respect to the nuclear polarization axis, respectively.<br />

Fig. 3-14. Schematic side view of the experimental setup for polarization by the TF method and<br />

β-NMR measurement<br />

Figure 3-15 shows: (a) the photos of the experimental setup, (b) a polystyrene foil<br />

on a non-magnetic copper ladder and (c) a stack of 20 polystyrene foils with tilt angle<br />

70 degrees, respectively. The area of each foil is 45 mm x 15 mm which corresponded<br />

to 15 mm x 15 mm at the tilted angle of 70 o in view of the beam direction. The area is<br />

about twice larger than the beam spot about 8 mm in diameter at the entrance of the<br />

foils, considering the broadening of beam spot due to the multiple scattering when the<br />

beam passes through in the multi-foils. The thickness of the foil is about 3 µg/cm 2 (~ 30<br />

nm), and therefore the number of the stacked foils could be increased three times more<br />

than those of the generally used carbon foils.<br />

Polarization of 8 Li: The TF method has been studied in detail at the <strong>TRIAC</strong> by using<br />

the 8 Li (τ1/2 = 838 ms, I π = 2 + , g = 0.8268) beam at various energies (140, 178 and 240<br />

keV/nucleon). The 8 Li nuclei with a typical intensity of ~10 5 pps were implanted into an<br />

67


annealed platinum foil (stopper foil), which is 10-µm thick and placed at room<br />

temperature in the magnetic field of B0 ~ 0.05 T for holding the nuclear polarization.<br />

The β-NMR technique was applied to the measurement of the nuclear polarization.<br />

Ten carbon foils, each of which is of 10 µg/cm 2 in thickness and distant by ~1 mm<br />

from each other, were used with a tilted angle of 70 o in the first experiment. The<br />

polarization was observed to be 3.7 ± 1.1 % but it was not saturated with the 10 sheets<br />

of foils. Although higher polarization is expected with increasing number of foils, the<br />

carbon foils are so thick that the 8 Li beam cannot pass through more than ten foils.<br />

Therefore, we developed thinner polystyrene foils (e.g. ~3 µg/cm 2 in thickness). In the<br />

second and third experiments, the polarization of 7.3 ± 0.5 % was achieved at the beam<br />

energy of 140 keV/nucleon by using 15 foils of polystyrene at the tilted-angle of 70 o .<br />

Figure 3-16 shows the polarization as a function of the number of polystyrene foils for<br />

various energies of incident 8 Li.<br />

(a) (b)<br />

(c)<br />

RIBs<br />

Stacked tilted-foils<br />

20 foils @ 70o 20 foils @ 70o Magnet<br />

B 0<br />

Nuclear<br />

polarization<br />

68<br />

45<br />

15<br />

Fig. 3-15. Photos of (a) the experimental<br />

setup for the polarization of heavy<br />

elements, (b) a polystyrene foil and (c) a<br />

stack of the polystyrene foils. For the<br />

polarization of 8 Li, we used a rather<br />

simple configuration because much lower<br />

holding field (Bo) was allowed (i.e. a<br />

coreless electromagnet was used).


8 Li polarization : PI (%)<br />

10<br />

8<br />

6<br />

4<br />

2<br />

140 keV/A ( 70 deg. )<br />

178 keV/A ( 70 deg. )<br />

240 keV/A ( 70 deg. )<br />

0<br />

0 5 10 15 20 25 30<br />

Number of foils<br />

Fig. 3-16. 8 Li polarization as functions of sheet-number of polystyrene foils measured for incident<br />

beam energies of 140, 178 and 240 keV/nucleon, respectively. The solid lines were given by fitting<br />

the formula for nuclear polarization (PI) to the data. The data points observed with largest number of<br />

foils at the beam energy of 140 and 240 keV/nucleon are largely deviated from the general tendency;<br />

it might be related to an unintentional geometrical mismatch in the foil system, especially at the exit<br />

of the foil (e.g. tilted angle, damped energy and broadened spatial distribution of the beam, etc.).<br />

Except for the two data points, the data points seem to be well reproduced with the formula.<br />

In the third experiment, we tested a new experimental set-up modified for the<br />

polarization of the heavy nuclei around 132 Sn, as shown in Fig. 3-15 (a): The tilted-foil<br />

system was modified to reduce the distance between the tilted-foil and stopper for<br />

making possible simultaneous implantation of beam of several charge states after<br />

passing through the stack of foils. The modified β-NMR system can operate at a high<br />

magnetic field B0 ~ 0.5T (Holding field) and a high frequency for RF magnetic field<br />

(oscillating transverse field). The operation of the system was confirmed by using the<br />

8 8<br />

Li beam. The polarized Li beam was applied for test of time reversal symmetry using<br />

polarized nuclei (K04) described in the next section.<br />

The solid lines in Fig. 3-16 show the fitted results using the formula, which is<br />

given above for the polarization where the multi-foil effect is taken into account. In<br />

fitting, we assumed the atomic spin of J = 1/2 for 8 Li ions, whose values actually govern<br />

the dependence of the polarization on the number of foils. For example, the formula<br />

becomes saturated with about 10 foils and cannot reproduce the experimental data when<br />

the atomic spin of higher than J = 1/2 is assumed. It indicates that 2p ( 2 P1/2) state may<br />

dominantly contribute to produce the polarization of 8 Li. In the presently used formula<br />

69


the observed CP-violating phenomena in K-meson and B-meson system, predicts a<br />

negligible effect on u-d quark (nuclear) systems. On the other hand, some models based<br />

on physics beyond the standard model, predict visible size of CP-violating effects, i.e.<br />

T-violating effects in normal nuclear system. Therefore, the observation of a non-zero<br />

T-violating effect directly implies the evidence for a new physics beyond the standard<br />

model. A triple vector correlation between nuclear spin J I<br />

r , electron spin σ e<br />

r and<br />

momentum Pe r is a T-odd variable. Therefore, assuming time reversal symmetry, beta<br />

r r r<br />

decay rate must not depend on the R-correlation, defined as; σ e ⋅ 〈 J I 〉 J × P I e E . In the<br />

e<br />

standard model, R is almost zero. If non-zero R-correlation exists, electrons emitted<br />

from the β-decaying polarized nuclei would have non-zero transverse polarization.<br />

Therefore, experimental investigation of time reversal symmetry can be accomplished<br />

by the measurement of electron momentum and its transverse polarization, using<br />

