This content was uploaded by our users and we assume good faith they have the permission to share this book. If you own the copyright to this book and it is wrongfully on our website, we offer a simple DMCA procedure to remove your content from our site. Start by pressing the button below!
[GeV/c]
Single Event K/rc ratio
Figure 5. Distibutions of (pt) and the ratio K/ir measured in individual events (data points) compared to expectations from finite number fluctuations (histograms).
size increases more than twofold. For this method of analysis a scheme was used which treats BE-correlations and inclusive particle spectra consistently. However, it overestimates the temperature at large /3j_. Further information on the effective equation of state of the produced matter can be deduced from the study of non-central collisions, where pressure driven flow in the initial almond-shaped interaction region may lead to an anisotropic azimuthal distribution at freezeout. This anisotropy, quantified by the Fourier coefficient Vi of the azimuthal distribution of particles 7 , is plotted in Fig. 4b as a function of the impact parameter b of the collision. Significant in-plane elliptic flow is observed; its order of magnitude is reproduced by the rescattering model RQMD 8 . Hydrodynamical calculations based on a pion gas 9 overpredict the measurements by a factor 5. 2.4
Fluctuations of temperature and flavor content
Fluctuations of average event properties reflect both the equilibration and thermodynamical properties of the produced matter as well as possible longrange correlations which may occur in a medium close to the phase transition. The distributions of the average transverse momentum (p t ) and the ratio of the numbers of charged kaons and pions K/iv per event are displayed in Fig. 5.
104
They closely agree with expectations from finite number statistics as shown by the histograms derived by mixing particles from different events. It appears that central P b + P b collisions produce rather uniform particle ensembles. The lack of dynamical fluctuations in temperature (upper limit on fluctuations in (pt) < 1%) and flavor composition (fluctuations in K/TT < 5%) indicates a high degree of equilibration in the produced matter. This naturally occurs in a QGP phase where equilibration processes are expected to proceed much faster than in a hadron gas. 3
Summary and outlook
Results of NA49 and other SPS experiments found predicted signatures of the QGP, unfortunately none of them are unequivocal. When the produced hadron density is extrapolated back to the initial stage of the reaction, these hadrons are found to be so densely packed that it is hard to imagine how they can exist as separate entities. Whether this dense state is a truly equilibrated QGP phase remains to be elucidated. In the remaining SPS heavy-ion program NA49 will attempt to locate the threshold for QGP production at lower beam energies and with lighter nuclei in order to more clearly confirm its fleeting existence. References 1. 2. 3. 4. 5. 6. 7. 8. 9.
S. Afanasiev et al, CERN-EP-99-001, NIM A430(1999)210 G.E. Cooper et al, Nucl. Phys. A661(1999)362c-365c H. Appelshaeuser et al, Phys. Rev. Lett. 82 (1999) 2471 M. Gazdzicki and M. Gorenstein, Acta Phys. Pol. B30(1999)2705 F. Becattini et al, Z. Phys.C76(1997)269; Eur. Phys. J. C5(1998)143 H. Appelshaeuser et al, Eur. Phys. J. C2 (1998) 661 A. Poskanzer and S. Voloshin, Phys. Rev. C58 (1998) 1671 H. Sorge, Phys. Rev. Lett. 82 (1999) 2048 J.-Y. Ollitrault, Phys. Rev. D46 (1992) 229
E X P L O R I N G T H E CHIRAL P H A S E T R A N S I T I O N IN H I G H - E N E R G Y COLLISIONS J 0 R G E N RANDRUP Nuclear Science Division, Lawrence Berkeley Laboratory, Berkeley, California 94720 E-mail: [email protected] Chiral symmetry plays an important role in nuclear physics but its manifestations have so far been testable only in the normal phase where the symmetry is spontaneously broken. More global explorations can be made by means of high-energy nuclear collisions, since a partial restoration of chiral symmetry is expected to occur within the highly agitated collision zone. Our current understanding of the chiral phase diagram will be reviewed, the special character of the induced nonequilibrium relaxation dynamics will be elucidated, and two areas of intensified future research will be highlighted.
1
Introduction
As the new century dawns, the field of nuclear physics encompasses a broad spectrum of research frontiers at which many promising research issues present themselves. Chiral symmetry constitutes a central concept in modern nuclear physics1 and the present discussion focusses on the prospects for exploring its manifestations over a wide range of environments by means of high-energy nuclear collisions. The strong interaction possesses chiral symmetry only to the degree that the basic quark masses can be neglected. Moreover, the self-interaction of the gluon fields encourages the spontaneous breaking of the symmetry in systems at low excitations, causing the familar vacuum to be a chiral condensate with a finite expectation value. Although chiral symmetry has proven to provide a successful framework for understanding many features in hadronic physics, the confrontation with experimental data has been primarily limited to the strongly broken phase. In the highly excited environments produced in high-energy collisions, the chiral condensate is significantly weakened and the system may (briefly) achieve approximate chiral symmetry. The subsequent non-equilibrium relaxation process may then produce observable signals that could be exploited to test our global understanding of chiral symmetry in strong interactions 2 ' 3,4 . This presentation reviews our current expectations regarding this novel phenomenon and identifies two important topics for intensified research, namely the development of a practical treatment of real-time non-equilibrium quantumfield dynamics in a non-uniform environment and the extension of the models from the ud SU(2) sector to the uds SU(3) sector of the strong interactions. 105
106
2
Equilibrium Properties
A simple theoretical tool for the global exploration of chiral symmetry in baryon-free systems is provided by the linear a model. It describes the 0(4) chiral field 0(r, i) = (a, n) by means of a simple effective quartic interaction, c = \d»<\>°d»4>-
2 \{4>°
+ Ha
, 0 o 0 = (r(r)2+7r(r)-7r(r) . (1)
The parameters, A, v, H, may be fixed to reproduce the pion decay constant fn, the free pion mass m w , and the mass of the schematic a meson, mCT. Useful insight can be gained by considering macroscopically uniform matter. The chiral field can then be decomposed, 0(r, t) — 0(t) + 64>(r,t), where 5(f) represents quasi-particle agitations relative to the (false) vacuum characterized by the 0(4) chiral order parameter >. Figure 1 {left) shows the associated free energy density FT = Vr(^o,Xo) _ TST(<{>O) along the a axis, with (To = !>ocosxo- At each temperature T, the minimum in FT represents the equilibrium value of the order parameter and it shrinks steadily as T is increased, thus producing a crossover from the strongly broken to the approximately symmetric chiral phase: Cooling of uniform matter 7"o=400 MeV
&100
-120 -80 -40 0 40 80 Order parameter a0 (MeV)
120
0
20 40 60 80 100 Order parameter $0 = l«l»l (MeV)
Figure 1: Left: The free energy density FT calculated with the semi-classical mean-field approximation to the linear a model, as a function of the order parameter tro for various values T of the quasi-particle gas. At each temperature, the solid dots indicate the equilibria. For T < To = V%v, the order parameter must exceed a certain minimum value before all quasiparticle modes are stable; the corresponding end points are connected by the dotted curve. The top dashed curve is the bare potential obtained for T = 0 . Right: The combined dynamical evolution of the order parameter ao = (c) and the field fluctuations {(50 2 } 1 / 2 . The dashed curve connects the equilibria from T=0 to above 500 MeV and the unstable region within which fi% < 0 is shown by the shaded region. Each system has been prepared in thermal equilibrium at To — 400 MeV and then subjected to a Rayeligh cooling, emulating a uniform expansion in D dimensions.
107
3
D C C Dynamics
If the field fluctuation is employed as a measure of the agitation, as in fig. 1 (right), the resulting diagram applies to any state of the system and is thus suitable for displaying non-equilibrium dynamics. In particular, it is possible to display the trajectories that result from the application of a Rayleigh cooling to initially hot chiral matter in equilibrium,
[D + A(0o0-« J )]0(r,t) =
He„---
(2)
which emulates a Bjorken-type scaling expansion in D dimensions.6 Two general features are apparent: 1) only very rapid cooling rates lead into the supercritical region where spontaneous pion creation occurs, thus suggesting that this interesting quench scenario may be dynamically unreachable; 2) for any cooling rate, the order parameter ultimately settles into rather regular damped oscillations around the true vacuum which may be expected to produce parametric amplification of pionic modes with frequencies u^ « \ma. More refined scenarios considering rods endowed with a longitudinal Bjorken scaling expansion confirm the above expectations, as is illustrated in fig. 2 for either the dynamical trajectory (left) or the effective pion mass (right)?
T„= 240 MeV
T, = 240 MeV
— Bjorken matter
*
- Bjorken rod {R0 =6 fm) — Equilibrium path — Critical boundary "\ — Bjorken matter ot=1,2,3,4,5,10,20,30 fm/c -- Bjorken rod (R0=6 fm) •t=1,2,3,4,5,10,20,30 tm/c
Order parameter rji0 (MeV)
11A
11
Proper time t (fm/c)
Figure 2: Left: Phase evolution of the interior of a Bjorken rod prepared with a bulk temperature of Tn=240 MeV and a radius of i?o=6 fm and the corresponding evolution of Bjorken matter (obtained in the limit of a very thick Bjorken rod, Ro —> oo) In order to be able to display an arbitrary non-equilibrium scenario, the average field fluctuation ( o ^ 2 ) 1 / 2 = (oV2 4- (57T2)1/2 has been used in place of the temperature on the vertical axis. Also shown are the equilibrium path and the boundary of the supercritical region. The rod has been probed over a hollow cylindrical volume, K p < 3 (fm). The paths are marked at successive proper times T=1,2,3,4,5,10,20,30 fm/c. The two evolutions are similar early on, but the rod then relaxes significantly faster due to the generated transverse expansion. Right: The corresponding evolutions of the squared effective pion mass ^ 2 .
108
4
Vacuum Fluctuations
A full quantum-field treatment of the chiral dynamics is beyond current reach and the dynamical simulation studies have employed classical fields. Although much valuable insight can been gained in this manner, it is important to recognize that such treatments are not quantitatively reliable. This is perhaps best illustrated by considering a free pionic mode with a time-dependent mass, [D + ojl{t)}
n£ nal = Xk [nt% + \] - | > Xkn™«
(3)
which exceeds the classical result X^nj^11', since the amplification coefficient Xk is generally larger than unity. The expression brings out the fact that the quantum fluctuations and the statistical fluctuations combine in a democratic fashion as seeds for the parametric amplification. A quantitative impression of their relative importance can be gained from fig. 3 (left). It is evident that the thermal occupancies are never large compared to one half. Consequently, the vacuum fluctuations are never negligible. Efforts to develop a suitable quantum-field treatment are underway10 and an illustration is given in fig. 3 (right). It is seen that pions with momenta around 200 MeV/c are produced, even if the initial state contained no quasiparticles at all. The initial presence of thermal excitations will then typically increase the yield several fold. Thermal Equilibrium
100 200 300 Temperature T(MeV)
400
-200
0
200
Momentum p (MeV/c)
Figure 3: Left: The temperature dependence of the thermal occupanices of the quasipions as obtained with either a semi-classicial treatment 5 or an optimized perturbation method. 9 Right: The final pion spectrum arising solely from amplification of the vacuum fluctuations in one-dimensional systems with given effective mass functions n\(x,t) of forms similar to those obtained for Bjorken rods (see fig. 2 (right)).10
109
5
Strangeness
Since the temperatures of interest easily exceed the mass of the s quark, it is expected that strangeness must be incorporated. Within the linear a model, this involves an extension from its familiar SU(2) form to SU(3). The number of meson fields then increases from 4 to 18,
ag,C; a,0
,n+
a; 7r°,7r
Vo;
•*'
> a0 ' K* ", K*+, K*°, K*°; -,n+,K~ , K+, K°, K°
(4)
and the equilibrium value of the order parameter acquires a strange component, a -> (a, £), where ( = (ss). As is evident from Fig. 4 (left), the inclusion of strangeness severely impedes the restoration of chiral symmetry as the temperature is raised, thus casting doubt on the standard scenario in which approximate restoration is assumed to occur once T exceeds a few hundred MeV. On the other hand, as illustrated in Fig. 4 (right), the dynamics of the order parameter becomes more intricate. As a result, the kaon modes may experience parametric amplification in analogy to what happens for the pions. The neutral kaon fraction fK = (K° + K°)/(K+ + K~ + K° + K°) then has a uniform distribution, PK(/K) = 1, for an idealized source, while the neutral pion fraction /„ = Tr°/(n+ + -K~ + TT°) retains its SU(2) distribution. 11
•.s 60
S * ' I
D=3
% 100
1 7
2
ft i
60
*
40
AT=100MeV
80
— • SU(2) — • SU(3). m„=eoo MeV ~ • SU(3), mo=800 MeV — • S U ( 3 ) , m„-1000MeV — • SU(3), no kaon fluct.
