You are on page 1of 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 p.regan@surrey.ac.uk

October 2003

Contents
1 Electromagnetic Probes of Nuclear Structure. 1.1 1.2 Gamma-Ray Decay Selection Rules. . . . . . . . . . . . . . . . . . Internal Conversion Electrons. . . . . . . . . . . . . . . . . . . . . 1.2.1 Electric Monopole Decays. . . . . . . . . . . . . . . . . . . 1.2.2 Magnetic Monopoles. . . . . . . . . . . . . . . . . . . . . . 2 3 4 7 9 10 10 11 12 16 17 19 20 23 24 25 29 32 35 39 39 41 41 44

2 Studies of Nuclear Structure at High Angular Momentum. 2.1 Fusion Evaporation Reactions. . . . . . . . . . . . . . . . . . . . . 2.1.1 2.1.2 2.1.3 2.1.4 2.1.5 2.1.6 2.2 Beam Currents, Energies and Target Thicknesses. . . . . . Compound Nucleus Excitation Energy and Maximum Angular Momentum. . . . . . . . . . . . . . . . . . . . . . . . Compound Nucleus Decay. . . . . . . . . . . . . . . . . . . Excitation Functions. . . . . . . . . . . . . . . . . . . . . . Spin Assignments: Gamma-ray Angular Distributions. . . Anisotropies and Gated Angular Distributions. . . . . . . .

2.1.7 DCO Ratios. . . . . . . . . . . . . . . . . . . . . . . . . . Determining the Intrinsic Structure of Rotational Bands. . . . . . 2.2.1 2.2.2 Rotational Frequency, Moments of Inertia and Alignments. Particle-Core Coupling. . . . . . . . . . . . . . . . . . . . .

2.3

2.2.3 Branching Ratios and g-Factors. . . . . . . . . . . . . . . . 2.2.4 Two State Mixing. . . . . . . . . . . . . . . . . . . . . . . Selected Topics in High Spin Nuclear Structure. . . . . . . . . . . 2.3.1 2.3.2 2.3.3 2.3.4 Shape Coexistence and Superdeformation. . . . . . . . . . Band Terminations. . . . . . . . . . . . . . . . . . . . . . . High K-Isomers and Pairing Reduction. . . . . . . . . . . Octupole Correlations. . . . . . . . . . . . . . . . . . . . .

2.4

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

45 48 48 49 49 51 52 52 54 56 56 56 63 63 65 66 69 75 78

3 Experimental Gamma-ray Spectroscopy. 3.1 Germanium Semi-Conductor Detectors. . . . . . . . . . . . . . . . 3.2 Gamma-Ray Spectroscopy with Germanium Detectors. . . . . . . 3.2.1 Response Function of Germanium Spectra. . . . . . . . . . 3.2.2 Germanium Detector Eciency. . . . . . . . . . . . . . . . 3.2.3 The Compton Suppressed Spectrometer (CSS). . . . . . . Gamma-Ray Arrays. . . . . . . . . . . . . . . . . . . . . . . . . . 3.3.1 3.3.2 3.3.3 3.3.4 Resolving Power and Total Photopeak Eciency. . . . . . Add-Backs from Clover/Cluster Detectors. . . . . . . . . . Polarization Measurements. . . . . . . . . . . . . . . . . . Gamma-ray Tracking. . . . . . . . . . . . . . . . . . . . .

3.3

4 Channel Selection In Fusion-Evaporation Reactions. 4.1 4.2 Inner Multiplicity Sum-Energy Balls. . . . . . . . . . . . . . . . . Studies of Very Neutron Decient Nuclei. . . . . . . . . . . . . . . 4.2.1 4.2.2 4.2.3 4.2.4 4.3 4.4 Charged Particle Balls. . . . . . . . . . . . . . . . . . . . . Kinematic Focussing and Conversion from Lab to COM Energies. . . . . . . . . . . . . . . . . . . . . . . . . . . . . Silicon Detectors. . . . . . . . . . . . . . . . . . . . . . . . CsI(Tl) Balls Using Pulse Shape Discrimination. . . . . . .

Neutron Detection. . . . . . . . . . . . . . . . . . . . . . . . . . . 91 Recoil Detectors. . . . . . . . . . . . . . . . . . . . . . . . . . . . 100 4.4.1 4.4.2 4.4.3 4.4.4 Recoil Mass Separators. . . . . . . . . . . . . . . . . . . . 100 Gas Filled Separators. . . . . . . . . . . . . . . . . . . . . 107 Recoil Decay Tagging. . . . . . . . . . . . . . . . . . . . . 108 Recoil Filter Detectors. . . . . . . . . . . . . . . . . . . . . 109

5 Measurement of Lifetimes of Bound Nuclear States. 113 5.0.5 Weisskopf Single Particle Estimates. . . . . . . . . . . . . 114 5.0.6 5.1 Determining Nuclear Quadrupole Deformation from Lifetimes of E2 Transitions. . . . . . . . . . . . . . . . . . . . 115

Electronic Timing Methods. . . . . . . . . . . . . . . . . . . . . . 116 5.1.1 Gamma-ray Spectroscopy Across Isomers . . . . . . . . . . 120

ii

5.2

The Recoil Distance Method. . . . . . . . . . . . . . . . . . . . . 124 5.2.1 Feeding Corrections and Gating From Above. . . . . . . . 128 5.2.2 The Dierential Decay Curve Method. . . . . . . . . . . . 133 The Doppler Shift Attentuation Method. . . . . . . . . . . . . . . 136 5.3.1 Lineshape Analysis. . . . . . . . . . . . . . . . . . . . . . . 139

5.3

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 Dis6.1.2 6.1.3 6.1.4 tribution Technique. . . . . . . . . . . . . . . . . . . . . . 146 Analysis of Precession Data with Limited Angles. . . . . . 148 Transient Field Measurements. . . . . . . . . . . . . . . . 149 Time Dierential Perturbed Angular Distributions. . . . . 150 153

7 Spectroscopy of Neutron Rich Nuclei. 7.1 7.2 7.3

Using Fusion Evaporation Reactions . . . . . . . . . . . . . . . . 153 Incomplete Fusion/Massive Transfer Reactions. . . . . . . . . . . 154 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 8.1.2 8.1.3 8.1.4 8.1.5 8.1.6 8.1.7 8.2 Projectile Fragmentation. . . . . . . . . . . . . . . . . . . 171 Particle Identication in Fragmentation. . . . . . . . . . . 173 Isomeric Ratios and Angular Momentum Population. . . . 184 Projectile Fission . . . . . . . . . . . . . . . . . . . . . . . 188 Intermediate Energy Coulex. . . . . . . . . . . . . . . . . . 189 Double Fragmentation and In-beam spectroscopy . . . . . 190 Beta-Decay Measurements. . . . . . . . . . . . . . . . . . . 190

ISOL Based Techniques. . . . . . . . . . . . . . . . . . . . . . . . 190 8.2.1 On-Line Mass Separators. . . . . . . . . . . . . . . . . . . 191 8.2.2 The
19

Ne +40 Ca Experiment at Louvain La Neuve. . . . . 198

Chapter 1 Electromagnetic Probes of Nuclear 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 electromagnetic interactions and can be investigated by detecting the emitted electromagnetic radiation as the state de-excites, ie. through the emission of either gamma-rays or conversion electrons. While the strong force is the dominant interaction 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 angular momentum and exotic proton to neutron ratios. We will investigate methods of producing exotic nuclear matter in-beam and dierent methods of selecting specic 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 determination of decay probablilities of nuclear states (ie. their lifetimes) gives direct information on the make-up of the initial and nal states and can reveal highly collective, deformed structures within the nucleus.

1.1

Gamma-Ray Decay Selection Rules.

For a gamma-ray decay between states of initial spin Ii and nal state spin If , 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 1 , gamma-ray transitions h 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).

(a)

+ ("mixed") M1, E2 M3, E4, M5, E6, M7, E8, M9 4 ("mixed") M1, E2, +

(pure) E2 M3, E4, M5, E6, M7, E8

M3, E4, M5, E6, M7

(b)

+ E1, (M2) E2

M3,....,E6 + 2

E3, M4, E5, M6, E7 E1, (M2) E3, M4, E5

Figure 1.1: Schematic decay scheme showing the eect of gamma-ray selection rules on allowed multipolarities. The parity of a magnetic transition of multipole order L is given by = 3

(1)L+1 , while that of an electric transition is given by = (1)L . The transition probability for a state decaying from state Ji to state Jf , separated by energy E , by a transition of multipole order L is given by [1, 7] Tf i (L) = 8(L + 1) hL ((2L + 1)!!)2 E hc
2L+1

B(L : Ji Jf )

(1.1.2)

where B(L : Ji Jf ) is called the reduced matrix element. As gure 1.1 shows schematically, typically, the lowest multipolarity transitions dominate the decays. This is an eect of the diering transition probabilties for dierent multipoles (see chapter on measuring nuclear lifetimes). The transition probability for a mixed multipolarity transition (usually restricted to M1/E2 decays for in-beam decays) can be calculated in terms of the multipole mixing ratio, [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]
2 E2/M 1 =

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] 1 E5 I (I = 1) B(M1) = 0.697 2 3 2 B(E2) E1 1 + E2/M 1 I (I = 2) (1.1.4)

T (E2 : J > J 1) T (M1 : J > J 1)

(1.1.3)

where I are the experimentally measured gamma-ray intensities for the competing transitions. (Note the I = 1 intensity contains both M1 and E2 admxitures).

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 eld 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 dierence 4

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 = E B.E. (1.2.5)

where E is the energy of the competing gamma decay and B.E. is the electron binding energy.
442.7

202Po 800 x102

gammas delayed (35 - 550 ns)


571.2

400
385.7 526.2

676.8

counts

0 442.7 electrons

571.2 4000 676.8

2000 385.7 526.2

0 400 600 800 transition energy (K) in Po [keV] 1000

Figure 1.2: Experimental internal conversion and gamma-ray spectra for transitions in 202 Po. The 386, 443, 571 and 677 are E2 decays, the 526 is an E1 and the 912 is an E3. Figure 1.2 shows a comparison of electron and gamma-ray spectra for decays in Po [8]. Note that while the gamma-ray transitions are single lines, the electrons come in groups. This reects the fact that the electrons come out with dierent 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
202

912.1

912.1

Experimental ICCs (T. Kibedi et al., Australian National University) 202Po 100

ICC [K shell]

10-1

M2 10-2 M1 E3 E2 E1 10-3 200 400 600 800 Transition energy [keV] 1000 1200

Figure 1.3: Experimental internal conversion coecients for transitions in 202 Po. These data were taken from the reaction 194 Pt(12 C,xn)202 Po at a beam energy of 76MeV [8].

I = I +
i

Iec = I 1 +
i

(1.2.6)

where i is the internal conversion coecient 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 coecients are very useful as they are dependent on the multipolarity of the transition [10, 11] and can thus give information on the spin and parity of nuclear states. Transition multipolarities can be assigned by either measuring the absolute internal conversion coecients i = or the ratio of the partial conversion coecients (eg. M ). L
e

In the case of mixed multipolarity I=1 transitions, the experimentally determined value of the electron conversion coecient directly gives the magntitude of the E2/M1 mixing ratio. Figure 1.3 shows the experimentally determined internal conversion coecients for 202 Po determined from the spectra shown in gure 1.2. The size of the electron conversion coecient 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

emitted between the beam bursts (see later). The emission of a conversion electron results in an electron vacancy being lled 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 coecients for electric (E) and magnetic (M) multipoles can be calculated using the following expressions [5], Z3 (EL) 3 n L L+1 e2 4ohc
4 4

2me c2 E
3 L+ 2

5 L+ 2

(1.2.7)

Z3 (ML) 3 n

e2 4o hc

2me c2 E

(1.2.8)

Note that there has been a recent report of the rst 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 (E0) = 1 = 2 0+ 0+ i f j (Z, k)
j

(1.2.9)

where is the (partial) lifetime for the E0 decay, is the (dimensionless) monopole strength parameter and j are the electronic factors (analagous to internal conversion coecients) [13]. The electronic factors are tabulated in reference [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 dened by [12] = < 0+ | f
2 ej rj |0+ > < 0+ |m(E0)|0+ > i i f = eR2 eR2 j
1

(1.2.10)

where R is the nuclear radius (1.2A 3 fm) and m(E0) is the electric monopole operator.

Figure 1.4: Out of beam decay spectra for the reaction 144 Sm(33 S,p2n)174 Ir at 153 MeV. Non-yrast E0 decays are observed in 174 Os which was populated via the beta-decay of 174 Ir. The beam was incident on a 1.3 mg/cm2 target and irradiated for 4 seconds followed by a 4 sec measuring cycles. The lines in 174 Os correspond to the following decays: 546 keV, 0+ 0+, pure E0 ; 532 keV, 2+ 2+ , mixed E2+M1+E0; and 555 kev 4+ 4+, mixed E2+M1+E0. Note the absence of a 532 keV line in the gamma-ray spectrum [16]. The single particle units for E0 decays are given by [17], 2 = 0.5A 3 sp
2

(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+ congurations [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 dierent nuclear states with dierent radii (ie. dierent deformation). The electric 8

monopole operator can be expanded [18, 296] in terms of the quadrupole and triaxial deformation parameters and respectively such that 3Z M(E0) = 4 4 5( 5) 2 3 cos + + 5 21 (1.2.12)

In the limit of simple two-state mixing between congurations with deformations 1 , 1 and 2 , 2 , if a is the mixing amplitude between the congurations, the resulting monopole strength is given by
2

3Z (E0) = 4
2

a
+

1a
+

2 1

2 2

5 5 + 21

2 3 1

cos 1

3 2 cos 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 2 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 rst vanishes for equal deformations of opposite sign. Note that E0 decays also occur between nuclear states with the same spin and parity (ie. J J ), 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].

Chapter 2 Studies of Nuclear Structure at High Angular Momentum.


In order to fully understand the physics of the nucleus, one needs to examine the eect of extreme conditions, (such as high temperature/excitation energy, distortions 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 dierent 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 dierent 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 excitation 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

inelastic scattering (Coulex) fusion R b DIC

elastic (Rutherford) scattering


Figure 2.1: Various types of heavy-ion collisions as a function of impact parameter. 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 gure 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 (1028 m2 ) 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 ) 3 f m2 ). This fusion cross-section drops dramatically for heavier nuclei where ssion 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 elec11
1

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 reduces 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 1067 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-)

Fusion ~ 1mb I~1 pnA beam


1111 0000 1111 0000 1111 0000 1111 0000 1111 0000 1111 0000 1111 0000 1111 0000 1111 0000 1111 0000 1111 0000

vaccum vessel ~ E-7 Torr

1 0 1 0 11111 00000 11111 00000 11111 00000 11111 00000 11111 00000 1111 0000 11111 00000 1111 0000 11111 1111 0000 11 00000 00 00000 11 00000 thin 00 11111 Pb/Au 11111 11 00 11 11111 00 00000 11 11111 00 00000 target 11111 00000 11111 00000 11111 00000 11111 00000 11111 00000 11111 00000 11111 00000 11111 00000 11111 00000 11111 00000 11111 00000 11111 00000 11111 00000 11111 00000 11111 00000 11111 00000

outgoing beam stopper

1 0 1 0 1 0 1 0 1 0 1 0 1 0 1 0 1 0 1 0 1 0 1111 0000 1 0 1111 0000 1 0 1111 0000 1 0 1111 0000 1 0 1111 0000 1 0 1111 0000 1111 0000 1111 0000 1111 0000 1111 0000 1111 0000 1111 0000 1111 0000 1111 0000 1111 0000 1111 0000 1111 0000

beam stop/ Farady cup

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

2.1.2

Compound Nucleus Excitation Energy and Maximum Angular Momentum.

For fusion to occur the beam nuclei must have sucient kinetic energy to overcome the Coulomb repulsion between the two positively charged nuclei. Fusionevaporation reactions require the formation of a compound nucleus. This describes a hot nuclear system which lives long enough (> 1020 s) for thermodynamic equilibrium to occur, during which time the compound system loses its 12

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

Eex = Ecm + Qf us

(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 MB VB = (MT + MB ) VR and by conservation of energy, 1 2 (MT + MB ) VR 2 2 substituting in for VR , and recalling that EB = 1 MB VB , we obtain 2 Ecm = EB Ecm = EB 1 (2.1.4) (2.1.3)

MB (2.1.5) MT + MB 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-o approximation which assumes the nuclei are hard spheres without a diuse surface). The fusion cross-section will be a sum of partial waves (depending on the size of the impact parameter). In the sharp cut o approximation, the assumption is that the transmission coecient 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, f can be written as a sum of partial waves upto lmax such that [1] f = 2
2 lmax l=0

(2l + 1)Tl

(lmax + 1)2

(2.1.6)

where is the wavelength of the entrance channel given by h = 2 2Ecm 14 (2.1.7)

Ecm is the kinetic energy of the collision in the centre of mass and is the A reduced mass of the system such that = ABB ATT where AB and AT are the +A masses of the beam and target nuclei respectively. The value of lmax , calculated using the reduced mass of the system, =
MT MB , MT +MB

and from conservation of energy and angular momentum is given by [1] hlmax = vR (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 2 v = Ecm Vc 2 Substituting in for v we obtain,
2 lmax =

(2.1.9)

2R2 (Ecm Vc ) (2.1.10) h2 where R is the maximum nucleus-nucleus distance for which a reaction can occur and is given (in fm) empirically by [1]
1 1 3 3 R = 1.36 AB + AT + 0.5

(2.1.11)

The Coulomb barrier energy Vc (in MeV) is given by ZB ZT (2.1.12) R 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. Vc = 1.44 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 parameters (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

1 lcr + 2

RB RT ZB ZT e2 (RT + RB )3 4 = RT + RB h2 (RT + RB )2

(2.1.13)

where 0.9 MeV fm2 is the surface tension of the nucleus.

2.1.3

Compound Nucleus Decay.

It typically takes around 102122 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 1020 second, after which the compound system decays by either high energy gamma-ray emission (such as giant resonance decays) and/or by nucleon evaporation, where neutrons, protons and -particles are emitted [1, 21]. Due to the eect 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 compound nucleus moves further to the neutron decient 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.
compound nucleus E-20 secs 1
11 00 11 00 11 00

Excitation energy

nucleon separation energy above yrast

11 00 11 00 11 00 11 00 11 00 11 00

E-15s

11 00 11 00 11 00

1111 0000 1111 0000 1111 0000 1111 0000 1111 0000 1111 0000 1111 0000 1111 0000 1111 0000 gammas 1111 0000

in residual

nuclei

111111111 000000000 yrast 111111111 000000000 111111111 000000000

line (locus of yrast states)

E-9 sec

Angular momentum

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

16

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 excitation energy of the compound system by around 5-8 MeV per nucleon, yet only remove 12 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 1015 seconds for the compound nucleus to decay into the residual nucleus. Note that as gure 2.3 shows, the nal nucleus created is determined by the entry point (relative to the yrast line) in the excitation energy/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 eect can be used to experimentally select transitions from specic evaporation channels using total energy/gamma-ray multiplicity detectors. The probability for a compound nucleus evaporating a particle (usually meaning a proton, neutron or alpha particle) is proportional to the density of nal states and a barrier (Coulomb usually) transmission coecient, given by [1] T (li, Ep (i)) = exp

where V is the height of the barrier and is its width. Note that the shape of the spectrum for the emitted particles is dierent for charged particles (protons and alphas) compared to neutrons due to the eect of the Coulomb barrier. In neutron decient 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.

