Avrigeanu 2008

Download as pdf or txt
Download as pdf or txt
You are on page 1of 25

Nuclear Physics A 806 (2008) 15–39

www.elsevier.com/locate/nuclphysa

Fast-neutron induced pre-equilibrium reactions


on 55Mn and 63,65Cu at energies up to 40 MeV
M. Avrigeanu a , S.V. Chuvaev b , A.A. Filatenkov b , R.A. Forrest c ,
M. Herman d , A.J. Koning e , A.J.M. Plompen f , F.L. Roman a ,
V. Avrigeanu a,∗
a “Horia Hulubei” National Institute for Physics and Nuclear Engineering, PO Box MG-6, 077125 Bucharest, Romania
b V.G. Khlopin Radium Institute, 2nd Murinski Ave. 28, St. Petersburg 194021, Russia
c EURATOM-UKAEA Fusion Association, Culham Science Centre, Abingdon OX14 3DB, United Kingdom
d Brookhaven National Laboratory, PO Box 5000, Upton, NY 11973-5000, USA
e Nuclear Reseb and Consultancy Group NRG, PO Box 25, 1755 ZG Petten, The Netherlands
f European Commission, Joint Research Center, Institute for Reference Materials and Measurements,
B-2440 Geel, Belgium
Received 25 July 2007; received in revised form 10 March 2008; accepted 17 March 2008
Available online 21 March 2008

Abstract
The experimental cross sections for all neutron-induced reactions on the stable isotopes of Mn and Cu
at incident energies up to 40 MeV are compared with results of pre-equilibrium (PE) and statistical model
calculations. Taking advantage of the large sensitivity of statistical model analysis to residual nuclei pa-
rameters, for energies up to 20 MeV, our work was then focused on assessing PE model assumptions by
data analysis in the energy range 20–40 MeV. The PE model reliance on the nuclear potential finite-depth
correction for description of particle–hole state densities (PSD), nuclear-shell effects within the PSD for-
mula, and single-particle level density is thus discussed with regard to the whole data basis. The related
variance of the model results has been found as large as the differences between the global and local model
calculations, while the data basis available between 20 and 40 MeV for the Mn and Cu isotopes is however
too scarce in this respect.
© 2008 Elsevier B.V. All rights reserved.

PACS: 24.10.-I; 24.60.Dr; 25.40.-h; 28.20.-v

* Corresponding author. Tel.: +40 21 4046125.


E-mail address: [email protected] (V. Avrigeanu).

0375-9474/$ – see front matter © 2008 Elsevier B.V. All rights reserved.
doi:10.1016/j.nuclphysa.2008.03.010
16 M. Avrigeanu et al. / Nuclear Physics A 806 (2008) 15–39

Keywords: N UCLEAR REACTIONS 55 Mn, 63,65 Cu(n, X), (n, γ ), E = 0.001–50 MeV; 52,53,54 Cr, 55 Mn, 54,56,57,58 Fe,
59 Co, 58,60,61,62,64 Ni, 63,65 Cu, 64,66,68 Zn(p, X), (p, n), (p, γ ), E = 1–20 MeV; analyzed total, activation and
capture cross sections and particle emission spectra using an optical model with emphasis on pre-equilibrium
emission. Comparison with all available data

1. Introduction

As part of a general investigation [1–4] of the reaction mechanisms of fast neutrons at low
and medium energies, we have analyzed the activation cross sections of the odd-mass isotopes
55 Mn and 63,65 Cu in the excitation-energy range up to 40 MeV.

The main purpose of this paper is to discuss some of the question marks associated with
the model calculations which combine pre-equilibrium emission (PE) with equilibrium decay
of the remaining compound nucleus. Although a former aim was to comply with the needs of
a sound, complete and reliable neutron-induced cross section data library to address safety and
environmental issues of the fusion programme [5,6], these results also enable a stringent test
of models for the above-mentioned nuclear processes. The odd-mass target nuclei within the
present work may be particularly useful in connection with the proven influence of the f7/2
neutron and proton shell closures on the particle PE spectra (e.g., [7]). Actually Koning and Dui-
jvestijn [8] pointed out that omission of the shell effects is probably the most important cause
of the remaining discrepancies in their large-scale comparison of the nucleon PE model with
angle-integrated nucleon spectra. Moreover, a systematic analysis by Mills et al. [9] in the same
mass range highlighted that some of the discrepancies observed in the yields of nuclides with
closed or nearly-closed nucleon shells may not affect the inherent validity of the relevant model
but follow the use of incorrect, e.g., average model parameters for certain nuclei involved in the
decay process. Thus, in order to gain insight into this problem, we have analyzed the activation
cross sections of 55 Mn and 63,65 Cu isotopes using the parameter databases obtained previously
by global optimization within the computer codes TALYS [10] and EMPIRE-II [11], as well as
a local parameter set within the STAPRE-H code [12]. No fine tuning was done to optimize the
description of the nucleon emission for all the cases, but for STAPRE-H a consistent set of local
parameters has previously been established or validated on the basis of independent experimental
information of, e.g., neutron total cross sections, proton reaction cross sections, low-lying level
and resonance data, and γ -ray strength functions based on neutron-capture data. The comparison
of various calculations, including their sensitivity to model approaches and parameters, has con-
cerned all the activation channels for which there are measured data. It has thus avoided the use
of model parameters which have been improperly adjusted to take into account properties pecu-
liar to specific nuclei in the decay cascade, considered to be the case for discrepancies observed
around the closures of both the f7/2 proton and neutron shells [9].
Last but not least, there is indeed great interest in modern nuclear physics at low energies
for microscopic dynamical models such as, e.g., DYWAN [13]. However, for the time being this
kind of models are used to gain a better understanding of the microscopic mechanisms involved
in nucleon-induced reactions in order to obtain valuable information about nuclear interaction
properties [13]. Extended/unified semi-classical models may have a rather similar role as it was
proved by Ref. [14] according to which no mixing of particle–hole configurations provides a
more consistent description of PE reactions. Thus, it will be more difficult to use such models
in systematic studies of nucleon-induced reactions as within the recent European program HIN-
DAS (e.g., [15] and references therein), or in the latest SPIRAL2-NFS [16] and n_TOF-Ph2 [17]
M. Avrigeanu et al. / Nuclear Physics A 806 (2008) 15–39 17

projects. The thorough model calculations reported hereafter, whose importance is definitely be-
hind the microscopic and advanced semi-classical models, will have still to be largely involved
within these programs and match better the experimental and model-analysis activities. More-
over, despite the apparent gap between the global and local model analyses, we emphasize a
closer link between them, aiming to bring together the convenience and celerity of the former
and the latter’s accuracy within eventually extended nuclear structure frontiers. As soon as pow-
erful computer codes (e.g., [10,11]) become available, thoroughgoing model results may validate
or lead to better theoretical assumptions provided that various parameters are correctly settled.
The cross sections for nuclear reactions induced by fast neutrons below 20 MeV are gener-
ally considered to be reasonably well known in spite of many fast neutron reactions for which
the available data are either conflicting or incomplete even around 14 MeV [18]. Consequently,
a more recent set of accurately measured cross sections around 14 MeV [19] as well as just below
[20] and above this energy [21,22] proves valuable. The model calculations at these energies are
actually most sensitive to the statistical model parameters related to the residual nuclei, while PE
processes dominate at higher energies. Therefore, the corresponding PE model assumptions are
indeed better investigated by analyzing the data above 20–30 MeV, which helpfully exist for the
stable isotopes of Mn and Cu. However, this final aim is reachable provided that (i) the statistical
parameters related to nuclei in the decay cascade are validated by the account of the data below
∼20 MeV, and (ii) a large body of data is described with no free but consistent model parameters,
properly established by the analysis of other independent data. Thus, in order to match these for-
mer constraints, the main assumptions and parameters of the model calculations with the three
computer codes are discussed in Section 2, including the additional nuclear structure and reaction
data analyzed in this respect. Then, consideration of the main outcomes of this work correspond-
ing to the former analysis below ∼20 MeV is given in Section 3, while Section 4 concerns the
focus made thus possible on the discussion of model assumptions at higher energies.