•••-
20
0
20
40
60
m =1 GeV 80
Order parameter a [MeV]
100
20
40
60
80
100
Order parameter a [MeV]
Figure 4: Left: The thermal path of the two order parameter (
110
6
Outlook
Chiral symemtry is a fundamental concept in strong-interaction physics and hence it is of general interest to explore its properties over a wide range of environments. High-energy collisions between heavy nuclei provide a promising means for this undertaking, as dynamical explorations with the linear a model suggest that specific signatures may be practically observable. It is therefore worthwhile to intensify the research efforts in this area. On the theoretical side, many challenges remain and the present discussion has merely highlighted two of them: 1) the need for taking account of the vacuum fluctuations by developing a proper real-time non-equilibrium quantum-field treatment of the chiral dynamics and 2) the need for including the strangeness degrees of freedom into the description. Although both problems pose considerable difficulties, formal as well as practical, the available results do provide specific guidance with respect to what may constitute particularly informative observables. These include enhancements in the soft part of the transverse pion spectra, anomalous fluctuations in their multiplicities, azimuthal correlation patterns, and a widening of the distribution of the neutral pion fraction.7 Thus (apart from this latter observable which may be beyond early reach), there is reason to hope for the appearance of interesting data in the upcoming first round of RHIC experiments. Acknowledgments This work was supported by the Director, Office of Energy Research, Office of High Energy and Nuclear Physics, Nuclear Physics Division of the U.S. Department of Energy, under Contract No. DE-AC03-76SF00098. References 1. 2. 3. 4. 5. 6. 7. 8. 9. 10. 11.
V. Koch, Int. J. Mod. Phys. E6, 203 (1997). K. Rajagopal, Quark-Gluon Plasma 2 (R. Hwa), World Scientific (1995). J.-P. Blaizot and A. Krzywicki, Acta Phys. Polon. B27 (1996) 1687. J.D. Bjorken, Acta Phys. Polon. B28 (1997) 2773. J. Randrup, Phys. Rev. D55, 1188 (1997); Nucl. Phys. A616, 531 (1997). J. Randrup, Phys. Rev. Lett. 77, 1226 (1996). T.C. Petersen and J. Randrup, Phys. Rev. C61, 024906 (2000). J. Randrup, Heavy Ion Phys. 9, 289 (1999). S. Chiku and T. Hatsuda, Phys. Rev. D58, 076001 (1998). J. Randrup, LBNL-00000 (2000) in preparation. J. Schaffner-Bielich and J. Randrup, Phys. Rev. C59, 3329 (1999).
N U C L E A R COLLECTIVE FLOW IN HEAVY ION COLLISIONS AT SIS ENERGIES
LPC
N . B A S T I D for t h e F O P I Collaboration a Clermont-Ferrand, IN2P3-CNRS and Universite Blaise 63177 Aubiere Cedex, FRANCE E-mail: [email protected]
Pascal,
Directed flow and elliptic flow have been studied in heavy ion collisions at SIS energies with the FOPI detector. A particular interest is devoted to the results obtained from the Fourier expansion method which is used to describe the anisotropies of the particle azimuthal distributions. Comparisons with the predictions of the IQMD transport model are performed.
1
Introduction
Collective flow effects in heavy ion collisions are of great interest since they allow to extract information on the complex dynamics with the goal of drawing conclusions about the hot and dense nuclear matter and its underlying Equation of State (EoS). The FOPI Collaboration has realized a systematic study of in-plane flow component and out-of-plane flow component providing a complete and accurate set of experimental data. In this contribution we focus on flow results of charged baryons in Ru + Ru collisions at 400 AMeV and 1.69 AGeV. 2
Experimental setup and analysis method
The data have been measured with the FOPI detector 1 at GSI-Darmstadt. The FOPI detector is an azimuthally symmetric apparatus made of several sub-detectors which provide charge and mass determination over nearly the full 47r solid angle. The central part (33° < 9iab < 150°) is placed in a superconducting solenoid and consists of a drift chamber surrounded by a barrel of plastic scintillators. The forward part (1.2° < 6iab < 30°) is composed of a wall of plastic scintillators and an other drift chamber. The events are sorted out according to their degree of centrality by imposing cuts on the multiplicity of charged particles 2 . The global variable Erat 3 a
F O P I Collaboration : NIPNE Bucharest, Romania ; KFKI Budapest, Hungary ; LPC Clermont-Ferrand, France ; GSI Darmstadt, Germany ; FZR Dresden, Germany ; Seoul Korea University, South Korea ; Heidelberg University, Germany ; I T E P Moscow, Russia ; KI Moscow, Russia ; IReS Strasbourg, France ; Warsaw University, Poland ; RBI Zagreb, Croatia
in
112
is used for the selection of the most central collisions. The reaction plane is reconstructed using the standard transverse momentum analysis method 4 . The flow observables are corrected for the reaction plane dispersion and for autocorrelation and momentum conservation effects according to the methods described in 5 . 3
Results
Flow results obtained from light and intermediate mass fragments measured in semi-central Au + Au collisions between 100 AMeV and 800 AMeV can be found i n 6 . We report here on recent flow results obtained with charged baryons measured in semi-central Ru + Ru collisions with the central part of the FOPI detector. - u
-0.1
< b > = 2.9 fm - 0.7< y0 < - 0.5
L
CM0.2
geo>
=
4
-7fm
- 0.3 < y0 < - 0.1
-0.2 -0.3 -0.4
f
-0.5 -0.6 '-
V
A
% -m * A ^m
>J
ir
****, T
-0.7 -0.8 C)
•
0.4
0.8
*VHe A->A = 3 • ->d •~*P 1.2 1.6 2
Pt°
0.4
0.8
1.2
1.6
2
Pt°
Figure 1: vi (left) and V2 (right) Fourier coefficients as a function of the scaled transverse momentum p° for the normalized rapidity bins j/o mentioned on the figures. p° is the particle transverse momentum normalized t o the projectile momentum in the center-of-mass system and j/o denotes the particle rapidity divided by the beam rapidity in the center-of-mass system. T h e results concern charged baryons in semi-central Ru + Ru reactions a t 400 AMeV.
The in-plane flow component and the out-of-plane flow component are investigated by means of a Fourier expansion of azimuthal distributions, (p being the azimuthal angle of a particle with respect to the reconstructed reaction plane,
113
the dN/dip distributions can be parametrized by VQ (1 + 2J2li=i vn cos{rup)) where vn = < cos(rup) > are the Fourier coefficients. The first and the second coefficient, ^i and V2, quantify the strength of directed and elliptic flow, respectively. This procedure allows to simultaneously study the two flow components as a function of the particle transverse momentum pt and rapidity. The comparisons of the results with the predictions of transport models should provide deeper insight into the properties of the hot and dense nuclear matter. Figure 1 displays the pt dependence of v\ and v% of charged baryons measured in semi-central Ru + Ru collisions at 400 AMeV. Charged baryons have a negative «i for all pt. Since the rapidity window is located in the backward hemisphere, that means that charged baryons are positively flowing. The magnitude of directed flow increases with pt and tends to saturate in the highest pt range. The V2 coefficient is plotted near mid-rapidity where the magnitude of elliptic flow is the largest. This second order Fourier coefficient is negative (i.e. out-of-plane preferred emission) and strongly correlated with pt. Both directed flow and elliptic flow increase with the particle mass hence confirming that composite particles provide a good probe for collective flow effects.
^ o,
-1.2
1
-0.9
,
-0.6 < y(0) < -0.3
1
0 0.20.40.60.8 0 0.20.40.60.8 0 0.20.40.60.8 1 pt (GeV/c) Figure 2: v\ versus pt for protons in semi-central Ru -I- Ru reactions at 1.69 AGeV. The points refer to the data and the curves correspond t o t h e IQMD predictions. The results are shown for three rapidity windows mentioned on the figures.
The pt dependence of the proton directed flow data is compared with the predictions of the IQMD (Isospin Quantum Molecular Dynamics) model 7 in figure 2 for semi-central Ru + Ru collisions at 1.69 AGeV and three rapidity windows in the backward hemisphere. Two parametrizations of the EoS have
114
been studied, soft and hard, both including momentum dependent interactions, labelled SM and HM, respectively. The magnitude of directed flow is maximum in the target rapidity region and decreases when going towards mid-rapidity both in the model and in the data. The model fails in consistently describing the pt dependence of v\, although a reasonable agreement is found when ptintegrated flow data are used 8 . This seems to be a general feature of transport model calculations since a similar discrepancy has been observed from the comparison of experimental data with the RQMD model predictions for Au + Au reactions at 11 AGeV 9 and with the RBUU model predictions for Ni + Ni reactions at 1.93 AGeV and Ru + Ru reactions at 1.69 AGeV 1 0 . 4
Conclusion
The present experimental results reveal detailed information on the various dependences of directed and elliptic flow at SIS energies. The study of the Pt-differential flow should allow to constrain the main parameters used in the microscopic models. References 1. A. Gobbi, FOPI Collaboration, Nucl. Instrum. Methods A 324, 156 (1993) ; J. Ritman, FOPI Collaboration, Nucl. Phys. B (Proc. Suppl.) 44, 708 (1995). 2. J.P. Alard, FOPI Collaboration, Phys. Rev. Lett. 69, 889 (1992). 3. W. Reisdorf, FOPI Collaboration, Nucl. Phys. A 612, 493 (1997). 4. P. Danielewicz and G. Odyniec, Phys. Lett. B 157, 146 (1985). 5. C.A. Ogilvie et al., Phys. Rev. C 40, 2592 (1989) ; J.Y. Ollitrault, Nucl-ex/9711003 (1997). 6. V. Ramillien, FOPI Collaboration, Nucl. Phys. A 587, 802 (1995) ; N. Bastid, FOPI Collaboration, Nucl. Phys. A 622, 573 (1997) ; P. Crochet, FOPI Collaboration, Nucl. Phys. A 624, 755 (1997) ; P. Crochet, FOPI Collaboration, Nucl. Phys. A 627, 522 (1997) ; F. Rami, FOPI Collaboration, Nucl. Phys. A 646, 367 (1999) ; A. Andronic, FOPI Collaboration, Nucl. Phys. A, in press. 7. J. Aichelin, Phys. Rep. 202, 233 (1991) ; C. Hartnack et al., Eur. Phys. Journ. A 1, 151 (1998). 8. G.Q. Li and G.E. Brown, Nucl. Phys. A 636, 487 (1998) ; P.K. Sahu et al., Nucl. Phys. A 640, 693 (1998). 9. J. Barrette et al., Phys. Rev. C 56, 3454 (1997). 10. P. Crochet, FOPI Collaboration, Phys. Lett. B 486, 6 (2000).