(2mp (V Ep )) 2

2 h

(2.1.14)

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

10

523 keV
5

10

4n ( Cd)

110

335 keV

(arb. units)

10

5n ( Cd)

109

798 keV 249 keV 191 keV

10

125 keV p3n ( Ag)


110

10

p4n ( Ag)

109

10

50

60

70 EBEAM (MeV)

80

90

Figure 2.4: Excitation function for various products of the reaction [22].

18

O+96 Zr

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 transitions from nuclei of interest in fusion of an 18 O beam on a 96 Zr target (forming the compound nucleus 114 Cd. Note how the relative intensity of the higher multiplicity channels such as the 5n to 109 Cd and p4n to 109 Ag increase with increasing beam energy in the region of this excitation function, while the four particle out channels to 110 Cd and 110 Ag peak at a beam energy of around 60 MeV, before falling o. 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

2.1.5

Spin Assignments: Gamma-ray Angular Distributions.

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 specic orientation in space. This is achieved in fusionevaporation reactions, where the angular momentum vector (l = r p) is in a direction (to a good approximation) perpendicular to the beam direction (see gure 2.5).

substate alignment after evaporation

reaction plane

beam direction

target

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 eect 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

W () =
k

Ak Pk (cos)

(2.1.15)

where W () is the gamma-ray intensity measured at angle 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 (cos) are the standard Legendre polynomials; and Ak is the angular distribution coecient. The Ak value depends on the substate or m-population distribution and the values of the initial and nal state spins [23]. By measuring the intensity of a gamma-ray transition as a function of detector angle about the beam direction, a full angular distribution can be obtained from which the values of A2 and A4 can be obtained by tting the distribution to equation 2.1.15. The Ak values can be used to experimentally distinguish between transitions of dierent multipolarities [23, 24]. Similarly, tting the experimentally observed A2 and A4 coecients 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 () = A0 {1 + A2 P2 (cos)} (2.1.16)

1 where P2 (cos) = 2 (3cos2 1) and A0 is the true intensity. For a quadrupole (I=2) transition (E2), the distribution will have the form,

W () = A0 {1 + A2 P2 (cos) + A4 P4 (cos)} where P4 (cos) =


1 8

(2.1.17)

(35cos4 30cos2 + 3)

2.1.6

Anisotropies and Gated Angular Distributions.

The xed (and often limited) angular granularity of modern gamma-ray arrays coupled to the complexity of the singles spectra often observed in fusionevaporation reactions mean that a full angular distribution analysis may not be viable. It is often enough to be able to tell the dierence in the anisotropy between 20

I=1 and I = 2 type transitions. In the case where only a few gamma-ray detector angles are used, the A4 coecient is set to a value of zero (eg. [25]). Alternatively, coincidence data can be used to form gated singles spectra at specic detector angles, from which an anisotropy can be taken (eg [22]). Pohl et al. dened 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 8 gamma-ray array.

1.0

0.5

0.0

Anisotropy
-0.5 I = 0 I = -1 I = -2 -1.0 -1.5

-90

-60

-30 0 30 arctan(E2/M1)

60

90

Figure 2.6: Theoretical values for various transitions for the gamma-ray anisotropy as dened in reference [22]. Figure 2.6 shows the theoretically expected value for this anistropy for various 21

multipolarity decays (assuming A4 = 0). Note the dierent 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 108 Pd and 109 Ag. The dierence between pure I = 2 (E2) and pure I = 1 (E1) transitions 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 multipolarities 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 108 Pd and 109 Ag. Note the clear separation between I = 2 and I = 1 transitions [22]. Also, note the lower values of for stretched E2 transition in 108 Pd resulting from a larger destruction of reaction alignment associated with the emission of an -particle.

22

2.1.7

DCO Ratios.

For coincidence measurements, the Directional Correlations from Oriented states (DCO) method can be used to infer the spin dierences between states observed by the measurement of the gamma-decay between them [26, 27, 28]. Fusionevaporation 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 gammaray detector angles are usually xed 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 discriminate between transitions of diering multipolarities. By measuring coincidence data from detectors are dierent angles, once can construct sets of angle gated gamma-gamma coincidence matrices, in oine sorting. These 2 dimensional spectra 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 multipolarities 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 61 Cu and projects transitions of known E2 and E1 multipolarity. Note the dierence 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 dierent multipolarities. Figure 2.9 shows the results of two dierent types of DCO ratios (dierent detector angles used) for 106 Cd, formed using two dierent fusion-evaporation reactions. Note that in both cases, there is a clear discrimination possible between the dierent types of transition.

23

Figure 2.8: Angle gated DCO correlation spectra showing the dierence between E2 and E1 transitions in 61 Cu from the reaction 24 Mg+40 Ca3p+61 Cu (AYEBALL data [106]).

2.2

Determining the Intrinsic Structure of Rotational Bands.

Using coincidence relationships and information from gamma-ray anisotropies and DCO ratios, one can come up with a (complicated) decay scheme with many dierent 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 observed, 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 experimental information obtained on nuclear states (ie. their spins and excitation 24

1.5

J=1 J=2 J->J

1.0 R 0.5

0.0

J=1 J=2 J->J

1.5 R 1.0

0.5 0 200 400 600 800 1000 Gamma-ray Energy (keV) 1200

Figure 2.9: DCO ratios for transitions in 106 Cd from (a) the CAESAR array and (b) the 8 array [3]. energies) and transform this into the intrinsic (rotational) frame of the nucleus.

2.2.1

Rotational Frequency, Moments of Inertia and Alignments.

It can be intuitive to observe the eect of increasing/decreasing rotational frequency on the structure of the nucleus. Using a simple rotational model, the rotational frequency, , 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

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 dierence in the excitation energies of the two states, E(I+1)E(I1), is simply the measured gamma-ray transition energy, E . The value of Ix can be calculated using Pythagoras theorem, such that Ix (I) = I(I + 1) K 2 I+ 1 2
2

dE(I) E(I + 1) E(I 1) dIx (I) Ix (I + 1) Ix (I 1)

(2.2.19)

K2

(2.2.20)

substituting in to the expression for , E (I +


3 2 ) 2

K2

(I 1 )2 K 2 2

(2.2.21)

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 gure 2.10.
ix Ix j j2 j1 R K c I

for decays from a state of spin I + 1 to one of I 1. Note that at high spins where I >> K, this expression simplies to E . 2

Ix

ix

11 00 11 00

11 00 11 00

Figure 2.10: Schematic of components of angular momentum for a deformed nucleus. It can be approximated by taking the total angular momentum and subtracting a xed amount of spin due to the rotational motion of the inert, deformed

26

nuclear core. The quasi-particle aligned angular momentum ix , is a function of rotational frequency and can be dened by ix () = Ix () Iref () (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) K 2 (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) 2 (2.2.24)

The Harris parameters are usually tted to states in the band to give a constant alignment above the rst band-crossing.

18 16 14 12
/ i [h]

106 Cd band 4

10 8 6 4 2 0 -2 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0 / h [MeV] 106 Cd 106 Cd 106 Cd 105 Cd band 1 band 2 band 3 h11/2 band

Figure 2.11: Alignment plots for band structures in 106 Cd and the h 11 band in 2 105 Cd using Harris parameters of I0 =7.0 2 /MeV and I1 =15.0 4 /MeV3 [3]. h h

27

Figure 2.11 shows the alignment plots as a function of rotational frequency for structures in 106 Cd and the rotational band built upon the neutron h 11 orbital in
2

the neighbouring 105 Cd. At a rotational frequency of approximately 0.45 MeV/ , h 106 band 1 in Cd shows a rapid decrease in rotational frequency and a large increase of approximately 10 in aligned angular momentum. This eect is known as a h backbend [77, 78] and consitutes the crossing of a the ground state band with the rst two-quasi-particle band, which comes lower in excitation energy compared to the simple ground state structure (ie becomes yrast) for spins above 10 . (This h is due to the two-quasi-particle band having a higher moment of inerta compared to the ground state conguration due to a reduction in pairing correlations). At higher spins still, there is another, smaller increase in alignment of approximately 3 h in the same band. This is interpreted as the alignment of a pair of neutrons occupying g 7 orbitals. If the structure of a band already contains on
2

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 band in 2 105 Cd already has a initial alignment of around 5 at the bandhead (from the h unpaired neutron), and thus the rst (h 11 )2 alignment (or crossing) observed in 2 106 Cd 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 106 Cd, both the h 11 and g 7 alignments observed
2 2

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-quasiparticle bands built consisting of both and unpaired h 11 and g 7 neutron coupled 2 2 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 dened by the simple relation, 28

h2 Erot (I) = (0) I(I + 1) 2I (I) The kinematic moment of inertia is given by I (1) (I) = while the dynamic moment is given by I (2) = 4 h dI d E I

(2.2.26)

(2.2.27)

(2.2.28)

Note a calculation of the dynamic moment of inertia requires only the dierence 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 studies 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 eects (Coriolis mixing and increased/decreased magnetic dipole moments) which show up in the decay schemes observed for such structures [35].

29

Figure 2.12: Rotational properties of the yrast band in 106 Cd, showing the backbend due to the ( 11 )2 neutron conguration crossing the yrast, ground state, 2 vacuum conguration, giving rise to a backbend. Note that the decrease in pairing correlations above the backbend gives an increased moment of inertia [3].

30

(a) j+4 j+2 j

j+5

(b) K+5 K+4 K+3 K+2 K Deformation aligned K bandhead Interband I=1 decays Strongly coupled

j+3 j+1

K+1

high-j, low-K Rotation aligned de-coupled band Large Coriolis mixing

11 00 11 00

1 0 1 0

Figure 2.13: Schematic of decays schemes and descriptions of strong and weak coupling 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 correspondingly 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 eects are large. The Coriolis operator can be written by [36] h2 (2.2.29) Hc = 2 I.j 2 I where j is the single particle angular momentum and 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 parity states, which intrude down from the next harmonic oscilator shell in the 31

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 1 of the odd-particle. In the case where K = 2 and If = Ii 1, the B(M1) decay strength is given by [35, 235] B(M1) = 3 2 (I K)(I + K) N (gK gR )2 K 2 4 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, , of a state of spin I with projection K by the expression [35, 88] by K2 (2.2.31) I +1 The value of gR , the g-factor of the core is approximated by the ratio of protons to total nucleons in the nucleus ( Z ). Thus, if the magnetic dipole moment can A be measured (see later) for a known spin, the K value of the state can be infered. = gR I + (gK gR )

2.2.3

Branching Ratios and g-Factors.

The intrinsic structure of a rotational band can also be inferred in a model dependent way by measuring the intensity on the I=1 and I=2 branches and their energies [3] 2K 2 (2I 1) E1 2 = 2 1+ (I + 1)(I 1 + K)(I 1 K) E2
5

I (I = 2) I (I = 1)

(2.2.32)

32

where is the E2/M1 mixing ratio, K is the projection of the angular momentum on the axis of symmetry and I is the state spin. E1 and E2 are the dipole and quadrupole transition energies in MeV and I are the relative -ray intensities.
g The magnitude of the quantity |gKQ0 R | , (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

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 = g = g + g (2.2.34)

0.93E1 |gK gR | = 2 Q0 I 1

(2.2.33)

where for is the projection of the orbital angular momentum on the axis of symmetry and = 1 is the single particle intrinsic spin projection. For 2 protons g =1 and for neutrons g =0. The g values are simply the g-factors for the single proton and neutron (+5.59 and 3.83 respectively), attenuated by a factor of 0.60.8 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 dierent congurations 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 coecient can be extracted and the sign of the mixing ratio inferred. The sign of the mixing ratio g is the same of the sign of the quantity gKQo R [39], ie. sgn () = sgn gK gR Q9 (2.2.35)

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, (from the measured angular distribution of the interband I=1 33

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

Figure 2.14: Comparison of gK gR values for the rotational band built on the high-K isomeric state in 136 Sm. The observed branching ratios are consistent with a K =8 assignment and rule out the K = 7 conguration [38]. transitions in the band). It demonstrates clearly that the K = 8 , two-quasineutron conguation is the only one consistent with the measured branching ratios. Donau and Frauendorfs Geometrical Model. Dnau and Frauendorf [40, 41, 3] used a simplied geometrical model to estimate o 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 2 ) by, N 3 K2 (g (1) gR ) I 2 K 2 i(1) g (2) gR i(2) x x 8 I 2
2

B(M1) =

(2.2.36)

where the g are g-factors for specic 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 2 3(I K)(I K 1)(I + K)(I + K 1) Q 16 o (2J 2)(2J 1)J(2J + 1) 34 (2.2.37)

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 2 3K 2 (J K)(J + K) Q 16 o (J 1)(J)(2J + 1)(J + 1)
2 2 (h11/2) (g 9/2) (h11/2)2 (g7/2d5/2) (g9/2)2

(2.2.38)

106Cd 102

2 ( /eb)

B(M1)/B(E2)

2 (h11/2) (g9/2g7/2) 101

100

2 (h11/2) (g7/2d5/2 )

19

21

23

25 Spin

27

29

31

Figure 2.15: Branching ratio comparison for dierent congurations in the mutliquasi-particle structure in 106 Cd. [3]. The experimental expression
B(M 1) B(E2)

branching ratios (in units of

2 N ) e2 b2

are given by the

5 E2 1 I (I = 1) B(M1) = 0.697 3 B(E2) E1 1 + 2 I (I = 2)

(2.2.39)
B(M 1) B(E2)

As gure 2.15 shows, by comparing the experimentally observed

values

with those extracted using the B(E2)s from the rotational model and the B(M1)s from the geometrical model, a specic assignment of a mutli-quasi-particle state can be made.

2.2.4

Two State Mixing.

Often two nuclear congurations of the same spin and parity can not be described in terms of exactly pure congurations but rather linear combinations of two dierent congurations which wavefunctions 1 and 2 [164]. We can write the wavefunction for the nal state in terms of admixtures of the two basis states 1 and 2 , ie 35

= + 1 2 h1 > 2V

E1

2 unperturbed basis states

h2 = + 1 2 E2

experimentally observed, mixed states

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

= 1 + 2 equation such that H = E

(2.2.40)

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

(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 eects) v. Then h1 v12 v12 h2 =E

h1 and h2 are the unperturbed energies that the states would have if there was no interaction between the two congurations, ie. the energies of the theoretical states corresponding to the pure wavefunctions 1 and 2 . 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 congurations. and are the eigenvectors. h1 E v12 v12 h2 E 36 =0

Since and are not zero, the determinant of the rst 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 which gives
2 (h1 E) (h2 E) v12 = 0

=0

(2.2.42)

This is a quadratic equation in E which has two roots (which are the two energy eigen-values) given by h1 + h2 2 h1 h2 2
2 2 + v12

E=

(2.2.43)

The wavefunctions of the experimentally observed states can then be written in terms of a linear combination of the basis states 1 and 2 . The strengths of the two nal states in terms of the basis states weighting, and can be calculated in terms of the h1 , h2 , E1 , E2 , and v12 by solving the expressions (h1 E) + v12 = 0 for the higher energy state and (h2 E) + v12 = 0 for the lower one and normalising with the condition that 2 + 2 = 1 (2.2.46) (2.2.45) (2.2.44)

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 = where 1Y 1 2 1Y 2 2 (2.2.47)

37

Y = and X=

X (1 + X 2 )

(2.2.48)

(h2 h1 ) 2v12

(2.2.49)

10 8

180 W |V|=141.5 keV

Alignment (h)

6 4 2 0 0 100 200 300 Rotational frequency (keV) 400

Figure 2.17: Alignment plots for the yrast and K =8+ band in 180 W showing the unperturbed bands [42]. From the experimentally observed (perturbed) energies of two states, an estimate 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 congurations which are thought to mix at the band crossing point. Figure 2.17 shows the bandcrossing of the yrast states in 180 W (decay scheme shown in gure 2.18) which has been attributed to an interaction between the K=0 ground state band and the K = 8+ , t-band [42]. Note the interaction between the even spin members of the two bands, while the odd-spin members of the K =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

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

h1 E v12 v13 v12 h2 E v23 = 0 v13 v23 h3 E

2.3

Selected Topics in High Spin Nuclear Structure.

In this short section, we will briey 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 potential energy surface corresponding to dierent shapes of the nuclear mean eld[44]. Such eects often occur in nuclei with oe near magic numbers of protons and

39

neutrons. Here, a spherical conguration competes with a deformed shape corresponding to particle-hole excitations across the shell gaps. This deformed conguration populates deformation driving orbitals (those which drop down rapidly with energy for increasing deformation). Good examples of studies of sphericalprolate shape coexistence in heavy nuclei for very neutron decient 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 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 A150 region are centred about 152 Dy86 , where both proton and neutron 66 numbers correspond to superdeformed magic numbers [52]. Similar regions of superdeformation have been identied around and 82 Sr44 [61]. 38
3000.0
Spin (h)
192Hg Band 2 194Hg Band 3

132 58 Ce74

[65, 59, 64],

192 80 Hg112

[50]

50 40 30 20 10

2000.0
(Mev/h ) J Spin (h)
2

130 120 110 100 90 1.2 1.0 0.8 0.6 200

Counts

1000.0

(2)

0.0 100.0

300.0

500.0
Energy (keV)

700.0

900.0

400 600 Egamma (keV)

800

Figure 2.19: Identical superdeformed bands in

192

Hg and

194

Hg [66].

A partilcularly interesting and puzzling aspect of the study of superdeformed 40

bands is the presence of rotational bands with almost identical energy spacing in SD bands of neighboring nuclei, the so called identical bands eect [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 A190 region [67, 68], however such links in the A80 and 150 regions currently remain elusive.

2.3.2

Band Terminations.

