Anda di halaman 1dari 34

Nuclear Physics A 793 (2007) 178211

Measurement and DWBA analysis of the 12C(6Li, d)16 O


-transfer reaction cross sections at 48.2 MeV. R-matrix
analysis of 12C(, )16 O direct capture reaction data
A. Belhout a , S. Ouichaoui a, , H. Beaumevieille a , A. Boughrara a ,
S. Fortier b , J. Kiener c , J.M. Maison b , S.K. Mehdi a , L. Rosier b ,
J.P. Thibaud c , A. Trabelsi a , J. Vernotte b
a USTHB-Facult de Physique, BP 32, El-Alia, 16111 Bab Ezzouar, Algiers, Algeria
b IPNO, IN2P3/CNRS-Universit de ParisXI, 91406 Orsay cedex, France
c CSNSM, IN2P3/CNRS-Universit de ParisXI, 91405 Campus Orsay, France

Received 22 May 2006; received in revised form 19 June 2007; accepted 21 June 2007
Available online 28 June 2007

Abstract
The angular distributions for low lying states of 16 O produced in the 12 C(6 Li, d)16 O -transfer re-
action at E6 Li (lab) = 48.2 MeV have been measured using a high energy resolution position sensitive
detection system. The measured cross section data have been analyzed by the FRDWBA theory with
a particular emphasis put on the states of astrophysical interest mainly the 1 (7.12 MeV) state. Ex-
tracted values of nuclear level parameters (excitation energy, Ex , line width, c.m. , -spectroscopic
factor, S , and reduced -width, 2 or 2 ) are reported and discussed in comparison to previous
ones. The used binding potential corresponds to a long nuclear interaction range (up to a channel
radius a = 7.7 fm), suggesting that the channel radius a = 6.5 fm is likely suitable for the cal-
culations of reduced -widths near the nuclear surface. However, given that previous DWBA re-
sults are available for comparison only near the channel radius a = 5.5 fm, calculations have been
carried out here for both three preceding a values. Formal values, 0.0036  2 (7.12)  0.152 and
0.019  2 (6.92)  0.096 of the reduced -widths (a = 6.5 fm) of the 1 (7.12 MeV) and 2+
(6.92 MeV) states have thus been derived. Furthermore, the R-matrix analysis of various data sets of
importance to the 12 C(, )16 O direct capture reaction has been carried out, leading to the range of
values for the 1 (7.12 MeV) state, 0.0096  2 (7.12)  0.0166, which is consistent with the pre-
ceding DWBA interval, and to a result for the E1 component of the reaction astrophysical factor,

* Corresponding author.
E-mail address: souichaoui@gmail.com (S. Ouichaoui).

0375-9474/$ see front matter 2007 Elsevier B.V. All rights reserved.
doi:10.1016/j.nuclphysa.2007.06.008
A. Belhout et al. / Nuclear Physics A 793 (2007) 178211 179

SE1 (0.3) = 80.6+17


16 keV b, very consistent with most constrained values previously reported in the lit-
erature.
2007 Elsevier B.V. All rights reserved.

PACS: 25.70.Hi; 24.10.Eq; 27.20.+n; 26.20.+f

Keywords: N UCLEAR REACTION 12 C(6 Li, d)16 O, E = 48.2 MeV. Measured differential cross sections and angular
distributions. Deduced spectroscopic factors. 16 O, deduced level parameters using DWBA analysis. 12 C(, )16 O,
E (c.m.); 5 MeV. Analysed existing data using R-matrix formalism. Deduced S-factor

1. Introduction

Previously, the 12 C(6 Li, d)16 O -transfer reaction angular distributions have been investigated
experimentally by many groups [16] (and references therein) at different incident energies.
Notably, Becchetti et al. [2,3] have measured deuteron angular distributions at E6 Li (lab) =
42.1 MeV [2] and 90.2 MeV [3] over a wide range of excitation energies in 16 O. More re-
cently, Brune et al. [4] have measured total cross sections for the J = 2+ (Ex = 6.92 MeV)
and 1 (7.12 MeV) states at sub-Coulomb energies while Drummer et al. [5] have reported an-

gular distribution and analyzing power data for the 0+ ground state produced in the 12 C(6 Li,
d)16 O reaction at E6Li

= 50 MeV. Recently, Keeley et al. [6] have reported analysing power and
cross section data for this state and the 3 (6.13 MeV), 2+ (6.92 MeV), 2 (8.87 MeV) and 4+
(10.35 MeV) excited states produced in the same reaction at E6Li
= 34 MeV and 50 MeV. These
authors have performed quantitative Coupled Channel Born Approximation/Coupled Reaction
Channel (CCBA/CRC) calculations assuming multi-step transitions to these states of 16 O, which
they compared to experimental data. Besides, similar studies as in Refs. [13] have been carried
out for the 12 C(7 Li, t)16 O -transfer reaction [79] (and references therein). Several interaction
mechanisms compete in both reactions: the compound nucleus formation presumed to be less
likely at sub-Coulomb barrier energies [4] and at energies substantially exceeding the Coulomb
barrier [3], the direct single-step -transfer mechanism, multi-step -transfers involving core
excitation (which require CCBA/CRC calculations [6,7]), and other possible direct reaction
mechanisms. In [13,79] and earlier works carried out at lower energies (E6,7 Li  30 MeV),
the experimental data were analyzed using the Zero-Range (ZR) or Finite-Range (FR) Distorted-
Waves Born Approximation (DWBA) assuming only single-step -transfers to dominate the
reaction cross sections at forward angles relative to the compound nucleus component evaluated
by the HauserFeschbach (HF) statistical model and normalized to the 2 (8.87 MeV) unnat-
ural parity state of 16 O. A large part of the published spectroscopic information for light and
medium mass nuclei has been derived following this classical approach. Concerning 16 O, the
previous studies aimed at two main objectives: (i) determining the level structure of this doubly
magic nucleus (-clustering, rotational bands, . . . ) by extracting the S spectroscopic factors,
and/or (ii) exploring to which extent some states participate to the stellar helium burning using
the level reduced -width, 2 (2 , see Eqs. (10)(13) in Section 4 for definitions), as the relevant
criterion. The structure of 16 O has been intensively investigated in previous works, both experi-
mentally and via theoretical models. These studies [2,3,8,10] (and references therein) have led to
establish that most of the states of this nucleus are involved into -cluster rotational bands built
upon intrinsic deformed states of different shell model particlehole configurations like, e.g., the
K = 0+ positive-parity band, based on the 0+ (6.05 MeV) first excited state and including the
2+ (6.92 MeV), 4+ (10.35 MeV) and 6+ (16.30 MeV) states of surface 4p4h configuration.
180 A. Belhout et al. / Nuclear Physics A 793 (2007) 178211

From a nuclear astrophysics point of view, the high cross section (6 Li, d) and (7 Li, t) reactions
provide an alternative way for extracting the reduced -widths for low lying states of 16 O cru-
cially involved in determining the total astrophysical S(E) factor of the challenging 12 C(, )16 O
-capture reaction. The rate of this reaction essentially yields the 12 C/16 O abundance ratio at the
end of helium burning and strongly influences the subsequent nucleosynthesis of heavier ele-
ments in massive stars (M  M ) and further evolution of these astrophysical sites [1118].
The 12 C(, )16 O reaction cross section is extremely small at stellar temperatures amounting
down to  1045 m2 at the Gamow peak energy, E (c.m.) = 0.3 MeV, corresponding to a stel-
lar temperature of T9  0.2 [18]. This is mainly due to the high Coulomb barrier in the + 12 C
reaction channel, i.e., Bc = 3.142 MeV for a channel radius value a = a0 (41/3 +121/3 ) = 5.5 fm.
Consequently, the reaction rate has remained excessively uncertain over about four decades com-
pared to that of the 3 12 C + triple-alpha process determined to within 1015% [19].
The 12 C + fusion (Q = 7.162 MeV) operates over an excitation energy region in 16 O around
Ex = 7.46 MeV. However, the reaction cross section is very sensitive to neighboring resonant
states of small -decay widths [20], i.e., the 1 (9.6 MeV) unbound state corresponding to the
broad resonance at ER  2.42 MeV which strongly influences the stellar energy region via its
low energy tail and the 1 (7.12 MeV) and 2+ (6.92 MeV) subthreshold states correspond-
ing to the resonances at ER = 45 keV and 245 keV, respectively. While the 2 values for
these states needed be determined with the highest possible precision, they actually remained
affected by long-standing too large uncertainties. Two g.s. -decay electric transition ampli-
tudes, a dipole and a quadrupole amplitudes (E1, E2) of comparable magnitudes [21], dominate
the 12 C(, )16 O capture reaction total astrophysical S(E) factor. The latter is generated by the
incoherent sum of these two components and of a much smaller cascade transition one due to
direct -captures of the same multipolarities, essentially E2 transitions to the 2+ (6.92 MeV)
and 0+ (6.05 MeV) states, i.e., S(E) = SE1 (E) + SE2 (E) + Scasc. (E). Notice that the capture to
the 0+ first excited state of 16 O has been estimated in a recent experiment [22] to be likely the
strongest contribution (15% at E = 0.3 MeV) to S(E) from cascade transitions. The first two
leading components (mainly determined by the two 1 states presumed to strongly interfere and
by the 2+ state, respectively) have been comprehensively investigated experimentally by many
groups at accessible (relatively high) energies, presently extending down to E = 0.89 MeV [23].
Experiments were performed via (i) direct measurements of the 12 C(, )16 O capture reaction
cross section ( -ray angular distribution data) in direct or inverse kinematics [19,21,2329] and
the elastic -scattering off 12 C yields (angular distribution and phase shift data) [3032], and
(ii) indirect measurements of the Coulomb break-up cross section for high energy 16 O ions pass-
ing nearby a heavy (208 Pb) target (e.g., [33]) and the -delayed -spectrum of 16 N [3436].
A survey of the achieved direct 12 C(, )16 O reaction measurements with indication of the de-
tection set-ups used is made in Ref. [29]. Both SE1 (0.3) and SE2 (0.3) values have been extracted
independently by extrapolating the corresponding measured data to the energy region of the reac-
tion Gamow window. The extrapolations were performed under the guidance of nuclear reaction
(K-matrix, R-matrix or hybrid R-matrix) theories [19,21,2328,3032,3439] or using nuclear
structure microscopic model calculations [40,41]. In the R-matrix analyses of various experimen-
tal data sets, the reduced -width amplitude of 16 O states was usually taken as a free fitting
parameter. This long-standing quest has led to strongly constrain the SE1 (0.3) value in a deci-
sive measurement [36] of the -delayed -spectrum of 16 N (sensitive to 2 (7.12)) with R-matrix
analysis of the data [36,37] while the SE2 (0.3) value (sensitive to 2 (6.92)) still remained much
less known. Over the following decade, further investigations have been devoted to also constrain
the latter component by measuring new direct 12 C(, )16 O cross section data [23,2729] and
A. Belhout et al. / Nuclear Physics A 793 (2007) 178211 181

elastic -scattering yield data [32] by using highly improved experimental techniques sometimes
involving 4 sr -ray detection geometry [29]. Yet, while theoretical astrophysics calculations re-
quire a relative precision of  10% on the reaction total S(0.3) value and stellar rate [13,17], the
currently attained experimental relative uncertainty is still heavily debated but surely not better
than 25% [23]. Meanwhile, due to the steadily large spreads on the laboratory measurements,
the 12 C(, )16 O reaction rate has been evaluated by astrophysics calculations [13], predicting a
value of S(0.3) = 170 50 keV b and a corresponding rate value exceeding by a factor of 1.7
the CF88 [12] value, and also from astrophysics observations [42].
As already stated, 2 values leading to higher 12 C(, )16 O reaction S(0.3) ones than rec-
ommended in [11,12], consistently with the predictions of improved stellar evolution models
[1317], are expected to be more easily derived from -transfer reactions on 12 C. However, short-
comings inherent to the DWBA calculation might induce large systematic errors on the nuclear
level parameter values extracted from the -transfer reaction experimental data using this theory.
To minimize such uncertainties arising from several sources (optical model potential parameters,
angular momentum mismatches affecting high excitation energy and high spin states, the over-
lap of different interaction mechanism, . . . ), relative values, R1 = 2 /2 (9.6), R2 = 2 /2 (6.92)
and R3 = 2 /2 (10.35) instead of absolute 2 values can be [2,3,8] more reliably determined
at as high possible incident energy. Relative (S , 2 ) values for the low lying states of 16 O
have been extracted previously from the measured 12 C(6 Li, d)16 O and 12 C(7 Li, t)16 O angular
distribution data via DWBA + HF theoretical analyses [1,2,7,8] (and earlier references therein).
According to Becchetti et al. [3] who investigated the latter reaction at E7 Li = 34 MeV [8],
70 MeV and 101 MeV [9], this reaction exhibits a more pronounced direct -transfer char-
acter at E  30 MeV than the former one whose study is complicated by compound nucleus
formation at such energies although, in both reactions, the higher the incident energy the more
prominent the direct -transfer interaction mechanism [2,3,8]. Therefore, the 2 results extracted
from the (6 Li, d) data at E6 Li = 90.2 MeV were presented [3] as more reliable than those de-
rived at E6 Li = 42.1 MeV [2]. This energy dependence of the interaction mechanisms seems to
manifest itself through the observed weaker populations of the 2 (8.87 MeV) unnatural parity
state and the 2+ (9.85 MeV) non--cluster state as the incident energy increases [2,3,8,9]. The
2 (7.12)  0.1 values for the 1 (7.12 MeV) state of 16 O extracted by these authors at several in-
cident energies were found significantly higher than those (2 (7.12)  0.1) derived via R-matrix
analyses of various experimental data related to the (, ) helium burning. However, although
being of the same order of magnitude, the relative reduced -width values reported in [2,3,8,9]
themselves still show substantial spreads, probably due to uncertainties arising from the over-
lap of interaction mechanisms. Besides, the 2 (7.12) values extracted from previous R-matrix
analyses of experimental data related to the 12 C(, )16 O reaction also show large spreads or are
often reported without errors quoted. They can hardly be considered presently to be known with
high precision despite the achieved significant reduction of the uncertainty on the 12 C(, )16 O
reaction SE1 (0.3) value. The close confrontation of the 2 values extracted by R-matrix analyses
of (, ) capture data with those derived via DWBA analyses of -transfer reaction data for the
astrophysically relevant states of 16 O should be very useful and seems to be highly desirable.
A previous discussion of this point can be found, e.g., in [37].
Therefore, among other objectives of interest to the 16 O nuclear structure, we aimed in this
work at extracting the most relevant range of 2 (7.12) values common both to the DWBA
analysis of (6 Li, d) -transfer reaction data and the R-matrix analysis of (, ) capture data,
consistently with the most constrained SE1 (0.3) values for the latter reaction reported in the lit-
erature [23,28,36]. Therefore, in view to constrain the S and 2 parameter values for low-lying
182 A. Belhout et al. / Nuclear Physics A 793 (2007) 178211