H A D R O N OBSERVABLES F R O M H A D R O N I C T R A N S P O R T MODEL W I T H JET P R O D U C T I O N AT RHIC YASUSHI N A R A RIKEN
BNL
Research
Center,
Brookhaven National Laboratory, 11974, USA E-mail: [email protected]
Upton, New
York
The hadronic transport model with jet production is applied to explore heavy ion collision at RHIC energies. Higher hadronic resonance states up to 2GeV are also included in this model to handle hadronic resonance gas system. After showing some results on SPS data, we will show the predictions at RHIC energies. It is found that pQCD like transverse momentum shape becomes exponential at pj_ < 4GeV by hadronic rescattering.
1
Introduction
Relativistic heavy ion collider (RHIC) experiments at Brookhaven National Laboratory have started to explore a new form of matter. Microscopic transport models, for example, RQMD 3 or UrQMD 4 have been well tested at SPS or AGS energies. They have been recognized as a powerful tool to study relativistic heavy collisions. Those models are also used to give some predictions at RHIC 5 ' 6 , especially, observables which is considered to be related to the soft physics. Event generators based on perturbative QCD (pQCD) are proposed such as HIJING 1 , VNI 2 , emphasizing the importance of mini-jet productions. In this work, the shape of the high momentum part in the transverse momentum distributions will be investigated by the transport model JAM 7 . Particle production in JAM is based on hadronic resonances or strings formation and decay and pQCD hard scattering to simulate large energy range of particle interactions (lGeV < v ^ < 200GeV). We are interested in the particle spectra affected by the interactions in the hadronic resonance gas stage of the heavy ion collisions. 2
Model description
The main elastic hh idea from excitation
features included in JAM are as follows. (1) At low energies, incollisions are modeled by the resonance productions based on the RQMD and UrQMD. (2) Above the resonance region, soft string is implemented along the lines of the HIJING model 1 . (3) Multiple 115
116
. Pb(158AGeV)+Pb h' b<3.2fm (a)
10 ^
> 3.7
O
10
10 Z -a 10
T3
1
2 p ± (GeV/c)
3
JAM, Au+Au at RHIC b<3.0fm pion (b)
10
2 4 Px (GeV/c)
Figure 1. Fig. (a) shows transverse momentum distributions of negative particles (left) in Pb(158AGeV) + P b , 6 < 3.2fm together with the NA49 experimental data from 1 0 . (b) shows transverse momentum distributions of pions in Au + Au, y/s = 200AGeV, b < 3.0fm. Solid lines correspond to the JAM results with rescattering among hadrons, dotted lines are for without meson-baryon and meson-meson collisions.
minijet production is also included in the same way as the HIJING model in which jet cross section and the number of jet are calculated using an eikonal formalism for perturbative QCD (pQCD) and hard parton-parton scatterings with initial and final state radiation are simulated using PYTHIA 9 program. (4) Rescattering of hadrons which have original constituent quarks can occur with other hadrons assuming the additive quark cross section within a formation time. Note that in this calculations, neither nuclear shadowing effects, nor the effect of jet quenching is not included. 3
Results
In Fig. 1(a), we compare JAM results to NA49 experimental data 10 for the central collision of P b + P b at 158AGeV. The calculations without mesonbaryon and meson-meson interactions (dotted lines) give totally different shape, however it can be seen that the calculations with full rescattering among hadrons well reproduce experimental data. In Fig. 1(b), we give a predictions from JAM calculations on pion distributions with and without rescattering. The shape of the pion transverse momentum below pi. < 3 - 4GeV/c becomes exponential from the pQCD like spectrum after hadronic rescatter-
117
60
Net proton (p-p)
(a) full no.reseat.
0 rapidity
Figure 2. Solid lines denote calculations with full rescattering. Dotted lines denote calculations without any meson-baryon and meson-meson collisions, (a) Rapidity distribution of net protons, (b) Rapidity distribution of net baryons. (c) Rapidity distribution of negative pions. (d) Rapidity distribution of antiprotons. (e) Rapidity distribution of negative kaons. (d) Rapidity distribution of positive kaons.
ing.
Figs. 2 show rapidity distributions of (a) net protons ,(b) net baryons, (c) negative pions, (d) antiprotons, (e) negative kaons, (d) positive kaons in central Au + Au collisions (b < 3fm) at ^/s = 200AGeV. Solid lines correspond to calculations with full hadronic rescattering, while dotted lines are for without
118
meson-baryon and meson-meson interactions. The yield of kaons are largely affected by rescattering. This is consistent with AMPT 8 and UrQMD 6 calculations. We have to care about this increase of the strangeness by hadronic interactions when we discuss about the enhancement of strangeness by QGP formation. We also see that antiproton yield is almost the same after finalstate hadronic interactions consisting with AMPT ampt and UrQMD 6 . 4
Conclusions
In conclusion, the JAM model has been used to study Au + Au reactions at RHIC energies to give predictions on final hadronic spectra. We saw the effect of rescattering on the transverse momentum spectra at both SPS and RHIC energies. Hadronic rescattering generally increase the slope of the transverse momentum spectra at both SPS and RHIC energies. References 1. X. N. Wang, Phys. Rep. 280, 287 (1997); X. N. Wang and M. Gyulassy, Comp. Phys. Comm. 83, 307 (1994); 2. K. Geiger, Phys. Rep. 258, 238 (1995); Comp. Phys. Comm. 104, 70 (1997); 3. H. Sorge, Phys. Rev. C 52, 3291 (1995). 4. S.A. Bass, M. Belkacem, M. Bleicher, M. Brandstetter, L. Bravina, C. Ernst, L. Gerland, M. Hofmann, S. Hofmann, J. Konopka, G. Mao, L. Neise, S. Soff, C. Spieles, H. Weber, L.A. Winckelmann, H. Stocker, W. Greiner, C. Hartnack, J. Aichelin and N. Amelin, Prog. Part. Nucl. Phys. 41, 225 (1998); 5. T. Schonfeld, H. Stocker, W. Greiner, and H. Sorge, Mod. Phys. Lett. A 8, 2631 (1993). 6. M. J. Bleicher, S. A. Bass, L. V. Bravina, W. Greiner, S. Soff, H. Stocker, N. Xu, and E. E. Zabrodin, Phys. Rev. C 62, 024904-1 (2000); 7. Y. Nara, N. Otuka, A. Ohnishi, K. Niita, S. Chiba, Phys. Rev. C 61, 024901 (2000); Y. Nara, Nucl. Phys. A 638, 555c (1998), http://quark.phy. bnl.gov/~ ynara/jam/. 8. B. Zhang, C. M. Ko, B. A. Li, and Z. Lin, Phys. Rev. C 61, 067901 (2000). 9. T. Sjostrand, Comp. Phys. Comm. 82, 74 (1994); PYTHIA 5.7 and JETSET 7.4 Physics and Manual, http://thep.lu.se/tf2/staff/torbjorn/Welcome.html. 10. NA49, H. Appelshaueser et al, Phys. Rev. Lett. 82, 2471 (1999).
SIMULTANEOUS HEAVY ION DISSOCIATION AT ULTRARELATIVISTIC ENERGIES I.A. P S H E N I C H N O V 1 ' 2 " , J . P . B O N D O R F 3 , S. M A S E T T I 2 , I.N. M I S H U S T I N 3 ' 4 , A. V E N T U R A 2 1 Institute for Nuclear Research, Russian Academy of Science, 117312 Moscow, Russia 2 Italian National Agency for New Technologies, Energy and Environment, 40129 Bologna, Italy 3 JViels Bohr Institute, DK-2100 Copenhagen, Denmark 4 Kurchatov Institute, Russian Research Center, 123182 Moscow, Russia We study the simultaneous dissociation of heavy ultrarelativistic nuclei followed by the forward-backward neutron emission in peripheral collisions at colliders. The main contribution to this particular heavy-ion dissociation process, which can be used as a beam luminosity monitor, is expected to be due to the electromagnetic interaction. The Weizsacker-Williams method is extended to the case of simultaneous excitation of collision partners which is simulated by the RELDIS code. A contribution to the dissociation cross section due to grazing nuclear interactions is estimated within the abrasion model and found to be relatively small.
1
Single and mutual electromagnetic dissociation
Let us consider a collision of heavy ultrarelativistic nuclei with the masses and charges (Ai,Z\) and {A^iZ-i) at the impact parameter b exceeding the sum of nuclear radii, b > Ri + R2. According to the Weizsacker-Williams (WW) method, the impact of the Lorentz-boosted Coulomb field of the nucleus At on the collision partner A^ is treated as the absorption of equivalent photons 1 . The mean number of photons absorbed by the nucleus Ai in such collision is: mA2(b) = JNz^Eutyo-A^E^dEx,
(1)
where the spectrum of virtual photons, Nz^iEx^b)l, and the total photoabsorption cross section, GA2{EX), are used. Assuming that the probability of multiphoton absorption is given by the Poisson distribution with the mean multiplicity m,i 2 (b), one has the cross section for the single electromagnetic dissociation to a given channel il: oo
a 1 (i) = 27r fbdbPAAb),
bc = 1.34{A\/3 + A\/3 - 0.75(A~1/3
°e-mail: [email protected] 119
+ A~1/3)),
(2)
120
At —* Ei y
At •
»
Ei
A*
A*
•"•2
" 2
Figure 1: Mutual electromagnetic dissociation of relativistic nuclei. The open and closed circles denote the elastic and inelastic vertices, respectively.
where bc is in fm and the probability of dissociation at impact parameter b is given by: PAl(b) = c -""i(») JdE1NZl(E1,b)aA2(E1)fA2(E1,i).
(3)
Here fAs (EI , i) is the branching ratio for the considered channel i in the absorption of a photon with the energy E\ on the nucleus A2. These values are calculated by photonuclear reaction models 2 ' 3 or taken from experiments. In the WW method, the graph for the mutual electromagnetic dissociation, Fig. 1, may be constructed from two graphs of the single dissociation by interchanging the roles of "emitter" and "absorber" at the secondary photon exchange. Such procedure is possible since the first emitted photon with Ei < Emax ~ 7/i? does not change essentially the total energy, EA — ~fMA, of the emitting nucleus, Emax/EA sa 1/RMA ~ 10~ 4 , and there are no correlations between the energies Ex and E2. In other words, both photon exchanges may be considered as independent processes. Moreover, at ultrarelativistic energies the collision time is much shorter then a typical deexcitation period when a nucleus loses its charge via the proton emission or fission. It means that the equivalent photon spectrum from the excited nucleus, A%, is equal to the spectrum from the nucleus in its ground state, A2, see Fig. 1. Therefore the cross section for the mutual electromagnetic dissociation of the nuclei Ai and A2 to given channels i and j is given by: oo
(4)
121
Substituting PA (b) for each of the nuclei one has:
<>m(i,j) = J J
dE1dE2Mm(EuE2)aA2(E1)aAl(E2)fA2(E1,i)fAl(E2,j), (5)
with the spectral function Mm{Ei,E2)
denned for the mutual dissociation:
oo
Nm (Ex, E2) = 2TT J bdbe-2m^NZl {E1, b)NZ2 (E2, b).
2
(6)
Nucleon removal in grazing nuclear collisions
The cross section for the abrasion of n neutrons and z protons from the projectile (Ax, Z\) in a peripheral collision with the target (A2, Z2) may be derived from the Glauber multiple scattering theory 4 ' 5 . A simple parameterization exists for the single neutron removal cross section 5 : —-r—
aG = 2TT\OC — — y
where A6 « 0.5 fm and Pesc w 0.75 is the probability for a neutron to escape without suffering the interaction with a spectator fragment. This can be extended to the case of mutual (In,In) emission:
<w(ln,ln) = ( ^ l - ) ( ^ i y 3
G
P l
c
.
(7)
Results and discussion
The results of the abrasion model for the charge changing cross sections of the single dissociation of 158A GeV 2 0 8 Pb ions are shown in Fig. 2 in addition to the electromagnetic contribution calculated by the RELDIS code. As one can see, the electromagnetic contribution dominates for few nucleon removal process. The interaction of knocked out nucleons with spectators and spectator de-excitation process itself were neglected in this version of the abrasion model. However, good agreement with the experimental data 6 is found. After this verification the model can be extrapolated to the energies of RHIC and LHC heavy-ion colliders. There is a proposal to use the simultaneous neutron emission for beam luminosity monitoring via the correlated registration of forward neutrons in zero degree calorimeters at RHIC 7 . The model predicts er m (ln, In) « 750 and 970 mb for AuAu and PbPb collisions at RHIC and LHC, respectively, for the correlated single neutron emission.