β-decaying polarized nuclei [3-23, -24]. Figure 3-18 shows the principle of determining<br />

electron transverse polarization. Electrons are emitted from the stopped nuclei and then<br />

injected into a polarimeter, which is installed outside of the vacuum chamber. The<br />

polarimeter utilizes the transverse analyzing power in the Mott scattering of electrons<br />

from heavy nuclei (gold or lead foils). By measuring the ratio between electrons<br />

scattered upward and downward, the transverse polarization can be determined.<br />

Fig. 3-18. Principle of determining electron transverse polarization by observing up-down<br />

asymmetry in backward scattering angular distribution.<br />

The electron polarimeter using a drift chamber has been developed. Figure 3-19<br />

shows the experimental setup. The polarimeter consists of trigger plastic scintillators<br />

counters, a planer multi-wire drift chamber, and an analyzer foil of Pb (113-mg/cm 2<br />

thick). The drift chamber is designed for tracking the trajectories of electrons; it<br />

observes not only the tracks of the incident electrons, but also the second tracks of<br />

electrons scattered backwards by the thin lead foil placed behind the chamber.<br />

72


In the first experiment, the performance test of the polarimeter was carried out by<br />

using unpolarized 8 Li beam of 178 keV/nucleon. The “V-track” events scattered from a<br />

thin gold foil (10-µm thick, 300 mm x 500 mm in area) were successfully identified and<br />

the observed angular distribution of the Mott scattering shows a good agreement with<br />

the Mont-Carlo simulation taking into account the full detector system.<br />

Polarimeter:<br />

Drift chamber<br />

Analyzing foil:<br />

Pb<br />

73<br />

Polarized<br />

8 Li beam<br />

Stopping plastic counters<br />

Fig. 3-19. Experimental setup. Electrons scattered at backward angles from the analyzer foil form<br />

“V-track”s, which is detected by the planer drift chamber.<br />

In the second experiment, the measurement of the R-correlation (i.e. the first<br />

experimental test of the time reversal symmetry in nuclear beta decay of a polarized<br />

nucleus) was examined. For the measurement, a new FPGA-based intelligent triggering<br />

system (DC trigger) was installed in order to suppress the background events such as<br />

Coulomb multiple scattering and X-ray radiation, which helps reduce significantly the<br />

backgrounds by utilizing a V-track reconstruction technique on an event-by-event basis.<br />

As a result, about 10 6 V-track events from polarized 8 Li nuclei have been successfully<br />

recorded in the measurement. Thanks to the DC trigger, the purity of the real V-tracks<br />

among the recorded events was improved by 10 times better than in the previous<br />

experiment, and the live time of the data taking system was found to be almost 100 %<br />

with a beam rate of ~10 5 pps. In order to cancel geometrical asymmetries of the<br />

polarimeter, the spin direction of incident 8 Li beam is flipped every 5 minutes by<br />

rotating the tilted foil angles. Totally 17-Giga 8 Li ions were implanted on the annealed<br />

platinum stopper, which sandwiched by spin holding permanent magnets. About<br />

1.6-Giga triggers by plastic scientillators were generated at a rate of around 12 kHz, and<br />

finally 13-Mega DC-triggered events were recorded at only around 0.1 kHz. After the


offline data analysis, about 1-Mega V-track events were successfully reconstructed.<br />

The Mott scattering angular distribution is shown in Fig. 3-20, for both spin-up and<br />

-down settings of the tilted-foil system for polarized 8 Li beam. If shapes of the<br />

distributions are different from each other, it means that the time reversal symmetry is<br />

violated. The value obtained from the preliminary analysis is concluded to be zero<br />

within the present statistics.<br />

Fig. 3-20. Angular distributions of the V-track events of electrons scattered backwards by the lead<br />

foil.<br />

The R&D for the test of the time reversal symmetry using polarized 8 Li nuclei at<br />

<strong>TRIAC</strong> [3-25] is successfully performed, as the first measurement using event-by-event<br />

V-track reconstruction without suffering from dominant backgrounds. The present<br />

system developed for the experiments at the <strong>TRIAC</strong>, shows a good performance as an<br />

electron transverse polarimeter which is going to be used at the TRIUMF. By utilizing<br />

the 8 Li beam with higher intensity and better polarization (10 7 pps with 80 %<br />

polarization) at the TRIUMF, the R-parameter will be determined with a precision of<br />

0.01%. The project has been already started at the TRIUMF as S1183 MTV experiment<br />

from 2009. First physics run is successfully performed in 2010 with highest statistical<br />

precision, utilizing the experimental technique developed at the <strong>TRIAC</strong>.<br />

3-2-3. Search for the highly excited state in 11 Be via 9 Li + d reaction (K08)<br />

An excited level at 18.2 MeV in 11 Be has been suggested from observations of<br />

beta-delayed deuteron, triton, and α emissions in 11 Li beta decay experiments [3-26].<br />

This level is above and close to the 9 Li+d threshold at 17.9 MeV and possibly enhances<br />

74


the beta-delayed deuteron emission probabilities [3-27] (Fig. 3-21). This level is also a<br />

candidate of a “halo analog state” of 11 Li, which is populated by a Gamow-Teller (GT)<br />

transition of the two halo neutrons in 11 Li, namely 11 Li ( 9 Li+2n) 11 Be* ( 9 Li+d) [3-28].<br />

20.68<br />

11 Li<br />

β −<br />

18.15<br />

0<br />

11 Be<br />

75<br />

17.916<br />

9 Li+d<br />

8 Li+t, 2α+3n,<br />

6 He+α+n,<br />

9 Be+2n, 10 Be+n<br />

Fig. 3-21. Level diagram of beta-delayed particle-emitting 11 Li.<br />

Clear determination of the energy, width, and spin-parity may be therefore useful for<br />

studying the GT transitions of 11 Li and its halo structure. However the signature of the<br />

level was not so clear in the beta decay experiments; the energy and width of the level<br />

were deduced from the spectrum, which contained unresolved deuterons and tritons<br />

without any well identified peaks, and from an unresolved 4,6 He spectrum in the decay<br />

11 Li (β) 11 Be * 6 He+α+n, which did not show any peak corresponding to the level<br />

because of the three body decay. Moreover, a recent observation of beta delayed<br />

deuterons did not indicate any resonance signature [3-29]. Further experimental<br />

investigation with a different approach is necessary.<br />

Fig. 3-22. Experimental setup for studying 9 Li+d reactions.<br />

In the present experiment we have studied 9 Li+d reactions to search for a<br />

resonance signature attributed to the 18.2-MeV level. We measured deuteron, triton,<br />

alpha spectra from 9 Li+d elastic scattering, 9 Li(d,t), and 9 Li(d,α) reactions, respectively.<br />

