+ All Categories
Home > Documents > nuc_expt_phr03

nuc_expt_phr03

Date post: 28-Nov-2014
Category:
Upload: patrick-regan
View: 1,311 times
Download: 6 times
Share this document with a friend
225
Post Graduate Nuclear Experimental Techniques (4NET) Course Notes By Dr. Paddy Regan, Department of Physics University of Surrey Guildford, GU2 7XH, UK e-mail [email protected] October 2003
Transcript
Page 1: nuc_expt_phr03

Post Graduate NuclearExperimental Techniques

(4NET) Course Notes

By

Dr. Paddy Regan,

Department of Physics

University of Surrey

Guildford, GU2 7XH, UK

e-mail [email protected]

October 2003

Page 2: nuc_expt_phr03

Contents

1 Electromagnetic Probes of Nuclear Structure. 2

1.1 Gamma-Ray Decay Selection Rules. . . . . . . . . . . . . . . . . . 3

1.2 Internal Conversion Electrons. . . . . . . . . . . . . . . . . . . . . 4

1.2.1 Electric Monopole Decays. . . . . . . . . . . . . . . . . . . 7

1.2.2 Magnetic Monopoles. . . . . . . . . . . . . . . . . . . . . . 9

2 Studies of Nuclear Structure at High Angular Momentum. 10

2.1 Fusion Evaporation Reactions. . . . . . . . . . . . . . . . . . . . . 10

2.1.1 Beam Currents, Energies and Target Thicknesses. . . . . . 11

2.1.2 Compound Nucleus Excitation Energy and Maximum An-

gular Momentum. . . . . . . . . . . . . . . . . . . . . . . . 12

2.1.3 Compound Nucleus Decay. . . . . . . . . . . . . . . . . . . 16

2.1.4 Excitation Functions. . . . . . . . . . . . . . . . . . . . . . 17

2.1.5 Spin Assignments: Gamma-ray Angular Distributions. . . 19

2.1.6 Anisotropies and Gated Angular Distributions. . . . . . . . 20

2.1.7 DCO Ratios. . . . . . . . . . . . . . . . . . . . . . . . . . 23

2.2 Determining the Intrinsic Structure of Rotational Bands. . . . . . 24

2.2.1 Rotational Frequency, Moments of Inertia and Alignments. 25

2.2.2 Particle-Core Coupling. . . . . . . . . . . . . . . . . . . . . 29

2.2.3 Branching Ratios and g-Factors. . . . . . . . . . . . . . . . 32

2.2.4 Two State Mixing. . . . . . . . . . . . . . . . . . . . . . . 35

2.3 Selected Topics in High Spin Nuclear Structure. . . . . . . . . . . 39

2.3.1 Shape Coexistence and Superdeformation. . . . . . . . . . 39

2.3.2 Band Terminations. . . . . . . . . . . . . . . . . . . . . . . 41

2.3.3 High K-Isomers and Pairing Reduction. . . . . . . . . . . 41

2.3.4 Octupole Correlations. . . . . . . . . . . . . . . . . . . . . 44

i

Page 3: nuc_expt_phr03

2.4 Branching Ratios and g-Factors in High-K Bands. . . . . . . . . . 45

3 Experimental Gamma-ray Spectroscopy. 48

3.1 Germanium Semi-Conductor Detectors. . . . . . . . . . . . . . . . 48

3.2 Gamma-Ray Spectroscopy with Germanium Detectors. . . . . . . 49

3.2.1 Response Function of Germanium Spectra. . . . . . . . . . 49

3.2.2 Germanium Detector Efficiency. . . . . . . . . . . . . . . . 51

3.2.3 The Compton Suppressed Spectrometer (CSS). . . . . . . 52

3.3 Gamma-Ray Arrays. . . . . . . . . . . . . . . . . . . . . . . . . . 52

3.3.1 Resolving Power and Total Photopeak Efficiency. . . . . . 54

3.3.2 Add-Backs from Clover/Cluster Detectors. . . . . . . . . . 56

3.3.3 Polarization Measurements. . . . . . . . . . . . . . . . . . 56

3.3.4 Gamma-ray Tracking. . . . . . . . . . . . . . . . . . . . . 56

4 Channel Selection In Fusion-Evaporation Reactions. 63

4.1 Inner Multiplicity Sum-Energy Balls. . . . . . . . . . . . . . . . . 63

4.2 Studies of Very Neutron Deficient Nuclei. . . . . . . . . . . . . . . 65

4.2.1 Charged Particle Balls. . . . . . . . . . . . . . . . . . . . . 66

4.2.2 Kinematic Focussing and Conversion from Lab to COM

Energies. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69

4.2.3 Silicon Detectors. . . . . . . . . . . . . . . . . . . . . . . . 75

4.2.4 CsI(Tl) Balls Using Pulse Shape Discrimination. . . . . . . 78

4.3 Neutron Detection. . . . . . . . . . . . . . . . . . . . . . . . . . . 91

4.4 Recoil Detectors. . . . . . . . . . . . . . . . . . . . . . . . . . . . 100

4.4.1 Recoil Mass Separators. . . . . . . . . . . . . . . . . . . . 100

4.4.2 Gas Filled Separators. . . . . . . . . . . . . . . . . . . . . 107

4.4.3 Recoil Decay Tagging. . . . . . . . . . . . . . . . . . . . . 108

4.4.4 Recoil Filter Detectors. . . . . . . . . . . . . . . . . . . . . 109

5 Measurement of Lifetimes of Bound Nuclear States. 113

5.0.5 Weisskopf Single Particle Estimates. . . . . . . . . . . . . 114

5.0.6 Determining Nuclear Quadrupole Deformation from Life-

times of E2 Transitions. . . . . . . . . . . . . . . . . . . . 115

5.1 Electronic Timing Methods. . . . . . . . . . . . . . . . . . . . . . 116

5.1.1 Gamma-ray Spectroscopy Across Isomers . . . . . . . . . . 120

ii

Page 4: nuc_expt_phr03

5.2 The Recoil Distance Method. . . . . . . . . . . . . . . . . . . . . 124

5.2.1 Feeding Corrections and Gating From Above. . . . . . . . 128

5.2.2 The Differential Decay Curve Method. . . . . . . . . . . . 133

5.3 The Doppler Shift Attentuation Method. . . . . . . . . . . . . . . 136

5.3.1 Lineshape Analysis. . . . . . . . . . . . . . . . . . . . . . . 139

6 Measurement of Magnetic Moments. 142

6.1 Measurement of Nuclear Magnetic Dipole Moments. . . . . . . . . 144

6.1.1 Corrections in the Ion-Implantation Perturbed Angular Dis-

tribution Technique. . . . . . . . . . . . . . . . . . . . . . 146

6.1.2 Analysis of Precession Data with Limited Angles. . . . . . 148

6.1.3 Transient Field Measurements. . . . . . . . . . . . . . . . 149

6.1.4 Time Differential Perturbed Angular Distributions. . . . . 150

7 Spectroscopy of Neutron Rich Nuclei. 153

7.1 Using Fusion Evaporation Reactions . . . . . . . . . . . . . . . . 153

7.2 Incomplete Fusion/Massive Transfer Reactions. . . . . . . . . . . 154

7.3 Deep Inelastic Reactions. . . . . . . . . . . . . . . . . . . . . . . . 156

7.3.1 Maximum Angular Momentum in DIC. . . . . . . . . . . . 159

7.3.2 Useful Formulae for Binary Reaction Studies. . . . . . . . 160

7.3.3 Doppler Correction. . . . . . . . . . . . . . . . . . . . . . . 164

8 Spectroscopy With Radioactive Ions Beams. 170

8.1 Production of Radioactive Beams. . . . . . . . . . . . . . . . . . . 171

8.1.1 Projectile Fragmentation. . . . . . . . . . . . . . . . . . . 171

8.1.2 Particle Identification in Fragmentation. . . . . . . . . . . 173

8.1.3 Isomeric Ratios and Angular Momentum Population. . . . 184

8.1.4 Projectile Fission . . . . . . . . . . . . . . . . . . . . . . . 188

8.1.5 Intermediate Energy Coulex. . . . . . . . . . . . . . . . . . 189

8.1.6 Double Fragmentation and In-beam spectroscopy . . . . . 190

8.1.7 Beta-Decay Measurements. . . . . . . . . . . . . . . . . . . 190

8.2 ISOL Based Techniques. . . . . . . . . . . . . . . . . . . . . . . . 190

8.2.1 On-Line Mass Separators. . . . . . . . . . . . . . . . . . . 191

8.2.2 The 19Ne∗+40Ca Experiment at Louvain La Neuve. . . . . 198

1

Page 5: nuc_expt_phr03

Chapter 1

Electromagnetic Probes ofNuclear Structure.

The focus of this course is the study of nuclear structure by the measurement

of particle bound nuclear state decays. These decays proceed mainly via elec-

tromagnetic interactions and can be investigated by detecting the emitted elec-

tromagnetic radiation as the state de-excites, ie. through the emission of either

gamma-rays or conversion electrons. While the strong force is the dominant in-

teraction within nuclei, the EM interaction is an excellent probe since (a) it is well

understood and (b) its weak nature compared to the strong force means that it

does not petrurb the system very much. The course will be based on methods of

exciting nuclei via interactions between an energetic beam of ions and a metalic

foil. In particular, we will look at the study of the nucleus at high values of angu-

lar momentum and exotic proton to neutron ratios. We will investigate methods

of producing exotic nuclear matter ‘in-beam’ and different methods of selecting

specific nuclei for study from a large ‘background’ of other nuclear species which

may be produced.

We can measure basic nuclear properties of excited nuclear states such as

excitation energy, angular momentum (spin) and parity using conservation laws

and electromagnetic selection rules. In addition, as we shall see later, the deter-

mination of decay probablilities of nuclear states (ie. their lifetimes) gives direct

information on the make-up of the initial and final states and can reveal highly

collective, deformed structures within the nucleus.

2

Page 6: nuc_expt_phr03

1.1 Gamma-Ray Decay Selection Rules.

For a gamma-ray decay between states of initial spin Iπi and final state spin Iπ

f ,

the gamma-ray selection rules are that the decay can proceed by a photon of

multipole order L where,

|Ii + If | ≥ L ≥ |Ii − If | (1.1.1)

Note that because the intrinsic spin of the photon is 1h, gamma-ray transitions

from 0+ → 0+ states are forbidden. These transition decay by electron conversion

and/or internal pair formation (if the transition energy is above 1.022 MeV).

M3, E4, M5,E6, M7, E8

5

3

4

+

+

+

("mixed")

(‘pure’)E2

M1, E2("mixed")

M7, E8, M9M3, E4, M5, E6,

M1, E2,

M3, E4, M5, E6, M7

-3

4

2

+

+

E1, (M2)

E3, M4, E5, M6, E7E2

M3,....,E6

E1, (M2)E3, M4, E5

(b)

(a)

Figure 1.1: Schematic decay scheme showing the effect of gamma-ray selectionrules on allowed multipolarities.

The parity of a magnetic transition of multipole order L is given by π =

3

Page 7: nuc_expt_phr03

(−1)L+1, while that of an electric transition is given by π = (−1)L.

The transition probability for a state decaying from state Ji to state Jf ,

separated by energy Eγ, by a transition of multipole order L is given by [1, 7]

Tfi(λL) =8π(L+ 1)

hL ((2L+ 1)!!)2

(

hc

)2L+1

B(λL : Ji → Jf ) (1.1.2)

where B(λL : Ji → Jf) is called the reduced matrix element.

As figure 1.1 shows schematically, typically, the lowest multipolarity transi-

tions dominate the decays. This is an effect of the differing transition probabilties

for different multipoles (see chapter on measuring nuclear lifetimes).

The transition probability for a mixed multipolarity transition (usually re-

stricted to M1/E2 decays for in-beam decays) can be calculated in terms of the

multipole mixing ratio, δ [1, 2]. The mixing ratio for ∆I=1, parity non-changing

transitions is given by the ratio of the reduced matrix elements for the E2 and

M1 components. This is related to their partial transition probabilities, T , by

the simple equation [2]

δ2E2/M1 =

T (E2 : J− > J − 1)

T (M1 : J− > J − 1)(1.1.3)

The experimentally measured branching ratio for competing ∆I = 2 (E2)

and ∆I = 1 and (M1/E2) transitions is related to the ratio of reduced transition

probabilities (B(M1) and B(E2)) by the expression [3, 4]

B(M1)

B(E2)= 0.697

E52

E31

1

1 + δ2E2/M1

Iγ(∆I = 1)

Iγ(∆I = 2)(1.1.4)

where Iγ are the experimentally measured gamma-ray intensities for the com-

peting transitions. (Note the ∆I = 1 intensity contains both M1 and E2 admxi-

tures).

1.2 Internal Conversion Electrons.

A competing process to gamma-ray emission in the decay of bound nuclear states

is internal conversion where an atomic electron is emitted. Here the EM field

of the nucleus interacts with an atomic electron and the energy released by the

nuclear decay is transferred to the electron causing it to be ejected from the

atom. The electron is released with a kinetic energy equal to the energy difference

4

Page 8: nuc_expt_phr03

between the nuclear states minus the atomic binding energy for the electron shell

from which it was emitted. Thus the kinetic energy of the conversion electron is

given by,

Ee− = Eγ − B.E. (1.2.5)

where Eγ is the energy of the competing gamma decay and B.E. is the electron

binding energy.

202Po

counts

transition energy (K) in Po [keV]

electrons

400 600 800 10000

2000

4000

385.7

442.7

571.2

526.2

676.8

912.1

gammas

2x10

0

400

800delayed (35 - 550 ns)

385.7

442.7

571.2

676.8

526.2

912.1

Figure 1.2: Experimental internal conversion and gamma-ray spectra for transi-tions in 202Po. The 386, 443, 571 and 677 are E2 decays, the 526 is an E1 andthe 912 is an E3.

Figure 1.2 shows a comparison of electron and gamma-ray spectra for decays

in 202Po [8]. Note that while the gamma-ray transitions are single lines, the

electrons come in groups. This reflects the fact that the electrons come out with

different energies depending on which shell they are emitted from. Electrons

from the 1s, or K-shell are most likely to be emitted, and are most bound, thus

causing them to have the lowest energy. Electrons from the L and M shells can

also be observed with higher energies (since their binding energies are less).

The total decay intensity from a particle bound state is given by I where,

5

Page 9: nuc_expt_phr03

Experimental ICC’s (T. Kibedi et al., Australian National University)

202Po

ICC [K shell]

Transition energy [keV]

200 400 600 800 1000 1200-310

-210

-110

010

E1

M1

E2

M2

E3

Figure 1.3: Experimental internal conversion coefficients for transitions in 202Po.These data were taken from the reaction 194Pt(12C,xn)202Po at a beam energy of76MeV [8].

I = Iγ +∑

i

Iec = Iγ

(

1 +∑

i

αi

)

(1.2.6)

where αi is the internal conversion coefficient for the ith electron shell. Inner

shell electrons (K,L,M) are more likely to be converted than outer lying ones, as

long as the energy of the transition is greater than the electron binding energy

for that shell.

Experimentally, electron conversion coefficients are very useful as they are

dependent on the multipolarity of the transition [10, 11] and can thus give infor-

mation on the spin and parity of nuclear states. Transition multipolarities can be

assigned by either measuring the absolute internal conversion coefficients αi = Γe

Γγ

or the ratio of the partial conversion coefficients (eg. αM

αL).

In the case of mixed multipolarity ∆I=1 transitions, the experimentally de-

termined value of the electron conversion coefficient directly gives the magntitude

of the E2/M1 mixing ratio. Figure 1.3 shows the experimentally determined in-

ternal conversion coefficients for 202Po determined from the spectra shown in

figure 1.2.

The size of the electron conversion coefficient increases with (a) decreasing

transition energy, (b) increasing Z of the nucleus and (c) increasing multipolarity.

The decays from isomeric (long lived) nuclear states are often accompanied by a

low energy transition and/or a large change in multipolarity. Such decays can be

well studied using pulsed beam techniques by observing the conversion electrons

6

Page 10: nuc_expt_phr03

emitted between the beam bursts (see later).

The emission of a conversion electron results in an electron vacancy being

filled and the subsequent emission of a characteristic X-ray which can be used to

identify the proton number of the nucleus of interest.

The internal conversion coefficients for electric (E) and magnetic (M) multi-

poles can be calculated using the following expressions [5],

α(EL) ≈ Z3

n3

(

L

L+ 1

)

(

e2

4πǫohc

)4 (2mec

2

)L+ 5

2

(1.2.7)

α(ML) ≈ Z3

n3

(

e2

4πǫohc

)4 (2mec

2

)L+ 3

2

(1.2.8)

Note that there has been a recent report of the first example of ‘bound state’

internal conversion, where the electron is not emitted from the atom, but rather

raised to a higher lying atomic bound state [6].

1.2.1 Electric Monopole Decays.

Electric monopole decays (E0) between two 0+ states decay only by electron

conversion (and/or internal pair formation for transition energies greater than

1.022 MeV).

The total transition probability for E0 decays is given by [12] is given by

Γ(E0) =1

τ= ρ2

(

0+i → 0+

f

)

j

Ωj (Z, k) (1.2.9)

where τ is the (partial) lifetime for the E0 decay, ρ is the (dimensionless)

monopole strength parameter and Ωj are the electronic factors (analagous to

internal conversion coefficients) [13]. The electronic factors are tabulated in ref-

erence [13] and depend on the Z of the nucleus and the energy of the transition.

In most cases, K-conversion dominates. The nuclear structure information is

contained in the monopole strength parameter which is defined by [12]

ρ =< 0+

f |∑

j ejr2j |0+

i >

eR2=< 0+

f |m(E0)|0+i >

eR2(1.2.10)

where R is the nuclear radius (1.2A1

3 fm) and m(E0) is the electric monopole

operator.

7

Page 11: nuc_expt_phr03

Figure 1.4: Out of beam decay spectra for the reaction 144Sm(33S,p2n)174Ir at 153MeV. Non-yrast E0 decays are observed in 174Os which was populated via thebeta-decay of 174Ir. The beam was incident on a 1.3 mg/cm2 target and irradiatedfor 4 seconds followed by a 4 sec measuring cycles. The lines in 174Os correspondto the following decays: 546 keV, 0+ → 0+, pure E0 ; 532 keV, 2+ → 2+, mixedE2+M1+E0; and 555 kev 4+ → 4+, mixed E2+M1+E0. Note the absence of a532 keV line in the gamma-ray spectrum [16].

The single particle units for E0 decays are given by [17],

ρ2sp = 0.5A− 2

3 (1.2.11)

where A is then nuclear mass number. This gives a useful scaling of E0

strengths, independent of mass number.

The lifetime of E0 decays can be used to infer the degree of mixing and/or

the change in deformation between two 0+ configurations [12, 14, 15, 17] and is

a very useful tool in the study of shape coexistence in nuclei.

The E0 matrix element can also be used to measure the admixture of differ-

ent nuclear states with different radii (ie. different deformation). The electric

8

Page 12: nuc_expt_phr03

monopole operator can be expanded [18, 296] in terms of the quadrupole and

triaxial deformation parameters β and γ respectively such that

M(E0) =(

3Z

)

[

5+ β2 +

(

5(√

5)

21√π

)

β3cosγ

]

(1.2.12)

In the limit of simple two-state mixing between configurations with deforma-

tions γ1, β1 and γ2, β2, if a is the mixing amplitude between the configurations,

the resulting monopole strength is given by

ρ2(E0) =(

3Z

)2

a2(

1 − a2)

[

(

β21 − β2

2

)

+

(

5√

5

21√π

)

(

β31 cos γ1 − β3

2 cos γ2

)

]2

(1.2.13)

Most observed 0+ → 0+ E0 decays are between states where at least one of the

states is predominantly spherical in nature [12, 14] and it is usual to keep terms

only up to order β2 in equations (2) and (3). However, it has been suggested [15]

that in the case of prolate oblate mixing, the second term may become important

since the first vanishes for equal deformations of opposite sign.

Note that E0 decays also occur between nuclear states with the same spin

and parity (ie. Jπ → Jπ), although these will also compete with higher mutipole

gamma-decays. A review of E0 decays can be found in [17].

1.2.2 Magnetic Monopoles.

In principle, decays from 0− to 0+ states could decay by magnetic monopole type

transitions [19]. Such decays are forbidden to decay by photon emission but it

has been proposed [20] that they may decay via mono-energetic electron emission

by a cascade of virtual E1 and M1 pairs. To date this decay mechanism has not

been observed, although several searches have been made, eg. [19].

9

Page 13: nuc_expt_phr03

Chapter 2

Studies of Nuclear Structure atHigh Angular Momentum.

In order to fully understand the physics of the nucleus, one needs to examine the

effect of extreme conditions, (such as high temperature/excitation energy, distor-

tions of the nuclear shape, exotic proton to neutron ratios and large rotational

stresses), on nuclear matter. This chapter will deal with the study of high values

of angular momentum on the nucleus. We shall examine the formation of high

spin states via the mechanics of fusion-evaporation reactions and look at some

of the ways of analysing the spectroscopic information gained in such pursuits to

characterise the different nuclear structures observed in the decay of high spin

states.

2.1 Fusion Evaporation Reactions.

In order to study high spin states, one requires a reaction which will impart the

largest possible angular momentum into the nucleus of interest. Figure 2.1 shows

schematically different types of nuclear reaction depending on the value of the

impact parameter, b.

Fusion-evaporation reactions are the best way experimentally of producing

high spin states with large cross-sections. In a fusion-evaporation reaction, the

kinetic energy of the collision in the centre of mass frame is converted into ex-

citation energy of the compound system. The amount of angular momentum

transferred into the compound nucleus is given by b× p where b is the impact

parameter and p is the linear momentum, mv, of the beam. The angular momen-

10

Page 14: nuc_expt_phr03

elastic (Rutherford)scattering

b

Rfusion

inelastic scattering(Coulex)

DIC

Figure 2.1: Various types of heavy-ion collisions as a function of impact param-eter.

tum transfered is simply l = mvb. Thus, the higher energy the beam particles,

the more angular momentum will be transfered into the compound system. Note,

however, that as figure 2.1 shows, fusion reactions only occur for small values of

impact parameter, with other nuclear reactions occuring at increased target-beam

distances.

2.1.1 Beam Currents, Energies and Target Thicknesses.

Due to the small size of the nucleus, most of the beam particles simply ‘miss’

the target nuclei. Total fusion cross-sections are usually of the order of 1 barn

(10−28 m−2) for beam energies around the Coulomb barrier (∼ 3 → 5 MeV/A).

(Note the geometrical area for the reaction of two nuclei will be approximately(

1.22(A1 + A2)1

3

)2fm2). This fusion cross-section drops dramatically for heavier

nuclei where fission begins to dominate over fusion-evaporation. Typical beam

currents for fusion-evaporation experiments are of the order of a few particle

nano-amps (∼1010 particles per second) and are used to bombard relatively thin

target foils of thicknesses of the order of 1 mg/cm2 (dividing by the target density

gives the physical thickness).

Higher beam currents can be used to study nuclei at energies below the

Coulomb barrier, but a limit is usually set by either (a) the production of the

machine supplying the beam or (b) the deadtime of the data acquisition elec-

11

Page 15: nuc_expt_phr03

tronics/detectors in the experiment. Targets often have a thicker gold or lead

backing to stop the recoiling nuclei within the view of the detector system. The

choice of these high-Z stoppers is due to their higher Coulomb barrier which re-

duces the likelihood of beam-induced fusion events in the backing. Also, higher Z

stoppers cause the nucleus to slow down faster. The beam and target assemblies

are housed in high vacuum of around 10−6→7 Torr. A beam stop, or Faraday cup,

is usually placed downstream, behind the target and acts as a monitor for the

beam current.

Figure 2.2 gives a schematic set-up of a typical, thick target fusion-evaporation

experiment.

radiation detectors(gammas, e-)

Pb/Au stopper

targetthin

~ E-7 Torr beam stop/Farady cup

vaccum vessel

beamI~1 pnA outgoing beam

Fusion ~ 1mb

Figure 2.2: Schematic of a typical ‘in-beam’ set-up for the study of high-spinstates using a fusion evaporation reaction.

2.1.2 Compound Nucleus Excitation Energy and Maxi-mum Angular Momentum.

For fusion to occur the beam nuclei must have sufficient kinetic energy to over-

come the Coulomb repulsion between the two positively charged nuclei. Fusion-

evaporation reactions require the formation of a compound nucleus. This de-

scribes a hot nuclear system which lives long enough (> 10−20 s) for thermody-

namic equilibrium to occur, during which time the compound system ‘loses its

12

Page 16: nuc_expt_phr03

memory’ of how it was formed in terms of the make up of the target and projectile

nuclei [1]. However, quantities such as total energy and angular momentum are

conserved. By conservation of energy, the compound nucleus will be formed at

an excitation energy which depends on the centre of mass kinetic energy of the

collision and the Q-value for compound nucleus formation such that

13

Page 17: nuc_expt_phr03

Eex = Ecm +Qfus (2.1.1)

Ecm is the kinetic energy of the collision which is transfered to the compound

system. It can be calculated by taking the kinetic energy of the beam, EB and

subtracting the kinetic energy of the recoiling compound system, ER. Thus

Ecm = EB − ER (2.1.2)

By conservation of momentum, for beam and target masses of MB and MT

respectively, the velocity of the recoiling compound, VR can be calculated using

MBVB = (MT +MB)VR (2.1.3)

and by conservation of energy,

Ecm = EB − 1

2(MT +MB)V 2

R (2.1.4)

substituting in for VR, and recalling that EB = 12MBV

2B, we obtain

Ecm = EB

(

1 − MB

MT +MB

)

(2.1.5)

The maximum angular momentum that can be transferred in a fusion-evaporation

reaction will occur when the two nuclei are just touching in a peripheral collision.

(This is the so-called sharp cut-off approximation which assumes the nuclei are

‘hard spheres’ without a diffuse surface). The fusion cross-section will be a sum

of partial waves (depending on the size of the impact parameter). In the sharp

cut off approximation, the assumption is that the transmission coefficient Tl for

nuclear penetration falls to zero for l > lmax and has a value of 1 for l ≤ lmax.

Thus, the total fusion reaction cross-section, σf can be written as a sum of partial

waves upto lmax such that [1]

σf = π

(

λ

)2 lmax∑

l=0

(2l + 1)Tl ≈ π

(

λ

)2

(lmax + 1)2 (2.1.6)

where λ is the wavelength of the entrance channel given by

λ =h

2√

2Ecmµ(2.1.7)

14

Page 18: nuc_expt_phr03

Ecm is the kinetic energy of the collision in the centre of mass and µ is the

reduced mass of the system such that µ = ABAT

AB+ATwhere AB and AT are the

masses of the beam and target nuclei respectively.

The value of lmax, calculated using the reduced mass of the system, µ =MT MB

MT +MB, and from conservation of energy and angular momentum is given by [1]

hlmax = µvR (2.1.8)

where the velocity v, can be calculated using conservation of energy in terms

of the kinetic energy of the collision in the centre of mass and the Coulomb barrier

(Vc), by the expression

1

2µv2 = Ecm − Vc (2.1.9)

Substituting in for v we obtain,

l2max =2µR2

h2 (Ecm − Vc) (2.1.10)

where R is the maximum nucleus-nucleus distance for which a reaction can

occur and is given (in fm) empirically by [1]

R = 1.36(

A1

3

B + A1

3

T

)

+ 0.5 (2.1.11)

The Coulomb barrier energy Vc (in MeV) is given by

Vc = 1.44ZBZT

R(2.1.12)

It is clear from equation 2.1.10 that those collisions which maximise the value

of the reduced mass (ie. symmetric reactions) will have the largest input angular

momentum for a given centre of mass energy.

The experimental data on compound nucleus reactions shows that at very

high bombarding energies, the angular momentum in the compound system is

somewhat less than given by lmax in equation 2.1.10. This is because for higher

energies, compound fusion formation can only occur for smaller impact parame-

ters (ie. not the peripheral collisions used to calculate lmax). The critical angular

momentum lcrit, is the maximum angular momentum for which fusion can occur

and can be estimated by the expression [1]

15

Page 19: nuc_expt_phr03

(

lcr +1

2

)2

=µ (RT +RB)3

h2

(

4πγRBRT

RT +RB

− ZBZT e2

(RT +RB)2

)

(2.1.13)

where γ ≈0.9 MeV fm−2 is the surface tension of the nucleus.

2.1.3 Compound Nucleus Decay.

It typically takes around 10−21→22 s for a beam nucleus to pass a target nucleus.

If the beam and target nuclei do interact and fuse together, thermodynamic

equilibrium occurs within about 10−20 second, after which the compound system

decays by either high energy gamma-ray emission (such as giant resonance de-

cays) and/or by nucleon evaporation, where neutrons, protons and α-particles are

emitted [1, 21]. Due to the effect of protons and alpha-particles having to tunnel

through the Coulomb barrier, charged particle emission is inhibited compared

to neutron evaporation for compound systems closer to stability. Once the com-

pound nucleus moves further to the neutron deficient side, the neutron separation

increases and the proton separation decreases allowing charged particle (proton

and alpha) emission to compete and often dominate over neutron evaporation.

nucleon separationenergy above yrast

yrast line(locus of yrast states)

gammas in residual nuclei

compound nucleus

Exc

itatio

n en

ergy

Angular momentum

4

3

2

1

E-20 secs

E-9 sec

E-15s

Figure 2.3: Schematic of the formation of high-spin residual nuclei from com-pound nucleus decay.

16

Page 20: nuc_expt_phr03

Due to the very high density of states in the highly excited compound system,

the evaporated particles have a statistical energy spectrum and reduce the exci-

tation energy of the compound system by around 5-8 MeV per nucleon, yet only

remove 1→2 h of angular momentum. Particle evaporation will continue until

the system reaches a state where the excitation energy is less than the particle

separation energy above the yrast line. The yrast state is the state of lowest

energy for a given value of angular momentum.

It takes around 10−15 seconds for the compound nucleus to decay into the

residual nucleus. Note that as figure 2.3 shows, the final nucleus created is de-

termined by the entry point (relative to the yrast line) in the excitation en-

ergy/angular momentum plane. Generally speaking, residual nuclei formed by

the emission of fewer evaporated particles have higher initial angular momenta

and excitation energy distributions than higher multiplicity evaporation channels.

As shown later, this effect can be used to experimentally select transitions from

specific evaporation channels using total energy/gamma-ray multiplicity detec-

tors.

The probability for a compound nucleus evaporating a particle (usually mean-

ing a proton, neutron or alpha particle) is proportional to the density of final

states and a barrier (Coulomb usually) transmission coefficient, given by [1]

T (li, Ep(i)) = exp

− −2h∆

(2mp (V − Ep))1

2

(2.1.14)

where V is the height of the barrier and ∆ is its width. Note that the shape of

the spectrum for the emitted particles is different for charged particles (protons

and alphas) compared to neutrons due to the effect of the Coulomb barrier. In

neutron deficient compound systems, the neutron separation energy is so high

(upto 15-20 MeV), that it is larger than the height of the Coulomb barrier at a

given excitation energy, so charged particle emission is then favoured.

2.1.4 Excitation Functions.

If an experiment is interested in studying a particular nucleus or set of nuclei, it

is usual to perform an excitation function to decide on the optimum beam energy

to maximise the cross-section and angular momentum input for the channel of

interest. Clearly, increasing the beam energy will both increase the maximum

17

Page 21: nuc_expt_phr03

50 60 70 80 90EBEAM (MeV)

101

102

103

104

105

106

σ (a

rb. u

nits

)

p4n (109

Ag)

p3n (110

Ag)

5n (109

Cd)

4n (110

Cd)

798 keV

249 keV

191 keV

125 keV

335 keV

523 keV

Figure 2.4: Excitation function for various products of the reaction 18O+96Zr[22].

input angular momentum, but will also increase the excitation energy of the

compound system, resulting in more particles being evaporated.

Figure 2.4 shows the relative intensity of various known gamma-ray transi-

tions from nuclei of interest in fusion of an 18O beam on a 96Zr target (forming

the compound nucleus 114Cd. Note how the relative intensity of the higher multi-

plicity channels such as the 5n to 109Cd and p4n to 109Ag increase with increasing

beam energy in the region of this excitation function, while the four particle out

channels to 110Cd and 110Ag peak at a beam energy of around 60 MeV, before

falling off. Note that the shape of the excitation function of all lines should be

similar for a given channel and thus, the variation of gamma-ray intensity of

a line with beam energy can be used to identify a transition with a particular

evaporation product.

18

Page 22: nuc_expt_phr03

2.1.5 Spin Assignments: Gamma-ray Angular Distribu-tions.

In order to determine the spins (and infer the parities) of excited nuclear states

formed in fusion-evaporation reactions, one can measure the angular distribution

of the gamma-ray transitions. In order to observe an anisotropic distribution

one needs to populate the nucleus in a way which gives rise to states of aligned

angular momentum with a specific orientation in space. This is achieved in fusion-

evaporation reactions, where the angular momentum vector (l = r × p) is in a

direction (to a good approximation) perpendicular to the beam direction (see

figure 2.5).

after evaporationsubstate alignment

beam direction

target

reaction plane

Figure 2.5: Schematic of initial orientation in fusion-evaporation reactions.

The orientation of the nucleus will be slightly attenuated by the emission of

evaporated particles (neutrons, protons and alpha-particles) and by the emission

of gamma-rays. The effect will be to provide a substate, or m-state alignment,

peaked symmetrically about the the m = 0 value corresponding to the reaction

plane.

The general formula for the angular distribution function is given by [1, 23]

19

Page 23: nuc_expt_phr03

W (θ) =∑

k

AkPk (cosθ) (2.1.15)

where W (θ) is the gamma-ray intensity measured at angle θ to the beam

direction; for parity conserving decays, such as gamma-ray emissions, k=even

numbers less than or equal to 2l where l is the angular momentum taken away

by the emitted photon; Pk (cosθ) are the standard Legendre polynomials; and Ak

is the angular distribution coefficient. The Ak value depends on the substate or

m-population distribution and the values of the initial and final state spins [23].