states of astrophysical interest in 16 O, mainly the 1 (7.12 MeV) state, we have undertaken a
new measurement of the 12 C(6 Li, d)16 O -transfer reaction angular distributions at a bombard-
ing energy of 48.2 MeV (lab). A high energy resolution experimental set up coupling a split
pole magnetic spectrometer to a position-sensitive detection system was used for this purpose.
Cross section data have been measured for the 0+ g.s. and the first ten excited states of 16 O,
i.e., the 0+ (6.05 MeV), 3 (6.13 MeV), 2+ (6.92 MeV), 1 (7.12 MeV), 2 (8.87 MeV),
1 (9.6 MeV), 2+ (9.85 MeV), 4+ (10.35 MeV) and 3+ + 4+ ( 11.09 MeV) states. Then,
they have been analyzed within the framework of the FRDWBA theory assuming only direct
-transfers to these states in the calculations. The obtained results for the 16 O level parameters
(Ex , c.m. , S , 2 ), extracted from the recorded deuteron spectra or via the theoretical analysis
of the measured cross section data, are reported below and discussed. To derive the most credible
2 (6.92) and 2 (7.12) values from the determined R2 ratio (see above), we have adopted DWBA
calculation independent values for the 2 (6.92) and 2 (7.12) reduced widths. Then, the 2 (7.12)
values extracted from our FRDWBA analysis of the (6 Li, d) angular distribution data have been
compared to those also derived in this work by R-matrix analyses of most recent [23,28] and
earlier [21,24,26,27] data sets for the (, ) direct capture reaction SE1 (E) factor, of 1,3 elastic
-scattering phase shift data [3032] and 16 N -delayed -spectrum data [36].

2. Experimental method, results and discussion

2.1. Experimental set up and data treatment

The experiment has been carried out at the Orsay-Institut de Physique Nuclaire high en-
ergy resolution (
E/E  2 104 ) MP tandem accelerator. A 6 Li3+ beam with an energy of
48.2 MeV (lab, i.e., E = 32.13 MeV) and an average current intensity I  100 nA was di-
rected onto 100 g/cm2 -thick self-supported natural carbon (98.9%12 C, 1.1%13 C) target foils
prepared by vacuum evaporation, placed at the center of the reaction chamber under high vac-
uum ( 106 Torr). The chosen incident energy was the highest one fitting the requirements of
the used split pole magnetic spectrometer. During the experiment, the beam intensity (measured
with a Faraday cup) and the target thickness were continuously monitored by means of a 100 m-
thick surface barrier Si-detector fixed at lab = 30 relative to the incident beam direction inside
the reaction chamber. The reaction products were, first, momentum analyzed by the split pole
magnetic spectrometer [43] with a mean solid angle
1.7 msr. Then, they were detected in
the spectrometer focal plane by a 70 cm wide detection system composed of three components
[44]: (i) a position and angle sensitive 128-anode wires drift chamber giving the position, X, of
the particle impacts followed by (ii) a proportional counter measuring the particle energy loss,

E, then by (iii) a plastic scintillator associated to a PM-tube through a light-guide measuring


the residual particle energy E
= (E
E). The first two detectors formed a unique and iso-
lated module filled with propane gas under a pressure of 340 mbar. Then, the energy E of each
particle (of mass M, momentum P and charge Q) deflected by the spectrometer magnetic field
to a bending radius was accurately determined from the measurement of the magnetic rigidity
given by the relation:
 1/2
B = P /Qe = 0.1438 E 2 + 2Mc2 E /Qc (1)
for relativistic particles, where B is the magnetic field intensity (in tesla units with E and
expressed in MeV and meter, respectively) and e, the electron charge.
A. Belhout et al. / Nuclear Physics A 793 (2007) 178211 183

A Sun-type computer ensured the acquisition of the reaction parameters that were stored and
visualized in form of two-dimensional (E, X), (
E, X) and (E,
E) maps for the sake of off-
line treatment. Such maps contained both deuteron groups from the 12 C(6 Li, d)16 O reaction and
triton groups from the 12 C(6 Li, t)15 O reaction. The former events were selected within contours,
then projected onto the X-axis to obtain the corresponding deuteron energy spectra. The peak
contents of the latter have been extracted by conveniently subtracting the background events and
de-convoluting the peak shapes when overlapped using Gaussian or Lorentzian fitting curves
generated by a computer programme. Angular distributions over the angular range 5  (lab) 
70 have been deduced mainly for the low lying states of 16 O within the excitation energy region
including the two 1 states of astrophysical interest at Ex = 7.12 MeV and 9.6 MeV. The energy
resolution (peak FWHM in the deuteron spectra) varied between 35 keV and 55 keV depending
on the observation angle. Based on the well known states of the 16 O nucleus, the calibration of the
energy spectra was straightforward. The overall systematic uncertainty in the measured absolute
d
differential cross sections, d ( ), was estimated to be of 15% while the relative uncertainties
(statistical and peak fitting) were of about 10%.

2.2. Energy spectra

Actually, the measured deuteron energy spectra covered an excitation energy region in 16 O
extending up to Ex 24 MeV and showed to be dominated by the K = 0+ and the K = 0 -
cluster rotational bands, the latter first negative parity band being composed of the 1 (9.6 MeV),
3 (11.6 MeV), 5 (14.6 MeV) and 7 (20.8 MeV) states. This attests that at the considered
bombarding energy, the 12 C(6 Li, d)16 O reaction substantially proceeds via the direct (one-step)
-transfer interaction mechanism for the majority of the produced states of 16 O which are thus
selectively populated at forward angles. This can be seen in Fig. 1(a) displaying a deuteron energy
spectrum taken at (lab) = 5 where is shown the low excitation energy region in 16 O. This
spectrum exhibits the following features: (i) the 0+ g.s. of 0p0h structure is weakly populated,
(ii) the 1 (7.12 MeV) state of spherical 1p1h configuration is well separated from the 2+
(6.92 MeV) state, (iii) the 1 (9.6 MeV) state is also clearly evidenced but due to its large natural
c.m. width, it partially overlaps with the 2+ (9.85 MeV) state as can be seen in Fig. 1(b) reporting
a spectrum portion that shows the measured data and corresponding fitted curves for the latter
two states and the 4+ (10.35 MeV) state, (iv) as expected, the 2 (8.87 MeV) unnatural parity
state, forbidden via a direct alpha transfer, the 2+ (9.85 MeV) state and the 3+ + 4+ (11.09 MeV)
doublet of states of small reduced -widths as being non--cluster states are weakly populated,
(v) the 1 (9.6 MeV) and 3 (11.59 MeV) broad states are also weakly populated although being
of natural parity, (vi) on the contrary, the 2+ (6.92 MeV) and 4+ (10.35 MeV) states of natural
parity, known with the 6+ (16.3 MeV) state (not shown in Fig. 1(a)), as indicated, to be 4p4h
-cluster states with large reduced -widths, are very prominent, (vii) this is also the case of the
3 (6.13 MeV) 1p1h -cluster state of natural parity but seemingly not the case of the K =
0+ band-head 0+ (6.05 MeV) first excited state of 16 O, the latter two states being fairly well
resolved, (viii) other level peaks seem to appear in the excitation energy region around 14 MeV
affected by a large reaction background. Indeed, the deuteron energy spectra were generally
affected by a continuous background starting around Ex = 11 MeV, then increasing to higher
energies and dominating the data at Ex  16 MeV. It presumably originates from the 6 Li +d
break-up in the Coulomb and nuclear fields of the 12 C target and subsequent nuclear reaction
processes involving fast neutrons (n + 15 O, n + 5 Li + 15 O, + n + p + 15 O, . . . ). Then, only
the excitation energy region extending up to 16 MeV has been rigorously analyzed, the states of
184 A. Belhout et al. / Nuclear Physics A 793 (2007) 178211

(a)

(b)

Fig. 1. (a) Deuteron energy (position) spectrum from the 12 C(6 Li, d)16 O reaction at lab = 5 . (b) Portion of (a) showing
the separation of the broad 1 (9.6 MeV) and 2+ (9.85 MeV) states.

16 O lying at higher excitation energies being treated only approximately. Upon analysis, states
of 16 O have been identified at Ex = 13.129 MeV (3 ), 13.869 MeV (4+ ), 13.980 MeV (2+ ),
14.302 MeV (4() ), 14.399 MeV (5+ ), 14.815 MeV (6+ ) and 14.926 MeV (2+ ). The high energy
resolution of the detection system allowed us to extract values of the c.m. line width for the
A. Belhout et al. / Nuclear Physics A 793 (2007) 178211 185

Table 1
Values of nuclear level parameters for the states of 16 O from this work and from the literature
Ex (MeV keV)a J c.m. (keV)b exp f Av g dir h
This work 90 MeVc 42 MeVd Acceptede (b) (b) (b)

0.00 0+ 122 5.3 116.7


6.05 0+ 77.4 18.3 51.9
6.13 3 572 147.4 424.6
6.92 2+ 882.5 111.3 771.2
7.12 1 126 23.7 102.3
8.87 2 < 35 < 20 75.5 21.6
9.6 30 1 338 60 330 30 400 50 420 20 108 25.2 82.8
9.846 1 2+ < 35 < 30 0.625 0.1 52.7 41.4
10.35 4+ < 35 34 5 26 3 1416 270 1146
11.09 3+ + 4+ < 35 < 30 0.28 0.05 263.5 315
11.60 3 620 100 770 90 800 100
13.12 12 3 123 10 110 30
a E values given with errors are those extracted from this work. The other ones are from [20]. b Measured
x
values: FWHM from the Gaussian fits to the peaks. Only the accepted values include -penetrabilities. c From
Ref. [3]. d From Ref. [2]. e From Ref. [20]. f Experimental cross section integrated from 6 to 84 and to
74 for the (3+ + 4+ ) states at 11.09 MeV. g Calculated averaged cross section (Eq. (3)). h Direct reaction cross
section: dir = exp HF .

-unbound states of 16 O, mainly the 1 (9.6 MeV) state of astrophysical interest. The inferred
Ex and c.m. level parameter values are listed in Table 1 together with previously accepted ones
reported in [20]. Notice, finally, in Fig. 1(a), that while the 3 (6.13 MeV) state is strongly
excited, the other states of negative parity also presumed to be of dominant 1p1h configuration,
i.e., the 7.12 MeV and 8.87 MeV states, are much less excited.

2.3. Cross section data

The measured 12 C(6 Li, d)16 O differential cross sections are reported in Fig. 2 where the er-
ror bars refer to statistical uncertainties and to errors originating from the fits to the deuteron
peaks in the deuteron energy spectra. They are very consistent with the shape and absolute nor-
malization of previously reported ones [1,2,6] over the common angular range explored, mainly
those taken at E6 Li = 42.1 MeV [2] and 50 MeV [6]. However, as can be noted, the cross sec-
tions from the previous study at E6 Li = 90.2 MeV [3] lie significantly below all the different
data sets and exhibit the most sharply forward-peaked angular distributions for the -cluster
states of 16 O. This energy dependence of the (6 Li, d) reaction can be confirmed by further scru-
tinising the present and previous cross section data [2,6]. Indeed, we have observed the two
following trends: (i) the current cross section values for the 0+ g.s. and the 3 (6.13 MeV), 2+
(6.92 MeV), 2 (8.87 MeV) and 4+ (10.35 MeV) states measured at E6 Li = 48.2 MeV and those
from Ref. [6] measured at E6 Li = 50 MeV are very close, our data being slightly lower, (ii) the
measured cross section values show a net decrease as the incident energy increases (see [2] and
Fig. 1 of [6]). This trend is in favour of the (somewhat controversial [36]) assumption that the
direct -transfer mechanism dominates the reaction while the compound nucleus formation re-
duces as the incident energy increases [2,3]. It also clearly manifests itself when observing the
12 C(7 Li, t)16 O angular distribution data for the 4+ (10.35 MeV) state measured previously at

E7 Li = 34 MeV [8] and (70, 101) MeV [9]. As can be seen, the 0+ g.s. and the 0+ (6.05 MeV),
186 A. Belhout et al. / Nuclear Physics A 793 (2007) 178211

Fig. 2. Experimental (scatter points) and calculated angular distributions of the first eleven states of 16 O. The solid curves
refer to the theoretical cross section (Eq. (3)). For the 1 (7.12 MeV) state, the dotted curve depicts the calculated cross
section for (N, L) = (4, 1) while for the 1 (9.6 MeV) and the 4+ (10.35 MeV) states, it represents the total calculated
cross section by Eq. (3) using the extrapolation method (see text). The dashed and dashed-dotted curves represent the
contributions to the reaction cross section of the HF component (normalized to the 0+ g.s.) and the averaged term from
Eq. (3), respectively.