The 9 Li+d reaction experiment was performed at the <strong>TRIAC</strong>. A primary beam of 7 Li at<br />

70 MeV from the JAEA tandem accelerator bombarded an ISOL target of BN.


Secondary 9 Li ions extracted from the ISOL target were accelerated up to 0.85<br />

MeV/nucleon. Figure 3-22 shows the experimental setup for the study of the 9 Li+d<br />

reactions. The beam ions of 9 Li were counted on an event-by-event basis using an MCP<br />

detector with a 0.1 mg/cm 2 Mylar foil. After passing through the MCP detector, the 9 Li<br />

beam bombarded a (CD2)n target of 2.3 mg/cm 2 with an average beam intensity of<br />

approximately 3×10 4 pps. To deduce the excitation functions of 9 Li(d,d), 9 Li(d,t), and<br />

9<br />

Li(d,α) reactions, we utilized a thick target method in inverse kinematics (TTIK),<br />

where the target thickness was chosen to be thick enough to stop most of the incident<br />

9<br />

Li ions in the target. The recoil deuteron, triton, and α particles suffered relatively<br />

small energy losses in the target and detected by a Si detector array (SDA) located<br />

26-cm downstream of the target with its center at 0°. The SDA consisted of 3×3 Si<br />

detectors, each of which was 300-µm thick and had a sensitive area of 28×28 mm 2 .<br />

TOF(ns)<br />

120<br />

115<br />

110 110<br />

105<br />

100<br />

95<br />

90<br />

85<br />

80<br />

75<br />

70<br />

0 2 4 6 8 10 12 14<br />

E (MeV)<br />

Fig.3-23. E-TOF spectrum for recoil particles.<br />

The particle identification was made by using time-of-flight (TOF) information<br />

between the MCP and SDA and energy (E) information of the SDA. Figure 3-23 shows<br />

an example of the measured TOF-E spectrum. There was a background component<br />

below 2 MeV, which was due to accidental α particles from the beta decay of 9 Li. The<br />

timing distribution of the accidental α particles was essentially uniform and their energy<br />

distribution could be easily estimated. Then, the accidental background component was<br />

subtracted from the deuteron, triton, and α spectra. Though analysis is in progress, a<br />

preliminary result of the 9 Li+d spectrum contains no sharp peak but shows<br />

monotonically increasing cross sections toward the low excitation energy side. A<br />

76


possibility of resonance contribution to this spectral shape will be examined by utilizing<br />

R-matrix analysis. The triton and α spectra will be analyzed on the same R-matrix basis<br />

by assuming some resonance contributions.<br />

3-2-4. Coulomb excitation of 142 Ba at 1.1 MeV/nucleon (KJ03)<br />

Barium isotopes are located in an island of octupole deformation. The B(E2) and<br />

B(E1) values in these nuclei give us crucial information on the correlations to the dipole<br />

or quadrupole moments in nuclei. A recent lifetime measurement of the first 2 + (2 + 1)<br />

state of 142 Ba gave a B(E2) value inconsistent with the previous experimental data<br />

[3-30]. Up to date, the B(E2) values of 142 Ba have been measured by using the lifetime<br />

measurements of the 2 + 1 state produced by the spontaneous fission of 252 Cf. The<br />

inconsistency among the previous data can be attributed to how to take into account the<br />

effects of cascade transitions and recoil velocities of fission products. Instead of lifetime<br />

measurement, the KJ03 proposed the Coulomb excitation measurement of 142 Ba<br />

employing the <strong>TRIAC</strong>. The <strong>TRIAC</strong> can accelerate radioactive isotopes extracted by the<br />

JAEA-ISOL up to 1.1 MeV/nucleon, which is far below the Coulomb barrier. Thus, the<br />

cascade transition can be negligible since the probability of multi-step excitation should<br />

be very small.<br />

As a test of the whole experimental system including charge-breeding electron<br />

cyclotron resonance ion source (<strong>KEK</strong>CB), we performed an experiment of Coulomb<br />

excitation with a stable 130 Ba beam of 1.1 MeV/nucleon. The energy of 2 + 1 state of 130 Ba<br />

is 359 keV, which is close to the 357 keV for 142 Ba. Therefore, the present experiment<br />

well simulates the measurement of 142 Ba.<br />

A 130 Ba 1+ beam was supplied by the JAEA-ISOL using the enriched BaCO3.<br />

Typical beam intensity was about 30 enA. The beam was injected into the <strong>KEK</strong>CB for<br />

charge-breeding the singly charged ions 130 Ba 1+ to multi-charged ones 130 Ba 19+ . The<br />

charge breeding efficiency was around 1%. The 130 Ba 19+ beam was transported to<br />

post-accelerators of the <strong>TRIAC</strong> and accelerated to 1.1 MeV/nucleon.<br />

At the final focal plane of the <strong>TRIAC</strong>, a secondary target of 2-µm thick nat V was<br />

placed and irradiated by the beam. 6 NaI(Tl) detectors of 9 x 9 x 18 cm 3 were placed to<br />

detect the de-excitation γ rays from the excited 130 Ba around the target. The NaI(Tl)<br />

detectors were mounted in the 2-cm thick lead box to suppress the background γ rays or<br />

X rays. To measure the recoiled particles, an annular type SSD of a large solid angle<br />

was placed 18-cm downstream of the target. The detector covered from 12 to 24 o in the<br />

laboratory frame, which corresponds to angles from 48 to 160 o in the center of mass<br />

frame.<br />

77


Fig.3-24. γ ray energy spectra coincident with SSD (solid line) and accidentally coincident with SSD<br />

(dashed line).<br />

Fig. 3-25. Set-up for the Coulex experiment. The set-up used for the 130 Ba experiment is given<br />

schematically as an inset. The picture of the set-up at the experimental hall is also given, where part of<br />

10 NaI(Tl) detectors surrounding the target (here Vanadium target) are visible. An annual SSD is<br />

installed in the chamber downstream of the target. The beam passing through the target is monitored a<br />

plastic scientillator for beta-decaying 142 Ba beam installed at the position of the FC for the stable 130 Ba<br />

beam.<br />

Figure 3-24 shows two γ-ray energy spectra measured by three NaI(Tl) detectors<br />

with different timing gates. Solid line indicates the γ rays which were coincident with<br />