By measuring the intensity of a gamma-ray transition as a function of detec-

tor angle about the beam direction, a full angular distribution can be obtained

from which the values of A2 and A4 can be obtained by fitting the distribution to

equation 2.1.15. The Ak values can be used to experimentally distinguish between

transitions of different multipolarities [23, 24]. Similarly, fitting the experimen-

tally observed A2 and A4 coefficients for transitions of known multipolarity (and

possibly mixing ratios) gives a measure of the degree of substate alignment for

that spin. Generally, in fusion evaporation reactions, transition multipolarities

can be restricted to angular momentum values of 2 or less (ie. usually only E2,

M1 or E1 decays are observed).

For a pure dipole ∆I = 1 transition (E1), the angular distribution will be

given by

W (θ) = A0 1 + A2P2 (cosθ) (2.1.16)

where P2 (cosθ) = 12(3cos2θ − 1) and A0 is the ‘true’ intensity.

For a quadrupole (∆I=2) transition (E2), the distribution will have the form,

W (θ) = A0 1 + A2P2 (cosθ) + A4P4 (cosθ) (2.1.17)

where P4 (cosθ) = 18(35cos4θ − 30cos2θ + 3)

2.1.6 Anisotropies and Gated Angular Distributions.

The fixed (and often limited) angular granularity of modern gamma-ray ar-

rays coupled to the complexity of the singles spectra often observed in fusion-

evaporation reactions mean that a full angular distribution analysis may not be

viable. It is often enough to be able to tell the difference in the anisotropy between

20

Page 24: nuc_expt_phr03

∆I=1 and ∆I = 2 type transitions. In the case where only a few gamma-ray

detector angles are used, the A4 coefficient is set to a value of zero (eg. [25]).

Alternatively, coincidence data can be used to form ‘gated singles spectra’ at

specific detector angles, from which an anisotropy can be taken (eg [22]).

Pohl et al. defined one particular type of anistropy as

A = 2

(

W (37) −W (79)

W (37) +W (79)

)

(2.1.18)

where the angles correspond to detector positions in the Chalk River 8π

gamma-ray array.

-90 -60 -30 0 30 60 90arctan(δE2/M1)

-1.5

-1.0

-0.5

0.0

0.5

1.0

Ani

sotr

opy

∆ I = 0∆ I = -1∆ I = -2

Figure 2.6: Theoretical values for various transitions for the gamma-rayanisotropy as defined in reference [22].

Figure 2.6 shows the theoretically expected value for this anistropy for various

21

Page 25: nuc_expt_phr03

multipolarity decays (assuming A4 = 0). Note the different values of A2 for ∆=1

decays depending on the value of the E2/M1 mixing ratio.

Figure 2.7 shows the experimentally measured anisotropies for 108Pd and109Ag. The difference between pure ∆I = 2 (E2) and pure ∆I = 1 (E1) transi-

tions is clear. Mixed M1/E2 transitions are generally easy to identify for small

values of mixing ratio, but note that for large, positive values of the mixing ratio,

the anisotropy can not be distinguished from a pure E2 transition. The multi-

polarities for such transitions are assigned on the basis of other branches to the

same level in the decay scheme.

Figure 2.7: Experimentally measured value of anisotropy for 108Pd and 109Ag.Note the clear separation between ∆I = 2 and ∆I = 1 transitions [22]. Also,note the lower values of for stretched E2 transition in 108Pd resulting from a largerdestruction of reaction alignment associated with the emission of an α-particle.

22

Page 26: nuc_expt_phr03

2.1.7 DCO Ratios.

For coincidence measurements, the Directional Correlations from Oriented states

(DCO) method can be used to infer the spin differences between states observed

by the measurement of the gamma-decay between them [26, 27, 28]. Fusion-

evaporation reactions produce many gamma-ray transitions and consequently,

using singles data to obtain angular distribution/anisotropy data is usually only

useful for (a) strong transitions and/or (b) strongly populated reaction channels.

In obtaining high-spin data, it is more common to measure coincidence data,

using an electronic master gate of detecting at least two transitions. By gating

on a known transition in a given nucleus, much cleaner, angle gated spectra can be

obtained from which an anisotropy can be obtained. However, since the gamma-

ray detector angles are usually fixed in such experiments, the extra condition

that at least two gamma-rays must be measured introduces angular correlations

into the data, thus altering the observed angular distribution compared to true

singles data.

Since such reactions are still aligned, one can use these correlations to discrim-

inate between transitions of differing multipolarities. By measuring coincidence

data from detectors are different angles, once can construct sets of angle gated

gamma-gamma coincidence matrices, in offline sorting. These 2 dimensional spec-

tra correlate the intensities of coincidence transitions in a single cascades by the

angle at which the gamma-ray was measured. By setting a software gate on a

transition energy of known multipolarity (usually E2) on both axes and projecting

the spectra for the two angles, it is possible to distinguish transition multipolari-

ties by the intensity of the projected transition for given angles. Figure 2.8 shows

the gamma-ray spectra gated on a known E2 transition in 61Cu and projects

transitions of known E2 and E1 multipolarity. Note the difference in relative

intensity between the two spectra for the two types of decay.

By taking the ratio of the number of counts for a given transition in these two

spectra, a DCO ratio can be obtained, which can be used to tell the distinguish

between different multipolarities. Figure 2.9 shows the results of two different

types of DCO ratios (different detector angles used) for 106Cd, formed using two

different fusion-evaporation reactions. Note that in both cases, there is a clear

discrimination possible between the different types of transition.

23

Page 27: nuc_expt_phr03

Figure 2.8: Angle gated DCO correlation spectra showing the difference be-tween E2 and E1 transitions in 61Cu from the reaction 24Mg+40Ca→3p+61Cu(AYEBALL data [106]).

2.2 Determining the Intrinsic Structure of Ro-

tational Bands.

Using coincidence relationships and information from gamma-ray anisotropies

and DCO ratios, one can come up with a (complicated) decay scheme with many

different structures observed. Since it has been shown experimentally that yrast

or near yrast are preferentially populated in fusion evaporation reactions, it is

usual to assume that the spins of states increase with increasing excitation energy.

In order to understand the intrinsic make-up of the various structures ob-

served, the information contained in the decay scheme has to be used to construct

variables which one can compare with theoretical predictions.

A common treatment in the study of deformed nuclei is to take the exper-

imental information obtained on nuclear states (ie. their spins and excitation

24

Page 28: nuc_expt_phr03

Gamma-ray Energy (keV)0 200 400 600 800 1000 1200

0.5

1.0

1.5

∆J=1∆J=2 J->J

R’

R’’

∆J=1∆J=2 J->J

0.0

0.5

1.0

1.5

Figure 2.9: DCO ratios for transitions in 106Cd from (a) the CAESAR array and(b) the 8π array [3].

energies) and transform this into the intrinsic (rotational) frame of the nucleus.

2.2.1 Rotational Frequency, Moments of Inertia and Align-ments.

It can be intuitive to observe the effect of increasing/decreasing rotational fre-

quency on the structure of the nucleus.

Using a simple rotational model, the rotational frequency, ω, for a transition

between states of spin I + 1 and I − 1, with projections along the symetery and

rotation axes of K and Ix respectively, is given by the expression [30, 31, 32],

25

Page 29: nuc_expt_phr03

ω =dE(I)

dIx(I)≈ E(I + 1) −E(I − 1)

Ix(I + 1) − Ix(I − 1)(2.2.19)

The value of K is usually taken as being equal to the spin of the lowest energy

state in the band, known as the bandhead. The difference in the excitation energies

of the two states, E(I+1)−E(I−1), is simply the measured gamma-ray transition

energy, Eγ. The value of Ix can be calculated using Pythagoras theorem, such

that

Ix(I) =√

I(I + 1) −K2 ≈√

(

I +1

2

)2

−K2 (2.2.20)

substituting in to the expression for ω,

ω ≈ Eγ√

(I + 32)2 −K2 −

(I − 12)2 −K2

(2.2.21)

for decays from a state of spin I + 1 to one of I − 1. Note that at high spins

where I >> K, this expression simplifies to ω ≈ Eγ

2.

The quasi-particle aligned angular momentum, ix is the spin generated by the

breaking of the core, valence quasi-particles along an axis perpendicular to the

axis of symmetry. This is shown schematically in figure 2.10.

ωc

ix

i x

ω

R

K

I

j2

j1

j

Ix

R

jI

K

Ix

Figure 2.10: Schematic of components of angular momentum for a deformednucleus.

It can be approximated by taking the total angular momentum and subtract-

ing a fixed amount of spin due to the rotational motion of the ‘inert’, deformed

26

Page 30: nuc_expt_phr03

nuclear core. The quasi-particle aligned angular momentum ix, is a function of

rotational frequency and can be defined by

ix(ω) = Ix(ω) − Iref(ω) (2.2.22)

where Ix is the projection of the total angular momentum along the axis of

rotation and can be calculated using the expression,

Ix =√

I(I + 1) −K2 (2.2.23)

Iref is the reference angular momentum (of the core) which will be subtracted

and is calculated in terms of the Harris parameters [34], such that

Iref =(

I(0) + I(2)ω2)

ω (2.2.24)

The Harris parameters are usually fitted to states in the band to give a con-

stant alignment above the first band-crossing.

[h]

i//

h [MeV]// ω0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0

-2

0

2

4

6

8

10

12

14

16

18

106Cd band 1106Cd band 2106Cd band 3105Cd h band 11/2

106Cd band 4

Figure 2.11: Alignment plots for band structures in 106Cd and the νh 11

2

band in105Cd using Harris parameters of I0=7.0h2/MeV and I1=15.0h4/MeV3 [3].

27

Page 31: nuc_expt_phr03

Figure 2.11 shows the alignment plots as a function of rotational frequency for

structures in 106Cd and the rotational band built upon the neutron h 11

2

orbital in

the neighbouring 105Cd. At a rotational frequency of approximately 0.45 MeV/h,

band 1 in 106Cd shows a rapid decrease in rotational frequency and a large increase

of approximately 10h in aligned angular momentum. This effect is known as a

‘backbend’ [77, 78] and consitutes the crossing of a the ground state band with the

first two-quasi-particle band, which comes lower in excitation energy compared to

the simple ground state structure (ie becomes yrast) for spins above 10h. (This

is due to the two-quasi-particle band having a higher moment of inerta compared

to the ground state configuration due to a reduction in pairing correlations).

At higher spins still, there is another, smaller increase in alignment of approx-

imately 3 h in the same band. This is interpreted as the alignment of a pair of

neutrons occupying g 7

2

orbitals. If the structure of a band already contains on

of the ‘aligning orbitals’, the alignment can not proceed by the Pauli exclusion

principle and the bandcrossing is said to be blocked. Note that the h 11

2

band in105Cd already has a initial alignment of around 5h at the bandhead (from the

unpaired neutron), and thus the first (h 11

2

)2 alignment (or ‘crossing’) observed in106Cd is blocked in this structure.

The intrinic alignments are additive [31] and this is useful when quantifying

more complex structures. That is to say, at a given frequency

ix(12) = ix(1) + ix(2) (2.2.25)

where ix(12) is the alignment of the two-quasi-particle structure and ix(1) and

ix(2) are the intrinsic alignments of the single quasi-particle components, 1 and

2. Note that bands 2 and 3 in 106Cd, both the h 11

2

and g 7

2

alignments observed

in band 1 appear to be absent. This, together with the large, initial alignement

of these bands, may be taken as suggesting that bands 2 and 3 are two-quasi-

particle bands built consisting of both and unpaired h 11

2

and g 7

2

neutron coupled

together.

Moments of Inertia.

There are three types of moment of inertia used to describe high-spin rotational

structures, the static (I(0)), kinematic (I(1)) and dynamic (I(2)). The static

moment of inertia is defined by the simple relation,

28

Page 32: nuc_expt_phr03

Erot(I) =h2

2I(0)(I)I(I + 1) (2.2.26)

The kinematic moment of inertia is given by

I(1)(I) =I

ω(2.2.27)

while the dynamic moment is given by

I(2) =dI

dω≈ 4h

∆Eγ(2.2.28)

Note a calculation of the dynamic moment of inertia requires only the differ-

ence in transition energy between two decays and has no inherent spin depedence.

It is thus a useful quantity to use in cases where the decay out of a band is not

measured and the spin of the states not well established (for example in stud-

ies of super-deformed nuclei). The variation of the dynamic moment of inertia

with rotational frequency can also be used to determine the intrinsic structure of

rotational bands (see for example Fallon et al. [33]).

2.2.2 Particle-Core Coupling.

For axially symetric nuclei, rotational bands can be characterised by the single

particle excitations upon which the bandheads are built. The way that the the

odd-particle couples to the nuclear core gives rise to a number of effects (Coriolis

mixing and increased/decreased magnetic dipole moments) which show up in the

decay schemes observed for such structures [35].

29

Page 33: nuc_expt_phr03

Figure 2.12: Rotational properties of the yrast band in 106Cd, showing the back-bend due to the (11

2)2 neutron configuration crossing the yrast, ground state,

vacuum configuration, giving rise to a backbend. Note that the decrease in pair-ing correlations above the backbend gives an increased moment of inertia [3].

30

Page 34: nuc_expt_phr03

K+1

K+5

K+3

K+4

K+2

Khigh-j, low-KRotation alignedde-coupled bandLarge Coriolis mixing

Deformation alignedK bandheadInterband I=1 decaysStrongly coupled

K

j

j+5

j+3

j+1

j

j+2

j+4

(a) (b)

Figure 2.13: Schematic of decays schemes and descriptions of strong and weakcoupling of an odd-nucleon to a deformed nuclear core.

Figure 2.13 shows the two extremes of particle core coupling for deformed

nuclei. For equatorial orbits, the odd-particle has a large angular momentum

projection on the axis of rotation (ie. a large intrinsic alignment ix) and a corre-

spondingly small projection on the symmetry axis, or K value.

In the deformation aligned case, ix is small and K is large. The bandhead is

taken to have spin K and the spectra show both odd and even spins.

The strong coupling limit breaks down in the case where the Coriolis effects

are large. The Coriolis operator can be written by [36]

Hc = −2h2

I(

I.j− Ω2)

(2.2.29)

where j is the single particle angular momentum and Ω is its projection on the

nuclear axis of symmtery. In the case of weak coupling (coming from population

of high-j orbitals with small K projections), the rotation of the core is directly

coupled to the odd-particle, and indeed, the bands built on such structures have

energy spacings strikingly similar to the even-even neighbouring nuclei [35, 36].

The orbitals involved in decoupled bands are usually high-j, unnatural par-

ity states, which ‘intrude’ down from the next harmonic oscilator shell in the

31

Page 35: nuc_expt_phr03

spherical shell model as a consequence of the spin-orbit interaction. They are

thus known as intruder orbitals. In such, decoupled bands, the even-spins of the

core which coupled to the single particle are depressed in energy compared to the

odd-spins (corresponding to states where the odd-particle angular momentum is

not maximally aligned with the rotation axis). Since, these states will then be

non-yrast, they are usually not observed in fusion-evaporation reactions.

Magnetic Dipole Transition Strengths.

The strength of in-band B(M1) decays depends on the single particle structure

of the odd-particle. In the case where K 6= 12

and If = Ii − 1, the B(M1) decay

strength is given by [35, 235]

B(M1) =3

4πµ2

N (gK − gR)2K2 (I −K)(I +K)

I(2I + 1)(2.2.30)

where gK and gR are the single particle and collective g-factors respectively.

Thus, measuring the lifetime of a state which decays by a pure M1 decay allows

a measurement of the B(M1) and thus a deduction of (gK − gR) which will be

particle dependent.

These particle and core g-factors gK and gR, relate to the magnetic dipole

moment, µ, of a state of spin I with projection K by the expression [35, 88] by

µ = gRI + (gK − gR)K2

I + 1(2.2.31)

The value of gR, the g-factor of the core is approximated by the ratio of protons

to total nucleons in the nucleus (ZA). Thus, if the magnetic dipole moment can

be measured (see later) for a known spin, the K value of the state can be infered.

2.2.3 Branching Ratios and g-Factors.

The intrinsic structure of a rotational band can also be inferred in a model de-

pendent way by measuring the intensity on the ∆I=1 and ∆I=2 branches and

their energies [3]

δ2

1 + δ2=

2K2(2I − 1)

(I + 1)(I − 1 +K)(I − 1 −K)

(

E1

E2

)5 Iγ(∆I = 2)

Iγ(∆I = 1)(2.2.32)

32

Page 36: nuc_expt_phr03

where δ is the E2/M1 mixing ratio, K is the projection of the angular mo-

mentum on the axis of symmetry and I is the state spin. E1 and E2 are the

dipole and quadrupole transition energies in MeV and Iγ are the relative γ-ray

intensities.

The magnitude of the quantity |gK−gR|Q0

, (where Q0 is intrinsic quadrupole

moment, and gK and gR are the g-factors for the intrinsic state and collective

core respectively), is given by

|gK − gR|Q0

=0.93E1

δ√I2 − 1

(2.2.33)

The gK value can be approximated in the large deformation limit (ie. that

K remains a good quantum number) for a multi-quasi-particle state, using the

expression

KgK =∑

ΩgΩ =∑

ΛgΛ +∑

ΣgΣ (2.2.34)

where for Λ is the projection of the orbital angular momentum on the axis

of symmetry and Σ = ±12

is the single particle intrinsic spin projection. For

protons gΛ=1 and for neutrons gΛ=0. The gΣ values are simply the g-factors for

the single proton and neutron (+5.59 and –3.83 respectively), attenuated by a

factor of 0.6→0.8 to account for the fact that they exist in the nuclear medium

and are thus not ‘free’ [37].

Note that equation 2.2.33 only gives the magnitude of the value of gK − gR

and therefore there may be two different configurations which give branching

ratios consistent with the experimentally measured values. This anomaly can be

overcome, if one can obtain good angular distribution information on the ∆I = 1

interband transitions from which an A2 angular distribution coefficient can be

extracted and the sign of the mixing ratio inferred. The sign of the mixing ratio

is the same of the sign of the quantity gK−gR

Qo[39], ie.

sgn (δ) = sgn

(

gK − gR

Q9

)

(2.2.35)

Figure 2.14 shows a gK − gR analysis for the high-K band in 136Sm [38]. This

work assumed a value for the quadrupole moment of +4.5 eb (from the measured

deformation of the ground state band in this nucleus) and a negative value for the

mixing ratio, δ (from the measured angular distribution of the interband ∆I=1

33

Page 37: nuc_expt_phr03

Figure 2.14: Comparison of gK − gR values for the rotational band built on thehigh-K isomeric state in 136Sm. The observed branching ratios are consistentwith a Kπ=8− assignment and rule out the Kπ = 7− configuration [38].

transitions in the band). It demonstrates clearly that the Kπ = 8−, two-quasi-

neutron configuation is the only one consistent with the measured branching

ratios.

Donau and Frauendorf’s Geometrical Model.

Donau and Frauendorf [40, 41, 3] used a simplified geometrical model to estimate

the B(M1) values for a multi-quasi-particle band. For bands with no signature

splitting, the B(M1) value can be estimated (in units of µ2N) by,

B(M1) =3

K2

I2

[

(g(1) − gR)(√

I2 −K2 − i(1)x

)

−(

g(2) − gR

)

i(2)x

]2(2.2.36)

where the g are g-factors for specific particles and ix is the contribution to

the aligned angular momentum from that particle. The superscripts (1) and (2)

correspond to deformation aligned and rotation aligned particles respectively.

In the rotational model, the B(E2 : I → I − 2) is given by

B(E2) =5

16πQ2

o

3(I −K)(I −K − 1)(I +K)(I +K − 1)

(2J − 2)(2J − 1)J(2J + 1)(2.2.37)

34

Page 38: nuc_expt_phr03

where Q0 is the intrinsic quadrupole moment. For an unstretched (J → J−1),

E2 transitions, the B(E2 : I → I − 1) is given by

B(E2) =5

16πQ2

o

3K2(J −K)(J +K)

(J − 1)(J)(2J + 1)(J + 1)(2.2.38)

106CdB(M1)/B(E2)

Spin

19 21 23 25 27 29 31

010

110

210

( µ/eb)2

(h ) (g d )ν ν11/2 7/2 5/22

(h ) (g d ) (g )ν ν π11/22

7/2 5/2 9/22

(h ) (g )ν π11/22

9/22

(h ) (g g ) ν π11/22

9/2 7/2

Figure 2.15: Branching ratio comparison for different configurations in the mutli-quasi-particle structure in 106Cd. [3].

The experimental B(M1)B(E2)

branching ratios (in units ofµ2

N

e2b2) are given by the

expression

B(M1)

B(E2)= 0.697

E52

E31

1

1 + δ2

Iγ(∆I = 1)

Iγ(∆I = 2)(2.2.39)

As figure 2.15 shows, by comparing the experimentally observed B(M1)B(E2)

values

with those extracted using the B(E2)s from the rotational model and the B(M1)s

from the geometrical model, a specific assignment of a mutli-quasi-particle state

can be made.

2.2.4 Two State Mixing.

Often two nuclear configurations of the same spin and parity can not be described

in terms of exactly pure configurations but rather linear combinations of two

different configurations which wavefunctions φ1 and φ2 [164].

We can write the wavefunction for the final state in terms of admixtures of

the two basis states φ1 and φ2, ie

35

Page 39: nuc_expt_phr03

unperturbed basisstates

φ1

φ 2

αφ + βφ1 2Ψ =

Ψ = −βφ + αφ1 2

experimentally observed,mixed states

E1

E2

h1

h2

> 2V

Figure 2.16: Effect of mixing two basis states to form a mixed initial state.

Ψ = αφ1 + βφ2 (2.2.40)

The general way this is written is in the matrix form of the Schrodinger

equation such that

HΨ = EΨ (2.2.41)

In matrix form this may be written in terms of the unperturbed energies of

the basis states (h1 and h2) and the interaction which mixes the two states, (such

as pairing or Coriolis effects) v. Then

(

h1 v12

v12 h2

)(

αβ

)

= E

(

αβ

)

h1 and h2 are the unperturbed energies that the states would have if there was

no interaction between the two configurations, ie. the energies of the theoretical

states corresponding to the pure wavefunctions φ1 and φ2. The eigen values, E are

the experimentally observed energies of the two states and v12 is the interaction

matrix element which corresponds to the strength of the interaction between the

two configurations. α and β are the eigenvectors.

(

h1 − E v12

v12 h2 − E

)(

αβ

)

= 0

36

Page 40: nuc_expt_phr03

Since α and β are not zero, the determinant of the first matrix must equal

zero. In the case of simple two state mixing, the energy of the observed states (the

eigenvalues) are related to the unperturbed energies and the interaction matrix

element by solving the determinant

[

h1 −E v12

v12 h2 − E

]

= 0

which gives

(h1 − E) (h2 − E) − v212 = 0 (2.2.42)

This is a quadratic equation in E which has two roots (which are the two

energy eigen-values) given by

E =

(

h1 + h2

2

)

±

(

h1 − h2

2

)2

+ v212 (2.2.43)

The wavefunctions of the experimentally observed states can then be written

in terms of a linear combination of the basis states φ1 and φ2. The strengths

of the two final states in terms of the basis states weighting, α and β can be

calculated in terms of the h1, h2, E1, E2, and v12 by solving the expressions

α(h1 − E) + v12β = 0 (2.2.44)

for the higher energy state and

α(h2 − E) + v12β = 0 (2.2.45)

for the lower one and normalising with the condition that

α2 + β2 = 1 (2.2.46)

Gathering the terms and summarising, the two mixed states with energies

given by equation 2.2.43 have wavefunctions which can be written in terms of

Ψ =

(

1 ∓ Y

2

)

φ1 ±√

(

1 ± Y

2

)

φ2 (2.2.47)

where

37

Page 41: nuc_expt_phr03

Y =X

(1 +X2)(2.2.48)

and

X =(h2 − h1)

2v12(2.2.49)

Alignment (h)

Rotational frequency (keV)0 100 200 300 400

0

2

4

6

8

10 180W|V|=141.5 keV

Figure 2.17: Alignment plots for the yrast and Kπ=8+ band in 180W showing theunperturbed bands [42].

From the experimentally observed (perturbed) energies of two states, an es-

timate of the minimum value of the mixing matrix element can be obtained,

and the unperturbed energies of the basis states obtained. These can then be

used to calculate the unperturbed alignments for the two configurations which

are thought to mix at the band crossing point. Figure 2.17 shows the bandcross-

ing of the yrast states in 180W (decay scheme shown in figure 2.18) which has

been attributed to an interaction between the K=0 ground state band and the

Kπ = 8+, t-band [42]. Note the interaction between the even spin members of

the two bands, while the odd-spin members of the Kπ=8+ remain unperturbed.

This approach can be expanded for more than two bands. For interactions

between three bands, the unperturbed energies can be obtained by diagonlising

the three dimensional determinant [43],

38

Page 42: nuc_expt_phr03

Figure 2.18: Partial decay scheme for 180W showing the ground state and Kπ =8+ band. The alignment in the yrast band is interpreted as a t-band crossing[42].

h1 −E v12 v13

v12 h2 − E v23

v13 v23 h3 − E

= 0

2.3 Selected Topics in High Spin Nuclear Struc-

ture.

In this short section, we will briefly discuss a few ‘hot topics’ in the high spin

study of nuclei. Interested parties are encouraged to look up the references for

the topics of interest.

2.3.1 Shape Coexistence and Superdeformation.

Shape coexistence occurs when two competeing minima exist in the nuclear poten-

tial energy surface corresponding to different shapes of the nuclear mean field[44].

Such effects often occur in nuclei with oe near magic numbers of protons and

39

Page 43: nuc_expt_phr03

neutrons. Here, a spherical configuration competes with a deformed shape corre-

sponding to particle-hole excitations across the shell gaps. This deformed config-

uration populates deformation driving orbitals (those which drop down rapidly

with energy for increasing deformation). Good examples of studies of spherical-

prolate shape coexistence in heavy nuclei for very neutron deficient nuclei around

the Z=82 magic number can be found in references [46, 45, 47, 48].

The study of very elongated or superdeformed nuclei is an extreme form of

shape coexistence [49, 50, 51]. Superdeformed decays are characterised by long

cascades of regularly spaced E2 transitions, with a large I(2) moment of inertia

very large, in-band B(E2) values, corresponding to highly collective decays. Using

the rotational model, these B(E) correspond to nuclear quadrupole deformations

of β ∼ 0.35 → 0.6 or major to minor axis ratios of between 3:2 [53, 54, 55, 56]

and 2:1 [57, 58]. The regions where superdeformed nuclei have been observed,

are close to those nuclei with proton and neutron numbers which correspond to

large shell gaps (regions of low level density) at large prolate deformations in

the deformed shall model potential. For example, the superdeformed nuclei in

the A∼150 region are centred about 15266 Dy86, where both proton and neutron

numbers correspond to superdeformed magic numbers [52]. Similar regions of

superdeformation have been identified around 13258 Ce74 [65, 59, 64], 192

80 Hg112 [50]

and 8238Sr44 [61].

100.0 300.0 500.0 700.0 900.0Energy (keV)

0.0

1000.0

2000.0

3000.0

Cou

nts

192Hg Band 2

194Hg Band 3

200 400 600 800Egamma (keV)

0.6

0.8

1.0

1.2

∆Spi

n (h

)

90

100

110

120

130

J(2) (

Mev

/h2 )

10

20

30

40

50

Sp

in (

h)

Figure 2.19: Identical superdeformed bands in 192Hg and 194Hg [66].

A partilcularly interesting and puzzling aspect of the study of superdeformed

40

Page 44: nuc_expt_phr03

bands is the presence of rotational bands with almost identical energy spacing in

SD bands of neighboring nuclei, the so called identical bands effect [62, 63, 64, 65].

This is surpring in light of the expected drop in pairing corrlations caused by the

presence of an odd, unpaired nucleon which in normal deformed nuclei, causes a

reduction in the I(2) moment of inertia by about 10-15%.

Much work has been invested over the past 10 years in identifying the direct

decays out of SD bands into the lower spin yrast states. This information is very

important as without it, the spin and excitation energy of the SD band memebers

can not be exerimentally determined with any certainty. There have been recent

ports of direct links from SD bands in the A∼190 region [67, 68], however such

links in the A∼80 and 150 regions currently remain elusive.

2.3.2 Band Terminations.

Since the nucleus is a finite system, the collective angular momentum observed in

a nucleus must have a microscopic basis. It is generated by the individual valence

particles. Thus, the maximum spin available will be determined by the maximal

coupling of the single particle spins for those particles outside the inert core. In

order to generate higher angular momentum valuesthe core must be broken and

the collective band structure will be ’terminated’ This can occur rapidly, over

one or two transitions [69, 70, 71] or more gradually over a number of transitions

[72, 73, 74]. The latter effect is termed ’soft’ of ’smooth’ band termination [75]

and is characterised by an increase in gamma-ray transition energy (corrsponding

to a reduction in the moment of inertia and collectivity) with increasing spin.

2.3.3 High K-Isomers and Pairing Reduction.

For deformed nuclei with valence nucleons with large angular momentum pro-

jections, Ω, on the nuclear axis of symmtery, we can assign a total projection of

the angular momentum on the symmtery axis, K, where K =∑

j Ωj . The K-

selection rule for gamma-decays is that for K remaining a good quantum number,

∆L ≥ ∆K, where ∆L is the multipolarity of the transition linking two bands

differering by ∆K. Thus, if ∆K is (≥ 2), then unusually high order multipoles

are required for such decays to occur, and an isomeric state results. For nuclei

in the vicinity of Z∼72 and N∼106, both the proton and neutron Fermi sur-

faces lie in the vicinity of high-Ω orbitals. (K-isomeric states have also been

41

Page 45: nuc_expt_phr03

Figure 2.20: Smooth band termination in 110Sb [76].

observed in N=74 isotones [90, 91, 92, 93, 94]). By breaking pairs of nucle-

ons, very high-Ω states can be obtained relatively cheaply in energy and these

compete energtically with the yrast states and a large number of K-isomeric

states and their associated rotational bands have been observed in this region

[37, 79, 80, 81, 82, 83, 84, 85, 86, 87, 88]. The study of the rotational bands

built in these states is of particular interest since, these configurations are built

on configuration consisting of upto eight unpaired nucleons [80, 9] and the effect

this has on the overall pairing correlation can be invetsigated in a configuration

dependent manner by obtaining the moment of inertia for the various seniority

band structures and noting the increase with number of broken pairs [80].

The reduction in the moment of inertia of a nucleus with pairing correlations,

compared to the rigid body value can be estimated using the Migdal formula [89],

Ip,n = Irig

(

1 − x+√

1 + x2

x√

1 + x2

)

(2.3.50)

where

x = δhω0

2∆p,n(2.3.51)

42

Page 46: nuc_expt_phr03

where ∆ is the pair gap parameter, hω0 is the harmonic oscilator frequency

and I is the moment of inertia.

Figure 2.21: Partial decay scheme for 178W showing 0, 2, 4, 6 and 8 quasi-particlestructures [9].

By investigating the moments of inertia of mutli-quasi-particle states (where

various numbers of valence pairs are blocked), the effect on pairing on the mo-

ment on inertia (and visa versa) can be investigated for various specific orbital

configurations. As an example, figure 2.21 shows the partial level scheme for178W, which shows a large number of different rotational bands built on 0, 2, 4, 6

and eight quasi-particle states. The moments of inertia of the various bands can

be extracted from the gamma-ray transition energies and these are plotted in fig-

ure 2.22. The increase in moment in inertia with increased blocking (ie. reduced

pairing correlations) is clear. However, note that even for the eight quasi-particle

structure, the value if still some way from the rigid body value.

43

Page 47: nuc_expt_phr03

0 5 10 15 20 25 30Bandhead Spin (h)

0.0

20.0

40.0

60.0

80.0

Mom

ent o

f in

ertia

(h

2 MeV

-1)

Exact calc.BCS calc.Expt.Rigid body

gsb

2-qp4-qp

6-qp8-qp

Figure 2.22: Experimental and calculated moments of inertia for various multi-quasi-particle bands in 178W [9].

2.3.4 Octupole Correlations.

The study of reflection assymetric, octupole deformed nuclei at high spins [95]

is centred around regions of the nuclear chart where there are sets of nuclear

orbitals which differ in both total spin J and orbital angular momentum l, by

three units, ie ∆J = ∆L = 3, such as the following pairs of spherical orbitals,

(j 15

2

, g 9

2

), (i 13

2

, f 7

2

), (h 11

2

, d 5

2

) and (g 9

2

, p 3

2

). Regions where octupole correlations are

particularly favoured to occur where both protons and neutron occupy regions of

the Fermi surface where these combinations of such orbitals lie closeby in energy.

For example, in the actinide region around Z∼88 (i 13

2

, f 7

2

) and N∼134 (j 15

2

, g 9

2

)

[95, 96, 97] and in the neutron rich rare earth region around Z ∼56 (h 11

2

, d 5

2

),

N ∼88 (i 13

2

, f 7

2

) [95, 99, 100].