Since the nucleus is a nite 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 eect 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 projections, , on the nuclear axis of symmtery, we can assign a total projection of the angular momentum on the symmtery axis, K, where K = j j . The Kselection 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 dierering 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 Z72 and N106, both the proton and neutron Fermi surfaces lie in the vicinity of high- orbitals. (K-isomeric states have also been 41

Figure 2.20: Smooth band termination in

110

Sb [76].

observed in N=74 isotones [90, 91, 92, 93, 94]). By breaking pairs of nucleons, very high- 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 congurations are built on conguration consisting of upto eight unpaired nucleons [80, 9] and the eect this has on the overall pairing correlation can be invetsigated in a conguration 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 where x= h0 2p,n (2.3.51) x + 1 + x2 1 x 1 + x2 (2.3.50)

42

where is the pair gap parameter, h0 is the harmonic oscilator frequency and I is the moment of inertia.

Figure 2.21: Partial decay scheme for 178 W showing 0, 2, 4, 6 and 8 quasi-particle structures [9]. By investigating the moments of inertia of mutli-quasi-particle states (where various numbers of valence pairs are blocked), the eect on pairing on the moment on inertia (and visa versa) can be investigated for various specic orbital congurations. As an example, gure 2.21 shows the partial level scheme for 178 W, which shows a large number of dierent 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 gure 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

80.0

Moment of inertia (h MeV )

-1 2

60.0 8-qp 6-qp 40.0 2-qp gsb 20.0 Exact calc. BCS calc. Expt. Rigid body 4-qp

0.0

10

15 Bandhead Spin

20 (h)

25

30

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

2.3.4

Octupole Correlations.

The study of reection assymetric, octupole deformed nuclei at high spins [95] is centred around regions of the nuclear chart where there are sets of nuclear orbitals which dier 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 , g 9 ), (i 13 , f 7 ), (h 11 , d 5 ) and (g 9 , p 3 ). Regions where octupole correlations are 2 2 2 2 2 2 2 2 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 Z88 (i 13 , f 7 ) and N134 (j 15 , g 9 ) 2 2 2 2 [95, 96, 97] and in the neutron rich rare earth region around Z 56 (h 11 , d 5 ),
2 2

N 88 (i 13 , f 7 ) [95, 99, 100]. 2 2 Nuclei with static octupole deformations (3 ) are characterised rotational bands of opposing parity interleaved by enhanced electric dipoles (E1) (see gure 2.23). Typical E1 strengths in heavy nuclei range between 104 and 107 Weisskopf units. For nuclei which exhibit octupole deformation, this increases to 44

(28 + )

518.2
(26 + )

(27 )

(28 + )

511.6
(25 )

492.5
(26 + )

(27 )

500.8
(24 + )

484.0
(25 )

496.8
(23 )

472.6
(24 + )

485.1
(22 + ) (20 + )

467.0
(23 ) (22 + )

479.8
(21 )

452.9
(22 + )

467.1
(19 ) (20 + )

448.2
(21 )

446.8
(20 + ) (19 )

476.7
(18 + )

462.8
(19 )

431.6
(20 + )

470.8
(17 )

447.1
(18 + )

427.4
221.8 189.5
214.9
176.5
(19 )

436.1
(18 + )

452.3
(16 + )

442.9
(17 )

411.3
18 +

408.6
(17 ) (18 + )

211.4

444.8
(15 )

424.4
(16 + )

404.4
17

422.4
(16 + )

425.5
(14 + )

419.4
(15 )

391.1
16 +

254.6
277.5 308.3

380.1
(15 ) (16 + )

412 390
(14 + )

192.1

415.0
173.1
(13 )

400.2
(14 + )

396.0
(12 + )

391.1
(13 )

369.8
14 +

202.1

378.4
15

403.8
(14 + )

167.6
180.4 166.5 148.3
173.3 188.6 210.5
4+ 2 + 67.7 0+

348.3
(13 )

363.9
(10 + )

206.2
157.4

379.6
(11 )

373.8
12 +

348.5
13

356.7
11

347.3
12 +

378.9
(12 + )

312.3
(11 )

364.1
(12 + )

330.1
(8 + )

180.9
149.3 140.1 153.1

338.3
(9 )

345.3
10 +

314.5
11

314.8
151.2 160.5
182.2 206.0 +
2+ 0 9

321.3
10 +

347.8
(10 + )

343.6
380.8 418.4
4+ 2 + 63.8 0+

272.0

289.0
229.3

313.7
8+ 6+ 4+ 84.5

275.3 231.0

(9 ) (7 ) (5 ) 656.0 (3 ) 537.6 (1 ) 474.1

332.2
(10 + )

292.9
(6 + ) 4+ 190.7 2+

(7 ) (5 ) 3 317.3 1 242.1

265.9
207.9

290.7
8+ 6+

309.0
(8 + ) (6 + )

225.0

276.1
228.6
166.5

293.6
(8 + ) (6 + )

248.4

172.2
0+

5 142.6 3 1 215.9

252.9
205.0
144.0

7 180.2 5 447.0 3 321.5 1 253.7

(5 ) 879.8 (3 ) 785.8 (1 ) 710.9

262.6
206.9
140.8

247.3 192.3
129.4
(4 + ) (2 + ) 57.4

111.2

0+

222

Ra 88

224

Ra 88

226

Ra 88

228

Ra 88

230

88

Ra

Figure 2.23: Decays schemes for octupole deformed Radium (Z=88) nuclei showing enhanced E1 decays between rotational structures of diering parities [97, 97]. between 103 and 102 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 octupole deformation. The electric dipole moment is related to the B(E1 : Ii If ) by the expression [95]

B(E1 : Ii If ) =

2 2 3D0 (I K)(I + K) 3D0 < Ii 010|If 0 >2 = 4 4 I(2I + 1)

(2.3.52)

2.4

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

The -ray branching ratio , for a given state in a rotational band, is given by, = P2 (E2) T2 (E2) = T1 (E2) + T1 (M1) P1 (E2) + P1 (M1) (2.4.53)

45

where T1 and T2 are the -ray intensities for spin changes of 1 and 2 respectively. P1 and P2 are the corresponding transition probabilities ( intensity). Note that the ratio T1 (M 1) = 12 , where is the quadrupole/dipole mixing ratio. The T1 (E2) transition probabilities1 are related to the reduced transition probabilities (B(XL)) by equations 2.4.54. 1 P (E2; Ii If ) 5 1.225 109 E 1 B(M1; Ii If ) = P (M1; Ii If ) 3 1.758 1013 E B(E2; 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, E2 E1
5

The reduced in-band transition probabilities1 are given by, B(E2; Ii K If K) =

B2 (E2) 1.225 109 2 B1 (E2) 1.225 109 + B1 (M1)E1 1.758 1013

(2.4.55)

5 2 2 e Qo | < Ii 2K0|If K > |2 16 5 2 2 e Qo | < Ii 1K0|If K > |2 (2.4.56) B(E2; Ii K If K) = 16 3 2 B(M1; Ii K If K) = e | < Ii 1K0|If K > |2 (gK gR )2 K 2 4 where Qo is the intrinsic quadrupole moment and gK and gR are the intrinsic and rotational gyromagnetic ratios respectively. The relevant Clebsch-Gordon coecients2 are given below. 3(I K)(I K 1)(I + K)(I + K 1) E2(I = 2) = (2I 2)(2I 1)I(2I + 1)
1/2

3(I K)(I + K) E2(I = 1) = K (I 1)I(2I + 1)(I + 1)

1/2

(2.4.57)
1/2

(I K)(I + K) M1(I = 1) = I(2I + 1)

1 K.E.G. Lbner in, The Electromagnetic Interaction in Nuclear Spectroscopy, W.D. Hamilo ton (Ed), North-Holland (1975) Chapter 5 2 The Theory of Atomic Spectra, Condon and Shortley (1935) reprinted (1963) p76-77

46

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 =,

5 2 2 3(IK)(IK1)(I+K)(I+K1) 1.225 109 e Qo 16 (2I2)(2I1)I(2I+1) 2 2 3(IK)(I+K) 3 1.225 109 + 4 e2 (gK gR ) K (IK)(I+K) 2 (I1)I(2I+1)(I+1) I(2I+1) E1

E2 E1

5 2 2 e Qo 16

1.758 (2.4.58)

1013

Rearranging these terms gives equation 2.4.59



3(IK1)(I+K1)(I+1) (2I1)2K 2 13 g 2 2 3 + 3 (gK Q2 R )2 K 1.75810 9 (I1)(I+1) 4 E1 1.22510 o 5 16

E2 E1

5 16

(2.4.59)

This leads to the following expression for the branching ratio. E2 E1


5

1+

4 5

(IK1)(I+K1)(I+1) (2I1)2K 2 (gK gR )2 (I 2 + 1)(1.435 2 Q 2 E1 o

104 )

(2.4.60)

Making this substitution gives equation 2.4.61

This equation is identical to that quoted by Alexander et al.3 To convert Qo from Qo units of efm2 to eb, there is an additional factor required, namely, Qo 100 .

(gK gR ) = Qo

5 E1 (0.8347) 4 (I 2 1)

1 E2 E1

(I K 1)(I + K 1)(I + 1) 1 2K 2 (2I 1) (2.4.61)

1/2

This result can also be expressed in terms of the quadrupole/dipole mixing ratio , E1 (gK gR ) (2.4.62) = 0.933 Qo (I 2 1) where is related to the quadrupole admixture q, by,
5 2 2K 2 (2I 1) E1 q= = 5 1 + 2 (I K 1)(I + K 1)(I + 1) E2

(2.4.63)

The factor
3

(gK gR ) Qo

should be a constant for a given rotational band.

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

47

Chapter 3 Experimental Gamma-ray Spectroscopy.


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, Eg (3.1.1) 2kT where Eg is the band gap, k is Boltzmanns constant and C is a constant which is material dependent. P (T ) = CT 2 exp
3

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 temperature). 48

If radiation enters the electrically neutral depletion layer, electron-hole pairs can be created. The electrons will ow towards the positive potential and the holes to the negative. This charge can be collected and converted to an output voltage by a pre-amplier. 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 137 Cs). General considerations for a good gamma-ray spectrometer device are that (a) it must have excellent energy resolution, (b) a good photopeak eciency and (c) good timing properties. While sodium iodide has a better eciency 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 dierent components. These include, 1. The Full Energy Peak. The peak corresponding to where all the incident radiations energy has been collected by photoelectric absorption (some 49

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 = hh = h =180 ) is given by E=
E 1+2
E me c2 h (1cos) me c2 1+ h 2 (1cos) me c

Therefore, the minimum energy removed by scattering an electron (when , where me c2 = rest mass of the electron= 511 keV. This minimum amount of energy being lost in a Compton scattering, gives rise to the Compton background being essentially cut o at an energy E below the full energy peak. Note for E >> 511 keV, 2 E me c 250 keV. 2 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

2511 keV gamma-rays. One or both of these 511 keV gammas may then escape from the detector with no further interaction. If the initial gammaray enegy has an energy E , then the escape peaks lie at energies E 511 keV and E 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 corresponds to the energy of the photon after it is scattered. This is simply the energy dierence between the full energy peak and the Compton edge. 6. Annihilation Peak. One observes a peak at 511 keV due to annihilation radiation from pair-production caused by the initial radiation in the surrounding material (assuming E > 1.022 MeV). The 511 keV is then measured in the detector.

50

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 Eciency.

(See Knoll p427.) The eciency of a germaium detector is usually given relative to that of a 3in 3in NaI(Tl) detector for the 1333 keV gamma-ray in 60 Ni (from a 60 Co source). Typical eciencies range from 30 % to 70 % for hyperpure germanium detectors. The eciency response varies as a function of energy and is usually empirically deduced using a variety of standard calibration sources such as 152 Eu and 133 Ba (see gure 3.1.)

Figure 3.1: Gamma-ray spectra for (a) sources.

152

Eu and (b)

133

Ba eciency calibration

51

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 eciency gamma-ray scintillator (usually Bi4 Ge3 O12 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 60 Co source for an unsuppressed detector is only about 20 %. This rises to 5060 % for a suppressed detector with almost no loss in the number of counts in the full energy peak.
Liquid nitrogen cold finger photomultiplier tubes

BGO

Ge

NaI

NaI

incoming gamma-ray

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 coincidence spectroscopy. For an event where two gamma-rays from a cascade are measured (a or 2 event), for two unsupressed detectors, the probability that both gamma-rays (for E 1 MeV) will be in the full energy peak is approximately 0.20.2=0.04. If however, both detectors are surrounded by a

suppression shield this rises to 0.60.6=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 in52

BGO

(a) Eurogam I

Number of Counts

Energy (keV)

Figure 3.3: Eect 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% eciency. In practice this is not possible, however, over the last 10 years there have been signicant advances in the development of gamma-ray spectrometers consisting of large numbers of germanium detectors (or arrays) such as TESSA3 [101], the 8 spectrometer [103], EUROGAM [174, 173] and GAMMASPHERE [104]. These devices have revolutionised the eld of gamma-ray spectroscopy by allowing hitherto unthinkable sensitivity to weak transitions. The following subsections will describe the important factors in array design and how the maximum performance can be achieved. 53

3.3.1

Resolving Power and Total Photopeak Eciency.

The resolving power [174] of a gamma-ray spectrometer is given by R where R= SE PT nal E (3.3.2)

SE is the average energy spacing consecutive -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 nal must be made on either the energy resolution, E , and/or the peak to total fraction, P T . nal E is the full width at half maximum resolution of -rays obtained in the detector spectrum. This is made up from a number of eects including the intrinsic, statistical resolution of the detector, the eect of Doppler broadening, the spread in the recoil cone and recoil velocity. The total resolution of a given -ray transition in a fusion-evaporation reaction is given by [174]
nal 2 2 2 2 E = Eint + Eopen + Eang + Evel
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 to the recoil velocity direction to a value Es , where, if E0 is the unshifted energy, Es () = Eo 1 ( v )2 c
v cos c

Since the geramnium detector measuring the gamma-ray has a nite opening angle, there will be a spread in the Doppler shifted energy across the face of the detector (see gure 3.4). If the opening angle of the detector is , then dierentiating equation 3.3.4 gives, dEs Es v Eo sin d c (3.3.5)

v E0 1 + cos c

(3.3.4)

54

germanium detector 111111 000000 111111 000000


111111 000000 111111 000000 111111 000000 111111 000000 111111 000000 111111 000000 111111 000000 111111 000000 111111 000000 111111 000000 111111 000000 111111 000000 111111 000000 111111 000000 111111 000000 111111 000000 111111 000000 111111 000000 111111 000000 111111 000000 111111 000000 111111 000000 111111 000000 111111 000000 111111 000000 111111 000000 111111 000000 111111 000000 111111 000000 111111 000000 111111 000000 111111 000000 111111 000000 111111 000000 111111 000000 111111 000000 111111 000000 111111 000000 111111 000000 111111 000000 111111 000000 111111 000000 111111 000000 111111 000000 111111 000000 111111 000000 111111 000000 111111 000000 111 000 111 000 111 000 111 000 111 000

beam direction, velocity, v Es = Eo (1 + v/c cos ) s = v/c sin


Figure 3.4: Schematic of the eect of nite opening angle for a detector causing Doppler broadening of the gamma-ray lineshape. The full expression, including broadening due to velocity spread (see g. 3.5), can be approximated by Es Eo cos v v Eo sin c c (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 eective 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 eciency, 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

Figure 3.6 shows the eect 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 dened 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 germanium detector, by improving and maximising the Compton suppression and/or the use of add-back such as in the cluster detector [179] (see gure 3.7). A further advantage of using a clover type detector [176, 178] comes when trying to look at higher energy gamma-rays. As gure addback shows, the majority of counts with energies less than 1 MeV are detected in a single element. However, 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 signicant increase in the full energy peaks obtained using the detector in addback 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 dierent for electric and magnetic type transitions. Thus, by comparing the eciency corrected intensities of added-back lines from horizontal and vertical scatters, the electric or magnetic (or mixed) nature of the transition can be eludidated. This is highlighted for experimental data taken from the EUROGAM 2 array in gures 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

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 oine, 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

Figure 3.5: Gamma-ray spectra taken from the reaction 28 Si on a 40 Ca target with a v of 1.8%. Note the doppler broadening is larger away from 90 in this c case due to the large spread in recoil energy associated with the alpha-particle evaporation [106] .

58

1600

Phase II FWHM = 4.8 keV

1200

800

Number of counts

400 1400 800


Phase I

1450

1500

FWHM = 6.1 keV

600

400

200 1400 1450 1500

-ray energy (keV)


Figure 3.6: Eect of using clover detectors in improving energy resolution in EUROGAM phase II [174].

59

(a)

(b)

(c)

(d)

(e)

(f)

(g)

(h)

Figure 3.7: Spectra gated on add-backs from cluster detectors at PEX using the reaction 28 Si+40 Ca [107].

60

0.8

ray polarisation P

0.6 0.4 0.2 0.0 0.2 0

Band 12 (E2)

Band 3 (M1) 5 10 15 20 25 30

Spin (h)
Fig. 5
/home/esp/text/figs/i117/pol3.psk Paul et al PRC

Figure 3.8: Polarization measurements showing the dierence between magnetic and electric type decays in 132 Ce [186, 185]. 61

1.5 (a) 1.0


117

I gated

Counts (10 )

0.5 0.0 10
Electric Magnetic Electric Magnetic

(b) Ungated

10 0.0
/home/esp/text/figs/i117/polspec.psk Paul et al NPA

0.5

E (MeV)

1.0

1.5 Fig. 4

Figure 3.9: Dierence spectra for horizontal and vertical scatters for transitions in 117 I showing how a polarization measurement can be use to dierentiate between electric and magnetic transitions (see [186, 185].

62

Chapter 4 Channel Selection In Fusion-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 nucleus. In order to associate gamma-rays with specic 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 separator, gas lter techniques) or by measuring the particles evaporated from the compound system (charged particle balls, neutron detectors). The former generally 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 eciency. 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 dierent for dierent reaction products and (typically) depends on the number of evaporated particles. Usually, residual nuclei formed by the evaporation of fewer particles leave the nal nucleus in a state of higher spin and excitation energy than channels with higher

63

particle emission multiplicities. If a high eciency, high granularity -ray detector (calorimeter) can be placed around the target position, this eect can be used to give a degree of channel selection between dierent residual species [101] (see gures 4.1 and 4.2). This form of channel selection is particularly eective 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 River 8 spectrometer for the reaction 76 Ge(34 Sn,xn)110x Cd. Spectra (a) and (b) are gated by the 540 keV 15 11 and 633 keV 2+ 0+ transitions in 105 Cd and 2 2 106 Cd respectively. Note the lower values of fold and sum energy observed for the higher neutron multiplicity evaporation channel [3]. For example the 8 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 energy of the nucleus), with the number of detectors ring, giving an estimate of the total gamma-ray multiplicity of the cascade (which is related to the initial angular momentum of the system). As gures 4.1 and 4.2 clearly demonstrate, the entry point of a nucleus in the spin/excitation place depends on the number of evaporated neutrons for the
76

Ge+36 S reaction. By setting software gates 64

Sum Energy (entrance excitation energy)

4n + 106Cd

5n + 105Cd

Fold (~ multiplicity ~ total input angular momentum)

Figure 4.2: Two dimensional Fold and Sum Energy spectra from the inner ball of the Chalk River 8 spectrometer for the reaction 76 Ge(34 Sn,xn)110x Cd gated by the 540 keV 15 11 and 633 keV 2+ 0+ transitions in 105 Cd and 106 Cd 2 2 respectively. in oine sorting on various fold/sum-energy conditions, the relative intensity of transitions from specic evaporation residues can be enhanced.