78


the SSD, while dashed line represents the accidentally coincident γ rays; only a peak<br />

around the energy of 360 keV was clearly observed as expected. Comparison of the<br />

cross sections between elastic and inelastic scatterings yielded the B (E2;2 + 0 + ) = 0.97<br />

(22) e 2 b 2 , which was in good agreement with the adopted value of 1.163(16) e 2 b 2 .<br />

The beam development of 142 Ba was also performed independently. The beam<br />

intensity can be expected to be several 10 3 particles/s at the secondary target. The<br />

measurement of B(E2) value of the 2 + 1 state of 142 Ba is scheduled to be carried out soon.<br />

For the experiment, the number of NaI(Tl) in increased from 6 to 10, by which the<br />

detection efficiency for γ-rays of 360 keV becomes 14 %. The experimental set-up is<br />

shown in Fig. 3-25.<br />

3-3. Materials science<br />

<strong>TRIAC</strong> can provide low-energy radioactive ion beams ranging from 2 keV/nucleon<br />

to 1.1 MeV/nucleon with a small beam size, for example less than 10 mm in diameter,<br />

but still with intensity high enough for experiments. Thanks to these characteristics, we<br />

can implant the RIBs into samples of interest, usually small in size, with an<br />

implantation depth well-suited for specific experimental requirement.<br />

Over more than half a century, diffusion studies in solids using radioactive tracers<br />

[3-31] have played an important role in understanding the underlying mechanism of<br />

atomic transport in solids, which is of great importance in a number of branches of<br />

materials science and engineering. Although the conventional radiotracer method for<br />

diffusion studies in conjunction with a serial sectioning technique [3-32] has yielded the<br />

most accurate diffusion coefficients, the method has not yet been applied for some<br />

elements because of no-availability of radiotracers with adequate lifetimes. Among<br />

them, 8 Li, the radioactive isotope of Li with a half-life of 0.84 s is of special interest for<br />

practical issues; how well Li ions move in the secondary Li ion batteries. Fast Li<br />

diffusion is desirable in battery materials, i.e., Li ionic conductors for materials of<br />

electrodes and solid electrolyte. For studies on the macroscopic diffusivity of Li in Li<br />

ionic conductors, various electro-chemical methods [3-33] have been usually adopted<br />

up to now. However, the diffusion coefficients are scattered over several orders of<br />

magnitude, strongly depending on the method used for the measurement [3-34].<br />

Therefore, the diffusion coefficients measured in different ways, e.g. by using the<br />

radiotracer of Li, are highly required to settle down such disagreements. Such an<br />

experimental knowledge on the Li diffusion in as-developed materials for the battery is<br />

also of importance in the recent general efforts to design the battery by simulations<br />

based on the first principle.<br />

79


In the experiments of RNB-J02/03 (Spokesperson: S.C. Jeong (<strong>KEK</strong>), H. Sugai /<br />

M. Sataka (<strong>KEK</strong>)), Collaboration: (<strong>KEK</strong> (8), JAEA (7), Aomori Univ. (2), NIMS (3),<br />

Osaka Prefecture Univ. (1), CNS (1)), it has been found that the short-lived radioactive<br />

beam of 8 Li at the <strong>TRIAC</strong> successfully operates as a diffusion tracer in the lithium ionic<br />

conductors. This kind of application has been performed for the first time over the<br />

world. In the experiment, we measured α-particles, emitted from β-decaying 8 Li,<br />

coming out of the sample of interest after implanting RIBs, and have found that the<br />

time-dependent yields of α-particle from the diffusing 8 Li primarily implanted is a good<br />

measure of the lithium diffusion in the sample.<br />

The radiotracer 8 Li decays through β-emission to 8 Be with a half lifetime of 0.84 s,<br />

which immediately breaks up into two α-particles with energies continuously<br />

distributed around 1.6 MeV with a full width at half maximum (FWHM) of 0.6 MeV<br />

[3-35].<br />

Figure 3-26 shows schematically the principle of the measurement: Implanting the<br />

beam of 8 Li with a properly adjusted energy into a depth, deeper than the average range<br />

of α-particles, we can make a situation where most α-particles stop in the sample. After<br />

implantation, since the primary implantation profile is broadening with time by<br />

diffusion, the α-particles emitted by 8 Li diffused toward the surface can survive and<br />

come out of sample with measurable energies. Then a charged particle detector, located<br />

close to the sample surface, could selectively detect α-particles from 8 Li diffusing<br />

toward the sample surface, since the implantation-depth is deeper than the average<br />

range of α-particles in the present case. Therefore, the temporal evolution of α-particle<br />

yields that come out of the sample is supposed to be a measure of the diffusivity of Li.<br />

Fig. 3-26. Principle of diffusion measurements by implanting 8 Li-RIBs.<br />

80


Provided by the <strong>TRIAC</strong>, the radiotracer beam 8 Li of about 4 MeV with an intensity<br />

of about 10 4 particles/s was periodically implanted to the sample of interest with a<br />

following time sequence; 1.6 s for implantation (beam-on) and of 5.1 s for subsequent<br />

diffusion (beam-off). The α particles coming out of the sample were measured as a<br />

function of time by an annular solid-state detector (SSD). The SSD with a central hole<br />

was installed close to the front surface of the sample as shown in Fig. 3-27. The<br />

sequence was repeated to obtain good statistics, where the time-zero was always at the<br />

beginning of the implantation. Before starting the measurement, the sample was set at a<br />

temperature where the diffusion coefficient was supposed to be measured. It should be<br />

noted that the present diffusion time is different from that of the conventional method<br />

[3-36] because the tracer in the present method diffuses all the time of the measurement.<br />

This is the reason why we call the present method an on-line measurement of diffusion.<br />

Fig. 3-27. Experimental set-up. The energetic and pulsed tracer beam of 8 Li are implanted into the<br />

sample and the decayed α particles are measured as a function of time by the solid state detector<br />

located in front of the sample. The condition of the tracer beam is given, which includes energy,<br />

intensity, repetition frequency of the pulsed beam and its duty factor.<br />

Figure 3-28 shows time spectra of the yield of α-particles measured at different<br />

temperature for the stoichiometric LiIn. Properly normalized for comparison, the spectra<br />

are presented by the ratios, i.e. time-dependent yields of α-particles divided by the α<br />

radioactivity of 8 Li at the time of interest. In this way was excluded the trivial<br />

time-dependency in the yield of α-particles just depending on the lifetime of 8 Li. If 8 Li<br />

does not diffuse at all, the values of the ratios should be constant over time. However,<br />

the experimental values as shown in Fig. 3-28 show different time structure depending<br />

on the temperature, well demonstrating that the time structure is a good measure of the<br />

macroscopic diffusivity of 8 Li at the temperature where the measurement has been<br />

performed. Based on this relative time-dependency of α-particle yields (time dependent<br />