Nuclei with static octupole deformations (β3) are characterised rotational

bands of opposing parity interleaved by enhanced electric dipoles (E1) (see fig-

ure 2.23). Typical E1 strengths in heavy nuclei range between 10−4 and 10−7

Weisskopf units. For nuclei which exhibit octupole deformation, this increases to

44

Page 48: nuc_expt_phr03

0+2+

4+

(6+)

(8+)

(10+)

(12+)

(14+)

(16+)

(18+)

(20+)

1– 242.13– 317.3

(5– )

(7– )

(9– )

(11– )

(13– )

(15– )

(17– )

(19– )

0+84.5 2+4+

6+

8+

10+

12+

(14+)

(16+)

(18+)

(20+)

(22+)

(24+)

(26+)

(28+)

1– 215.93–5–

7–

9–

11–

(13– )

(15– )

(17– )

(19– )

(21– )

(23– )

(25– )

(27– )

0+2+ 67.7

4+

6+

8+

10+

12+

14+

16+

18+

(20+)

(22+)

(24+)

(26+)

(28+)

1– 253.73– 321.55– 447.0

7–

9–

11–

13–

15–

17–

(19– )

(21– )

(23– )

(25– )

(27– )

111.2

190.7

248.4

292.9

330.1

363.9

396.0

425.5

452.3

476.7

229.3

289.0

338.3

379.6

415.0

444.8

470.8

166.5

228.6

276.1

313.7

345.3

373.8

400.2

424.4

447.1

467.1

485.1

500.8

518.2

142.6

207.9

265.9

314.8

356.7

391.1

419.4

442.9

462.8

479.8

496.8

511.6

144.0

205.0

252.9

290.7

321.3

347.3

369.8

391.1

411.3

431.6

452.9

472.6

492.5

180.2

231.0

275.3

314.5

348.5

378.4

404.4

427.4

448.2

467.0

484.0

153.1

149.3

157.4

173.1

192.1

211.4

172.2

140.1

180.9

206.2

206.0

182.2

160.5

151.2

210.5

188.6

173.3

166.5

176.5

189.5

167.6

148.3

180.4

202.1

214.9

221.8

222

88Ra

224

88Ra

226

88Ra

0+2+ 63.8

4+

(6+)

(8+)

(10+)

(12+)

(14+)

(16+)

(18+)

(20+)

(22+)

(1– ) 474.1(3– ) 537.6(5– ) 656.0

(7– )

(9– )

(11– )

(13– )

(15– )

(17– )

(19– )

0+ (2+) 57.4(4+)

(6+)

(8+)

(10+)

(12+)

(14+)

(16+)

(18+)

(1– ) 710.9(3– ) 785.8(5– ) 879.8

140.8

206.9

262.6

309.0

347.8

378.9

403.8

422.4

436.1

446.8

225.0

272.0

312.3

348.3

380.1

408.6

129.4

192.3

247.3

293.6

332.2

364.1

390

412

418.4

380.8

343.6

308.3

277.5

254.6

228

88Ra

230

88Ra

Figure 2.23: Decays schemes for octupole deformed Radium (Z=88) nuclei show-ing enhanced E1 decays between rotational structures of differing parities [97, 97].

between 10−3 and 10−2 Wu. The E1/E2 branching ratio can be used to estimate

the intrinsic electric dipole moment of the nucleus, D0, which is related to the oc-

tupole deformation. The electric dipole moment is related to the B(E1 : Ii → If)

by the expression [95]

B(E1 : Ii → If ) =3D2

0

4π< Ii010|If0 >2=

3D20

(I −K)(I +K)

I(2I + 1)(2.3.52)

2.4 Branching Ratios and g-Factors in High-K

Bands.

The γ-ray branching ratio λ, for a given state in a rotational band, is given by,

λ =T2(E2)

T1(E2) + T1(M1)=

P2(E2)

P1(E2) + P1(M1)(2.4.53)

45

Page 49: nuc_expt_phr03

where T1 and T2 are the γ-ray intensities for spin changes of 1 and 2 respectively.

P1 and P2 are the corresponding transition probabilities (∝ intensity). Note

that the ratio T1(M1)T1(E2)

= 1δ2 , where δ is the quadrupole/dipole mixing ratio. The

transition probabilities1 are related to the reduced transition probabilities (B(XL))

by equations 2.4.54.

B(E2; Ii → If) =1

1.225 × 109E5γ

P (E2; Ii → If)

B(M1; Ii → If) =1

1.758 × 1013E3γ

P (M1; Ii → If) (2.4.54)

where E1 and E2 are the M1 and E2 transition energies, in MeV, respectively.

Substituting for the transition probabilities in equation 2.4.53 using equations 2.4.54

yields,

λ =(

E2

E1

)5 B2(E2) × 1.225 × 109

B1(E2) × 1.225 × 109 +B1(M1)E−21 × 1.758 × 1013

(2.4.55)

The reduced in-band transition probabilities1 are given by,

B(E2; IiK → IfK) =5

16πe2Q2

o| < Ii2K0|IfK > |2

B(E2; IiK → IfK) =5

16πe2Q2

o| < Ii1K0|IfK > |2 (2.4.56)

B(M1; IiK → IfK) =3

4πe2| < Ii1K0|IfK > |2(gK − gR)2K2

where Qo is the intrinsic quadrupole moment and gK and gR are the intrinsic

and rotational gyromagnetic ratios respectively. The relevant Clebsch-Gordon

coefficients2 are given below.

E2(∆I = 2) =

[

3(I −K)(I −K − 1)(I +K)(I +K − 1)

(2I − 2)(2I − 1)I(2I + 1)

]1/2

E2(∆I = 1) = −K[

3(I −K)(I +K)

(I − 1)I(2I + 1)(I + 1)

]1/2

(2.4.57)

M1(∆I = 1) = −[

(I −K)(I +K)

I(2I + 1)

]1/2

1K.E.G. Lobner in, The Electromagnetic Interaction in Nuclear Spectroscopy, W.D. Hamil-ton (Ed), North-Holland (1975) Chapter 5

2The Theory of Atomic Spectra, Condon and Shortley (1935) reprinted (1963) p76-77

46

Page 50: nuc_expt_phr03

where I is the angular momentum of the initial state. K is assumed to be a

good quantum number throughout. Substituting equations 2.4.57 and 2.4.58

into equation 2.4.55 results in λ =,

(

E2

E1

)5

516πe2Q2

o

[

3(I−K)(I−K−1)(I+K)(I+K−1)(2I−2)(2I−1)I(2I+1)

]

× 1.225 × 109

516πe2Q2

o

[

3(I−K)(I+K)(I−1)I(2I+1)(I+1)

]

× 1.225 × 109 + 34π

[

e2 (gK−gR)2K2

E2

1

(I−K)(I+K)I(2I+1)

]

× 1.758 × 1013

(2.4.58)

Rearranging these terms gives equation 2.4.59

λ =(

E2

E1

)5

516

[

3(I−K−1)(I+K−1)(I+1)(2I−1)2K2

]

516

[

3(I−1)(I+1)

]

+ 34

[

(gK−gR)2K2

Q2oE2

1

(

1.758×1013

1.225×109

)]

(2.4.59)

This leads to the following expression for the branching ratio.

λ =(

E2

E1

)5

[

(I−K−1)(I+K−1)(I+1)(2I−1)2K2

]

1 + 45

[

(gK−gR)2

Q2oE2

1

(I2 + 1)(1.435 × 104)]

(2.4.60)

This equation is identical to that quoted by Alexander et al.3 To convert Qo from

units of e·fm2 to e·b, there is an additional factor required, namely, Qo → Qo

100.

Making this substitution gives equation 2.4.61

(gK − gR)

Qo=

5

4(0.8347)

E1√

(I2 − 1)

(

1

λ

(

E2

E1

)5[

(I −K − 1)(I +K − 1)(I + 1)

2K2(2I − 1)

]

− 1

)1/2

(2.4.61)

This result can also be expressed in terms of the quadrupole/dipole mixing ratio

δ,(gK − gR)

Qo= 0.933

E1

δ√

(I2 − 1)(2.4.62)

where δ is related to the quadrupole admixture q, by,

q =δ2

1 + δ2=

2K2(2I − 1)

(I −K − 1)(I +K − 1)(I + 1)

E51

E52

λ (2.4.63)

The factor (gK−gR)Qo

should be a constant for a given rotational band.

3P. Alexander, F. Boehm, and E. Kankeleit, Phys. Rev. 133 (1964) B284

47

Page 51: nuc_expt_phr03

Chapter 3

Experimental Gamma-raySpectroscopy.

3.1 Germanium Semi-Conductor Detectors.

The basic idea behind using semiconductor materials to detect radiation is that

through interactions with the radiation, it is possible to excite electrons from

the valence band into the conduction band (assuming that the energy of the

radiation is larger than the semiconductor band gap). This leaves a hole behind

in the valence band, thus a electron-hole pair has been created. The liberated

charge can be then swept away by an applied voltage.

For temperatures greater than absolute zero, thermal energy is shared by the

electrons in the crystal lattice and thus it is possible to for an electron to be

thermally excited across the band gap into the conduction band. The probabilty

per unit time of thermally creating an electron-hole pair, P (T ), as a function of

absolute temperature, T , is given by a Boltzmann function,

P (T ) = CT3

2exp(

− Eg

2kT

)

(3.1.1)

where Eg is the band gap, k is Boltzmann’s constant and C is a constant

which is material dependent.

Thus for small value of a band gap (as is the case for semiconductors), there is

a large probability of thermal excitation, which in a detector would be a source of

unwanted noise. In order to reduce this thermal noise, semiconductor detectors

should be operated at low temperature (usually 77 K, liquid nitrogen tempera-

ture).

48

Page 52: nuc_expt_phr03

If radiation enters the electrically neutral depletion layer, electron-hole pairs

can be created. The electrons will flow towards the positive potential and the

holes to the negative. This charge can be collected and converted to an output

voltage by a pre-amplifier. The number of electron-hole pairs created, and thus

the size of he output voltage, is proportional to the energy of the radiation.

3.2 Gamma-Ray Spectroscopy with Germanium

Detectors.

Gamma-ray spectroscopy is used to (a) identify the quantum levels in a nucleus

to probe the physics of nuclear structure and (b) identify radioactive substances

by measuring their characteristic decay gamma-rays (eg 662 keV line in 137Cs).

General considerations for a good gamma-ray spectrometer device are that (a)

it must have excellent energy resolution, (b) a good photopeak efficiency and (c)

good timing properties.

While sodium iodide has a better efficiency than germanium, the excellent

energy resolution of germanium (better than 0.2 % at 1.333 MeV) makes it the

gamma detector of choice for high resolution studies. The main problems with

germanium are

• The most probable interaction for most gamma-rays is Compton scattering

and a sizeable portion of gamma-rays that enter the detector will scatter

out of the detector before the full energy has been absorbed in the crystal.

This gives rise to a large Compton background.

• Germanium detectors must be kept at liquid nitrogen temperature for good

resolution. This means that bulky liquid nitrogen dewars must be included

in the detector apparatus.

3.2.1 Response Function of Germanium Spectra.

(Knoll p289-293, p301.)

The typical germanium spectrum is made up from a number of different com-

ponents. These include,

1. The Full Energy Peak. The peak corresponding to where all the incident

radiation’s energy has been collected by photoelectric absorption (some

49

Page 53: nuc_expt_phr03

fraction will have been Compton scattered before p-e absorption). In a

perfect, idealised detector, all the counts would be in this peak.

2. The Compton Background. The background of counts with energies

less than the full energy peak where some of the incident radiation has

been Compton scattered out of the detector.

3. The Compton Edge. In a Compton scattering event, the energy removed

by the electron is given by ∆E, where ∆E = hν−hν ′ = hν( hν

mec2(1−cosθ)

1+ hν

mec2(1−cosθ)

)

.

Therefore, the minimum energy removed by scattering an electron (when

θ=180) is given by ∆E= Eγ

1+2Eγ

mec2

, where mec2 = rest mass of the elec-

tron= 511 keV. This minimum amount of energy being lost in a Compton

scattering, gives rise to the Compton background being essentially cut off

at an energy ∆E below the full energy peak. Note for Eγ >> 511 keV,

∆E → mec2

2≈ 250 keV.

4. Escape Peak(s). For photon energies greater than twice the electron

rest mass energy (511 keV ×2 = 1.022 MeV) there is a probabilty of pair

production where an electron-positron pair is created. The positron may

then recombine with an atomic electron in the detector and decay back to

2×511 keV gamma-rays. One or both of these 511 keV gammas may then

escape from the detector with no further interaction. If the initial gamma-

ray enegy has an energy Eγ , then the escape peaks lie at energies Eγ–511

keV and Eγ–1.022 MeV

5. Backscatter Peak. These correspond to gamma-rays which have been

scattered backwards in material surrounding the detector. The energy for

back scattered gamma-rays is approximately equal for all incident energies

at between 200 and 250 keV. The energy of the backscatter peak corre-

sponds to the energy of the photon after it is scattered. This is simply the

energy difference between the full energy peak and the Compton edge.

6. Annihilation Peak. One observes a peak at 511 keV due to annihila-

tion radiation from pair-production caused by the initial radiation in the

surrounding material (assuming Eγ > 1.022 MeV). The 511 keV is then

measured in the detector.

50

Page 54: nuc_expt_phr03

7. X-ray Escape Peaks. A characteristic X-ray is emitted by the material in

the photoelectric process. This is usually reabsorbed, but occasionally it can

escape the detector. Thus a peak with an energy equal to the photo-peak

energy minus the X-ray energy can appear. This is only really a problem for

(a) low gamma-ray energies and (b) detectors with large surface to volume

ratios.

3.2.2 Germanium Detector Efficiency.

(See Knoll p427.)

The efficiency of a germaium detector is usually given relative to that of a

3in × 3in NaI(Tl) detector for the 1333 keV gamma-ray in 60Ni (from a 60Co β-

source). Typical efficiencies range from 30 % to 70 % for hyperpure germanium

detectors.

The efficiency response varies as a function of energy and is usually empirically

deduced using a variety of standard calibration sources such as 152Eu and 133Ba

(see figure 3.1.)

Figure 3.1: Gamma-ray spectra for (a) 152Eu and (b) 133Ba efficiency calibrationsources.

51

Page 55: nuc_expt_phr03

3.2.3 The Compton Suppressed Spectrometer (CSS).

(Knoll p421.)

In order to reduce the (unwanted) Compton continuum events in a germanium

gamma-ray spectrum, and thus increase the signal to noise ratio, the germanium

detector can be surrounded by a high efficiency gamma-ray scintillator (usually

Bi4Ge3O12 or BGO). This shield acts as veto for Compton events which scatter

out of the germanium detector. The peak to total for the 1173+1333 keV lines

from the 60Co source for an unsuppressed detector is only about 20 %. This rises

to 50–60 % for a suppressed detector with almost no loss in the number of counts

in the full energy peak.

Liquid

nitrogen

‘cold finger’

incoming

gamma-ray

photomultiplier

tubes

GeBG

O

BG

O

NaI NaI

Figure 3.2: A Compton suppressed germanium detector

The improvement in the peak to total achieved by using anti-Compton shields

is even more crucial when using more than one germanium detector at a time to

measure more mulitiple gamma-rays emitted in a cascade. This is known as co-

incidence spectroscopy. For an event where two gamma-rays from a cascade are

measured (a γ − γ or γ2 event), for two unsupressed detectors, the probability

that both gamma-rays (for Eγ ∼1 MeV) will be in the full energy peak is

approximately 0.2×0.2=0.04. If however, both detectors are surrounded by a

suppression shield this rises to 0.6×0.6=0.36, almost an order of magntitude

improvement. The increase is even larger for higher multiplicity events.

3.3 Gamma-Ray Arrays.

Much nuclear spectroscopy is carried out by measuring the gamma-rays emitted

following a fusion-evaporation reaction. This tends to form the nucleus of in-

52

Page 56: nuc_expt_phr03

Energy (keV)

Nu

mb

er o

f C

ou

nts

(a) Eurogam I

Figure 3.3: Effect of higher fold gating on incresing the signal to noise [174].

terest in a high spin state (typically around 30-50 h) which decays by emitting

a cascade of gamma-rays. The ideal tool for gamma-ray spectroscopy would be

able to measure all of the individual gamma-rays in a cascade with 100% effi-

ciency. In practice this is not possible, however, over the last 10 years there have

been significant advances in the development of gamma-ray spectrometers con-

sisting of large numbers of germanium detectors (or ‘arrays’) such as TESSA3

[101], the 8π spectrometer [103], EUROGAM [174, 173] and GAMMASPHERE

[104]. These devices have revolutionised the field of gamma-ray spectroscopy by

allowing hitherto unthinkable sensitivity to weak transitions. The following sub-

sections will describe the important factors in array design and how the maximum

performance can be achieved.

53

Page 57: nuc_expt_phr03

3.3.1 Resolving Power and Total Photopeak Efficiency.

The resolving power [174] of a gamma-ray spectrometer is given by R where

R =

(

SEγ

∆Efinalγ

)

PT (3.3.2)

SEγ is the ‘average’ energy spacing consecutive γ-ray transitions in a ‘typical’

cascade. This is clearly reaction dependent and can not be ‘improved’ upon.

Therefore, in order to improve the resolving power on the array improvements

must be made on either the energy resolution, ∆Efinalγ , and/or the peak to total

fraction, PT .

∆Efinalγ is the full width at half maximum resolution of γ-rays obtained in

the detector spectrum. This is made up from a number of effects including the

intrinsic, statistical resolution of the detector, the effect of Doppler broadening,

the spread in the recoil cone and recoil velocity.

The total resolution of a given γ-ray transition in a fusion-evaporation reaction

is given by [174]

∆Efinalγ =

(

∆E2int + ∆E2

open + ∆E2ang + ∆E2

vel

) 1

2 (3.3.3)

Doppler Shifts and Doppler Broadening.

If the nucleus of interest recoils out of the target with a velocity v, its energy

is Doppler shifted when measured in a detector at angle θ to the recoil velocity

direction to a value Es, where, if E0 is the ‘unshifted’ energy,

Es(θ) = Eo

1 − (vc)2

1 − vccosθ

≈ E0

(

1 +v

ccosθ

)

(3.3.4)

Since the geramnium detector measuring the gamma-ray has a finite opening

angle, there will be a spread in the Doppler shifted energy across the face of the

detector (see figure 3.4).

If the opening angle of the detector is ∆θ, then differentiating equation 3.3.4

gives,

dEs

dθ≈ ∆Es

∆θ≈ Eo

v

csinθ (3.3.5)

54

Page 58: nuc_expt_phr03

germanium detector

E = Eo (1 + v/c cos )θs

θ

beam direction, velocity, v

∆θ∆θ

γ

∆Ε = Εοv/c sin θs ∆θ

Figure 3.4: Schematic of the effect of finite opening angle for a detector causingDoppler broadening of the gamma-ray lineshape.

The full expression, including broadening due to velocity spread (see fig. 3.5),

can be approximated by

∆Es ≈ Eocosθ∆v

c− Eo

v

csinθ∆θ (3.3.6)

Clovers and Segmented Clover Detectors.

One method which can be used to improve the energy resolution deteriation from

Doppler broadening is to segment the germanium detector, thereby reducing the

effective solid angle of its face. Clover detectors use four 25% germanium crystals

housed in the same can. The detectors are smaller than the typical 70% germaium

crystals and thus have a smaller intrinsic efficiency, however, this limitation can be

overcome by adding together gamma-rays measured in two neighboring segments

(from events which Compton scatter from one crystal to the next) [176, 177].

55

Page 59: nuc_expt_phr03

Figure 3.6 shows the effect of clover detectors at 90 in the EUROBALL

array compared to the same reaction using 70% detectors. The improvement in

resolution, and thus sensitivity is clear. In some more modern cases, the clover

elements are segmented electrically to give an even better defined localisation

[178]. This is particularly useful for fast moving recoils and has been implemented

for example in the EXOGAM array for use with radioactive beams [105].

3.3.2 Add-Backs from Clover/Cluster Detectors.

Improvements in the peak to total can be achieved by using larger volume germa-

nium detector, by improving and maximising the Compton suppression and/or

the use of add-back such as in the cluster detector [179] (see figure 3.7).

A further advantage of using a clover type detector [176, 178] comes when try-

ing to look at higher energy gamma-rays. As figure addback shows, the majority

of counts with energies less than 1 MeV are detected in a single element. How-

ever, for energies above this value, the probability of Compton scattering from

one crystal into a neighboring one (spectrum (c)) is large enough that there is a

significant increase in the full energy peaks obtained using the detector in add-

back mode. This increases the sensitivity of the array by improving the overall

signal to noise (or peak to total) of the device.

3.3.3 Polarization Measurements.

Clover detectors can also be used to remove the anomaly between electric and

magnetic dipole transition by measuring the polarization of the gamma-rays

[186, 185, 187]. This type of analysis uses the fact that the direction of Compton

scatters is different for electric and magnetic type transitions. Thus, by com-

paring the efficiency corrected intensities of ‘added-back’ lines from horizontal

and vertical scatters, the electric or magnetic (or mixed) nature of the transi-

tion can be eludidated. This is highlighted for experimental data taken from the

EUROGAM 2 array in figures 3.8 and 3.9.

3.3.4 Gamma-ray Tracking.

Work is currently underway on investigating the possibilities of using algorthyms

to track Compton scatters through a highly segmented germanium shell to make

56

Page 60: nuc_expt_phr03

the next generation of gamma-ray spectrometers (eg. the GRETA project [180,

181, 182]). The idea uses the Compton formula to determine whether two events

in a germanium shell are two separate gamma-rays or, two ‘tracks’ of a single

events which has Compton scattered. Using an algorythm (see eg. [183]) a

decision can be made offline, as to whether to ‘add-back’ or ‘reject’ Compton

scatters. Such algorythms depend on a good knowledge of the interaction points

of the gamma-rays in the crystals, which can be further improved by looking at

the charge collection times of the signals and their pulse shapes [184].

57

Page 61: nuc_expt_phr03

Figure 3.5: Gamma-ray spectra taken from the reaction 28Si on a 40Ca targetwith a v

cof 1.8%. Note the doppler broadening is larger away from 90 in this

case due to the large spread in recoil energy associated with the alpha-particleevaporation [106] .

58

Page 62: nuc_expt_phr03

1400 1450 1500

400

800

1200

1600

1400 1450 1500

200

400

600

800

γ-ray energy (keV)

Num

ber

of c

ount

s

Phase I

Phase IIFWHM = 4.8 keV

FWHM = 6.1 keV

Figure 3.6: Effect of using clover detectors in improving energy resolution inEUROGAM phase II [174].

59

Page 63: nuc_expt_phr03

(f) (e) (h)

(a) (b) (c)

(g)

(d)

Figure 3.7: Spectra gated on add-backs from cluster detectors at PEX using thereaction 28Si+40Ca [107].

60

Page 64: nuc_expt_phr03

0 5 10 15 20 25 30

–0.2

0.0

0.2

0.4

0.6

0.8

Band 3 (M1)

Band 12 (E2)

Spin (–h)

Fig. 5

γ–ra

y po

lari

satio

n P

/home/esp/text/figs/i117/pol3.psk Paul et al PRC

Figure 3.8: Polarization measurements showing the difference between magneticand electric type decays in 132Ce [186, 185].

61

Page 65: nuc_expt_phr03

0.0 0.5 1.0 1.5

–10

0

10

0.0

0.5

1.0

1.5

Eγ (MeV) Fig. 4

Cou

nts

(104 )

/home/esp/text/figs/i117/polspec.psk Paul et al NPA

(a) 117

I gated

(b) Ungated

↓ Magnetic

↑ Electric

↓ Magnetic

↑ Electric

Figure 3.9: Difference spectra for horizontal and vertical scatters for transitions in117I showing how a polarization measurement can be use to differentiate betweenelectric and magnetic transitions (see [186, 185].

62

Page 66: nuc_expt_phr03

Chapter 4

Channel Selection InFusion-Evaporation Reactions.

The statistical nature of nucleon emission from the hot, compound system means

that fusion-evaporation reactions tend to populate more than one residual nu-

cleus. In order to associate gamma-rays with specific nuclei, some form of extra

channel selection is required. There are two main ways to determine the nucleus

which emitted the gamma-rays, either by identifying the nucleus itself (recoil sep-

arator, gas filter techniques) or by measuring the particles evaporated from the

compound system (charged particle balls, neutron detectors). The former gen-

erally have the advantage of being ‘cleaner’ and less susceptible to contaminant

nuclei (from for example isotopic impurities in the target) while the latter tend

to have a larger detection efficiency.

In this section we will discuss four main types of channel selection; (a) inner

sum-energy balls [101]; (b) light charged particle detection arrays [109, 110, 111,

112, 113, 114, 124]; (c) neutron detectors [136, 25] and (d) recoil detectors and

separators [143, 144, 146, 147].

4.1 Inner Multiplicity Sum-Energy Balls.

The entry point for fusion evaporation products in the excitation energy versus

total angular momentum plane is generally different for different reaction prod-

ucts and (typically) depends on the number of evaporated particles. Usually,

residual nuclei formed by the evaporation of fewer particles leave the final nu-

cleus in a state of higher spin and excitation energy than channels with higher

63

Page 67: nuc_expt_phr03

particle emission multiplicities. If a high efficiency, high granularity γ-ray detec-

tor (calorimeter) can be placed around the target position, this effect can be used

to give a degree of channel selection between different residual species [101] (see

figures 4.1 and 4.2). This form of channel selection is particularly effective when

the majority of the fusion cross-section goes into evaporation channels consisting

of only neutrons (ie. 3n,4n,5n,...etc).

Figure 4.1: Fold and Sum Energy spectra from the inner ball of the Chalk River8π spectrometer for the reaction 76Ge(34Sn,xn)110−xCd. Spectra (a) and (b) are

gated by the 540 keV 152

− → 112

−and 633 keV 2+ → 0+ transitions in 105Cd and

106Cd respectively. Note the lower values of fold and sum energy observed for thehigher neutron multiplicity evaporation channel [3].

For example the 8π gamma-ray spectrometer at Chalk River contains a 71

element inner ball made up of BGO elements [112]. This can be used to measure

the total energy emitted in a decay cascade (related to the inital excitation en-

ergy of the nucleus), with the number of detectors firing, giving an estimate of

the total gamma-ray multiplicity of the cascade (which is related to the initial

angular momentum of the system). As figures 4.1 and 4.2 clearly demonstrate,

the entry point of a nucleus in the spin/excitation place depends on the num-

ber of evaporated neutrons for the 76Ge+36S reaction. By setting software gates

64

Page 68: nuc_expt_phr03

Fold (~ multiplicity ~ total input angular momentum)

Sum

Ene

rgy

(ent

ranc

e ex

cita

tion

ener

gy)

5n + 105Cd4n + 106Cd

Figure 4.2: Two dimensional Fold and Sum Energy spectra from the inner ballof the Chalk River 8π spectrometer for the reaction 76Ge(34Sn,xn)110−xCd gated

by the 540 keV 152

− → 112

−and 633 keV 2+ → 0+ transitions in 105Cd and 106Cd

respectively.

in offline sorting on various fold/sum-energy conditions, the relative intensity of

transitions from specific evaporation residues can be enhanced.

4.2 Studies of Very Neutron Deficient Nuclei.

For very neutron deficient compound nuclear systems, the reduced proton sepa-

ration energies compared to compound nuclei closer to stability mean that it is

easier for protons and α-particles to tunnel through the Coulomb barrier. This

allows charged particle evaporation to compete with and finally dominate over

neutron emission, and the population of between 10 and 20 different nuclei in a

single reaction is possible. The effectiveness of the particular method of channel

selection used depends on the final nucleus of interest. For example, in studies

of A∼80 nuclei along the N=Z line, these nuclei are accessible via a 2-neutron

evaporation channel [138, 139, 140, 141, 142] for which a charged particle detec-

tor array would not be suitable, whereas a recoil-separator and split-anode gas

ionisation-chamber, or neutron-detection system might yield useful information.

65

Page 69: nuc_expt_phr03

4.2.1 Charged Particle Balls.

The main job of charged particle detector arrays is to detect and disciminate

between the light charged particles evaporated from a compound nucleus. The

types of detectors are usually either some form of fast-slow scintilator sandwich

known as a phoswich detector [110, 111, 109], a thin silicon ∆E detector [134,

113, 114] or some form of scintilator such as CsI(Tl) which uses pulse shape

discrimination [112, 124].

66

Page 70: nuc_expt_phr03

Figure 4.3: (a) Total projection of the 28Si+40Ca reaction from PEX. (b) Samespectrum gated on the 2p condition. (c) 2p gated spectrum with 3p componentsubtracted. (d) 2p gated spectrum with both 3p and 4p components subtracted[106].

67

Page 71: nuc_expt_phr03

Figure 4.4: Particle gated identification spectra gated numbers of emitted protonsfrom the ANU particle detector ball with higher multiplicities subtracted [38].

68

Page 72: nuc_expt_phr03

The ideal charged particle detector has almost 4π coverage around the target

position (with gaps for the beam to come and possibly go out), good timing

properties, good granularity (to reduce the probability of two charged particles

hitting the same detector) and good discrimination between different types of

particles. In practice however, there are gaps and dead regions between the

detectors, and this coupled with kinematic focussing effects means that such

arrays are never 100% efficient. The result is for example a γ ray from a 3

proton evaporation channel will be present in the 3p, 2p 1p and 0p gated spectra.

Careful, normalised subtractions of higher multiplicity charged particle events

must be applied to achieve clean spectra (see figure 4.3). However, as shown in

figure 4.4, once these contaminatnts from higher multiplicity channels have been

subtracted, very clean identification spectra can be obtained

4.2.2 Kinematic Focussing and Conversion from Lab to

COM Energies.

Fusion evaporation reactions evaporate light particles isotropically in the rest

frame of the compound nucleus. However, as shown schematically in figure 4.5,

the fact that the recoil is moving forward in the laboratory frame means that

the angular distribution of emitted of charged particles in a fusion evaporation is

forward focussed in the laboratory frame see figure 4.5.

θPlab

Vcm

V P lab

V R lab

Vpcm

VRcm

Figure 4.5: Schematic of the kinematics of particle evaporation following a fusion-evaporation reaction. Note how the recoil direction is altered by the particleevaporation.

If the beam energy is EB and EB = 12MBv

2B, where MB is the mass of the

beam nuclei and vB is the beam velocity in the lab. frame, then by conservation

69

Page 73: nuc_expt_phr03

of momentum, the velocity of the centre of mass vcm is given by

70

Page 74: nuc_expt_phr03

MBvB = (MB +MT ) vcm (4.2.1)

and by substitution

vcm =MB

(MB +MT )

2EB

MB

(4.2.2)

where MT is the mass of the target nuclei. The velocity of an evaporated

proton in the rest frame of the compound nucleus can be calculated (by the

cosine rule) such that

v2pcm

= v2cm + v2

plab− 2vcmvplab

cosθlab (4.2.3)

where vplabis the emitted proton velocity in the laboratory frame and θlab is

the lab angle between the beam direction and the direction of the emitted proton.

By substitution

v2pcm

=MB

(MB +MT )22EB + v2plab

− 2MB

(MB +MT )

2EB

MBcosθlab (4.2.4)

The proton energy is the centre of mass (intrinsic frame of the compound

system) is given by

Epcm=

1

2Mpv

2pcm

(4.2.5)

Multiplying both sides of equation 4.2.4 by 12Mp gives

1

2mpv

2pcm

= Epcm=

1

2Mp

MB

(MB +MT )22EB+1

2Mpv

2plab

−1

2

Mp2MB

(MB +MT )

2EB

MBvplab

cosθlab

(4.2.6)

Given that vplab=√

Eplab

Mp, substituting in and collecting the terms of equa-

tion 4.2.6, we obtain the following expression for the energy of the evaporated

proton in the centre of mass frame in terms of the measured proton lab energy

(Eplab), the angle of emission in the lab frame (θlab), the beam energy (EB) and

the masses of the beam (MB) and target (MT ) nuclei respectively,

Epcm=

MpMB

(MB +MT )2EB + Eplab− 2

(MB +MT )

MPMBEBEplabcosθlab (4.2.7)

71

Page 75: nuc_expt_phr03

A similar expression can be obtained for α-particle emission.

Note one problem with using charged particle balls for channel selection are

the presence of target impurities. In particular, relatively light ions such as car-

bon and oxygen are common target contaminants. However, the lighter mass of

these nuclei, means that in general, the effect of kinematic focussing for particles

evaporated from beam induced reactions on these nuclei is greater than for the

(typically) heavier target nuclei. Thus, the angular distributions of the emitted

charged particles for such contaminants are usually more forward focussed (in the

laboratory frame) compared to the ‘true’ reaction products [132] and thus they

can be distinguished.

Scintilators.

The basic premise of a scintilator detector is that the incoming radiation interacts

with the electrons in the scintilator, exciting them into higher energy states. As

these states decay, they emit X-rays or light, which is converted into a charge by

means of a light-pipe, photo-cathode and photomultiplier tube assembly.

prom

pt

fluo

resc

ence delayed

fluorescence

(phosphorescence)

abso

rptio

n

S0

S3

S2

S1

SINGLET STATES

T2

T3

TRIPLET STATES

T1

Figure 4.6: Schematic of the decay of prompt and delayed light from an organicscintilator.

Often, depending on which quantum levels the electron decays through, this

72

Page 76: nuc_expt_phr03

light is emitted with two characteristic decay lifetimes, a fast one known as

prompt fluorescence and a slower type known as delayed fluorescence (see fig-

ure 4.6). The relative intensities of these two types of light depends on the type

of radiation which excited the scintilator in the first place and as we shall see

later can be used to distinguish between different types of radiation.

Phoswich Detectors.

Phoswich detectors are combinations of two detectors, in optical contact. The

front detector consists of a thin (∼ 100 µm) ∆E scintilator detector, which has

a fast (∼ 5ns) decay component to its light output, which is particle dependent.

Typically, the heavier the detector particle, the larger the fast signal. The second

detector (thickness ∼5 mm), placed behind the thin ∆E detector, is used to stop

the particle and usually gives out light with a much longer decay time (∼ 1µs)

which is particle independent. The total signal from the two detectors is taken via

a light pipe, into a photomultiplier tube. The anode signal from the PMT is then

integrated in two different time regions by setting two separate time discriminator

gates on the short (∆E) and long (E) components of the signal (see figure 4.7).

By making a 2-dimensional array of ∆E versus E, different evaporation particles

can be identified. In order to stop signals from scattered beam entering the

phoswich detector, thin metal absorber foils (Al or Au) are placed on the front

face of the detector. These foils are thick enough to stop the (higher Z) beam

particles, but thin enough to allow through evaporated protons and α-particles.

Examples of phoswich detectors are the ANU particle detector ball [38] and

the ERICIUS array [111] which use a combination of fast and slow plastic scin-

tilators and the Penn 4π array [110, 132, 287] which uses fast plastic as the ∆E

detector with CaF2(Eu) as the stopping detector.