122
158AGeVPbPb
- .!•-•"'"I
70
72
74
76
78
80
82 Z
Figure 2: Charge changing cross sections of 2 0 8 P b ions at CERN SPS energies. The solid and dashed-line histograms are the RELDIS code results with and without contribution of the abrasion process, respectively. The points are experimental d a t a 6 .
We found that an important part of the dissociation events leading to the single neutron emission is accompanied by the emission of charged particles and nuclear fragments. Such events are due to the equivalent photon absorption above the Giant Resonance region. For grazing nuclear collisions Eq. (7) gives anuc(ln, In) as 100 mb. Our detailed consideration of the mutual nuclear dissociation is in progress. We are grateful to A.S. Botvina, G. Dellacasa, J.J. Gaardh0je, G. Giacomelli, A.B. Kurepin and S. White for useful discussions. I.A.P. and I.N.M. are indebted to the Organizing Committee of Bologna 2000 Conference for the kind hospitality and financial support. References 1. LA. Pshenichnov et al., Phys. Rev. C60, 044901 (1999) and references therein. 2. M.B. Chadwick et al., Acta Phys. Slov. 45, 633 (1995). 3. A.S. Iljinov et al., Nucl. Phys. A616, 575 (1997). 4. J. Hiifner et al., Phys. Rev. C12, 1888 (1975). 5. C.J. Benesh et al., Phys. Rev. C40, 1198 (1989). 6. H. Dekhissi et al., Nucl. Phys. A662, 207 (2000). 7. A.J. Baltz et al., Nucl. Instrum. Methods A417, 1 (1998).
D I R E C T P H O T O N P R O D U C T I O N I N 158 A G E V 2 0 8 p B + 2 0 8 p B COLLISIONS
T. PEITZMANN University
of Minster,
D-48149
Miinster,
Germany
F O R T H E WA98 C O L L A B O R A T I O N A measurement of direct photon production in Pb+ P b collisions at 158 A GeV has been carried out in the CERN WA98 experiment. The invariant yield of direct photons in central collisions is extracted as a function of transverse momentum in the interval 0.5 < px < 4 GeV/c. A significant direct photon signal, compared to statistical and systematical errors, is seen at py > 1.5 GeV/c. The results constitute the first observation of direct photons in ultrarelativistic heavy-ion collisions which could be significant for diagnosis of quark gluon plasma formation.
1
Introduction
The observation of a new phase of strongly interacting matter, the quark gluon plasma (QGP), is one of the most important goals of current nuclear physics research. To study QGP formation photons (both real and virtual) were one of the earliest proposed signatures 3 ' 4 . They are likely to escape from the system directly after production without further interaction, unlike hadrons. Thus, photons carry information on their emitting sources from throughout the entire collision history, including the initial hot and dense phase. Recently, it was shown by Aurenche et al. 5 that photon production rates in the QGP when calculated up to two loop diagrams, are considerably greater than the earlier lowest order estimates 6 . Following this result, Srivastava 7 has shown that at sufficiently high initial temperatures the photon yield from quark matter may significantly exceed the contribution from the hadronic matter to provide a direct probe of the quark matter phase. A large number of measurements of prompt photon production at high transverse momentum (p? > 3GeV/c) exist for proton-proton, protonantiproton, and proton-nucleus collisions (see e.g. 8 ) . First attempts to observe direct photon production in ultrarelativistic heavy-ion collisions with oxygen and sulphur beams found no significant excess 9>10>u>12. The WA80 collaboration 12 provided the most interesting result with a p y dependent upper limit on the direct photon production in S+Au collisions at 200AGeV. In this paper we report on the first observation of direct photon production in ultrarelativistic heavy-ion collisions.
123
124
2
Data Analysis
The results are from the CERN experiment WA98 13 which consists of large acceptance photon and hadron spectrometers. Photons are measured with the WA98 lead-glass photon detector, LEDA, which consisted of 10,080 individual modules with photomultiplier readout. The detector was located at a distance of 21.5 m from the target and covered the pseudorapidity interval 2.35 < rj < 2.95 (ycm = 2.9). The particle identification was supplemented by a charged particle veto detector in front of LEDA. The results presented here were obtained from an analysis of the data taken with Pb beams in 1995 and 1996. The 20% most peripheral and the 10% most central reactions have been selected from the minimum bias cross section (a-minbia, » 6300mb) using the measured transverse energy ET- In total, sb 6.710 6 central and » 4.3-106 peripheral reactions have been analyzed. The extraction of direct photons in the high multiplicity environment of heavy-ion collisions must be performed on a statistical basis by comparison of the measured inclusive photon spectra to the background expected from hadronic decays. Neutral pions and 77 mesons are reconstructed via their 77 decay branch. For a detailed description of the detectors and the analysis procedure see 14 . The final measured inclusive photon spectra are then compared to the calculated background photon spectra to check for a possible photon excess beyond that from long-lived radiative decays. The background calculation is based on the measured 7T° spectra and the measured 7//7r°-ratio. The spectral shapes of other hadrons having radiative decays are calculated assuming rayscaling 15 with yields relative to 7r°'s taken from the literature. It should be noted that the measured contribution (from 7r° and 77) amounts to « 97% of the total photon background. 3
Results
Fig. 1 shows the fully corrected inclusive photon spectra for peripheral and central collisions. The spectra cover the pr range of 0.3 — 4.0 GeV/c (slightly less for peripheral collisions) and extend over six orders of magnitude. Fig. 1 also shows the distributions of neutral pions which extend over a similar momentum range with slightly larger statistical errors. The ratio of measured photons to calculated background photons is displayed in Fig. 2 as a function of transverse momentum. The upper plot shows the ratio for peripheral collisions which is seen to be compatible with one, i.e. no indication of a direct photon excess is observed. The lower plot shows the
125
158AGeV ^ P b + ^ P b 10I. Central
cir 10
5
o o o a o a
10' 10
o
a
•
••
'• o
10
Peripheral
o
o
o
a
*
"•••ttt
••
10
'% •
tit til
10 10
0
0.5
2
7.5
2
2.5
J
i.5
4
Transverse Momentum (GeV/c)
Figure 1. The inclusive photon (circles) and w° (squares) transverse momentum distributions for peripheral (open points) and central (solid points) 158 A GeV 2 0 8 P b + 2 0 8 P b collisions. The d a t a have been corrected for efficiency and acceptance. Only statistical errors are shown.
same ratio for central collisions. It rises from a value of » 1 at low pr to exhibit an excess of about 20% at high PT • A careful study of possible systematical errors is crucial for the direct photon analysis. The largest contributions are from the 7 and 7T° identification efficiencies and the uncertainties related to the 77 measurement. It should be emphasized that the inclusive photon and neutral meson (the basis for the background calculation) yields have been extracted from the same detector for exactly the same data sample. This decreases the sensitivity to many detector related errors and eliminates all errors associated with trigger bias or absolute yield normalization. Full details on the systematical error estimates are given in 14 . The total py-dependent systematical errors are shown by the shaded regions in Fig. 2. A significant photon excess is clearly observed in
126
l.S
-
1.6
-
1.4
r
1.2
-
158AGeV ^ P b + ^ P b Peripheral Collisions
^aa»y»T*«^fr"|J£
1
l°*
'-
Z . 0.6
-
i:
1.4
Central Collisions
:
1.2
ii
S
I 7
•
-
•
••
-
^
IT ;
0.8 0.6
i • • .i,in.i,i, , i
'
0.5
I
1.5 2 2.5 3 3.5 Transverse Momentum (GeV/c)
4
Figure 2. The TMeas/^Bkgd ratio as a function of transverse momentum for peripheral (part a)) and central (part b)) 158 A GeV 2 0 8 P b + 2 0 8 P b collisions. The errors on the d a t a points indicate the statistical errors only. The p-r-dependent systematical errors are indicated by the shaded bands.
central collisions for pr > 1.5GeV/e. The final invariant direct photon yield per central collision is presented in Fig. 3. The statistical and asymmetric systematical errors of Fig. 2 are added in quadrature to obtain the total upper and lower errors shown in Fig. 3. An additional px-dependent error is included to account for that portion of the uncertainty in the energy scale which cancels in the ratios. In the case that the lower error is less than zero a downward arrow is shown with the tail of the arrow indicating the 90% confidence level upper limit ("/Excess + 1-28 C[/ pper ). No published prompt photon results exist for proton-induced reactions at the y/s of the present measurement. Instead, prompt photon yields for proton-induced reactions on fixed targets at 200 GeV are shown in Fig. 3 for comparison 16 . 17 . 18 . These results have been scaled for comparison with the present measurements according to the calculated average number of nucleon-
127
1.5
2
2.5
3
3.5
4
4.5
Transverse Momentum (GeV/c)
Figure 3. The invariant direct photon multiplicity for central 158 A GeV 2 0 8 P b + 2 0 8 P b collisions. The error bars indicate the combined statistical and systematical errors. Data points with downward arrows indicate unbounded 90% CL upper limits. Results of several direct photon measurements for proton-induced reactions have been scaled to central 208 P b - | - 2 0 8 P b collisions for comparison.
nucleon collisions (660) for the central P b + P b event selection and according to the beam energy under the assumption that Ed^a-^/dp3 = f(xT)/s2, where XT = 2pr/y/s 19 . This comparison indicates that the observed direct photon production in central 2 0 8 Pb+ 2 0 8 Pb collisions has a shape similar to that expected for proton-induced reactions at the same -y/s but a yield which is enhanced. 4
Summary
The first observation of direct photons in ultrarelativistic heavy-ion collisions has been presented. While peripheral P b + P b collisions exhibit no significant
128
photon excess, the 10% most central reactions show a clear excess of direct photons in the range of pr greater than about 1.5GeV/c. The invariant direct photon multiplicity as a function of transverse momentum was presented for central 208 pb-(- 208 Pb collisions and compared to proton-induced results at similar incident energy. The comparison indicates excess direct photon production in central 2 0 8 Pb+ 2 0 8 Pb collisions beyond that expected from protoninduced reactions. The result suggests modification of the prompt photon production in nucleus-nucleus collisions, or additional contributions from preequilibrium or thermal photon emission. The result should provide a stringent test for different reaction scenarios, including those with quark gluon plasma formation, and may provide information on the initial temperature attained in these collisions. References 1. 2. 3. 4. 5. 6. 7. 8. 9. 10. 11. 12. 13. 14. 15. 16. 17. 18. 19.
M.C. Abreu et al., Phys. Lett. B 410, 337 (1997). E. Andersen et al., Phys. Lett. B 449, 401 (1999). E.L. Feinberg, Nuovo Cimento 34 A, 391 (1976). E. Shuryak, Phys. Lett. B 78, 150 (1978). P. Aurenche, F. Gelis, H. Zaraket, and R. Kobes, Phys. Rev. D 58, 085003 (1998). J. Kapusta, P. Lichard, and D. Seibert, Phys. Rev. D 44, 2774 (1991). D.K. Srivastava, Eur. Phys. J. C 10, 487 (1999). W. Vogelsang and M.R. Whalley, J. Phys. G: Nucl. Part. Phys. 23, Al (1997). HELIOS Collaboration, T. Akesson et al., Z. Phys. C 46, 369 (1990). WA80 Collaboration, R. Albrecht et al., Z. Phys. C 5 1 , 1 (1991). CERES Collaboration, R. Baur et al, Z. Phys. C 71, 571 (1996). WA80 Collaboration, R. Albrecht et al., Phys. Rev. Lett. 76, 3506 (1996). WA98 Collaboration, Proposal for a large acceptance hadron and photon spectrometer, 1991, Preprint CERN/SPSLC 91-17, SPSLC/P260 WA98 Collaboration, M.M. Aggarwal et al, nucl-ex/0006007, submitted to Phys. Rev. C. M. Bourquin and J.-M. Gaillard, Nucl. Phys. B 114, 334 (1976). E704 Collaboration, D.L. Adams et al., Phys. Lett. B 345, 569 (1995). E629 Collaboration, M. McLaughlin et al., Phys. Rev. Lett. 5 1 , 971 (1983). NA3 Collaboration, J. Badier et al., Z. Phys. C 3 1 , 341 (1986). J.F. Owens, Rev. Mod. Phys. 59, 465 (1987).