3 (6.13 MeV), 2+ (6.92 MeV), 1 (7.12 MeV), 1 (9.6 MeV), and 4+ (10.35 MeV) states
exhibit sharp forward-peaked angular distributions, a fact a priory supporting the hypothesis that
these states are populated by direct reactions. In contrast, the angular distributions of the 2
(8.87 MeV) , 2+ (9.85 MeV) and 3+ + 4+ (11.09 MeV) states do not show such forward peak-
ing although the 2 state clearly indicates an increase towards forward angles. Indeed, these
states were previously assumed to be of pure compound nucleus formation (the 2 state) or
essentially of non-direct -cluster structures [2,3,8,9] but their possible (partial) formation via
multistep -transfers should not be excluded [6]. Notice the net oscillatory behavior of the angu-
lar distribution of the weakly populated 2+ (9.85) MeV state, a trend also observed previously
in the same reaction at E6 Li = 42 MeV [2]; it originates not only from the peak overlap for
this state with that of the neighboring broad 1 (9.6 MeV) state (Fig. 1(b)) but also probably
from a more complex multi-step -transfer process. Notice also the structured shape around 45
(c.m.) in the angular distribution of the 2 (8.87 MeV) unnatural parity state usually assumed
A. Belhout et al. / Nuclear Physics A 793 (2007) 178211 187

to be of pure compound nucleus formation. This observation could support the contribution of
the multi-step -transfer mechanism to this state [6]. The inferred values of the integrated exper-
imental cross section (exp ), the averaged cross section (Av ) simulating the contribution of all
interaction processes at backward angles (see next section) and the direct single-step -transfer
cross section (dir ) deduced as the simple difference between the former two ones are reported
in Table 1. Comparing the behavior of these data from this work (E6 Li = 48.2) with those of the
data from previous studies at E6 Li = 42.1 [2] and 90.2 MeV [3] (overestimating the compound
nucleus cross section, CN , see next section) clearly confirms a net energy dependence of the
12 C(6 Li, d)16 O reaction cross sections:
dir increases relative to Av (CN ) as the incident energy
increases for the majority of the examined low lying states of 16 O.

3. Analysis

3.1. FR-DWBA calculation

As indicated, at the relatively high bombarding energy of our experiment, one expects the
12 C(6 Li, d)16 O reaction to proceed substantially via the single-step-transfer interaction mecha-
nism for the states of 16 O known to have an -cluster character; indeed, these states such the 2+
(6.92 MeV) and 4+ (10.35) MeV members of the K = 0+ rotational band, appear to be selec-
tively populated at forward angles. Therefore, the angular distribution data have been analyzed,
essentially at forward angles, in terms of the FR-DWBA theory in post representation including
the full complex remnant potential term in the residual interaction [45,46]. The calculations have
been carried out using the computer code FRESCO [47] assuming that only single-step transfers
of -clusters in the (0s) relative state from the 6 Li projectile onto the 12 C target occur in the
12 C(6 Li, d)16 O reaction. Similarly as has been pointed out in a recent study of the 13 C(6 Li, d)17 O

reaction [46], we have checked that the remnant potential term does not change the relative cross
section values for the different states of 16 O. The optical model potential parameters adopted
to describe the distorted waves in the 6 Li + 12 C and d + 16 O entrance and exit channels are
reported in Table 2. They were obtained from theoretical best fits to 12 C(6 Li, 6 Li)12 C elastic
scattering differential cross section and analysing power experimental data measured previously
at E6 Li = 50 MeV [5] (for the entrance channel) and from best fits to 16 O(d, d)16 O elastic scatter-
ing differential cross section data taken at Ed = 52 MeV [48] (for the exit channel). The bound
state wave functions were computed using SaxonWoods potential form factors to describe the
binding of the -cluster to the deuteron and 12 C cores in the entrance and exit channels, respec-
tively, while the potential well depths for the + d and + 12 C systems have been adjusted to
reproduce the corresponding experimental -particle separation energies. The bound state radial
wave functions for the different states were calculated for a fixed set of the (N , L) quantum num-

Table 2
Optical model parameters for the elastic scattering reaction channels
Elastic scattering VR rR aR W rW aW VSO a rSO aSO
(MeV) (fm) (fm) (MeV) (fm) (fm) (MeV) (fm) (fm)
6 Li + 12 C at E(6 Li) = 50 MeV [5] 186.0 1.13 0.83 10.0b 2.26b 0.62b 2.0 1.16 0.39
d + 16 O at E(d) = 52 MeV [48] 76.8 1.25 0.75 11.9c 1.25c 0.75c
a The spinorbit interaction parameters are relative to the 6 Li projectile. b Volume absorption term. c Surface
absorption term (derivative form).
188 A. Belhout et al. / Nuclear Physics A 793 (2007) 178211

bers representing the number of nodes and the orbital angular momentum of the wave function
for the transferred -particle relative to the d and 12 C cores, respectively. In the calculations, the
nodes at the origin were included while those at infinity were excluded. These quantum numbers
are related by the TalmiMoshinsky formula [45], i.e.:
2N + L = Q, (2)
where Q denotes the number of excitation quanta. Notice that some ambiguity might affect the
choice of the N values depending on the assumed structures for the nuclear states (see next sec-
tion). The wave function proposed by Kubo and Hirata [49] for the + d overlap with a value
a = 1.9 fm of the potential radius and the -particle being in a 2s, (N, L) = (1, 0) state relative to
the d core has been adopted for 6 Li. Lehman and Rajan [50] have predicted that the s-component
of the 6 Li wave function should have a node located at r 1.65 fm corresponding, according
to our estimation, to a value a = 2.05 fm of the potential radius. However, the calculations per-
formed using this radius value have led to practically the same results as those obtained using the
former potential radius value. The a = 4.85 fm value has been used in our calculations for the
+ 12 C system.
Among the unbound states of 16 O lying at Ex > 7.16 MeV, the 1 (9.6 MeV) and
+
4 (10.35 MeV) -cluster states have been treated in the following way: (i) first, FR-DWBA
cross section values for these two states have been generated for positive binding energy values
10  B  300 keV, then they were extrapolated to negative energies down to the corresponding
-separation energies, (ii) in a second approach, the wave functions for these states were cal-
culated by solving the radial wave equation at the corresponding resonance energies. The depth
of the potential well describing the interaction of the + 12 C system was set to the respective
values V = 51.965 MeV and 50.2 MeV for which the phase shift passed by /2 at the resonance
energies. In this case, an upper cut-off radius value of 200 fm has been used in order to warrant
the convergence of the radial integral.
A major problem one has to consider in performing DWBA calculations [2,3,5,6,8,46] is the
high sensitivity of the results to the optical model potential parameters in the entrance channel
that should be determined with as high possible accuracy. In particular, the extracted S values
may considerably differ depending on whether the elastic scattering angular distribution data are
available over a narrow (forward) angular range or a wider one extending to backward angles.
E.g., in the study of the 13 C(6 Li, d)17 O reaction [46], theoretical fits to such data generated in
the former case proved to be unphysical when extended to higher angles where experimental
data are lacking, leading to an inadequate potential model and then to incorrect values of the in-
ferred 17 O nuclear level parameters. The situation is much more favorable for extracting realistic
level parameters for the 16 O nucleus from the 12 C(6 Li, d)16 O reaction due the long-standing and
comprehensive coverage accomplished both in the measurement and the theoretical fitting of the
6 Li + 12 C elastic scattering cross section data over extended angular and incident energy ranges

[5,6,46] (and references therein). Consequently, the optical model potential parameters required
for a reliable DWBA analysis of angular distribution data for this reaction were determined with
a very high accuracy. Their use in the current analysis thus warrants the reliability of the results
presented in Section 4 below, e.g., the extracted S and 2 values for the studied states of 16 O.

3.2. Total theoretical differential cross section and fitting procedure

The presence in the deuteron spectra of the 2 (8.87 MeV), 2+ (9.85 MeV) and (3+ + 4+ )
(11.09 MeV) states of small reduced -widths presumed to be of non--cluster structure [2,3,8]
A. Belhout et al. / Nuclear Physics A 793 (2007) 178211 189

can be considered as the signature of reaction mechanisms other than the single-step -transfer
which dominates the measured forward angle cross section data in the case of -cluster states,
like the 2+ (6.92 MeV) and 4+ (9.6 MeV) members of the K = 0+ rotational band. Indeed, the
former states seem to show rather complex angular distributions, mainly the 2 state (see Fig. 2).
It must be noted that single particle transfer reactions similarly showing no forward peaked angu-
lar distribution data were previously interpreted to be due either to compound nucleus formation
[51] or to multi-step processes [52]. Both two theoretical models predicted nearly isotropic cross
section angular distributions for sd shell nuclei which were found to be compatible with ex-
perimental data [51,52]. In the current more complex 12 C(6 Li, d)16 O -cluster transfer reaction,
although they cannot be considered as dominating the cross section, the compound nucleus and
multi-step -transfer mechanisms rigorously need to be evaluated in the calculation in order
to correctly account for the measured angular distribution data. While the former mechanism,
favoured at relatively high excitation energies where the level density is high, can be easily
evaluated by the HF statistical model [53], CRC/CCBA calculations for describing multi-step
-transfers are considerably more complex to handle [6]. One is forced to admit, however, that
even the HF model calculation obviously suffers serious problems arising from the rather ar-
bitrary normalization to a given particular state, which leads one to misestimate the calculated
d J
( d )CN differential cross sections for other states. In previous works [2,7,8] (and references
therein), the 2 (8.87 MeV) unnatural parity state of 16 O was systematically presumed to be of
predominant compound nucleus formation and then taken as the reference state for normalizing
the calculated HF cross sections, scaled by a 2J + 1 factor to fit the data for other states. In con-
trast, the compound nucleus character of this state has been strongly cast into doubt recently in
the work of Keeley et al. [6], as discussed below. Therefore, we have finally adopted the follow-
ing procedure in our FRDWBA analysis of the measured (6 Li, d) cross section data. A constant
d
component, ( d )Av , obtained by averaging, for each state of 16 O, the measured differential cross
section at backward angles (c.m.  60 ) where single-step -transfers can be neglected, has been
incoherently added to the FRDWBA differential cross section to simulate the theoretical total dif-
ferential cross section. Thus, the following model expression was fitted to the total experimental
differential cross section at forward angles, i.e.:
     
d d d
= C 2 S
S + . (3)
d exp d DWBA d Av

In this expression, C is a target nucleus isospin coupling ClebschGordon coefficient and S

and S are the phenomenological -particle spectroscopic factors of the projectile and target
nuclei, respectively, for the transfer of an -particle onto 12 C to form a given state of 16 O. These
factors take into account the probability that the 6 Li and 16 O nuclei are likely composed of an
-cluster and a deuteron or of an -cluster and a 12 C core, respectively. They are measures of the
overlaps of the respective initial and final state wave functions, i.e.:

  2
S
= 6 Lid , (4)
  2
S = 16 O12 C . (5)

First, theoretical differential cross section curves were generated as in Eq. (3), then fitted to
the corresponding experimental values such as they best accounted essentially for forward angle
190 A. Belhout et al. / Nuclear Physics A 793 (2007) 178211

data. To check the quality of the theoretical fits to the measured cross section data, a reduced
chi-2 function, defined as:
   d 
N
d
1 d exp. (i ) d th. (i ) 2
red. =
2
 d  , (6)
N
d exp. (i )
i=1
d
where N is the number of angular distribution data points (angles) and
( d )exp. (i ) the ab-
solute uncertainty on each measured differential cross section value, has been minimized via the
least-squares method. The S = S
S product involved in the DWBA cross section (Eq. (3)) was
varied as the only fit parameter for each state of 16 O with values ranging between S and S
S.
The relative uncertainty on S was evaluated by the relation:

S
S
S

= +
, (7)
S S S
with that on S
being fixed for all the states of 16 O as will become clearer below.

4. Results and discussions

4.1. Angular distributions

The calculated differential cross sections for the first 11 low-lying states of 16 O produced
in the 12 C(6 Li, d)16 O reaction are compared to their experimental counterparts in Fig. 2. It is
remarkable, first, that the experimental data for the 0+ g.s., the 0+ (6.05 MeV), 1 (7.12 MeV)
and 1 (9.6 MeV) states are quite acceptably reproduced by the calculations. The data are less
satisfactorily accounted for by calculation concerning the 4+ (10.35 MeV) -cluster state. This
state and the 1 (9.6 MeV) state have been treated using the two methods described above
leading to similar quality fits to the corresponding data (see Fig. 2), but to different values of the
associated S spectroscopic factors for these two unbound states. The measured data for the 3
(6.13 MeV) and 2+ (6.92 MeV) states are also acceptably accounted for by the calculations, in
average. Nevertheless, as has also been found in all previous theoretical analyses, the calculated
curves for these two states evolve in opposite direction relative to experiment at the smallest
angles (c.m.  10 ).
Prior to adopting the method based on Eq. (3) for analysing the (6 Li, d) angular distribution
d J d
data, we have performed HF calculations by normalizing ( d )CN {replacing ( d )Av in Eq. (3)}
+
either to the 2 (8.87 MeV) state as in previous works [2,7,8] or to the 0 g.s. backward an-
gle data. In the latter case, the calculated HF cross section curves consistently lay below the
measured data for all other considered states of 16 O, in contrast to the former case, and corre-
sponding final (DWBA + HF) fits to experimental data of general good quality are obtained as
when using Eq. (3). The generated HF curves for the 2 (8.87 MeV), 2+ (9.85 MeV) and (3+ +
4+ ) (11.09 MeV) states via normalization to the 0+ g.s. are also shown in Fig. 2 (dashed curves)
together with the averaged cross section components of Eq. (3) (dashed-dotted curves). As can
be observed, the compound nucleus component represents only a small fraction of the measured
total cross section for these states presumed, in previous works [2,3,79], to be of either nearly
pure compound nucleus formation (the 2 state) or essentially non--cluster character. One can
also see that the anomalous structure at c.m.  45 in the angular distribution data of the 2
(8.87 MeV) state is clearly irrelevant to the compound nucleus mechanism, the calculated HF
angular distribution being unstructured. Therefore, manifestly high-order sequential -transfer
A. Belhout et al. / Nuclear Physics A 793 (2007) 178211 191

Table 3
Spectroscopic factor values for the 16 O states relative to that of the 2+ (6.92 MeV) state
J (Ex , MeV) (N, L) S /S (2+ )
6 Li 6 Li 6 Li 7 Li SU(3)b
(48 MeV)a (42 MeV)b (90 MeV)c (34 MeV)d
0+ (0.00) (2, 0) 0.91 0.25 7.4 1.85 2.3 1.26
0+ (6.05) (4, 0) 0.34 0.13 0.6 0.87 0.6 1.04
3 (6.13) (1, 3) 0.78 0.42 0.8 0.89 0.5 0.79
2+ (6.92) (3, 2) 1.00 1.0 1.00 1.0 1.00
1 (7.12) (2, 1) 0.44 0.14 0.8 0.63 0.5 0.20
(4, 1) 0.23 0.09 0.3 0.24 0.2