2. Nuclear models and calculations

2.1. Global approach

The two sets of global calculations within the direct–reaction, PE and statistical Hauser–
Feshbach (HF) models, performed by means of the computer codes TALYS [10] and EMPIRE-
II [11], have mainly used systematics based on global phenomenological analysis. Thus their
results are firstly predictions of the reaction cross sections which should be considered from the
point of view of the global parameters involved in the corresponding calculations. Actually, such
blind calculations typically produce a correct shape for the excitation functions, while there is as
much underprediction as overprediction when the results are compared with data for all nuclides
of the periodic table of elements. Moreover, for a true evaluation, a normalization of the curves
can always be performed with nuclear model parameters that have an intrinsic uncertainty, such
as average radiative widths, level density parameters and pre-equilibrium matrix elements. How-
ever, for large-scale data evaluations based on nuclear model calculations, the performance of the
corresponding global estimations of these data are also quite important. The main assumptions
and parameters involved in this work for both sets of global calculations have been recently de-
scribed [4], while detailed descriptions were lately given [8,11] too. Therefore we give here only
some specific points which have arisen in the meantime.
A similar approach using the code TALYS was applied in this work to the Mn and Cu iso-
topes as that reported in Refs. [2,3] for Co, Ni, and Mo isotopes. However, a new version (0.72)
18 M. Avrigeanu et al. / Nuclear Physics A 806 (2008) 15–39

of the TALYS code was used although the previous description regarding the choices for the
optical model potential (OMP) [23], the direct reaction (using ECIS97 [24]), the level density
model [25,26] including the damping of shell effects at high excitation energies, and especially
the PE contributions with the two-component Exciton model using Kalbach systematics [27]
and particle–hole state densities including surface effects [28,29] which depend on the type of
projectile and the target mass [8] has been not altered for calculations that were performed in
this work. The discrete level schemes are adopted from the RIPL-2 database [25]. We note that
for this paper TALYS was only used for the global approach. A full description of the models
and methods used in TALYS can be found in Refs. [15,30–32], where also the applicability of
the code for the local approach, i.e. optimized parameters for each nucleus, is demonstrated. On
the other hand, the 2.19 version of the nuclear reaction code EMPIRE-II has been used for this
work due to its advantage of including the PE exciton model for cluster emission [11]. At the
same time, besides the adoption of default parameters, the Hybrid Monte Carlo simulation ap-
proach has been selected for the nucleon PE due to our interest in the neutron energies higher
than 30 MeV.

2.2. Local approach

The particular properties of various target nuclei and reaction channels have been considered
by using a consistent local parameter set, established on the basis of various independent data in
a small range of mass and charge numbers. A generalized Geometry-Dependent Hybrid (GDH)
model [33,34] for PE processes in STAPRE-H version of the original code [35] includes the
angular-momentum conservation [36] and the α-particle and deuteron emission based on a pre-
formation probability ϕ [37] with the values in the present work of 0.2 for α-particles and 0.4 for
deuterons [2]. The same optical potential and nuclear level density parameters have been used in
the framework of the OM [38], GDH and HF models, for calculation of the intra-nuclear transi-
tion rates and single-particle level (s.p.l.) densities at the Fermi level [34,39,40], respectively, in
the former case.
The nucleon optical potential of Koning and Delaroche [23], used by default in both TALYS
and EMPIRE codes, has obviously been the first option. However, a basic point revealed by
these authors is that their global potential does not reproduce the minimum around the neutron
energy of 1–2 MeV for the total neutron cross sections of the A ∼ 60 nuclei. Following also their
comment on the constant geometry parameters which may be responsible for this aspect, we have
applied the SPRT method [41] for determination of the OMP parameters over a wide neutron
energy range through analysis of the s- and p-wave neutron strength functions, the potential
scattering radius R  and the energy dependence of the total cross section σT (E). The recent
RIPL-2 recommendations [25] for the low-energy neutron scattering properties and the available
measured σT data (Fig. 1) have been used in this respect, and we found that it is necessary
to consider the energy dependence of the real potential geometry at lower energies shown in
Table 1. These potentials were used also for the calculation of the collective inelastic scattering
cross sections by means of the direct-interaction distorted-wave Born approximation (DWBA)
method and a local version of the computer code DWUCK4 [42]. The weak coupling model was
adopted in this respect for the odd nuclei 55 Mn and 63,65 Cu using the collective state parameters
of Kalbach [28]. Typical ratios of the direct inelastic scattering to the total reaction cross sections
in the energy range from few to 60 MeV decrease from ∼11 to 5%, for the 55 Mn nucleus, and
from ∼8 to 3% for the Cu isotopes.
M. Avrigeanu et al. / Nuclear Physics A 806 (2008) 15–39 19

Fig. 1. (Color online.) Comparison of experimental and calculated neutron total cross sections for 55 Mn and 63,65,nat Cu
target nuclei, by using the global (dotted curves) and local (dashed curves) OMP parameter sets of Koning and De-
laroche [23], and the changes of the latter given in Table 1 (solid curves). The experimental data are taken from the
EXFOR database [18].

Table 1
Comparison of experimental [25] and calculated neutron scattering parameters of 55 Mn and 63,65 Cu isotopes at the
neutron energies of 100, 80 and 50 keV, respectively, and (bottom) the changes of OMP parameters [23] which provide
the best SPRT results, where the energies are in MeV and geometry parameters in fm
Potential Target
55 Mn 63 Cu 65 Cu

S0 × 104 S1 × 104 R S0 × 104 S1 × 104 R S0 × 104 S1 × 104 R


Exp. 4.4(6) 0.3(1) 2.1(3) 0.44(7) 2.2(3) 0.47(8)
Ref. [23]-global 3.8 0.70 6.2 2.2 0.81 7.0 2.1 0.81 7.4
Ref. [23]-local 4.1 0.58 6.2 2.1 0.77 7.1 1.76 0.75 7.4
Ref. [23]-local + changes: 3.8 0.48 4.6 2.2 0.48 6.0 1.92 0.48 6.8
rV = 1.260 − 0.02E, E < 3 rV = 1.251−0.016E, E < 3
aV = 0.563+0.02E, E < 5 aV = 0.213+0.15E, E < 3 aV = 0.303 + 0.12E, E < 3

The OMP of Koning and Delaroche [23] was considered also for the calculation of proton
transmission coefficients on the residual nuclei, i.e. the isotopes of Cr and Ni, while a former
trial of this potential concerned the proton reaction cross sections σR [43]. Since these data are
20 M. Avrigeanu et al. / Nuclear Physics A 806 (2008) 15–39

Fig. 2. (Color online.) Comparison of the measured [43] and calculated proton reaction cross sections on all stable
isotopes of Mn, Fe, Co, Ni, Cu and Zn, by using either the local OMP predictions of Koning and Delaroche when they
are available in Table 8 of Ref. [23] or otherwise their proton global OMP (dotted, dash-dotted and dashed curves) and
the modified parameter set mentioned in the text for the target nuclei 56 Fe and 58 Ni (solid curves).

missing for the Cr nuclei, our local analysis involved the isotopes of Mn, Fe, Co, Ni, Cu and
Zn, for lower energies important in statistical emission from excited nuclei. The comparison of
these data and results of either the local OMP predictions when they are available in Table 8
of Ref. [23] or otherwise their proton global OMP is shown in Fig. 2. A very good agreement
exists apart from the isotopes of Fe and in particular Ni, with the data overpredicted by about or
higher than 10%. In order to obtain the agreement with the corresponding σR data (Fig. 2) we
have found necessary to replace the constant real potential diffusivity aV = 0.663 fm [23] by the
energy-dependent forms aV = 0.563 + 0.002E up to 50 MeV for the target nucleus 56 Fe, and
aV = 0.463 + 0.01E up to 20 MeV for 58 Ni, where the energy E is in MeV and the diffusivity in
fm. A final validation of both the original OMP and the additional energy-dependent aV has been
obtained by analysis of the available (p, γ ) and (p, n) reaction data up to Ep ∼ 12 MeV on Cr
(Fig. 3) and Ni isotopes (Fig. 4) while the other statistical model parameters are the same as in the
rest of the present work. It can be seen that these reaction data have been quite well reproduced,
with an increase of the related accuracy within 10% provided by the energy dependence adopted
for the real potential diffusivity at lower energies.
M. Avrigeanu et al. / Nuclear Physics A 806 (2008) 15–39 21

Fig. 3. (Color online.) Comparison of the measured [18] and calculated proton reaction cross sections (dash-dotted
curve), (p, γ ) and (p, n) reaction cross sections up to Ep ∼ 5–16 MeV on Cr isotopes by using the OMP parameter sets
mentioned for Fig. 2.