D e f o r m a t i o n a n d O r i e n t a t i o n Effects i n U r a n i u m - o n - U r a n i u m C o l l i s i o n s at R e l a t i v i s t i c E n e r g i e s
Bao-An Li Department of Chemistry and Physics, P.O. Box 419> Arkansas State University, State University, AR 72467-0419, USA Email: [email protected] Deformation and orientation effects on compression, elliptic flow and particle production in uranium on uranium collisions (UU) at relativistic energies are studied within the transport model ART. The density compression in tip-tip UU collisions is found to be about 30% higher and lasts approximately 50% longer t h a n in bodybody or spherical UU reactions. The tip-tip UU collisions are thus more probable to create the QGP at AGS and SPS energies while the body-body UU collisions are more useful for studying properties of the QGP at higher energies.
1
Introduction
Prospects for new physics in uranium-on-uranium (UU) collisions due to deformation and orientation effects have recently generated much interests in the relativistic heavy-ion community 1 ' 2 ' 3 ' 4 . In particular, UU collisions have been proposed to extend beyond P b + P b collisions to better understand the J/ip suppression mechanism at C E R N ' s SPS. Many other outstanding issues m a y also be resolved by studying deformation and orientation effects in UU collisions at relativistic energies. One of the most critical factors to all of these issues is the m a x i m u m achievable energy density in UU collisions. Because of the deformation, UU collisions at the same b e a m energy and impact parameter b u t different orientations are expected to form dense m a t t e r with different compressions and lifetimes. In particular, the large deformation of u r a n i u m nuclei lets one gain particle multiplicity, energy density a n d longer reaction time by aligning the two u r a n i u m nuclei with their long axes head-on (tip-tip). We report here some results of a study on UU collisions within the relativistic transport model ART 6 . 2
D e f o r m a t i o n a n d O r i e n t a t i o n Effects o n R e a c t i o n D y n a m i c s
Uranium is approximately an ellipsoid with a long and short semi-axis Ri — R • (1 + | 6 ) and R, = R • (1 — | 6 ) , where R is the equivalent spherical radius and 6 is the deformation parameter. For 238U, one has 6 = 0.27 and thus a long/short axis ratio of about 1.3. Among all possible orientations between two colliding u r a n i u m nuclei, the tip-tip (with long axes head-on)and body-body 129
130
Figure 1: The evolution of central baryon density in Au-Au (filled circles), tip-tip (solid line), body-body (dotted line) and sphere-sphere (dashed line) UU collisions at a beam energy of 20 GeV/nucleon and an impact parameter of 0 (upper panel) and 6 fm (lower panel), respectively.
(with short axes head-on and long axes parallel in the reaction plane) collisions are the most interesting ones. Shown in Fig. 1 are the evolution of central baryon densities in the UU collisions at a beam energy of 20 GeV/nucleon and an impact parameter of 0 and 6 fm, respectively. While the body-body UU collisions lead to density compressions comparable to those reached in the Au-Au and spherical UU collisions, a 30% more compression is obtained in the tip-tip UU collisions at both impact parameters. The high density phase (i.e., with p/po > 5) in the tip-tip collisions lasts about 3-5 fm/c longer than the body-body collisions. The higher compression and longer passage time render the tip-tip UU collisions the most probable candidates to form the QuarkGluon-Plasma (QGP). 3
Deformation and Orientation Effects on Elliptic Flow
The elliptic flow reflects the anisotropy in the particle transverse momentum (pt) distribution at midrapidity, V2=<(pl-P2y)/P2t>,
(1)
where px (py) is the transverse momentum in (perpendicular to ) the reaction plane and the average is taken over all particles in all events. Shown in Fig. 2 are the nucleon elliptic flow as a function of impact parameter in UU collisions with different orientations at a beam energy of 10 GeV/nucleon and an impact parameter of 6 fm. We initialized the two uranium nuclei such that their long axes are in the reaction plane in both tip-tip and body-body collisions. It is seen that both the tip-tip and sphere-sphere collisions lead to a strong "in-plane
131
a
XJ-*-Ba**TJ. t i p — t i p
E / A = 1 0
GOV
a p h o r o - ^ s g l i o i
a^8*,
Figure 2: The impact parameter dependence of nucleon elliptic flow in the tip-tip (solid line), body-body (dotted line) and sphere-sphere (dashed line) UU collisions at a beam energy of 10 GeV/nucleon.
flow" (positive t^) while the body-body reactions result in a large "squeezeout" (negative i^)- Unique to the body-body UU collisions, the strength of elliptic flow is the highest in the most central collisions where the shadowing effect in the reaction plane is the strongest. While in tip-tip and sphere-sphere UU collisions the elliptic flow vanishes in the most central collisions due to symmetry. Similar results are also found for pions.
4
Deformation and Orientation Effects on Particle Production
Shown in Fig. 3 are the multiplicities of pions and positive kaons as a function of impact parameter. The maximum impact parameter for the tip-tip and body-body UU collisions are 2RB and 2Ri, respectively. As one expects the central (with 6 < 5 fm) tip-tip UU collisions produce more particles due to the higher compression and the longer passage time of the reaction. While at larger impact parameters, the smaller overlap volume in the tip-tip collisions leads to less particle production than the body-body and sphere-sphere reactions. Also as one expects from the reaction geometry, the multiplicities in the bodybody collisions approach those in the sphere-sphere collisions as the impact parameter reaches zero. In the most central collisions, the tip-tip UU collisions produce about 15% (40%) more pions (positive kaons) than the body-body and sphere-sphere UU collisions. Compared to pions, kaons are more sensitive to the density compression since most of them are produced from second chance particle (resonance)-particle (resonance) scatterings at the energies studied here.
132 a o o i • . . - — i — • • • • ! • • — • • ! • • • • — i
• • • •—r
*> CTm.) Figure 3: The impact parameter dependence of pion (upper panel) and positive kaon (lower panel) production in UU collisions.
5
Summary
In summary, using A Relativistic Transport model we have studied the deformation and orientation effects on the compression, elliptic flow and particle production in uranium on uranium (UU) collisions at relativistic energies. The compression in the tip-tip UU collisions is about 30% higher and lasts approximately 50% longer than in the body-body or spherical UU collisions. Moreover, we found that the nucleon elliptic flow in the body-body UU collisions have some unique features. We would like to thank P. Braun-Munzinger, W.F. Henning, C M . Ko, A.T. Sustich, Matt Tilley and B. Zhang for helpful discussions. This work was supported in part by the National Science Foundation under Grant No. PHY0088934 and Arkansas Science and Technology Authority Grant No. OO-B-14. References 1. W.F. Henning, private communications. 2. P. Braun-Munzinger, Memo, to RHIC management, Sept. 18, 1992. 3. B.A. Li, Phys. Rev. C61, 021903 (2000); E.V. Shuryak, Phys. Rev. C61, 034905 (2000). 4. Ben-Hao Sa and A. Tai, nucl-th/9912028; S. Das Gupta and C. Gale, nucl-th/0003005; P.F. Kolb, J. Solfrank abd U. Heinz, hep-ph/0006129; C. Nonaka, E. Honda and S. Muroya, hep-ph/0007187. 5. B. A. Li and C. M. Ko, Phys. Rev. C52, 2037 (1995).
S U B T H R E S H O L D HEAVY - M E S O N A N D A N T I P R O T O N P R O D U C T I O N IN THE N U C L E U S - N U C L E U S COLLISIONS
V.G. Khlopin
Radium
A. T . D ' Y A C H E N K O Institute, 197022 St. Petersburg,
Russia
Production of K- mesons and antiprotons in heavy ion collisions is considered at energies below the production threshold in the free nucleon - nucleon collisions. The fluid dynamics model was used to describe the experimental data.For strange particles and antiparticles the associative character of the production was taken into account, wich reduced considerably the absolute value of the production crosssection.
1
Introduction
The collisions of heavy nuclei at intermediate energies (800-2000 MeV/nucleon) are of great interest when studying nuclear matter at high temperature and density. The determination of characteristics of secondaries allows to determine the thermal and the collective energy of the fireball, which are connected with an equation of state (EOS) of nuclear matter and the module of compression of nuclear matter [1-4]. The thresholds of the K+ and K~ - meson production in nucleon - nucleon collisions are equal to 1.58 GeV and 2.5 GeV respectively, the threshold of the p - production amounts to 5.6 GeV. The production of K - mesons and antiprotons at subthreshold energies is a result of the collective dynamics in the nucleus - nucleus collisions. The associative character of the strange particle production and antiparticle production processes should be included in the thermodynamic description of the system [5]. 2
T h e M o d e l for Calculations
The fluid dynamics model is most suitable for studying the nuclear matter EOS ( the dependence of pressure and density of energy on temperature and nucleon density ). The equation of state determining the dependence of the pressure P and the energy density e on the density p is a sum of kinetic terms and interaction terms (P = Pkin + Pint and e = efc,„ + eint). The invariant differential crosssection of the reaction A + B —>• K (p) + X with the emission of kaons and antiprotons can be written as
E
$=jS^IidiI
d d
* ^ -pvcose)f-
133
(1)
134
Here / is a distribution function for K - mesons (antiprotons). It reads as f(E,-p,t)=g(exp(-
,l(E — pv cos6 — u). -^ —)±l)-\
. i (2)
where g = 1 for kaons, and 2 for antiprotons, p is the momentum of the produced particle, E = y/p2 + m2, v(r, t), T(r, t) are the fields of velocites and temperatures respectively, which are the solutions of relativistic fluid dynamics equations, 7 = (1 — v 2 ) - ? is the Lorentz factor, I is the impact parameter, <j> is the azimuthal angle, \i is the chemical potential. The sign ± concerns fermions or bosons respectively. The absorption of particles, interacting with the environment can be taken into account by introducing the factor
Q = exp(-^p) )
(3)
where < R > is the mean radius of the area, from which the secondary particles are emitted, A being the particle free-path-length. 3
Results
The calculation of the K- meson and antiproton production cross-sections was carried out according to expressions (1) - (3) with Skyrme's interaction parameters equal to 60 = -1033 MeV fm 3 , 63 = 14000 MeV fm 6 [6,7]. The chosen parametrization of the effective forces corresponds to realistic values of compression module K = 398 MeV (hard EOS) and the normal density po = 0.17 fm"3. The associative character of the production process should be included in the thermodynamic description by the replacement of the local equilibrium distribution function (2) with the conditional distribution function [5] The comparison of the calculated differential K+ - meson production cross-section with experimental data [9] is shown in fig. 1 for the reaction 58 Ni + 5 8 Ni at angles 40° < 9 < 48° and energies of the Ni ions equal to 1.8 GeV/nucleon(l), 1 GeV/nucleon(2) and 0.8 GeV/nucleon(3), respectively. The effective free-path-length of the K+ - mesons in the nuclei was taken equal to 6 fm. As seen from fig. 1, not only the calculated and the experimental spectra slopes are in agreement, but also the absolute values of the experimental and the calculated differential cross-sections of the K+ - meson production. In fig. 2 the calculated and the experimental [10] invariant differential production cross-sections in 28Si + 2 8 Si collisions (2 GeV/nucleon) are shown for 7T_ - mesons (1), K+ - mesons (2), K~~ - mesons (3) and antiprotons(4) as
135
0.0
0.1
0.2
0.3
Fie. 1.