0.3 0.09e
1 (9.6) (4, 1) 0.91 0.28 0.6 0.58 1.1 1.05
2+ (9.85) (2, 2)  0.02  0.05  0.01  0.01  0.01
4+ (10.35) (2, 4) 0.44 0.16 0.4 0.94 1.8 0.88
4+ (11.09) (2, 4)  0.08  0.1  0.05  0.11  0.06
a This work: S spectroscopic factors relative to that for the 2+ (6.92 MeV) state, S (2+ ) = 0.37 0.11. b From

Ref. [2]. c From Ref. [3]. d From Ref. [13]. e Value obtained for mixing configuration (see text).

transitions occur to this state [6] forbidden in single-step -transfers. The 2+ (9.85 MeV) and
4+ (11.09 MeV) states also, in contrast to the members of the 4p4h, K = 0+ -cluster rota-
tional band, cannot be populated by single-step -transfer transitions, whence the inability of the
FRDWBA calculations to reproduce the corresponding angular distributions.
For all studied states, the (N , L) quantum numbers assigned (see Table 3) to the c.m. motion
of the -cluster within the 16 O nucleus are those expected to be dominant given by the SU(3)
group model [54] or those derived by the orthogonality condition model (OCM) [55]. In the
former model, the 1 (7.12 MeV) state has been assigned the values (N, L) = (2, 1) correspond-
ing to the 1p1h configuration. Besides, assuming a different structure, calculations have been
performed for the values (N, L) = (4, 1) leading also to a satisfactory fit of the corresponding
angular distribution data with a smaller reduced -width value. However, by considering the
-emission following the -delayed decay of 16 N to this state, we have checked that the 1p1h
configuration is, indeed, the most probable. Furthermore, assuming a more realistic configuration
mixing for this state, we have also performed a corresponding FRDWBA calculation (see below).
The results derived here from our analysis of the (6 Li, d) angular distribution data at E6 Li =
48.2 MeV thus clearly show evidence that interaction mechanisms other than the direct single-
step -transfer also occur for the majority of studied states of 16 O: the 3 (6.13 MeV), 2+
(6.92 MeV), 2 (8.87 MeV), 2+ (9.85 MeV) and 3+ + 4+ (11.09 MeV) states (see Fig. 2). They
likely consist of compound nucleus formation, multistep -transfers into these states and other
direct reaction processes.
As noted above, assuming multi-step -transfers, Keeley et al. [6] have recently compared
their CRC calculations for the transitions to the 0+ g.s. and the 3 (6.13 MeV), 2+ (6.92 MeV),
2 (8.87 MeV) and 4+ (10.35 MeV)

excited states of 16 O with measured analysing power and
cross section data for the 12 C(6 Li, d)16 O reaction at E6Li

= 34 MeV and 50 MeV. The follow-
ing two main results have been inferred in that work: (i) multi-step -transfer transitions were
found unambiguously significant only to the 0+ g.s. and the 3 (6.13 MeV) state while such
transitions to the 2+ (6.92 MeV) and 4+ (10.35 MeV) states proved to be weaker at both 6 Li
ion energies, suggesting that the latter states could be modeled by a direct single-step -transfer,
192 A. Belhout et al. / Nuclear Physics A 793 (2007) 178211

(ii) evidence has been found that the 2 (8.87 MeV) unnatural parity state (forbidden as a di-
rect -transfer) would only be populated via multistep -transfer processes. These results have
led the authors to the conclusion that the DWBA theory would be inappropriate to describe the
-transitions to the 0+ g.s. and the 3 (6.13 MeV) state, and to put into question the reliability
of the hypothesis that the 2 (8.87 MeV) state of 16 O is of a predominant compound nucleus
character as previously assumed for deriving + 12 C spectroscopic factors via the analysis of
-transfer angular distribution data by the classical DWBA + HF method. The latter claim was
supported by the high values of the iT11 vector analyzing power measured by these authors for
this state at E6Li

= 50 MeV, indeed not compatible with compound nucleus formation. It also
gains credit through the observed somewhat structured shape of the angular distribution for the
2 state never fully reproduced in previous works by HF calculations normalized to this state.
However, the cross section curves generated for this state by CRC calculations [6] assuming only
transfers via the 2+ (4.44 MeV) state of 12 C lay very far below the corresponding experimental
data: by a factor of 10 at E6 Li = 34 MeV and a factor of 70 at E6 Li = 50 MeV! Such ob-
served very large discrepancies probably could not be removed out only by including additional
multi-step -transfer paths in the CRC calculation, such the transitions via the 3 (9.64 MeV)
state of 12 C or via other states of the 16 O nucleus itself. Consequently, it turns out that the 2
(8.87 MeV) state has obviously a much more complex structure than being of predominant com-
pound nucleus formation (see Fig. 2) as believed earlier [2,3,7,8] or only populated by multi-step
-transfers as suggested in Ref. [6]. The non-vanishing values of iT11 for the 8.87 MeV state
combined with its unnatural parity can be interpreted as an evidence that direct reactions (not
including one-step -transfers) are dominant, given that multi-step -transfer transitions to this
state are visibly weak like compound nucleus formation. In addition to the latter two mecha-
nisms, direct reaction processes such as sequential 3 He and neutron transfers, (6 Li, t)(t, d), likely
contribute in populating the 2 (8.87 MeV) state of 16 O. Besides, it is remarkable that the 0+
g.s. of 16 O is quite satisfactorily reproduced by our DWBA calculations while it was claimed in
[6] to be significantly populated by multi-step -transfers. Notice also that probably due to the
limited energy resolution of the detection system used in the experiment reported in Ref. [6], no
independent experimental data have been reported for the 0+ (6.05 MeV) and 2+ (9.85 MeV)
states nor for the two astrophysically relevant 1 (7.12 MeV) and (9.6 MeV) states of 16 O for
which multistep CRC calculations have then not been performed.
Notice, finally, that by normalizing the compound nucleus cross section (used in Eq. (3) in-
stead of the average constant term) to that of the 2 (8.87 MeV) state as in earlier works, we
obtained the best quality fits possible to the angular distribution data, given that the HF model
yields increasing values at forward angles which considerably improves the agreement between
theory and experiment. However, as stated, this procedure strongly overpredicts the calculated
HF cross sections leading to rather unrealistic theoretical fits to the measured data.

4.2. Spectroscopic factors

4.2.1. Asymptotic normalization coefficient for 6 Li


The value S
= 0.8 for the spectroscopic factor of the ( + d) clustering in 6 Li deduced
previously from an analysis [56] of the three body break-up cross sections and the analyzing
p)2 H reaction has been adopted here in Eq. (3). This value is comprised
powers in the 6 Li(p,
between that (S 0.7) predicted by the three body model [50,57] and the value (S
0.9)

obtained from microscopic and resonating group models [5862], and can thus be considered as
A. Belhout et al. / Nuclear Physics A 793 (2007) 178211 193

reliable to within
S
= 0.1, which leads to a relative uncertainty
S
/S
= 0.125, taken the
same for all the considered states of 16 O.
In term of the so-called asymptotic normalization coefficient (ANC, denoted C2 , see below for
definition), the preceding S
value for the 6 Li ion corresponds to a value C2 = 5.14 0.64 fm1 ,
which is in very good agreement with the value C2 = 5.3 0.5 fm1 reported in the previous
literature (see Ref. [4]).

4.2.2. Results for -cluster states of 16 O


The S factor values and associated uncertainties deduced in the present analysis for the 16 O
states, are reported in Table 3 relative to that (S = 0.37 0.11) of the 2+ (6.92 MeV) state.
They are compared to those derived by Becchetti et al. in their (6 Li, d), (7 Li, t) experiments at
E6 Li = 42 MeV [2] and 90 MeV [3], and at E7 Li = 34 MeV [8] as well as to values derived by the
SU(3) group model [54]. First, as expected, significantly lower values are obtained in this work
as in [2,3,8] for the 2+ and 4+ non--cluster states of 16 O at Ex 9.85 MeV and 11.09 MeV,
respectively, compared to those for the other states reported in Table 3. One can clearly see that
while the current and previous S /S (2+ ) values for all states are of the same order of mag-
nitude (S /S (2+ )  1) as those derived by the SU(3) model calculation [54], our results are
in rather fair agreement (within the involved uncertainties) with the latter, except for the 0+
(6.05 MeV) state. Our values are therefore comparable, although with a significant spreading,
to those reported in Refs. [2,3,8] at different Li ion energies, except the case of the 0+ g.s. for
which Becchetti et al. have found values largely exceeding unity at E6 Li = 42 MeV and 90 MeV
[2,3] and E7 Li = 34 MeV [8]. In particular, our value (0.776 0.42) for the 3 (6.13 MeV) state
is consistent with previous ones. Concerning the 1 (9.6 MeV) and 4+ (10.35 MeV) unbound
states, the following values of the relative S factor have been derived in the current analy-
sis: S /S (2+ ) = 0.50 and 1.82, respectively, via the cross section extrapolation method, and
S /S (2+ ) = 0.91 0.28 and 0.44 0.16, respectively, by considering both two states as effec-
tively unbound. Note that in terms of the S absolute values, large differences were also observed
between our value for the 0+ g.s. (S = 0.34 0.07) and those from Refs. [2,3] (S = 10 at
E6 Li = 42 MeV, S = 7.6 at E6 Li = 90 MeV) while our value is consistent with that (S = 0.38)
reported in [8] for E7 Li = 34 MeV. In contrast, our absolute value for this state is very consistent
with that (S = 0.32) obtained by Drummer et al. [5] for the same value of the + 12 C potential
radius (a = 4.85 fm).

4.2.3. Special case of the 1 (7.12 MeV) state of 16 O


As can be seen in Table 3, the S /S (2+ ) values obtained for the 1 (7.12 MeV) state of
astrophysical concern are consistent with those reported in Refs. [2,3,8], especially for the minor
contribution of (N, L) = (4, 1) quantum number values. However, this state is more likely due to
a configuration mixing [7] involving both (N, L) = (2, 1) and (4, 1) assignments. This configu-
ration mixing has been considered in the calculated DWBA cross section as a linear combination
of the two (N, L) independent contributions. This leads, indeed, to a better theoretical fit to ex-
perimental data, accounting very satisfactorily for the corresponding whole angular distribution
as can be seen in Fig. 3(a). This improved fit corresponds to mixing proportions of 80% and
20%, respectively, as derived from the reduced chi-2 minimization search (see the red 2 contours
+
in Fig. 3b). The resulting associated relative spectroscopic factor value, S /S (2 ) = 0.3 0.09,
fortuitously equals the mean value obtained from the two separate contributions.
It must be noted that we have also observed, as has been demonstrated previously [2,3,8],
that while the absolute S values are very model-dependent, the S /S (2+ ) or S /S (4+ ) rela-
194 A. Belhout et al. / Nuclear Physics A 793 (2007) 178211

Fig. 3. (a) Experimental (scatter points) and calculated (Eq. (3)) angular distributions of the 1 (7.12 MeV) state: the
dashed and dotted curves represent the total calculated cross section for the (N, L) = (4, 1) and (2, 1) assignments,
respectively, while the solid curve represents the total calculated cross section obtained by considering a configuration
mixing with proportions 1,2 = 80% for (N, L) = (2, 1) and 1,4 = 20% for (N, L) = (4, 1) yielding the best fit to the
2 function versus
data as shown by (b) the contour plot representing the variation of the red 1,4 and 1,2 .

tive values show less such a model dependence. However, the least model-dependent parameters
that can be determined in (6 Li, d) or (7 Li, t) transfer reaction experiments [2,3,8] are the reduced
-widths discussed below, which are the relevant quantities both for -cluster model calculations
in nuclear physics and for nuclear astrophysics calculations. The indirect method based on ex-
tracting the asymptotic normalization coefficients is currently used [6368] to perform reliable
calculations.

4.3. Reduced -widths

4.3.1. ANCs and associated R-matrix reduced -widths


As has been suggested first in [63,64], the spectroscopic factors extracted by DWBA analysis
of measured particletransfer reaction cross section data can be converted into ANCs in view
to determine reduced particlewidth parameters which can be used to extract astrophysical fac-
A. Belhout et al. / Nuclear Physics A 793 (2007) 178211 195

tors for corresponding radiative particlecapture reaction processes populating loosely bound
final states. Several nuclear astrophysics works have been carried out using the ANC method
[6368] (and references therein). Contrarily to the equivalent reduced particlewidths, the
ANCs present the advantage of being independent on the somewhat arbitrary channel radius a.
A necessary condition to be fulfilled, however, is that this method applies only to reactions hav-
ing an essentially peripheral character. This nuclear surface interaction property expectably holds
for incident energy values lying below the Coulomb barrier [4], and should gradually attenuate
and vanish as the incident energy increases above the potential barrier unless a strong absorption
of the projectile occurs at the nuclear surface, in which case the reaction above the barrier can
also be peripheral [66] (and Ref. [7] therein).
In the case of the current study of the (6 Li, d) transfer reaction at E6 Li = 48.2 MeV, difficulties
are predictable in this respect, in addition to binding energy considerations. For the 0+ g.s., 0+
(6.05 MeV) and 3 (6.13 MeV) tightly bound states of 16 O, the -particle separation energy
amounts up to 7.162 MeV (for the 0+ g.s.), and may thus be too large for the (6 Li, d) transfer
reaction to conserve a peripheral character. One obviously expects the situation to be much less
critical for the moderately bound 2+ (6.92 MeV) and 1 (7.12 MeV) states of astrophysical
interest in 16 O. In view to derive the most reliable as possible spectroscopic information on these
two states and better estimate the 12 C(, )16 O capture reaction astrophysical SE1 (E) factor, we
have then examined the possibility of converting our preceding (6 Li, d) reaction S factors into
ANCs, C 2 (J ). The latter two relevant quantities relate [4,6567] as:
 