One problem with phoswich detectors is that due to kinematic focussing,

charged particles emitted at backward angles in the lab frame tend to have much

reduced energies compared to forward angles. If the lab energy is too small, and

the ∆E detector is too thick, the particles will stop in the ∆E fast scintilator

detector. The thickness of the ∆E detector is then a compromise between particle

identification resolution and the low energy threshold. Also, the light pipes and

photomultiplier tubes associated with converting the light output from a phoswich

detector to an electrical pulse constrict the geometery and available space which

73

Page 77: nuc_expt_phr03

for example germanium gamma-ray detectors may take up.

74

Page 78: nuc_expt_phr03

NE102A3 mmCaF2 (Eu)

0.1 mm

"DE""E"

light

out

put /

vol

tage

time

slowfast

α

p

0-40 ns 60 - 250 ns

slow (E)

fast

(D

E)

p

α

Figure 4.7: Schematic of the operation of a phoswich detector for discriminationbetween protons and alpha particles.

4.2.3 Silicon Detectors.

Si Energy Loss Detectors.

Thin (∼150–200 µm) silicon detectors have also been used to give channel se-

lection between protons and α-particles emitted in fusion evaporation reactions

[113, 114, 134, 345]. Particle identification is achieved due to the differential en-

ergy loss of protons and α-particles. An advantage of such detectors is that they

can be made to be very compact and do not require the photo-multiplier tubes

that phoswich detector systems do. They can easily be made into 4π geometry

and the thin nature of the detector means that attenuation of reaction gamma-

rays in the ball material is kept to a minimum. This is an important factor in

keeping the low-energy gamma-ray detection efficiency of such experiments at a

75

Page 79: nuc_expt_phr03

premium.

76

Page 80: nuc_expt_phr03

Figure 4.8 shows a schematic of the NBI silicon inner ball used in the PEX

array [114]. Note the extra granularity of the forward angle detectors. This is to

take into account the effect of kinematic focussing of the reaction products, and

to attempt to keep the count rate for each individual element at a reasonable

level. (If the count rate in an individual element is too high, the dead time for

the whole system is increased and the beam current and thus total reaction rate

in the experiment has to be reduced).

Due to the lower Coloumb barrier, evaporated protons tends to have lower

energies than evaporated α-particles. Thus to first order, the energy deposited

in the detector can be used to distinguish protons and α-particles. The problem

is the high energy tail of the proton distribution leaks into the α-particle energy

regime. This is compensated for in the Si ball type detector by making the

detector thin enough that higher energy protons are not stopped in the detector

and thus leave only a fraction of their energy. As figure 4.9 shows, a good degree

of discrimination is available using such a device.

The main disadvantages of such detectors are that they are rather sensitive

to radiation damage and also, in general, they do not retain good information on

particle energies.

PSD with Si Detectors.

It has long been established that pulse shape discrimination between electron

and hole associated rise times can be used in silicon detectors to provide some

form of disrcimination between different types of radiation [115, 116]. There have

been a number of recent developments which use pulse shape discrimination for

with planar silicon detectors to discriminate between different heavy ions, from

protons, alphas and heavier ions, including the RoSiB detector [118, 119, 117],

which uses a zero cross-over taken from a bipolar amplifier output, verses the

total energy deposited in the detector, to differentiate between different types of

ions.

(E,∆E) Discrimination with Silicon Balls.

The Bethe equation as used in phoswhcih style detectors can also be employed for

silicon detector telesccope to provide good proton/alpha discrimination in fusion-

evaporation reactions. The EUCLIDES and ISIS arrays [120] are constructed in

77

Page 81: nuc_expt_phr03

TARGET

(a)

(c)

1611

20(b)

OUT

IN

(d)

5

31

OUT IN1

23

4

7,8,9

6,10

11,15

16,20

12,13,14

17,18,1923,24

21,22,25

26,27,30

28,29

IN

31

28

22 27

26

21

23

29

24

30

25

1 5

42

3

13 138

12

1712

7

11 6 10

15

15

1914

914

18

Figure 4.8: Schematic of the PEX Si inner ball geometry for selection of lightcharged particles [107, 114].

a ‘football’ geometry of interlocking pentagonal and hexagonal silicon detector

telescopes made up of an energy loss, ∆E, and residual energy E detectors.

Note the separation of the protons and alpha particles in figure 4.10 and also

the locii corresponding to events where more than one evaporated particle has hit

a single telescope (such as αp. Note that due to the kinematics of typical fusion-

evaporation reactions, α particles are often not observed at backward angles in

the laboratory frame, ie. they are more forward focussed than protons.

4.2.4 CsI(Tl) Balls Using Pulse Shape Discrimination.

Some of the best inner balls for channel selection are made of elements consisting

of single crystals of CsI(Tl) scintillator [112, 124]. The light output from CsI(Tl)

has two parts, a fast component (τ = 0.4 →1.0 µs), the amplitude of which

78

Page 82: nuc_expt_phr03

Figure 4.9: Energy spectra for charged particles from the PEX inner Si ball forthe reaction 28Si+40Ca showing the effect of kinematic focussing [106].

is radiation type dependent and a slower (τ = 7 µs), radiation independent

component. Particle identification can be achieved using a similar pulse shape

discrimination technique as outlined above for the phoswich type detectors. High

efficiency, high granularity devices based on this method, such as the ALF-ball

[112] and the MICROBALL [124] have been used very successfully inside the 8π

and GAMMASPHERE gamma-ray spectrometer arrays, to give high degrees of

channel selection. In both cases, to save space and reduce the material between

the target position and surrounding germaniun detectors, the light is converted

into charge using silicon PIN photodiodes.

Figure 4.13 demonstrates the proton/alpha-particle particle discrimination

afforded for an element of the ALF ball. The two axes in figure 4.13 are total

light output (which is related to the total energy deposited in the detector) verses

a zero-cross-over time. This corresponds to the time difference between a T=0

79

Page 83: nuc_expt_phr03

reference time (such as beam burst of prompt-γ-ray signal) and response time of

the CsI(Tl) detector (which is radiation type dependent). This response time can

be obtained by taking an amplitude invariant zero-crossing point (obtained by

taking the preamplified output pulse from the detector through a bipolar shaping

amplifier [112]).

80

Page 84: nuc_expt_phr03

Figure 4.10: Particle hit pattern rom the ISIS silicon E,∆E array for differentangles in the lab frame for the reaction 40Ca on 40Ca. Note the general forwardfocussing of the particles and the absence of alpha particles at backward angles[121]. The lower detector numbers correspond to detectors aty forward angles.

Channel Selection Using Proton/Alpha COM Energies.

While a charged particle detection system can be used to separate evaporation

channels consisting of purely charged particles, another method must be used

to distinguish between channels containing evaporated neutrons as well as the

charged particles. For example a spectrum gated on the condition that 2 pro-

tons are detected will contain counts from 2p,2pn,3p,3pn,4p,α2p,...etc channels.

The transitions from higher charged particle multiplicities will come into such a

spectrum (due to the fact that such devices are never 100% efficient) but can be

identified by their presence in higher multiplicity gated spectra (eg. 3p gated will

contain 3p,3pn,4p,...etc. but should not contain the 2p and 2pn channels).

81

Page 85: nuc_expt_phr03

φ = 108.0°

10 20 30

E (MeV)

2

4

6

8

10

12

14 ∆

E (

MeV

)

φ = 148.3°

10 20 30

E (MeV)

2

4

6

8

10

12

14

2p

p p

Figure 4.11: Particle identification spectra from the ISIS silicon E,∆E array fordifferent angles in the lab frame for the reaction 40Ca on 40Ca for backward ISISangles in the lab frame. Note the general forward focussing of the particles andthe absence of alpha particles at backward angles [120, 121, 122].

φ = 31.7°

10 20 30

E (MeV)

2

4

6

8

10

12

14

∆E

(M

eV)

φ = 69.9°

10 20 30

E (MeV)

2

4

6

8

10

12

14

α p

α

2p 2p

α

p p

Figure 4.12: Particle identification spectra from the ISIS silicon E,∆E array forforward angles in the lab frame for the reaction 40Ca on 40Ca [120, 121].

82

Page 86: nuc_expt_phr03

Tim

e re

spon

se o

f C

sI(T

l)

Light Output

p

α

Figure 4.13: Time verses total energy spectrum for one element of the chalk RiverALF ball for the reaction 76Ge+34S. Note the protons and alpha particle locii areclearly separated.

One method of distinguishing between for example 2p and 2pn events is to

compare the average evaporated proton energy (in the centre of mass frame).

The 2p proton channel will tend to have a higher average proton energy than

the 2pn channel (since some of the available excitation energy is removed by the

neutron).

Another way of thinking of this is to realise that the total excitation energy

of the compound nucleus is fixed (by the kinetic energy of the reaction and the

Q-value for the reaction). By measuring the proton energies in the centre of mass

frames and subtracting them from the total compound excitation energy, one is

left with the ‘residual’ excitation energy of the system. This can be removed by

either the emission of an (unobserved) neutron or by gamma-ray emission (if the

excitation energy of the residual nucleus is below the particle evaporation thresh-

old). The probability of emitting a further particle depends on the excitation

energy the system is left in after a particle emission. For example, if the nucleus

emits two protons with relatively high energies, the residual system will be left

in a lower energy state than for cases where the two protons remove less energy.

Since there is more residual energy left in the latter case, it is more likely that

83

Page 87: nuc_expt_phr03

such a scenario will result in the emission of a further nucleon (neutron) to bring

the final nucleus below the particle evaporation threshold. This is shown to be

true experimentally in the work by Balamuth et al. and Pohl et al. [132, 135].

Similarly, the average energy of protons emitted in the 2p channel is larger

than the average energy of protons emitted in the 3p channel and so on [110, 123,

135, 112]. Thus, even using a single charged particle detector, and setting off-

line gates on measured average proton energy will yield some selection between

different proton evaporation channels.

Selection of High Spin Cascades by Particle Evaporation Spectra

Pohl et al. [135] showed that there is a correlation between the proton emission

spectra (and thus the total entrance excitation energy of the residual system)

and the population of high spin states. This is a simple effect to understand, the

less excitation energy taken away by the emitted particles, the more is available

for gamma decay in the residual nucleus. The experimental data taken from the58Fe(27Al,3p)82Sr reaction [135] highlighting this effect is shown in figure 4.15.

The Total Energy Plane Method.

As pointed out previously, a problem with using charged particle balls to iden-

tify weak, low-multiplicity channels in very neutron deficient systems, such as

the 2p channel, is that the raw 2 proton gated spectrum will be ‘contaminated’

with lines from higher charged particle multiplicity evaporation channels (eg,

3p,4p,α2p,α2pn...) If the low-multiplicity channel of interest is populated weakly

in the reaction, the breakthrough lines from the higher multiplicity channels will

dominate the spectra. A clever method has been suggested by Svensson et al.

[133] to overcome this problem.

In a fusion-evaporation reaction, the total energy in the centre of mass frame

for a given exit channel, ECM , is given by

ECM = TCM +Q (4.2.8)

where TCM is the centre of mass kinetic energy of the beam-target collision

(ie. beam energy in the centre of mass frame) and Q is the reaction Q-value for

the different (specific) exit channels.

84

Page 88: nuc_expt_phr03

This total excitation energy bought into the compound system is then emitted

in the form of γ-ray decays and emitted nucleon kinetic energies such that

85

Page 89: nuc_expt_phr03

Figure 4.14: Centre of mass proton energy spectra for 27Al+58Fe reaction at abeam energy of 92 MeV, gated on transitions in 82Sr and 81Sr from the 3p and3pn channels respectively [135]. The lower average proton energy for the 3pnchannel is clearly demonstrated.

86

Page 90: nuc_expt_phr03

Figure 4.15: Entry-point excitation energy centroids in coincidence with discretegamma-rays depopulating states at excitation energies between 0 and 8 MeV in82Sr. Note that the higher excitation energy states, typically come from nucleiformed with higher energy entry points (ie. lower proton evaporation energies)[135].

87

Page 91: nuc_expt_phr03

ECM = Hγ + Tpart (4.2.9)

where Hγ is the total energy emitted as gamma-rays and Tpart is the total

emitted particle kinetic energy in the centre of mass frame. Note that ECM is

constant for a specific channel. Therefore if Hγ and Tpart can be measured event

by event and plotted in a 2 dimensional array (called the total energy plane

or TEP). Since the total available ECM is constant for a given channel, events

corresponding to a channel where all the evaporated particles have been measured

should lie on a straight line in this TEP (see figure 4.16).

Tot

al P

artic

le K

inet

ic E

nerg

y

Total gamma-ray energy

3p

Ecm

Ecm

a3p 3pn

4p

etc

Figure 4.16: Schematic Total Energy Plane for a 3p out channel showing a con-stant line for 3p events and clear separation from the higher multiplicity chargedparticle events [133].

Events from higher multiplicity channels where some of the Tpart has not

been measured will lie below this line on the TEP plane. Using a 4π charged

88

Page 92: nuc_expt_phr03

particle detector (to obtain the centre of mass energies of the emitted particles)

in conjunction with a 4π gamma-ray inner ball (to obtain an estimate for the

Hγ) one can increment such a plane event by event. By setting software gates

on various regions of the TEP plane, excellent channel selection between particle

multiplicities can be obtained.

Kinematic Corrections for Doppler Effects.

Figure 4.17: Spectra showing the SD band in 85Nb using the reaction58Ni(36Ar,α2p)85Nb at an energy of 180 MeV. The improvement in signal tonoise due to particle gating and kinematic corrections are clear [166].

By conservation of momentum, the evaporation of particles from the com-

pound system may have the effect of causing the recoiling nucleus to alter its

direction and (velocity) with respect to the beam direction. In the case of thin

target experiments, where the recoiling nucleus is not stopped in view of the

89

Page 93: nuc_expt_phr03

detectors, this spread in velocities and increase of the recoil cone can cause a

deteriation of the gamma-ray energy resolution [174].

This is particularly a problem for light nuclei (≤ 100) which are formed by the

emission of α-particles. However, by measuring the energy of the emitted charged

particles and converting back to the centre of mass energy, the momentum of

the evaporated particles can be obtained. By conservation of momentum, the

momentum of the recoiling nucleus (in the centre of mass frame) can then be

reconstructed on an event-by-event basis. By converting this back into the lab

frame, the change in direction and velocity of the recoil due to the charged particle

emission can be calculated and used to improve the gamma-ray energy resolution

of the emitted gamma-rays [131].

The measured gamma-ray energy, Eγ will be shifted from its correct energy,

E0, by (a) the Doppler shift and (b) the change in direction/velocity due to the

emission of light charged particles. In general, for i emitted particles of mass mi,

this can be written as

Eγ = E0

(

1 + ∆E0 −n∑

i=1

∆Ei

)

(4.2.10)

where

∆E0 =MCN

MR

vCN.dγ

c(4.2.11)

and

∆Ei =Mi

MR

vi.dγ

c(4.2.12)

where vCN.dγ is the vector dot product between the velocity vector of the

recoil and the unit vector of the direction of gamma-ray emission (ie. VCN times

the cosine of the angle between the recoil and gamma-ray directions in the lab

frame).

Figure 4.17 shows the improvement in energy resolution (and thus peak to

total) obtained for the superdeformed band in 85Nb by kinematically correcting

the spectra for charged particle evaporation.

90

Page 94: nuc_expt_phr03

4.3 Neutron Detection.

The detection of neutrons evaporated from compound nucleus reactions is usu-

ally achieved using organic scintilators containing a large amount of hyrdrogen.

A common type of detector consists of the liquid scintilator NE213 [136]. Dif-

ferentiation between neutron and gamma-ray induced events is usually achieved

by one of two methods: (a) time of flight and/or (b) pulse shape disrimination.

Figure 4.18 shows a time spectrum for events in a neutron detector with respect

to the RF of the beam pulse for the reaction 40Ca+24Mg from the AYEBALL

array at Argonne National laboratory. The gamma-ray induced events all have

a common time profile (since all photons travel at the same speed), while the

neutrons are slower and arrive later. Note the time spectrum can also be thought

of as a neutron energy spectrum.

Figure 4.18: Neutron time of flight spectra showing clear separation betweengamma-ray and neutron event. From the reaction 40Ca+24Mg on AYEBALL[106].

In general, the larger the distance between the neutron detectors and the

target position, the better the TOF discrimination between gamma-rays and

neutrons. However, by placing the detectors further back, one loses detection

efficiency. In order to maximise the neutron detection efficiency, the detec-

tors must be placed closer to the target position and pulse shape discrimination

[136, 127, 125] is used to separate gamma and neutron events (see figure 4.19).

As with the phoswich detector, the prompt component of the scintilator light

is particle dependent (neutron induced, proton scattering events have a larger

prompt component than gamma-rays events). (Note that the neutron detectors

91

Page 95: nuc_expt_phr03

can also be used as a gamma-ray multiplicity detector [136]).

gammasneutronsN

eutr

on T

ime

of F

light

Zero-crossing time

Figure 4.19: Time of flight verses z component for neutrons measured in the28Si+40Ca reaction at PEX [106].

In the studies of very neutron deficient nuclei, it is often desirable to combine

both charged particle and neutron channel selection devices to obtain clean, high-

resolution gamma-ray spectra for a specific, weakly populated channel. In modern

arrays, different forms of channel selection are often used in tandem to give a

maximum effect. A good example of this is the use of both charged particle

inner balls and neutron detectors to study very neutron deficient nuclei around

N∼Z∼100 at the Niels Bohr Institute, Denmark [126, 127, 128, 129]. The PEX

array (see figure 4.21) is an extension of this project, as is the EUROBALL

neutron wall [130].

Figure 4.20 shows the effect of using both charged particle and neutron se-

lection in identifying transitions in the neutron deficient nucleus 6231Ga from the

reaction 40Ca+28Si, via the αpn evaporation channel. It is clear that the signal to

92

Page 96: nuc_expt_phr03

noise for the lines associated with the nucleus of interest improves dramatically

with each extra level of channel selection.

Once gamma-rays from such a channel have been identified, gamma-gamma

coincidence data such as that shown in figure 4.22 can be used to work out the

decay sequence of the transitions to form a decay scheme.

93

Page 97: nuc_expt_phr03

Figure 4.20: Particle and neutron identification spectrum showing different resid-ual channels in the reaction 28Si+40Ca from PEX [106].

94

Page 98: nuc_expt_phr03

Figure 4.21: Schematic of the PEX array, used in conjunction with a small, siliconinner ball and a wall of neutron detectors.

95

Page 99: nuc_expt_phr03

Figure 4.22: Alpha-neutron gated γ − γ coindicence gates showing transitions in62Ga from the reaction 28Si+40Ca taken from PEX [106]. The matrix was gatedon the condition that one α-particle and either one proton and/or one neutronwere also detected.

96

Page 100: nuc_expt_phr03

Determining Charged Particle and Neutron Multiplicity.

Figure 4.23: Efficiency corrected spectra used to identify states in 61Co. (a) Thepre-scaled γ singles data, (b) proton–γ (c) proton–neutron–γ and (d) the α − γcoincidence spectra from the 16O+48Ca reaction. [25].

For evaporation channels containing more than one type of the same particle,

the probability of measuring at least one of these particles increases with the

particle multiplicity. For example, the probability of measuring a proton from

a 2p channel is approximately twice as large as measuring a proton from a 1p

channel, simply because in the former case there are twice as many ‘chances’ to

measure the proton. This can be useful when identifying gamma-rays with specific

evaporation channels. By comparing the intensities of gamma-ray transitions in

the raw, ungated spectra with the same spectrum gated by the condition of

measuring one type of particle, the various evaporation channels fall into discrete

groups depending on the evaporated particle multiplicity of that channel [126,

25, 127, 128].

Figure 4.23 shows identification spectra for the reaction 16O+48Ca gated on

charged particles and neutrons taken from the U. Penn 4π array [110, 25]. The

main channels of interest were the pxn channels to the 62Co and 61Co channels

97

Page 101: nuc_expt_phr03

via the pn and p2n channels respectively. The ratio of counts for lines between

the 1p and 1p1n gated spectra are shown in figure 4.24 and the lines from the

different neutron multiplicities are clearly separated.

Figure 4.24: Plot of ratio of relative intensities of states in proton–neutron–γ andproton–γ spectra, identified in (a) the 16O+48Ca and (b) the 18O+48Ca reactions.Note how the neutron multiplicities, used to identify the residual decay productsare clearly separated and the effect of differing Q-values on the neutron detectionefficiency in the two reactions [25].

Figure 4.24 also shows the same plot for the evaporation channels for the

same target but using an 18O reaction. Note that the apparent neutron efficiency

is different for the 16O induced reaction. This is a kinematic focussing effect

due to the different Q-values for the two reactions. Thus the recoil cones of the

neutrons in the lab frame for the two reactions are different. In the 18O case,

the evaporated neutrons typically have higher energies than those in the 16O

98

Page 102: nuc_expt_phr03

reaction. Thus the neutrons from the 18O reaction are less forward focussed in

the laboratory frame. Since the neutron detector was placed at forward angles,

their detection efficiency is reduced relative to the 16O data.

99

Page 103: nuc_expt_phr03

4.4 Recoil Detectors.

As shown above, the use of ancillary detectors to measure evaporation products

from compound nucleus reactions results in a dramatic improvent in the signal to

noise and allows the identification of transitions in channels produced with low

cross-sections (≤ 10µb). However, one problem with such a method is that one

is not actually measuring the nucleus of interest but rather inferring it from the

evaporated particles. Isotopic impurities in the target at the level of only one per

cent of so can can cause problems with identification of very weak channels due

to the large differences in for example, neutron evaporation probability between

isotopic compound nuclei differing by only a few neutrons.

A more direct way of determining which nucleus specific gamma-ray tran-

sitions come from is to detect the recoiling nucleus itself. This is the premise

behind the use of recoil separator type devices. Their main job is to collect the

recoils and separate them from the (much more intense) flux of beam particles. In

addition, there may be other contaminant nuclei which one would like to remove

such as products from fusion-induced fission or deep-inelastic type reactions. The

rejection of non-evaporation type events is usually achieved by passing the par-

ticles through some form of electric and magnetic fields which separate the ions

by their mass over charge state (AQ

). There are usally two electro-magnetic fields,

one to do the main job of ‘dumping’ the beam and the second one to give disper-

sion in the focal plane with the AQ

of the evaporation residues. There is usually a

position sensitive detector (such as a PPAC or microchannel plate) at the focal

point and the position of the recoil on this detector yields information on its AQ

ratio.

4.4.1 Recoil Mass Separators.

The Daresbury recoil mass separator [143] used two Wien filters (crossed magnetic

and electric fields) to separate the beam from the evaporation residues. Crossed

E and B fields are velocity filters and will only allow through particles with a

certain velocity. Since the recoils are slower than the beam particles, setting the

fields so that the residues passed through the Wien filter means that the faster

beam particles are deflected. The final AQ

separation was achieved using a dipole

magnet which dispersed the recoils along a focal plane where they are detected

100

Page 104: nuc_expt_phr03

using a microchannel plate.

Three more modern devices are the FMA or Argonne Fragment Mass Analyser

[144] (see figure 4.25), the Oak Ridge RMS [145] and CAMEL [146] at the INFN,

Legnaro, Italy. They work on the basis of two electric dipole fields which give

the beam separation with a magnetic quadrupole magnet in the middle to give

steering and focussing.

DipoleBeamStop

ElectricField

MagneticField

ElectricField

NeutronDetectors

Target

Ion-Chamber

Dipole

F.M.A

Chamber

Ge Detectors

AYEBALL

LinacTandem

ATLAS

P.P.A.C

Figure 4.25: Schematic of the AYEBALL array used in conjunction with an arrayof neutron scintilator detectors, the Argonne Fragment Mass Analyser and a splitanode ionisation chamber [29].

One problem with recoil separator devices is that the recoiling ions have a

distribution of charge states when they leave their thin production target. The

dispersion allowed by the final dipole magnet and the limit of the size of the

postion detector (micro-channel plate or PPAC) typically limits to twothe number

of charge states which can be focussed through such a device and identified at the

final focus. It is usual in experiments using recoil separators to initially perform

a charge state sweep, where the field settings of the electro-magnetic devices

are set to the charge state which will maximise the yield of a specific isobar.

Figure 4.26 shows the results of a charge state sweep for A=87 recoils for the

reaction 36Ar+54Fe from the FMA.

Clearly, there will be a susbantial loss in detection efficiency due to the fact

that not all ionic charge states can be measured.

However, such devices do allow very clean isotopic identification of recoils with

mass resolutions of upto one part in 300. Figure 4.27 shows a two-dimensional

spectrum taking from the FMA of the x-position on the PPAC verses the energy

loss in the PPAC for the reaction 24Mg+40Ca. Clear identification of different

isobaric species is allowed. Note the presence of two charge states on the focal

plane.

By setting gates around these mass locii in offline sorting, one can associate

gamma-ray transitions with specific recoils. Figure 4.28 shows the results of these

101

Page 105: nuc_expt_phr03

Figure 4.26: Results of a charge state sweep for A=87 recoils for the reaction36Ar+54Fe [29].

mass gates for the 24Mg+40Ca reaction and there is a clear difference between

the spectra.

However, due to the fact that there is some overlap between the mass locii in

the focal plane, there are some bleed throughs from one mass gate to another.

This is a problem when the contaminant nucleus is produced much more readily

than the one of interest (eg. the production cross-section for the A=61 recoils

is an order of magnitude greater than for the A=62). If there are sufficient

statistics this problem can be overcome by subtracting normalised portions of

higher mass spectra from the spectrum of interest to obtain ‘clean’ or isobarically

pure gamma-ray identification spectra as shown in figure 4.29

Z-Separation Using Split Anode Ionisation Chambers.

The discussion above shows how one can obtain the mass of a nucleus using

a separator. However, for full identication, the atomic number must also be

obtained. In the case of high-Z nuclei, where the main contamination comes

from the large fission background, it is often possible to determine the Z of the

specific recoils by the energies of the co-incident X-rays (associated with electron

conversion decays) observed in the gamma-ray identification spectra [48, 45, 46].

However, for lower-Z recoils, where the X-ray energies are much decreased, a

different way of acertaining the Z of the recoil must be used.

One successul method has been to put a split anode ionisation chamber behind

the position focus of a separator. The energy loss in the first part of this chamber

102

Page 106: nuc_expt_phr03

∆Ε

A/Q

5762

Q=11

61

60

59 58 59

Q=10

Figure 4.27: Argonne FMA two-dimensional spectrum of X-position versus energyloss in the PPAC for the reaction 40Ca+24Mg showing clear separation of thevarious isobars. Note the presence of two charge states on the focal plane [29].

(∆E) can then be plotted against the total energy deposited in the chamber by

the recoil E. If the velocity of the ions is high enough, it has been shown that

recoils of the same mass but different Z will deposit different amounts of energy

in the first part of the ion-chamber [154, 155]. The ion chamber associated with

the Daresbury recoil separator [143] was used very successfully to identify the

first excited states in very neutron deficient N=Z nuclei from 64Ge upto 84Mo

[139, 140, 141, 142] by detecting the very weakly populated 2-neutron evaporation

channel. The ion-chamber was used to separate out events from the same mass

but from the (much larger) 2p and pn channels.

Figure 4.30 shows the ion-chamber ∆E signal for A=58 gated recoils from the

reaction 40Ca+24Mg from the AYEBALL array with the FMA. There is a clear

separation between the ∆E signals associated with the 58Ni and 64Zn (6411

∼ 5810

recoils from the α2p ands 2p2n+44Ca channels resepctively.

103

Page 107: nuc_expt_phr03

Figure 4.28: Raw, mass gated spectra from the reaction 40Ca+24Mg [29].

By gating on these different portions of the ion-chamber ∆E spectrum in

offline sorting and performing normalised subtractions of contaminants to decon-

volute the different overlapping channels, one can achieve the isotopically ‘pure’

identification spectra shown for 58Ni in figure 4.31.

104

Page 108: nuc_expt_phr03

Figure 4.29: A=58 gated spectra from the reaction 40Ca+24Mg with and withoutsubtractions from higher mass contaminants [29].

Figure 4.30: Rotated ∆E ion chamber signals gated on known transitions in 58Niand 64Zn. The separation by the Z of the recoil is clearly demonstrated [29].

105

Page 109: nuc_expt_phr03

Figure 4.31: Gamma-ray spectra from the AYEBALL array for the reaction40Ca+24Mg showing the effect of FMA and ion-chamber gating. The transitionfrom the A

Q= 5.8 recoils 58Ni and 64Zn are clearly separated [29].

106

Page 110: nuc_expt_phr03

4.4.2 Gas Filled Separators.

The main problem with recoil separator type devices is that since they trans-

mit the recoils in more than one charge state, they have a limited transmission

efficiency (≤ 10 → 15 %). In cases where the main aim of the experiment is

to simply identify the fusion-evaporation events and separate them from beam

particles and possibly fission background, the transmission efficiency can be dra-

matically improved by using a gas filled separator such as RITU (see figure 4.32

[147]).

Figure 4.32: Design drawing of the RITU gas filled separator [147].

RITU is a QDQQ type device (ie. a quadrupole magnet for collection and

focussing, followed by a dipole magnet for the separation, followed by two further

quadrupole magnets for final focussing. In vaccum mode, the mass resolution of

this device is around a part in 100.

The main idea behind such a device is to use an dipole magnet to separate the

fusion-evaporation products from the background, as in the usual vaccum-mode

recoil separators, but to fill this field region with a dilute gas such as helium.

Collisions between the reaction products and the gas atoms lead to a charge state

107

Page 111: nuc_expt_phr03

focussing effect. Ions then follow trajectories approximately determined by the

average charge state of the ion through the gas (independent of the initial charge

state of the ions as they exit the thin production target).

The magnetic rigidity of the dipole magnet for a gas filled separator is given

approximately by the expression [147, 148]

Bρ =mv

eqav=

mv(

evv0

) 1

3

= 0.0227A

Z1

3

Tm (4.4.13)

where mv is the momentum of the reaction product and vo is the Bohr velocity

(=2.19×106 ms−1). Typical gas pressures of about 1mb of helium are used.

4.4.3 Recoil Decay Tagging.

If the neutron deficient nucleus of interest has a substantial decay width by either

α-decay or direct proton emission, this can be used as an experimental tag with

which to identify gamma-ray transitions from excited decays in that nucleus.

This technique is known as recoil decay tagging [156, 157, 158, 159, 160, 161].

The basic idea is shown schematically in figure 4.33.

beam

EDMD

ED

8.2m 0.4m

implantedrecoils

PPACFMA

target

AYEBALL

DSSD

Figure 4.33: Schematic of using the FMA for RDT experiments [159].

Fusion residues are selected using online mass separation such as the FMA or

RITU. A pixellated charged particle detector is placed at the final focus of the

separator a charged particle detector, which gives good position resolution for the

incoming recoils. Gamma-rays emitted from fusion-evaporation reactions can be

correlated with an evaporation residue detected at the focus of the separator.

The pixel that the recoil hit is recorded and if a second event is measured in the

same pixel (but without a recoil signal through the separator) it is assumed that

this second event is the α (or proton) decay of the recoil.

108

Page 112: nuc_expt_phr03

Figure 4.34: Proton and alpha-particle energy spectra from the decays of theproducts from the 92Mo+58Ni reaction at a beam energy of 260 MeV [161].

Since α and direct proton decays have discrete energies, these are definitive

signatures of a particular nucleus and can be correlated with specific recoils by

their position in the DSSD detector. The limiting factor in such experiments is

that the average time between 2 recoils hitting the same pixel should be large

compared to the lifetime of the α or direct proton decay.

Using this method, in-beam spectroscopy of neutron deficient nuclei with

fusion-evaporation cross-sections as low as 10 µb can be achieved [158].

Figure 4.36 shows the comparison between the recoils mass gated and RDT

gated γ − γ coincidence spectra for very neutron deficient Rn isotopes. The

extra degree of channel selection afforded by gating on the alpha-decay of the

compound nucleus allows this weak channel to be identified.

4.4.4 Recoil Filter Detectors.

Another way of improving the signal to noise and measured resolution of gamma-

ray lines emitted in fusion-evaporation reactions using a thin target is to detect

the recoils directly using an array of thin scintilators [162, 163] to detect those

recoils which are scattered in the target out of the beam direction and into the

recoil cone.

Figure 4.37 shows schematically how such a detector works. Using a pulsed

beam, one can measure the time of flight for the recoils (and other reactions

products such as scattered beam, fission products, etc) to reach the filter detector.

109

Page 113: nuc_expt_phr03

Figure 4.35: (a) Gamma-ray spectra gated on A=147 residues by the FMA, (b)Gamma-ray spectra gated by the tagged proton decaying from the h 11

2

of 147Tm

(c) the proton d 3

2

isomeric state in 147Tm [161].

The (slower moving) evaporation residues can be selected in off-line anaylysis by

gating on this time of flight signal. Note that if the target/detector distance

is known, this time of flight gives a direct measure of the recoil velocity. This

coupled with the angular information obtained by knowing which recoil filter

detector element was hit, allows excellent correction for Doppler broadening due

to the spread in recoil cone [163].

This type of detector has also proved immensely useful in identifying low

cross-section evaporation products in heavy nuclei where the background from

fission products is a large problem [47]. In these cases, fission can make up 99%

or more of the total fusion yield.

110

Page 114: nuc_expt_phr03

0

2

4

6

8

100 200 300 400 500 600 700 800gamma energy (keV)

0

100

200

300

400

500

coun

tsco

unts

selected usingrecoil’s alpha decay

selected usingrecoil mass only(for comparison)

433 keV

504 keV

Hf Coulex176

(4 2 )

(2 0 )+ +

+→ +

Figure 4.36: Comparison of FMA-γ − γ projections for (a) no recoil conditionand (b) using RDT using the reaction 176Hf+28Si at a beam energy of 142 MeV[159].

111

Page 115: nuc_expt_phr03

RFD

evap.residues.

pulsedbeam

gamma-rayarray

γ

beam

scat

tere

d be

am

evap

.re

sidu

es

coul

exfi

ssio

n

Time of Flight (ns)

0 100 200

Figure 4.37: Schematic of the design and operation of the recoil filter detectorwhich uses time of flight to distinguish between evaporation residues and un-wanted background events.

112

Page 116: nuc_expt_phr03

Chapter 5

Measurement of Lifetimes ofBound Nuclear States.