The optical potential which is used in this work for calculation of the α-particle transmission
coefficients was established previously [44] for emitted α-particles, and supported recently by
semi-microscopic analysis for A ∼ 90 nuclei [45]. On the other hand, by comparison of the
present calculations and measured data [18] for the target nuclei 63,65 Cu we found that the real
well diffuseness aR of the above-mentioned global OMP should be changed to 0.67 fm. This
reduction is rather similar to that found necessary for the target nuclei 59 Co, and 58,60,62 Ni [2],
so that it has been taken into account also for 55 Mn. Lastly, the calculation of the deuteron
transmission coefficients has been carried out by using the global OMP of Lohr and Haeberli [46]
and validated throughout analysis of the deuteron-emission spectra at 14.8 MeV [47].
The back-shifted Fermi gas (BSFG) formula has been used for the excitation energies be-
low the neutron-binding energy, with the parameters a and  (Table 2) obtained by a fit of
the recent experimental low-lying discrete levels [48] and s-wave nucleon resonance spacings
D0 [25]. Actually the same approach basis [50–53] and similar parameters have been used as
previously within this mass range [1,2], updated by means of the new structure data published
in the meantime. Concerning the particle–hole state density playing for PE description the same
role as the nuclear-level density for statistical model calculations, a composite formula [40] was
22 M. Avrigeanu et al. / Nuclear Physics A 806 (2008) 15–39

Fig. 4. (Color online.) As for Fig. 3 but for the Ni isotopes.

involved within the GDH model. Thus no s.p.l.-density free parameter except for the α-particle
state density [37] gα = A/10.36 MeV−1 was used for the PE account.
The modified energy-dependent Breit–Wigner (EDBW) model [54,55] was used for the elec-
tric dipole γ -ray strength functions fE1 (Eγ ) of main importance for calculation of the γ -ray
transmission coefficients, also as previously within this mass range [1,2]. The corresponding
M. Avrigeanu et al. / Nuclear Physics A 806 (2008) 15–39 23

Table 2
The low-lying levels number Nd up to excitation energy Ed [48] used in Hauser–Feshbach calculations, and the low-
exp
lying levels and s-wave nucleon-resonance spacings D0 (Ref. [25] except otherwise noted) in the nucleon energy range
E above the respective binding energy B, for the target-nucleus ground-state spin I0 , fitted in order to obtain the BSFG
level-density parameter a and ground-state shift  (corresponding to a spin cut-off factor calculated with a variable
moment of inertia between the half and 75% of the rigid-body value, for the excitation energies from g.s. to the nucleon
binding energy, and the reduced radius r0 = 1.25 fm)
Nucleus Nd Ed Fitted level and resonance data a 
exp
Nd Ed B + E
2 I0 D0
(MeV) (MeV) (MeV) (keV) (MeV−1 ) (MeV)
50 Ti 19 4.940 19 4.94 11.059 7/2 4.0(8) 5.55 1.20
51 Ti 22 4.187 18 3.77 6.565 0 125(70) 6.05 0.40
47 V 23 2.810 23 2.81 7.750 0 36.0(48)a 5.50 −1.15
48 V 23 1.781 23 1.78 5.95 −1.85
49 V 25 2.408 25 2.41 9.559 0 10.6(10)a 5.35 −1.80
50 V 32 2.162 46 2.65 5.95 −1.75
51 V 37 3.683 54 4.12 10.646 0 7.9(6)a 5.65 −0.68
11.071 6 2.3(6)
52 V 20 1.843 20 1.84 7.361 7/2 4.1(6) 6.15 −1.60
53 V 25 2.967 25 2.97 5.65 −1.03
54 V 19 1.752 17 1.54 5.95 −1.85
50 Cr 32 4.363 32 4.36 5.40 0.00
51 Cr 41 3.448 85 4.29 9.561 0 13.3(13)a 5.50 −1.20
52 Cr 17 4.100 17 4.10 5.55 0.20
53 Cr 31 3.617 27 3.44 8.432 0 43.40(437) 5.35 −0.90
54 Cr 33 4.458 33 4.46 9.817 3/2 7.8(8) 5.55 0.10
55 Cr 24 2.895 24 2.90 6.696 0 54.4(8)a 6.02 −0.82
62.0(8)
50 Mn 6 1.143 6 1.14 5.85 −1.40
51 Mn 20 2.984 28 3.29 5.55 −0.85
52 Mn 20 2.337 17 2.13 6.00 −1.20
53 Mn 36 3.555 42 3.73 5.35 −1.10
54 Mn 24 1.925 24 1.93 6.05 −1.81
55 Mn 32 2.953 45 3.07 10.497 0 7.1(7)a 5.70 −1.55
56 Mn 23 1.384 37 1.88 7.374 5/2 2.3(4) 6.10 −2.30
57 Mn 21 2.233 21 2.33 6.20 −1.25
58 Mn 11 0.882 11 0.88 6.65 −1.95
58 Fe 42 4.350 60 4.72 10.139 1/2 6.5(10) 6.15 0.15
59 Fe 28 2.856 28 2.86 6.755 0 25.4(49) 6.70 −0.70
60 Fe 21 3.714 21 3.71 6.15 0.15
61 Fe 3 0.391 13b 1.75 6.85 −1.00
55 Co 23 3.775 23 3.78 5.35 −0.40
56 Co 28 2.969 20 2.61 6.20 −0.78
57 Co 34 3.296 70 4.11 8.819 0 19.4(24)a 5.75 −0.98
9.591 13.3(11)a
58 Co 29 1.606 41 1.93 6.50 −2.23
59 Co 38 3.090 68 3.67 10.217 0 4.3(4)a 6.40 −0.85
60 Co 35 1.833 41 1.98 7.542 7/2 1.25(15) 6.95 −1.70
61 Co 24 2.499 28 2.64 6.85 −0.75
62 Co 12 0.920 16 1.27 7.30 −1.55
63 Co 11 2.191 11 2.19 7.30 −0.30
64 Co 8 0.953 17b 1.36 7.75 −1.30
(continued on next page)
24 M. Avrigeanu et al. / Nuclear Physics A 806 (2008) 15–39

Table 2 (continued)
Nucleus Nd Ed Fitted level and resonance data a 
exp
Nd Ed B + E
2 I0 D0
(MeV) (MeV) (MeV) (keV) (MeV−1 ) (MeV)
57 Ni 22 4.374 22 4.37 5.70 0.46
58 Ni 32 4.752 32 4.75 5.90 0.65
59 Ni 36 3.196 57 3.64 9.411 0 13.4(9) 5.90 −1.10
12.5(9)a
60 Ni 31 4.116 45 4.54 11.394 3/2 2.0(7)a 6.10 0.20
61 Ni 21 2.129 21 2.13 8.045 0 13.8(9) 6.40 −1.24
13.9(15)a
62 Ni 25 3.860 47 4.46 10.631 3/2 2.10(15) 6.40 0.27
63 Ni 19 2.353 19 2.35 7.117 0 16(3) 7.35 −0.52
7.238 15(2)a
64 Ni 29 4.285 49 4.76 6.85 0.79
65 Ni 20 2.520 20 2.52 6.398 0 19.6(30) 7.80 −0.20
59 Cu 15 2.715 15 2.72 6.25 −0.45
60 Cu 17 1.007 22 1.43 7.00 −1.75
61 Cu 35 3.092 35 3.09 6.55 −0.67
62 Cu 18 1.077 56 1.92 7.10 −2.00
63 Cu 36 2.978 40 3.10 9.026 0 5.9(7)a 7.08 −0.50
64 Cu 45 1.918 84 2.42 7.993 3/2 0.95(9) 7.25 −1.78
65 Cu 21 2.669 51 3.36 7.70 −0.15
66 Cu 22 1.439 22 1.44 7.166 3/2 1.30(11) 7.95 −1.35
a Ref. [49]. b Levels of similar isotope in the close neighbouring.

fE1 (Eγ ) values have been checked within the calculations of capture cross sections of Mn and
Cu isotopes in the neutron energy range from keV to 3–4 MeV, by using the OMP and nuclear
level density parameters described above and global estimations [50] of the γ -ray strength func-
tions for multipoles λ  3. Thus we found that the fE1 (Eγ ) strength functions corresponding to
exp
the experimental [25] average radiative widths Γγ 0 provide an accurate description of the cap-
ture data for the Cu isotopes (Fig. 5) while an increased value Γγ 0 ∼ 1300 meV has been required
in the same respect for the 55 Mn nucleus. Finally, the accuracy of the γ -ray strength functions
adopted in this work is shown also by the above-mentioned analysis of the (p, γ ) reaction cross
sections (Figs. 3–4).