C 2 J = S b 2 , (8)
where b is the single-particle ANC defining the amplitude of the tail of the bound state wave
function, u(r), relative to the Whittaker function, W (r), in the asymptotic region, i.e.:
u(r)  bW (r), r > a. (9)
The detailed connection between the ANCs and nuclear level fit parameters within the R-
matrix theory is made in Ref. [65]. E.g., the observed reduced -width of a subthreshold bound
state of 16 O, 2 (a), can be derived by the relation:
h 2 W 2 (a) 2
2 (a) = |C| , (10)
2 a
where is the reduced mass of the interacting nuclei. As the corresponding ANC (Eqs. (8), (9)),
the reduced -width of a nuclear state of 16 O thus gives the probability of its fragmentation into
the + 12 C channel via the strength of the radial wave function u (r) in the asymptotic external
region. By scaling 2 (a) to the Wigner single-particle limit, W2 (a) = 3h 2 /2a 2 , one gets the
dimensionless relative reduced -width:
2 (a)
2 (a) = . (11)
W2 (a)
Notice that the R-matrix theory usually gives the particlewidths as formal quantities (2 ,
from which the observed widths (2 , 2 ) are obtained through division of the former by the
2 )
factor (1 + l2 dSc /dE) [21,38,65] due to the energy dependence of the resonance level shift
function, Sc (E), in reaction channel c ( ) [38,65].
To check the peripheral character of the studied (6 Li, d) transfer reaction for the considered
kinematics, two methods have been used [66,67]. First, following the procedure described in
Ref. [67], we have examined the behavior of the calculated distorted wave (DW) transfer cross
196 A. Belhout et al. / Nuclear Physics A 793 (2007) 178211

Fig. 4. DWBA cross sections integrated over the ranges = 025 and = 012 for the 2+ (6.92 MeV) (! symbol)
and 1 (7.12 MeV) (P symbol) states in, respectively, function of the cutoff radius.

sections (integrated over the first maximum of the reaction angular distribution) versus the cut-
off radius for the two relevant subthreshold states of 16 O. The results found are reported in Fig. 4
which clearly shows that the DW values essentially lie inside the nuclear interaction region
(1.4(A6 Li +A12 C ) = 5.75 fm). In a second test, we have explored the variation of the b2 /DW (5 )
1/3 1/3

ratio versus the used binding potential form factor parameters. While one expects this ratio to
remain nearly constant when the reaction is peripheral even while DW varies relatively strongly
[66], we found it to actually exhibit substantial variations. Therefore, the (6 Li, d) reaction appears
to be of non-peripheral character for the present kinematics (E6 Li = 48.2 MeV).
Converting, nevertheless, our above S results into ANCs for the 1 (7.12 MeV) and 2+
(6.92 MeV) weakly bound states of 16 O, we obtained the following results, respectively, in fm1
units: C 2 (1 ) = 25.69 6.49 1028 , C 2 (2+ ) = 11.87 3.53 1010 . In their study of the same
(6 Li, d) -transfer reaction at sub-Coulomb barrier energies, Brune et al. [4] have extracted ex-
perimental values of C 2 (1 ) = 4.33 0.84 1028 fm1 and C 2 (2+ ) = 1.29 0.23 1010 fm1 .
Other experimental estimates derived for the 2+ (6.92 MeV) state via R-matrix analyses are the
following, in fm1 units: C 2 (2+ ) = 2.372 0.55 1010 [32], C 2 (2+ ) = 16.16 1010 (given
without error quoted) [69] and C 2 (2+ ) = 5.198+1.61
1.55 10 [70]; the first two values were ex-
10
12
tracted [32,69] from elastic -scattering on C yield and phase-shift data, while the latter one
was obtained [70] from 12 C(, )16 O direct capture cascade transition data. Besides, theoretical
values of C 2 (2+ ) = 2.088 0.24 1010 fm1 and C 2 (2+ ) = 1.796 1010 fm1 have been ob-
tained from inverse potential quantum scattering theory [68] and microscopic multi-configuration
cluster model [41] calculations, respectively.
Besides, studying the variations of the reduced -widths given by Eq. (10) for the above two
subthreshold states of 16 O versus the channel radius, we obtained the results shown in Fig. 5.
A considerably long range of the nuclear interaction is thus visible in the (6 Li, d) transfer reaction
for the current kinematics, extending up to a channel radius a = 7.7 fm (corresponding to a
Wigner single-particle limit W2  0.35 MeV) where the interaction potential is vanishing. Our
preceding ANC values for the two states imply the following reduced -width values: 2 = 0.16
(for a = 5.5 fm corresponding to W2  0.69 MeV) and 0.014 (a = 7.7 fm) for the 1 (7.12 MeV)
state, and 2 = 1.22 (a = 5.5 fm) and 0.071 (a = 7.7 fm) for the 2+ (6.92 MeV) state. It is
A. Belhout et al. / Nuclear Physics A 793 (2007) 178211 197

Fig. 5. Variation of the observed reduced -widths 2 (7.12) (a) and 2 (6.92) (b) solid curves, Eq. (12), dashed curves,
Eq. (10) and of the binding potential of the system + 12 C (c) in function of the channel radius.

remarkable that the derived 2 (6.92) value for a = 5.5 fm is too high, exceeding the Wigner
single-particle limit. In contrast, the expectably more reliable values for both two states for the
channel radius a = 7.7 fm are compatible with those resulting from the R-matrix analyses of
12 C(, )16 O data (see Section 5).

The validity of the ANC estimates from previous works is discussed in Ref. [68]. Likely due
to the apparently non-peripheral character of the (6 Li, d) transfer reaction at E6 Li = 48.2 MeV,
our ANC values appear to differ by at least a factor of 3.5 within the involved uncertainties
from previous ones reported in Refs. [4,41,68] and judged as reliable values [68]. As noted by
the author of Ref. [68], who argued that the uncertainty on the C 2 (2+ ) experimental value from
Ref. [32] is significantly underestimated, the extraction of C 2 (2+ ) from experiment may be very
delicate. It is the case, indeed, not only from elastic -scattering on 12 C yield and phase shift
data but also from -transfer reaction data depending on the adopted reaction kinematics.
In the following of this section, we, first, proceed to extract (and discuss) the reduced -widths
for low-lying states of 16 O from our DWBA analysis of the measured (6 Li, d) angular distribution
data according to the classical method used, e.g., in [2,3,5,7,8,10,71]. Then, the obtained results
will be confronted with those extracted, in a second step, via the R-matrix analysis of data sets
for the 12 C(, )16 O helium burning reaction presented in Section 5.
198 A. Belhout et al. / Nuclear Physics A 793 (2007) 178211

4.3.2. Reduced -widths following the DWBA method


The observed absolute reduced -width, 2 (a), associated with a bound state formed by a
partial wave of transferred orbital angular momentum, , from the + 12 C channel is defined by
[72,73]:
h 2  2
,
2
(a) = u (a) , (12)
2a
where u (a) is the relative motion radial wave function (the same as in Eq. (9)). The radial wave
function usually utilized in the DWBA model calculations [54,71], RDWBA (a), relates to the
preceding one (R (r) = u (r)/r) via the S spectroscopic factor, i.e.:

R (a) = S RDWBA (a). (13)

The 2 (a) {2 (a)} nuclear level structure parameter, Eqs. (12), (13), is thus determined by the
value of the exact internal radial wave function at channel radius r = a, and represents a mea-
sure of the probability of -clustering at the nuclear surface where -transfer reactions induced
by the strongly absorbed 6,7 Li cluster projectiles are primarily sensitive to u (r). When evalu-
ated at a conveniently chosen channel radius appropriate to the considered nuclear reaction and
defining the interaction surface, then 2 (a) should be independent on the precise form of the
DWBA model wave function [71]. In particular, for unbound states it should be insensitive to
extrapolation into the continuum as the corresponding (partial) -width given by:
= 22 P (a), (14)
where P (a) is the Coulomb barrier penetrability, to be calculated at the considered resonance
energy.
The definition of 2 (a) via Eqs. (12), (13) is obviously equivalent to that given by Eqs. (10),
(8) for r > a, provided the involved channel radius a corresponds to the effective nuclear interac-
tion surface. Notice that the latter, defined by the channel radius, differs from the nuclear surface
associated with the nuclear interaction potential usually defined via the diffuseness parameter,
aR , giving its thickness corresponding to potential strength fractions between 90% and 10%
of the potential height around the nuclear radius, R (see Table 2). The determination of the re-
duced -widths by Eqs. (12), (13) is also subject to the problems of the reaction under study not
being very peripheral.
The spectroscopic factors obtained in the present DWBA calculations (through Eqs. (12), (13))
have been used to determine the variations of the reduced -widths in function of the channel
radius a. This is illustrated by solid curves in Figs. 5(a), (b) in the cases of the 2+ (6.92 MeV) and
1 (7.12 MeV) states of 16 O. These variations together with that of the interaction potential also
shown in Fig. 5(c) indicate that the asymptotic behavior of the wave functions can be considered
to start at the channel radius a = 6.5 fm for which the amplitude of our WoodsSaxon nuclear
potential reaches 10% of its maximum value, and vanishes beyond the channel radius value
a = 7.7 fm. Therefore, one expects the effective nuclear surface to lie, with some diffuseness,
between these two channel radii. Also, in the following, calculations will be carried out for these
two a values, and discussed in combination with our forthcoming phenomenological R-matrix
analysis (Section 5).
The choice of the adequate value of the channel radius a for performing valuable DWBA or
R-matrix calculations has been a subject often discussed in the literature. Thus, a values ranging
from 4.5 fm to 7.5 fm have been proposed [2,3,5,8,9,21,23,27,32,3639,54,71,74,75] with the
A. Belhout et al. / Nuclear Physics A 793 (2007) 178211 199

Table 4
Reduced -width values for the 16 O states relative to that of the 2+ (6.92 MeV) state (R2 ratio), for the channel radius
a = 5.5 fm
J (Ex , MeV) (N, L) 2 /2 (2+ )
6 Li 6 Li 6 Li 7 Li 7 Li 7 Li
(48 MeV)a (42 MeV)b (90 MeV)c (34 MeV)d (38 MeV)e (70 MeV)f
0+ (0.00) (2, 0) 0.59 0.16 0.93 0.18 0.29
0+ (6.05) (4, 0) 0.32 0.12 0.38 1.10 0.68
3 (6.13) (1, 3) 0.78 0.42 0.23 0.22 0.14
2+ (6.92) (3, 2) 1.00 1.00 1.00 1.00
1 (7.12) (2, 1) 0.48 0.15 0.53 0.39 0.30 0.18 0.17 0.05
(4, 1) 0.13 0.05 0.24 0.30 0.21
0.3 0.09g
1 (9.6) (4, 1) 0.52 0.16 0.30 0.60 0.76
4+ (10.35) (2, 4) 0.58 0.21 0.25 0.47 1.09
a This work: 2 reduced widths relative to that for the 2+ (6.92 MeV) state, 2 (2+ ) = 0.185

0.055. b Ref. [2]. c Ref. [3]. d Ref. [8]. e Ref. [7]. f Ref. [9]. g Value obtained for mixing configuration
(see text).

value a = 6.5 fm being often adopted in R-matrix analyses [4,21,23,36,37,69] of data sets of
interest to the 12 C(, ) problem.
In the previous DWBA analyses of the (6 Li, d) and (7 Li, t) transfer reaction data, however,
Becchetti et al. [2,3,8,9] have adopted the value a = 5.4 fm that has also been used (together
+
value a = 7.2 fm) by Drummer et al. [5] in their study of the 0 g.s. of O produced in
with the 16

the ( Li, d) reaction at E6 Li = 50 MeV. Besides, a careful R-matrix analysis of the 12 C(, )12 C
6

elastic scattering yield data has previously been performed [32] for a = 5.5 fm, and this channel
radius value has also been adopted, e.g., in Refs. [27,38] and more recently in Ref. [22]. In
order to compare our results to previous ones, we then also consider the channel radius value
a = 5.5 fm in our calculations.
Notice that in the major part of the previous R-matrix analyses of the 12 C(, )16 O data,
the choice of the adopted interaction radius value was made only on the basis of 2 function
minimizations. Furthermore, the reduced -widths, as well as other fit parameters, were not
calculated from basis functions.

4.3.2.1. Calculation for a = 5.5 fm The obtained 2 parameter values for the studied -cluster
states of 16 O, scaled to that of the 2+ (6.92 MeV) state, are listed in Table 4. The comparison
of these results to previously reported ones derived via FRDWBA + HF analyses [2,7,8] (or
only by FRDWBA analyses [3,9] where the S and 2 values have been more reliably derived
without use of overpredicted compound nucleus cross section contributions) shows the following
trends: (i) similarly as for the S /S (2+ ) results, all the (present and previous) 2 /2 (2+ ) values
for each given state of 16 O are of the same order of magnitude, although a general significant
dispersion is observed between them, (ii) our values show an overall consistency with those
derived by Becchetti et al. from their (6 Li, d) study at 90.2 MeV [3], except for the 0+ g.s., the 0+
(6.05 MeV) and 3 (6.13 MeV) states: indeed, while our values for the first two states are higher
by a factor of 3.4, the contrary holds for the latter state whose angular distribution data are the
best accounted for by our DWBA calculation (see Fig. 2), (iii) a high relative uncertainty of 54%
results from our DWBA calculation in the case of the 3 (6.13 MeV) state whose cross section
data are not well fitted neither by the previous [2,3,6] nor by the current (see Fig. 2) theoretical
200 A. Belhout et al. / Nuclear Physics A 793 (2007) 178211

analyses, a fact which is concordant with likely significant multi-step -transfer transitions to
this state [6], (iv) our values for the two 1 (7.12 MeV) and (9.6 MeV) astrophysically relevant
states are fairly consistent with previous ones, mainly those from Ref. [3] (E6 Li = 90.2 MeV).
Therefore, one can consider that values in overall consistency with those derived in previous
experiments on the (6 Li, d) and (7 Li, t) -transfer reactions [2,3,8,9] are obtained for the states
of 16 O with (N, L) assignments mainly based on the SU(3) group model.
To minimize the dependence of the results on the DWBA model calculation, we defined the
following ratios, e.g., for the 1 (7.12 MeV) state [2,3,8,9]:
2 (7.12)
R1 = , (15)
2 (9.6)
2 (7.12)
R2 = , (16)
2 (6.92)
and
2 (7.12)
R3 = , (17)
2 (10.35)
for which the respective values R1 = 0.57 0.13, R2 = 0.30 0.09 and R3 = 0.51 0.15 have
been obtained.
Due to the facts that the 1 (7.12 MeV) and 2+ (6.92 MeV) states are close in excitation
energy (then they are not subject to Q-value effects) and that the latter state is much less affected
by angular momentum mismatches than the 1 (9.6 MeV) and 4+ (10.35 MeV) states, the R2
ratio is expected to better minimize the dependence on the FRDWBA calculations [2,3,8]. While
the absolute reduced width values are strongly channel radius dependent, the R13 relative values
should be less sensitive to a [38], as will be explicited below. The R2 value found here can
be, first, compared to previous ones extracted from similar studies of the (6 Li, d) [2,3] and the
(7 Li, t) [79] -transfer reactions performed at E6 Li,7 Li  30 MeV. Our value is consistent with
the values obtained from the 12 C(6 Li, d)16 O reaction at E6 Li = 42.1 MeV [2] and 90.2 MeV [3],
i.e., 0.2 < R2 < 0.6 depending on the (N , L) assignments and R2 = 0.39 for the (N, L) = (2, 1)
assignments to the 1 (7.12 MeV) state, respectively. It is also consistent with the value one
can deduce from the Brune et al. DWBA analysis of their (6 Li, d) and (7 Li, t) total cross section
data measured at sub-Coulomb energies, R2 = 0.42 [4], the Plaga et al. R-matrix analysis of the
12 C(, ) phase shift data, R = 0.46 [31] and the recent Tischhauser et al. R-matrix analysis of
2
the 12 C(, ) excitation function data, R2 = 0.36 [32]. In contrast, our R2 value is larger by a
factor of 1.9 than that extracted from the previous Barker and Kajino R-matrix analysis of the
12 C(, )16 O capture reaction data, R = 0.16 [38].
2
From Eq. (12) through Eqs. (11), (13), one gets the dimensionless reduced -width values
for a = 5.5 fm: 2 (6.92) = 0.185 0.055 and 2 (7.12) = 0.055 0.012 (the absolute errors
on the 2 values include the
S
/S
= 0.125 error on the spectroscopic factor of the (d + 6 Li)
clustering). The latter value for the astrophysically relevant 1 (7.12 MeV) thus corresponds to
a range of strictly DWBA inferred values of:
0.043  2 (7.12)  0.067. (18)
However, to ensure the independence of the 2 (7.12) parameter value on the DWBA model
calculation and obtain a normalized DWBA value, one must use in Eq. (16), 2 (6.92) val-
ues from other adequate sources. E.g., assuming (10.35) = c.m. = 26 keV [20] for the 4+
A. Belhout et al. / Nuclear Physics A 793 (2007) 178211 201