In this chapter we will investigate various methods for the measurement of nuclear

lifetimes. (The review by Nolan and Sharpey-Schafer [202] is an excellent source

of information on these topics). The deduction of lifetimes is important for a

number of reasons. The lifetime of the nuclear state τ is related to its intrinsic

width by the Heisenberg uncertainty principle such that

Γτ ≥ h (5.0.1)

The probability of decay is proportional to the intrinsic energy width Γ and

depends soley on the matrix element between the initial and final states and the

operator which governs the decay between them, such that [202]

Γ α | < φf |M |φi > |2 (5.0.2)

where M is the operator for the decay and φf and φi are the wavefunctions

of the initial and final states respectively. The lifetimes of the decay thus re-

veal information on the nature of the states. The lifetime can be compared with

Weisskopf single particle estimates to help deduce the spin difference between

the initial and final states, or if this is already known (from for example an-

gular distribution/correlation data) can be used to deduce other effects such as

an enhancement in E2 strength consistent with a highly collective (deformed)

structure.

By measuring the lifetime of a nuclear state, one is really measuring the decay

probability from one quantum state to another. For electromagnetic decays, the

113

Page 117: nuc_expt_phr03

transitions probability from a state Ji to a state Jf (summed over all possible

magnetic substates) by a transition of energy Eγ is given by [7]

Tfi(λL) =8π(L+ 1)

hL ((2L+ 1)!!)2

(

hc

)2L+1

B(λL : Ji → Jf ) (5.0.3)

where B(λL : Ji → Jf) is called the reduced matrix element.

Measuring the lifetime (decay probability) of a nuclear state thus gives a value

for the B(λL : Ji → Jf).

For lifetimes, τ in units of seconds where the transition probability per unit

second, T = 1τ, (Eγ in MeV),

T (E1) = 1.587 × 1015E3γB(E1) (5.0.4)

T (E2) = 1.223 × 109E5γB(E2) (5.0.5)

T (E3) = 5.698 × 102E7γB(E3) (5.0.6)

T (M1) = 1.779 × 1013E3γB(M1) (5.0.7)

T (M2) = 1.371 × 107E5γB(M2) (5.0.8)

T (M3) = 6.387 × E7γB(M3) (5.0.9)

The units ofB(Eλ) are e2fm2λ and the units ofB(Mλ) are (eh2Mc)2 (fm)2λ−2.

Due to the long range involved in the lifetimes of nuclear states, different

techniques must be employed in order to measure nuclear states. In this chapter

we will deal with methods to deduce nuclear lifetimes in the region from 10−15 →10−3 seconds.

5.0.5 Weisskopf Single Particle Estimates.

Lifetimes of nuclear states are sometimes described in terms of of Weisskopf units

(W.u.), which give a yardstick as to the lifetime range expected for a typical decay

of a fixed multipolarity. The Weisskopf single particle estimates are based on a

114

Page 118: nuc_expt_phr03

single proton in a spherical orbit. The expressions for the single particle estimates

for the reduced transition probabilities are given in table 5.1. To convert these

into lifetimes, simply substitute in the single particle B(Mλ) values into the

equations for the transition rates given in the previous section.

The general equations for the single particle estimates for the reduced prob-

ability matrix element are [35],

B(Wu :EL) =1.22L

(

3

L+ 3

)2

A2L3 e2fm2L (5.0.10)

for electric transitions and

B(Wu :ML) =10

π1.22L−2

(

3

L+ 3

)2

A2L−22

(

eh

2Mc

)2

fm2L−2 (5.0.11)

for magnetic ones. M is the single nucleon mass and A is the atomic mass

number.

Transition Multipolarity T 1

2

(1 spu) (seconds)

E1 6.76 × 10−6E−3γ A− 2

3

E2 9.52 × 106E−5γ A− 4

3

E3 2.04 × 1019E−7γ A−2

E4 6.50 × 1031E−9γ A− 8

3

M1 2.20 × 10−5E−3γ

M2 3.10 × 107E−5γ A− 2

3

M3 6.66 × 1019E−7γ A− 4

3

M4 2.12 × 1032E−9γ A−2

Table 5.1: Weisskopf single particle estimates for transition half-lives [190, 191].Note, Eγ is in keV and A is the mass number.

Typcal recommended upper limits rate for the different types of transitions

can be found for the different mass regions in references [192, 193, 194, 195].

5.0.6 Determining Nuclear Quadrupole Deformation fromLifetimes of E2 Transitions.

For deformed nuclei, the deformation, or deviation from sphericity, is related to

the intrinsic quadrupole moment of the nucleus, Qo, which is in turn related to

the B(E2) of the collective, E2 transitions in the system by the expression,

115

Page 119: nuc_expt_phr03

B(E2) =5

16πQ2

o| < JiK20|JfK > |2 (5.0.12)

where < JiK20|JfK > is a Clebsh-Gordon coefficient for the transition de-

caying from a state of spin I to one of I − 2 is given by

< JiK20|JfK >=

3(J −K)(J −K − 1)(J +K)(J +K − 1)

(2J − 2)(2J − 1)J(2J + 1)(5.0.13)

Thus by susbtitution, the lifetime is related to the quadrupole moment by the

expression [53],

1

τ= 1.223E5

γ

5

16Q2

o| < JiK20|JfK > |2 (5.0.14)

The quadrupole moment can be related to the quadrupole deformation pa-

rameter β2 by the expression [108],

Q0 =3√5πZR2β2

1 +1

8

5

πβ2....

(5.0.15)

Therefore, assuming the rotational model, measuring the intrinsic lifetime of

a state in a stretched E2 cascade (rotational band) can allow the deduction (in a

model dependent way) of the nuclear deformation.

5.1 Electronic Timing Methods.

The law of radioactive decay states that at time t, the number of nuclei left in a

particular state of lifetime τ (given that there were N0 at t = 0), is given by

N(t) = N0exp(

− t

τ

)

(5.1.16)

If the lifetime of the nuclear state we wish to measure is long compared to

the intrinsic timing properties of the (germanium) detector, one can simply use

pulsed beam techniques to determine the number of decays of a state as a function

of time.

The general experimental idea is summarised in figure 5.1. A thick or backed

target is irradiated by a beam to form the nucleus of interest for a period of time

which is short compared to the lifetime of the isomer we wish to measure. (Note

116

Page 120: nuc_expt_phr03

the recoil must stop in the view of the gamma-ray detectors). The beam-on period

is also usually used to start a clock, ie. to define a population at time t = 0. The

beam is then switched off for a period and gamma-rays decaying from the isomer

(called delayed, or out-of-beam decays) are measured in the germanium detectors.

The relative time difference between when the target was irradiated and when

the gamma-ray is measured (usually done with a time to amplitude converter

or TAC, or a time to digital conver, TDC). Over a period of time, a full time

(decay) spectrum for the state can be obtained. For a single, long lived state,

with no long lived feeding transitions, the time spectrum will be an exponential

(ie. a straight line on a log scale), which can be simply fitted to the radioactive

decay law to obtain the lifetime of the state.

prompt, in beam gammas

isomer, T1/2 > few ns

(in delayed portion of TDC/TAC)delayed, out of beam gammas

(in prompt peak)

time

Log

N

time

Log

N

beam off period(delayed gammas)

next beam pulse,

beam

inte

nsity

stop clock

beam pulse(prompt gammas)

start clockreset clock

t=0

Figure 5.1: Schematic using pulsed beams to measure nuclear lifetimes electron-ically.

Electronic timing methods have been extensively used to measure lifetimes of

K-isomers in the mass 130 [93, 94, 38] and 180 regions [83, 86, 87, 84, 9].

Figure 5.2 shows the time spectra gated by gamma-ray lines below the T 1

2

=6µs,

Kπ = 8− isomer in 138Gd [93]. Figure 5.3 shows the gamma-gamma coincidence

117

Page 121: nuc_expt_phr03

data for 138Gd lines observed out of beam. The first 4 decays in the yrast band

are observed, together with the 583 keV transition which decays out of the iso-

mer. Note that the 10+ → 8+ 616 keV decay previously observed in prompt

decay studies [189] is not present.

138Gd

counts

time [ s]µ

0 10 20 30010

110

210

Figure 5.2: Decay curve following the decay of the T 1

2

=6 µs, Kπ = 8− isomer in138Gd [93].

118

Page 122: nuc_expt_phr03

138Gd

counts

energy [keV]0 200 400 600 800 1000

0

100

200

221

<---- 384 (gate)

489

556

583 0.0 0 +

220.8 2 +

605.2 4 +

1094.3 6 +

1649.9 8 +

2233.1 (8 ) -

220.8

384.4

489.1

555.6

583.2

T = 6 1/2 s

Figure 5.3: Out of beam, gamma-ray coincidence spectrum showing transitionsin coincidence with the 384 kev 4+ → 2+ transition in 138Gd [93].

119

Page 123: nuc_expt_phr03

Other examples of such work are to be found in studies of high spin, yrast

trap decays in trans-lead nuclei by electric octupole (E3) decays [196, 197, 246,

198, 199, 200, 201]. A particularly good example of this technique at high spins is

the proposed 8.5 MeV, spin 34+ state in 212Fr, which decays by an E3 transition

with a meanlife of 34±3 µs [196].

5.1.1 Gamma-ray Spectroscopy Across Isomers

In order to establish prompt decays into isomeric states one needs to be able to

correlate coincidences between transitions above and below isomeric states. For

lifetimes below around 20 µs, this is possible using a pulsed beam (with the beam

pulses separated by the order of 2 µs) by allowing each gamma-ray detector to

have its own individual time signal with respect to the beam pulse (or other fixed

time reference such as an RF signal).

Figure 5.4: Single TDC and TDC difference spectra for reaction 11B+176Yb usinga pulsed beam of width 1 ns separated by 1400 ns [292]. Note the clear separationbetween in-beam and out-of-beam event and the well defined regions correlatingto co-incidences above and below isomeric states.

120

Page 124: nuc_expt_phr03

The upper portion of figure 5.4 shows a typical TDC time spectrum for

gamma-rays detected in a pulsed beam measurement. Note that the majority

of counts occur during the ‘in beam’ period. The bump to the right of the

prompt, in-beam peak is due to low energy gamma-rays (or X-rays) which occur

in the beam pulse but have poor timing properties (due to poor charge collection

times in the detector). These thus ‘walk’ out of the prompt time gate. This effect

can be corrected for in off-line software analysis.

Figure 5.5: (a) Spectrum of ‘earlies’ obtained from a sum of transitions belowthe 12+ isomer in 106Cd. (b) Out-of-beam,prompt-γ − γ spectrum gated on the633 keV 2+ → 0+ transition in 106Cd [3].

One can software gate on the various portions of this spectrum to define those

gamma-rays which occur ‘in-beam’ or ‘out-of-beam’. It is thus possible to create

a gamma-gamma coincidence matrix of gamma-rays which occur in the in-beam

period correlated by delayed coincidences with gamma-rays which are detected

in the out-of-beam period. By gating on in-beam or out-of-beam portions of

the TDC singles spectra, one can determine whether a gamma-ray was measured

in-beam (early) or out-of-beam (delayed) (see figure 5.5).

121

Page 125: nuc_expt_phr03

0 0 +

633 2 +

1494 4 +

6 + 24922504 6 +

8 + 3044

3367 8 +3354 7 +

8 + 3788

4121 9 +

4816 10 +

5419 12 +

6227 14 +

7119 16 +

8100 18 +

9250 20 +

10561 22 +

12049 24 +

13726 (26 ) +

15584 (28 ) +

21052331 5 +

7 + 3084

3409 7 - 9 - 3678

4324 11 -

5214 13 -

6264 15 -

7518 17 -

8884 19 -

10350 21 -

11941 23 -

5 - 2629

3319 6 -3507 8 -

4106 10 -

4967 12 -

5976 14 -

7121 16 -

8411 18 -

9877 20 -

4436 10 +4660 12 +

5253 (13 ) +

5573

5987 5912

55585770

6101

6858

7480

1716 2 +

2371 3 -

4183 8 +

9319 + X (18 ) +

9722 + X (19 ) +

10161 + X (20 ) +

10664 + X (21 ) +

11168 + X (22 ) +

11740 + X (23 ) +

12311 + X (24 ) +

12951 + X (25 ) +

13614 + X (26 ) +

14322 + X (27 ) +

15065 + X (28 ) +

15861 + X (29 ) +

632.8

861.3

997.7

874.7

1009.4

540.9 552.8

1284.5

1076.7

1028.4695.3

703.3

1471.5

602.4

807.9

892.3

980.8

1150.6

1310.6

1487.6

1677.6

1857.4

433.4

862.31295.9

610.6

225.8

836.4

592.5 754.2

171.1311.6633.9

269.1

906.0 917.6 780.6

1134.8

524.2

690.5422.

8 187.6

827.4

598.3

861.3

1008.6

1145.3

1290.1

1466.0

645.6

889.8

1050.4

1253.6

1366.5

1465.8

1591.4

315.0335.7223.6

592.8

659.7319.9

414.2

304.9

542.6

517.7

330.5

757.7

621.9

403.6

1622.6

438.5

502.6

504.4

572.5

571.0

639.7

663.3

718.8

733.0

795.6

842.4

1007.0

1143.8

1302.6

1452.1

941.5

1077.2

1211.4

1382.3

1529.0

= 90 ns τ

Cd106 48 58

14

3

2

Figure 5.6: Decay scheme for 106Cd [3].

122

Page 126: nuc_expt_phr03

By taking the time difference between two detectors, one can create the type

of time difference spectrum shown in the lower portion of figure 5.4. If for example

one wanted to look at gamma-rays in prompt coincidence with each other, but

delayed with respect to the beam pulse (to look at decay cascades below an

isomer), one would gate on the prompt-prompt region of the TDC difference

spectrum (thus ensuring that the gamma-rays were in prompt coincidence with

respect to each other) but with the extra condition on the individual TDC spectra

such that the transitions were measured out of beam.

Similarly, one can put the condition on the TDC difference spectrum to look

for coincidences between early and delayed gamma-rays. Figure 5.5 shows the

gamma-ray spectra for transitions above and below the 12+ isomer in 106Cd [3].

counts

time [ns]0 500 1000 1500

110

210

310

410

510

610

1950(150) ns

145,170,194,217,236 and 710 keVb)

counts

time [ns]0 200 400 600 800

010

110

210

310

410

222(8) ns

145, 170 or 194 keV (start)131 keV (stop)

a)

Figure 5.7: Time spectra for isomeric decays in 175Ta (a) shows the lifetime of

the 92

−[514] isomer in 175Ta (TDC difference spectrum) (b) shows the lifetime of

the 212

−isomer in 175Ta (pulsed beam TAC singles spectrum) [87].

Using the gamma-gamma-time coincidence techniques, one can gate on any

123

Page 127: nuc_expt_phr03

two gamma-ray energies in a gamma-gamma-time-difference cube (3D-coincidence

matrix) and project the time difference between any two gamma-rays. Figure 5.7

shows the time difference spectra gated by transitions above and below the 92

isomer in 175Ta [87]. The upper portion of figure 5.7 is a TDC difference spec-

trum gated on transitions above and below the 228 ns 92

−isomer, while the lower

spectrum is a singles TAC spectrum showing the lifetime of the 212

−, 1950 ns

isomer [87]. For reference, the decay scheme of 175Ta is shown in figure 5.8.

619 13/2 +

0 7/2 +130 9/2 +

284 11/2 +

461 13/2 +

658 15/2 +

872 17/2 +

1101 19/2 +

1341 21/2 +

1592 23/2 +

1850 25/2 +

2118 27/2 +

2394 29/2 +

2681 31/2 +

2979 33/2 +

3287 35/2 +

3611 37/2 +

3947 39/2 +

4297 41/2 +

4661(43/2 ) +

5037(45/2 ) +

9/2 - 5/2 -

131 9/2 -

276 11/2 -

446 13/2 -

640 15/2 -

856 17/2 -

1093 19/2 -

1350 21/2 -

1621 23/2 -

1909 25/2 -

2205 27/2 -

2515 29/2 -

2825 31/2 -

3143 33/2 -

3457 35/2 -

1552 17/2 +1651 19/2 +

1793 21/2 +

1969 23/2 +

2173 25/2 +

2402 27/2 +

2656 29/2 +

2931 31/2 +

3225 33/2 +

1566 21/2 -

1877 23/2 -

2202 25/2 -

2537 27/2 -

2879 29/2 -

3231(31/2 ) -

1729 21/2 +

1895 23/2 +

2086 25/2 +

2298 27/2 +

2531 29/2 +

2782 31/2 +

3052 33/2 +

3338 35/2 +

3640 37/2 +

3957 39/2 +

4282 41/2 +

4619(43/2 ) +

4966(45/2 ) +

1279 (15/2)

3216(31/2 ) -

3526(33/2 ) -

3762

4041

4329

4635

129.6

154.3

176.9

196.9

214.4

228.8

240.5

250.1

258.8

284.3

373.9

443.0

490.6

526.0

563.1

606.7

659.2

714.0

331.6

411.3

469.3

509.0

543.7

584.5

632.3

686.2

740.0

144.6

170.0

193.6

216.6

236.4

256.9

271.1

288.7

295.7

309.6

310.4

318.3

314.5

410.4

493.4

559.9

605.3

627.9

363.9

453.1

528.3

584.6

619.0

632.0

131.471.9

833.4

98.8

142.7

175.5

204.1

229.4

253.6

275.1

293.7

379.8

482.9

569.1

318.6

433.7

528.2912.1

695.7458.9

680.3

894.2

1090.6

932.4

311.2

325.2

334.5

342.5

351.9

635.9

677.2

659.5

694.0

216.5

473.3

709.6

165.7

191.0

212.4

232.3

251.3

269.5

286.6

302.1

316.3

325.5

337.4

346.0

403.7

483.3

555.9

618.7

662.9

357.1

444.4

520.4

588.8

641.1

684.0

163.4

235.7

279.1

288.0

305.8

646.8

336.4

678.8

310.4

Ta175 73 102

7/2 [404] +

5/2 [402] +

9/2 [514] - 222ns

5ns 1950ns0.9ns

( 0.5ns)

( 2ns)(35/2,37/2)

(37/2,39/2)

(39/2,41/2)

(41/2,43/2)

1/2 [541] - 123.451.5+X

( 1ns)

Figure 5.8: Partial decay scheme for 175Ta [87].

5.2 The Recoil Distance Method.

In order to measure lifetimes in the region between 10−8 → 10−12 s the most

commonly used method is the Recoil distance or ‘plunger’ method [202]. The

nuclei are formed in a thin (∼500 µg/cm2) target usually by a fusion-evaporation

reaction. The recoils then fly out of the thin target towards a thick, stopper or

‘plunger’ where they are stopped.

124

Page 128: nuc_expt_phr03

The Recoil Distance Doppler-Shift Method(RDM/RDDS)

Target Stopper

Eu E = E (1+ v/c cos( ))us θ

θ

Detector

Decay Curve

v ~ 1-2 % cv

u: unshifteds: shifted

d

Figure 5.9: Schematic of the Recoil Distance Method.

The idea behind this technique is shown graphically in figure 5.9. The premise

of the RDM is to measure the difference in the intensity of gamma-rays decaying

either in flight of when stopped in the plunger as a function of target-stopper

distance. Gamma-rays which are observed in a detector at angle θ to the recoil

direction will have their energies shifted by the Doppler effect such that to first

order,

Es(θ, t) = Eo

1 − vc

1 − vccosθ

≈ Eo

(

1 +v

ccosθ

)

(5.2.17)

125

Page 129: nuc_expt_phr03

where Eo is the unshifted transition energy and v is the recoil velocity.

The gamma-rays which are emitted between the target and the plunger will

decay from a recoil with velocity v, while those that decay while stopped in the

plunger will have their unshifted energy E0. The gamma-ray lineshape will be

split into two parts, the shifted and stopped component. The intensity in the

shifted peak Is is given by

Is = N0

(

1 − exp

(

− d

))

(5.2.18)

where N0 is the total number of decays (total number in the shifted and

stopped peaks) d is the recoil distance (ie. target–stopper distance) and τ is the

measured, apparent lifetime of the state. The nuclei which decay in the stopper

(emitting the gamma-ray at the stopped energy, E0) will give rise to a line with

intensity,

Io = Noexp

(

− d

)

(5.2.19)

Since v and d are known, the lifetime can be determined using R where,

R =Io

Io + Is= exp

(

− d

)

(5.2.20)

Descriptions of various plungers can be found in references [222, 202]. The

target-stopper distance can be measured using a micrometer screw [202] or using

the capacitance between target and stopper (in effect parallel plates).

126

Page 130: nuc_expt_phr03

Figure 5.10: Schematic of the NORDBALL plunger taken from [228] .

127

Page 131: nuc_expt_phr03

The velocity of the recoil can be calculated directly from the difference in

energy between the shifted and unshifted peak values at a given angle using

equation 5.2.17. Good examples of RDM experiments can be found in references

[217, 226, 227, 228, 229, 230]

5.2.1 Feeding Corrections and Gating From Above.

The observed lifetime of a nuclear state depends on the lifetimes of the states

which feed it. The relationship between the intrinsic lifetime of state and the

observed or apparent lifetime is given by the Bateman equations of radioactive

decay [203], such that

dNi(t)

dt=

N∑

j=1+1

Njλij −Ni

i−1∑

j=1

Niλi (5.2.21)

where Ni(t) is the population of level i at time t.

In order to deduced the intrinsic lifetime of a state from the measured or

‘apparent’ lifetime, one must first correct for the lifetimes of the states which

feed into the level of interest. In the case of fusion-evaporation reactions, often

much of the side-feeding intensity comes from unresolved continuum states of

which the lifetime is not known and this can cause problem with the fitting of

RDM data.

In singles measurements, the decay properties and intensities of all the feeding

transitions for the state of interest must be known in order to solve the Bate-

man equations. In practice, this is very difficult to achieve and can give rise to

erroneous values for the lifetimes. The advent of high-efficiency gamma-ray ar-

rays has somewhat alleviated this problem. If a number of discrete gamma-rays

are in a connected cascade, one can determine the feeding completely by setting

a gamma-gamma coincident gate on the shifted component of a line above (ie.

higher spin) than the state of interest. The coincidence requirement also has

the large advantage of cleaning the spectra up, thus reducing the possibility of

contaminant transitions giving false values for the shifted or unshifted intensities.

Figure 5.11 gives a good example of the use of the coincidence technique

in the RDM for high spin states in 110Cd [229]. the spectra are gated on the

shifted component of the 335 keV 10+ → 8+ yrast transition. Note the change

in the relative intensities of the stopped and shifted components for the various

128

Page 132: nuc_expt_phr03

lines with stopper distance The decay scheme for this nucleus and the fits to the

lifetime data are shown for reference in figures 5.12 and 5.13.

129

Page 133: nuc_expt_phr03

Figure 5.11: Coincident plunger spectra gated on the 10+ → 8+, 335 keV transi-tion in 110Cd. Taken from Piiparinen et al. [229].

130

Page 134: nuc_expt_phr03

0 +0

2 +658

4 +1542

6 +2480

8 +32758 +3440

10 +3611

12 +4172

14 +5026

16 +6100

18 +7325

6 –28968 –3056

10 –3824

12 –4931

8 –3428

10 –4182

509312 –

7 –2879

9 –3346

11 –4173

524913 –

15 –6181

5 – 25405 –2660

10 +4078

12 +4889

14 +5857

14 +5676

8 +3187

10 +4620

0 +1473

2 +1783

4 +2251

6 +2877

658

885

938

795

171

561

854

1075

1224

159

768

1107

755

911

467

827

1076

932

811

968

1433

467

626

1126

70813

3539

7

399563

802

637

707

424

787

1504

335164

960

265

1155

998

1117

399339219236356

177

531 548290

478

9.2 ps

< 3 ps

< 3 ps

< 4 ps

< 4 psτ = 670 ps

80 ps

1.0 ps

2.0 ps

12.0 ps

2.0 ps

< 1.5 ps

750 ps

71 ps

3.0 ps

< 2.0 ps

8.6 ps

1.5 ps

4.7 ps

5.0 ps

A B

C

S

G

E

F

H

Cd6248110

Figure 5.12: Decay scheme for 110Cd by Piiparinen et al. [229].

131

Page 135: nuc_expt_phr03

10+(3611) → 8+Gate: 561

τ = 670(35) ps

2+(658) → 0+Gate: 795 + 885 + 811

τ = 9.2(6) ps

12+(4172) → 10+Gate: 795 + 335

τ = 12.0(6) ps

9-(3345) → 7-Gate: 658 + 885

+ 938 + 399

τ = 71(4) ps

100Mo(13C,3n)110Cd E = 44 MeV Θ = 143°

Shifted peakUnshifted peakFit

Distance µm

Num

ber

of C

ount

s

10 100 1000 10 100 1000

102

103

104

103

104

102

103

104

102

103

104

Figure 5.13: Fits to RDM data for 110Cd from Piiparinen et al. [229].

132

Page 136: nuc_expt_phr03

5.2.2 The Differential Decay Curve Method.

The usual method of analysis of high-spin data in RDM experiments is to fit the

lifetimes of state including the effect of feeding from high spin states using the

Bateman equations. The individual lifetimes in a cascade are generally calculated

using a χ2 minimization program. This works well when the nature and intensity

of the levels feeding the state of interest are well known, however, if this informa-

tion is less well known, systematic errors can be introduced given erroneous values

for the calculated lifetimes. The Differential Decay Curve Method [231, 232, 218]

allows one to eliminate the effects of the feeding lifetimes for the transition by

gating from transitions above the transition of interest (see figure 5.14).

(gating transition)(detemines feeding)

side

feed

ing

"direct feeding transition"

level, i

decay from level of interest

Z

X

Y

XsXu

Zs

Zu

YuYs

Figure 5.14: Schematic of the gating in a cascade for use with the differentialdecay curve method.

The basis of the DDCM is that the Bateman equations can be reduced to the

expression [232, 218],

τi =−Ni(t) +

h bhiNh(t)d(Ni(t))

dt

(5.2.22)

where τi is the intrinsic lifetime of the state of interest, Ni is the population

of that state, Nh are the populations of those states which feed the state i and

bhi are the branching ratios for the feeding transitions. d(Ni(t))dt

is the differential

of Ni with respect to time at time t.

133

Page 137: nuc_expt_phr03

If there are three transitions in cascade, Y ,Z and X, (see figure 5.14) where

X is the transition out of the state on interest (level i), Z is the transition

which directly feeds state i and Y is a transition in the cascade above Z and

X, then the lifetime of the state i can be calculated directly, (ie accounting for

feeding) by setting a gamma-ray co-incidence gate on the shifted component of

the higher lying, transition Y and measuring the intensities of the moving and

shifted components of the transitions X (decaying from the state of interest) and

Y (directly feeding the state of interest).

µ31 m

300

0

200

0

300

0

200 240 250 290 320290

cou

nts

Energy [keV]

213 keV 256 keV 298 keV

0 20 40 60 80 100

50

0

150

100

0

100

200

300

400

d[ m]

co

un

ts

µ

Lifetimes in the SD band of Pb194

Decay out ~ 15%

decay curves (shifted)

µ22 m

µ97 m

298

256

213

R. Krucken et al. PRL 73, 3359 (94)

:

Figure 5.15: Decay curves for the decay of the SD band in 194Pb [219].

This is given by [232, 218]

τ(ti) =IXs(tk) − αIZu(tk)

d(IXs(tk)dt

(5.2.23)

where the I is the intensity measured in the gate on the shifted component of

the higher lying transition Y and the subscripts u and s correspond the unshifted

134

Page 138: nuc_expt_phr03

(stopped) and shifted components of the lineshape respectively.

α =IXu + IXs

IZu + IZs(5.2.24)

135

Page 139: nuc_expt_phr03

and

d(IXs(tk))

dt=IXs(tk+1) − IXs(tk−1)

tk+1 − tk−1(5.2.25)

and t = dv.

This tecnhique has the advantages [232] that (a) only directly measured in-

tensities are involved in the analysis, no lifetimes or feeding intensities have to

be known; (b) only flight-time differences are important in the analysis, the ab-

solute target-stopper distance is not involved, thus a systematic error in this

measurement does not affect the results

5.3 The Doppler Shift Attentuation Method.

If the lifetime of the state is of the same order of the slowing down time in a

target/backing (∼ 10−12 sec) then the RDM can no longer be used to determine

the lifetime (as the target–stopper distance has to be made too short). However,

if the experimenter has a knowledge of the slowing down process of the recoil in

the target (and target backing), one can use the lineshape of the transition of

interest as a function of detector angle to determine the nuclear lifetime. This

method is known as the Doppler Shift Attentuation Method.

Recalling that the observed gamma-ray energy for a transition of energy Eo,

emitted from a recoil of velocity v at an angle θ to the detector is given by Es

where

Es ≈ Eo

(

1 +v

ccosθ

)

(5.3.26)

The gamma-ray energy spectrum will now have a lineshape depending on what

velocity (between vo and zero) the recoil had when it emitted the gamma-ray.

For example, on average, decays from faster (∼ 10−15s) levels will have hardly

slowed down in the target/backing material when the gamma-ray is emitted,

while decays from states with slower lifetimes (∼ 10−12s) maybe almost totally

stopped. Measuring the centroid of the total lineshape gives a measure of the

average velocity at which the gamma-ray was emitted. The centroid results are

usually expressed in terms of a Doppler Shift Attenuation Factor F (τ), where

F (τ) =vav

vo=

1

v0τ

∫ ∞

0v(t)exp

(

− t

τ

)

dt (5.3.27)

136

Page 140: nuc_expt_phr03

where v0 is the initial recoil velocity and vav is the average recoil velocity when

the gamma-ray is emitted.

The slowing down of the recoil with as a function of time (v(t)) is usually

taken from tables of stopping powers such as those by Braune [204], Northcliffe

and Schilling or Ziegler [205].

The stopping power is usually separated into two effects, electronic and nuclear

stopping. In electronic stopping, the recoil slows down due to interactions with

atomic electrons in the stopping material. Since the mass of the recoil is much

larger than the mass of the electrons, many collisions are required to remove all

the energy of the recoil.

In the nuclear stopping process, the recoils lose energy in a small number of

discrete steps due to nuclear collisions. These can cause the nucleus to scatter

and alter its direction, which must be accounted for (since a change in direction

will alter the measured, Doppler shifted energy). This is usually accounted for

using either the Blaugrund formulism [206] or using Monte Carlo simulations of

the recoil velocity profiles [207, 208, 209].

137

Page 141: nuc_expt_phr03

700 900 1100 1300 1500 1700 1900 2100

Energy (keV)

-200

200

600

1000

1400Cou

nts

-200

200

600

1000

1400

Ce band 1132

Ce band 1132

31.7 + 37.4o o

o o142.6 + 148.3

Figure 5.16: DSAM spectra for forward and backward angles for the yrast su-perderformed band in 132Ce taken from GAMMASPHERE [56].

138

Page 142: nuc_expt_phr03

0.7 1.2 1.7Energy (MeV)

0.4

0.6

0.8

1.0

F(τ

)

132Ce 1

131Ce 1

0.7 1.2 1.7Energy (MeV)

132Ce 2

132Ce 3

0.7 1.2 1.7 Energy (MeV)

131Ce 2

a) b) c)

Figure 5.17: Measured and calculated fractional Doppler shifts for hte highlydeformed bands in 131,132Ce [56].

5.3.1 Lineshape Analysis.

Note that the lifetime of the state can also be gained directly from the gamma-ray

lineshape (dNdE

) since this gives directly the velocity distribution of the recoils as

they emit the gamma-ray.

dN(t) =1

τexp

(

− t

τ

)

dt (5.3.28)

A knowledge of the slowing down process or stopping power of the recoil, dvdt

,

thus allows the lineshape to be calculated for a given value of τ . (Note that since

Es and v have a linear relationship, it is simple to translate from the measureddNdE

to the calculated dNdv

, the dNdt

= dNdv.dvdt

).

DSAM with Thin Targets for Very Fast Decays.

In the case of very fast transitions (∼10−15s) such as those observed in the su-

perdeformed bands in the A∼80 region the effect of the recoil slowing down in

139

Page 143: nuc_expt_phr03

395.0 405.0 415.00.0

500.0

1000.00.0

200.0

400.0

coun

ts

0.0

500.0

1000.0

435.0 445.0 455.0Energy [keV]

0.0

200.0

400.0

600.0

0.0

100.0

200.0

300.00.0

200.0

400.0

600.0

455.0 465.0 475.00.0

200.0

400.0

600.0

0.0

100.0

200.0

300.00.0

200.0

400.0

600.0

197Pb (1)

403 keV 446 keV 467 keV

35o+50

o

90o

130o+145

o

Figure 5.18: DSAM lineshape fits to transitions in the M1 band of 197Pb [216].

the thin target can be used to obtain a Doppler shift between the forward and

backward angles of a large array, from which a fractional Doppler shift can be

obtained [213, 214, 167]. As in the usual DSAM, the difference in gamma-ray

energy measured between detectors at forward and backward angles can be used

to infer the velocity at which the recoil was moving when the gamma-ray was

emitted.

Figure 5.19 shows the angle gated spectra for the SD band in 87Nb from

GAMMASPHERE [167]. For the highest spin transitions there is a clear differ-

ence in centroid for the forward and backward angle detectors. From this shift,

an F (τ) value can be deduced from which an estimate of the lifetime (and thus

quadrupole moment) can be obtained.

140

Page 144: nuc_expt_phr03

(a)

(b)

Figure 5.19: (a) DSAM spectra for thin target data on the SD band in 87Nb. (b)F(τ) curves for the superdeformed and normal deformation bands in 87Nb [167].

141

Page 145: nuc_expt_phr03

Chapter 6

Measurement of MagneticMoments.

The magnetic dipole moment, µ, is defined classically as the vector cross-product

of a current I and area about which the charge circulates, A. In the nuclear

case, the measurement of a magnetic dipole moment gives extremely sensitive

information on the nature of particles causing the current (ie. whether protons or

neutrons) and the single particle orbital(s) which the active nucleon(s) occupy. As

such they are extremely sensitive probes of nuclear wavefunctions. One can draw

the semi-classical analogy of higher angular momentum orbitals being further on

average from the centre of the nucleus and thus sweeping out larger areas, giving

rise to larger magnetic dipole moments.