3. Outlook of the pre-equilibrium reactions account

The comparison of the experimental data with the model calculations in Figs. 6–8 concerns in
general measurements which were carried out after the end of ’60s, differing overall by ∼20%
excepting the reaction 65 Cu(n, p)65 Ni where the divergence goes to ∼50%. The more recent data
[19–22] provide a better confined range of these reaction cross sections around 14 MeV, within
an accuracy of 3–4% which is most important for the validation of model calculations at a similar
level. In spite of the smaller weight of PE processes at these energies, of around 15% of the total
reaction cross section, this increased level of accuracy is already acting as a rigorous assessment
of the model parameters. The rather similar agreement with these data is a good point for all three
calculations, while larger differences come out just above this energy range. The good agreement
of the local calculation, especially with reference to the simultaneous description of the data for
M. Avrigeanu et al. / Nuclear Physics A 806 (2008) 15–39 25

Fig. 5. (Color online.) Comparison of the measured [18] neutron-capture cross sections of 55 Mn and 63,65,nat Cu target
nuclei, for incident energies up to 3–4 MeV, and calculated values by using the computer codes TALYS-0.72 (dashed
curves) and EMPIRE-II (dash-dotted curves) with default global parameters, and the local analysis with γ -ray strength
functions fE1 (Eγ ) within the EDBW model corresponding to either the experimental [25] average s-wave radiative
widths Γγ 0 (dotted curves), or Γγ 0 values corresponding to a fit of experimental neutron capture data (solid curves).

all reaction channels and target nuclei, is mostly due to the local set up of level density parameters
by fitting of the recent resonance data and low-lying level schemes.
The formerly-mentioned needs of complete and reliable neutron-induced cross section data
for Mn and Cu also enabled a stringent test of the various nuclear models as well as their cor-
responding account of particular effects. However, the key points in this respect are related to
the PE description which becomes increasingly significant at higher energies. Thus, it seems of
relevance to look for answers which may be provided by the data analysis firstly below and then
above the incident energy of 20 MeV. Since this is the upper energy limit for neutron data which
are generally considered to be reasonably well known, the present discussion may also reveal
the eventual need for more measurements at higher energies. It has been carried out within the
local approach, based on the use of a consistent parameter set already established on the basis
of ancillary independent data. On the other hand, the insight of calculated results correspond-
ing to distinct parameter values or model assumptions may contribute to the understanding of
variance shown by the three model calculations in Figs. 6–8. Actually, an ultimate goal of this
investigation is to increase the global predictions accuracy to the level of local analysis.
26 M. Avrigeanu et al. / Nuclear Physics A 806 (2008) 15–39

Fig. 6. (Color online.) Comparison of measured neutron-activation cross sections for the target nucleus 55 Mn [19,21,
56–73] and calculated values obtained using the computer codes TALYS-0.72 (dashed curves) and EMPIRE-II (dash-
dotted curves) with default global parameters, and STAPRE-H (solid curves) with the local parameter set given in this
work. For the production of only an isomeric or unstable ground state, the corresponding spin, parity and lifetime are
shown between square brackets.

3.1. Calculated cross-section sensitivity below 20 MeV to optical potential parameters

A first point, following the optical potential analysis described in Section 2.2, should concern
the effects of the neutron and proton modified OMPs on the calculated reaction cross sections.
The larger amount of data existing for the (n, 2n) and (n, p) reactions on 65 Cu have been involved
in this respect as it is shown in Fig. 9. Thus, firstly one may note that the modified neutron
potential (Table 1) is leading to a decrease of ∼5% for the (n, 2n) reaction calculated cross
sections, with respect to the results obtained by using the original OMP [23]. Similarly, the
M. Avrigeanu et al. / Nuclear Physics A 806 (2008) 15–39 27

Fig. 7. (Color online.) As for Fig. 6 but for the target nucleus 63 Cu and measured cross sections [19,20,22,47,60,61,66,
74–99].
28 M. Avrigeanu et al. / Nuclear Physics A 806 (2008) 15–39

Fig. 8. (Color online.) As for Fig. 6 but for the target nucleus 65 Cu and measured cross sections [19,20,47,61,64,72,76,
77,80,82,87,88,93,96,98,100–118].
M. Avrigeanu et al. / Nuclear Physics A 806 (2008) 15–39 29

Fig. 9. (Color online.) Comparison of measured (see Fig. 8) and calculated cross sections within the local approach for the
(n, 2n) and (n, p) reactions on the target nucleus 65 Cu, by using the OMP parameter sets of Koning and Delaroche [23]
(dotted curves), and corresponding changes for the proton OMP (dashed curve) and neutron OMP (solid curves).

modified proton potential of Koning and Delaroche leads to a decrease of ∼20% of the calculated
(n, p) reaction cross sections. On the other hand, joining together the two changes in the case
of the (n, p) reaction, results in a compensation of the latter one and a reduced final change of
∼10% for the calculated cross sections. Therefore, the additional analysis of the nucleon OMP
improved the accuracy of the calculated cross sections from ∼20%, for the smaller cross sections,
to around 5% for the major reaction channels. In fact, this better precision is closer to the above-
mentioned range of 3–4% accuracy of the more recently measured data around 14 MeV, together
making possible an effective trial of the PE model parameters which are responsible for ∼15%
of the total reaction cross section at these energies.
It was noted in the previous section that, by comparison of the present calculations and mea-
sured data [18] for the target nuclei 63,65 Cu, it was found that the real well diffuseness aR of
the global OMP [44] for emitted α-particles should be decreased to 0.67 fm. Since this re-
duction is rather similar to that found recently to be needed for the target nuclei 59 Co and
58,60,62 Ni [2], it has been taken into account also for the target nucleus 55 Mn. However there

are a couple of key points related to this matter. First, by using just the global OMP [44] for
emitted α-particles, one would obtain also within the local approach calculated cross sections
for the reaction 55 Mn(n, α)52 V closed to the results provided by the code EMPIRE-II and in
very good agreement with the measured cross sections also by Bostan and Qaim [56]. On the
other hand, there is no way to explain the rest of corresponding data above the incident energy
of 12 MeV, by pure statistical including PE emission.
Second, Yamamuro [119] pointed out, with respect to the clear difference of the α-particle
OMPs which are needed for calculation of the (α, n) and (n, α) reaction cross sections, that
it is found for closed shell nuclei but not for odd target nuclei such as 53 Cr, 57 Fe, 61 Ni, and
67 Zn. However, the present case of the 63,65 Cu nuclei comes in addition of those mentioned in

Ref. [2], at variance with Yamamuro’s statement. Alternatively one may consider the possible
enhancement related to the position of a giant quadrupole resonance (GQR) at the excitation
energies concerned in these nuclei. Although generally the decay of the GQR is observed with
nucleon emission, recent work shows [120] that an appreciable (non-statistical) decay through
α-particle emission can occur. An extra yield which could be understood as decay from giant
resonances populated via neutron capture has been found as well for the Mo isotopes [45]. Thus
30 M. Avrigeanu et al. / Nuclear Physics A 806 (2008) 15–39

it follows that further analysis is required, making use also of microscopic DF potentials based
on temperature-dependent nuclear density distributions for the description of (n, α) excitation
functions [45].

3.2. Charged-particle emission spectra sensitivity

Actually one may note the same level of 5–20% differences between the global predictions
and the measured cross sections, around the incident energy of 14 MeV, as the OMP has an effect
on the calculated cross sections at these energies. Larger divergence occurs at higher energies,
where it could be related to the continuously growing importance of PE assumptions and key
quantities. However, just the assessment within this energy range of the consistent parameter set
involved in the local approach allows for a further focus on differences between the measured
and calculated cross sections especially above 20 MeV, in order to establish the correctness of
the adopted PE formalism.
Moreover, in spite of the well-known reduced usefulness of 14 MeV neutron reaction data to
validate PE calculations for medium-mass nuclei [33], the suitable description of related charged-
particle emission spectra may have a twofold outcome. The lowest-energy region of spectra,
corresponding to a second emitted particle from a fully equilibrated compound nucleus, may
truly validate the OMP used for emitted particles as well as the level density parameters of the
excited nuclei. On the other hand, the emission of high-energy charged particles is entirely due
to the PE processes. Thus, advanced pairing and shell corrections of particle–hole state densities
could be eventually confirmed by the PE model account of this emission–spectrum energy region.
Comparison of measured angle-integrated proton and α-particle emission spectra from 9
[18,86], 14.1 [121,122] and 14.8 MeV [47] neutron-induced reactions on 55 Mn and 63,65 Cu nuclei
and calculated values within the local approach is shown in Fig. 10. The two goals mentioned
above can be considered as being satisfied, with a couple of additional comments. While the
measured particle spectra of Ref. [47] are given in the laboratory system, their conversion to the
center-of-mass system is equivalent to a shift of the spectrum to higher energies, of up to one
MeV for the most energetic α-particles (see, e.g., Refs. [123,124]). Thus a good agreement is
seen between the measured and calculated α-particle emission spectra for 63,65 Cu nuclei, apart
from considerably higher measured data for the high energy parts of the spectra corresponding
to excitations below ∼2 MeV in the residual nuclei 60,62 Co. The same effect is seen in the case
of the target nucleus 55 Mn, for which the experimental α-particle spectra [121,122] are given as
function of channel energy and no further conversion is necessary for comparison with the model
calculation. This underestimation was noted as well for other target nuclei in this mass region
[123,124], indicating that there may be considerable direct excitation of residual nuclei low ly-
ing levels beyond the validity of the PE models. Concerning the additional underestimation of
the α-particle spectrum on 55 Mn at lowest energies, one should note that the measurements are
complicated in this energy range by a rather large background [124].