(10.35 MeV) state of 16 O and using Eq. (14), one obtains 2 (10.35) = 0.27 (a = 5.5 fm). Then,
from the above value of R3 (Eq. (17)), one gets the estimate 2 (7.12) = 0.14 0.04, i.e., the
range of values 0.10  2 (7.12)  0.18.
Besides, in the R-matrix analysis of their 12 C(, )12 C elastic scattering data, Tischhauser et
al. [32] extracted the (formal) value 12 = 0.47 0.06 MeV1/2 of the reduced width amplitude
for the 2+ (6.92 MeV) state at the same typical channel radius a = 5.5 fm, from which we
deduce the R-matrix formal value 2 (6.92) = 0.32 0.08, i.e., an observed value of 2 (6.92) =
0.310 0.073. Adopted in Eq. (16), the latter value leads to:
2 (7.12) = 0.093 0.050, (19a)
i.e., to the range of 2 (7.12) values:
0.043  2 (7.12)  0.143 (19b)
(a = 5.5 fm), including the preceding one and corresponding to a range of the reduced width
amplitude values:
0.18  (7.12)  0.32, (20)
in MeV1/2 units.
Noting that the direct single-step -transfer reaction mechanism is more favoured relative to
the compound nucleus formation in the (7 Li, t) reaction than in (6 Li, d) and in both two reactions
as the bombarding energy increases [2,3,7,8], we compare our final 2 (7.12) result (Eqs. (19))
only to previous ones that have been derived from the studies of these reactions at incident ener-
gies E6,7 Li  30 MeV at which the compound nucleus component is expected to be minimized.
Since the 1 (7.12 MeV) and (9.6 MeV) states are largely separated in excitation energy, the R1
ratio may be very DWBA model dependent. Its value (R1 = 0.57 0.13) derived here can be
considered as not reliable, and this ratio then needs be redetermined. Using in Eq. (14) the level
width value, c.m. = (c.m.) = (338 60) keV, for the 1 (9.6 MeV) state determined in this
work, one gets the value:
2 (9.6) = 0.45 0.08.
Note that while Becchetti et al. [8] similarly determined a comparable value 2 (9.6)  0.6 from
the level -width of this state in the 12 C(7 Li, t)16 O reaction, our value amounts to one half
earlier accepted values (2 (9.6) = 0.85 from [8,10,11], 2 (9.6) = 0.794 in [38]). Introducing
this value in Eq. (15) with the above determination of 2 (7.12), one obtains R1 = 0.21 0.15,
i.e.:
0.06  R1  0.36. (21)
Our results for 2 (7.12) and R1 (Eqs. (19) and (21)) can, first, be compared to previous ones
derived, for the channel radius a = 5.4 fm, via DWBA analyses of (7 Li, t) and (6 Li, d) transfer
reactions data sets, which are the following:
(a) in their study of the (6 Li, d) reaction, Becchetti et al. [2,3] obtained: 0.1  2 (7.12)  0.4
and R1  0.4, determined from the measured (c.m.) width of the 9.6 MeV state at E6 Li =
42.1 MeV [2], 0.05  2 (7.12)  0.3 and 0.3 < R1 < 0.6 at E6 Li = 90.2 MeV [3];
(b) while their studies [8,9] of the 12 C(7 Li, t)16 O reaction led to: 0.1  2 (7.12)  0.18
(weighted mean value) and 0.22 < R1 < 0.48 at E7 Li = 34 MeV [8], 0.28 < R1 < 0.42 and
2 (7.12 ) not reported at E7 Li = 70 and 101 MeV [9].
202 A. Belhout et al. / Nuclear Physics A 793 (2007) 178211

Table 5
Reduced -width values for the 16 O states relative to that of the 2+ (6.92 MeV) state (R2 ratio) from the present DWBA
analysis and for the channel radii a = 6.5 and 7.7 fm
J (Ex , MeV) (N, L) 2 /2 (2+ )a
a = 6.5 fm a = 7.7 fm
0+ (0.00) (2, 0) 0.098 0.049 0.025 0.012
0+ (6.05) (4, 0) 0.350 0.173 0.332 0.166
3 (6.13) (1, 3) 0.261 0.166 0.141 0.091
2+ (6.92) (3, 2) 1.00 1.00
1 (7.12) 0.230 0.092b 0.2530.101b
1 (9.6) (4, 1) 0.611 0.257 2.28 0.95
4+ (10.35) (2, 4) 0.069 0.025 0.048 0.017
a 2 reduced widths relative to 2 (2+ ) = 0.20 0.06 and 0.071 0.021 for the channel radii 6.5 and 7.7 fm,

respectively. b Value obtained for mixing configuration (see text).

Fig. 6. Variation of the R2 = 2 (7.12)/2 (6.92) ratio in function of the channel radius a. The solid curve represents the
DWBA calculation with corresponding lower and upper limits depicted by dotted curves. R2 values deduced from other
works are also reported (a [31], " [32], 2 [76], Q [4]).

Thus, while an overlapping between our results and previous ones is observed, our values
for the 2 (7.12) parameter show to be rather included within the range of values extracted by
Becchetti et al. [3] from their study of the (6 Li, d) reaction at E6 Li = 90.2 MeV.

4.3.2.2. Calculation for a = 6.5 fm The 2 values of the studied low lying -cluster states of
16 O, scaled to that of the 2+ (6.92 MeV) state, derived for the channel radii a = 6.5 fm and
7.7 fm, are listed in Table 5. Notice that no similar results from DWBA analyses are available in
the previous literature, to our knowledge.
The R2 ratio calculated using Eq. (12) for the DWBA , 2 (a) reduced widths is much more

independent on the interaction radius than the reduced -widths themselves. This is clearly il-
lustrated in Fig. 6 showing a smooth variation of this ratio beyond the channel radius value
a = 5.5 fm, this variation not exceeding 8.7% between a = 6.5 fm and 7.7 fm. Notice that for
the same channel radius a = 6.5 fm, the R2 value (0.12 0.06) reported in Ref. [76] and ob-
tained by R-matrix calculation is inconsistent with that (R2 = 0.42 0.15) derived by Brune et
A. Belhout et al. / Nuclear Physics A 793 (2007) 178211 203

al. [4] from a DWBA analysis of the (6 Li, d) reaction cross section data at sub-Coulomb energies
involving a determination of ANCs (see Fig. 6). Our own value, R2 = 0.23 0.094, calculated
using our strict DWBA reduced -widths {2 (7.12) = 0.047 0.011, 2 (6.92) = 0.2 0.085}
lies between the latter two values and partially overlaps with them. As has been done above for
the channel radius a = 5.5 fm, this R2 value must be multiplied by a DWBA model independent
2 (6.92) value in order to deduce an also least model independent 2 (7.12) value. For this pur-
pose, we have used the value 2 (6.92) = 0.04 0.0072 of Brune et al. [4], which leads to the
following result:
2 (7.12) = 0.0093 0.0059 (22)
herein denoted as normalized DWBA value.
Alternatively, the inverse of the R2 ratio may be multiplied by an adequately chosen non-
DWBA 2 (7.12) value in view to infer a least DWBA model dependent 2 (6.92) value. In
this case, using the 2 (7.12) = 0.0125+0.004
0.003 value derived in our own R-matrix analysis of the
12 C(, )16 O data (see next section), we obtain the normalized DWBA value:

2 (6.92) = 0.054+0.039
0.035 , (23)
which is in agreement within the involved uncertainties with the above quoted value of Brune et
al. [4] and with the value 2 (6.92) = 0.091 0.020 of Fey et al. [76]. These determinations of
2 (6.92) and 2 (7.12) correspond to the following ANC values:
 
C 2 2+ = 1.96+1.41
1.27 10 ,
10
(24a)

2

C 1 = 3.48 2.0 10 , 28
(24b)
in fm1 units, which are consistent with the (already cited) previous determinations from
Refs. [4,32,68]. It is worthwhile noting the consistently larger error on our C 2 (2+ ) result in
comparison to those affecting the previous experimental values from Refs. [4,32], in connection
with the still large indeterminacy of the 12 C(, )16 O SE2 (E) factor at the stellar helium burning
energy.

4.3.2.3. Calculation for a = 7.7 fm Our 2 values derived for this channel radius from Eq. (12)
through Eqs. (11), (13) are also reported in Table 5. For the 1 (7.12 MeV) astrophysically rele-
vant state of 16 O, the obtained value in this case, R2 = 0.25 0.1, corresponds to strict DWBA
values of 2 (6.92) = 0.07 0.03 and 2 (7.12) = 0.018 0.004, which are, indeed, very close to
those calculated above (without errors) for the same radius using Eqs. (10), (11). As can be seen
in Fig. 5, the reduced -width decreases as the interaction radius increases over the asymptotic
region, and the value a = 7.7 fm seems to be a higher limit value of the channel radius. The
inverse ratio R21 combined, as above, with our R-matrix value, 2 (7.12) = 0.0011+0.0005
0.0008 (see
+0.0063
next section), leads to a normalized DWBA value, (6.92) = 0.00440.0035 , which therefore
2

may be considered as a lower reduced -width limit.


The comparison of our DWBA results with those derived via R-matrix analyses of data sets
for the 12 C(, )16 O helium burning reaction is made in the following section.

5. R-matrix analysis of data sets for the 12 C(, )16 O helium burning

To derive an estimate of the E1 part of the 12 C(, )16 O reaction total astrophysical factor
value at the Gamow peak energy and constrain the preceding DWBA 2 (7.12) values Eqs. (19),
204 A. Belhout et al. / Nuclear Physics A 793 (2007) 178211

Table 6
Best fit R-matrix parameter values and SE1 (0.3) results for the channel radii a = 5.5, 6.5 and 7.7 fm
a (fm) 5.5 6.5 7.7
+0.012
l (7.12) (MeV1/2 ) 0.177 0.03 0.07870.010 0.0198 0.004
+0.004
2 (7.12) 0.045 0.015 0.01260.003 (1.1 0.45) 103
SE1 (0.3) (keV b) 85+21
19
+17
80.616 59.8a
2 6.92 7.42 8.42
a Error due to the variation of 2 (7.12) 10 keV b.

(20), we have performed in this work a three-level R-matrix analysis of the most recent [23,
28] and earlier [21,24,26,27] data sets for the 12 C(, ) direct capture SE1 (E) factor, the 1,3
elastic -scattering phase shift [31] and the -spectrum following 16 N -delayed decay [36].
The SE1 (E) excitation function, 1 phase shift and the p-wave contribution in the -spectrum
were calculated using the formalism given in Refs. [39,72] assuming the contributions of the
two 1 (7.12 MeV) and (9.6 MeV) states of 16 O essentially responsible for the E1 capture and
of an additional broad, highly excited 1 background level. Similarly, the 3 phase shift and
the f -wave part of the -spectrum were also determined by a three-level R-matrix calculation
considering the 3 (6.13 MeV) and (11.6 MeV) states of 16 O, and an additional 3 background
level. In the calculations, the (arbitrary) boundary condition parameter was taken equal to the
shift factor at the subthreshold resonance energies, i.e., Bc = Sc (ER ). The observed reduced
-width amplitudes, l , and the R-matrix formal ones, l2 , are related by:
 1
l2 = l2 1 + l2 dS/dE E . (25)
R
This relation holds only for states for which the preceding requirement on Bc is satisfied,
i.e., only for the subthreshold states of 16 O in our case, as already stated above.
2 quality
To appreciate the agreement of the theoretical fits to experimental data, a reduced red
function was defined similarly as in Eq. (6).
As has been done above in the case of our DWBA calculations, the present R-matrix analyses
have been carried out for the three channel radii a = 5.5 fm, 6.5 fm and 7.7 fm. Theoretical
fits to the different experimental data sets of quite good quality have been obtained in the three
cases. The results corresponding to the best-fit curves to the measured data for the channel radius
a = 6.5 fm are reported in Fig. 7. The results concerning the reduced -width amplitudes for the
1 (7.12 MeV) state and extrapolated SE1 (0.3) values for both three channel radii are given in
Table 6, together with the corresponding best fit red2 values. Considering the strong dependence

of the reduced -widths on the channel radius, the choice of the latter should be subjected to
other restrictions in addition to that based on the 2 function optimization. As discussed above,
an extended range of the strong interaction can be involved in the studied (6 Li, d) transfer reac-
tion depending on the reaction kinematics conditions. This makes it seemingly difficult to come
to a definite conclusion concerning the exact channel radius to be adopted for performing the
most reliable nuclear level parameter determination both in the DWBA and the R-matrix analy-
ses of experimental data. However, the channel radius a = 6.5 fm (for which valuable previous
results are available and can be used here for comparison) appears to be more realistic than the
5.5 fm and 7.7 fm ones. Therefore, mainly the results corresponding to this channel radius are
commented below.
Our R-matrix (formal) reduced -width results are compared in Table 7 to the current DWBA
calculated ones and to those reported in the literature. The R-matrix 2 (7.12) value from this
A. Belhout et al. / Nuclear Physics A 793 (2007) 178211 205