4.2

Studies of Very Neutron Decient Nuclei.

For very neutron decient compound nuclear systems, the reduced proton separation energies compared to compound nuclei closer to stability mean that it is easier for protons and -particles to tunnel through the Coulomb barrier. This allows charged particle evaporation to compete with and nally dominate over neutron emission, and the population of between 10 and 20 dierent nuclei in a single reaction is possible. The eectiveness of the particular method of channel selection used depends on the nal nucleus of interest. For example, in studies of A80 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 detector array would not be suitable, whereas a recoil-separator and split-anode gas ionisation-chamber, or neutron-detection system might yield useful information.

65

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

Figure 4.3: (a) Total projection of the 28 Si+40 Ca reaction from PEX. (b) Same spectrum gated on the 2p condition. (c) 2p gated spectrum with 3p component subtracted. (d) 2p gated spectrum with both 3p and 4p components subtracted [106].

67

Figure 4.4: Particle gated identication spectra gated numbers of emitted protons from the ANU particle detector ball with higher multiplicities subtracted [38].

68

The ideal charged particle detector has almost 4 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 dierent types of particles. In practice however, there are gaps and dead regions between the detectors, and this coupled with kinematic focussing eects means that such arrays are never 100% ecient. The result is for example a 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 gure 4.3). However, as shown in gure 4.4, once these contaminatnts from higher multiplicity channels have been subtracted, very clean identication 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 gure 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 gure 4.5.

V R V cm P lab

lab

1111 0000 1111 0000 1111 0000 1111 0000 1111 0000 1111 0000 1111 0000 1111 0000

Rcm

V P lab

Vp cm
11 00 11 00 11 00

Figure 4.5: Schematic of the kinematics of particle evaporation following a fusionevaporation reaction. Note how the recoil direction is altered by the particle evaporation.
2 If the beam energy is EB and EB = 1 MB vB , where MB is the mass of the 2 beam nuclei and vB is the beam velocity in the lab. frame, then by conservation

69

of momentum, the velocity of the centre of mass vcm is given by

70

MB vB = (MB + MT ) vcm and by substitution

(4.2.1)

MB 2EB (4.2.2) (MB + MT ) MB where MT is the mass of the target nuclei. The velocity of an evaporated vcm = proton in the rest frame of the compound nucleus can be calculated (by the cosine rule) such that
2 2 2 vpcm = vcm + vplab 2vcm vplab coslab

(4.2.3)

where vplab is the emitted proton velocity in the laboratory frame and lab is the lab angle between the beam direction and the direction of the emitted proton. By substitution MB 2MB 2 2 2EB + vplab (MB + MT ) (MB + MT ) 2EB coslab MB

2 vpcm =

(4.2.4)

The proton energy is the centre of mass (intrinsic frame of the compound system) is given by 1 2 Epcm = Mp vpcm 2 Multiplying both sides of equation 4.2.4 by 1 Mp gives 2 1 1 MB 1 1 Mp 2MB 2 2 mp vpcm = Epcm = Mp 2 2EB + Mp vplab 2 2 2 2 (MB + MT ) (MB + MT ) Given that vplab =
Eplab , Mp

(4.2.5)

2EB vp coslab MB lab (4.2.6)

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 (lab ), the beam energy (EB ) and the masses of the beam (MB ) and target (MT ) nuclei respectively, 2 Mp MB MP MB EB Eplab coslab (4.2.7) 2 EB + Eplab (MB + MT ) (MB + MT ) 71

Epcm =

A similar expression can be obtained for -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 carbon and oxygen are common target contaminants. However, the lighter mass of these nuclei, means that in general, the eect 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.
SINGLET STATES S3 TRIPLET STATES T3 S2

T2

S1 T1

prompt fluorescence

S0

Figure 4.6: Schematic of the decay of prompt and delayed light from an organic scintilator. Often, depending on which quantum levels the electron decays through, this 72

absorption

delayed fluorescence (phosphorescence)

light is emitted with two characteristic decay lifetimes, a fast one known as prompt uorescence and a slower type known as delayed uorescence (see gure 4.6). The relative intensities of these two types of light depends on the type of radiation which excited the scintilator in the rst place and as we shall see later can be used to distinguish between dierent types of radiation. Phoswich Detectors. Phoswich detectors are combinations of two detectors, in optical contact. The front detector consists of a thin ( 100 m) 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 ( 1s) 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 dierent time regions by setting two separate time discriminator gates on the short (E) and long (E) components of the signal (see gure 4.7). By making a 2-dimensional array of E versus E, dierent evaporation particles can be identied. 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 -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 scintilators and the Penn 4 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 identication 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

for example germanium gamma-ray detectors may take up.

74

3 mm CaF2 (Eu)

0.1 mm NE102A

fast (DE)

"E"

"DE" slow (E) time

light output / voltage

fast 0-40 ns slow 60 - 250 ns

Figure 4.7: Schematic of the operation of a phoswich detector for discrimination between protons and alpha particles.

4.2.3

Silicon Detectors.

Si Energy Loss Detectors. Thin (150200 m) silicon detectors have also been used to give channel selection between protons and -particles emitted in fusion evaporation reactions [113, 114, 134, 345]. Particle identication is achieved due to the dierential energy loss of protons and -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 4 geometry and the thin nature of the detector means that attenuation of reaction gammarays in the ball material is kept to a minimum. This is an important factor in keeping the low-energy gamma-ray detection eciency of such experiments at a 75

premium.

76

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 eect 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 -particles. Thus to rst order, the energy deposited in the detector can be used to distinguish protons and -particles. The problem is the high energy tail of the proton distribution leaks into the -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 gure 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 dierent 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 dierent 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 amplier output, verses the total energy deposited in the detector, to dierentiate between dierent 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 fusionevaporation reactions. The EUCLIDES and ISIS arrays [120] are constructed in 77

23 29 28 31 IN 22 27 26

24

IN

30 25

OUT 21

(c)

(a)

18 13 12 17 7 12 2 1 11 16 6 11 5 10 15 15 20 8 14 3 4 9 14
OUT 1 5

13 19
2 3 4

17,18,19 12,13,14 7,8,9

23,24 28,29

TARGET

31 IN

6,10 11,15 16,20

26,27,30 21,22,25

(b)

(d)

Figure 4.8: Schematic of the PEX Si inner ball geometry for selection of light charged 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 gure 4.10 and also the locii corresponding to events where more than one evaporated particle has hit a single telescope (such as p. Note that due to the kinematics of typical fusionevaporation reactions, 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 ( = 0.4 1.0 s), the amplitude of which 78

Figure 4.9: Energy spectra for charged particles from the PEX inner Si ball for the reaction 28 Si+40 Ca showing the eect of kinematic focussing [106]. is radiation type dependent and a slower ( = 7 s), radiation independent component. Particle identication can be achieved using a similar pulse shape discrimination technique as outlined above for the phoswich type detectors. High eciency, high granularity devices based on this method, such as the ALF-ball [112] and the MICROBALL [124] have been used very successfully inside the 8 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 aorded for an element of the ALF ball. The two axes in gure 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 dierence between a T=0

79

reference time (such as beam burst of prompt--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 preamplied output pulse from the detector through a bipolar shaping amplier [112]).

80

Figure 4.10: Particle hit pattern rom the ISIS silicon E, E array for dierent angles in the lab frame for the reaction 40 Ca on 40 Ca. Note the general forward focussing 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 protons are detected will contain counts from 2p,2pn,3p,3pn,4p,2p,...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% ecient) but can be identied 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

14 = 108.0 12 10 8 6 4 2

14 = 148.3 12 10 8 6 4

E (MeV)

2p p
10 E (MeV) 20 30

p
10 E (MeV) 20 30

Figure 4.11: Particle identication spectra from the ISIS silicon E, E array for dierent angles in the lab frame for the reaction 40 Ca on 40 Ca for backward ISIS angles in the lab frame. Note the general forward focussing of the particles and the absence of alpha particles at backward angles [120, 121, 122].

14 = 31.7 12 10 8 6 4 2

14 = 69.9 12 10

E (MeV)

8 6

2p p
10 E (MeV) 20 30

2p p
10 E (MeV) 20 30

4 2

Figure 4.12: Particle identication spectra from the ISIS silicon E, E array for forward angles in the lab frame for the reaction 40 Ca on 40 Ca [120, 121].

82

Time response of CsI(Tl)

Light Output

Figure 4.13: Time verses total energy spectrum for one element of the chalk River ALF ball for the reaction 76 Ge+34 S. Note the protons and alpha particle locii are clearly 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 xed (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 threshold). 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

such a scenario will result in the emission of a further nucleon (neutron) to bring the nal 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 oline gates on measured average proton energy will yield some selection between dierent 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 eect 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 the
58

Fe(27 Al,3p)82 Sr reaction [135] highlighting this eect is shown in gure 4.15.

The Total Energy Plane Method. As pointed out previously, a problem with using charged particle balls to identify weak, low-multiplicity channels in very neutron decient 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,2p,2pn...) 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 dierent (specic) exit channels.

84

This total excitation energy bought into the compound system is then emitted in the form of -ray decays and emitted nucleon kinetic energies such that

85

Figure 4.14: Centre of mass proton energy spectra for 27 Al+58 Fe reaction at a beam energy of 92 MeV, gated on transitions in 82 Sr and 81 Sr from the 3p and 3pn channels respectively [135]. The lower average proton energy for the 3pn channel is clearly demonstrated.

86

Figure 4.15: Entry-point excitation energy centroids in coincidence with discrete gamma-rays depopulating states at excitation energies between 0 and 8 MeV in 82 Sr. Note that the higher excitation energy states, typically come from nuclei formed with higher energy entry points (ie. lower proton evaporation energies) [135].

87

ECM = H + Tpart

(4.2.9)

where H 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 specic channel. Therefore if H 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 gure 4.16).

Total Particle Kinetic Energy

Ecm

3p 4p 3pn a3p etc

Ecm Total gamma-ray energy


Figure 4.16: Schematic Total Energy Plane for a 3p out channel showing a constant line for 3p events and clear separation from the higher multiplicity charged particle 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 4 charged 88

particle detector (to obtain the centre of mass energies of the emitted particles) in conjunction with a 4 gamma-ray inner ball (to obtain an estimate for the H ) 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 Eects.

Figure 4.17: Spectra showing the SD band in 85 Nb using the reaction 58 Ni(36 Ar,2p)85 Nb at an energy of 180 MeV. The improvement in signal to noise due to particle gating and kinematic corrections are clear [166]. By conservation of momentum, the evaporation of particles from the compound system may have the eect 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

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 -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, E 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
n

E = E0 1 + E0 where E0 = and

Ei
i=1

(4.2.10)

MCN vCN .d MR c

(4.2.11)

Mi vi .d (4.2.12) MR c where vCN .d is the vector dot product between the velocity vector of the Ei = 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 the spectra for charged particle evaporation.
85

Nb by kinematically correcting

90

4.3

Neutron Detection.

The detection of neutrons evaporated from compound nucleus reactions is usually achieved using organic scintilators containing a large amount of hyrdrogen. A common type of detector consists of the liquid scintilator NE213 [136]. Differentiation between neutron and gamma-ray induced events is usually achieved by one of two methods: (a) time of ight 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 40 Ca+24 Mg from the AYEBALL array at Argonne National laboratory. The gamma-ray induced events all have a common time prole (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 ight spectra showing clear separation between gamma-ray and neutron event. From the reaction 40 Ca+24 Mg 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 eciency. In order to maximise the neutron detection eciency, the detectors must be placed closer to the target position and pulse shape discrimination [136, 127, 125] is used to separate gamma and neutron events (see gure 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

can also be used as a gamma-ray multiplicity detector [136]).

neutrons

gammas

Neutron Time of Flight

Zero-crossing time
Figure 4.19: Time of ight verses z component for neutrons measured in the 28 Si+40 Ca reaction at PEX [106]. In the studies of very neutron decient nuclei, it is often desirable to combine both charged particle and neutron channel selection devices to obtain clean, highresolution gamma-ray spectra for a specic, weakly populated channel. In modern arrays, dierent forms of channel selection are often used in tandem to give a maximum eect. A good example of this is the use of both charged particle inner balls and neutron detectors to study very neutron decient nuclei around NZ100 at the Niels Bohr Institute, Denmark [126, 127, 128, 129]. The PEX array (see gure 4.21) is an extension of this project, as is the EUROBALL neutron wall [130]. Figure 4.20 shows the eect of using both charged particle and neutron selection in identifying transitions in the neutron decient nucleus 62 Ga from the 31 40 28 reaction Ca+ Si, via the pn evaporation channel. It is clear that the signal to 92

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 identied, gamma-gamma coincidence data such as that shown in gure 4.22 can be used to work out the decay sequence of the transitions to form a decay scheme.

93

Figure 4.20: Particle and neutron identication spectrum showing dierent residual channels in the reaction 28 Si+40 Ca from PEX [106].

94

Figure 4.21: Schematic of the PEX array, used in conjunction with a small, silicon inner ball and a wall of neutron detectors.

95

Figure 4.22: Alpha-neutron gated coindicence gates showing transitions in 62 Ga from the reaction 28 Si+40 Ca taken from PEX [106]. The matrix was gated on the condition that one -particle and either one proton and/or one neutron were also detected.

96

Determining Charged Particle and Neutron Multiplicity.

Figure 4.23: Eciency corrected spectra used to identify states in 61 Co. (a) The pre-scaled singles data, (b) proton (c) protonneutron and (d) the coincidence spectra from the 16 O+48 Ca 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 specic 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 identication spectra for the reaction 16 O+48 Ca gated on charged particles and neutrons taken from the U. Penn 4 array [110, 25]. The main channels of interest were the pxn channels to the 97
62

Co and

61

Co channels

via the pn and p2n channels respectively. The ratio of counts for lines between the 1p and 1p1n gated spectra are shown in gure 4.24 and the lines from the dierent neutron multiplicities are clearly separated.

Figure 4.24: Plot of ratio of relative intensities of states in protonneutron and proton spectra, identied in (a) the 16 O+48 Ca and (b) the 18 O+48 Ca reactions. Note how the neutron multiplicities, used to identify the residual decay products are clearly separated and the eect of diering Q-values on the neutron detection eciency in the two reactions [25]. Figure 4.24 also shows the same plot for the evaporation channels for the same target but using an 18 O reaction. Note that the apparent neutron eciency is dierent for the 16 O induced reaction. This is a kinematic focussing eect due to the dierent Q-values for the two reactions. Thus the recoil cones of the neutrons in the lab frame for the two reactions are dierent. In the 18 O case, the evaporated neutrons typically have higher energies than those in the 16 O 98

reaction. Thus the neutrons from the 18 O reaction are less forward focussed in the laboratory frame. Since the neutron detector was placed at forward angles, their detection eciency is reduced relative to the
16

O data.

99

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 identication of transitions in channels produced with low cross-sections ( 10b). 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 identication of very weak channels due to the large dierences in for example, neutron evaporation probability between isotopic compound nuclei diering by only a few neutrons. A more direct way of determining which nucleus specic gamma-ray transitions 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) ux of beam particles. In addition, there may be other contaminant nuclei which one would like to remove such as products from fusion-induced ssion or deep-inelastic type reactions. The rejection of non-evaporation type events is usually achieved by passing the particles through some form of electric and magnetic elds which separate the ions A by their mass over charge state ( Q ). There are usally two electro-magnetic elds, one to do the main job of dumping the beam and the second one to give disperA sion in the focal plane with the Q of the evaporation residues. There is usually a position sensitive detector (such as a PPAC or microchannel plate) at the focal A point and the position of the recoil on this detector yields information on its Q ratio.

4.4.1

Recoil Mass Separators.

The Daresbury recoil mass separator [143] used two Wien lters (crossed magnetic and electric elds) to separate the beam from the evaporation residues. Crossed E and B elds are velocity lters and will only allow through particles with a certain velocity. Since the recoils are slower than the beam particles, setting the elds so that the residues passed through the Wien lter means that the faster A beam particles are deected. The nal Q separation was achieved using a dipole magnet which dispersed the recoils along a focal plane where they are detected

100

using a microchannel plate. Three more modern devices are the FMA or Argonne Fragment Mass Analyser [144] (see gure 4.25), the Oak Ridge RMS [145] and CAMEL [146] at the INFN, Legnaro, Italy. They work on the basis of two electric dipole elds which give the beam separation with a magnetic quadrupole magnet in the middle to give steering and focussing.
AYEBALL ATLAS
Target Chamber Neutron Detectors Dipole Electric Field

F.M.A
Electric Field P.P.A.C Beam Stop

Magnetic Field

Dipole

Tandem

Linac Ion-Chamber Ge Detectors

Figure 4.25: Schematic of the AYEBALL array used in conjunction with an array of neutron scintilator detectors, the Argonne Fragment Mass Analyser and a split anode 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 nal 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 identied at the nal focus. It is usual in experiments using recoil separators to initially perform a charge state sweep, where the eld settings of the electro-magnetic devices are set to the charge state which will maximise the yield of a specic isobar. Figure 4.26 shows the results of a charge state sweep for A=87 recoils for the reaction 36 Ar+54 Fe from the FMA. Clearly, there will be a susbantial loss in detection eciency due to the fact that not all ionic charge states can be measured. However, such devices do allow very clean isotopic identication 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 24 Mg+40 Ca. Clear identication of dierent isobaric species is allowed. Note the presence of two charge states on the focal plane. By setting gates around these mass locii in oine sorting, one can associate gamma-ray transitions with specic recoils. Figure 4.28 shows the results of these 101

Figure 4.26: Results of a charge state sweep for A=87 recoils for the reaction 36 Ar+54 Fe [29]. mass gates for the
24

Mg+40 Ca reaction and there is a clear dierence 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 sucient 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 identication spectra as shown in gure 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 ssion background, it is often possible to determine the Z of the specic recoils by the energies of the co-incident X-rays (associated with electron conversion decays) observed in the gamma-ray identication spectra [48, 45, 46]. However, for lower-Z recoils, where the X-ray energies are much decreased, a dierent 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 rst part of this chamber 102

A/Q

61

Q=11
60 59 62

Q=10

57

58

59

Figure 4.27: Argonne FMA two-dimensional spectrum of X-position versus energy loss in the PPAC for the reaction 40 Ca+24 Mg showing clear separation of the various 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 dierent Z will deposit dierent amounts of energy in the rst part of the ion-chamber [154, 155]. The ion chamber associated with the Daresbury recoil separator [143] was used very successfully to identify the rst excited states in very neutron decient N=Z nuclei from 64 Ge upto 84 Mo [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 40 Ca+24 Mg from the AYEBALL array with the FMA. There is a clear 64 separation between the E signals associated with the 58 Ni and 64 Zn ( 11 5810

recoils from the 2p ands 2p2n+44 Ca channels resepctively.