4. Model-analysis support above 20 MeV

Above the neutron incident energy of 20 MeV the PE model becomes increasingly important
in determining the reaction cross section. The lack of free parameters within the corresponding
GDH model, as well as the consistent use of the same optical potential and nuclear level density
parameters as the HF model, make possible a focus on the correctness of the main related quantity
which is the particle–hole state density.
M. Avrigeanu et al. / Nuclear Physics A 806 (2008) 15–39 31

Fig. 10. (Color online.) Comparison of measured [47,86,121,122] angle-integrated proton and α-particle emission spectra
from 9, 14.1 and 14.8 MeV neutron-induced reactions on the 55 Mn and 63,65 Cu nuclei and calculated values within the
local approach for the PE emission (dashed curves), statistical first- (dash-dotted curves) and second-emitted particles
(dotted curves) from equilibrated compound nuclei, and their sum (solid curves).

4.1. Nuclear potential finite-depth correction

The original GDH formalism [33,34] considered a Fermi distribution for the nuclear matter
density, with the Fermi energy F = 40 MeV at saturation density. On this basis it takes into ac-
count the nuclear surface effects by means, firstly, of the sum of contributions due to different
entrance channel partial waves l for the first projectile-target interaction. The relevant parame-
ters in this case are averaged over the nuclear densities corresponding to the entrance-channel
trajectories from a point at which the nuclear density is ∼1/150 of its saturation value to the
radius Rl = λ (l + 12 ). Secondly, lower local-density Fermi energies [125] calculated for each of
these trajectories, F1 (Rl ), have been considered within the particle–hole state densities (PSD)
and limited the hole degrees of freedom. They correspond to a finite well depth correction which
has been included [29] in the PSD equidistant spacing model at the same time as the advanced
pairing [126] and shell corrections [127] added to the Pauli correction, and the non-equidistant
single-particle levels [128]. All of the above were included in a PSD composite formula [40]
added to the GDH model within the STAPRE-H code, and were part of previous studies carried
out in a similar way [1–4] but at incident energies up to 20 MeV. The extension of the present
analysis to 40 MeV, by means of the measured data put together in Figs. 11–12, is able to check
the importance of the finite-depth correction in the frame of the PSD composite formula [40].
Thus, the vanishing of this correction is obtained by replacing the local-density Fermi en-
ergies F1 (Rl ) with the Fermi energy central value F = 40 MeV. The results of this exercise
are shown in Fig. 11, the most apparent and direct view corresponding to the 65 Cu(n, p)65 Ni
reaction. The GDH l-dependent finite-depth corrections F1 (Rl ) allow the opening only with
the energy increase of the PE contribution due to each higher partial wave. This attribute, to-
32 M. Avrigeanu et al. / Nuclear Physics A 806 (2008) 15–39

Fig. 11. (Color online.) Comparison of measured (see Figs. 6–8) neutron-activation cross sections for the target nuclei
55 Mn and 63,65 Cu up to 40 MeV, and calculated values with the local parameter set given in this work (solid curves)
except for replacement of either the local-density Fermi energies F1 (Rl ) with the Fermi energy central value F (dash-
dotted curves), or the average energy-dependent s.p.l. densities with the constant value g(F ) (dotted curves), as well as
for removal of the shell correction S in the PSD composite formula (dashed curves).
M. Avrigeanu et al. / Nuclear Physics A 806 (2008) 15–39 33

Fig. 12. (Color online.) As for Fig. 11 but for removal of the ‘continuum effect’ (CE) of the s.p.l. density within the
particle–hole state density calculation (dotted curves), and taking into account for this effect the nucleon binding energy
B either alone (dash-dotted curves) or together with the Coulomb barrier BC (dashed curves), while the solid curves
correspond to consideration of also the centrifugal barrier BCF .
34 M. Avrigeanu et al. / Nuclear Physics A 806 (2008) 15–39

gether with the decreasing total reaction cross section σR with energy increasing, leads to a
rising fraction σP E /σR but a rather constant (n, p) reaction cross section above the incident
energy ∼20 MeV where the emission of a second neutron becomes possible. By raising the
local-density Fermi energies to the central value, the PE contributions of all partial waves be-
come possible from the beginning, in the limit set by the corresponding transmission coefficients.
Thus the fraction σP E /σR will increase faster while, e.g., the (n, p) reaction cross section will
decrease continuously after getting a higher maximum. The latter two attributes are both opposed
to the experimental data, even if their energy dependence above 20 MeV is only fairly accurate.
The same findings follow the analysis of the other data above 20 MeV shown in Fig. 11, as well
as the reaction 65 Cu(n, 2n)64 Cu added for completion. Changes of the calculated reaction cross
sections above the incident energy of 20 MeV, corresponding to this finite-depth correction, are
going from ∼50% to more than 100%. A similar result was noticed by Korovin et al. [129],
within a modified GDH model and using the former PSD formula of Ericson, while the present
results are based on the composite formula [40]. However, excepting the (n, 4n) and (n, pα)
reactions on 55 Mn and the (n, 3n) reaction on 63 Cu, the need for more accurate measured data at
least up to 40 MeV is obvious.

4.2. Single-particle levels density effects

The Fermi-gas model (FGM) energy dependence of the s.p.l. density has been used within the
PSD composite formula [40], in the present local approach as well as in the recent similar anal-
yses [1–4], following the study and conclusions of Herman et al. [128]. Actually average values
of the s.p.l. of excited particles and holes, gp (p, h) = g(F + ūp ) and gp (p, h) = g(F − ūh ) re-
spectively, have been obtained corresponding to the average excitation energies for particles and
holes ūp and ūh [40]. The s.p.l. density value at the Fermi level has been derived on the basis
of its relation to the nuclear level density parameter, g(F ) = π62 a, by using the parameter values
given in Table 2. By replacing the above-mentioned average energy-dependent s.p.l. gp(h) (p, h)
with the constant value g(F ) result in the changes shown in Fig. 11. They are quite small below
20 MeV, increasing for the incident energies up to 40 MeV from ∼5% to 20%.
Two features should be pointed out at this time. Firstly, the energy dependence of the s.p.l.
density is much less important than its value at the Fermi energy [4]. This is a consequence of
the fact that the PE cross section is determined by a ratio of the particle–hole level densities cor-
responding to exciton configurations which differ by one excited particle [33,34,39]. Secondly,
this change may become significant at energies higher than 40 MeV.