81


atio), a diffusion coefficient was estimated by comparing with the simulation where<br />

one-dimensional Fickian (Gaussian) diffusion was assumed [3-34,-37].<br />

Fig. 3-28. Time spectra of α-particle yields measured at various temperatures for LiIn. The spectra<br />

were corrected for removing trivial time dependence and further normalized properly for easy<br />

comparison.<br />

Fig. 3-29. Temperature-dependence of diffusion coefficients (closed symbols) and electrical<br />

resistivity (open symbols) for β-LiGa with 44 at. % Li, and β-LiIn with 48 % at. Li.<br />

The diffusion coefficients of Li in β-LiGa and β-LiIn with a near stoichiometeric<br />

composition of Li, are displayed in Fig. 3-29 as a function of inverse temperature. For<br />

both samples, the diffusion coefficients suddenly change at a certain temperature,<br />

following Arrhenius behavior in the region of higher temperature. The sudden change in<br />

the value of the diffusion coefficient occurs at the temperature where the anomalous<br />

electrical resistivity is observed. The resistivity measurements were carried out using a<br />

van der Pauw method as used for β-LiAl [3-38]. This observation is closely related to<br />

82


the thermal properties of the structural defects, already observed as the anomalies in<br />

heat capacity [3-39, -40] and nuclear-spin lattice relaxation [3-41] at 233 K near the<br />

critical composition of the Li-deficient β-LiGa. It has been suggested that these<br />

phenomena are related to order-disorder transformation of the Li vacancies in the<br />

compounds.<br />

The ordering of the vacancies would produce a sharp drop in the Li diffusion<br />

coefficients below the ordering temperature since the vacancies are supposed to be the<br />

carriers of Li atom. The observed amount of change, more than two orders of magnitude<br />

in the value of diffusion coefficients in the case of β-LiGa, is quite impressive as<br />

compared to those observed in the measurement of electrical resistivities where just a<br />

small change (at most 1/10) can be seen at the transformation temperature. The<br />

transition temperature is known to shift to lower temperature, and the change in the<br />

electrical resistivity becomes invisibly smaller with increasing Li composition, where a<br />

large fraction of the Li vacancies forms complex defects with anti-structural Li atoms<br />

on Ga sublattices [3-42]. Therefore, with higher sensitivity, the present method could be<br />

applied for better investigating the characteristics of the transformation that are<br />

supposed to be correlated with the concentrations of structural defects strongly<br />

depending on the Li compositions.<br />

At the lower temperature followed by a sudden change around 234 (±2) K for<br />

β-LiGa, the diffusion coefficients are observed as a constant, which is the lower limit of<br />

diffusion coefficients accessible by the present method; for diffusion coefficients less<br />

than about 10 -10 cm 2 /s, any significant effect in the yields of α-particles due to the<br />

diffusing 8 Li could not be observed because of the short life-time of the radiotracer.<br />

In Fig. 3-30, the time-dependent normalized α-particle yields are compared for<br />

different Li compositions, i.e. 43.6, 50.0 and 53.2 at. % Li. Referring to the time<br />

dependence of the α-particle yields, the stoichiometric β-LiGa has the highest<br />

diffusivity of Li among three, most quickly rising and falling down. This observation is<br />

quite different from those observed for β-LiAl and β-LiIn [3-43] which are<br />

iso-structural with the β-LiGa.<br />

The abnormal behavior of Li diffusivity in β-LiGa is well identified in Fig.3-31,<br />

where the diffusion coefficients in β-LiGa at room temperature are presented as a<br />

function of Li composition together with the data for β-LiAl and β-LiIn reported in Ref.<br />

3-43. As shown in Fig. 3-31, the Li diffusion coefficients in β-LiAl and β-LiIn decrease<br />

monotonously with increasing Li composition, but the correlation is changing due to the<br />

coexistence of vacant Li sites VLi and antistructure Li atoms on the aluminum (or<br />

indium) sites LiAl (or LiIn), as discussed earlier in terms of the size effect of constituent<br />

83


Although the data for β-LiAl and β-LiIn were available only in a limited range of<br />

Li composition from 48.3 to 53 at. % Li, it would be interesting to note that the<br />

diffusion coefficients in β-LiGa varied in a similar way (increase with decreasing Li<br />

content) between those values in β-LiAl and β-LiIn in the corresponding range of Li<br />

composition. This would be intuitively understandable, since the iso-structural β-LiGa<br />

with a comparable size of the constituent atoms should have an interaction between VLi<br />

and LiGa with intermediate strength as compared to those discussed in the case of β-LiAl<br />

and β-LiIn [3-43]. Therefore, the composition-dependence of Li diffusivity in β-LiAl<br />

and β-LiIn could be considered as the lower and upper limits of the Li diffusivity in<br />

β-LiGa over the Li composition range of interest, respectively: two complementary<br />

explanations might be possible by referring to the respective composition-dependency<br />

observed in β-LiAl and β-LiIn.<br />

Referring to the tendency in β-LiIn, the diffusion of Li in very Li-poor β-LiGa<br />

seems to be suppressed. This could happen, for example, by assuming the formation of<br />

defect complex such as VLi-VLi and/or VLi-LiGa-VLi, since the concentration of VLi<br />

become much (almost three times at maximum) larger [3-44] in the more Li-poor<br />

composition than investigated in β-LiIn [3-43]. It should be noted that the number of<br />

vacant Li sites in a unit cell volume (8 for Li and 8 for Ga) is about two for the most<br />

Li-poor β-LiGa (43.6 at. % Li), whereas there exist about one vacant Li site in every<br />

two-unit cells for the most Li-poor β-LiAl (48.3 at. % Li) and β-LiIn (48.4 at. % Li) in<br />