103

Figure 4.28: Raw, mass gated spectra from the reaction

40

Ca+24 Mg [29].

By gating on these dierent portions of the ion-chamber E spectrum in oine sorting and performing normalised subtractions of contaminants to deconvolute the dierent overlapping channels, one can achieve the isotopically pure identication spectra shown for 58 Ni in gure 4.31.

104

Figure 4.29: A=58 gated spectra from the reaction 40 Ca+24 Mg with and without subtractions from higher mass contaminants [29].

Figure 4.30: Rotated E ion chamber signals gated on known transitions in 58 Ni and 64 Zn. The separation by the Z of the recoil is clearly demonstrated [29].

105

Figure 4.31: Gamma-ray spectra from the AYEBALL array for the reaction 40 Ca+24 Mg showing the eect of FMA and ion-chamber gating. The transition A from the Q = 5.8 recoils 58 Ni and 64 Zn are clearly separated [29].

106

4.4.2

Gas Filled Separators.

The main problem with recoil separator type devices is that since they transmit the recoils in more than one charge state, they have a limited transmission eciency ( 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 ssion background, the transmission eciency can be dramatically improved by using a gas lled separator such as RITU (see gure 4.32 [147]).

Figure 4.32: Design drawing of the RITU gas lled 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 nal 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 ll this eld region with a dilute gas such as helium. Collisions between the reaction products and the gas atoms lead to a charge state 107

focussing eect. 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 lled separator is given approximately by the expression [147, 148] B = mv mv A = 1 Tm 1 = 0.0227 eqav ev 3 Z3
v0

(4.4.13)

where mv is the momentum of the reaction product and vo is the Bohr velocity (=2.19106 ms1 ). Typical gas pressures of about 1mb of helium are used.

4.4.3

Recoil Decay Tagging.

If the neutron decient nucleus of interest has a substantial decay width by either -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 gure 4.33.
AYEBALL ED MD beam implanted recoils FMA ED PPAC DSSD

8.2m

0.4m

target

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 nal 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 (or proton) decay of the recoil. 108

Figure 4.34: Proton and alpha-particle energy spectra from the decays of the products from the 92 Mo+58 Ni reaction at a beam energy of 260 MeV [161]. Since and direct proton decays have discrete energies, these are denitive signatures of a particular nucleus and can be correlated with specic 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 or direct proton decay. Using this method, in-beam spectroscopy of neutron decient nuclei with fusion-evaporation cross-sections as low as 10 b can be achieved [158]. Figure 4.36 shows the comparison between the recoils mass gated and RDT gated coincidence spectra for very neutron decient Rn isotopes. The extra degree of channel selection aorded by gating on the alpha-decay of the compound nucleus allows this weak channel to be identied.

4.4.4

Recoil Filter Detectors.

Another way of improving the signal to noise and measured resolution of gammaray 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 ight for the recoils (and other reactions products such as scattered beam, ssion products, etc) to reach the lter detector. 109

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 of 147 Tm 2 (c) the proton d 3 isomeric state in 147 Tm [161].
2

The (slower moving) evaporation residues can be selected in o-line anaylysis by gating on this time of ight signal. Note that if the target/detector distance is known, this time of ight gives a direct measure of the recoil velocity. This coupled with the angular information obtained by knowing which recoil lter 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 ssion products is a large problem [47]. In these cases, ssion can make up 99% or more of the total fusion yield.

110

433 keV
6

(2+ 0+)

selected using recoils alpha decay

counts

4 2 0

504 keV
(4+ 2 +)

500 400
176

Hf Coulex

selected using recoil mass only (for comparison)

counts

300 200 100 0 100 200 300 400 500 600 700 800

gamma energy (keV)

Figure 4.36: Comparison of FMA- projections for (a) no recoil condition and (b) using RDT using the reaction 176 Hf+28 Si at a beam energy of 142 MeV [159].

111

gamma-ray array

RFD

pulsed beam

1 0 1 1 1 0 0 0 1 1 0 0 11 00 1 0 1111 0000 1 0 1111 0000 1 0 1 0 1 0 11 1 1 00 0 0 1 0 11 1 1 00 0 0 1111111111111111111111111 11 0000000000000000000000000 00 1111111111111111111 0000000000000000000 1 0 11 00 1 0 1111111111111111111 0000000000000000000 1 0 11 00 1 0 1111111111111111111 0000000000000000000 1 0 1 1 0 0 11111111111111111111 1 1 00000000000000000000 0 0 1 0 11111111111111111111 1 1 00000000000000000000 0 0 1 0 1 0 11111111111111111111 1 1 00000000000000000000 0 0 1 0 11111111111111111111 1 1 00000000000000000000 0 0 1 0 1111111111111111111 0000000000000000000 11 00 1111111111111111111 1 1 0000000000000000000 0 0 1 0 evap. 11 1 00 0 11 1 00 0 1 0 1 0 residues.

scattered beam fission coulex

beam

100 Time of Flight (ns)

evap. residues
200
112

Figure 4.37: Schematic of the design and operation of the recoil lter detector which uses time of ight to distinguish between evaporation residues and unwanted background events.

Chapter 5 Measurement of Lifetimes of Bound 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 is related to its intrinsic width by the Heisenberg uncertainty principle such that h (5.0.1)

The probability of decay is proportional to the intrinsic energy width and depends soley on the matrix element between the initial and nal states and the operator which governs the decay between them, such that [202] | < f |M|i > |2 (5.0.2)

where M is the operator for the decay and f and i are the wavefunctions of the initial and nal states respectively. The lifetimes of the decay thus reveal information on the nature of the states. The lifetime can be compared with Weisskopf single particle estimates to help deduce the spin dierence between the initial and nal states, or if this is already known (from for example angular distribution/correlation data) can be used to deduce other eects 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

transitions probability from a state Ji to a state Jf (summed over all possible magnetic substates) by a transition of energy E is given by [7] 8(L + 1) Tf i (L) = hL ((2L + 1)!!)2 E hc
2L+1

B(L : Ji Jf )

(5.0.3)

Measuring the lifetime (decay probability) of a nuclear state thus gives a value for the B(L : Ji Jf ). For lifetimes, in units of seconds where the transition probability per unit 1 second, T = , (E in MeV),
3 T (E1) = 1.587 1015 E B(E1)

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

(5.0.4)

5 T (E2) = 1.223 109 E B(E2)

(5.0.5)

7 T (E3) = 5.698 102 E B(E3) 3 T (M1) = 1.779 1013 E B(M1) 5 T (M2) = 1.371 107 E B(M2)

(5.0.6)

(5.0.7)

(5.0.8)

7 T (M3) = 6.387 E B(M3)

(5.0.9)

The units of B(E) are e2 fm2 and the units of B(M) are (e 2Mc)2 (fm)22 . h Due to the long range involved in the lifetimes of nuclear states, dierent 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 1015 103 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 xed multipolarity. The Weisskopf single particle estimates are based on a 114

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(M) values into the equations for the transition rates given in the previous section. The general equations for the single particle estimates for the reduced probability matrix element are [35], B(Wu :EL) = for electric transitions and 10 2L2 3 1.2 L+3
2

1.22L 4

3 L+3

A 3 e2 fm2L

2L

(5.0.10)

B(Wu :ML) =

A2L2 2

e h 2Mc

fm2L2

(5.0.11)

for magnetic ones. M is the single nucleon mass and A is the atomic mass number. Transition Multipolarity E1 E2 E3 E4 M1 M2 M3 M4 T 1 (1 spu) (seconds)
2

3 6.76 106 E A 3 4 5 9.52 106 E A 3 7 2.04 1019 E A2 8 9 6.50 1031 E A 3 3 2.20 105 E 2 5 3.10 107 E A 3 4 7 6.66 1019 E A 3 9 2.12 1032 E A2

Table 5.1: Weisskopf single particle estimates for transition half-lives [190, 191]. Note, E is in keV and A is the mass number. Typcal recommended upper limits rate for the dierent types of transitions can be found for the dierent mass regions in references [192, 193, 194, 195].

5.0.6

Determining Nuclear Quadrupole Deformation from Lifetimes 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

5 2 Q | < Ji K20|Jf K > |2 (5.0.12) 16 o where < Ji K20|Jf K > is a Clebsh-Gordon coecient for the transition decaying from a state of spin I to one of I 2 is given by B(E2) = 3(J K)(J K 1)(J + K)(J + K 1) (2J 2)(2J 1)J(2J + 1)

< Ji K20|Jf K >=

(5.0.13)

Thus by susbtitution, the lifetime is related to the quadrupole moment by the expression [53], 1 5 5 = 1.223E Q2 | < Ji K20|Jf K > |2 (5.0.14) 16 o The quadrupole moment can be related to the quadrupole deformation parameter 2 by the expression [108], 3 1 Q0 = ZR2 2 1 + 8 5

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 2 ....

(5.0.15)

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 (given that there were N0 at t = 0), is given by t (5.1.16) If the lifetime of the nuclear state we wish to measure is long compared to N(t) = N0 exp 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 gure 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

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 dene a population at time t = 0. The beam is then switched o 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 dierence 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 tted to the radioactive decay law to obtain the lifetime of the state.

beam intensity

beam off period (delayed gammas) stop clock beam pulse (prompt gammas) start clock next beam pulse, reset clock

Log N

t=0

prompt, in beam gammas (in prompt peak) isomer, T1/2 > few ns delayed, out of beam gammas (in delayed portion of TDC/TAC)

time

Log N
time

Figure 5.1: Schematic using pulsed beams to measure nuclear lifetimes electronically. 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 =6s,
2

K = 8 isomer in

138

Gd [93]. Figure 5.3 shows the gamma-gamma coincidence 117

data for 138 Gd lines observed out of beam. The rst 4 decays in the yrast band are observed, together with the 583 keV transition which decays out of the isomer. Note that the 10+ 8+ 616 keV decay previously observed in prompt decay studies [189] is not present.

138Gd 102

counts
101 100

10

20 time [ s]

30

Figure 5.2: Decay curve following the decay of the T 1 =6 s, K = 8 isomer in 2 138 Gd [93].

118

138Gd 200

T1/2 = 6 s

221

(8-)

2233.1

8+ 555.6 6+

.2 583
1649.9 1094.3 489.1 4+ 605.2 384.4

counts

100

489

<---- 384 (gate)

556

0 0 200 400 600 energy [keV] 800 1000

Figure 5.3: Out of beam, gamma-ray coincidence spectrum showing transitions in coincidence with the 384 kev 4+ 2+ transition in 138 Gd [93].

119

583

2+ 220.8 220.8 0+ 0.0

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 212 Fr, which decays by an E3 transition with a meanlife of 343 s [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 s, this is possible using a pulsed beam (with the beam pulses separated by the order of 2 s) by allowing each gamma-ray detector to have its own individual time signal with respect to the beam pulse (or other xed time reference such as an RF signal).

Figure 5.4: Single TDC and TDC dierence spectra for reaction 11 B+176 Yb using a pulsed beam of width 1 ns separated by 1400 ns [292]. Note the clear separation between in-beam and out-of-beam event and the well dened regions correlating to co-incidences above and below isomeric states.

120

The upper portion of gure 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 eect can be corrected for in o-line software analysis.

Figure 5.5: (a) Spectrum of earlies obtained from a sum of transitions below the 12+ isomer in 106 Cd. (b) Out-of-beam,prompt- spectrum gated on the 633 keV 2+ 0+ transition in 106 Cd [3]. One can software gate on the various portions of this spectrum to dene 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 gure 5.5). 121

(28+)

1
15584 1857.4

(29+)

4
15861 + X 795.6 15065 + X 733.0

(28+) 1529.0

(27+) 718.8 (26+) 13726 (26+) 1382.3 663.3 1677.6 (25+) 639.7 (24+) 1211.4 24+ 12049 (23+) 1487.6 572.5 (22+) 1077.2 504.4 22+ 10561 (21+) 502.6 (20+) 941.5 1310.6 20+ 1150.6 438.5 (19+) 9250 (18+)
.6 22 16

1452.1 14322 + X

13614 + X

1302.6 12951 + X 12311 + X 571.0 1143.8 11740 + X 11168 + X 1007.0 10664 + X 2110161 + X 842.4 9722 + X 403.6 9319 + X 1466.0 198884 188411 201465.8 10350 1591.4 23-

3
11941

2
9877

18+ 980.8 16+ 892.3

8100 7480 7119 621.9 6858 757.7 17-

1366.5 1290.1 7518 161253.6 1145.3 155987 6101 5912 659.7 304.9 4816 542.6 5558 592.8 223.6 4436 4660 119335.7 4324 645.6 3788 104106 889.8 861.3 330.5 5770 517.7 5253 135214 124967 1050.4 1008.6 414.2 6264 145976 7121

14+ 807.9 12+ 602.4 10+

6227

5419

5573 (13+) 319.9

695.3 8+ 4183 8+ 71284.5 + 8 32+ 6+ 2371 1009.4 1716 4+ 906.0 540.9 2504 6+ 1076.7 1028.4 9+

10+

= 90 ns12+
315.0 4121

433.4 7+ 3354 862.3

598.3 3678 269.1 311.6 3507 8- 171.1 187.6 703.3 + 3409 8 3367 2.8 6633.9 3319 42 1295.9 3084 3044 7+ 917.6 780.6 4 690.5 874.7 7. 592.5 82 552.8 754.2 52629 2492 524.2 5+ 2331 225.8 2105 1134.8 997.7 836.4 610.6 1494 861.3 1471.5

2+ 632.8 0+ 0

633

106Cd 48 58

Figure 5.6: Decay scheme for 122

106

Cd [3].

By taking the time dierence between two detectors, one can create the type of time dierence spectrum shown in the lower portion of gure 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 dierence 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 dierence 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 106 Cd [3].

104 103

a)

145, 170 or 194 keV (start) 131 keV (stop) 222(8) ns

counts

102 101 100 106 b) 105 145,170,194,217,236 and 710 keV

200 400 time [ns]

600

800

counts

104 103 102 101 0 500 1000 time [ns] 1500 1950(150) ns

Figure 5.7: Time spectra for isomeric decays in 175 Ta (a) shows the lifetime of the 9 [514] isomer in 175 Ta (TDC dierence spectrum) (b) shows the lifetime of 2 the 21 isomer in 175 Ta (pulsed beam TAC singles spectrum) [87]. 2 Using the gamma-gamma-time coincidence techniques, one can gate on any 123

two gamma-ray energies in a gamma-gamma-time-dierence cube (3D-coincidence matrix) and project the time dierence between any two gamma-rays. Figure 5.7
9 shows the time dierence spectra gated by transitions above and below the 2 isomer in 175 Ta [87]. The upper portion of gure 5.7 is a TDC dierence spec9 trum gated on transitions above and below the 228 ns 2 isomer, while the lower spectrum is a singles TAC spectrum showing the lifetime of the 21 , 1950 ns 2 isomer [87]. For reference, the decay scheme of 175 Ta is shown in gure 5.8.

(45/2+)

5037

(45/2+) 346.0

4966

(43/2+)

740.0

4661

(41/2,43/2) 305.8

4635

(43/2+) 337.4

684.0

4619

41/2+

714.0

4297

(39/2,41/2) 288.0

4329

41/2+

662.9 325.5

4282

(37/2,39/2) 39/2+ 686.2 3947 (35/2,37/2) 37/2+ 659.2 3611 35/235/2+ 632.3 3287 33/2+ 293.7 33/2+ 606.7 2979 31/2+ 29/2+ 253.6 29/2+ 27/2+ 25/2+ 23/2+ 250.1 21/2+ 19/2+ 228.8 17/2+ 15/2+ 196.9 13/2+ 11/2+ 9/2+ 7/2+ 154.3 284.3 129.6 130 0 131.4 373.9 176.9 331.6 284 11/29/2314.5 144.6 131 461 443.0 214.4 411.3 658 13/2+ 872 490.6 240.5 469.3 1101 1341 526.0 258.8 509.0 1592 563.1 2394 27/2+ 25/2+ 204.1 23/2+ 1850 21/2+ 19/2+ 17/2+ 379.8 175.5 318.6 142.7 98.8 1793 1651 1552 23/221/2256.9 19/217/2216.6 619 15/213/2170.0 276 9/25/2410.4 193.6 363.9 446 493.4 236.4 453.1 856 640 1093 559.9 271.1 1969 25/2288.7 1621 528.3 482.9 229.4 543.7 433.7 2118 2173 2402 27/2605.3 295.7 584.6 1909 569.1 275.1 31/2+ 584.5 2681 528.2 2656 29/2309.6 2205 2931 31/2627.9 310.4 619.0 2515 318.3 2825 3225 (33/2-) 3457 279.1

4041

39/2+

641.1 316.3

3957

235 .7

3762

( 2ns)
37/2+ 3526 618.7

3640 302.1

( 0.5ns)
310. 4
35/2+ 3216 3231 351.9 (31/2-)

588.8 286.6

3338

33/2-

632.0

3143

646 .8

(31/2-)

678.8

4 6. 33
694.0 2879

33/2+ 31/2+

555.9 269.5 520.4 251.3 483.3 232.3 444.4 212.4 403.7 191.0 357.1 165.7

3052

29/2342.5 27/2677.2 334.5 25/2325.2 23/2635.9 311.2 21/25 216. .3 473

2782 2531 2298 2086 1895 1729

2537

29/2+ 27/2+

659.5

2202 25/2+ 1877 23/2+ 21/2+ 1566

4 163.

0.9ns

5ns
680 894. .3 1090.6 2

1950ns
(15/2) 1279

.9 7 458 695. 12.1 9

1350

70 9. 6

( 1ns)

932.4

4 3. 83

5/2+[402]

9/2-[514] 222ns

71.9

7/2+[404]

175Ta 73 102

1/2-[541]

123.4 51.5+X

Figure 5.8: Partial decay scheme for

175

Ta [87].

5.2

The Recoil Distance Method.

In order to measure lifetimes in the region between 108 1012 s the most commonly used method is the Recoil distance or plunger method [202]. The nuclei are formed in a thin (500 g/cm2 ) target usually by a fusion-evaporation reaction. The recoils then y out of the thin target towards a thick, stopper or plunger where they are stopped.