4.3. Nuclear-shell effects

The PSD composite formula [40] included the advanced pairing [126] and shell correc-
tions [127] added to the Pauli correction, by taking into account the nuclear-shell effects through
an additional back-shift S of the effective excitation energy [127]. It has been connected, for the
excitation energies lower than the binding energy, to the BSFG virtual ground-state shift param-
eter  = Up + S, where the former term is a constant pairing correction corresponding to the
PSD closed formula [127]. The washing out of shell effects above the neutron binding was taken
into account also for the back-shift S value by using the shell correction within the approach of
Junghans et al. [52], derived as mentioned in Section 2.2, with a similar smooth transition be-
tween the two energy range as for the nuclear level density. Obviously, the largest effect of this
PSD correction corresponds to the lowest excitation energies, the related S-values causing up the
M. Avrigeanu et al. / Nuclear Physics A 806 (2008) 15–39 35

high-energy limit of the emitted-particle spectra (e.g., Fig. 10). Since the shift  is around and
less than zero value for the odd-A and odd–odd nuclei, respectively, the back-shift S is negative
for most of the nuclei involved in the present work. It leads to enlarged effective excitation en-
ergies for the PSD calculation at lower excitations, finally increasing the PE cross sections. In
Fig. 11 is also shown the effect of removing the shell correction in the PSD composite formula,
the subsequent decrease of the PE component leading to, e.g., (n, 2n) and (n, 3n) reaction cross
sections decreased by 5–10% around the maximum of their excitation functions but (n, p) reac-
tion cross sections which decrease by ∼20%, around the incident energy of 14 MeV, up to more
than 50% at 40 MeV. The effect is obviously less important for (n, 4n) reactions, at least at the
energies involved in this work, where multiple PE processes are not important [8]. Nevertheless,
consideration of the nuclear-shell effects proves to be quite important within the present analy-
sis, in addition to the influence of the f7/2 neutron and proton shell closures on the particle PE
spectra [7] already noted.

4.4. The s.p.l.-density ‘continuum effect’

The ‘continuum effect’ (CE), i.e. the s.p.l. density decreasing with energy in the continuum
region [130–132], can be described basically using the corrected s.p.l. density formula [40]
    
ūp 1/2 ūp − B 1/2
gp (p, h) = g(F ) 1 + − θ (ūp − B) , (1)
F F
where B is the nucleon binding energy. However, for the present reaction cross section calcu-
lation, one should also take into account the Coulomb and centrifugal barriers (e.g., [38]). On
the other hand, one should note that the role of this effect will be major in the case of composite
nucleus particle–hole configurations excited at higher energies with respect to the residual nuclei.
The progressive addition of these barriers to the binding energy B, as well as the removal
of the continuum effect within the PSD calculation, are shown in Fig. 12. The most apparent
and direct view can be seen once more in the reaction 65 Cu(n, p)65 Ni. The decrease of the s.p.l.
densities due to the consideration of binding energy alone, with respect to no CE presumed, leads
to the increase of PE cross sections. The addition of the Coulomb barrier actually decreases the
CE weight, which remains visible only above the incident energy of ∼25 MeV. Finally, the
inclusion also of the centrifugal barrier reduces even more the CE size within the energy range
discussed in the present work. The CE complete treatment may again play an important role at
higher energies, the steps of its partial account in this analysis being able to shed some light on
the expected consequences at these energies.

5. Summary and conclusions

The experimental neutron-induced reaction cross sections are compared with the results of
model calculations for all activation channels for 55 Mn and 63,65 Cu target nuclei and incident
energies up to 40–50 MeV. The increased accuracy of recent cross sections [19–22] below 20–
21 MeV has made possible an effective trial of the PE model parameters at the same level, even
if this reaction mechanism is responsible at these energies for only ∼15% of the total reaction
cross section. It should be also noted that similar differences of 5–20% exist between the global
predictions and the measured cross sections, in the same energy range, as the OMP effects on
the calculated cross sections. On the other hand, this assessment of the consistent parameter set
36 M. Avrigeanu et al. / Nuclear Physics A 806 (2008) 15–39

involved in the local analysis below 20 MeV allows a further focus at higher energies on differ-
ences between the measured and calculated cross sections related to the PE model assumptions.
The production of both low- and high-spin isomers supports the assumptions adopted, e.g. within
local-parameter calculations, for the level density angular-momentum distribution as well as the
γ -ray strength functions. In the case of α-particle emission channels this work has additionally
validated the angular-momentum conservation within the PE model.
Larger divergence between the measured and calculated cross sections occurs mainly for
global predictions at higher energies, where the importance of PE assumptions and key quantities
is continuously increasing. Since an ultimate goal of this investigation is to increase the global
prediction accuracy to the level of local analysis, we have looked for the significant effects related
to distinct parameter or model assumptions. The most important is found to correspond within
the GDH model to the nuclear potential finite-depth correction taken into account for description
of particle–hole state densities. Its omission leads to a large increase of the PE weight as well as
to reaction cross section changes going from ∼50% to more than 100%. However, the need for
more accurate measured data at least up to the incident energy of 40 MeV is obvious. A similar
case is shown by consideration of the nuclear-shell effects within the PSD formula. On the other
hand, there are effects such as the s.p.l.-density energy dependence and inclusion of the ‘contin-
uum effect’ which may however become significant at energies higher than 40 MeV. Therefore,
this discussion is also pointing out the need of further measurements of neutron activation cross
sections at energies up to 40 [16] or even 100 MeV [17].

Acknowledgements

Work supported in part by European Community EFDA under Contract of Association


EURATOM-MEdC and MEdC Contract No. CEEX-05-D10-48.

References

[1] P. Reimer, V. Avrigeanu, A.J.M. Plompen, S.M. Qaim, Phys. Rev. C 65 (2001) 014604.
[2] V. Semkova, V. Avrigeanu, T. Glodariu, A.J. Koning, A.J.M. Plompen, D.L. Smith, S. Sudar, Nucl. Phys. A 730
(2004) 255.
[3] P. Reimer, V. Avrigeanu, S. Chuvaev, A.A. Filatenkov, T. Glodariu, A.J. Koning, A.J.M. Plompen, S.M. Qaim,
D.L. Smith, H. Weigmann, Phys. Rev. C 71 (2005) 044617.
[4] V. Avrigeanu, R. Eichin, R.A. Forrest, H. Freiesleben, M. Herman, A.J. Koning, K. Seidel, Nucl. Phys. A 765
(2006) 1.
[5] R.A. Forrest, J. Kopecky, Fus. Eng. Design 82 (2007) 73.
[6] R.A. Forrest, J. Kopecky, J.-Ch. Sublet, Report UKAEA FUS 535, UKAEA Culham Science Centre, March, 2007.
[7] M. Blann, S.M. Grimes, L.F. Hansen, T.T. Komoto, B.A. Pohl, W. Scobel, M. Trabandt, C. Wong, Phys. Rev. C 28
(1983) 1475.
[8] A.J. Koning, M.C. Duijvestijn, Nucl. Phys. A 744 (2004) 15.
[9] S.J. Mills, G.F. Steyn, F.M. Nortier, in: S.M. Qaim (Ed.), Proceedings of the International Conference on Nuclear
Data for Science and Technology, Jülich, May 1991, Springer-Verlag, Berlin, 1992, p. 708.
[10] A.J. Koning, S. Hilaire, M.C. Duijvestijn, in: R.C. Haight, M.B. Chadwick, T. Kawano, P. Talou (Eds.), Proceed-
ings of the International Conference on Nuclear Data for Science and Technology, 2004, Santa Fe, in: AIP Conf.
Proc., vol. 769, American Institute of Physics, New York, 2005, p. 1154;
www.talys.eu.
[11] M. Herman, P. Oblozinsky, R. Capote, M. Sin, A. Trkov, A. Ventura, V. Zerkin, in: R.C. Haight, M.B. Chadwick,
T. Kawano, P. Talou (Eds.), Proceedings of the International Conference on Nuclear Data for Science and Tech-
nology, 2004, Santa Fe, in: AIP Conf. Proc., vol. 769, American Institute of Physics, New York, 2005, p. 1184;
EMPIRE-II v.2.19, http://www-nds.iaea.org/empire/.
M. Avrigeanu et al. / Nuclear Physics A 806 (2008) 15–39 37