0.4 Ecm
'
0.5 GeV
0
200
400
600
Fig. 2.
800 Ecm
'
1000 MeV
Figure 1: The calculated (solid lines) and experimental (points) invariant differential A""*" - meson production cross-sections [9], in the reaction 58 iVi + 5 8 JVt at the angles 40 — 48° and energies of the Ni ions, equal to 1.8 GeV/nucleon (1), 1.0 GeV/nucleon (2) and 0.8 GeV/nucleon (3). Figure 2: The calculated (solid lines) and the experimental (points) [12] invariant differential production cross-sections in the reaction 2 8 5i + 2 8 Si at the energy of the Si ions 2 GeV/nucleone for 7r— - mesons (diamonds) (1), K+ - mesons (squares) (2), K - mesons (triangles) (3), antiprotons (circles) (4) as functions of the energy of the particles in the centre of mass system.
functions of energy of the particles in the center of mass system of the colliding nuclei. The calculation reproduces the effective experimental spectra slopes. The solid curve (1) obtained at A = 5 fm is closer to the experimental data, than the dashed curve (2) obtained at A = 3 fm. The pion production from Adecay as well as the thermal pion production should be taken into account. The calculated production cross-sections for K~ -mesons are overestimated compared with experimental data (dashed curve), which can be connected with a smaller antikaon free-path-length in the nucleus equal to 2 fm (see, e.g.[8]) (solid curve) compared with that for K+ - meson (Aj^ + ~ 6 fm). The calculation reproduces the absolute value of the experimental antiproton production cross-section. The free-path-length of the antiproton in the nucleus was taken to be equal to 1.6 fm. It is more difficult for antiprotons compared with mesons to leave the interaction zone. The calculation reproduces the energy spectrum of the produced particles as well as the absolute value of the K~ - meson production cross-section at the energies 1.85 GeV/nucleon and 1.66 GeV/nucleon
136
for the Ni + Ni reactions if the antikaon free-path-lenght is equal to A = 2 fm [11]. A m o m e n t u m dependence of the antiproton production cross-section at various energies and combinations of the colliding nuclei [12] is reproduced in the calculation. T h e relative value of the cross-section for various reactions is reproduced also [11] within the experimental errors, though experimental d a t a are rather scarce. Thus, the calculation within the framework of relativistic fluid dynamics model of heavy ion interactions can reproduce the experimental energy spect r a of the subthreshold antiprotons and K- mesons at various energies and various combinations of the colliding nuclei. For strange particle production and antiparticle production the associative character of the reaction was taken into account, which reduced considerably the absolute value of the A'-meson production cross-section. Though the experimental points are not too numerous, the calculation reproduces the experimental antiproton energy spectrum as well as the relative value of the production cross-section for various energies and various combinations of the colliding nuclei. A cknowledgment s T h e author expresses his gratitude to Professors O.V. Lozhkin, V.P. Eismont and A.A. Rimski-Korsakov for their interest in the work and support. References 1. H. Stocker and W . Greiner, Phys. Rep. 137, 277 (1986). 2. W . Cassing, V. Metag, U. Mosel and K. Niita, Phys. Rep. 1 8 8 , 363 (1990). 3. I.N. Mishustin, V.N. Russkikh and L.M. Satarov, Yad. Fiz. 5 4 , 429 (1991). 4. C. Hartnack, J. Aichelin, H. Stocker and W . Greiner, Phys. Rev. Lett. 72, 3767 (1994). 5. K.K. G u d i m a a n d V.D. Toneev, Yad. Fiz, 4 2 , 645 (1985). 6. A . T . D'yachenko, Bull. Russ. Acad. Set. (Phys.Ser.) 6 0 , No. 2, 195 (1996). 7. A . T . D'yachenko, Nucl. Phys. A 6 2 6 , 273 (1997). 8. W . Zwermann and B. Schiirmann Nucl. Phys. A 4 2 3 , 525 (1984). 9. R. B a r t h et al. (KaoS collaboration) Phys. Rev. Lett. 78, 4007 (1997). 10. A. Shor A et al. Phys. Rev. Lett. 6 3 , 2192 (1989). 11. A . T . D'yachenko, J. Phys. G: Nucl. Part. Phys. 26, 861 (2000). 12. A. Schrotter et al. Z. Phys. A 3 5 0 , 101 (1994).
P I O N I M A G I N G AT T H E AGS S.Y. PANITKIN, 7 N.N. AJITANAND, 12 J. ALEXANDER, 12 M. ANDERSON, 5 D. BEST, 1 F.P. BRADY, 6 T. CASE, 1 W. CASKEY, 6 D. CEBRA, 5 J. CHANCE, 5 P. CHUNG, 12 B. COLE, 4 K. CROWE, 1 A. DAS, 10 J. DRAPER, 5 M. GILKES, 11 S. GUSHUE, 2 M. HEFFNER, 5 A. HIRSCH, 11 E. HJORT, 1 1 L. HUO, 6 M. JUSTICE, 7 M. KAPLAN, 3 D. KEANE, 7 J. KINTNER, 8 J. KLAY, 5 D. KROFCHECK, 9 R. LACEY, 12 M.A. LISA, 10 H. LIU, 7 Y.M. LIU, 6 R. MCGRATH, 12 Z. MILOSEVICH, 3 , G. ODYNIEC, 1 D. OLSON, 1 C. PINKENBURG, 1 2 N. PORILE, 1 1 G. RAI, 1 H.G. RITTER, 1 J. ROMERO, 5 R. SCHARENBERG, 11 L.S. SCHROEDER, 1 B. SRIVASTAVA,11 N.T.B. STONE, 2 T.J.M. SYMONS, 1 S. WANG, 7 R. WELLS, 1 0 J. WHITFIELD, 3 T. WIENOLD, 1 R. WITT, 7 L. WOOD, 5 X. YANG, 4 W.N. ZHANG, 6 Y. ZHANG 4 (E895 COLLABORATION) 1 Lawrence Berkeley National Laboratory, Berkeley, California 94720 2 Brookhaven National Laboratoryj Upton, New York 11973 3 Carnegie Mellon University, Pittsburgh, Pennsylvania 15213 4 Columbia University, New York, New York 10027 5 University of California, Davis, California 95616 6 Harbin Institute of Technology, Harbin 150001, P. R. China 7 Kent State University, Kent, Ohio 44242 6 St. Mary's College of California, Moraga, California 94575 9 University of Auckland, Auckland, New Zealand 10 The Ohio State University, Columbus, Ohio 43210 11 Purdue University, West Lafayette, Indiana 47907 12 State University of New York, Stony Brook, New York 11794 Source imaging analysis was performed on two-pion correlations in central Au + Au collisions at beam energies between 2 and 8A GeV. We apply the imaging technique by Brown and Danielewicz, which allows a model-independent extraction of source functions with useful accuracy out to relative pion separations of about 20 fm. We found that extracted source functions have Gaussian shapes. Values of source functions at zero separation are almost constant across the energy range under study. Imaging results are found to be consistent with conventional source parameters obtained from a multidimensional HBT analysis.
1
Introduction
Measurements of two-particle correlations offer a powerful tool for studies of spatial and temporal behaviour of the heavy-ion collisions. The complex nature of the heavy-ion reactions requires utilization of different probes and different analysis techniques in order to obtaine reliable and complete picture
137
138
of the system created in the collision. We will discuss application of imaging technique of Brown-Danielewicz which allows one to reconstruct the entire source function, phase space density and entropy from any like-pair correlations in a model independent way. We will present recent results of the two-pion imaging analysis performed by the E895 Collaboration and discuss natural connection between imaging and traditional HBT. 1.1
E895 Experiment and Analysis Details
Detailed description of the experiment E895 can be found elsewhere 1 . The Au beams of kinetic energy 1.85, 3.9, 5.9 and 7.9A GeV from the AGS accelerator at Brookhaven were incident on a fixed Au target. The analyzed ir~ samples come from the main E895 subsystem — the EOS Time Projection Chamber (TPC) 2 , located in a dipole magnetic field of 0.75 or 1.0 Tesla. Results of the E895 w~ HBT analysis were reported earlier 3 , and in this paper, we focus on an application of imaging to the same datasets. 1.2
Source Imaging
In order to obtain the two-pion correlation function Ci experimentally, the standard event-mixing technique was used 4 . Negative pions were detected and reconstructed over a substantial fraction of 47r solid angle in the centerof-mass frame, and simultaneous measurement of particle momentum and specific ionization in the TPC gas helped to separate ir~ from other negatively charge particles, such as electrons, K~ and antiprotons. Contamination from these species is estimated to be under 5%, and is even less at the lower beam energies and higher transverse momenta. Momentum resolution in the region of correlation measurements is better than 3%. Event centrality selection was based on the multiplicity of reconstructed charged particles. In the present analysis, events were selected with a multiplicity corresponding to the upper 11% of the inelastic cross section for Au -I- Au collisions. Only pions with pr between 100 MeV/c and 300 MeV/c and within ±0.35 units from midrapidity were used. Another requirement was for each used 7r - track to point back to the primary event vertex with a distance of closest approach (DCA) less than 2.5 cm. This cut removes most pions originating from weak decays of long-lived particles, e.g. A and K°. Monte Carlo simulations based on the RQMD model 5 indicate that decay daughters are present at a level that varies from 5% at 4A GeV to 10% at 8A GeV, and they he preferentially at pr < 100 MeV/c. Finally, in order to overcome effects of track merging, a cut on spatial separation of two tracks was imposed. For pairs from both
139
2.5 it Ih
2
IS IS
O
-*-•-•-
_»
• * * * - •
1
• • • • • • •
• 2GeV
OS
• 4GeV
,
0
j
2 IS
" -W-
1
-^M^MM>
03
* 6GeV
0
i
T U T
'
0.05 0.1 Q lnv ,(GeWc)
-ChCbO^
o 8GeV
, 0.05 0.1 Q^GeV/c)
0.15
Figure 1. Measured two-pion correlation functions for Au + Au central collisions at four beam energies.
"true" and "background" distributions, the separation between two tracks was required to be greater than 4.5 cm over a distance of 18 cm in the beam direction. This cut also suppresses effects of track splitting 3 . Correlation functions were corrected for pion Coulomb interaction 3 . Imaging 6 ' 7 was used to extract quantitative information from the measured correlation functions. It has been shown 8 that imaging allows robust reconstruction of the source function for systems with non-zero lifetime, even in the presence of strong space-momentum correlations. The main features of the method are outlined below; see Refs 6 ' 7 ' 9 for more details. The two-particle correlation function may be expressed as follows:
CP(Q)-1 =
JdrK(Q,r)SP(T),
(1)
140
•3
10
• • • O
2GeV 4GeV 6GeV SGeV
10
"'«|
"MM «!
GO
10
i
1A
0
2
4
. .
i i i i i i
6
8
10
12
14
16
18
20
r,(fm) Figure 2. Relative source functions extracted from the pion correlation data at 2, 4, 6 and 8A GeV.
where AT = |$Q^(r)| 2 — 1 andQ^ is the relative wavefunction of the pair. The source function, Sp(r), is the distribution of relative separations of emission points for the two particles in their center-of-mass frame. The total momentum of the pair is denoted by P , the relative momentum by Q = p i — P2, and the relative separation by r. Since the correlation data is already Coulomb corrected, we may approximate the relative wavefunction of the pions with a noninteracting spin-0 boson wavefunction:
*¥(r) = ±(eW+e-«*'*').
(2)
The goal of imaging is the determination of the relative source function (Sp(r) in Eq. (1)), given Cp(Q). The problem of imaging then becomes the problem of inverting K(q, r) 1 0 . For pions, one might think that because Eq. 1 is a Fourier cosine transform it may be inverted analytically 6 to give the source
141 Table 1. Radius parameters of Gaussian fits to the extracted source (Rs) measured correlation functions(Rc , 2) for different energies.