The g-factor relates the magnetic dipole moment, µ, and the spin of the state

by the expression [1, 235],

g =µ

I(6.0.1)

In the case of odd-A nuclei, the Schmidt model [235] gives an estimate of the

values of g-factors for pure, single nucleon states moving in shell model orbits,

independent of the nuclear core.

For a single, independent nucleon, the single particle magnetic dipole moment,

µ can be calculated using the expression [233],

µ =gl[l(l + 1) + j(j + 1) − s(s+ 1)] + gs[s(s+ 1) + j(j + 1) − l(l + 1)]

2(j + 1)(6.0.2)

Substituing j = l ± s and s = 12, the Schmidt values for g-factors for pure

142

Page 146: nuc_expt_phr03

protons and neutron orbitals can be calculated. For spherical shell model orbitals

where j = l + 12

(eg. p 3

2

, d 5

2

...)

g =1

j

1

2gs +

(

j − 1

2

)

gl

(6.0.3)

and for orbitals with j = l − 12

(p 1

2

, d 3

2

, g 7

2

...),

g =1

j + 1

−1

2gs +

(

j +3

2

)

gl

(6.0.4)

Where gl=1 and gs=+5.587 for protons and gl=0, gs=-3.826 for neutrons

repsectively. The emprical data suggested that the gs values need to be attenuated

by a factor of approximately 0.7 to account for fact that the odd nucleon is not

‘free’ but included in the nuclear medium.

As table 6.1 shows the emprically g-factors for single particle shell model

states.

Particle Orbital g-factorNeutrons h 11

2

–0.21

g 9

2

–0.24

g 7

2

+0.21

d 5

2

–0.33

d 3

2

+0.44

s 1

2

–1.76

Protons h 11

2

+1.17

g 9

2

+1.27

g 7

2

+0.72

f 5

2

+0.54

d 5

2

+1.38

d 3

2

+1.33

p 1

2

–0.23

s 1

2

+2.90

Table 6.1: g-factors for intrinsic spherical orbitals [234].

In reality, most nuclear states are not pure single particle states but super-

positions of a number of different configurations which are mixed together. A

measurement of the magnetic dipole moment can be used to infer the underlying

single particle structure and purity of such a state.

143

Page 147: nuc_expt_phr03

6.1 Measurement of Nuclear Magnetic Dipole

Moments.

If a magnetic field is applied to a nucleus with magnetic dipole moment, µ, a

torque, T , will occur which will cause the nucleus to twist around or ‘precess’.

This torque can be calculated using the vector cross product of the magnetic

dipole moment and the applied magnetic field. The rotational frequency of this

precession, is given by the Larmor frequency, ωL where [1]

ωL =µB

hI= −gµNB

h(6.1.5)

110Cd

angular distribution

335 keV ext field up

0 10 20 30 40 50 60 70 80 90

0.8

1.0

1.2

= -107 24 mrad + -ωτ

θ

335 keV ext field down

0.8

1.0

1.2

a = +0.31 2

a = -0.08 4

Figure 6.1: Angular distributions used to measure the g-factor of the yrast 10+

state in 110Cd (E2 decay) [236].

144

Page 148: nuc_expt_phr03

110Cdangular distribution

399 keV ext field up

0 10 20 30 40 50 60 70 80 90

0.8

1.0

1.2

θ

= -302 37 mrad + -ωτ

399 keV ext field down

0.8

1.0

1.2

a = -0.23 2

a = 0.00 4

Figure 6.2: Angular distributions used to measure the g-factor of the yrast 7−

state in 110Cd (E1 decay) [236].

For a state with lifetime τ , the effect of this torque will be to shift the measured

gamma-ray angular distribution by an angle ∆θ (essentially because the nucleus

twists around relative to the reaction plane, before it decays) where

∆θ = ωLτ = −gτ BµN

h(6.1.6)

Thus, if the lifetime of a nuclear state is known, by measuring the shift in

the angular distribution in the presence of a known magnetic field, compared to

the usual in-beam distribution, the g-factor can be deduced. Figures 6.1 and 6.2

show the shifts in the observed angular distributions for both field up and down

directions for two different configurations in 110Cd.

The size of the applied magnetic field, B required to cause the nucleus to

145

Page 149: nuc_expt_phr03

precess sufficiently so that a rotation of angle ∆θ can be measured, for a state

decaying with lifetime τ , can thus be estimated using the expression [1],

Bτ =∆θh

gµN(6.1.7)

Thus, the shorter the nuclear lifetime, the greater the required magnetic field

to observe a shift in the angular distribution function. A rule of thumb from

equation 6.1.7, is that for a shift of 100 mrads, a Bτ value of approximately

2×10−9 Ts is needed. Thus for liftimes of the order of a nanosecond, external

fields of around 1 Tesla are required, which is at the limit of what can be provided

using an external magnet. For excited states with lifetimes of ∼ 10−12 s, required

larger fields than can be obtained using such magnets, and the very large internal

(or transient) fields of the ion moving through a ferromagnetic crystal have to be

used.

In the case where the applied magnetic field is perpendicular to the reaction

plane, the angular distribution, for a nucleus precessing with the Larmor fre-

quency for time t, will have a, perturbed angular distribution, given by [1, 235]

W (θ, t, B) =∑

k

AkPk (cosθ − ωLt) (6.1.8)

If the lifetime of the state of interest τ is small compared to the electronic

time resolution of the system, the intergrated precession angle, (∆θ = ωLτ) over

the lifetime of the state is measured and the angular distribution will have the

form, [235]

W (θ, t, B) =∑

k

∫ ∞

0

1

τexp

(

− t

τ

)

AkPk (cosθ − ωLt) (6.1.9)

6.1.1 Corrections in the Ion-Implantation Perturbed An-gular Distribution Technique.

The ion-implantation, integral perturbed-angular-distribution (IMPAD) technique

[235, 236, 237], assumes that following a heavy-ion reaction, for an ion coming to

rest in a magnetic host, the total precession of the angular distribution, ∆θ, is

given by,

∆θ = ωLτ + ∆θtr + ∆θfeed, (6.1.10)

146

Page 150: nuc_expt_phr03

external static fieldof ~ 0.1 ->1 Tesla to provide B x T.

productiontarget

S

Nthin ferro-mag. layer(gives rise to large transient B)

external mag fieldto provide directionalpolarisation of transientfield

N

S

stopper

beam beam

(a) Static field, for t > 1ns (b) Transient field, for t~10 ps

Figure 6.3: Schematic methods of the (a) static field and (b) transient fieldmethods of measuring g-factors of exited nuclear states.

where ∆θtr is the precession due to the transient field, which acts on the nucleus

as it slows down in the ferromagnetic medium and ∆θfeed is the average static-

field precession accumulated in states which feed the state of interest. The static-

field precession of the level of interest is given by ωLτ , where

ωL = −gµN

hBst (6.1.11)

and Bst is the static hyperfine magnetic field, while g and τ are the the mean

life of the level and its gyromagnetic ratio, respectively.

The magnitude of the transient field contribution may be estimated using the

expression [236],

∆θtr = −〈g〉µN

h

∫ Ts

0Btr(t)dt (6.1.12)

where Btr(t) denotes the transient magnetic field, Ts is the time the ion takes to

come to rest in the ferromagnetic layer, and 〈g〉 denotes the average g-factor of the

(continuum) states that are populated while the nuclei experience the transient

field.

The precession induced in the state of interest due to precessions induced

in higher-lying feeding transitions can be be deduced by measuring the shift in

147

Page 151: nuc_expt_phr03

the angular distributions of the direct feeding transitions directly in the same

experiment.

6.1.2 Analysis of Precession Data with Limited Angles.

It is often not necessary to measure the entire shift in the angular distribution to

determine the g-factor. In the case where there are a limited number of gamma-

ray detectors, one can measure the ‘up-down’ intensity ratio by placing detectors

at angles symmetric to the beam direction and measuring the ratio of counts

in the precession shifted decay with the field in both up and down directions

[236, 237, 238, 239, 240]. To determine the field up-down counting asymmetry

and reduce possible systematic errors, the double ratio, ρ, is defined for a ±θdetector pair by

ρ =

N+θ(↑) ×N−θ(↓)N+θ(↓) ×N−θ(↑)

(6.1.13)

where N+θ(↑) denotes the number of counts in the θ = +θ detector for field up

direction, etc. For small precessions, ∆θ ≤ 100 mrad, the precession angle may

be obtained from

∆θ =ǫ

S(6.1.14)

where

ǫ =1 − ρ

1 + ρ(6.1.15)

and S, the logarithmic derivative of the angular distribution, is given by

S =1

W (θ)

dW

dθ(6.1.16)

The angular distribution has the usual form:

W (θ) = A0 1 + a2P2(cosθ) + a4P4(cosθ) (6.1.17)

One explicit form for evaluation of Eq. 6.1.16 is [236]

S =−8 sin(2θ)[12a2 + 5a4 + 35a4 cos(2θ)]

64 + a2[16 + 48 cos(2θ)] + a4[9 + 20 cos(2θ) + 35 cos(4θ)](6.1.18)

148

Page 152: nuc_expt_phr03

Therefore, if the angular distribution coefficients of the decaying transition

are known, one can calculate the precession angle (and thus the g-factor is the

lifetime of the state in known) directly from equation 6.1.14.

6.1.3 Transient Field Measurements.

For nuclear lifetimes of excited nuclear states of the order or 10−11→12 s, the

product of the lifetime and static field strength is not usually enough to cause

a measureable precession. In order to obtain a measureable shift in the angular

correlation/distribution, the nucleus has to be caused to experience a much larger

‘transient’ field caused by its movement through a ferromagentic medium. The

basis of the transient field technique [235, 241, 242] is shown schematically in

figure 6.3b. The nucleus is formed in a thin production target and then passes

through a thin ferromagnetic layer (usually iron or gadolinium), where it experi-

ences a strong transient field of the order of 1000 T. A non-magnetic stopper is

then placed behind the ferromagnetic layer to stop the recoil, where it emits the

gamma-rays from which a shift in the angular distribution/correlation function

can be deduced.

One problem with this technique is that it is rather hard to correct for the

shifts in the angular correlations caused by interactions with states above the state

of interest. One method proposed to remove this problem is to use a ‘plunger’

set-up, as used to measure lifetimes in the recoil distance method [241].

The size of the transient field experienced by the nucleus as it passes through

the polarised ferro-magnetic layer can be hard to measure and where possible, a

calibration of the field strength is performed using the measurement of a state

with a known g-factor. The size of the transient field depends on the ratio of the

recoil velocity as it passes through the ferromagnetic layer and velocity of the 1s

orbital electons in the medium. An estimate of the transient field as a function

of the recoil of proton number Z and recoil velocity, v, can be given by [242, 243],

BTR(v) = αZ(

v

v0

)

exp(

−β vv0

)

(6.1.19)

where v0 = c137

is the Bohr velocity and α and β are constants obtained fits

to experimentally determined transient fields [244].

149

Page 153: nuc_expt_phr03

Magnetic Moment Measurements Using A Plunger and the TransientField Technique.

A problem with deducing magnetic moments from the transient field technique

is correcting for the effect of the precessions of states which feed the state of

interest. This problem can be solved using a plunger [241]. The basic idea is

that the recoils come out of the thin production target with velocity v and are

stopped in a backing target placed a distance d behind the production target.

Those gamma-rays which are emitted while the nucleus is in flight between the

production target and the stopper, are, as observed in a fixed angle detector,

shifted from their true energies by the Doppler effect.

Gamma-gamma coincidence data is taken and a gate is set on the shifted

component of transition above the state of interest. This ensures that the nucleus

was moving when this state decayed and thus had not yet entered the stopper

(and thus had not yet experienced the transient field). The angular correlation

of the stopped component of the transition of interest is then measured and a

precession frequency extracted, from which a g-factor can be deduced.

6.1.4 Time Differential Perturbed Angular Distributions.

For nuclear state lifetimes of between 1 ns and 1 µs, it is possible to measure

the Larmor frequency directly using an in-beam method. The basis of the Time

Differential Perturbed Angular Distribution (also known as the ‘spin rotation’)

technique [235, 234, 247, 246, 248], is to use a pulsed beam to irradiate the target

and form the isomeric state, which is implanted into a ferromagnetic medium,

where it is subjected to an applied static magnetic field, Bst whose value is known.

The decay of the transition(s) out of the state are measured as a function of time

delay after the target was irradiated. The time spectra for such experiments show

the typical exponential decay associated with a nuclear lifetime, but a smooth

oscilation, with the periodicity of the Larmor frequency, is superimposed upon

this decay spectrum. This effect is caused by the nucleus precessing around in

the presence of the magnetic field, which causes the angular distribution (and

thus measured intensity at a fixed detector angle) to oscillate with time. The ob-

served intensity of an observed γ-ray, measured at angle θ to the beam direction,

decaying from a state of lifetime, τ at time t is then given by [235]

150

Page 154: nuc_expt_phr03

I(θ, t, B) = Ioexp(

− t

τ

)

W (θ, t, B) (6.1.20)

where W (θ, t, B) is the angular distribution function for the state given in

equation 6.1.8.

A good example of the use of the TDPAD technique to measure the g-factors

of high-spin isomeric states is shown in figure 6.4. Note that the Larmor frequency

can be shown up to a larger extent by measuring the gamma-ray intensity as a

function of time using two detectors at symmetric angles ±θ to the beam direction

and plotting the intensity ratio, R, defined by R = I(+θ)−I(−θ)I(+θ)+I(−θ)

151

Page 155: nuc_expt_phr03

214Fr

R(t)

time (ns)

472 keV ratio

0 100 200 300 400 500 600

-0.4

-0.2

0.0

0.2

0.4

214Fr

counts

472 keV (135 deg)

310

410

510

472 keV (-135 deg)

310

410

510

Figure 6.4: Time differential perturbed angular distribution spectra showing theoscillation in the time spectra associated with the Larmor frequency [246].

152

Page 156: nuc_expt_phr03

Chapter 7

Spectroscopy of Neutron RichNuclei.

The structure of very neutron deficient nuclei extends to nuclei at and in some

cases even beyond the proton drip line. Using stable beam/target combinations,

it is possible to use fusion-evaporation reactions to form very neutron deficient

compound systems. Unfortunately, this very fact, means that the study of very

neutron-rich systems at high spins using this method is not possible. However,

there are a number of in-beam techniques for the study of nuclei with a neutron

excess which will be discussed in this chapter.

7.1 Using Fusion Evaporation Reactions

It is possible to study a few specific cases of β-stable and a few nucleons to the

neutron rich side of the valley of stability, using fusion-evaporation reactions.

The basic idea is to form as neutron-rich a compound system as possible and use

a charged particle detector to detect any evaporated charged particles such as

protons and α-particles [249, 22, 25]. Such events are inhibited by the Coulomb

barrier, however, the clean selection allowed by modern day charged particle de-

tectors means that very clean charged particle gated, γ-ray identification spectra

can be obtained.

Figures 7.1 highlights the effectiveness of this technique for identifying the

yrast states of the neutron rich nucleus 63Co [25] (note that the β-stable neutron

rich isotope is 59Co). The gamma-ray transitions associated with cobalt isotopes

are identified by having the gating condition that an evaporated proton must be

153

Page 157: nuc_expt_phr03

(a)

(b)

Figure 7.1: (a) Singles identification spectra and (b) proton gated γ −γ coincidence data on the neutron rich nucleus 63

27Co33 populated via the18O(48Ca,p2n)63Co reaction [25].

measured. The isotope is then identified by means of either, the average proton

energy, an excitation function or a measurement of neutron multiplicity.

The same method can also be used to identify high-spin states in the most

neutron rich stable nuclei [22, 251]. These are often nuclei where many of the non-

yrast structures are known from neutron capture and pick-up/transfer reaction

studies. The addition of knowledge of high-spin states allows a full spectroscopic

study of such nuclei.

7.2 Incomplete Fusion/Massive Transfer Reac-

tions.

Another way of studying the near yrast states of slightly-neutron to medium

spins is to use a light projectile which partially fuses with a heavy target nucleus,

154

Page 158: nuc_expt_phr03

close to the nucleus of interest [80]. This method, known as massive transfer

or incomplete fusion has been used to to study quasi-particle alilgnments in the

A∼100 region [250] and more recently to study high-K structures in stable A∼180

nuclei [81].

0

5

10

15

0

4

8

0 200 400 600 8000

1

2

Eγ(keV)

R55

(F/M

B)

R60

(F/M

B)

RF(

60/5

5)(a)

(b)

(c)

179Hf

178Hf

177Hf

178Hf

177Hf

177Hf

178Hf

Figure 2.

Figure 7.2: Angular distributions of α-particles observed in the reaction176Yb+9Be. Note that the incomplete fusion reactions to populate higher spinsin 178Hf have a more forward focussed distribution compared to the pure fusionevaporation products [81].

While charged particles emitted in fusion-evaporation reactions have an isotropic

distribution in the centre of mass frame, the break up products from incomplete

fusion are more forward focussed. In general, the higher the spin state popu-

lated in an incomplete fusion reaction, the more peripheral the collision and the

more foward focussed the outgoing particle distribution. Figure 7.2 shows the

α-particle anisotropies for various channels observed in the 176Yb+94Be5 reaction

[81]. Note that the higher the gamma-ray energy corresponding in general to the

decay of higher spin states in the rotational bands of 178Hf), the more forward

focussed the α-particle anisotropy. Note that the lower spin states are probably a

combination of fusion-evaporation and massive transfer, so the anisotropy effect

155

Page 159: nuc_expt_phr03

is more pronounced for higher spins. Figure 7.2 shows α-particle gated γ-ray

spectra obtained for the reaction 176Yb+94Be5, which is proposed to the band

built on top of the τ=31 year, Kπ = 16+ isomer in 178Hf, identified as high-spin

states by the α-particle anisotropy.

7.3 Deep Inelastic Reactions.

For neutron-rich nuclei, it is difficult to observe the yrast sequence to high spins

(such as through the backbend) due to the preferential population of neutron-

deficient species in fusion-evaporation reactions and the low angular momentum

involved in the fission process. The use of deep-inelastic reactions to populate

near yrast states in slightly neutron-rich nuclei is now well established [252, 254,

255, 256, 257, 258, 259, 96, 97, 98, 264, 265, 266, 267, 268, 269, 270, 271, 272] and

provides an efficient way of studying the yrast states of stable and slightly neutron

rich nuclei. Figure 7.3 shows schematically the use of deep inelastic reactions to

populate neutron rich nuclei.

136

Xe

Neutronsevaporated

232

Th

Newbeam like

nuclei

Newtargetlike

nuclei

Target

Beam

N:Z equilibration

Nucleons exchangedEb ~ 10-15%above Ec

Figure 7.3: Schematic of the use of deep inelastic collisions to populate yraststates of neutron rich nuclei.

It has been shown experimentally that the binary system equilibrates into

systems with approximately equal N:Z ratios [260, 261, 262, 263]. Thus, the

extra neutron excess of heavy, stable targets, means that bombarding these with

lighter beams usually results in an overall flow of neutrons onto the lighter, beam

like fragments. Since these are neutron rich nuclei, charged particle evaporation

is strongly hindered by the Coulomb barrier and thus the total Z of the compound

system (=Zbeam+Ztarget) is usually conserved in the break up. Typically, between

2 and 6 neutrons will be evaporated from the two hot binary fragments and thus

a specific nucleus will be accompanied by a number of binary partner nuclei,

comprising of between 2 and 4 isotopes of the same element. If transitions in

156

Page 160: nuc_expt_phr03

these binary partners are known, the relative intensity of the various binary

partner products can be used to identify the isotope of other, possibly unknown

neutron rich fragment.

A major problem with using deep inelastic reactions to study beam like prod-

ucts is their large recoil velocity. One approach to solve this problem is to use a

thick or backed target which stops the beam-like product within a few ps. Thus,

transitions from decays with apparent lifetimes greater than this stopping time

have no Doppler shift and can be clearly resolved. Figure 7.4 shows the region

of the Segre chart populated with the binary reaction 86Pd+110Pd [259].

66

64

Cd

Pd

Ru

Zr

Ag

Rh

Tc

Y

Kr

Br

Rb

Sr

Nb

Mo

50 52 54 56 58 60 6248

Se

= STABLE NUCLEUS

100

104Ru

108Pd

Mo

86Kr

110Pd

= BEAM / TARGET NUCLEI

Figure 7.4: Region of the Segre chart poplated by binary collisions using thereaction 110Pd+86Kr [259].

Figure 7.5: Spectra from the 86Kr+110Pd binary reaction. Comparison of align-ments between odd and even-N Mo, Pd and Ru nuclei populated in this reaction.

Figure 7.5 shows the spectra for 104Ru studied using the binary reaction86Pd+110Pd [259]. The 104Ru lines of interest can be clearly selected, as can

157

Page 161: nuc_expt_phr03

the presence of the Sr binary partner lines. The data in figure 7.5 extends the

yrast band of 104Ru upto spin 14+. As figure 7.5 shows, thus is high enough to

obtain useful insights into the alignment processes in these nuclei.

In heavier deformed nuclei, relatively high spins can be obtained using binary

reactions with thick/backed targets and a high efficiency gamma-ray arrays [98,

271, 272]. Figure 7.6 shows the high spin cascade in 234Th and 230Th observed

using GAMMASPHERE obtained from the binary reaction 136Xe+232Th [97].

Note that the higher multiplicity gamma-ray gating afforded by using the higher

efficiency arrays gives rise to extremely clean spectra, with only the nucleus of

interest (and possibly a few binary partners) observed,

0

1

2

3

x 10

4 C

ount

s

50 150 250 350 450 550 650Energy (keV)

0

1

2

3

x 10

4 C

ount

s

182.

5T

h

Th

Th

237.

4

285.

7

327.

9

365.

339

7.9

426.

845

2.3

Xe

483

.4

Th

493.

747

6.3

427.

639

6.8

361.

8322.

2

277.

8

228.

2

173.

4

Th

X-r

ays

X-r

ays

120.

8T

h

(a)

(b)

454.

1

Xe

589

.0

Figure 7.6: Gamma-ray spectra produced for the binary reaction 136Xe+232That beam energy of 833 MeV. (a) double gated spectrum showing ground state-band transitions in 234Th. (b) spectrum showing ground state-band transitionsin 230Th. Transitions labelled ‘Xe’ are cross-coincident gamma-rays attributedto 138Xe, the projectile-like partner of 230Th. Gamma-ray peaks marked ‘Th’ inboth spectra are contaminant transitions in the yrast band in Coulomb-excited232Th [96, 97].

158

Page 162: nuc_expt_phr03

7.3.1 Maximum Angular Momentum in DIC.

Spin-wise, the use of deep inelastic reactions are not particularly efficient in gen-

erating very high spin states. (Recent studies of (discrete-line) spin-input us-

ing thin target deep-inelastic reactions can be found in reference [268]). Unlike

fusion-evaporation reactions, not all of the input angular momentum of the reac-

tion goes into the intrinsic angular momentum of the final products. Most of the

angular momentum goes into the relative spin of the two fragments with respect

to each other. It can be shown by a semi-classical argument that in the pres-

ence of strong friction between two rolling spheres (ie. the target and beam like

nuclei), known as the rolling limit, 57

of the maximum initial maximum angular

momentum (for a grazing collision), Lmax goes into the relative motion of the two

fragments while only 27

goes into the intrinsic spins of the beam and target like

fragments respectively [253].

The maximum input spin into the beam and target like fragments of masses

AB and AT respectively can be estimated using the following semi-classical ex-

pressions [253, 254, 280]. Figure 7.7 shows how increasing the input beam energy,

increases the angular momentum input of the target and beam nuclei. Note also,

that these fold distributions are ‘two-peaks’ corresponding to low-fold reactions

such as Coulex and higher fold reactions such as deep-inelastic collisions. This

is highlighted further in figure 7.8 which shows the fold distribution for states

gated on the yrast band in 100Mo for increasing spin. The lower spin states are

populated mostly by direct Coulomb excitation, while the higher spin states are

populated via transfer followed by neutron emission, and have a considerably

higher entry spin.

LBLF =2

7

1

1 +(

AT

AB

) 1

3

Lmax (7.3.1)

LTLF =2

7

1

1 +(

AB

AT

) 1

3

Lmax (7.3.2)

159

Page 163: nuc_expt_phr03

7.3.2 Useful Formulae for Binary Reaction Studies.

The following section is a summary of useful formula for studies of transfer/binary

style reactions and was compiled by Prof. P.A. Butler of Liverpool University.

Elastic scattering

Projectile mass M1 on target M2 with lab. energy E0.

M1 scatters at lab angle ψ, c.m. angle θ,

M2 scatters at lab angle ζ , c.m. angle φ = π − θ = 2ζ

Lab. energy of scattered nucleus M1 is

E1

E0

= 1 − 2M1M2

(M1 +M2)2(1 − cosθ) = (

M1

M1 +M2

)2cosψ ± [(M2

M1

)2 − sin2ψ]1

22

Use only plus sign unless M1 > M2, in which case ψmax = sin−1(M2

M1

)

Lab. energy of recoil nucleus M2 is

E2

E0

= 1 − E1

E0

=4M1M2

(M1 +M2)2cos2ζ

where ζ ≤ π2

Angles

sinζ = (M1E1

M2E2

)1

2 sinψ

tanψ =sin2ζ

M1

M2

− cos2ζ=

sinθM1

M2

+ cosθ

θ = ψ + sin−1(M1

M2)sinψ = π − 2ζ

cosθ = 1 − 2cos2ζ

ψ = π/2 − ζ if M1 = M2

160

Page 164: nuc_expt_phr03

Rutherford Scattering Cross Sections (in mb/sr).

dΩcm= 1.3(

Z1Z2

Elab

M1 +M2

M2)2 1

sin4 θcm

2

(Elab is in MeV)

dΩscat,lab=

dΩcm

sin2θ

sin2ψcos(θ − ψ)

dΩrecoil,lab=

dΩcm4sin

θ

2

Inelastic scattering.

Same notation as for elastic scattering except that lab energies of scattered beam

and recoil are respectively E3 and E4. Define the following terms:

Q = (M1 +M2 −M3 −M4)c2

ET = E0 +Q = E3 + E4

A =M1M4

(M1 +M2)(M3 +M4)

E0

ET

B =M1M3

(M1 +M2)(M3 +M4)

E0

ET

C =M2M3

(M1 +M2)(M3 +M4)(1 +

M1

M2

Q

ET) =

E4′ET

D =M2M4

(M1 +M2)(M3 +M4)(1 +

M1

M2

Q

ET) =

E3′ET

The total scattered energy in the c.m. is

E′ = E3′ + E4′ =M2

M1 +M2(ET +

M1

M2Q) = Ecm +Q

161

Page 165: nuc_expt_phr03

Lab energy of light product

E3

ET

= B +D + 2(AC)1

2 cosθ

Lab energy of heavy product

E4

ET= A + C + 2(AC)

1

2 cosφ

Angles:

sinζ = (M3E3

M4E4)

1

2 sinψ

sinθ = (E3/ET

D)

1

2 sinψ

Deep Inelastic or Fission

with initial kinetic energy, put M2 = 0 and

Q ≈ 0.107Z2

A1

3

+ 22.2

(A is in u and Q in MeV) so that

E4 =M4E3

M3 +M4[cosζ ±

M3Q

M4E3− sin2ζ ]2

Use only + sign unless M3QM4E3

< 1, when

ζmax = sin−1(M3Q

M4E3

)1

2

Compound Nucleus

Projectile M1 with lab energy E0, target M2 as above.

Lab energy of compound nucleus

ECN =M1

M1 +M2

E0

Excitation energy of compound nucleus

Ex = (E0 − ECN) +Q =M2

M1 +M2E0 +Q = Ecm +Q

162

Page 166: nuc_expt_phr03

Coulomb Barrier

Interaction radius is

R = 1.16(A1

3

1 + A1

3

2 + 2)

where A is in u and R is in fm, then

Vcm = 1.44Z1Z2

R

Vlab = (1 +M1

M2

)Vcm , Elab = (1 +M1

M2

)Ecm

Maximum Angular Momentum.

ℓmax = 0.219R[µ(Ecm − Vcm)]1

2

where

µ =M1M2

M1 +M2

Grazing Angle.

The grazing angle [274], where the cross-section for peripheral, deep-inelastic

reactions will be maximised, is defined as the angle at which the distance of

closest approach, d, given by [5, 273],

d = [Z1Z2e

2

4πǫ0Ek].[1 + cosec(

θ

2)] (7.3.3)

(where Z1 and Z2 are the atomic numbers of the two nuclei involved and E

is their kinetic energy) is equal to the sum of the nuclear radii (ie. when the two

nuclei are just touching), given by,

d = 1.25(A1

3

1 + A1

3

2 )fm (7.3.4)

where A1 and A2 are the nuclear mass numbers.

A quick estimate of the grazing angle can be obtained using,

1

2(1 + cosec

θc.m.

2) =

E

V

163

Page 167: nuc_expt_phr03

Reaction Cross Section (in barns).

σR ≈ 0.67

µEcmℓ2max

Velocity in % c.

v = 4.634

E

A

(E in MeV, A in u)

For transfer reactions the two nucleon transfers are strongly dependent on the

Q-value,

Q = Qgg −Q(∆Z)

where,

Q(∆Z) =[(

Z3Z4

Z1Z2

)

− 1]

Vcm

Z1 and Z2 correspond to the projectile and target before the reaction. Z3 and

Z4 correspond to the projectile-like and target-like fragments after the reaction.

The more positive the Q-value the higher the population. Note that Q(∆Z)=0

for only neutron transfers. Qgg is the mass difference between initial and final

products.

7.3.3 Doppler Correction.

By stopping the beam like products in the target, transitions with lifetimes less

than a few ps are Doppler shifted and thus can not be resolved. In order to observe

the highest spins of neutron rich products populated in deep inelastic collisions,

a thin target is required, which allows the beam-like products to decay in flight.

In this case, a method is required to obtain the recoil velocity and emission angle

of the beam like fragment so an appropriate Doppler correction can be applied

to reduce the overall Doppler broadening of the gamma-ray transitions [?].

The HIPS detector, shows in figure 7.9 gives a position sensitive signal which

gives the recoil direction and uses also uses the time of flight to give a degree

of mass resolution for beam-like products in binary reactions. Figure 7.10 shows

the improvement in the gamma-ray energy resolution of beam-like products in

164

Page 168: nuc_expt_phr03

the binary reaction using a 86Kr beam on a 116Sn target by detecting the beam

like fragment in the HIPS detector and correcting for the Doppler effect.

Similarly, a recoil filter detector [47, 163] can be used to obtain similar infor-

mation and Doppler corrections.

The CHICO detector [276] is a large solid angle set of PPACS which uses

time of flight to get some mass infomation on the fragments and allows excellent

Doppler correct when used with the GAMMASPHERE array. In addition to in-

beam experiments, it has been used to extend the decays schemes of a number

of neutron-rich fission products around A ∼100 upto discrete spins of 20h using

a thin, fission source [277, 278, 279]

Determination of Angular Corrections in DIC.

If the recoiling beam (or target like fragment) is detected at angles θ1 and φ1 to

the beam axis, and the (doppler shifted) gamma-ray emitted from the recoiling

fragment is detected in a germanium detector and angles θ2, φ2 to the same axis,

in order for the Doppler correction to be applied, the angle between the emitted

gamma-ray and charged particle fragment directions (θdop) must be known. This

can be calculated using simple geometry assuming two unit vectors, v1 and v2,

where the (x, y, z) components of these vectors are given by x1 = sin(θ1)cos(φ1),

y1 = sin(θ1)sin(φ1), z1 = cos(θ1) and x2 = sin(θ1)cos(φ2), y2 = sin(θ2)sin(φ2),

z1 = cos(θ2) respectively. The angle between these two vectors can be calculated

using the vector dot product, ie. v1.v2 = v1v2cos(θdop). Therefore

cos(θdop) = sin(θ1)cos(φ1)sin(θ2)cos(φ2)+sin(θ1)sin(φ1)sin(θ2)sin(φ2)+cos(θ1)cos(θ2)

(7.3.5)

collecting the terms and recalling the trigonometric identity cos(A − B) =

cosAcosB + sinAsinB, the above expresssion reduces to

cos(θdop) = sin(θ1)sin(θ2)[cos(φ1 − φ2)] + cos(θ1)cos(θ2) (7.3.6)

165

Page 169: nuc_expt_phr03

Figure 7.7: Fold distributions gated on the low-lying 601 keV transition in 100Mo,following the binary reaction 100Mo +136Xe at beam energies of 650, 700 and750 MeV respectively. Note that the overall spin of the fragments increasesdramatically with beam energy [281]. Data taken from the 8π array at Berkeley.

166

Page 170: nuc_expt_phr03

Figure 7.8: Fold distributions gated on the low-lying yrast transitions in 100Mo,following the binary reaction 100Mo +136Xe at a beam energy of 700 MeV. Notethat low-lying states are predominantly populated vua Coulomb excitation, whilethe higher ones have a higher overall multiplcity distribution, associated with adeep-inelastic collision, probably followed by neutron emission [281]. Data takenfrom the 8π array at Berkeley.

167

Page 171: nuc_expt_phr03

ygri

d

cato

de

dela

y-ch

ip

xgri

d

10MΩ

10MΩ

10MΩ

10MΩ

10MΩ

10MΩ

xgri

d

ygri

d

cato

de

dela

y-ch

ip

housing

Stop

Detector

Ionisation

Chamber

Foils &Support Grids

Field Rings

FrischGrid

Anodes

10

Gate-valve

Anode HT

Signal

Signal

Signal

Signal

Signal

Grid HT

0 10 20 30 40 50cm

Start-Detector

S.Schwebel et.al.Manchester University

Figure 7.9: Scale drawing of the HIPS vessel for detecting recoils from deep-inelastic reactions [275].