[12] M. Avrigeanu, V. Avrigeanu, STAPRE-H95 Computer Code, IPNE Report NP-86-1995, Bucharest, 1995, and
references therein;
News NEA Data Bank 17 (1995) 22.
[13] F. Sebille, V. de la Mota, I.C. Sagrado Garcia, J.F. Lecolley, V. Blideanu, Nucl. Phys. A 791 (2007) 313.
[14] C.A. Soares Pompeia, B.V. Carlson, Nucl. Phys. A 787 (2007) 211c.
[15] M.C. Duijvestijn, A.J. Koning, Ann. Nucl. Eng. 33 (2006) 1196.
[16] X. Ledoux, S. Simakov, in: The SPIRAL2 Project. Neutrons For Science GANIL, Caen, 2006, http://www.ganil.fr/
research/developments/spiral2/.
[17] A. Mengoni, F. Käppeler, E. Gonzales Romero, et al., n_TOF Collaboration, n_TOF-Ph2 Activities at the Neutron
Time of Flight Facility at CERN, CERN, Geneva, 2005, http://www.cern.ch/ntof/Documents/LoI_2005/.
[18] EXFOR Nuclear Reaction Data, http://www-nds.iaea.or.at/exfor, http://www.nndc.bnl.gov/nndc/exfor.
[19] A.A. Filatenkov, S.V. Chuvaev, V.G. Khlopin, Radium Institute Report RI-252, CNIIatominform, Moscow, 1999
(in Russian);
A.A. Filatenkov, et al., Data file EXFOR-31496, dated 2002-03-21, as Ref. [18].
[20] W. Mannhart, D. Schmidt, Data file EXFOR-22902, dated 2005-12-27, as Ref. [18].
[21] A. Fessler, A.J.M. Plompen, D.L. Smith, J.W. Meadows, Y. Ikeda, Nucl. Sci. Eng. 134 (2000) 171.
[22] A.J.M. Plompen, et al., Data file EXFOR-22684, dated 2004-01-12, as Ref. [18].
[23] A.J. Koning, J.P. Delaroche, Nucl. Phys. A 713 (2003) 231.
[24] J. Raynal, Tech. Rep. CEA-N-2772, CEA, Saclay, 1994.
[25] IAEA-CRP Reference Input Parameter Library (RIPL-2), http://www-nds.iaea.or.at.
[26] A.V. Ignatyuk, G.N. Smirenkin, A.S. Tishin, Yad. Fiz. 21 (1975) 485;
A.V. Ignatyuk, G.N. Smirenkin, A.S. Tishin, Sov. J. Nucl. Phys. 21 (1976) 255.
[27] C. Kalbach, Phys. Rev. C 33 (1986) 818.
[28] C. Kalbach, Phys. Rev. C 62 (2000) 44608.
[29] C. Kalbach, Phys. Rev. C 32 (1985) 1157.
[30] A.J. Koning, M.C. Duijvestijn, Nucl. Instrum. Methods B 248 (2006) 197.
[31] A.J. Koning, M.C. Duijvestijn, S.C. van der Marck, R. Klein Meulekamp, A. Hogenbirk, Nucl. Sci. Eng. 156
(2007) 397.
[32] A.J. Koning, M.C. Duijvestijn, J. Nucl. Sci. Techn. 44 (2007) 823.
[33] M. Blann, H.K. Vonach, Phys. Rev. C 28 (1983) 1475.
[34] M. Blann, Nucl. Phys. A 213 (1973) 570.
[35] M. Uhl, B. Strohmaier, Report IRK-76/01, IRK, Vienna, 1981.
[36] M. Avrigeanu, M. Ivascu, V. Avrigeanu, Z. Phys. A 329 (1988) 177;
M. Avrigeanu, M. Ivascu, V. Avrigeanu, Z. Phys. A 335 (1990) 299.
[37] E. Gadioli, E. Gadioli-Erba, Z. Phys. A 299 (1981) 1.
[38] O. Bersillon, Centre d’Etudes de Bruyeres-le-Chatel, Note CEA-N-2227, 1981.
[39] M. Avrigeanu, V. Avrigeanu, J. Phys. G: Nucl. Part. Phys. 20 (1994) 613.
[40] M. Avrigeanu, V. Avrigeanu, Comput. Phys. Commun. 112 (1998) 191;
A. Harangozo, I. Stetcu, M. Avrigeanu, V. Avrigeanu, Phys. Rev. C 58 (1998) 295.
[41] J.P. Delaroche, Ch. Lagrange, J. Salvy, IAEA-190, vol. 1, IAEA, Vienna, 1976, p. 251.
[42] P.D. Kunz, DWUCK4 user manual, OECD/NEA Data Bank, Issy-les-Moulineaux, France, 1984;
www.nea.fr/abs/html/nesc9872.html.
[43] R.F. Carlson, At. Data Nucl. Data Tables 63 (1996) 93.
[44] V. Avrigeanu, P.E. Hodgson, M. Avrigeanu, Phys. Rev. C 49 (1994) 2137.
[45] M. Avrigeanu, W. von Oertzen, V. Avrigeanu, Nucl. Phys. A 764 (2006) 246.
[46] J.M. Lohr, W. Haeberli, Nucl. Phys. A 232 (1974) 381.
[47] S.M. Grimes, R.C. Haight, K.R. Alvar, H.H. Barschall, R.R. Borchers, Phys. Rev. C 19 (1979) 2127;
Data file EXFOR-10827, dated 2003-06-02, as Ref. [18].
[48] Evaluated Nuclear Structure Data File (ENSDF), http://www.nndc.bnl.gov/ensdf/index.jsp.
[49] H. Vonach, M. Uhl, B. Strohmaier, B.W. Smith, E.G. Bilpuch, G.E. Mitchell, Phys. Rev. C 38 (1988) 2541.
[50] C.H. Johnson, Phys. Rev. C 16 (1977) 2238.
[51] V. Avrigeanu, T. Glodariu, A.J.M. Plompen, H. Weigmann, J. Nucl. Sci. Tech. Suppl. 2 (2002) 746.
[52] A.R. Junghans, M. de Jong, H.-G. Clerc, A.V. Ignatyuk, G.A. Kudyaev, K.-H. Schmidt, Nucl. Phys. A 629 (1998)
635.
[53] A.J. Koning, M.B. Chadwick, Phys. Rev. C 56 (1997) 970.
[54] D.G. Gardner, F.S. Dietrich, Report UCRL-82998, LLNL-Livermore, 1979.
38 M. Avrigeanu et al. / Nuclear Physics A 806 (2008) 15–39