124

The Recoil Distance Doppler-Shift Method (RDM/RDDS)


Target v Stopper v ~ 1-2 % c

Detector

u: unshifted s: shifted

Eu

Es = E u (1+ v/c cos( ))

Decay Curve

Figure 5.9: Schematic of the Recoil Distance Method. The idea behind this technique is shown graphically in gure 5.9. The premise of the RDM is to measure the dierence in the intensity of gamma-rays decaying either in ight of when stopped in the plunger as a function of target-stopper distance. Gamma-rays which are observed in a detector at angle to the recoil direction will have their energies shifted by the Doppler eect such that to rst order, Es (, t) = Eo 1
v c

v cos c

Eo 1 +

v cos c

(5.2.17)

125

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 v (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. targetstopper distance) and 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 = No exp d v (5.2.19)

Since v and d are known, the lifetime can be determined using R where, R= Io d = exp Io + Is v (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 eect parallel plates).

126

Figure 5.10: Schematic of the NORDBALL plunger taken from [228] .

127

The velocity of the recoil can be calculated directly from the dierence 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
N i1 dNi (t) = Nj ij Ni Ni i dt j=1+1 j=1

(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 rst 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 tting 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 Bateman equations. In practice, this is very dicult to achieve and can give rise to erroneous values for the lifetimes. The advent of high-eciency gamma-ray arrays 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 110 Cd [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

lines with stopper distance The decay scheme for this nucleus and the ts to the lifetime data are shown for reference in gures 5.12 and 5.13.

129

Figure 5.11: Coincident plunger spectra gated on the 10+ 8+ , 335 keV transition in 110 Cd. Taken from Piiparinen et al. [229].

130

S
7325 18 +

C
6181
115
932

1224

15 6100 < 1.5 ps


1075

16 + 5676 5857 14 +

E
14 +
968

A
4931
1107

B
5093 4.7 ps 12
911

5249 < 2.0 ps 12


1076

787

5026 2.0 ps
854

14 +

150

13

4889 2.0 ps
811

12 +

F
4620 10 +
1433

4182 1.5 ps
755

10

4173 3.0 ps
827

11

4172 12.0 ps
561

12 +

3824 5.0 ps
768

8.6 ps 3428
531

8
548

339 2540 5

2480 < 3 ps
998

6+
1335 708

2251
467

626

397

290 3056 8 2896 159 177 6 236 2660 219 356

3346 71 ps 2879

4+ 2+ 0+

938 1542 < 3 ps 885 4+

1783 1473

1126

110 48

Cd 62

658 9.2 ps 658 0

2+

0+

Figure 5.12: Decay scheme for

110

Cd by Piiparinen et al. [229].

131

707

3611 10 + = 670 ps 265 3440 171 + 335 < 4 ps 8 164 3275 9 8+ < 4 ps 467 7 750 ps 795
960

802

10
478

637

4078 1.0 ps

10 +

424

2877

3187 80 ps 6+

8+

563 399

399

1117

100Mo(13C,3n)110 Cd
10 4

E = 44 MeV

= 143

10 4

= 12.0(6) ps

= 71(4) ps

Number of Counts

10 3

9-(3345) 7Gate: 795 + 335

10 4

10 2

12+(4172) 10+

Gate: 658 + 885 + 938 + 399

10 3

= 9.2(6) ps

10 4

Gate: 561

10 3

10+(3611) 8+

2+(658) 0+
Gate: 795 + 885 + 811

10 3

= 670(35) ps

10 2

10 2

Shifted peak Unshifted peak Fit 10 100 1000

10

100

1000

Distance m

Figure 5.13: Fits to RDM data for

110

Cd from Piiparinen et al. [229].

132

5.2.2

The Dierential Decay Curve Method.

The usual method of analysis of high-spin data in RDM experiments is to t the lifetimes of state including the eect of feeding from high spin states using the Bateman equations. The individual lifetimes in a cascade are generally calculated using a 2 minimization program. This works well when the nature and intensity of the levels feeding the state of interest are well known, however, if this information is less well known, systematic errors can be introduced given erroneous values for the calculated lifetimes. The Dierential Decay Curve Method [231, 232, 218] allows one to eliminate the eects of the feeding lifetimes for the transition by gating from transitions above the transition of interest (see gure 5.14).
Ys Y (gating transition) (detemines feeding) Yu

sidefeeding

Zu Z "direct feeding transition" Zs

level, i X decay from level of interest

Xs Xu

Figure 5.14: Schematic of the gating in a cascade for use with the dierential decay curve method. The basis of the DDCM is that the Bateman equations can be reduced to the expression [232, 218], i = Ni (t) +
h bhi Nh (t) d(Ni (t)) dt

(5.2.22)

where i 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)) is the dierential dt of Ni with respect to time at time t.

133

If there are three transitions in cascade, Y ,Z and X, (see gure 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).

Lifetimes in the SD band of 194 Pb


300 213 keV 256 keV 298 keV 22 m 0

counts

200 0 300

31 m

97 m 0 200 240 250 290 290 Energy [keV] 320

decay curves (shifted)


counts

400 300 200 100 0 150 50 100 0 0 20 40 60 80 d[m] 100

298

256 213

Decay out ~ 15%


R. Krucken et al. PRL 73, 3359 (94)

Figure 5.15: Decay curves for the decay of the SD band in This is given by [232, 218] (ti ) = IXs (tk ) IZu (tk )
d(IXs (tk ) dt

194

Pb [219].

(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

(stopped) and shifted components of the lineshape respectively. = IXu + IXs IZu + IZs (5.2.24)

135

and d(IXs (tk )) IXs (tk+1 ) IXs (tk1 ) = dt tk+1 tk1 (5.2.25)

d and t = v . 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 ight-time dierences are important in the analysis, the absolute target-stopper distance is not involved, thus a systematic error in this measurement does not aect 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 ( 1012 sec) then the RDM can no longer be used to determine the lifetime (as the targetstopper 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 to the detector is given by Es where v Es Eo 1 + cos (5.3.26) c 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 ( 1015 s) levels will have hardly slowed down in the target/backing material when the gamma-ray is emitted, while decays from states with slower lifetimes ( 1012 s) 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 ( ), where F ( ) = 1 vav = vo v0
0

v(t)exp

t dt

(5.3.27)

136

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], Northclie and Schilling or Ziegler [205]. The stopping power is usually separated into two eects, 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 proles [207, 208, 209].

137

1400

1000

132 Ce band 1 o 31.7 + 37.4o

600

Counts

200

-200 1400

1000

132 Ce band 1 o o 142.6 + 148.3

600

200

-200 700

900

1100

1300

1500

1700

1900

2100

Energy (keV)

Figure 5.16: DSAM spectra for forward and backward angles for the yrast superderformed band in 132 Ce taken from GAMMASPHERE [56].

138

a) 1.0

b)

c)

0.8

F()
0.6
132 131

Ce 1 Ce 1

132 132

Ce 2 Ce 3

131

Ce 2

0.4 0.7

1.2 1.7 Energy (MeV)

0.7

1.2 1.7 Energy (MeV)

0.7

1.2 1.7 Energy (MeV)

Figure 5.17: Measured and calculated fractional Doppler shifts for hte highly deformed bands in 131,132 Ce [56].

5.3.1

Lineshape Analysis.

Note that the lifetime of the state can also be gained directly from the gamma-ray lineshape ( dN ) since this gives directly the velocity distribution of the recoils as dE they emit the gamma-ray. t 1 dt (5.3.28) exp A knowledge of the slowing down process or stopping power of the recoil, dv , dt thus allows the lineshape to be calculated for a given value of . (Note that since dN(t) = Es and v have a linear relationship, it is simple to translate from the measured dN to the calculated dN , the dN = dN . dv ). dE dv dt dv dt DSAM with Thin Targets for Very Fast Decays. In the case of very fast transitions (1015 s) such as those observed in the superdeformed bands in the A80 region the eect of the recoil slowing down in 139

197

Pb (1) 600.0 400.0 200.0 0.0 300.0 446 keV 467 keV 200.0 90
o

1000.0 600.0 400.0 200.0 0.0 400.0 403 keV 200.0 200.0 0.0 300.0 35 +50
o o

500.0

counts

100.0

100.0

1000.0 0.0

0.0 600.0 400.0 200.0

0.0 600.0 400.0 200.0 0.0 455.0 130 +145


o o

500.0

0.0 395.0

405.0

415.0

0.0 435.0

445.0 Energy [keV]

455.0

465.0

475.0

Figure 5.18: DSAM lineshape ts to transitions in the M1 band of

197

Pb [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 dierence 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
87

Nb from

GAMMASPHERE [167]. For the highest spin transitions there is a clear dierence in centroid for the forward and backward angle detectors. From this shift, an F ( ) value can be deduced from which an estimate of the lifetime (and thus quadrupole moment) can be obtained.

140

(a)

(b)

Figure 5.19: (a) DSAM spectra for thin target data on the SD band in 87 Nb. (b) F( ) curves for the superdeformed and normal deformation bands in 87 Nb [167].

141

Chapter 6 Measurement of Magnetic Moments.


The magnetic dipole moment, , is dened 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, , and the spin of the state by the expression [1, 235], (6.0.1) I 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, g= independent of the nuclear core. For a single, independent nucleon, the single particle magnetic dipole moment, can be calculated using the expression [233], gl [l(l + 1) + j(j + 1) s(s + 1)] + gs [s(s + 1) + j(j + 1) l(l + 1)] (6.0.2) 2(j + 1)

Substituing j = l s and s = 1 , the Schmidt values for g-factors for pure 2 142

protons and neutron orbitals can be calculated. For spherical shell model orbitals where j = l + 1 (eg. p 3 , d 5 ...) 2
2 2

g=

1 1 1 gl gs + j j 2 2
1 2

(6.0.3)

and for orbitals with j = l g=

(p 1 , d 3 , g 7 ...),
2 2 2

1 1 3 gs + j + gl j+1 2 2

(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 Neutrons Orbital g-factor h 11 0.21 2 g9 0.24 2 g7 +0.21 2 d5 0.33 2 d3 +0.44 2 s1 1.76 2 h 11 +1.17 2 g9 +1.27 2 g7 +0.72 2 f5 +0.54 2 d5 +1.38 2 d3 +1.33 2 p1 0.23 2 s1 +2.90
2

Protons

Table 6.1: g-factors for intrinsic spherical orbitals [234]. In reality, most nuclear states are not pure single particle states but superpositions of a number of dierent congurations 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

6.1

Measurement of Nuclear Magnetic Dipole Moments.

If a magnetic eld is applied to a nucleus with magnetic dipole moment, , 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 eld. The rotational frequency of this precession, is given by the Larmor frequency, L where [1] L = B N B = g hI h (6.1.5)

110Cd 1.2

335 keV ext field down

1.0

angular distribution

0.8

a2 = +0.31 a4 = -0.08 335 keV ext field up

1.2

1.0 = -107 + 24 mrad 0.8 0 10 20 30 40 50 60 70 80 90

Figure 6.1: Angular distributions used to measure the g-factor of the yrast 10+ state in 110 Cd (E2 decay) [236].

144

1.2

110Cd a2 = -0.23 a4 = 0.00

399 keV ext field down

1.0

angular distribution

0.8

1.2

399 keV ext field up

1.0

0.8 = -302 + 37 mrad 0 10 20 30 40 50 60 70 80 90

Figure 6.2: Angular distributions used to measure the g-factor of the yrast 7 state in 110 Cd (E1 decay) [236]. For a state with lifetime , the eect of this torque will be to shift the measured gamma-ray angular distribution by an angle (essentially because the nucleus twists around relative to the reaction plane, before it decays) where BN (6.1.6) h Thus, if the lifetime of a nuclear state is known, by measuring the shift in = L = g the angular distribution in the presence of a known magnetic eld, 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 eld up and down directions for two dierent congurations in 110 Cd. The size of the applied magnetic eld, B required to cause the nucleus to 145

precess suciently so that a rotation of angle can be measured, for a state decaying with lifetime , can thus be estimated using the expression [1], h (6.1.7) gN Thus, the shorter the nuclear lifetime, the greater the required magnetic eld 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 B value of approximately B = 2109 Ts is needed. Thus for liftimes of the order of a nanosecond, external elds 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 1012 s, required larger elds than can be obtained using such magnets, and the very large internal (or transient) elds of the ion moving through a ferromagnetic crystal have to be used. In the case where the applied magnetic eld is perpendicular to the reaction plane, the angular distribution, for a nucleus precessing with the Larmor frequency for time t, will have a, perturbed angular distribution, given by [1, 235] W (, t, B) =
k

Ak Pk (cos{ L t})

(6.1.8)

If the lifetime of the state of interest is small compared to the electronic time resolution of the system, the intergrated precession angle, ( = L ) over the lifetime of the state is measured and the angular distribution will have the form, [235] W (, t, B) =
k 0

t 1 Ak Pk (cos{ L t}) exp

(6.1.9)

6.1.1

Corrections in the Ion-Implantation Perturbed Angular 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, , is given by, = L + tr + feed , 146 (6.1.10)

(a) Static field, for t > 1ns (b) Transient field, for t~10 ps

S beam

S production target beam


1 0 0 1

stopper

1 0 0 1

thin ferro-mag. layer (gives rise to large transient B)

external static field of ~ 0.1 ->1 Tesla to provide B x T.

external mag field to provide directional polarisation of transient field

Figure 6.3: Schematic methods of the (a) static eld and (b) transient eld methods of measuring g-factors of exited nuclear states. where tr is the precession due to the transient eld, which acts on the nucleus as it slows down in the ferromagnetic medium and feed is the average staticeld precession accumulated in states which feed the state of interest. The staticeld precession of the level of interest is given by L , where N B (6.1.11) h st and Bst is the static hyperne magnetic eld, while g and are the the mean life of the level and its gyromagnetic ratio, respectively. The magnitude of the transient eld contribution may be estimated using the L = g expression [236], N Ts tr = g Btr (t)dt (6.1.12) h 0 where Btr (t) denotes the transient magnetic eld, 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 eld. 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

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 gammaray 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 eld in both up and down directions [236, 237, 238, 239, 240]. To determine the eld up-down counting asymmetry and reduce possible systematic errors, the double ratio, , is dened for a detector pair by = N+ () N () N+ () N () (6.1.13)

where N+ () denotes the number of counts in the = + detector for eld up direction, etc. For small precessions, 100 mrad, the precession angle may be obtained from = where 1 (6.1.15) 1+ and S, the logarithmic derivative of the angular distribution, is given by = S= 1 dW W () d (6.1.16) S (6.1.14)

The angular distribution has the usual form: W () = A0 {1 + a2 P2 (cos) + a4 P4 (cos)} One explicit form for evaluation of Eq. 6.1.16 is [236] 8 sin(2)[12a2 + 5a4 + 35a4 cos(2)] 64 + a2 [16 + 48 cos(2)] + a4 [9 + 20 cos(2) + 35 cos(4)] 148 (6.1.17)

S=

(6.1.18)

Therefore, if the angular distribution coecients 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 101112 s, the product of the lifetime and static eld 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 eld caused by its movement through a ferromagentic medium. The basis of the transient eld technique [235, 241, 242] is shown schematically in gure 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 experiences a strong transient eld 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 eld 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 eld strength is performed using the measurement of a state with a known g-factor. The size of the transient eld 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 eld as a function of the recoil of proton number Z and recoil velocity, v, can be given by [242, 243], v v exp (6.1.19) v0 v0 c where v0 = 137 is the Bohr velocity and and are constants obtained ts to experimentally determined transient elds [244]. BT R (v) = Z

149

Magnetic Moment Measurements Using A Plunger and the Transient Field Technique. A problem with deducing magnetic moments from the transient eld technique is correcting for the eect 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 ight between the production target and the stopper, are, as observed in a xed angle detector, shifted from their true energies by the Doppler eect. 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 eld). 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 Dierential Perturbed Angular Distributions.

For nuclear state lifetimes of between 1 ns and 1 s, it is possible to measure the Larmor frequency directly using an in-beam method. The basis of the Time Dierential 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 eld, 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 eect is caused by the nucleus precessing around in the presence of the magnetic eld, which causes the angular distribution (and thus measured intensity at a xed detector angle) to oscillate with time. The observed intensity of an observed -ray, measured at angle to the beam direction, decaying from a state of lifetime, at time t is then given by [235]

150

t W (, t, B) (6.1.20) where W (, t, B) is the angular distribution function for the state given in equation 6.1.8. I(, t, B) = Io exp A good example of the use of the TDPAD technique to measure the g-factors of high-spin isomeric states is shown in gure 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 to the beam direction
I(+)I() I(+)+I()

and plotting the intensity ratio, R, dened by R =

151

105 104
counts

214Fr

472 keV (-135 deg)

103 105 104 103


0.4 0.2
R(t)

472 keV (135 deg)

214Fr

472 keV ratio

0.0 -0.2 -0.4 0 100 200 300 time (ns) 400 500 600

Figure 6.4: Time dierential perturbed angular distribution spectra showing the oscillation in the time spectra associated with the Larmor frequency [246].

152

Chapter 7 Spectroscopy of Neutron Rich Nuclei.


The structure of very neutron decient 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 decient 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 specic cases of -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 -particles [249, 22, 25]. Such events are inhibited by the Coulomb barrier, however, the clean selection allowed by modern day charged particle detectors means that very clean charged particle gated, -ray identication spectra can be obtained. Figures 7.1 highlights the eectiveness of this technique for identifying the yrast states of the neutron rich nucleus 63 Co [25] (note that the -stable neutron rich isotope is 59 Co). The gamma-ray transitions associated with cobalt isotopes are identied by having the gating condition that an evaporated proton must be

153

(a)

(b)

Figure 7.1: (a) Singles identication spectra and (b) proton gated coincidence data on the neutron rich nucleus 63 Co33 populated via the 27 18 O(48 Ca,p2n)63 Co reaction [25]. measured. The isotope is then identied 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 nonyrast 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 Reactions.

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

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 A100 region [250] and more recently to study high-K structures in stable A180 nuclei [81].
15

(a)

179Hf

R55(F/MB)

10
178Hf

5
177Hf

0 8

R60(F/MB)

(b)

178Hf

4 0 2

177Hf

RF(60/55)

(c)

177Hf

1
178Hf

0 0

200

Figure 2.

400 E(keV)

600

800

Figure 7.2: Angular distributions of -particles observed in the reaction 176 Yb+9 Be. Note that the incomplete fusion reactions to populate higher spins in 178 Hf have a more forward focussed distribution compared to the pure fusion evaporation 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 populated in an incomplete fusion reaction, the more peripheral the collision and the more foward focussed the outgoing particle distribution. Figure 7.2 shows the -particle anisotropies for various channels observed in the 176 Yb+9 Be5 reaction 4 [81]. Note that the higher the gamma-ray energy corresponding in general to the decay of higher spin states in the rotational bands of 178 Hf), the more forward focussed the -particle anisotropy. Note that the lower spin states are probably a combination of fusion-evaporation and massive transfer, so the anisotropy eect 155

is more pronounced for higher spins. Figure 7.2 shows -particle gated -ray spectra obtained for the reaction 176 Yb+9 Be5 , which is proposed to the band 4 built on top of the =31 year, K = 16+ isomer in states by the -particle anisotropy.
178

Hf, identied as high-spin

7.3

Deep Inelastic Reactions.