168

Page 172: nuc_expt_phr03

0 500 1000 1500 2000 2500 3000 3500Channel Number

0

50

100Doppler corrected spectrum

0

50

100

150

HIPS-Ge coincidence γ-ray spectra86

Kr + 116

Sn

Raw Ge spectrum

Cou

nts

per

chan

nel

Figure 7.10: HIPS-γ coincidence spectra for the binary reaction 86Kr+116Sn high-lighting the effect of Doppler correction on the detected lineshape [275].

169

Page 173: nuc_expt_phr03

Chapter 8

Spectroscopy With RadioactiveIons Beams.

The restriction of using beams of ions which are stable against radioactive decay

places constraints on those nuclei which can be produced in the laboratory for

study. Figure 8.1 shows the Segre chart and shows the predicted proton and

neutron drip lines. There around predicted to be around 7,000 nuclei which lie

within these driplines, of which less than 300 are stable against radioative decay.

N=Z

0 20 40 60 80 100 120 140

Neutron number N

0

20

40

60

80

100

Ato

mic

num

ber

Z

Figure 8.1: Segre chart showing the stable isotopes predicted proton and neutrondriplines [282].

Figure 8.2 shows those compound nuclei which can be formed using stable

170

Page 174: nuc_expt_phr03

beam target combinations. Note, virtually all of these systems are neutron defi-

cient with respect to the stable isotopes.

Proton Drip Line

Neutron Drip Line

0 20 40 60 80 100 120 140

Neutron number N

0

20

40

60

80

100

Ato

mic

num

ber

Z

Figure 8.2: Segre chart showing the compound nuclei which can be formed infusion-evaporation reactions using stable beam/target combinations [282].

In order to study the nuclear proporties of the entire nuclear chart, the ability

to induce nuclear reactions with radioactively unstable nuclear beams is impor-

tant. This chapter will look at methods of producing beams of radioactive ions

of the desired energy and intensity to be useful in nuclear structure studies.

8.1 Production of Radioactive Beams.

There are two main methods of producing radioactive nuclear beams [283], pro-

jectile fragmentation [284] and isotope separation on-line or ISOL [285, 286]. As

table 8.1 shows, both techniques have advantages and disadvantages depending

of the energy regime and beam intensity requirements of the experiment.

8.1.1 Projectile Fragmentation.

In a projectile fragmentation reaction, a high energy, heavy ion beam bomards a

target at energies of 30 MeV/A and above [294]. The result of a nuclear collision

at these energies is often that the beam loses a number of protons and neutrons

171

Page 175: nuc_expt_phr03

Method Advantages DisadvantagesProjectile Fast delivery times (∼ µs) low beam intensityFragmentation No chemical contraints Final beam deceleration difficult

Reliable operation Limited target thicknessHigh collection efficiency Large energy spreadSimple target design Moderate isobaric purity

ISOL Thick target Decay losses due to slow releaseHigh beam purity Needs post accelerationLow energy spread Radiation contamination in target‘Useful’ beam energies Complex target design

Yield depends on beam/target chem.

Table 8.1: Comparison of the ISOL and projectile fragmentation methods ofradioactive beam production.

but a beam like fragment carries on with a velocity similar to the iniital beam.

In this way, many radioactive species can be created using a single beam/target

combination. The beam-like fragment is then separated from the primary beam

using a set of electro-magnetic focussing and steering devices provided by for

example the LISE3 spectrometer [293] at the GANIL facility, France.

Figure 8.3 shows the identification spectrum (time of flight verses energy

loss) for the fragmentation of a 92Mo beam on a nickel target at an energy of

60 MeV per nucleon using the LISE3 spectrometer at GANIL. Note, the particle

identification afforded in fragmentation reactions using the time of flight (AQ

)

and energy loss (Z) technique means that each nuclear species produced can be

individually identified.

This technique of radioactive beam production has the advantages of high

beam purities and high collection efficiency (all the products are very forward

focussed in the lab frame due to the large beam velocity), but has disadvantages

for use in providing beams for fusion-evaporation reactions in that (a) the beam

intensities are generally smaller than are useful for evaporation reactions and

(b) a set of degraders have to be introduced to slow down the beam to energies

around the Coulomb barrier ∼ 4 − 5 MeV/A causes a spread in beam velocity.

However, this method is useful in experiments (as discussed below) where the

radioactive beam itself is the nucleus of interest.

172

Page 176: nuc_expt_phr03

Zr80

66As

__qA = 2

T Z

/23/ 15/2 012 2

B = 1.9068ρ Tm

Mo 42Zr 40Sr 38

states of beamdifferent charge

"beam"

Z

Kr 36

Figure 8.3: Particle identification spectrum from the fragmentation of 92Mo atGANIL with no degrader or Wien filter selection [287].

8.1.2 Particle Identification in Fragmentation.

One of the initial benefits of projectile fragmentation is that it enables the identi-

fication of nuclei on an event-by-event basis, and also can be used to show where

the drip-lines occur, by the absence of any nuclei in specific places on the particle

identification plot. For example, the nuclei 81Nb and 85Tc are thought not to be

bound against direct proton emission due to their absence in the fragmentation

of 92Mo [321].

The parameters used to identify the beam-like fragments are: (a) the mag-

netic rigidity of the dipoles which select the ions, Bρ; (b) the time-of-flight,

TOF,assuming that the distance, L, between the production target and the fo-

cal point, where the ions are implanted, is constant for every fragment; (c) the

173

Page 177: nuc_expt_phr03

energy-loss, ∆E and (d) the total kinetic energy, K, of the beam-like fragments.

From these quantities, the calculation of the fragment atomic mass number, A,

atomic number, Z, and charge, Q, is possible. Charged particles are deviated

by a magnetic field, B. This deviation is characterized by the bending radius,

ρ. This depends on the linear momentum, p, and the charge, Q = qe. From the

Lorentz law, one can derive that:

Bρ =p

qe(8.1.1)

Substituting for the linear momentum, this equation becomes

Bρ =Mβγc

qe, (8.1.2)

where M is the mass of the fragment, c is the speed of light in vacuum and

γ2 = 11−β2 with β = v

c, where v is the fragment velocity. If the mass is expressed

in atomic mass units (a.m.u.), M = Au where u is the atomic mass unit, and Bρ

is expressed in Tesla-meters (Tm), then:

Bρ = 3.105 × Aβγ

q. (8.1.3)

The measurement of the total kinetic energy of an implanted fragment, K, is

expressed in relativistic mechanics as:

K = Mc2(γ − 1). (8.1.4)

Therefore, it is possible to directly obtain the atomic mass number, A, for the

implanted fragment:

A =K

931.5 × (γ − 1). (8.1.5)

where K is expressed in MeV and A is in atomic mass units. Similarly, the

charge state, q is deduced from equations (8.1.3) and (8.1.5):

q =3.105 × βγK

931.5 × (γ − 1)Bρ. (8.1.6)

The total kinetic energy for each fragment, K, can be measured by adding

the energy deposited in the silicon ∆E and stopping detectors. The time-of-flight

from the production target to the final focus can be measured by taking the time

174

Page 178: nuc_expt_phr03

difference between a fast signal (rise time of a few nanoseconds) extracted from

the ∆E detector and the cyclotron radiofrequency (RF = 11 MHz). Calibrations

of the energy-loss in the silicon detectors and of the time-of-flight are performed by

simulating the velocity of the fragments at the LISE3 focal point and analyzing

the energy deposited by each fragment with a code based on Bethe’s formula

[320]. The atomic number, Z, is calculated using the energy-loss in a ∆E-silicon

detector. Relativistic corrections to Bethe’s formula lead to the expression:

∆E = a1Z2

β2[ln (a2β

2γ2) + a3β2 + a4] + a5 (8.1.7)

where the an are constants that can be obtained through calibration. The

atomic number, Z, can be expressed after integration of equation ([320]) as:

Z = c1

∆E

Y+ c2 + c3

∆E

Y+ c4β (8.1.8)

where:

Y =ln (5930 × β2γ2)

β2− 1 (8.1.9)

and c1, c2, c3 and c4 are constants which can be fitted.

The complete reconstruction of A, q and Z is therefore possible from the

measurement of the time-of-flight, the energy loss, and the total kinetic energy.

Figure 8.4 shows a calibrated particle identification spectrum following the frag-

mentation of a 60 MeV per nucleon 92Mo beam at GANIL [320].

Gamma-Ray Spectroscopy Using Projectile Fragmentation.

The products from projectile fragmentation reactions may populate long lived

excited states. If the lifetime of these states is long compared to the flight path

of the separator, their decay can be measured at the end of the separator by an

array of germanium detectors [287, 290].

Figure 8.5 shows a more selective region of the identification data from fig-

ure 8.3, with and without the condition that a delayed gamma-ray must be mea-

sured with 100 µ seconds of he fragment stopping in a silicon telescope detector.

Those fragments which are transmitted in an isomeric excited state will decay in

the silicon telescope and their gamma-rays detected in the surrounding germa-

nium detector array. By gating on the nuclear species of interest in the particle

175

Page 179: nuc_expt_phr03

1 2 4 8 16 32 64 128 256 512 1024 2048 4096

1.950 1.975 2.000 2.025 2.050 2.075 2.100 2.125

A/q

26

28

30

32

34

36

38

40

42

44

46

48

Z

N=Z

Nb

Tc

Y

Rb

Figure 8.4: Calibrated particle identfication spectra from the fragmemtation of92Mo [320].

identification spectra one can project the gamma-ray energy and time spectrum

for the isomeric decay.

176

Page 180: nuc_expt_phr03

(a) all (b) gammas

Kr (36)Rb (37)

Ge (32)As (33)

Se (34)Br (35)

Tz =3/2 1 1/2

time of flight (A/Q)

ener

gy lo

ss

Kr-73

Se-69

‘GRZYWACZ PLOT’ TO ID ISOMERS

92-Mo @ 60 MeV/A on Ni target at GANIL

Figure 8.5: Particle identification spectrum from the fragmentation of 92Mo atGANIL. Note the extra condition of observing a delayed gamma-ray highlightsthose nuclei in which isomeric decays have been observed.

177

Page 181: nuc_expt_phr03

Figure 8.6 shows a two-dimensional plot of time verses gamma-ray energy

taken for the fragmentation of a 92Mo beam from GANIL [291]. Note the lines

extending out corresponding to gamma-rays from isomeric decays. Note also,

the presence of the line corresponding to ‘prompt’ radiation, which serves as a

useful ‘time-zero’ calibration. The spectrum on the right of figure 8.6 has been

corrected for low-energy time walk [291].

0 500

Time Difference (ns)

0

500

1000

1500

2000

2500

3000

3500

4000

Gam

ma–

ray

ener

gy (

keV

)

γ

0 500

Time Difference (ns)

0

500

1000

1500

2000

2500

3000

3500

4000

Gam

ma–

ray

ener

gy (

keV

)

E = 734 keV

a) b)

Figure 8.6: Two-dimensional gamma-ray energy verses time spectra showing thedecays of the isomeric states following the fragmentation of a 60 MeV/A beamof 92

42Mo. [291]. Note the correction for ‘time-walk’ for lower energy gamma-raysin the right hand spectrum.

Figure 8.7 shows the gamma-ray and time spectra associated with the decay

of the 42 ns isomer in 74Kr obtained by gating on the identification spectra shown

in figure 8.5.

This is a very powerful technique in identifying isomeric decays in exotic nuclei

and can be used for both neutron deficient [287, 288, 290, 291] and neutron rich

systems [308, 313, 314, 315, 316]. Figures 8.8 and 8.9 show the gamma-ray energy

and time spectra for decays associated with isomeric states in the neutron rich

nucleus 6828Ni40, produced by the fragmentation of a 86Kr beam. Note the presence

of the 511 keV gamma-ray line, corresponding to the 1770 keV, 0+ → 0+ decay

by internal pair formation.

It has been observed [288] that the population of yrast and non-yrast iso-

178

Page 182: nuc_expt_phr03

τ = 42 ± 8 ns

150 200 250 300 350

time difference (ns)

100

101

coun

ts

Figure 8.7: Gamma-ray energy and associated time spectra showing the decay ofthe isomeric state in 74

36Kr from the fragmentation of a 60 MeV/A beam of 9242Mo.

[296, 291].

meric states in intermediate energy projectile fragmentation reactions varies sig-

nificantly, with yrast states being favoured. Values of the isomeric ratio for nuclei

produced using fragmentation reactions have been found to range dramatically

from case to case [288]. Indeed, the production of nuclei in their isomeric state

has been found to be dependent on the reaction mechanism and the velocity of

the fragment compared to that of the beam [297].

Note, the effective lifetimes of the nuclei through a separator are also extended

due to relativistic effects, such that through the separator, the effective lifetime

is

τrmeff = τo1√

1 − β2(8.1.10)

where β is vc.

Fragmentation can also be used at relativistic energies corresponding to pri-

mary beams of hundred of MeV per nucleon [309, 310, 308, 312]. This is particu-

larly useful for the identification of heavy, exotic fragments [308, 312, 316]. The

high beam velocity, as provided using the SIS accelerator at GSI, allows both a

larger production of the most exotic nuclei (since thicker targets can be used).

179

Page 183: nuc_expt_phr03

T1/2=290 ns (exp)

T (us)

68m2Ni

Figure 8.8: Time spectrum gated on 511 keV decays in 68Ni from the fragmenta-tion of 86Kr at GANIL. The lifetime suggests an E0 decay proceeding by internalpair formation [295].

The FRagment Separator (FRS) at GSI can be used to separate and identify

on an event by event basis the nuclei transmitted [307] using time of flight and

energy loss measurements [307, 309, 311]. In addition, the higher primary beam

velocity means that the ions have a higher probability of being fully stripped of

electrons through the spectrometer. A knowledge of the charge state is important

in order to resolve charge state ambiguities in the time of flight signal (∼ AQ

, see

fig. 8.11.) Figure 8.11 shows how the charge states of the transitted nuclei can

be obtained by measuring their relative position through the spectrometer as a

function of their time of flight. Fully stripped ions do not change their position

through the spectrometer, whereas hydrogen and helium like one do, causing a

longer effective time of flight (see figure 8.11).

High spin spectroscopy can be achieved following isomeric decays in these

fragments. Figures 8.12 shows the gamma-ray spectrum gated on (a) 179W show-

ing the decay of the Kπ = 352

−isomer. A prescription of how to estimate the

much angular momentum transfered to nucleus in a projectile fragmentation re-

action can be found in reference [317]. Figure 8.14 shows the predicted average

spins of the fragments for a Pb on Pb collision at 1 GeV per nucleon.

180

Page 184: nuc_expt_phr03

68m2Ni

Figure 8.9: Gamma-ray gate on 68Ni from the fragmentation of 86Kr showing the511 keV lines associated with internal pair formation [295].

181

Page 185: nuc_expt_phr03

Auc

IC:time

IC:dE,Q

γ

MW:x,y

βeρB

Q

Segmented Clover

=

+

Degrader

SCI

TOF

Super Cloversmall Clover

catcher

SCI

primary beam:

Pb @ 1 GeV

production target

Degrader

dipole : Bρmiddle focus S2

end focus S4

Array

Figu

re8.10:

Sch

ematic

ofth

eset-u

pusin

gth

eG

SIFragm

ent

Sep

aratorto

inves-

tigatedecay

sfrom

isomeric

states[315].182

Page 186: nuc_expt_phr03

2.56 2.58 2.60 2.62 2.64

A/Q

–40

–20

0

SC

21x

(mm

)

fully–stripped hydrogen–like helium–like

2.56 2.58

A/Z

–60

–40

–20

0

20

40

60

80

SC

41x

(mm

)

187

Ta 188

Ta

189

W 190

W

191

Re

192

Re 193

Re

194

Os

195Os

196Os

2.60 2.62

A/(Z–1)

192

Re 193

Re

194

Os 195

Os 196

Os

197

Ir 198

Ir 199

Ir

202

Pt 200

Pt 201

Pt

203

Au

2.64

A/(Z–2)

200

Pt 201

Pt

203

Au 204

Au

205

Hg 206

Hg

209

Tl

Figure 8.11: Fragment identification spectra following the fragmentation of a1 GeV/u 208Pb, highlighting the different charge states of the ions for an FRSsetting centred on 191W ions [316].

183

Page 187: nuc_expt_phr03

Figure 8.15 shows background subtracted gamma-ray spectra associated with

the decay of isomers populated in a variety of neutron-rich nuclei around A ∼190,

for different charge states. Note in particular, the 15 ns isomer in 200Pt, which

is observed due to the effective lengthening of the decay half-life through the

spectrometer by the ‘switching off’ of the electron-conversion component of the

decay [291, 316].

Figure 8.12: Gamma-ray spectra associated with the fragmentation of a 1 GeV/u208Pb, for fully stripped ions around 177Ta [313].

Elemental identification in relativistic fragmentation is achieved by a com-

bination of energy loss (usually measured in gas detector, known as a MUSIC

chamber, see fig 8.16) or by position on a variety of scintilation detectors through

the spectrometer.

8.1.3 Isomeric Ratios and Angular Momentum Popula-

tion.

The isomeric ratio, R, is defined as the probability that in the reaction a nucleus

is produced in an isomeric state [322, 323, 324].

These can be determined in the following way.

184

Page 188: nuc_expt_phr03

Firstly, the observed decay yield is calculated :

Y =Nγ (1 + αtot)

ǫeff bγ, (8.1.11)

where Nγ is the number of counts in the gamma line depopulating the isomer

of interest, αtot is the total conversion coefficient for this transition, ǫeff is the

effective efficiency and bγ is the probability that the decay proceeds through this

transition (i.e. the absolute branching ratio). The isomeric ratio is then given by

:

R =Y

Nimp F G, (8.1.12)

where Nimp is the number of implanted heavy ions, F is a correction factor for

the in-flight isomer decay losses and the factor G corrects for the finite detection

time of gamma radiation. The factor F is calculated from :

F = exp

[

−(

λq1TOF1

γ1+ λq2

TOF2

γ2

)]

, (8.1.13)

where TOF1 (TOF2) is the time of flight through the first (second) stage of

the spectrometer (eg. the FRS), γ1 (γ2) is the corresponding Lorentz factor and

λq1 (λq2) is the decay constant for the ion in the charge state q1 (q2). In most of

the experiment using relativistic ions, the in-flight ions were highly charged (in

most of the cases studied, the ions are fully stripped), and the decay constants

λq can differ considerably from the value for an electrically neutral atom, λ.

For the fully stripped ion, λ0 can be calculated from :

λ0 = λ∑

i

bγi

1 + αitot

, (8.1.14)

where the summation is over all the decay branches depopulating the isomer.

Finally, the correction factor G is calculated using:

G = exp(−λ ti) − exp(−λ tf), (8.1.15)

where ti and tf are the gamma delay-time limits set in the off-line analysis to

produce the delayed gamma spectrum.

When more than one gamma-ray line was observed to depopulate an isomer,

the isomeric ratio was calculated separately for the strongest lines and then av-

eraged.

185

Page 189: nuc_expt_phr03

In some cases, more than one isomer in the same nucleus is populated in the

reaction and a lower lying isomer may be partly fed by the delayed decay of a

higher lying metastable state. We adopt here the definition of isomeric ratio as

the probability that a state is populated promptly after production of the nucleus

in the reaction. Then, it can be shown that in a case where the upper state decays

with the probability (branching) bUL to the lower one, the isomeric ratio for the

latter can be calculated by :

RL =YL

Nimp FLGL

− bULRU

FLGL

[

λLGU − λUGL

λL − λU

FU +λ0

U

λ0L − λ0

U

GL (FU − FL)

]

,

(8.1.16)

where the indexes ’L’ and ’U ’ refer to the lower and the upper states, respec-

tively, and the second term on the right side represents the correction due to

feeding from the upper state.

The Sharp Cut-Off Model

To describe the population of an isomeric state in a fragmentation reaction we

separate the process into two steps. In the ablation phase of the reaction, highly

excited prefragments evaporate nucleons until the final fragment is formed with

an excitation energy below the particle emission threshold. Subsequently, a sta-

tistical gamma cascade proceeds down to the yrast line and then along this line

to the ground state. If a long-lived state lies on this decay path, part of the

cascade may be hindered or stopped depending on the lifetime of the isomer.

The isomeric ratio is equal to the probability that gamma decay from the initial

excited fragments proceeds via this isomeric state.

The crucial aspect of the first step is the distribution of the angular momentum

in the ensemble of the excited fragments just prior to the gamma de-excitation

step. This problem was addressed by de Jong, Ignatyuk and Schmidt [298] who

applied the statistical abrasion-ablation model [299] of fragmentation. Assum-

ing that any angular momentum taken away by evaporating particles is small

and can be neglected, they calculated the angular momentum distribution of

the final fragment as the superposition of the angular momenta of all prefrag-

ments contributing to the final fragment of interest using the ABRABLA code

[299]. Furthermore, they have shown that for a large mass difference between the

186

Page 190: nuc_expt_phr03

projectile and the fragment this distribution can be approximated by a simple

analytical formula :

PJ =2J + 1

2σ2f

exp

[

−J(J + 1)

2σ2f

]

, (8.1.17)

where σf , the so called spin-cutoff parameter of the final fragments, is given

by :

σ2f = 〈j2

z 〉(Ap − Af ) (νAp + Af )

(ν + 1)2(Ap − 1). (8.1.18)

Ap and Af denote the projectile and fragment mass numbers respectively, ν

is the mean number of evaporated nucleons per abraded mass unit and 〈j2z 〉 is the

average square of the angular-momentum projection of a nucleon in the nucleus.

It is generally assumed that the abrasion of one nucleon induces an excitation

energy of about 27 MeV [301], whereas the evaporation of a nucleon decreases

the energy by about 13 MeV, hence the parameter ν = 2 is taken. Values of 〈j2z 〉,

estimated on the basis of a semi-classical consideration of the angular-momentum

distribution in the Saxon-Woods potential [300, 298], are written as :

〈j2z 〉 = 0.16A2/3

p (1 − 2

3β), (8.1.19)

where β is the quadrupole deformation parameter.

Given the angular momentum distribution of the final fragment, one can con-

sider the probability that gamma decays will lead to an isomeric state of spin

Jm. First, we assume that the initial excitation energies are well above the exci-

tation energy of the isomer. One can make the extreme simplifying assumption

that all states with J ≥ Jm, and only those, decay to the isomer. A similar

approach, known in the literature as the ‘sharp cut-off model’, has been used

in studies of angular momentum distributions in compound nuclei [302] and in

fission fragments [303]. From Eq. 10 it follows :

Rth =∫ ∞

Jm

PJ dJ = exp

[

−Jm (Jm + 1)

2σ2f

]

. (8.1.20)

Substituting ν = 2 and introducing ∆A = Ap−Af , the above equations yield:

σ2f = 0.0178 (1 − 2

3β)A2/3

p

∆A (3Ap − ∆A)

Ap − 1. (8.1.21)

187

Page 191: nuc_expt_phr03

8.1.4 Projectile Fission

A similar method of producing high energy radioactive beams to projectile frag-

mentation is that of projectile fission. Here a heavy beam, such as 238U bombards

a light target (such as 1H). The heavier element usually fissions in such a reac-

tion, giving rise to a large number of neutron rich fragments moving forward in

the lab frame. These fragments are subsequently collected and identified using a

mass separator such as the FRS at GSI, Darmstadt, Germany [304]. Using this

technique, many new neutron rich isotopes have been identified for the first time

[318].

188

Page 192: nuc_expt_phr03

Figure 8.13: Predicted spin distribution for 179W fragments in the fragmentationof a 1 GeV per nucleon 208Pb on a 9Be target using the abrasion-ablation model[310], taken from ref [315].

8.1.5 Intermediate Energy Coulex.

Another technique which can employed to study the collective states of very ex-

otic nuclei produced in projectile fragmentation reactions is to Coulomb excite

the products on a heavy stopper (such as Pb) and measure the detected gamma-

rays. The large fragment velocity means that Doppler broadening effects can be

large and rather degrade the energy resolution. However, since the number of

lines is generally quite small, and gamma-ray detection efficiency is at a premium

due to the relatively small number of events, a higher efficiency, lower resolution

detetctor, such as NaI(Tl) is used instead of germanium. Work using this tech-

nique has can give direct measurements of both the energy of the lowest 2+ state

and the associated B(E2) for it decay to the ground state in even-even nuclei very

far from stability. Recent work using this technique on the neutron rich N=28

nuclei has suggested a break down of this ‘magic number’ for very neutron rich

species, such as 4416S [305, 306].

189

Page 193: nuc_expt_phr03

Neutron Number

Prot

on N

umbe

r

Figure 8.14: Predicted average fragments spins using the abrasion-ablation model[310] for 208Pb on 208Pb at 1 GeV per nucleon.

8.1.6 Double Fragmentation and In-beam spectroscopy

For double fragmentation at RIKEN in 34Mg see [326]. In-beam spectroscopy at

GSI see [327]. For in-beam fragmentation gamma-ray spectroscopy at GANIL,

see [328].

8.1.7 Beta-Decay Measurements.

The same technique as used for isomer decays can also be applied to measurement

of fast β-decay lifetimes [319, 320] (see figures 8.17 and 8.18).

8.2 ISOL Based Techniques.

The isotope separation on-line method uses a primary reaction to create the ra-

dioactive nuclei of interest (such as a fusion-evaporation reaction or the spallation

of a heavy nuclei by protons or beam induced fission). The radioactive nuclei must

then diffuse out of their production target where they are ionised and selected

by mass (using a dipole magnet). The ions must then be accelerated to useful

experimental energies and thus some form of post-accelerator is required.

For many years, ISOL based experiments have been performed to investigate

the ground state and beta-decaying properties of exotic nuclei, using instruments

such as the GSI On-Line Mass Separator [333, 334, 335, 336, 337, 338, 339, 340,

341, 342, 343, 344].

190

Page 194: nuc_expt_phr03

Figure 8.15: Gamma-ray spectra associated with isomers populated following thefragmentation of a 1 GeV/u 208Pb beam, for fully stripped, hydrogen line andhelium-like ions around 191W [316].

8.2.1 On-Line Mass Separators.

A typical on-line mass separator (see ref. [330] for a review) is effectively a dipole

magnet which separates ions emitted from a thermal (or similar type) ion-source

[331, 332]. The ions are usually singly ionized and extracted by a voltage, V .

Thus they have energies QeV , where Q is usually 1.

Following the Lorentz equation, these ions are bent in a radius ρ, via a mag-

netic field according to the equation,

Bρ =Auv

Qe(8.2.22)

where A is the nuclear mass number, u is the atomic mass units (approx1.6×10−27 Kg), e is the electron charge, Q is the ionic charge state and v is the ion

191

Page 195: nuc_expt_phr03

Figure 8.16: Energy loss signals in the music chamber at the end of the FragmentRecoil Separator, which allows elemental identification of the fragments. Thesespectra are all for the fragmentation of a 208Pb beam at 1 MeV/a, but withdifferent settings of the spectrometer to transmit different nuclei the final focus[313].

velocity. The ion velocity, v can be obtained since the total kinetic energy, E, of

the ions can be deduced from the expression,

E =1

2Auv2 (8.2.23)

substituting in to equation 8.2.22, we obtain

Bρ =Au√

2EAu

Qe=

√2EAu

Qe(8.2.24)

where Q is usually equal to unity.

Figure 8.19 shows gamma-ray spectra for mass separated nuclei with A=176,177

and 178 from the GSI on-line mass separator, formed using identical reaction,

ion-source and tape cycle time conditions [329].

Channel Selection, Grow-in Curves and Tape Systems.

In addition to the mass selection provided by the on-line mass separator, other

channel selection can be provided by, (1) the choice of reaction to form the nuclei

192

Page 196: nuc_expt_phr03

1

2

4

8

16

32

64

128

256

512

1024

1.960 1.970 1.980 1.990 2.000 2.010 2.020

A/q

34

36

38

40

42

44

46

ZN=Z

Tc

Nb

Y

Rb

Figure 8.17: Particle id-spectra as in figure 8.4, but with hardware beam-offcondition for N=Z nuclei [320].

of interest; (2) the use of various types of ions sources such as FEBIAD and

TIS [331, 332]; and (3) the use of different tape cycle speeds to select decays of

different lifetimes (typically the furthest from stability are the shortest lived).

The extracted ions are often sent to a tape counting station, where the gamma-

rays from the radioactivity (with liofetimes usually longer than hundreds of mil-

liseconds) is collected. In a tape drive system, the tape can then be caused to

move with different frequencies, which means that the detection system becomes

sensitive to different rates of decays. For example, if the tape moves the ra-

dioactivity away quickly, only the fast decaying products will be observed in the

detection system (since the longer lived decays will be moved away from the sight

of the detectors before they decay).

One Component Grow-ins.

For a single component decay, the grow-in curve, which represents the decay rate

of a specific activity is given by assuming a constant implantation rate, I0, then,

193

Page 197: nuc_expt_phr03

01020304050607080

0 200 400 600 800 1000

time (ms)

counts 74Rb

72(18) ms

0

50

100

150

200

250

300

350

0 200 400 600 800 1000

time (ms)

counts

82Nb52(6) ms

0

10

20

30

40

50

0 200 400 600 800 1000

time (ms)

counts

86Tc45(12) ms

020406080

100120140160180

0 200 400 600 800 1000

time (ms)

counts 78Y

50(8) ms

Figure 8.18: Time spectra for delayed β+ decay in the same strip as the detectedN=Z recoil, with the beam-off condition provided by figure 8.17 [320].

dN

dt= I0 − λN (8.2.25)

where λN is the decay rate, with a decay constant, λ.

Trying a solution of the form

N = a(1 − e−λt) (8.2.26)

where a is a constantto be determined, then differentiating equation 8.2.26,

we obtain

dN

dt= λae−λt = I0 − λa(1 − e−λt) (8.2.27)

therefore,

I0 = λa (8.2.28)

194

Page 198: nuc_expt_phr03

which gives us the general solution for a one component grow-in curve of

N =I0λ

(1 − exp(−λt)) (8.2.29)

Thus the count rate, λN = I0(1 − exp(−λt)).

Two component Grow-ins.

In the case of a two component decay curve, where for example a β decay feeds

into a long lived isomeric state, a two component fit for the grow-in curve must

be performed.

If I0 is the production/implantation rate as above, the higher lying state has

a decay rate of λ1N1 and the lower lying state (ie. the one fed by state 1) has an

intrinsic decay rate of λ2N2, then the total decay rate as measured for the two

states is given by the decay rate of state 1 (to produce state 2), minus the decay

rate of state 2, ie.

dN2

dt= λ1N1 − λ2N2 (8.2.30)

The solution to equation 8.2.30 is given by trying a solution of the form,

N2 = a(1 − e−λ1t) + be−λ2t − c (8.2.31)

with the boundary conditions that at t = 0, N2 = 0 = b− c.

therefore, differentiating equation 8.2.31, we obtain,

dN2

dt= aλ1e

−λ1t − bλ2e−λ2t = λ1N1 − λ2N2 (8.2.32)

and therefore

dN2

dt= I0(1 − e−λ1t) − λ2a(1 − e−λ1t) − λ2be

−λ2t + λ2c (8.2.33)

re-arranging and collecting the terms

aλ1e−λ1t − bλ2e

−λ2t = I0 − λ2(a− c) − (I0 − λ2a)e−λ1t − λ2be

−λ2t (8.2.34)

Equation 8.2.34 requires that at t = ∞, I0 = λ2(a − c) and comparing the

terms for the e−λ1t term, aλ1 = (I0 − λ2a). Rearranging, this gives that

195

Page 199: nuc_expt_phr03

a =I0

λ2 − λ1

(8.2.35)

also, collecting and comparing the non-exponential terms in equation 8.2.34

gives

I0 = λ2(I0

λ2 − λ1

− c) (8.2.36)

re-arranging this gives

c = I0(1

λ2 − λ1− 1

λ2) = b (8.2.37)

196

Page 200: nuc_expt_phr03

Figure 8.19: Gamma-ray spectra for different masses taken from the GSI on-linemass separator, following the binary collision of a 11.4 MeV/u 136Xe beam witha Ta target. In each case, the tape cycle time was taken to be 8 seconds. [329].

197

Page 201: nuc_expt_phr03

Substituting the results of equations 8.2.35 and 8.2.37 into equation 8.2.33

gives that the decay rate of level 2, is given by

λ2N2 =λ2I0

λ2 − λ1(1 − e−λ1t) − I0(

λ2

λ2 − λ1− 1)(1 − e−λ2t) (8.2.38)

Note that the limits of equation 8.2.38 give when λ2t >> 0 and λ2 >> λ1 (ie.

upper level is much longer lived than lower one,

λ2N2 ≈ I0(1 − e−λ1t) (8.2.39)

ie. the decay rate depends on the lifetime of the upper state.

Similarly, when λ1t >> 0 and λ1 >> λ2

λ2N2 ≈ I0(1 − e−λ2t) (8.2.40)

If we substitute r = λ2

λ2−λ1, into equation 8.2.38, it reduces to

count rate = const.× r(1 − e−λ1t) − (r − 1)(1 − e−λ2t) (8.2.41)

Figure 8.20 shows the grow-in curves for the decays from the ground state

and excited states of 177Lu into states in 177Hf, following a binary reaction at GSI

[329].

Figure 8.21 shows the effect of the different tape cycle times for A=177 nuclei

following the binary reaction of 136Xe on a natural Ta target at the GSI on-line

mass separator. Note the presence of longer lived decays in the spectra with the

extended cycle times.

8.2.2 The 19Ne∗+40Ca Experiment at Louvain La Neuve.

To date there is one report of the use of a (short lived) radiaoctive ion beam

used to induce a fusion evaporation reaction [345]. This used the cyclotrons at

Louvain La Neuve to initially produce radioactive 19Ne (T 1

2

=17 s) by accelerating

protons onto a 19F target and using the charge exchange reaction 19F(p,n)19Ne.

The 19Ne beam was then injected into a second cyclotron where it was acceler-

ated to Coulomb barrier energies and used to bombard a target of 40Ca. The

target position was surrounded with germanium detectors to measure discrete

gamma-rays and a thin silicon charged particle detector (LEDA) to identify any

198

Page 202: nuc_expt_phr03

evaporated α-particle and protons from fusion-evaporation events. Typical 19Ne

beam currents were around 0.1 pnA.