Table 7
Formal reduced -width values determined in this work compared to those from the literature for the channel radius
a = 6.5 fm
J (Ex , MeV) 2
This work Literature
R-matrix Strict DWBA Normalized DWBA R-matrix DWBA + ANC
3 (6.13) 0.012 0.17 0.09 0.013a
+0.13 +0.041 +0.013 b
2+ (6.92) 0.220.09 0.0550.036 0.0920.019 0.043 0.008d
+0.004
1 (7.12) 0.01260.003 0.047 0.01 0.0094 0.0058 0.013 ; 0.011 0.003 0.017 0.003d
a,c b

1 (9.6) 0.22 0.22a ; 0.21d


a Ref. [36]. b Ref. [76]. c Ref. [37]. d Ref. [4].

work is in very good agreement with those inferred in other works via R-matrix analyses [36,37,
76] whereas the Brune et al. [4] value, deduced from their ANC calculation, is slightly higher
than the other values. Likely due to the model dependence of our strict DWBA 2 (7.12) and
2 (6.92) values, they are substantially different (much higher) than the other values contrarily
to our normalized DWBA values which are remarkably consistent, within the associated un-
certainties, with the previous results derived both by R-matrix analyses [36,37,76] of data sets
for the 12 C(, ) capture reaction and by DWBA analysis of -transfer reaction data at sub-
Coulomb energies [4]. Besides, it is remarkable that our R-matrix 2 (7.12) result is compatible
with early astrophysics calculations [77,78] predicting reduced -width values  0.1 for the 1
(7.12 MeV) state.
Our best-fit R-matrix 2 values for the 1 (9.6 MeV) and 3 (6.13 MeV) states are also given
in Table 7 (for a = 6.5 fm). As can be seen, they are very consistent with the values reported
in Refs. [4,36], derived by R-matrix analyses for the same channel radius. Our strict DWBA
inferred 2 (6.13 MeV) value is much larger, however, than the R-matrix ones, even after nor-
malization, a behavior also shown by the DWBA values reported in Refs. [2,3]. As discussed
in the preceding section, this may be ascribed to the complex interplay of different interaction
mechanisms making the DWBA calculation unable to fully reproduce the measured -transfer
reaction cross section data for this state (as for other states subject to interaction mechanism over-
laps), while the R-matrix calculations only deal with compound resonance reactions. Besides, in
contrast, Descouvemont et al. [40] derived a much lower value (absolute 2 = 3.3 103 ) from
a microscopic generator coordinate method (GCM) calculation, attributing this too low value (as
that also very small found for the 0+ g.s.) to the smallness of the 3 (6.13 MeV) state rms ra-
dius value used in the calculation (see Ref. [40]). Finally, one observes that our reduced -width
results for both the 3 (6.13 MeV), 1 (7.12 MeV) and 1 (9.6 MeV) states of 16 O derived via
R-matrix analysis (i.e., 2 = 0.012, 0.0126+0.004
0.003 and 0.22, respectively) are essentially identi-
cal to those obtained in the previous analysis of Azuma et al. [36] for the same channel radius
a = 6.5 fm (i.e., 2 = 0.013, 0.013 and 0.22, respectively, see Table 7); the small differences seen
are likely due only to the fact that more data for the 12 C(, ) capture reaction are considered in
the current analysis. Whence again, our normalized DWBA value for the 1 (7.12 MeV) state
{2 (7.12) = 0.0094 0.0058} is quite consistent with the values from both R-matrix analyses.
Our SE1 (0.3) = 80.6+17
16 keV b estimate, derived by R-matrix analysis for the channel radius
a = 6.5 fm, is very consistent with previously derived R-matrix values reported in the literature
for the same channel radius (see Table 6). These values are, e.g., SE1 (0.3) = 79 16 keV b [19],
7620 keV b [28], 7921 keV b [36], and the value 7717 keV b recently reported in Refs. [23,
206 A. Belhout et al. / Nuclear Physics A 793 (2007) 178211

76]. Besides, our value, SE1 (0.3) = 85+21 19 keV b, derived for the channel radius a = 5.5 fm
is consistent with that (80 21 keV b) obtained previously for this radius from an analysis of
elastic -scattering data [32]. Finally, our result for a = 7.7 fm, SE1 (0.3) = 59.8 keV b, is rather
consistent with those for the other two channel radii, all values overlapping within the involved
uncertainties. Indeed, as can be seen in Table 6, the astrophysical factor is actually much less
sensitive to the variation of the channel radius than the reduced -widths.

6. Summary and conclusions

The high energy resolution detection system used in the current 12 C(6 Li, d)16 O experiment
at E6 Li = 48.2 MeV allowed us to achieve a clear separation in the deuteron spectra of the
peaks for the states of 16 O lying at excitation energies Ex  11 MeV. Values of Ex and the
level width, c.m. , for states up to 13.2 MeV excitation are extracted and compared to those
reported in the literature [20]. In particular, we find a value c.m. = 338 60 keV for the line
width of the astrophysically relevant 1 (9.6 MeV) unbound state, very consistent with that
(338 30 keV) derived by Becchetti et al. in their study of the (6 Li, d) -transfer reaction at
E6 Li = 90.2 MeV [3].
The measured angular distribution data exhibit a sharp forward peaking for the 0+ g.s. and
the 0+ (6.05 MeV), 3 (6.13 MeV), 2+ (6.92 MeV), 1 (7.12 MeV), 1 (9.6 MeV) and 4+
(10.35 MeV) states presumed to have an -cluster character. In contrast, no noticeable such
trend is observed for the 2 (8.87 MeV) and 2+ (9.85 MeV) states and the 3+ + 4+ doublet at
Ex 11.09 MeV assumed to be of non--cluster structure. Assuming the earlier DWBA + HF
analyses of the (6 Li, d), (7 Li, t) -transfer cross section data to likely overestimate the compound
nucleus component normalized to the 2 (8.87 MeV) state [2,7,8], we have analyzed our (6 Li, d)
angular distribution data for the first 11 states of 16 O by the FRDWBA theory used within a
simulation method aiming at globally accounting for the complex interplay of several interaction
mechanisms visibly manifested by the measured data. A differential cross section term, aver-
aged over backward angles (c.m.  60 ) where single-step -transfers can be neglected, and
assumed to describe all reaction processes (direct single-step -transfers, multi-step -transfers,
compound nucleus formation, . . . ), has been incoherently added to the FRDWBA component to
simulate the theoretical total differential cross section fitted to the measured cross section data
mainly at forward angles. Concerning the -cluster states, fits to experimental data of quite good
quality for the 0+ g.s., 0+ (6.05 MeV, best fitted), 1 (7.12 MeV) and 1 (9.6 MeV) states have
been obtained while the fits to lowest angle data for the 3 (6.13 MeV), 2+ (6.92 MeV) and 4+
(10.35 MeV) states are of less good quality.
Besides, the averaged cross section term simulating the contribution of all involved interaction
mechanisms at backward angles seems to globally well reproduce the data for the 2+ (9.85 MeV)
state and the 3+ + 4+ doublet at 11.09 MeV but not the data for the 2 (8.87 MeV) state at
(c.m.)  50 which exhibit a significant increase of the reaction cross section. Furthermore,
the compound nucleus components for these non--cluster states, more suitably evaluated here
by HF calculation relative to the 0+ g.s., turn out to represent only a small fraction of the total
differential cross section in contrast with previous evaluations [1,2,7,8]. According to the recent
analysing power measurement and CRC calculations of Ref. [6], multi-step -transfer transitions
seem to be important to the 0+ g.s., 3 (6.13 MeV) and 2 (8.87 MeV) states of 16 O but weaker
to the 2+ (6.92 MeV) and 4+ (10.35 MeV) states, suggesting that the latter two states could be
fully populated via direct single-step -transfers. Thus, our above results seem to confirm these
statements, except for the 0+ g.s., quite acceptably described here by our FRDWBA calculation.
A. Belhout et al. / Nuclear Physics A 793 (2007) 178211 207

S spectroscopic factors and 2 reduced widths scaled to those of the 2+ (6.92 MeV) state
have been extracted for the studied -cluster states of 16 O, and seem to us as being more reliable
than previous ones derived by FRDWBA + HF analyses of -transfer reaction data [1,2,7,8]
where the calculated HF cross sections have been overestimated.
The extracted relative S values are found to be different but of the same order of magnitude
as those from Refs. [2,3,8] while our absolute S value for the 0+ g.s. (S = 0.34 0.07) is
found to be very consistent with that (0.32) obtained by Drummer et al. [5] for the same value
of the + 12 C bound state potential radius, a = 4.85 fm. Besides, except for the 0+ (6.05 MeV)
state, our values show to be in fair agreement within the involved uncertainties with SU(3) model
calculations [54], contrarily to those of Becchetti et al. [2,3,8]. The case of the 1 (7.12 MeV)
astrophysically relevant state of 16 O has been treated in a particular manner assuming a con-
figuration mixing for its cross section in proportions of 80% and 20% to the (N, L) = (2, 1)
and (4, 1) assignments, respectively. This leads to an improved best theoretical fit to the cor-

Fig. 7. Experimental data for (a) the 12 C(, )16 O reaction astrophysical SE1 (E) factor from [21,23,24,2628], (b) the
-delayed -spectrum of 16 N from [36] and (c) the 1,3 elastic -scattering phase shifts from [31]. The solid curves
represent the best R-matrix fits for the channel radius a = 6.5 fm while the dashed curves in (c) correspond to the p- and
f -wave contributions to the -spectrum.
208 A. Belhout et al. / Nuclear Physics A 793 (2007) 178211

Fig. 7. (continued)

responding angular distribution experimental data for a relative value 0 of the -spectroscopic
factor S /S (2+ ) = 0.3 0.09, which is fortuitously the same as the relative reduced -width,
2 /2 (2+ ), for this state.
The variation of the reduced -widths in function of the channel radius for the states of 16 O
of astrophysical interest has been studied, showing the following features:
(i) the nuclear interaction potential range extends up to a = 7.7 fm, the potential amplitude
reaching 10% of its maximum value at a = 6.5 fm (a = 7.2 fm in Ref. [5]),
(ii) in the asymptotic region starting at a 6.5 fm, the R2 = 2 (7.12)/2 (6.92) ratio presents
a smooth variation versus the channel radius a. As already stated, this ratio minimizes the influ-
ence of the uncertainties inherent to the DWBA model, and was then used to extract normalized
DWBA reduced -widths, in the absence of a better alternative way,
(iii) while for the DWBA calculations the channel radii a = 5.5 fm and 7.7 fm stand for ex-
treme limit ones (relative to the R2 ratio variation), the channel radius a = 6.5 fm appears to be
more realistic for use both in DWBA and R-matrix calculations. Noting that previous DWBA
results are available only for channel radii close to a = 5.5 fm, such calculations have been car-
ried out for both three preceding channel radii, although the nuclear potential used here extends
beyond the interaction radius a = 5.5 fm.
Concerning our reduced -width results for the channel radius a = 5.5 fm, an overlapping
with the previous ones is observed, our 2 (7.12) values being rather contained in the range of
values extracted by Becchetti et al. [3] in their study of the (6 Li, d) reaction at E6 Li = 90.2 MeV
[3] involving only a DWBA analysis of experimental data without consideration of a compound
nucleus contribution to the reaction cross section. Calculated (formal) normalized DWBA val-
ues of 0.044  2 (7.12)  0.11 have been obtained for this channel radius. Besides, the value
0.03  2 (7.12)  0.06 has been extracted in the current R-matrix analysis of various data sets
(SE1 (E) factor, 1,3 phase shifts and -delayed -spectrum of 16 N) for the 12 C(, )16 O he-
lium burning reaction. A significant overlapping is thus observed between the DWBA values and
R-matrix ones.
A. Belhout et al. / Nuclear Physics A 793 (2007) 178211 209

For the other limit channel radius a = 7.7 fm, a normalized DWBA value of 2 (6.92) =
0.0044+0.0063
0.0035 have been extracted using the value, (7.12) = 0.0011 0.00045, deduced from
2
12 16
our R-matrix analysis of the C(, ) O capture reaction.
As for the likely more adequate channel radius a = 6.5 fm, normalized DWBA values of
0.0036  2 (7.12)  0.152 and 0.019  2 (6.92)  0.096 have been obtained. They correspond
to ANC values of C 2 (1 ) = 3.48 2.2 1028 fm1 and C 2 (2+ ) = 1.96+1.41 1.27 10
10 fm1 ,

respectively, which are consistent both with previous experimental determinations [4,32] and
theoretical model calculations [41,68] (for C 2 (2+ )).
Besides, our R-matrix analysis of the helium burning reaction for this radius yields the values,
0.0096  2 (7.12)  0.0166, which are almost contained within the range of our normalized
DWBA ones. They correspond to an extrapolated value of the E1 component of the reaction
astrophysical factor at the Gamow peak energy, SE1 (0.3) = 80.6+17 16 keV b. This result, thus af-
fected by a relative uncertainty
SE1 /SE1 of 20%, is very consistent with previous estimates
reported in the literature [19,23,28,32,36].
Hence, while in previous DWBA analyses, overestimated values (2 (7.12)  0.1) much
higher than those (2 (7.12)  0.1) derived via the main previous R-matrix analyses were ob-
tained, we find here DWBA values in better consistency with the latter results.
However, the DWBA method can be surely trustable if only single-step -transfers are in-
volved in -transfer reaction experiments, which seems quite impossible to achieve in practice.
Also, the privileged way to experimentally derive spectroscopic and astrophysical informations
on the 12 C(, )16 O helium burning reaction ultimately resides in performing the most possible
accurate direct capture measurements, currently envisaged in two directions via: (i) -nucleus
coincidences for separating the different interplaying contributions to the tiny reaction cross sec-
tion using the available recoil nucleus separators and associated high efficiency -ray detectors
[22,79,80] at high -particle energies (E  1.95.4 MeV) and (ii) similar future experiments
at energies much lower than the presently attained ones (E  0.89 MeV [23,76]). The existing
most powerful theoretical tool used to consistently extrapolate the measured high energy data
to the Gamow peak energy of 0.3 MeV probably consists in R-matrix calculations. The latter
also could be reliably handled only if the interaction radius is precisely identified. This is not
usually achieved in phenomenological R-matrix analyses of experimental data for the 12 C(, )
capture where the reduced -widths are deduced via fit procedures constraining the choice of the
channel radius only through the 2 function minimization, without paying the necessary care to
the effective nuclear interaction range and effective nuclear interaction surface.
Notice, in contrast, that in the DWBA analyses the interaction potential is primarily defined
and is then used to calculate the + 12 C system wave functions for determining the nuclear
parameters at a channel radius assumed to correspond to the nuclear interaction surface. But this
has not generally been done in previous works within the asymptotic region, to our knowledge.
Only at this condition (i.e., provided that the correct interaction radius is adopted), the DWBA
calculations should provide reduced -widths corresponding to their definition in the R-matrix
theory. To be reliable, the R-matrix analyses also should be similarly associated to the detailed
variations of the nuclear interaction potential and reduced -widths to precisely determine the
most appropriate interaction radius to be adopted in the calculations.