Eb (AGeV) i?s(fm) Rc2 (fm)
2 6.70±0.04 6.39±0.2
4 6.35±0.03 6.05±0.1
6 5.56±0.03 5.51±0.15
functions and
8 5.53±0.05 5.61±0.28
function directly in terms of the correlation function: 5 P (r) = J^ysj
cos (Q • r) (C P (Q) - 1).
(3)
While this might work for vanishing experimental uncertainty, for realistic situations, this is a poor way to perform the inversion. When Fourier transforming, it is impossible to distinguish between statistical noise and real structure in the data n . Instead, we proceed as in 13 and expand the source in a Basis Spline basis: S(r) = E • SjBj(r). With this, Eq. (1) becomes a matrix equation d = J2j KijSj with a new kernel: Kij = jdrKiQi^Bjir).
(4)
Imaging reduces to finding the set of source coefficients, Sj, that minimize the X2. Here, X 2 = E<(<* " E ^ ^ / A 2 ^ . This set of source coefficients is Sj = Yli[(KT(A2C)-1K)-1KTB]ji(Ci - 1) where KT is the transpose of the kernel matrix. The uncertainty of the source is the square-root of the diagonal elements of the covariance matrix of the source, A 2 S = ( J ftT T (A 2 C)- 1 i£:)- 1 . It has been shown 9 that for the case of noninteracting spin-zero bosons with Gaussian correlations, there is a natural connection between the imaged sources and "standard" HBT parameters. Indeed, the source is a Gaussian with the "standard" HBT parameters as radii: S(r) =
_
X
exp (- \nri[R2]^11
,
,
(5)
where A is a fit parameter traditionally called the chaoticity or coherence factor in HBT analyses. Here, [-R2] is the real symmetric matrix of radius parameters [R2] =
Rl
Rl R\t
.
(6)
142
• S"V->0) o
S^fromHBT
£ 10'
6 , (GeV)
Figure 3. Values of inverse relative source functions at zero separation as a function of beam energy, obtained using imaging (solid circles) and HBT (open circles).
Eq. (5) is the most general Gaussian one may use, but usually one assumes a particular symmetry of the single particle source so that there is only one non-vanishing non-diagonal element R^. The asymptotic value of the relative source function at zero separation S(T -»• 0) is related to the inverse effective volume of particle emission, and has units of fm - 3 : S(r - • 0) =
A
(2V5F) V d e t [R>]
(7)
As it was shown in 9 5(r -+ 0) is an important parameter needed to extract the space-averaged phase-space density. Figure 1 shows measured angle-averaged two-pion correlation functions for central Au + Au collisions at 2, 4, 6 and &A GeV. Figure 2 shows relative source functions S(r) obtained by applying the imaging technique to the measured two-pion correlation functions. Note that the plotted points are for representation of the continuous source function and
143
Table 2. Fit parameters for the Bertsch-Pratt HBT parameterization of the pion correlation functions for E895 beam energies used for evaluation of S(r —• 0).
Eb (AGeV) A R0 (fm) Ra (fm) Ri (fm)
Rl (fm)
2 0.99±0.06 6.22±0.26 6.28±0.20 5.15±0.19 -2.43±1.71
4 0.74±0.03 5.79±0.16 5.37±0.11 5.15±0.14 0.43±1.03
6 0.65±0.03 5.76±0.23 5.05±0.12 4.72±0.18 2.17±1.20
8 0.65±0.05 5.49±0.31 4.83±0.21 4.64±0.24 -0.65±1.85
hence are not statistically independent of each other as the source functions are expanded in Basis Splines 13 . Since the source covariance matrix is not diagonal, the coefficients of the Basis spline expansion are also not independent which is taken into account during x2 calculations. With this technique, we may reconstruct the distribution of relative pion separations with useful accuracy out to r ~ 20 fm. The images obtained at each of the four E895 beam energies are rather similar in shape, and upon fitting with a Gaussian function, values of x 2 per degree of freedom between 0.9 and 1.2 are obtained. Results of fits to the source function and correlation functions are shown in Table 1. One can see that source radii extracted via both techniques are similar, further confirming the validity of a Gaussian source hypothesis. Figure 3 compares values related to effective volumes of pion emission inferred from the standard Bertsch-Pratt pion HBT parametrization (open circles) with the effective volumes derived from the image source functions shown in Fig. 2 (solid circles). Essentially, this figure plots the inverse of the left- and right-hand sides of Eq. 7. Results of the multidimensional Bertsch-Pratt fit to the pion correlation data have already been published in Ref. 3 and are reproduced in Table 2. It can be seen from Fig. 3 that the agreement between imaging and the HBT parametrization is fairly good. Values of source functions at zero separations estimated via either technique are approximately constant within errors across the 2 to 8A GeV beam energy range. In summary, we present measurements of one-dimensional correlation functions for negative pions emitted at mid-rapidity from central Au + Au collisions at 2, 4, 6 and 8A GeV. These correlation functions are analyzed using the imaging technique of Brown and Danielewicz. It is found that relative source functions S(r) have rather similar shapes and zero-separation intercepts S(r —> 0). Distributions of relative separation have been measured out to 20 fm, and the extracted source functions are approximately Gaussian. We have performed the first experimental check of the predicted connection
144
between imaging and traditional meson interferometry techniques and found that the two methods are in good agreement. This agreement paves the way for applications of the imaging method to the interpretation of pair correlations among strongly interacting particle such as protons, antiprotons, etc. Values of source functions at zero separation which are related to the pion effective volumes of emission are almost constant across the range of bombarding energies under study. Acknowledgments Parts of this paper are based on work done in collaboration with D. Brown and G.F. Bertsch 9 . This research is supported by US DOE, NSF and other funding, as detailed in Ref. 3 ' 9 . References 1. 2. 3. 4. 5. 6. 7. 8. 9. 10. 11. 12. 13.
G. Rai et al., proposal LBL-PUB-5399 (1993). G. Rai et al., IEEE Trans. Nucl. Sci. 37, 56 (1990). E895 Collaboration, M.A. Lisa et al., Phys. Rev. Lett. 84, 2798 (2000). G. Kopylov, Phys. Lett. 50B, 472 (1974). H. Sorge, Phys. Rev. C 52, 3291 (1995). D.A. Brown and P. Danielewicz, Phys. Lett. B398, 252 (1997). D.A. Brown and P. Danielewicz, Phys. Rev. C57, 2474 (1998). S.Y. Panitkin and D.A. Brown, Phys. Rev. C61, 021901 (2000). D.A. Brown, S.Y. Panitkin and G.F. Bertsch, Phys. Rev. C62, 014904 (2000). A. Tarantola, Inverse Problem Theory, Elsevier, (1987). D. Brown, nucl-th/9904063. D.A. Brown, nucl-th/0003021. D.A. Brown and P. Danielewicz, in preparation; C. de Boor, A Practical Guide to Splines, Springer-Verlag, (1978).
D I R E C T E D F L O W IN 4.2A G E V / C C + C A N D COLLISIONS
C+TA
J . M I L O S E V I C A N D L J . SIMIC Institute
of Physics, E-mail:
P.O. Box 68, 11080 Belgrade, Yugoslavia [email protected]
The directed flow of protons and ir— in 4.2A GeV/c C + C and C + T a collisions is studied using the Fourier analysis of the azimuthal distributions. It is found that protons exhibit strong directed flow. In C + C collisions, the directed flow of pions is too weak to be detected experimentally, while in C + T a collisions pions show significant directedflow. Quark-gluon-string model calculations confirm these results. This model suggests that the decay of resonances and rescattering of secondaries determine the proton and 7r~ directed flow.
T h e directed flow of protons a n d x - in 4.2A G e V / c C + C and C + T a collisions is studied using the Fourier analysis of the azimuthal distributions *. We analysed 9500 C + C and 1000 C + T a semicentral and central collisions obtained with the 2m propane bubble chamber, exposed a t JINR, D u b n a synchrophasotron 2 . T h e same type of analysis is performed using the events generated by the Quark-Gluon-String Model (QGSM) 3 . Additionally, we used QGSM to clarify which physical processes are responsible for flow effects. For analysing directed flow, a Fourier expansion — - « — [l + 2t>! cos(<£) + 2v2 cos{2
(1)
Zit
of the azimuthal distribution of particles is performed. T h e coefficients v 1 =(cos(<^)) and t>2=(cos(2<£)) quantify the directed and elliptic flow respectively. In Eq. (1), 4> is the particle azimuthal angle measured with respect to the (true) reaction plane. Since b o t h Pbeam and b are available in Q G S M , they are used to determine the true reaction plane. In the experiment, the reaction plane is estimated, for the i-th proton in each event, using the vector 4 ' 5
where rrij and p ^ j represent mass and transverse m o m e n t u m of the j'-th proton emitted in the forward (y > ycm +6, Wj = 1), or backward (y < ycm — 6, ujj = — 1) hemisphere. Here, ycm denotes the center of mass rapidity while the quantity 8 (=0.2) removes the protons emitted around the ycm which are not contributing to the determination of the reaction plane. In Eq. (2),
145
146
mays is the sum of the projectile and target masses. Determining the reaction plane in this way, we avoid an autocorrelation as well as a correlation due to momentum conservation. Because of the finite resolution of the experimentally determined reaction plane, we need to correct the Fourier coefficients v[ obtained with respect to this reaction plane. The relation between the 'true' Fourier coefficients v\ and the coefficients v[ is t>i = v[/ (cos(A#)). For the determination of the correction factor (cos(A<&)) we used the subevent method 1>6. We found (cos(A*))=0.56 (0.59) for C+C (C+Ta) collisions.
C+Ta protons 0.2
:
/ $ &
f+ y* -0.2 -
1
0
i
.
.
.
1 JC
.
i
.
.
.
.
i
,
,
Figure 1. Rapidity dependence of vi for protons and n— for 4.2A GeV/c C+C and C+Ta collisions. Filled circles represent the experimental v\ values, while the solid (dashed) line represents QGSM calculation for «i relative to the true (estimated) reaction plane.
0.1 . • data
=r QGSM
-1 o , , ,
i ,
,
, .
i
,
,
,
,
i
.
.
.
1
-1
rapidity Fig. 1 shows the experimentally determined v\ coefficient vs. y for protons and ic~ together with the v\ calculated by QGSM relative to the true and to the estimated reaction plane for both collision systems. The values of the two QGSM results for v\ are quite close for protons. The dependence of t;i on rapidity is characterized by a curve with a positive slope and with a zero-crossing at y = 0. This indicates positive directed flow with magnitude v\ fa 0.15 at beam rapidity. QGSM reproduces satisfactorily the shape and the magnitude of the v\ (y) data. For protons we also determined the slope F = d(px)/dy at midrapidity, using the relation vi = {px)/(pr)- After the normalisation to the mass number of the colliding systems we obtained the so called scaled flow, Fs = F/{Ayz + A\lz) = 22 ± 2 MeV/c (26 ± 2) MeV/c for C+C (C+Ta) collisions. These are in agreement with the observed trend 7 of a slow decrease of the directed flow values with increasing beam energy. For ir~ the experimental values of Vi indicate that the directed flow is non-existent
147
in C + C collision. This result is confirmed by Q G S M calculations of vi with respect t o t h e estimated reaction plane. However, t h e model calculations of v\ with respect t o t h e true reaction plane show t h e existence of a very weak negative directed flow of pions with value of 0.02 a t t h e target rapidity. This further suggests t h a t in t h e collisions of light nuclei, like C + C , t h e very weak flow can not b e observed experimentally because of the limited accuracy in the determination of the reaction plane. For C + T a collisions, v-± values are positive a t all rapidity bins. It means t h a t pions are preferentially emitted in t h e reaction plane from t h e target t o projectile. This behaviour could be explained by t h e shadowing effect of the T a target. Q G S M calculations show the same behaviour, b u t they overestimate t h e flow values.