Fig. 3-23. Here, we assumed the random distribution of the vacancies over the available<br />

sites.<br />

Alternatively, assuming the diffusivity of Li in the most Li-poor β-LiGa as a good<br />

extension of the diffusivity of Li in the Li-poor β-LiAl (48.3 at. % Li) because of nearly<br />

zero concentrations of LiAl and LiGa in the corresponding Li composition, the diffusivity<br />

of Li observed around the stoichiometric β-LiGa could be considered to be enhanced by<br />

the coexisting defects of VLi and LiGa. Under the coexistence of VLi and LiGa, the motion<br />

of the vacancies on the Li site, supposedly the carriers of Li atom, seems to be strongly<br />

promoted. This suggests that the interaction between VLi and LiGa would be stronger and<br />

rather repulsive than expected by the atomic size effect, but not as strong as observed in<br />

β-LiIn. In addition, as a specific characteristic of the interaction in β-LiGa, we found<br />

strong composition-dependence of the interaction that appears to become stronger when<br />

a comparable amount of two types of point defects exists.<br />

In the present measurement, although we found abnormal Li diffusion in very<br />

Li-poor composition of β-LiGa, the characteristic of the anomaly in diffusion, i.e.<br />

whether the diffusion is suppressed or enhanced, is not conclusive. A detailed theoretical<br />

85


consideration is highly required for quantitative discussion. From the experimental point<br />

of view, however, it would be interesting to extend the measurement to the more<br />

Li-poor composition of LiIn recently confirmed to have the β-phase in the same<br />

composition [3-45] as investigated for β-LiGa in the present work.<br />

Using the method mentioned above, tracer diffusion coefficient of lithium ions in<br />

the perovskite-type structured lithium ion conductors (La2/3-xLi3xTiO3; LLTO) have also<br />

been measured in the experiment of RNB-K07 (Spokesperson: S. Takai (Tottori Univ.),<br />

Collaboration: Tottori Univ. (1), <strong>KEK</strong> (8), JAEA (4)). The diffusion coefficients were<br />

consistent with those obtained by neutron radiography method [3-46] measured above<br />

200°C (Fig. 3-32). The tendency that lower lithium concentration with higher A-site<br />

vacancy possesses higher diffusion coefficients was confirmed around room<br />

temperature. Additional experiments for the sample of x=0.166, which is A-site stuffed<br />

composition, has been performed and the data analysis is in progress. In the experiments,<br />

a quenched sample with higher crystal symmetry was also studied and would help<br />

clarify the contribution of disordering of the vacancy.<br />

s -1<br />

D / cm 2<br />

10 -5<br />

10 -6<br />

10 -7<br />

Temperature, t /<br />

500 300 200 100 50<br />

o C<br />

10<br />

1 2 3<br />

-8<br />

T -1 / kK -1<br />

Fig. 3-32. Arrhenius plots of diffusion coefficients of LLTO measured by radiography () and the<br />

present 8 Li () methods. Red, blue and black symbols are for x = 0.066, 0.116 and 0.166,<br />

respectively.<br />

86


In the experiment of RNB-K06 ((Spokesperson: W. Sato (Kanazawa Univ.),<br />

Collaboration: Kanazawa Univ. (1), KURRI (1), <strong>KEK</strong> (4), JAEA (2)), local fields at<br />

111<br />

Cd nuclei in Highly Oriented Pyrolytic Graphite (HOPG) has been investigated by<br />

means of the Time-Differential Perturbed Angular Correlation (TDPAC) method. This<br />

experiment was designed to investigate the physical and chemical states of isolated<br />

atoms introduced in carbon allotropes. For the introduction of the probe 111 Cd nuclei, an<br />

HOPG sheet was irradiated with a very-low energy (30 keV) radioactive nuclear beam<br />

of A = 111 as parents. The observed TDPAC spectra were discussed in terms of lattice<br />

damages produced when introducing the probe into the sample.<br />

As an extension of the present method, the proposal of RNB-KJ02 considers the<br />

possibility to trace the Li macroscopic behavior across the interface in hetero structural<br />

Li ionic conductors in micrometer scale, to be termed 8 Li microscope.<br />

The time spectra shown in Fig. 3-28 represent dynamical movement of Li in the<br />

sample between as-implanted position and surface during a cycle for measurement. For<br />

a single layer (diffusion in homogeneous sample) as discussed in case of Fig.3-26, the<br />

present method can trace, most efficiently, the Li movement within one-dimensional<br />

distance of about 10 µm for about 7 s. For slow diffusion, Li is still moving toward the<br />

surface for the time of measurement, the α-particle yields are monotonically increasing<br />

with time. For moderate diffusivity, a maximum is observed when Li is reflected by the<br />

surface. After the maximum, the yield is simply decreasing with time since Li is<br />

diffusing into the bulk. For double layers (hetero structural sample) whose interface<br />

exist in between, i.e. introducing an interface between the as-implanted position and the<br />

surface, we could observe time structure different from the case of single layer. In other<br />

words, we could observe how Li interacts with the interface, e.g. if Li is precipitated,<br />

perfect or half reflected on the interface. This idea could be applied to take a dynamical<br />

picture of Li in Li ion micro-batteries consisting of thin films of several µm in thickness.<br />

Therefore, the present method could be called 8 Li microscope by analogy with the<br />

neutron transmission image of Li in a secondary Li ion battery, where the picture of Li,<br />

actually 6 Li, in the battery was taken by the neutron radiography with a resolution of<br />

mm [3-46].<br />

For higher sensitivity, from microscale to nanoscale, the proposal of RNB-KJ02<br />

also considers a coincident measurement of two α-particles emitted from position of<br />

β-decaying 8 Li. The micrometer sensitivity of the method discussed so far is partly<br />

coming from the broad energy distribution of α-particles on decaying. Therefore, the<br />

coincident measurement is supposed to dramatically improve the sensitivity since the<br />

87


coincident α-particles have the same energy at the decaying position. In diffusion<br />

tracing method by 8 Li, however, a special attention is paid on the energy loss of<br />

α-particles subjected to the actual path length in the sample of interest from the<br />

decaying to the detection positions. For diffusion in a nanoscale, the diffusion length is<br />

too short to give a significant change in the energy loss of α-particles. We further apply<br />

a limitation in emission (detection) angles to the coincident measurement as shown in<br />