For neutron-rich nuclei, it is dicult to observe the yrast sequence to high spins (such as through the backbend) due to the preferential population of neutrondecient species in fusion-evaporation reactions and the low angular momentum involved in the ssion 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 ecient 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.
Neutrons evaporated
New beam like nuclei

136

Beam

Xe
232

Th

Target
New target like nuclei

Eb ~ 10-15% above Ec

Nucleons exchanged

N:Z equilibration

Figure 7.3: Schematic of the use of deep inelastic collisions to populate yrast states 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 ow 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 specic 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

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 products 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 86 Pd+110 Pd [259].

Cd Ag = STABLE NUCLEUS Ru Tc Mo Nb Zr Y Sr Rb Kr Br Se 48 50 52 54 56 58 60 62
86 Kr 100 Mo

Pd Rh
104 Ru

108 Pd

110 Pd

66

64

= BEAM / TARGET NUCLEI

Figure 7.4: Region of the Segre chart poplated by binary collisions using the reaction 110 Pd+86 Kr [259].

Figure 7.5: Spectra from the 86 Kr+110 Pd binary reaction. Comparison of alignments between odd and even-N Mo, Pd and Ru nuclei populated in this reaction. Figure 7.5 shows the spectra for
86 104

Ru studied using the binary reaction

Pd+

110

Pd [259]. The

104

Ru lines of interest can be clearly selected, as can 157

the presence of the Sr binary partner lines. The data in gure 7.5 extends the yrast band of 104 Ru upto spin 14+ . As gure 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 eciency gamma-ray arrays [98, 271, 272]. Figure 7.6 shows the high spin cascade in 234 Th and 230 Th observed using GAMMASPHERE obtained from the binary reaction 136 Xe+232 Th [97]. Note that the higher multiplicity gamma-ray gating aorded by using the higher eciency arrays gives rise to extremely clean spectra, with only the nucleus of interest (and possibly a few binary partners) observed,
322.2

228.2

277.8

(a)
361.8 396.8 427.6

x 10 Counts

X-rays

Th

173.4 Th

285.7

454.1 476.3 493.7

Th

237.4

(b)
327.9 365.3 397.9 426.8 Xe 483.4 Xe 589.0
550 650

x 10 Counts

X-rays 120.8

Th Th

182.5

50

150

250

350

450

Energy (keV)

Figure 7.6: Gamma-ray spectra produced for the binary reaction 136 Xe+232 Th at beam energy of 833 MeV. (a) double gated spectrum showing ground stateband transitions in 234 Th. (b) spectrum showing ground state-band transitions in 230 Th. Transitions labelled Xe are cross-coincident gamma-rays attributed to 138 Xe, the projectile-like partner of 230 Th. Gamma-ray peaks marked Th in both spectra are contaminant transitions in the yrast band in Coulomb-excited 232 Th [96, 97].

158

452.3

Th

7.3.1

Maximum Angular Momentum in DIC.

Spin-wise, the use of deep inelastic reactions are not particularly ecient in generating very high spin states. (Recent studies of (discrete-line) spin-input using thin target deep-inelastic reactions can be found in reference [268]). Unlike fusion-evaporation reactions, not all of the input angular momentum of the reaction goes into the intrinsic angular momentum of the nal 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 presence of strong friction between two rolling spheres (ie. the target and beam like nuclei), known as the rolling limit, 5 of the maximum initial maximum angular 7 momentum (for a grazing collision), Lmax goes into the relative motion of the two 2 fragments while only 7 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 expressions [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 gure 7.8 which shows the fold distribution for states gated on the yrast band in 100 Mo 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. 2 7 2 7

LBLF =

1 1+
AT AB
1 3

LT LF =

1 1+
AB AT
1 3

Lmax

(7.3.1)

Lmax

(7.3.2)

159

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 , c.m. angle , M2 scatters at lab angle , c.m. angle = = 2 Lab. energy of scattered nucleus M1 is
1 2M1 M2 M2 2 M1 E1 =1 )2 {cos [( ) sin2 ] 2 }2 (1 cos) = ( 2 E0 (M1 + M2 ) M1 + M2 M1

Use only plus sign unless M1 > M2 , in which case max = sin1 ( M2 ) M1 Lab. energy of recoil nucleus M2 is E1 4M1 M2 E2 =1 = cos2 2 E0 E0 (M1 + M2 ) where Angles sin = ( M1 E1 1 ) 2 sin M2 E2 sin + cos
2

tan =

M1 M2

sin2 = cos2

M1 M2

= + sin1 (

M1 )sin = 2 M2

cos = 1 2cos2 = /2 if M1 = M2 160

Rutherford Scattering Cross Sections (in mb/sr). d Z1 Z2 M1 + M2 2 1 = 1.3( ) dcm Elab M2 sin4 cm 2 (Elab is in MeV) d dscat,lab = d sin2 dcm sin2 cos( ) = d 4sin dcm 2

d drecoil,lab Inelastic scattering.

Same notation as for elastic scattering except that lab energies of scattered beam and recoil are respectively E3 and E4 . Dene the following terms: Q = (M1 + M2 M3 M4 )c2 ET = E0 + Q = E3 + E4 E0 M1 M4 (M1 + M2 )(M3 + M4 ) ET M1 M3 E0 (M1 + M2 )(M3 + M4 ) ET

A=

B=

C=

M2 M3 M1 Q E4 (1 + )= (M1 + M2 )(M3 + M4 ) M2 ET ET M2 M4 M1 Q E3 (1 + )= (M1 + M2 )(M3 + M4 ) M2 ET ET

D=

The total scattered energy in the c.m. is E = E3 + E4 = M1 M2 (ET + Q) = Ecm + Q M1 + M2 M2 161

Lab energy of light product


1 E3 = B + D + 2(AC) 2 cos ET

Lab energy of heavy product


1 E4 = A + C + 2(AC) 2 cos ET

Angles: sin = ( M3 E3 1 ) 2 sin M4 E4

sin = ( Deep Inelastic or Fission

E3 /ET 1 ) 2 sin D

with initial kinetic energy, put M2 = 0 and Q 0.107 (A is in u and Q in MeV) so that E4 = M4 E3 [cos M3 + M4
M3 Q M4 E3

Z2 A3
1

+ 22.2

M3 Q sin2 ]2 M4 E3

Use only + sign unless

< 1, when M3 Q 1 )2 M4 E3

max = sin1 ( Compound Nucleus

Projectile M1 with lab energy E0 , target M2 as above. Lab energy of compound nucleus ECN = M1 E0 M1 + M2

Excitation energy of compound nucleus Ex = (E0 ECN ) + Q = M2 E0 + Q = Ecm + Q M1 + M2 162

Coulomb Barrier Interaction radius is


1 1 3 3 R = 1.16(A1 + A2 + 2)

where A is in u and R is in fm, then Vcm = 1.44 Z1 Z2 R

Vlab = (1 +

M1 M1 )Vcm , Elab = (1 + )Ecm M2 M2

Maximum Angular Momentum. max = 0.219 R[(Ecm Vcm )] 2 where = Grazing Angle. The grazing angle [274], where the cross-section for peripheral, deep-inelastic reactions will be maximised, is dened as the angle at which the distance of closest approach, d, given by [5, 273], Z1 Z2 e2 ].[1 + cosec( )] (7.3.3) 40 Ek 2 (where Z1 and Z2 are the atomic numbers of the two nuclei involved and E d=[ is their kinetic energy) is equal to the sum of the nuclear radii (ie. when the two nuclei are just touching), given by,
1 1 3 3 d = 1.25(A1 + A2 )fm 1

M1 M2 M1 + M2

(7.3.4)

where A1 and A2 are the nuclear mass numbers. A quick estimate of the grazing angle can be obtained using, c.m. E 1 (1 + cosec )= 2 2 V 163

Reaction Cross Section (in barns). R Velocity in % c. E A 0.67 2 Ecm max

v = 4.634

(E in MeV, A in u) For transfer reactions the two nucleon transfers are strongly dependent on the Q-value, Q = Qgg Q(Z) Z3 Z4 1 Vcm Z1 Z2 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 Q(Z) = for only neutron transfers. Qgg is the mass dierence between initial and nal products. where,

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 ight. 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 gure 7.9 gives a position sensitive signal which gives the recoil direction and uses also uses the time of ight 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

the binary reaction using a 86 Kr beam on a 116 Sn target by detecting the beam like fragment in the HIPS detector and correcting for the Doppler eect. Similarly, a recoil lter detector [47, 163] can be used to obtain similar information and Doppler corrections. The CHICO detector [276] is a large solid angle set of PPACS which uses time of ight to get some mass infomation on the fragments and allows excellent Doppler correct when used with the GAMMASPHERE array. In addition to inbeam experiments, it has been used to extend the decays schemes of a number of neutron-rich ssion products around A 100 upto discrete spins of 20 using h

a thin, ssion source [277, 278, 279]

Determination of Angular Corrections in DIC. If the recoiling beam (or target like fragment) is detected at angles 1 and 1 to the beam axis, and the (doppler shifted) gamma-ray emitted from the recoiling fragment is detected in a germanium detector and angles 2 , 2 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 (dop ) 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(1 )cos(1 ), y1 = sin(1 )sin(1 ), z1 = cos(1 ) and x2 = sin(1 )cos(2 ), y2 = sin(2 )sin(2 ), z1 = cos(2 ) respectively. The angle between these two vectors can be calculated using the vector dot product, ie. v1 .v2 = v1 v2 cos(dop ). Therefore

cos(dop ) = sin(1 )cos(1 )sin(2 )cos(2 )+sin(1 )sin(1 )sin(2 )sin(2 )+cos(1 )cos(2 ) (7.3.5) collecting the terms and recalling the trigonometric identity cos(A B) = cosAcosB + sinAsinB, the above expresssion reduces to cos(dop ) = sin(1 )sin(2 )[cos(1 2 )] + cos(1 )cos(2 ) (7.3.6)

165

Figure 7.7: Fold distributions gated on the low-lying 601 keV transition in 100 Mo, following the binary reaction 100 Mo +136 Xe at beam energies of 650, 700 and 750 MeV respectively. Note that the overall spin of the fragments increases dramatically with beam energy [281]. Data taken from the 8 array at Berkeley.

166

Figure 7.8: Fold distributions gated on the low-lying yrast transitions in 100 Mo, following the binary reaction 100 Mo +136 Xe at a beam energy of 700 MeV. Note that low-lying states are predominantly populated vua Coulomb excitation, while the higher ones have a higher overall multiplcity distribution, associated with a deep-inelastic collision, probably followed by neutron emission [281]. Data taken from the 8 array at Berkeley.

167

S.Schwebel et.al. Manchester University

Stop Detector

Ionisation Chamber

Anode HT
delay-chip

Foils & Support Grids Gate-valve Start-Detector housing

catode ygrid

xgrid

Field Rings

10M

Signal

10M

Signal

10M

Signal Anodes
10M

Signal

Frisch Grid

10M

Signal

catode ygrid

delay-chip

Grid HT

10

20

30

40

50cm
10 10M

Figure 7.9: Scale drawing of the HIPS vessel for detecting recoils from deepinelastic reactions [275].

168

xgrid

HIPS-Ge coincidence -ray spectra


86

Kr +

116

Sn

150 Raw Ge spectrum 100

50

Counts per channel

0 Doppler corrected spectrum 100

50

500

1000

1500 2000 Channel Number

2500

3000

3500

Figure 7.10: HIPS- coincidence spectra for the binary reaction 86 Kr+116 Sn highlighting the eect of Doppler correction on the detected lineshape [275].

169

Chapter 8 Spectroscopy With Radioactive Ions 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 Segr chart and shows the predicted proton and e 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.

Atomic number Z

80 60 40 20 0 0 20 40 60 80 100 120 140

Neutron number N

Figure 8.1: Segre chart showing the stable isotopes predicted proton and neutron driplines [282]. Figure 8.2 shows those compound nuclei which can be formed using stable 170

100

=Z

beam target combinations. Note, virtually all of these systems are neutron decient with respect to the stable isotopes.

100

Atomic number Z

80 Proton Drip Line 60 40 Neutron Drip Line 20 0 0 20 40 60 80 100 120 140

Neutron number N

Figure 8.2: Segre chart showing the compound nuclei which can be formed in fusion-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 important. 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], projectile 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

Method Advantages Projectile Fast delivery times ( s) Fragmentation No chemical contraints Reliable operation High collection eciency Simple target design ISOL Thick target High beam purity Low energy spread Useful beam energies

Disadvantages low beam intensity Final beam deceleration dicult Limited target thickness Large energy spread Moderate isobaric purity Decay losses due to slow release Needs post acceleration Radiation contamination in target Complex target design Yield depends on beam/target chem.

Table 8.1: Comparison of the ISOL and projectile fragmentation methods of radioactive 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 identication spectrum (time of ight verses energy loss) for the fragmentation of a 92 Mo beam on a nickel target at an energy of 60 MeV per nucleon using the LISE3 spectrometer at GANIL. Note, the particle A identication aorded in fragmentation reactions using the time of ight ( Q ) and energy loss (Z) technique means that each nuclear species produced can be individually identied. This technique of radioactive beam production has the advantages of high beam purities and high collection eciency (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

TZ "beam" different charge states of beam


52 2 32 1 12 0 / / /
80 Zr

Mo Zr Sr Kr

Z 42 40 38 36

66As

B = 1.9068 Tm

A __ = 2 q

Figure 8.3: Particle identication spectrum from the fragmentation of GANIL with no degrader or Wien lter selection [287].

92

Mo at

8.1.2

Particle Identication in Fragmentation.

One of the initial benets of projectile fragmentation is that it enables the identication 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 specic places on the particle identication plot. For example, the nuclei 81 Nb and 85 Tc are thought not to be bound against direct proton emission due to their absence in the fragmentation of 92 Mo [321]. The parameters used to identify the beam-like fragments are: (a) the magnetic rigidity of the dipoles which select the ions, B; (b) the time-of-ight, TOF,assuming that the distance, L, between the production target and the focal point, where the ions are implanted, is constant for every fragment; (c) the

173

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 eld, B. This deviation is characterized by the bending radius, . This depends on the linear momentum, p, and the charge, Q = qe. From the Lorentz law, one can derive that: B = p qe (8.1.1)

Substituting for the linear momentum, this equation becomes B = Mc , qe (8.1.2)

where M is the mass of the fragment, c is the speed of light in vacuum and
1 2 = 1 2 with = v , where v is the fragment velocity. If the mass is expressed c in atomic mass units (a.m.u.), M = Au where u is the atomic mass unit, and B is expressed in Tesla-meters (Tm), then:

B = 3.105 expressed in relativistic mechanics as:

A . q

(8.1.3)

The measurement of the total kinetic energy of an implanted fragment, K, is

K = Mc2 ( 1). implanted fragment: A= K . 931.5 ( 1)

(8.1.4)

Therefore, it is possible to directly obtain the atomic mass number, A, for the

(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 K . 931.5 ( 1)B (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-ight from the production target to the nal focus can be measured by taking the time 174

dierence 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-ight 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 Bethes formula [320]. The atomic number, Z, is calculated using the energy-loss in a E-silicon detector. Relativistic corrections to Bethes formula lead to the expression: Z2 E = a1 2 [ln (a2 2 2 ) + a3 2 + 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 where: Y = ln (5930 2 2 ) 1 2 (8.1.9) E E + c2 + c3 + c4 Y Y (8.1.8)

and c1 , c2 , c3 and c4 are constants which can be tted. The complete reconstruction of A, q and Z is therefore possible from the measurement of the time-of-ight, the energy loss, and the total kinetic energy. Figure 8.4 shows a calibrated particle identication spectrum following the fragmentation of a 60 MeV per nucleon 92 Mo 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 ight 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 identication data from gure 8.3, with and without the condition that a delayed gamma-ray must be measured with 100 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 germanium detector array. By gating on the nuclear species of interest in the particle 175

48 46 44 42 40

N=Z
Tc Nb Y Rb
4096 2048 1024 512 256 128 64 32 16 8 4 2 1

38 36 34 32 30 28 26

1.950 1.975 2.000 2.025 2.050 2.075 2.100 2.125 A/q

Figure 8.4: Calibrated particle identcation spectra from the fragmemtation of 92 Mo [320]. identication spectra one can project the gamma-ray energy and time spectrum for the isomeric decay.

176

GRZYWACZ PLOT TO ID ISOMERS 92-Mo @ 60 MeV/A on Ni target at GANIL (a) all


Rb (37) Kr (36) Br (35) Se (34) As (33) Ge (32)

(b) gammas

energy loss

Kr-73

Se-69

Tz =3/2 1 1/2

time of flight (A/Q)

Figure 8.5: Particle identication spectrum from the fragmentation of 92 Mo at GANIL. Note the extra condition of observing a delayed gamma-ray highlights those nuclei in which isomeric decays have been observed.

177

Figure 8.6 shows a two-dimensional plot of time verses gamma-ray energy taken for the fragmentation of a 92 Mo 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 gure 8.6 has been corrected for low-energy time walk [291].

4000 3500 3000 2500 2000 1500 1000 500 0 0

a)
Gammaray energy (keV)

4000 3500 3000 2500 2000 1500

b)

Gammaray energy (keV)

E = 734 keV
1000 500

500 Time Difference (ns)

500 Time Difference (ns)

Figure 8.6: Two-dimensional gamma-ray energy verses time spectra showing the decays of the isomeric states following the fragmentation of a 60 MeV/A beam of 92 Mo. [291]. Note the correction for time-walk for lower energy gamma-rays 42 in the right hand spectrum. Figure 8.7 shows the gamma-ray and time spectra associated with the decay of the 42 ns isomer in 74 Kr obtained by gating on the identication spectra shown in gure 8.5. This is a very powerful technique in identifying isomeric decays in exotic nuclei and can be used for both neutron decient [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 68 Ni40 , produced by the fragmentation of a 86 Kr beam. Note the presence 28 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 iso178

= 42 8 ns

counts

101

100

150

200

250

300

350

time difference (ns)

Figure 8.7: Gamma-ray energy and associated time spectra showing the decay of the isomeric state in 74 Kr from the fragmentation of a 60 MeV/A beam of 92 Mo. 36 42 [296, 291]. meric states in intermediate energy projectile fragmentation reactions varies signicantly, 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 eective lifetimes of the nuclei through a separator are also extended due to relativistic eects, such that through the separator, the eective lifetime is rmef f = o where is v . c 1 1 2 (8.1.10)

Fragmentation can also be used at relativistic energies corresponding to primary beams of hundred of MeV per nucleon [309, 310, 308, 312]. This is particularly useful for the identication 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

68m2

Ni

T1/2=290 ns (exp)

Figure 8.8: Time spectrum gated on 511 keV decays in 68 Ni from the fragmentation of 86 Kr at GANIL. The lifetime suggests an E0 decay proceeding by internal pair 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 ight 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
A in order to resolve charge state ambiguities in the time of ight signal ( Q , see g. 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

T (us)

function of their time of ight. Fully stripped ions do not change their position through the spectrometer, whereas hydrogen and helium like one do, causing a longer eective time of ight (see gure 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) 179 W show ing the decay of the K = 35 isomer. A prescription of how to estimate the 2 much angular momentum transfered to nucleus in a projectile fragmentation reaction 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

68m2

Ni

Figure 8.9: Gamma-ray gate on 68 Ni from the fragmentation of 86 Kr showing the 511 keV lines associated with internal pair formation [295].