References

[1] A. Cunsolo, A. Foti, G. Pappalardo, G. Raciti, N. Saunier, Phys. Rev. C 18 (1978) 856.
[2] F.D. Becchetti, J. Jnecke, C.E. Thorn, Nucl. Phys. A 305 (1978) 313.
210 A. Belhout et al. / Nuclear Physics A 793 (2007) 178211

[3] F.D. Becchetti, D. Overway, J. Jnecke, W.W. Jacobs, Nucl. Phys. A 344 (1980) 336.
[4] C.R. Brune, W.H. Geist, R.W. Kavanagh, K.D. Veal, Phys. Rev. Lett. 83 (1999) 4025.
[5] T.L. Drummer, E.E. Bartosz, P.D. Cathers, M. Fauerbach, K.W. Kemper, E.G. Myers, K. Rusek, Phys. Rev. C 59
(1999) 2574.
[6] N. Keeley, et al., Phys. Rev. C 67 (2003) 044604.
[7] M.E. Cobern, D.J. Pisano, P.D. Parker, Phys. Rev. C 14 (1976) 491.
[8] F.D. Becchetti, E.R. Flynn, D.L. Hanson, J.W. Sunier, Nucl. Phys. A 305 (1978) 293.
[9] F.D. Becchetti, M.L. Dowell, P.M. Lister, J.W. Jnecke, A. Nadasen, J.S. Winfield, J. Phys. Soc. Jpn. 58 (1989) 635.
[10] F. Puhlhofer, et al., Nucl. Phys. A 147 (1970) 258272;
R. Weibezahn, H. Freiesleben, F. Puhlhofer, R. Bock, Nucl. Phys. A 305 (1971) 645656;
J.L. Honsaker, T.H. Hsu, W.J. McDonald, G.C. Neilson, Nucl. Phys. A 144 (1970) 473480;
J. Lowe, A.R. Barnett, Nucl. Phys. A 187 (1972) 323336.
[11] G.R. Caughlan, et al., At. Data Nucl. Data Tables 32 (1985) 197;
W.A. Fowler, G.R. Caughlan, B.A. Zimmerman, Annu. Rev. Astron. Astrophys. 5 (1967) 525;
W.A. Fowler, G.R. Caughlan, B.A. Zimmerman, Annu. Rev. Astron. Astrophys. 13 (1975) 69.
[12] G.R. Caughlan, W.A. Fowler, At. Data Nucl. Data Tables 40 (1988) 283.
[13] T.A. Weaver, S.E. Woosley, Phys. Rep. 227 (1993) 65.
[14] L. Buckmann, Astrophys. J. 468 (1996) L127;
L. Buckmann, Astrophys. J. 479 (1997) L153.
[15] R. Kunz, M. Jaeger, A. Mayer, J.W. Hammer, G. Staudt, S. Harissopulos, T. Paradellis, Astrophys. J. 567 (2002)
643.
[16] T. Rauscher, A. Heger, R.D. Hoffman, S.E. Woosley, Astrophys. J. 576 (2002) 323;
A. Heger, S.E. Woosley, Astrophys. J. 567 (2002) 532.
[17] S.E. Woosley, A. Heger, T. Rauscher, R.D. Hoffman, Nucl. Phys. A 718 (2003) 3c;
www.supersci.org.
[18] C.E. Rolfs, W.S. Rodney, Cauldrons in the Cosmos, University of Chicago Press, 1988.
[19] J.M.L. Ouellet, M.N. Betler, H.C. Evans, H.W. Lee, J.R. Leslie, J.D. MacArthur, W. McLachie, H.-B. Mak, P.
Skensved, J.L. Witton, X. Zahao, T.K. Alexander, Phys. Rev. C 54 (1996) 1982.
[20] D.R. Tilley, H.R. Weller, C.H. Cheves, Nucl. Phys. A 564 (1993) 1, Revised Manuscript, September 2002.
[21] A. Redder, H.W. Barker, C. Rolfs, H.P. Trautvetter, T.R. Donoghue, T.C. Rinckel, J.W. Hammer, K. Langanke,
Nucl. Phys. A 462 (1987) 385.
[22] C. Mattei, et al., Phys. Rev. Lett. 97 (2006) 242503.
[23] J.W. Hammer, et al., Nucl. Phys. A 752 (2005) 514c. See also Ref. [76] below.
[24] P. Dyer, C.A. Barnes, Nucl. Phys. A 223 (1974) 495.
[25] R.M. Kremer, et al., Phys. Rev. Lett. 60 (1988) 1475.
[26] J.M. Ouellet, et al., Phys. Rev. Lett. 69 (1992) 1896.
[27] G. Roters, et al., Eur. Phys. J. A 6 (1999) 451.
[28] R. Kunz, et al., Phys. Rev. Lett. 86 (2001) 3244;
R.W. Kunz, Ph.D. Thesis, Stuttgart University, Germany, 2002.
[29] M. Assano, et al., Phys. Rev. C 73 (2006) 055801.
[30] M. DAgostino Bruno, et al., Nuovo Cimento A 27 (1975) 1.
[31] R. Plaga, et al., Nucl. Phys. A 465 (1987) 291.
[32] P. Tischhauser, et al., Phys. Rev. Lett. 88 (2002) 072501.
[33] J. Kiener, V. Tatischeff, P. Auger, G. Bogaert, A. Coc, D. Disdier, L. Kraus, A. Lefebvre, I. Link, W. Mittig, T.
Motobayashi, F. de Oliveira-Santos, P. Roussel-Chomaz, C. Stephan, J.P. Thibaud, Nucl. Phys. A 621 (1997) 173.
[34] Z. Zhao, R.H. France, K.S. Lai, S.L. Rugari, M. Gai, E.L. Wilds, Phys. Rev. Lett. 70 (1993) 2066.
[35] L. Buchmann, R.E. Azuma, C.A. Barnes, J.M. DAuria, M. Dombsky, U. Giesen, K.P. Jackson, J.D. King, R.G.
Korteling, P. McNeely, J. Powell, G. Roy, J. Vincent, T.R. Wang, S.S.M. Wong, P.R. Wrean, Phys. Rev. Lett. 70
(1993) 726.
[36] R.E. Azuma, L. Buchmann, F.C. Barker, C.A. Barnes, J.M. DAuria, M. Dombsky, U. Giesen, K.P. Jackson, J.D.
King, R.G. Korteling, P. McNeely, J. Powell, G. Roy, J. Vincent, T.R. Wang, S.S.M. Wong, P.R. Wrean, Phys. Rev.
C 50 (1994) 1194.
[37] L. Buchmann, R.E. Azuma, C.A. Barnes, J. Humblet, K. Langanke, Phys. Rev. C 54 (1996) 393.
[38] F.C. Barker, T. Kajino, Aust. J. Phys. 44 (1991) 369.
[39] F.C. Barker, Aust. J. Phys. 24 (1971) 777;
F.C. Barker, Aust. J. Phys. 25 (1972) 341.
A. Belhout et al. / Nuclear Physics A 793 (2007) 178211 211

[40] P. Descouvemont, D. Baye, Nucl. Phys. A 430 (1984) 426.


[41] P. Descouvemont, D. Baye, Phys. Rev. C 36 (1987) 1249;
P. Descouvomont, Nucl. Phys. A 270 (1987) 309;
P. Descouvomont, Phys. Rev. C 47 (1993) 210.
[42] T.S. Metcalfe, M. Salaris, D.E. Winget, Astrophys. J. 573 (2002) 803.
[43] H.A. Enge, Nucl. Instrum. Methods 28 (1964) 119;
H.A. Enge, Nucl. Instrum. Methods 28 (1964) 119;
J. Spencer, H.A. Enge, Nucl. Instrum. Methods 49 (1967) 181.
[44] G. Rotbard, G. Berrier-Ronsin, O. Constantinescu, S. Fortier, S. Gales, M. Hussonnois, J.B. Kim, J.M. Maison, L.H.
Rosier, J. Vernotte, C. Briancon, R. Kulessa, Y.T. Oganessian, S.A. Karamian, Phys. Rev. C 48 (1993) R2148.
[45] G.R. Satchler, Direct Nuclear Reactions, Clarendon, Oxford, 1983, pp. 734736.
[46] N. Keeley, K.W. Kemper, D.T. Khoa, Nucl. Phys. A 726 (2003) 159.
[47] I.J. Thompson, Comput. Phys. Rep. 7 (1988) 167.
[48] F. Hinterberger, G. Mairle, U. Schmidt-Rohr, G.J. Wagner, P. Turek, Nucl. Phys. A 111 (1968) 265.
[49] K.I. Kubo, M. Hirata, Nucl. Phys. A 187 (1972) 186.
[50] D.R. Lehman, M. Rajan, Phys. Rev. C 25 (1982) 2743.
[51] G. Brown, J.G.B. Haigh, D.L. Watson, Nucl. Phys. A 232 (1974) 125.
[52] A.K. Abdullah, T. Udagawa, T. Tamura, Phys. Rev. C 8 (1973) 1855.
[53] W. Hauser, H. Fesbach, Phys. Rev. 87 (1952) 366;
F. Ajzenberg-Selove, Nuclear Spectroscopy, part B, Academic Press, 1960.
[54] M. Ichumura, A. Arima, E.C. Halbert, T. Terasawa, Nucl. Phys. A 204 (1973) 225.
[55] Y. Suzuki, Prog. Theor. Phys. 55 (1976) 1751;
Y. Suzuki, Prog. Theor. Phys. 56 (1976) 111.
[56] A. Cunsolo, A. Foti, G. Pappalardo, G. Raciti, N. Saunier, Phys. Rev. C 18 (1978) 856;
R.E. Warner, N.S. Chant, P.G. Roos, C. Samanta, S. Kakigi, N. Koori, M. Fujiwara, N. Matsuoka, K. Tamura,
E. Kubo, K. Ushiro, Nucl. Phys. A 641 (1998) 3.
[57] V.I. Kukulin, V.M. Krasnopolsky, V.T. Voronchev, P.B. Sazonov, Nucl. Phys. A 417 (1984) 128.
[58] R. Beck, F. Dickmann, R.G. Lovas, Nucl. Phys. A 446 (1985) 703.
[59] R.G. Lovas, A.T. Kruppa, R. Beck, F. Dickmann, Nucl. Phys. A 474 (1987) 451.
[60] K. Varga, R.G. Lovas, Phys. Rev. C 43 (1991) 1201.
[61] A. Cst, R.G. Lovas, Phys. Rev. C 46 (1992) 576.
[62] H. Walliser, Y.C. Tang, Phys. Lett. B 135 (1984) 344.
[63] A.M. Mukhamedzhanov, N.K. Timofeyuk, JETP Lett. 51 (1990) 282.
[64] H.M. Xu, et al., Phys. Rev. Lett. 73 (1994) 2027.
[65] A.M. Mukhamedzhanov, C.A. Gagliardi, R.E. Tribble, Phys. Rev. C 63 (2001) 024612;
A.M. Mukhamedzhanov, R.E. Tribble, Phys. Rev. Lett. 59 (1999) 3418.
[66] D. Beaumel, et al., Phys. Lett. B 514 (2001) 226.
[67] P.F. Bertone, et al., Phys. Rev. C 66 (2002) 055804.
[68] J.-M. Sparenberg, Phys. Rev. C 69 (2004) 034601.
[69] C. Angulo, P. Descouvemont, Phys. Rev. C 61 (2000) 064611(R).
[70] L. Buchmann, Phys. Rev. C 44 (2001) 022801(R).
[71] F.L. Milder, J. Janecke, F.D. Becchetti, Nucl. Phys. A 276 (1977) 7292.
[72] A.M. Lane, R.G. Thomas, Rev. Mod. Phys. 30 (1958) 257.
[73] J.M. Blatt, V.F. Weisskopf, Theoretical Nuclear Physics, John Wiley and Sons, New York, 1952, p. 420;
A.G. Sitenko, Theory of Nuclear Reactions, World Scientific, Singapore, 1990, p. 137.
[74] H. Baba, Nucl. Phys. A 159 (1970) 625.
[75] S.E. Koonin, T.A. Tombrello, G. Fox, Nucl. Phys. A 220 (1974) 221232.
[76] M. Fey, Nucl. Phys. A 718 (2003) 131c;
M. Fey, Ph.D. Thesis, Stuttgart University, Germany, 2004, http://elib.uni-Stuttgart.de/opus/volltexte/2004/1683.
[77] W.D. Arnett, Astrophys. J. 176 (1972) 681;
W.D. Arnett, Astrophys. J. 170 (1971) L43.
[78] W.D. Arnett, Annu. Rev. Astron. Astrophys. 11 (1973) 73.
[79] D. Rogalla, et al., Nucl. Instrum. Methods A 513 (2003) 573578.
[80] D. Schrmann, et al., Eur. Phys. J. A 26 (2005) 301.

Anda mungkin juga menyukai