Fig. 3-33. With small emission angles, the actual path length of α-particles in the<br />

sample of interest can be made significantly longer than implanted depth, giving rise to<br />

a considerable energy loss difference against a nanoscale diffusion length toward the<br />

surface. Along with the idea, we have performed a simulation to examine the feasibility<br />

and search for an observable most sensitive to diffusion profiles with a diffusion<br />

coefficient of 10 -12 cm 2 /s, a specific goal of present consideration. The simulation was<br />

performed in a similar way discussed in Ref. 3-34, by using the energy loss and<br />

straggling of α-particles provided by the SRIM-code [3-47]. We assumed the diffusion<br />

coefficient as 10 -12 cm 2 /s, the sample thickness as 50 nm, the emission angle as 5 o , and<br />

the implantation energy as 1 keV.<br />

Fig. 3-33. Schematic layout for a coincident measurement of two α-particles emitted with angles of<br />

θ1=θ2 relative to the surface of sample. The actual path lengths (L1 and L2) in the sample are<br />

enhanced 10 times as compared to those considered in Fig. 3-26 when applying limitation to the<br />

emission angle of 6 o .<br />

As a result of the simulation shown in Fig.3-34, we found that the energy<br />

difference between two coincidently measured α-particles could provide nanoscale<br />

sensitivity. The concentration profiles of 8 Li simulated sequentially with a condition<br />

given above are shown in Fig. 3-34(a); the as-implanted profile is broadening with time<br />

by diffusion. Simulated in the same time sequence as done for profiles, the spectra of<br />

energy difference of two α-particles to be measured coincidently are compared in Fig.<br />

3-34(b). There is a good correspondence between the diffusion profiles and energy<br />

88


difference spectra, more specifically a clear one-to-one correspondence between<br />

decaying position of 8 Li and energy difference of two coincidently measured α-particles.<br />

For example, looking at the time evolution of the counts of coincident events with zero<br />

energy difference in the energy difference spectra should be a good measure of the time<br />

when 8 Li is across the middle of the sample, because the zero energy difference means<br />

that the path lengths experienced by two coincident α-particles are identical. We can<br />

conclude that the sensitivity against the diffusion of 10 -12 cm 2 /s could be easily achieved<br />

by the coincident measurement of two α-particles with the geometry assumed in the<br />

simulation.<br />

Fig. 3-34 (a) Simulated concentration profiles of 8 Li implanted into LiCoO2 with a thickness of<br />

50nm with an energy of 1keV and their evolution by diffusion simulated at 1, 2, 3 and 4 seconds<br />

after implantation with a diffusion coefficient of 10 -12 cm 2 /s. (b) Spectra of energy difference of<br />

α-particles coincidently emitted at angles of 5 o relative to the surface of the sample. They are<br />

respectively simulated in the same time sequence as done in Fig. 3-34 (a).<br />

The proposal of KJ02 has met difficulties in the fabrication of the samples with an<br />

appropriate thickness and surface roughness, especially for the coincidence<br />

measurements of two α-particles where the sample is required to have a form of a<br />

self-support foil with a thickness of about 100 nm. After all considerations, the proposal<br />

has been modified to confirm the feasibility with a reduced sensitivity (e.g. 10 -10 ~10 -11<br />

cm 2 /s), as a simple extension of the previously developed method, by focusing the<br />

time-evolution of the energy spectra of α-particles emitted with a large angle from<br />

β-decaying 8 Li implanted with an energy of about 10 keV. An experimental set-up was<br />

newly fabricated and is ready for an experiment to be carried out soon.<br />

89


3-4. Experiments with the ISOL beam (without re-acceleration by the <strong>TRIAC</strong>)<br />

The ISOL ion sources developed for the <strong>TRIAC</strong> could provide good opportunities<br />

to perform some experiments by using the beam without re-acceleration. In this<br />

subsection, two experiments for nuclear decay-spectroscopy study using the SIS with<br />

UCx, and an experiment using the FEBIAD with an enriched 96 Mo target for materials<br />

science are described.<br />

3-4-1. Qβ measurements of 160-165 Eu and 163 Gd<br />

Fig. 3-35. Fermi-Kurie plots of the measured total absorption spectra [3-48]. The data points used<br />

for fitting and adopted lines are shown by closed circles and solid lines, respectively. The endpoint<br />

energies and the deduced Qβ values are given in the figures.<br />

90


Mass, which is one of the most fundamental quantities for atomic nuclei, provides<br />

good insight on the evolution of the nuclear structure. In addition, the mass is one of the<br />

key parameters to build the scenarios of nucleosynthesis in universe. The mass of<br />

160-165 163<br />

Eu and Gd isotopes have been determined for the first time by measuring Qβ<br />

values using a total absorption BGO detector. Among them, isotopes of 163-165 Eu were<br />

newly observed.<br />

The isotopes produced by proton-induced fission of 238 U of about 600 mg/cm 2<br />

thickness were ionized by the SIS and mass-separated by the JAEA-ISOL. The isotopes<br />

were implanted onto an aluminum-coated Mylar tape, and were periodically moved to<br />

the center position of the total absorption BGO detector. The detector composed of twin<br />

BGO scintillation detectors of 120 mm in diameter and 100mm in length which faced<br />

each other with a distance of 4 mm. The solid angle of the detector was 96 %. For<br />

measurements of 160-165 Eu, the time interval of the tape transport was set to be two or<br />

three times longer than the half-lives of these isotopes, while, in the case of 163 Gd, the<br />

times for collection, cooling, and measurement were set by 4, 2, and 2 minutes,<br />

respectively, to allow the decay out of the parent nucleus 163 Eu (T1/2 = 6 s), which has a<br />

larger Qβ value than 163 Gd.<br />

The Fermi-Kurie plots of the total absorption spectra of 158, 159 Pm, 159, 161 Sm<br />

160-165 163 166<br />

Eu, Gd and Tb are shown in Fig. 3-35. Qβ values of 158, 159 Pm, 159, 161 Sm, and<br />

166<br />

Tb were found to be consistent with our previous experimental results. The evaluated<br />

Qβ values given by Audi et al. [3-49] were in agreement with the present results. The<br />

mass predictions by Duflo et al. [3-50] and Dussel et al. [3-51] were also consistent<br />

with the experimental values.<br />

3-4-2. Lifetime measurements of 162, 164 Gd (2 + )<br />

The collectivity is one of useful quantities to explore the vibration of the nuclear<br />

structure. In the case of the even-even nuclei, the transition probability from the first 2 +<br />

state reflects the quadrupole collectivity. The lifetimes of the first 2 + states at 71.5 keV<br />

in 162 Gd and 73.3 keV in 164 Gd were measured by means of β–γ delayed coincidence<br />

technique for mass-separated 162 Eu and 164 Eu isotopes.<br />

As described previously, the fission products of 162,164 Eu were ionized by SIS<br />

embedded UCx and transported to the tape system. The tape was moved every 20 s and<br />