181

Figure 8.10: Schematic of the set-up using the GSI Fragment Separator to investigate decays from isomeric states [315].

production target

middle focus S2 primary beam: Pb @ 1 GeV


Degrader dipole : B IC:time IC:dE,Q MW:x,y

182

TOF

Segmented Clover Array

Degrader

SCI

end focus S4

A B e = Q c u
small Clover

catcher

SCI

Super Clover
+

fullystripped
0

hydrogenlike

heliumlike

SC21x (mm)

20

40

2.56

2.58

2.60 A/Q

2.62

2.64

80
209

Tl

60
195 194

203 196

Au
202

Os

Os
200

40

Os
193

Pt Pt

201

Pt

206 205

Hg Hg
204

SC41x (mm)

20
191

192

Re
198 197

Re Ir

199

Ir
203

Ir

Au

Re
190

Au

0
189

W
194

196 195

Os
200

20

W
188

Os

201

Pt

187

Ta

Os
193

Pt

40

Ta
192

Re
2.62 2.64 A/(Z2)

Re

60

2.56 A/Z

2.58

2.60 A/(Z1)

Figure 8.11: Fragment identication spectra following the fragmentation of a 1 GeV/u 208 Pb, highlighting the dierent charge states of the ions for an FRS setting centred on 191 W ions [316].

183

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 dierent charge states. Note in particular, the 15 ns isomer in 200 Pt, which is observed due to the eective lengthening of the decay half-life through the spectrometer by the switching o of the electron-conversion component of the decay [291, 316].

Figure 8.12: Gamma-ray spectra associated with the fragmentation of a 1 GeV/u 208 Pb, for fully stripped ions around 177 Ta [313]. Elemental identication in relativistic fragmentation is achieved by a combination of energy loss (usually measured in gas detector, known as a MUSIC chamber, see g 8.16) or by position on a variety of scintilation detectors through the spectrometer.

8.1.3

Isomeric Ratios and Angular Momentum Population.

The isomeric ratio, R, is dened 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

Firstly, the observed decay yield is calculated : Y = N (1 + tot ) , ef f b (8.1.11)

where N is the number of counts in the gamma line depopulating the isomer of interest, tot is the total conversion coecient for this transition, ef f is the eective eciency and b is the probability that the decay proceeds through this transition (i.e. the absolute branching ratio). The isomeric ratio is then given by : , (8.1.12) Nimp F G where Nimp is the number of implanted heavy ions, F is a correction factor for the in-ight isomer decay losses and the factor G corrects for the nite detection time of gamma radiation. The factor F is calculated from : F = exp q1 T OF2 T OF1 + q 2 1 2 , (8.1.13) R= Y

where T OF1 (T OF2) is the time of ight through the rst (second) stage of the spectrometer (eg. the FRS), 1 (2 ) is the corresponding Lorentz factor and q1 (q2 ) is the decay constant for the ion in the charge state q1 (q2 ). In most of the experiment using relativistic ions, the in-ight ions were highly charged (in most of the cases studied, the ions are fully stripped), and the decay constants q can dier considerably from the value for an electrically neutral atom, . For the fully stripped ion, 0 can be calculated from : 0 =
i

bi , i 1 + tot

(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( ti ) exp( tf ), (8.1.15)

where ti and tf are the gamma delay-time limits set in the o-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 averaged. 185

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 denition 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) bU L to the lower one, the isomeric ratio for the latter can be calculated by : YL RU 0 L GU U GL bU L FU + 0 U 0 GL (FU FL ) , Nimp FL GL FL GL L U L U (8.1.16) where the indexes L and U refer to the lower and the upper states, respec-

RL =

tively, and the second term on the right side represents the correction due to feeding from the upper state. The Sharp Cut-O 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 nal fragment is formed with an excitation energy below the particle emission threshold. Subsequently, a statistical 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 rst 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. Assuming that any angular momentum taken away by evaporating particles is small and can be neglected, they calculated the angular momentum distribution of the nal fragment as the superposition of the angular momenta of all prefragments contributing to the nal fragment of interest using the ABRABLA code [299]. Furthermore, they have shown that for a large mass dierence between the

186

projectile and the fragment this distribution can be approximated by a simple analytical formula : PJ = 2J + 1 J(J + 1) exp , 2 2 2f 2f (8.1.17)

where f , the so called spin-cuto parameter of the nal fragments, is given by :


2 2 f = jz

Ap and Af denote the projectile and fragment mass numbers respectively,


2 is the mean number of evaporated nucleons per abraded mass unit and jz is the average square of the angular-momentum projection of a nucleon in the nucleus.

(Ap Af ) (Ap + Af ) . ( + 1)2 (Ap 1)

(8.1.18)

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 2 the energy by about 13 MeV, hence the parameter = 2 is taken. Values of jz , estimated on the basis of a semi-classical consideration of the angular-momentum distribution in the Saxon-Woods potential [300, 298], are written as : 2 2 jz = 0.16 A2/3 (1 ), (8.1.19) p 3 where is the quadrupole deformation parameter. Given the angular momentum distribution of the nal fragment, one can consider 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 excitation 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-o model, has been used in studies of angular momentum distributions in compound nuclei [302] and in ssion fragments [303]. From Eq. 10 it follows : Rth =
Jm

PJ dJ = exp

Jm (Jm + 1) . 2 2f

(8.1.20)

Substituting = 2 and introducing A = Ap Af , the above equations yield: 2 A (3Ap A) 2 f = 0.0178 (1 ) A2/3 . p 3 Ap 1 187 (8.1.21)

8.1.4

Projectile Fission

A similar method of producing high energy radioactive beams to projectile fragmentation is that of projectile ssion. Here a heavy beam, such as 238 U bombards a light target (such as 1 H). The heavier element usually ssions in such a reaction, giving rise to a large number of neutron rich fragments moving forward in the lab frame. These fragments are subsequently collected and identied using a mass separator such as the FRS at GSI, Darmstadt, Germany [304]. Using this technique, many new neutron rich isotopes have been identied for the rst time [318].

188

Figure 8.13: Predicted spin distribution for 179 W fragments in the fragmentation of a 1 GeV per nucleon 208 Pb on a 9 Be 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 exotic nuclei produced in projectile fragmentation reactions is to Coulomb excite the products on a heavy stopper (such as Pb) and measure the detected gammarays. The large fragment velocity means that Doppler broadening eects can be large and rather degrade the energy resolution. However, since the number of lines is generally quite small, and gamma-ray detection eciency is at a premium due to the relatively small number of events, a higher eciency, lower resolution detetctor, such as NaI(Tl) is used instead of germanium. Work using this technique 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
44 16 S

[305, 306].

189

Proton Number

Neutron Number

Figure 8.14: Predicted average fragments spins using the abrasion-ablation model [310] for 208 Pb on 208 Pb at 1 GeV per nucleon.

8.1.6

Double Fragmentation and In-beam spectroscopy


34

For double fragmentation at RIKEN in

Mg 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 -decay lifetimes [319, 320] (see gures 8.17 and 8.18).

8.2

ISOL Based Techniques.

The isotope separation on-line method uses a primary reaction to create the radioactive nuclei of interest (such as a fusion-evaporation reaction or the spallation of a heavy nuclei by protons or beam induced ssion). The radioactive nuclei must then diuse 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

Figure 8.15: Gamma-ray spectra associated with isomers populated following the fragmentation of a 1 GeV/u 208 Pb beam, for fully stripped, hydrogen line and helium-like ions around 191 W [316].

8.2.1

On-Line Mass Separators.

A typical on-line mass separator (see ref. [330] for a review) is eectively 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 , via a magnetic eld according to the equation, B = Auv Qe (8.2.22)

where A is the nuclear mass number, u is the atomic mass units (approx1.6 10 Kg), e is the electron charge, Q is the ionic charge state and v is the ion
27

191

Figure 8.16: Energy loss signals in the music chamber at the end of the Fragment Recoil Separator, which allows elemental identication of the fragments. These spectra are all for the fragmentation of a 208 Pb beam at 1 MeV/a, but with dierent settings of the spectrometer to transmit dierent nuclei the nal 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, 1 E = Auv 2 2 substituting in to equation 8.2.22, we obtain B = Au
2E Au

(8.2.23)

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

46 44

N=Z
Tc Nb

1024 512 256 128

42 40

64 32 16

Y
38 36 34

8 4

Rb

2 1

1.960

1.970

1.980

1.990 A/q

2.000

2.010

2.020

Figure 8.17: Particle id-spectra as in gure 8.4, but with hardware beam-o condition 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 dierent tape cycle speeds to select decays of dierent lifetimes (typically the furthest from stability are the shortest lived). The extracted ions are often sent to a tape counting station, where the gammarays from the radioactivity (with liofetimes usually longer than hundreds of milliseconds) is collected. In a tape drive system, the tape can then be caused to move with dierent frequencies, which means that the detection system becomes sensitive to dierent rates of decays. For example, if the tape moves the radioactivity 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 specic activity is given by assuming a constant implantation rate, I0 , then,

193

80 70 60 50 40 30 20 10 0

74

Rb

72(18) ms

200 400 600 800 1000

180 160 140 120 100 80 60 40 20 0

counts

counts

78

50(8) ms

200 400 600 800 1000

time (ms)

time (ms)

counts

counts

350 300 250 200 150 100 50 0 0

50 40 30 20 10

82

Nb

86

Tc

52(6) ms

45(12) ms

200 400 600 800 1000

200 400 600 800 1000

time (ms)

time (ms)

Figure 8.18: Time spectra for delayed + decay in the same strip as the detected N=Z recoil, with the beam-o condition provided by gure 8.17 [320].

dN = I0 N dt where N is the decay rate, with a decay constant, . Trying a solution of the form N = a(1 et )

(8.2.25)

(8.2.26)

where a is a constantto be determined, then dierentiating equation 8.2.26, we obtain dN = aet = I0 a(1 et ) dt therefore, I0 = a 194 (8.2.28) (8.2.27)

which gives us the general solution for a one component grow-in curve of I0 (1 exp(t) ) Thus the count rate, N = I0 (1 exp(t) ). N= Two component Grow-ins. In the case of a two component decay curve, where for example a decay feeds into a long lived isomeric state, a two component t 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 1 N1 and the lower lying state (ie. the one fed by state 1) has an intrinsic decay rate of 2 N2 , 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 = 1 N1 2 N2 (8.2.30) dt The solution to equation 8.2.30 is given by trying a solution of the form, N2 = a(1 e1 t ) + be 2 t c with the boundary conditions that at t = 0, N2 = 0 = b c. therefore, dierentiating equation 8.2.31, we obtain, dN2 = a1 e1 t b2 e2 t = 1 N1 2 N2 dt and therefore dN2 = I0 (1 e1 t ) 2 a(1 e1 t ) 2 be2 t + 2 c dt re-arranging and collecting the terms (8.2.33) (8.2.31) (8.2.29)

(8.2.32)

a1 e1 t b2 e2 t = I0 2 (a c) (I0 2 a)e1 t 2 be2 t

(8.2.34)

Equation 8.2.34 requires that at t = , I0 = 2 (a c) and comparing the terms for the e1 t term, a1 = (I0 2 a). Rearranging, this gives that 195

I0 (8.2.35) 2 1 also, collecting and comparing the non-exponential terms in equation 8.2.34 gives a= I0 = 2 ( re-arranging this gives c = I0 ( I0 c) 2 1 (8.2.36)

1 1 )=b 2 1 2

(8.2.37)

196

Figure 8.19: Gamma-ray spectra for dierent masses taken from the GSI on-line mass separator, following the binary collision of a 11.4 MeV/u 136 Xe beam with a Ta target. In each case, the tape cycle time was taken to be 8 seconds. [329].

197

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 2 N2 = 2 I0 2 (1 e1 t ) I0 ( 1)(1 e2 t ) 2 1 2 1 (8.2.38)

Note that the limits of equation 8.2.38 give when 2 t >> 0 and 2 >> 1 (ie. upper level is much longer lived than lower one, 2 N2 I0 (1 e1 t ) ie. the decay rate depends on the lifetime of the upper state. Similarly, when 1 t >> 0 and 1 >> 2 2 N2 I0 (1 e2 t ) If we substitute r =
2 , 2 1

(8.2.39)

(8.2.40)

into equation 8.2.38, it reduces to (8.2.41)

count rate = const. r(1 e1 t ) (r 1)(1 e2 t )

Figure 8.20 shows the grow-in curves for the decays from the ground state and excited states of 177 Lu into states in 177 Hf, following a binary reaction at GSI [329]. Figure 8.21 shows the eect of the dierent tape cycle times for A=177 nuclei following the binary reaction of 136 Xe 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

19

Ne +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 19 Ne (T 1 =17 s) by accelerating
2

protons onto a 19 F target and using the charge exchange reaction 19 F(p,n)19 Ne. The 19 Ne beam was then injected into a second cyclotron where it was accelerated to Coulomb barrier energies and used to bombard a target of 40 Ca. 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

evaporated -particle and protons from fusion-evaporation events. Typical beam currents were around 0.1 pnA.

19

Ne

Figure 8.22 highlights the problems associated with doing in beam gammaray 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 + decay of 19 Ne). However, as gure 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 gure 8.24 shows, the eect of charged particle gating is less dramatic than performing a subtraction using the time spectra gating using the out-ofbeam spectra, however, a signicant 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 19 Ne induced reaction compared to a 19 F (stable beam) reaction. While, it is apparent that the cross-section of the most neutron decient nuclei increases by using a more neutron decient (radioactive) beam (19 Ne), 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 reactions will require both more exotic beams (more than one nucleon from stability) coupled to intensities of a least 1 pnA.

199

Figure 8.20: Grow-in curves for the beta-decay of the high-spin isomer in 177 KLu (upper panel) and ground state of 177 Lu into 177 Hf. The upper panel populates a second isomeric state in 177 Hf and therefore requires a two component t, while the lower panel is a simple one component t with a half-life of 1.9 hours [329].

200

Figure 8.21: Eect of the tape cycle time in selecting dierent decay half-lifes. A=177 spectra with identical reaction and source conditions, but dierent tape cycle times [329].

201

Figure 8.22: Dierents types of channel selection to pick out transitions from fusion products for the 19 Ne+40 Ca reaction [345].

202

Figure 8.23: TDC time spectra from the (a) 19 Ne and (c) 19 F beams at LLN. Note the increase in counts for the beam on period corresponding to gamma-rays from beam-induced reactions.

203

Figure 8.24: Channel selection aorded by gating on evaporated charged particles in the 19 Ne+40 Ca experiment [345].

204

Figure 8.25: Comparison of yields from various residual channels using both stable (19 F) and radioactive 19 Ne beams on a 40 Ca target [345].

205

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. Kibdi, private communication e [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. Rsel et al. At. Data. Nucl. Data. Tab. 21 (1978) p291 o [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, Department of Physics, University of Jyvskyl, Research Report No. 2/1986, a a unpublished 206

[14] K. Heyde and R.A. Meyer Phys. Rev. C37, 2170 (1988) [15] H. Mach et al., Phys. Rev. C42, 793 (1990) [16] T. Kibdi et al. Nucl. Phys. A567 (1994) p183 e [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. Steen and R.M. Wheeler, Nuclear Data Tables 11 (1973) p351 [27] A. Krmer-Flecken et al. Nucl. Inst. Meth. A275 (1989) p333 a [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

[35] P. Ring and R. Schuck The Nuclear Many Body Problem, (1980) SpingerVerlag, 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. Dnau, Nucl. Phys. A471 (1987) p469 o [41] D. Ward et al. Nucl. Phys. A529 (1991) p315 [42] P.M. Walker et al. Phys. Lett. B309 (1993) p17 [43] J.M. ODonnell 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

[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

[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

[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. OLeary 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. Lidn et al. Nucl. Inst. Meth. Phys. Res. A273 (1988) p240 e [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

[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; erratum 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

[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. Weizog 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

[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.Giord 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. Krll and D. Bazzacco Nucl. Instr. Meth. A463 (2001) p227 o [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

[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, Selected Topics in Nuclear Structure, (1990), vol.2, p3 (World Scientic), 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 (NorthHolland, 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

[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. Zamr 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

[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. Bhm et al. Nucl. Inst. Meth. Phys. Res. A329 (1993) p248 o [233] J.M. Reid The Atomic Nucleus (1986) Manchester University Press [234] T. Lnnroth et al. Z. Phys. A317 (1984) p215 o [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. Weiszog et al. Nucl. Phys. A584 (1995) p133 [240] M. Weiszog 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. Husser et al. Nucl. Phys. A412 (1984) p141 a [244] O. Husser Phys. Lett. B144 (1984) p341 a [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

[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

[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 NeutronRich Nuclei, Sanibel Island, Florida 1997, eds. J.H. Hamilton and A.V. Ramayya, World Scientic (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, University of Manchester (1993) [281] A. Yamamoto, nal year project, University of Surrey, (2000) unpublished. [282] R.D. Page private communication [283] H. Geissel, G.M nzenberg and K. Riisager, Ann. Rev. Nucl. Part. Sci. 45 u (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

[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. Pf tzner, Acta. Phys. Pol. 28 (1997) p289 u [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. Pf tzner et al., Phys. Lett. B444 (1998) p32 u [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. Podolyk et al. Phys. Lett. B491 (2000) p225 a 220

[313] Z. Podolyk et al. Proceedings of the International Conference on Neua tron Rich Nuclei and Fission, St. Andrews, Scotland, (1999) p156. Eds. J.H. Hamilton, W.R. Phillips and H.K. Carter, World Scientic, London [314] M. Pf tzner et al.Proceedings of the International Conference on Experu 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 Conferences, Berkeley, USA (2001), in press.

221

[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

You might also like