[55] M. Avrigeanu, V. Avrigeanu, G. Cata, M. Ivascu, Rev. Roum. Phys. 32 (1987) 837.
[56] M. Bostan, S.M. Qaim, Phys. Rev. C 49 (1994) 266.
[57] R. Prasad, D.C. Sarkar, Data file EXFOR-30336, dated 1993-02-08, as Ref. [18].
[58] B. Minetti, A. Pasquarelli, Data file EXFOR-21345, dated 1980-02-22, as Ref. [18].
[59] M. Matuzda, R. Tanaka, Y. Yamamoto, N. Mori, I. Murata, A. Takahashi, K. Ochiai, T. Nishitani, Data file EXFOR-
22839, dated 2004-07-05, as Ref. [18].
[60] T.S. Soewarsono, Y. Uwamino, T. Nakamura, Data file EXFOR-22335, dated 1997-02-14, as Ref. [18].
[61] Y. Uwamino, H. Sugita, Y. Kondo, T. Nakamura, Data file EXFOR-22703, dated 2001-10-17, as Ref. [18].
[62] I. Kimura, K. Kobayashi, Data file EXFOR-22214, dated 1993-03-29, as Ref. [18].
[63] H.-L. Lu, J. Zhou, P.-G. Fan, J.-Z. Huang, Data file EXFOR-30615, dated 1990-10-05, as Ref. [18].
[64] Y. Ikeda, C. Konno, K. Oishi, T. Nakamura, H. Miyade, K. Kawade, H. Yamamoto, T. Katoh, Data file EXFOR-
22089, dated 2004-06-10, as Ref. [18].
[65] K. Kobayashi, I. Kimura, Data file EXFOR-22093, dated 1989-05-30, as Ref. [18].
[66] L.R. Greenwood, Data file EXFOR-12977, dated 1986-09-25.
[67] G.F. Auchampaugh, D.M. Drake, L.R. Veeser, Data file EXFOR-12936, dated 1985-10-10, as Ref. [18].
[68] A. Grallert, J. Csikai, Cs.M. Buczko, Data file EXFOR-31496, dated 2002-03-21, as Ref. [18].
[69] B.M. Bahal, R. Pepelnik, Data file EXFOR-21936, dated 1986-07-22, as Ref. [18].
[70] R. Vaenskae, R. Pieppo, Data file EXFOR-21893, dated 1984-08-07, as Ref. [18].
[71] E. Zupranska, K. Rusek, J. Turkiewicz, P. Zupranski, Data file EXFOR-30581, dated 1984-09-11, as Ref. [18].
[72] P.N. Ngoc, S. Gueth, F. Deak, A. Kiss, Data file EXFOR-30562, dated 1988-12-05, as Ref. [18].
[73] U. Garuska, J. Dresler, H. Malecki, Data file EXFOR-30479, dated 1990-02-08, as Ref. [18].
[74] H. Sakane, Y. Kasugai, M. Shibata, K. Kawade, A. Takahashi, T. Fukahori, Data file EXFOR-22662, dated 2005-
12-16, as Ref. [18].
[75] G.R. Pansare, V.N. Bhoraskar, Data file EXFOR-33001, dated 2006-05-31, as Ref. [18].
[76] F. Ghanbari, J.C. Robertson, Data file EXFOR-12950, dated 1986-08-14, as Ref. [18].
[77] W. Ryves, Data file EXFOR-20772, dated 1986-09-26, as Ref. [18].
[78] R.A. Jarjis, Data file EXFOR-21673, dated 1980-11-04, as Ref. [18].
[79] M. Majumder, B. Mitra, Data file EXFOR-30296, dated 2001-04-17, as Ref. [18].
[80] R.A. Sigg, Data file EXFOR-10776, dated 1987-01-22, as Ref. [18].
[81] M. Valkonen, Data file EXFOR-10776, dated 1987-01-22, as Ref. [18].
[82] S.M. Qaim, Data file EXFOR-20536, dated 1980-08-08, as Ref. [18].
[83] S.M. Qaim, S. Spellerberg, F. Cserpak, J. Csikai, Data file EXFOR-22418, dated 2000-07-24, as Ref. [18].
[84] D.L. Bowers, L.R. Greenwood, Data file EXFOR-13132, dated 1997-05-16, as Ref. [18].
[85] S.M. Qaim, N.I. Molla, Data file EXFOR-20721, dated 1986-04-16, as Ref. [18].
[86] C. Tsabaris, E. Wattecamps, G. Rollin, C. Papadopoulos, Nucl. Sci. Eng. 128 (1998) 47.
[87] M. Ahmad, C.E. Brient, P.M. Egun, S.M. Grimes, S. Saraf, H. Satyanarayana, Data file EXFOR-12960, dated
1987-05-05, as Ref. [18].
[88] J.W. Meadows, D.L. Smith, L.R. Greenwood, R.C. Haight, Y. Ikeda, C. Konno, Data file EXFOR-13586, dated
1996-08-02, as Ref. [18].
[89] Y. Ikeda, C. Konno, Y. Kasugai, A. Kumar, Data file EXFOR-22658, dated 2002-02-26, as Ref. [18].
[90] C. Konno, Y. Ikeda, K. Oishi, K. Kawade, H. Yamamoto, Data file EXFOR-22637, dated 2004-06-10, as Ref. [18].
[91] H.-L. Lu, W. Zhao, W. Yu, X. Yuan, Data file EXFOR-31407, dated 1993-02-08, as Ref. [18].
[92] W. Yongchang et al., Data file EXFOR-32576, dated 1992-07-06, as Ref. [18].
[93] G. Winkler, D.L. Smith, J.W. Meadows, Data file EXFOR-10921, dated 1985-07-08, as Ref. [18].
[94] U. Garuska, J. Dresler, H. Malecki, Data file EXFOR-30553, dated 1989-10-05, as Ref. [18].
[95] Y. Kasugai, H. Yamamoto, K. Kawade, T. Iida, Data file EXFOR-22434, dated 2007-03-30, as Ref. [18].
[96] F. Cserpak, S. Sudar, J. Csikai, S.M. Qaim, Data file EXFOR-22382, dated 1999-05-27, as Ref. [18].
[97] A. Paulsen, Data file EXFOR-20390, dated 1984-01-09, as Ref. [18].
[98] J. Kantele, D.G. Gardner, Data file EXFOR-11196, dated 1987-01-16, as Ref. [18].
[99] D.W. Kneff, B.M. Oliver, H. Farrar IV, L.R. Greenwood, Data file EXFOR-10933, dated 1986-05-20, as Ref. [18].
[100] V. Semkova, A.J.M. Plompen, D.L. Smith, as Ref. [11], p. 1019.
[101] N.I. Molla, R.U. Miah, S. Basunia, S.M. Hossain, M. Rahman, Data file EXFOR-31449, dated 2007-12-07, as
Ref. [18].
[102] J. Csikai, Data file EXFOR-30640, dated 1993-04-26, as Ref. [18].
[103] R. Spangler, E.L. Draper Jr., T.A. Parish, Data file EXFOR-12956, dated 1986-03-14, as Ref. [18].
[104] W. Mannhart, H. Vonach, Data file EXFOR-20611, dated 1984-01-27, as Ref. [18].
M. Avrigeanu et al. / Nuclear Physics A 806 (2008) 15–39 39

[105] J.C. Robertson, B. Audric, P. Kolkowski, Data file EXFOR-20799, dated 1981-01-20, as Ref. [18].
[106] J. Araminowicz, J. Dresler, Data file EXFOR-30264, dated 1993-04-26, as Ref. [18].
[107] T. Shimizu, H. Sakane, M. Shibata, K. Kawade, T. Nishitani, Data file EXFOR-22837, dated 2005-02-18, as
Ref. [18].
[108] J.W. Meadows, D.L. Smith, M.M. Bretscher, S.A. Cox, Data file EXFOR-12969, dated 1988-09-06, as Ref. [18].
[109] R. Pepelnik, B. Anders, B.M. Bahal, M. Farooq, Data file EXFOR-21999, dated 1986-04-11, as Ref. [18].
[110] M. Bormann, F. Drayer, U. Zielinski, Data file EXFOR-20899, dated 2007-03-20, as Ref. [18].
[111] E.T. Bramlitt, R.W. Fink, Data file EXFOR-11590, dated 1987-01-20, as Ref. [18].
[112] O.I. Artemev, I.V. Kazachevskiy, V.N. Levkovskiy, V.L. Poznyak, V.F. Reutov, Data file EXFOR-41321, dated
1998-11-21, as Ref. [18].
[113] T. Clator, Data file EXFOR-11536, dated 1986-05-02, as Ref. [18].
[114] I.L. Preiss, R.W. Fink, D.G. Gardner, Data file EXFOR-11789, dated 1983-09-27, as Ref. [18].
[115] S.M. Qaim, G. Stoecklin, Data file EXFOR-20536, dated 1976-04-21, as Ref. [18].
[116] D.C. Santry, J.P. Butler, Data file EXFOR-11776, dated 1983-09-27, as Ref. [18].
[117] M. Bormann, S. Cierjacks, R. Langkau, H. Neuert, Data file EXFOR-21339, dated 1980-02-22, as Ref. [18].
[118] B. Joensson, K. Nyberg, I. Bergqvist, Data file EXFOR-20513, dated 1983-11-21, as Ref. [18].
[119] N. Yamamuro, Nucl. Sci. Eng. 122 (1996) 374.
[120] M. Fallot, J.A. Scarpaci, N. Frascaria, Y. Blumenfeld, A. Chbihi, Ph. Chomaz, P. Desesquelles, J. Frankland,
E. Khan, J.L. Laville, E. Plagnol, E.C. Pollacco, P. Roussel-Chomaz, J.C. Roynette, A. Shrivastava, T. Zerguerras,
Phys. Lett. B 613 (2005) 128.
[121] R. Fischer, G. Traxler, H. Vonach, M. Uhl, Rad. Effects 96 (1986) 309.
[122] R. Fischer, G. Traxler, M. Uhl, H. Vonach, Phys. Rev. C 34 (1986) 460.
[123] C. Derndorfer, R. Fischer, P. Hille, H. Vonach, P. Maier-Komor, Z. Phys. A 301 (1981) 327.
[124] R. Fischer, G. Traxler, M. Uhl, H. Vonach, Phys. Rev. C 30 (1984) 72.
[125] E. Gadioli, E. Gadioli Erba, P.G. Sona, Nucl. Phys. A 217 (1973) 589.
[126] C. Kalbach, Nucl. Sci. Eng. 95 (1987) 70;
C. Kalbach, Z. Phys. A 332 (1989) 397.
[127] C.Y. Fu, Nucl. Sci. Eng. 86 (1984) 344.
[128] M. Herman, G. Reffo, R. Rosetti, G. Giardina, A. Italiano, Phys. Rev. C 40 (1989) 2870.
[129] Yu.A. Korovin, A.Yu. Konobeyev, P.E. Pereslavtsev, A.Yu. Stankovsky, C. Broeders, I. Broeders, U. Fischer,
U. von Möllendorff, Nucl. Instrum. Methods A 463 (2001) 544.
[130] S. Shlomo, Nucl. Phys. A 539 (1992) 17.
[131] Ye.A. Bogila, V.M. Kolomietz, A.I. Sanzhur, S. Shlomo, Phys. Rev. C 53 (1996) 855.
[132] S. Shlomo, V.M. Kolomietz, H. Dejbakhsh, Phys. Rev. C 55 (1997) 1972.

You might also like