9 s for 162 Gd and 164 Gd, respectively. The detection position was equipped with a<br />

Pilot-U 60 mm × 63 mm × 1-mm thickness plastic scintillator and a 38-mm diameter ×<br />

5-mm thickness BaF2 scintillator to detect β and γ rays, respectively.<br />

The time interval spectra for the first 2 + state in 162 Gd and in 164 Gd were obtained<br />

91


as shown in Fig. 3-36 (a) and (b), respectively. These spectra include the prompt time<br />

components from Compton events caused by higher-energy γ rays. The lifetimes were<br />

deduced by fitting an exponential decay curve, a exp(-t/τ)+b , by a least-χ 2 method.<br />

Contributions of lifetimes of higher-lying states are neglected because their lifetimes are<br />

expected to be much shorter than that of the 2 + states. The lifetimes were determined to<br />

be 3.98(8) ns and 4.0(2) ns, for 162, 164 Ga, respectively. Here, the errors were estimated<br />

from the spread of various fitted values obtained by changing the fitting range. In<br />

addition to these nuclei, the lifetime of the 2 + state in 160 Ga was also measured from β<br />

delayed γ rays from 140 Eu. The resultant lifetime was 3.94(12) ns, which was in good<br />

agreement with the previous result of 3.91(8) ns, well verifying the present experiment.<br />

The measured lifetime were converted to B (E2;0 + 2 + ) values by taking into<br />

account the total internal conversion coefficients αT taken from the ICC code in the<br />

program [3-52], and the energies of the first 2 + state. The obtained B (E2;0 + 2 + )<br />

values were deduced to be 5.45(11) e 2 b 2 and 5.2(3) e 2 b 2 for 162,164 Gd, respectively.<br />

Fig. 3-36. Time interval spectra between β rays and γ rays from the fist 2 + state in 162 Gd (a) and<br />

164 Gd (b), respectively [3-52]. The results of the fitting were demonstrated by the solid lines. See<br />

the text for the details.<br />

3-4-3. Time differential perturbed angular correlation in highly oriented pyrolytic<br />

graphite (K06)<br />

When a radioactive nucleus with a large quadrupole moment is implanted into a<br />

material, the emission of radiation is affected by the (neighboring) magnetic field and/or<br />

an electric field gradient at the implanted place. In addition, if the radionuclide emits<br />

two gamma rays in cascade via an intermediate state whose lifetime is in order of ns, the<br />

emission pattern reflects the dynamics of the internal material. In other words, we can<br />

deduce the nano-scale structure concerning the extranuclear charge distribution, and<br />

magnetic properties of the material by measuring the angular correlation between the<br />

92


gamma rays. 111 In, which decays to 111 Cd with T1/2 = 2.8 days by emitting two cascade<br />

gamma rays of 171 and 245 keV via an excited state of 85 ns, is well known as a probe<br />

complying with the above condition. By accelerating 111 In by the <strong>TRIAC</strong>, the<br />

implantation depth can be controlled, enabling us<br />

to investigate the properties of material at<br />

different depths. As the first step, we implanted<br />

111<br />

In into a highly oriented pyrolytic graphite<br />

(HOPG) with 30 keV employing only the<br />

JAEA-ISOL.<br />

An enriched 96 Mo target of 3 mg/cm 2<br />

thickness was irradiated with the 19 F beam of 95<br />

MeV from the JAEA tandem accelerator. 111 In<br />

produced by the fusion evaporation reaction, 19 F<br />

( 96 Mo, 2p2n), was implanted into a HPOG plate.<br />

Typical beam intensity was about 10 4 pps. After<br />

20-hours implantation, we measured the<br />

time-differential perturbed angular correlation<br />

(TDPAC) of the cascade gamma rays.<br />

The TDPAC spectra at four temperatures are<br />

presented in Fig. 3-37. The anisotropy of the two<br />

gamma rays is represented as R(t). Contrary to<br />

the expectation of the oscillation due to the<br />

electric quadrupole interaction between 111 In and<br />

the extranuclear charge distribution, fast spectral<br />

damping was observed at the respective<br />

temperatures. The larger final isotropy at 673 K<br />

implies the recombination and movement of<br />

111<br />

In in the materials. For further information,<br />

measurements at higher temperatures are<br />

necessary.<br />

References<br />

[3-1] M.J. Balbes et al., Nucl. Phys. A 584 (1995) 315.<br />

[3-2] Z. H. Li et al., Phys. Rev. C 71 (2005) 0528018(R).<br />

[3-3] H. Ishiyama et al., Phys. Lett. B 640 (2006) 82.<br />

[3-4] T. Hashimoto et al., Phys. Lett. B 674 (2009) 276.<br />

93<br />

Fig. 3-37. TDPAC spectra of 111 Cd<br />

( 111 In) in HPOG at 673, 473, 373, 298<br />

K, respectively.


[3-38] H. Sugai et al., Phys. Rev. B 52 (1995) 4050.<br />

[3-39] H. Hamanaka et al., Solid State Ionics 113-115 (1998) 69.<br />

[3-40] K. Kuriyama et al., Phys. Rev. B 33 (1986) 7291.<br />

[3-41] K. Nakamura et al., Faraday Discuss. 134 (2007) 343.<br />

[3-42] K. Kuriyama et al., Phys. Rev. B 33 (1986) 7291.<br />

[3-43] J. C. Tarczon et al., Mater. Sci. Eng. A101 (1988) 99.<br />

[3-44] K. Kuriyama et al., Phys. Rev. B54 (1996) 6015.<br />

[3-45] Y. Asano, Master’s Thesis, College of Engineering, Hosei University, Koganei,<br />

Tokyo, Japan, 2000.<br />

[3-46] S. Takai et al., Solid State Ionics, 176 (2005) 2227.<br />

[3-47] http://www.srim.org/<br />

[3-48] H. Hayashi et al., Eur. Phys. J. A 34 (2007) 363.<br />

[3-49] G. Audi, A.H. Wapstra, C. Thibault, Nucl. Phys. A 729 (2003) 337.<br />

[3-50] J. Duflo, A.P. Zuker, Phys. Rev. C 52 (1995) R23.<br />

[3-51] G. Dussel, E. Caurier, A.P. Zuker, At. Data Nucl. Data Tables 39 (1988) 205.<br />

[3-52] D. Nagae et al., AIP Conf. Proc. 1224 (2010) 156.<br />

95

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!