Figure 8.22 highlights the problems associated with doing in beam gamma-

ray spectroscopy with intense radiaoctive beams. The raw gamma-ray spectra are

completely dominated by transitions associated with the decay of the radioactive

beam (in this case 511 keV annihilation gamma-rays coming from the β+ decay

of 19Ne).

However, as figure 8.22 shows if the beam is pulsed and the time structure

of the measured gamma-rays can be recorded with respect to the beam pulses,

most of this unwanted ‘radioactive’ background can be subtracted, leaving a

‘pure’ spectrum of transitions associated with fusion-evaporation events.

As figure 8.24 shows, the effect of charged particle gating is less dramatic

than performing a subtraction using the time spectra gating using the out-of-

beam spectra, however, a significant improvement in the signal to noise is clearly

observed.

Figure 8.25 shows a comparison of the relative yields of the various evaporation

residues for the 19Ne induced reaction compared to a 19F (stable beam) reaction.

While, it is apparent that the cross-section of the most neutron deficient nuclei in-

creases by using a more neutron deficient (radioactive) beam (19Ne), the increase

is negated by the decrease in beam intensity. It is clear that future radioactive

beam facilities which will desire to use such beams for fusion-evaporation reac-

tions will require both more exotic beams (more than one nucleon from stability)

coupled to intensities of a least 1 pnA.

199

Page 203: nuc_expt_phr03

Figure 8.20: Grow-in curves for the beta-decay of the high-spin isomer in 177KLu(upper panel) and ground state of 177Lu into 177Hf. The upper panel populates asecond isomeric state in 177Hf and therefore requires a two component fit, whilethe lower panel is a simple one component fit with a half-life of 1.9 hours [329].

200

Page 204: nuc_expt_phr03

Figure 8.21: Effect of the tape cycle time in selecting different decay half-lifes.A=177 spectra with identical reaction and source conditions, but different tapecycle times [329].

201

Page 205: nuc_expt_phr03

Figure 8.22: Differents types of channel selection to pick out transitions fromfusion products for the 19Ne+40Ca reaction [345].

202

Page 206: nuc_expt_phr03

Figure 8.23: TDC time spectra from the (a) 19Ne and (c) 19F beams at LLN. Notethe increase in counts for the ”beam on” period corresponding to gamma-raysfrom beam-induced reactions.

203

Page 207: nuc_expt_phr03

Figure 8.24: Channel selection afforded by gating on evaporated charged particlesin the 19Ne+40Ca experiment [345].

204

Page 208: nuc_expt_phr03

Figure 8.25: Comparison of yields from various residual channels using bothstable (19F) and radioactive 19Ne beams on a 40Ca target [345].

205

Page 209: nuc_expt_phr03

Bibliography

[1] H. Ejiri and M.J.A. de Voigt, Gamma-ray and Electron Spectroscopy in

Nuclear Physics, Clarendon Press, Oxford (1989)

[2] E.S. Paul et al. J. Phys. G17 (1991) p605

[3] P.H. Regan et al. Nucl. Phys. A586 (1995) p351

[4] A.P. Byrne et al. Nucl. Phys. A548 91992) p131

[5] K.S.Krane, Introductory Nuclear Physics, John Wiley and sons, New York,

(1988)

[6] T. Carreyre et al. Phys. Rev. C62 (2000) 024311

[7] A. Bohr and B. Mottelson Nuclear Structure vol. 2 (1975)

[8] T. Kibedi, private communication

[9] C.S Purry et al. Phys. Rev. Lett. 75 (1995) p406

[10] Table of Isotopes eds. C.M. Lederer and V.S. Shirley, (Wiley) (1978)

[11] F. Rosel et al. At. Data. Nucl. Data. Tab. 21 (1978) p291

[12] J. Kantele, Heavy Ions and Nuclear Structure, vol. 5. Nuclear Science

Research Conference Series, (1984) p391, Harwood academic publishers

(1984), edited by B. Sikora and Z. Wilhemi

[13] D.A. Bell, C.E. Aveledo, M.G. Davidson and J.P. Davidson, Canadian

Journal of Physics 44, 2542 (1970); A. Passoja and T. Salonen, Depart-

ment of Physics, University of Jyvaskyla, Research Report No. 2/1986,

unpublished

206

Page 210: nuc_expt_phr03

[14] K. Heyde and R.A. Meyer Phys. Rev. C37, 2170 (1988)

[15] H. Mach et al., Phys. Rev. C42, 793 (1990)

[16] T. Kibedi et al. Nucl. Phys. A567 (1994) p183

[17] J.L. Wood et al. Nucl. Phys. A651 (1999) p323

[18] J.P. Davidson, Rev. Mod. Phys. 37 (1965) p105

[19] A. Kuhnert et al. Phys. Rev. C47 (1993) p2386

[20] V.A. Krutov and O.M. Knyazkov, Ann. Phys. 25 (1970) p10

[21] M.J.A. de Voigt, J. Dudek and Z. Szymanski, Rev. Mod. Phys. 55 (1983)

p949

[22] K.R. Pohl et al. Phys. Rev. C53 (1996) p2682

[23] E Der Mateosian and A.W. Sunyar, Atomic Data and Nuclear Data Tables

13 (1974) p407

[24] B. Crowell et al, Phys. Rev. C45 (1992) p1564

[25] P.H. Regan et al. Phys. Rev. C54 (1996) p1084

[26] K.S. Krane, R.M. Steffen and R.M. Wheeler, Nuclear Data Tables 11 (1973)

p351

[27] A. Kramer-Flecken et al. Nucl. Inst. Meth. A275 (1989) p333

[28] C. Bargholtz et al. Nucl. Inst. Meth. A256 (1987) p513

[29] S. Mohammadi PhD thesis, University of Surrey (1997)

[30] R. Bengtsson and S. Frauendorf Nucl. Phys. A327 (1979) p139

[31] S. Frauendorf. Physica Scripta 24 (1981) p349

[32] R.Bengtsson et al. At. Data. Nucl. Data. Tab. 35 (1986) p15

[33] P. Fallon et al. Phys. Lett. B218 (1989) p137

[34] G.I. Harris et al. Phys. Rev. 139 (1965) p1113

207

Page 211: nuc_expt_phr03

[35] P. Ring and R. Schuck The Nuclear Many Body Problem, (1980) Spinger-

Verlag, New York

[36] F.S. Stephens et al. Phys. Rev. Lett. 29 (1972) p438

[37] P.M. Walker et al. Nucl. Phys. A568 (1994) p397

[38] P.H. Regan et al. Phys. Rev. C51 (1995) p1745

[39] K. Nakai, Phys. Lett. B34 (1971) p269

[40] F. Donau, Nucl. Phys. A471 (1987) p469

[41] D. Ward et al. Nucl. Phys. A529 (1991) p315

[42] P.M. Walker et al. Phys. Lett. B309 (1993) p17

[43] J.M. O’Donnell et al. Phys. Rev. C38 (1988) p2047

[44] J.L. Wood et al. Phys. Rep. 215 (1992) p101

[45] G.J. Lane et al. Nucl. Phys. A586 (1995) p316

[46] G.J. Lane et al. Phys. Lett. B324 (1994) p14

[47] J. Heese et al. Phys. Lett. B302 (1993) p390

[48] A.M. Baxter et al. Phys. Rev. C48 (1993) R2140

[49] P.J. Twin et al. Phys. Rev. Lett. 57 (1986) p811

[50] R.V.F. Janssens and T.L. Khoo Ann. Rev. Nucl. Part. Sci. 41 (1991) p321

[51] P.J. Nolan and P.J. Twin Ann. Rev. Nucl. Part. Sci. 38 (1988) p533

[52] J. Dudek et al. Phys. Rev. Lett. 59 (1987) p1405

[53] P.H. Regan et al. J. Phys. G18 (1992) p847

[54] R. Diamond et al. Phys. Rev. C41 (1990) R1327

[55] S.M. Mullins et al. Phys. Rev. C45 (1992) p2683

[56] R.M. Clark et al. Phys. Rev. Lett. 76 (1996) p3510

208

Page 212: nuc_expt_phr03

[57] M.A. Bentley et al. Phys. Rev. Lett. 59 (1987) p2141

[58] D. Nisius et al. Phys. Lett. B393 (1997) p18

[59] J. Wilson et al. Phys. Rev. Lett. 74 (1995) p1950

[60] G. Hackman et al. Phys. Rev. C52 (1995) R2293

[61] C. Baktash et al. Phys. Rev. Lett. 74 (1995) p1946

[62] C. Baktash et al. Ann. Rev. Nucl. Part. Sci. 45 (1995) p485

[63] T. Byrski et al. Phys. Rev. Lett. 64 (1990) p1650

[64] D. Santos et al. Phys. Rev. Lett. 74 (1995)

[65] R.M. Clark et al. Phys. Lett. B343 (1995) p59

[66] P. Fallon private communication

[67] T.L. Khoo et al. Phys. Rev. Lett. 76 (1996) p1583

[68] A. Lopez-Martens et al. Phys. Lett. B380 (1996) p18

[69] S.J. Gale et al. J. Phys. G21 (1995) p193

[70] H. Timmers et al. J. Phys. G20 (1994) p287

[71] J. Simpson et al. Phys. Lett. B262 (1991) p388

[72] R. Wadsworth et al. Phys. Rev. C50 (1994) p483

[73] R. Wadsworth et al. Nucl. Phys. A559 (1993) p461

[74] V. Janzen et al. Phys. Rev. Lett. 72 (1994) p1160

[75] I. Ragnarsson et al. Phys. Rev. Lett. 74 (1995) 3935; A.V. Afanasjev and

I. Ragnarsson, Nucl. Phys. A586 (1995) 387

[76] G.J. Lane et al. Phys. Rev. C55 (1997) R2127

[77] J. Simpson et al. Phys. Rev. Lett. 53 (1984) p141

[78] R.V.F. Janssens et al. Phys. Lett. B106 (1991) p167

209

Page 213: nuc_expt_phr03

[79] R.G. Helmer and C.W. Reich Nucl. Phys. A114 (1968) p649; A211 91973)

p1

[80] G.D. Dracoulis et al. J. Phys. G23 (1997) p1191; G.J. Lane et al. Phys.

Rev. C60 (1999) 067301

[81] S.M. Mullins et al. Phys. Lett. B393 (1997) p279

[82] T.L. Khoo et al. Phys. Rev. Lett 37 (1976) p823

[83] B. Crowell et al. Phys. Rev. C53 (1996) p1173

[84] M. Dasgupta et al. Phys. Lett. B328 (1994) p16

[85] P. Choudhury et al. Nucl. Phys. A485 (1988) p136

[86] F.G. Kondev et al. Phys. Rev. C54 (1996) R459

[87] F.G. Kondev et al. Nucl. Phys. A601 (1996) p195

[88] F.G. Kondev et al. Nucl. Phys. A617 (1997) p91

[89] A.B. Migdal, Nucl. Phys. 13 (1959) p655

[90] H.F. Brinkmann et al. Nucl. Phys. A133 (1969) p648

[91] D. Ward et al. Nucl. Phys. A117 (1968) p309

[92] D. Parkinson et al. Nucl. Phys. A194 (1972) p443

[93] A.M. Bruce et al. Phys. Rev. C55 (1997) p620

[94] A.M. Bruce et al. Phys. Rev. C50 (1994) p480

[95] I. Ahmad and P.A. Butler Ann. Rev. Nucl. Part. 43 (1993) p71

[96] J.F.C. Cocks et al. Acta Physica Polonica B27 (1996) p213

[97] J.F.C. Cocks et al. Phys. Rev. Lett. 78 (1997) p2920

[98] J.F.C. Cocks et al. J. Phys. G26 (2000) p23

[99] W. Urban et al. Phys. Lett. B274 (1990) p238

210

Page 214: nuc_expt_phr03

[100] J. Vermeer et al. Phys. Rev. C42 (1990) R1183

[101] J.F. Sharpey-Schafer and J. Simpson Prog. Part. Nucl. Phys. 21 (1988)

p293

[102] B. Herskind et al. Nucl. Phys. A447 (1985) p353c

[103] J.P. Martin et al. Nucl. Inst. Meth. A257 (1987) p301

[104] I.Y. Lee Nucl. Phys. A520 (1990) p641c

[105] J. Simpson et al. Heavy Ion Physics 11 (2000) p159

[106] S.M. Vincent, D.Phil University of Surrey, UK (1998)

[107] C. O’Leary private communication

[108] K.E.G. Lobner et al. Nuclear Data Tables A7 (1970) p495

[109] P.H. Regan et al. Phys. Rev. C51 (1995) p1745

[110] T. Chapuran et al. Nucl. Inst. Meth. Phys. Res. A272 (1988) p767

[111] F. Liden et al. Nucl. Inst. Meth. Phys. Res. A273 (1988) p240

[112] A. Galindo-Uribarri, Prog. Part. Nucl. Phys. 28 (1992) p463

[113] S. Mitarai et al. Nucl. Inst. Meth. A277 (1989) p491

[114] T. Kuroyanagi et al. Nucl. Inst. Meth. A316 (1992) p289

[115] Annerlan et al. Nucl. Instr. Meth. 22 (1963) p189

[116] J.B.A. England et al. Nucl. Instr. Meth. A280 (1989) p291

[117] Pausch et al. Nucl. Instr. Meth. A443 (2000) p304

[118] Pausch et al. Nucl. Inst. Meth. A365 (1995) p176

[119] Pausch et al. IEEE Trans. Nucl. Sci. 44 (1997) p1040

[120] E. Farnea et al. Nucl. Inst. Meth. A400 (1997) p87

[121] C. Chandler PhD Thesis, University of Surrey, UK (1999)

211

Page 215: nuc_expt_phr03

[122] J. Garces Narro PhD Thesis, University of Surrey, UK (2000)

[123] J.J. Simpson et al. Nucl. Phys. A287 (1977) p362

[124] D. Sarantites et al. Nucl. Inst. Meth. A381 (1996) p418

[125] J. Bialkowski et al. Nucl. Inst. Meth. A300 (1991) p303

[126] D. Seweryniak et al. Z. Phys. A345 (1993) p243

[127] A. Johnson et al. Nucl. Phys. A557 (1993) p401c

[128] C. Fahlander et al. Nucl. Phys. A577 (1994) p773

[129] M. Lipoglavsek et al. Z. Phys. A356 (1996) p239

[130] O. Skeppstedt et al. Nucl. Inst. Meth. A421 (1999) p531

[131] D. Seweryniak et al. Nucl. Inst. Meth. Phys. Res. A340 (1994) p353

[132] D.P. Balamuth Nucl. Inst. Meth. Phys. Res. A275 (1989) p315

[133] C.E. Svensson et al. Nucl. Inst. Meth. Phys. Res. A396 (1997) p228; erra-

tum A403 (1998) p57

[134] T. Davinson et al. Nucl. Inst. Meth. A288 (1990) p245

[135] K.R. Pohl et al. Phys. Rev. C49 (1994) p1372

[136] S. Arnell et al. Nucl. Inst. Meth. A300 (1991) p303

[137] W. Piel Jr. et al. Phys. Rev. C28 (1983) p209

[138] W. Gelletly, Acta Physica Polonica B26 (1995) p323

[139] P.J. Ennis et al. Nucl. Phys. A535 (1991) p392

[140] C.J. Lister et al. Phys. Rev. C42 (1990) R1191

[141] C.J. Lister et al. Phys. Rev. Letts. 59 (1987) p1270

[142] W. Gelletly et al. Phys. Lett. B253 (1991) 287

[143] A.N. James et al. Nucl. Inst. and Meth. A267 (1988) p144

212

Page 216: nuc_expt_phr03

[144] C.N. Davids et al. Nucl. Instr. and Meth., B70 (1992) p358

[145] C.J. Gross et al. Nucl. Inst. and Meth. A450 (2000) p12

[146] P. Spolare et al. Nucl. Inst. Meth. Phys. Res. A359 (1995) p500

[147] M. Leino et al, Nucl. Inst. Meth. B99 (1995) p653

[148] A. Ghiorso et al. Nucl. Inst. Meth. A269 (1988) p192

[149] B.J. Min et al. Nucl. Phys. A530 (1991) p211

[150] M. Weizflog et al. Z. Phys. A342 (1992) p257

[151] C.J. Gross et al. Nucl. Phys. A535 (1991) p203

[152] Ch. Winter et al. Nucl. Phys. A535 (1991) p137

[153] D. Rudolph et al J.Phys. G17 (1991) L113

[154] A.N.James et al. Nucl. Instr. and Meth., 212 (1983) 545.

[155] A.N.James, K.A.Connell and R.A.Cunningham, Nucl. Instr. and Meth.,

B53 (1991) 349

[156] E.S. Paul et al. Phys. Rev. C51 (1995) p78

[157] R.S. Simon et al. Z.Phys.A325 (1986) p197

[158] K. Helariutta et al. Phys. Rev. C54 (1996) R2799

[159] R.B. Taylor et al. Phys. Rev. C54 (1996) p2926

[160] M.P. Carpenter et al. Phys. Rev. Lett. 78 (1997) p3650

[161] D. Seweyniak et al. Phys. Rev. C55 (1997) R2137

[162] M. Rejmund et al. Acta. Phys. Pol. B27 (1996) p151

[163] K. Spohr et al. Acta. Phys. Pol. B26 (1995) p297

[164] R.‘Bark, J. Phys. G17 (1991) p1209

[165] P.J. Ennis and C.J. Lister, Nucl. Instr. and Meth. A313 (1992) 413

213

Page 217: nuc_expt_phr03

[166] D. LaFosse privrate communication

[167] D. LaFosse Phys. Rev. Lett. 78 (1997) p614

[168] P.J. Nolan and P.J. Twin, Ann. Rev. Nucl. Part. Sci. 38 (1988) p533

[169] P.J.Nolan, D.W.Gifford and P.J.Twin, Nucl. Instr. Meth. A236 (1985) 95.

[170] J.Simpson et al. Nucl. Instr. Meth. A269 (1988) 209.

[171] C.W. Beausang et al. Nucl. Instr. and Meth., A313 (1992) 37

[172] A.M. Baxter et al. Nucl. Instr. and Meth., A317 (1992) 101

[173] P.J. Nolan et al. Ann. Rev. Nucl. Part. Sci. 45 (1994) p561

[174] C.W.Beausang and J.Simpson, J.Phys. G 22 (1996) 527.

[175] R. Wyss Nucl. Inst. Meth. A256 (1987) p499

[176] G. Duchene et al. Nucl. Inst. Meth. A432 (1999) p90

[177] P.M. Jones et al. Nucl. Inst. Meth. A357 (1995) p458

[178] S.L. Shepherd et al. Nucl. Inst. Meth. A434 (1999) p373

[179] J. Eberth et al. Nucl. Inst. Meth. A369 (1996) p135

[180] M.A. Deleplanque et al. Nucl. Inst. Meth. A430 (1999) p292

[181] G.J. Schmidt et al. Nucl. Instr. Meth. A430 (1999) p69; erratum A434

(1999) p481

[182] K. Vetter et al. Nucl. Instr. and Meth. A452 (2000) p105

[183] J. van der Marel and B. Cederwall Nucl. Inst. Meth. A437 (1999) p538

[184] Th. Kroll and D. Bazzacco Nucl. Instr. Meth. A463 (2001) p227

[185] E.S. Paul et al. Nucl. Phys. A619 (1997) p177

[186] E.S. Paul et al. Phys. Rev. C59 (1999) p1984

[187] G.J. Schmid Nucl. Instr. Meth. A417 (1998) p95

214

Page 218: nuc_expt_phr03

[188] W.R. Leo, Techniques for Nuclear and Particle Physics Experiments,

Springer-Verlag, 1994, chapter 7

[189] E.S. Paul et al. J. Phys. G20 (1994) p751

[190] R.B. Firestone and V.S. Shirley Table and Isotopes, eighth edition, volume

II John Wiley and Sons, New York (1996) Appendix 1.

[191] J.M. Blatt and V.F. Wiesskopf Theoretical and Nuclear Physics, John Wiley

and Sons, New York ((1952) p627

[192] P.M. Endt At. Data. Nuc. Data. Tab. 55 (1993) p171

[193] P.M. Endt At. Data. Nuc. Data. Tab. 23 (1979) p547

[194] P.M. Endt At. Data. Nuc. Data. Tab. 26 (1981) p47

[195] M.J. Martin Nuc. Data. Sheets. 74, ix (1995)

[196] A.P. Byrne et al. Phys. Rev. C42 (1990) R6

[197] G.D. Dracoulis et al. Phys. Lett. B246 (1990) p31

[198] A.M. Baxter et al. Nucl. Phys. A515 (1990) p493 qu

[199] A.R. Poletti et al. Nucl. Phys. A442 (1985) p153

[200] G.D. Dracoulis et al. J. Phys. G17 (1991) p1795

[201] G.D. Dracoulis, Proceedings of the XXV Zakopane School on Physics, Se-

lected Topics in Nuclear Structure, (1990), vol.2, p3 (World Scientific),

Edited by J. Styczen and Z. Stachura

[202] P.J. Nolan and J.F. Sharpey-Schafer, Rep. Prog. Phys. 42 (1979) p1

[203] H. Bateman Proc. Camb. Phys. Soc. 15 (1910) p423

[204] D. Pelte andD. Schwalm, Hveay Ion Collisions edited by R. Bock (North-

Holland, Amsterdam, 1982) Vol.3 Chap.1

[205] J.F. Ziegler, Handbook of Stopping Cross-Sections For Energetic Ions in All

Elements (Pergamon, New York, 1980)

215

Page 219: nuc_expt_phr03

[206] A.E. Blaugrund, Nucl. Phys. 88 (1966) p501

[207] W.M. Currie, Nucl Inst. Meth. 73 (1969) p173

[208] J. Bacelar et al. Phys. Rev. Lett. 57 (1986) p3019

[209] J. Bacelar et al. Phys. Rev. C35 (1987) p1170

[210] R.M. Clark and N. Rowley J. Phys. 18 (1992) p1515

[211] P. Tikkanen et al. Phys. Rev. C42 (1990) p2431

[212] J.R. Hughes et al. Phys.Rev. Lett. 72 91994) p824

[213] H.-Q. Jin et al. Phys. Rev. Lett. 75 (1995) p1471

[214] B. Cederwall et al. Nucl. Inst. Meth. Phys. Res. A354 (1995) p591

[215] A.G. Smith et al. Phys. Rev. Lett. 73 (1994) p2540

[216] R.M. Clark private communication.

[217] S.A. Forbes et al. Z. Phys. A352 (1995) p15

[218] R. Krucken et al. Nucl. Phys. A589 (1995) p475

[219] R. Krucken et al. Phys. Rev. C55 (1997) R1625

[220] D. Zainea et al. Z. Phys. A352 (1995) p365

[221] P. Petkov et al. Nucl. Phys. A568 (1994) p572

[222] A. Dewald et al. J. Phys. G19 (1993) L177

[223] N.V. Zamfir et al. Z. Phys. A344 (1992) p21

[224] R. Kuhn et al. Phys. Rev. C55 (1997) R1002

[225] P. Willsau et al. Z. Phys. A355 (1996) p129

[226] R.M. Clark et al. Phys. Rev. C50 (1994) p84

[227] G.D. Dracoulis et al. Phys. Rev. C29 (1984) p1576

[228] I. Thorslund et al. Nucl. Phys. A568 (1994) p306

216

Page 220: nuc_expt_phr03

[229] M. Piiparinen et al. Nucl. Phys. A565 (1993) p671

[230] A. Dewald et al. Nucl. Phys. A545 (1992) p822

[231] A. Dewald Z. Phys. A334 (1989) p163

[232] G. Bohm et al. Nucl. Inst. Meth. Phys. Res. A329 (1993) p248

[233] J.M. Reid The Atomic Nucleus (1986) Manchester University Press

[234] T. Lonnroth et al. Z. Phys. A317 (1984) p215

[235] E. Recknagel, Nuclear Spectroscopy and Reactions, Part C, ed. J. Cerny

(Academic Press, New York, 1974) p. 93

[236] P.H. Regan et al. Nucl. Phys. A591 (1995) p533

[237] S. Harissopulos et al. Phys. Rev. C52 (1995) p1796

[238] J. Billowes et al. Phys. Lett. B178 (1987) p145

[239] M. Weiszflog et al. Nucl. Phys. A584 (1995) p133

[240] M. Weiszflog et al. J. Phys. G20 (1994) L77

[241] E. Lubkiewicz et al. Z. Phys. A335 (1990) p369

[242] U. Birkental et al. Nucl. Phys. A555 (1993) p643

[243] O. Hausser et al. Nucl. Phys. A412 (1984) p141

[244] O. Hausser Phys. Lett. B144 (1984) p341

[245] A.E. Stuchbery et al. Phys. Rev. Lett. 76 (1996) p2246

[246] A.P. Byrne et al. Nucl. Phys. A567 (1994) p445

[247] H. Bertschat et al. Nucl. Phys. A222 (1974) p399

[248] D.A. Volkov et al. Soviet Journal of Physics 44 (1986) p547

[249] D.P. Balamuth et al. Phys. Rev. C48 (1993) p2648

[250] D.R. Haenni et al. Phys. Rev. C33 (1986) p1543

217

Page 221: nuc_expt_phr03

[251] A. Savelius et al. Acta. Phys. Pol. B28 (1997) p173

[252] M.W. Guidry et al. Phys. Lett. B163 (1985) p79

[253] R. Bock et al. Nucl. Phys. A388 (1982) p334

[254] H. Takai et al. Phys. Rev C38 (1988) p1247

[255] R. Broda et al. Phys. Lett. 251 (1990) p245

[256] R. Broda et al. Phys. Rev. Lett. 74 (1995) p868

[257] B. Fornal et al. Phys. Rev. C49 (1995) p2413.

[258] B. Fornal et al. Acta Physica Polonica B26 (1995) p357

[259] P.H. Regan et al. Phys. Rev. C55 (1997) p2305

[260] H. Freiesleben and J.V. Kratz, Phys. Rep. 106 (1984) p1

[261] L. Corradi et al. J. Phys. G23 (1997) p1485

[262] L. Corradi et al. Phys. Rev. C59 (1999) p261

[263] L. Corradi et al. Phys. Rev. C61 (2000) 024609

[264] I.Y. Lee, et al. Acta Physica Polonica B28 (1997) p257

[265] S. Juutinen et al. Phys. Lett. 386B (1996) p80

[266] C. Wheldon et al. Phys. Lett. B425 (1998) p239

[267] S.J. Asztalos et al. Phys. Rev. C61 (2000)14602

[268] S.J. Asztalos et al. Phys. Rev. C60 (1999) 044307

[269] I.Y. Lee et al. Phys. Rev. 56 (1997) p753

[270] J.F.C. Cocks et al. J. Phys. G26 (2000) p23 ; Nucl. Phys. A645 (1999)

p61

[271] N. Amzal et al. J. Phys. G25 (1999) p831

[272] J. Wilson et al. Eur. Phys. J. A9 (2000) p183

218

Page 222: nuc_expt_phr03

[273] , Introductory Nuclear Physics P.E. Hodgson, E. Gadioli and E. Gadioli

Erba, Oxford Science Pulblications, (2000)

[274] R. Bass, in Springer Lecture Notes: Heavy Ion Reactions.

[275] S. Schwebel private communication

[276] M.W. Simon et al. Nucl. Instr. Meth. A452 (2000) p205

[277] M.W. Simon et al. Proc. Int. Conf. on Fission and Properties of Neutron-

Rich Nuclei, Sanibel Island, Florida 1997, eds. J.H. Hamilton and A.V. Ra-

mayya, World Scientific (1998) p270

[278] D. Cline, Acta Physica Polonica B30 (1999) 1291;

[279] C.Y. Wu et al. Phys. Rev. C61 (2000) 021305

[280] I. Hibbert, Gamma-rays from Deep-Inelastic Reactions, PhD Thesis, Uni-

versity of Manchester (1993)

[281] A. Yamamoto, final year project, University of Surrey, (2000) unpublished.

[282] R.D. Page private communication

[283] H. Geissel, G.Munzenberg and K. Riisager, Ann. Rev. Nucl. Part. Sci. 45

(1995) p163

[284] D. Morrisey, Nucl. Phys. A616 (1997) p45c

[285] J.D. Garrett, Nucl. Phys. A616 (1997) p3c

[286] S. Kubono et al. Nucl. Phys. A616 (1997) p21c

[287] P.H. Regan et al. Acta Physica Polonica B28 (1997) p431

[288] R. Grzywacz et al. Phys. Lett. 355B (1995) p439

[289] K. Rykaczewski et al. Phys. Rev. bf C52 (1995) R2310

[290] R. Grzywacz et al. Phys. Rev C55 (1997) p1126

[291] C. Chandler et al. Phys. Rev. C61 (2000) 044309

219

Page 223: nuc_expt_phr03

[292] C.J. Pearson et al. Phys. Rev. Lett. 79 (1997) p605

[293] A.C. Mueller and R. Anne, Nucl. Inst. Meth. Phys. Res. B56/57, 559

(1991)

[294] A.C. Mueller and B. Sheryl, Ann. Rev. Nucl. Part. Sci. 43, 529 (1993)

[295] R. Grzywacz private communication.

[296] C. Chandler et al. Phys. Rev C56 (1997) R2924; 61 (2000) 044309

[297] J.M. Daugas et al. Bormio Conference Proceedings, (1999),

[298] M. de Jong, A.V. Ignatyuk and K.-H. Schmidt, Nucl. Phys. A613, 435

(1997)

[299] J.-J. Gaimard and K.-H. Schmidt Nucl. Phys. A 531 (1991) p709

[300] A.V. Ignatyuk, Statistical Properties of Excited Nuclei (Energoatomizdat,

Moscow, 1983) [Russian].

[301] K.-H. Schmidt et al., Phys. Lett. B 300, 313 (1993)

[302] I.S. Grant and M. Rathle, J. Phys. G5 (1979) p1741 and references therein.

[303] H. Naik et al. Nucl. Phys. A 648 (1999) p45

[304] M. Pfutzner, Acta. Phys. Pol. 28 (1997) p289

[305] T. Glasmacher et al. Phys. Lett. B395 (1997) p163

[306] T. Glasmacher Ann. Rev. Nucl. Part. Sci. 48 (1998) p1

[307] H. Geissel et al. Nucl. Inst. Meth. B70 (1992) p286

[308] M. Pfutzner et al., Phys. Lett. B444 (1998) p32

[309] M. de Yong et al., Nucl. Phys. A628 (1998) 479.

[310] M. de Jong et al., Nucl Phys. A613 (1997) 435.

[311] J. Benlliure et al. Nucl. Phys. A660 (1999) p87

[312] Zs. Podolyak et al. Phys. Lett. B491 (2000) p225

220

Page 224: nuc_expt_phr03

[313] Z. Podolyak et al. Proceedings of the International Conference on Neu-

tron Rich Nuclei and Fission, St. Andrews, Scotland, (1999) p156. Eds.

J.H. Hamilton, W.R. Phillips and H.K. Carter, World Scientific, London

[314] M. Pfutzner et al.Proceedings of the International Conference on Exper-

imental Nuclear Physics in Europe, Sevilla, Spain 1999, eds. B. Rubio,

M. Lozano and W. Gelletly, AIP Conf. Proc. 494 p113-116

[315] Ch. Schlegel et al. Physica Scripta T88 (2000) p72

[316] M. Cammano et al. Nucl. Phys. A682 (2001) p175

[317] M. de Jong, Nucl. Phys. A613 (1997) p435

[318] M. Bernas et al. Nucl. Phys. A616 (1997) p352c

[319] C. Longour et al. Phys. Rev. Lett. 81 (1998) p3337

[320] J. Garces Narro et al. Phys. Rev. C63 (2000) 044307

[321] Z. Janas et al. Phys. Rev. Lett. 82 (1999) p295

[322] J.M. Daugas et al. Phys. Rev. C63 (2001) 064609

[323] M. Pfutzner et al. in press Phys. Rev. C (June 2001).

[324] M. Pfutzner et al. Acta Physica Polonica B32 (2001) p2507

[325] G. Georgiev et al. Physics of Atomic Nuclei 64 (2001) p1181: Yad. Fiz. 64

(2001) p1258

[326] K. Yoneda et al. Phys. Lett. B499 (2001) p233

[327] J. Gerl Acta Phys. Pol. 32 (2001) p1379; S. Wan et al. Z. Phys. A358

(1997) p213

[328] M. Belleguic et al. Nucl. Phys. A682 (2001) p136; Physica Scripta T88

(2000) p122

[329] S. Al Garni et al. Proceedings of the International Nuclear Physics Confer-

ences, Berkeley, USA (2001), in press.

221

Page 225: nuc_expt_phr03

[330] J.H. Hamilton Rep. Prog. Phys. 45 (1985) p632; and Probing Nuclei Far

From Stability with Heavy Ions, Chapter 4, in Heavy Ion Collisions, volume

3, edited by R. Bock, North-Holland Publishing (1982)

[331] R. Kirchner Nucl. Instr. Meth. 186 (1981) p295

[332] R. Kirchner Nucl. Instr. Meth. 186 (1981) p275

[333] K. Schmidt et al. Eur. Phys. J. A8 (2000) p303

[334] M. Oinonen et al. Eur. Phys. J. A5 (1999) p151

[335] M. Karny et al. Nucl. Phys. A640 (1998) p3

[336] M. Ramdhane et al. Phys. Lett. 432B (1998) p22

[337] Z. Janas et al. Nucl. Phys. A527 (1997) p119

[338] K. Schmidt et al. Nucl. Phys. A624 (1997) p185

[339] Z. Janas Phys. Scr. T56 (1995) p262

[340] K. Rykaczewski et al. Nucl. Phys. A499 (1989) p529

[341] R. Kirchner et al. Nucl. Phys. A378 (1982) p549

[342] R. Kirchner et al. Nucl. Instr. Meth. 234 (1985) p224

[343] E. Runte Nucl. Phys. A399 (1983) p163

[344] J. Eidens et al. Nucl. Phys. A141 (1970) p289

[345] W.N. Catford et al. Nucl. Inst. Meth. Phys. Res. A371 (1996) p449

222


Recommended