C+C protons 0.2
J? * * * * * * ^ « ••*
S^* 0 * -0.2
C+Ta
_t3r
•PTi
. . , , i , , , , i , ,
-1
0
-••
0.1
1
1
0
0.4 -
fefifcogU±>To '^^FfcjfcjK* Vo^, -0.1 -•.
.
.
i
.
.
1
,
.
,
,
i
1
,
,
• •
0.2
.
- 1 0
i
JC"-*-
Jl"
Figure 2. R a p i d i t y d e p e n d e n c e of v\ for p r o t o n s a n d ir~ [solid line); for p r o t o n s a n d ir — originati n g from r e s o n a n c e s d e c a y (stars), p r i m a r y (full circles), a n d s e c o n d a r y (open circles) n o n - r e s o n a n t interact i o n s for 4 . 2 A G e V / c C + C and C + T a events generated with QGSM.
,
"
'£. -1
0
1
rapidity According t o Q G S M , 39% (43%) of the protons a n d 7 1 % (83%) of t h e TT~ in C + C ( C + T a ) interactions originate from decays of A ' s , p, w, t) a n d rf resonances. T h e rest originates from t h e 'non-resonant' primary a n d secondary interactions. Some of the protons a n d ir~ from primary interactions (5% in C + C a n d 1% in C + T a collisions) escape t h e collision zone without further rescattering. Therefore, in Q G S M , we separately evaluated t h e flow of protons a n d v~ originating from: (i) decay of resonances, (ii) primary and ( m ) secondary non-resonant interactions. Fig. 2 shows distributions of vi vs. y for protons a n d ir~ from these sources. In C + C collisions, protons (%~) from resonances decay a n d rescattered protons (ic~) show directed flow
148
(antiflow). For nonrescattered protons the vi(y) distribution is relatively flat, while nonrescattered pions show strong negative directed flow at the level of 0.12 at beam rapidity. These particles are produced a t the early stage of the collision. Antiflow of pions can be explained by the shadowing efFect. However, the shadowing m a t t e r is different for pions at different collision stages. At early times (less t h a n passing time) pions are shadowed by the cold spect a t o r s . Later, after the spectator m a t t e r leaves the collision zone, pions are shadowed by the participant nucleons. This m a y be the underlying mechanism t h a t leads to a different behavior of Vi for ir~ and protons. As in the case of C + C collisions, the decay of resonances a n d rescattering of secondaries determine the proton and ir~ flow. For nonrescattered particles, shadowing of the cold spectators is much stronger because of the much bigger size of the Ta target. This leads to the large positive vi values for nonrescattered particles. In summary, the directed flow of protons and ic~ in 4.2A G e V / c C + C a n d C + T a collisions was examined using the Fourier analysis of the azimuthal distributions. Additionally, we used Q G S M events t o clarify which physical processes are responsible for flow effects. It was found t h a t the protons exhibit strong positive directed flow in b o t h collisional systems. Q G S M reproduces satisfactorily the magnitude of this flow. For ir~ in C + C collisions, a directed flow was not found experimentally due to the poor reaction plane resolution. But, Q G S M calculations show a very small negative directed flow. In C + T a collisions, because of the target shadowing, v\ values of -K~ are positive at all rapidities. According to Q G S M two factors t h a t dominantly determine the proton and ir~ flow are the decay of resonances and rescattering of secondaries. Shadowing of the cold nuclear m a t t e r affects only the flow of nonrescattered particles produced at the early stage of the reaction, except fot protons in C + C collisions because of the smallness of the colliding nuclei. References 1. A. M. Poskanzer a n d S. A. Voloshin, Phys. Rev. C 5 8 , 1671 (1998). 2. E. A b d r a h m a n o v et al, Z. Phys. C 5, 1 (1980); H. Agakishiev et al, Z. Phys. C 12, 283 (1982); H. Agakishiev et al, Z. Phys. C 16, 307 (1983). 3. N. S. Amelin, K. K. G u d i m a and V. D. Toneev, Yad. Fiz. 5 1 , 512 (1990); Sov. J. Nucl. Phys. 5 1 , 327 (1990). 4. P. Danielewicz and G. Odyniec, Phys. Lett. B 157, 146 (1985). 5. C. A. Ogilvie et al, Phys. Rev. C 4 0 , 2592 (1989). 6. J.-Y. Ollitrault, nucl-ex/9711003. 7. J. Chance et al, EOS Collaboration, Phys. Rev. Lett. 7 8 , 2535 (1997).
F R A G M E N T A T I O N OF VERY HIGH E N E R G Y HEAVY IONS
Dipartimento
M. G I O R G I N P di Fisica and INFN, Sezione di Bologna, V. It C. Berti Pichat 1-40127 Bologna, Italy E-mail: [email protected]
Radiation E-mail:
6/2,
S. M A N Z O O R 0 Division, PINSTECH, P. 0. Nilore, Islamabad, Pakistan, [email protected] , [email protected] Physics
A stack of CR39 ( C i 2 H i s 0 7 ) n nuclear track detectors with a Cu target was exposed to a 158 A GeV lead ion beam at the CERN-SPS, in order to study the fragmentation properties of lead nuclei. Measurements of the total, break-up and pick-up charge-changing cross sections of ultrarelativistic P b ions on Cu and CR39 targets are presented and discussed.
1
Introduction
We present experimental results on fragmentation charge-changing cross sections of 158 A GeV lead ions (charge Z = 82e) incident on Cu and CR39 targets. To detect and identify the relativistic ions, the nuclear track detector CR39 was used. When an ion crosses a nuclear track detector foil, it produces damages at the level of molecular bonds, forming the so called "latent track". During the chemical etching of the detector in a basic water solution, etch-pit cones are formed on both sides of the foil. The base area and the height of each cone are functions of the Restricted Energy Loss (REL) of the incident ion and thus of its charge Z1'2. 2
Experimental procedure
A stack made of CR39 nuclear track detectors with a Cu target was exposed in November 1996 at the CERN-SPS to a beam of 158 A GeV Pb ions. The exposure was performed at normal incidence. The total number of lead ions incident on the stack was about 7.8 x 104, distributed in 8 spots. The central density in each spot was around 1500 ions/cm 2 . The stack had the following composition: 12 CR39 sheets ~ 0.6 mm thick, a Cu target ~ 10 mm thick; 38 CR39 sheets ~ 0.6 mm thick. In the present °for the EMU18 Coll.: G. Giacomelli, M. Giorgini, H.A. Khan, G. Mandrioli, S. Manzoor, L. Patrizii, V. Popa, I.E. Qureshi, M.A. Rana, M. Sajid, P. Serra, M.I. Shahzad, G. Sher and V. Togo. 149
150
350 Z = 82
300 3
250
£
200
£
150
C O
Z = 80
6 Z
Z = 65 Z = 70
100
83
Z = 75
50
20
40
60
80
100
120
140
160
Cone height (arbitrary units) Figure 1: Cone height distribution for P b ions and heavy fragments measured on one face of the CR39 sheet immediately after the Cu target.
analysis, the CR39 sheets immediately before and after the Cu target and the last sheet of the stack were used. After exposure, the sheets were etched for 72 h in a 4N KOH water solution at a temperature of 45 °C. Previous calibrations of the detectors have shown that for high Z nuclei, the height of the etched cone is more sensitive to Z than its base area or diameter 3 . In order to separate the lead ions from the nuclear fragments with charge Z > 63e, we performed manual measurements of about 6300 cone heights using an optical Zeiss microscope with a magnification of 40 x. Fig. 1 shows the cone height distribution of Pb ions and heavy fragments measured on a single face of the CR39 sheet located after the Cu target. The charge resolution obtained is about 0.2e. 3
Total charge-changing cross sections
Using the survival fraction of lead ions for the Cu and CR39 targets, we measured the total charge-changing cross sections of lead ions using the formula: Nin AT (rtot = — Z T T l n TT~ (!) PTtrNA Nout where Ni„ and Nout are the numbers of lead ions before and after the target, respectively; NA is Avogadro's number; pr, AT,
151
mass and the thickness of the target. The data are indicated by the black points in Fig. 2, the uncertainties are statistical only.
Figure 2: Measured total fragmentation charge-changing cross sections otot of 158 A GeV P b projectiles versus the target mass number A?: the black points are our d a t a on Cu and CR39, the open points refer to d a t a obtained by a similar experiment 4 using the same beam. The solid line represents the fit of all the d a t a to formula (2) of ref. [4]. Results from a 10 A GeV Au beam incident on various targets I s - 7 ! are also shown.
As shown in Fig. 2, the data are in agreement with previous data obtained by a similar experiment using the same beam and different targets with atomic masses ranging from 4.7 a.m.u. (CH 2 ) to 207 a.m.u. (Pb) 4 . The solid line in Fig. 2 is the fit of all the data to formula (2) of ref. [4] which yields X2/D.o.F. — 0.7. Results from other experiments using a 10 A GeV Au beam incident on various targets t5_7J are also shown in Fig. 2. 4
Partial fragmentation charge-changing cross sections
The partial fragmentation charge-changing cross sections of Pb ions yielding fragments with charge 64e < Z < 82e were calculated for the Cu and CR39 targets using the formula 8 : aZ
—
AT
(9)
~ prtTNA N82 where Z = 64e -i- 81e, Nz is the number of fragment nuclei with charge Z produced in the target, 7^82 is the number of unfragmented beam nuclei and PT, AT, *T> NA have the same meaning as in Eq. (1). In this procedure, the successive fragmentation processes are neglected. The results for the partial fragmentation cross sections of incident lead ions on Cu and CR39 targets are shown versus AZ in Fig. 3.
152
i i i i i i i i i I i i i i I i i i i I i i i i I i i i i I i i i i I i i i i I i i i
: • • o-Cu
.
- O D
CTCR39
^
0
D
'n*
E
102
*5 « S . .
r-*
b
1
-20
5
• • • I • • • • I ' • • • I • • • • I • ' ' ' I • ' ' • I ' ' • ' I • • ' ' I '
-17.5
-15
-12.5
-10
-7.5
-5
-2.5
' '
0
AZ Figure 3: Partial fragmentation charge-changing cross sections for incident lead ions, uz versus AZ for the Cu and CR39 targets computed with Eq. (2). The black points refer to &z for Cu, the open points refer to oz for CR39. The squares refer to the pick-up cross sections. The errors are only statistical.
The square points in Fig. 3 refer to the charge pick-up cross sections, determined using Eq. (2) where Nz is the number of nuclei with Z = 83e produced in the target. Acknowledgments We thank the CERN SPS staff for the good performance of the exposure, the technical staff of the Bologna INFN and of the PINSTECH Laboratory. References 1. R. Fleischer, P.B. Price and R.M. Walker in Nuclear Tracks in Solids, (University of California Press, 1975). 2. S. Cecchini et al, Nuovo Cimento A 109, 1119 (1996). 3. G. Giacomelli et al., Nucl. lustrum. Methods A 411, 41 (1998). 4. H. Dekhissi et al, Nucl. Phys. A 662, 207 (2000). 5. S.E. Hirzebruch et al., Phys. Rev. C 5 1 , 2085 (1995). 6. L.Y. Geer et al., Phys. Rev. C 52, 334 (1995). 7. Y. He and P.B. Price, Z. Phys. A 348, 105 (1994). 8. D.P. Bhattacharyya et al., Nuovo Cimento C 16, 249 (1993).
Section II Liquid-Gas Phase Transitions in Nuclear Matter
Phase Transition in Finite Systems Philippe CHOMAZW, Veronique DUFLOT^ 1 - 2 ) and Francesca GULMINELLI^ 2 ) f1' G.A.N.I.L.(CEA-DSM/IN2P3-CNRS), BP 5027, U076 Caen cedex 5, France LPC Caen, (IN2P3-CNRS/ISMRA et Universite), F-14050 Caen cedex, France
(2)
In this paper we present a review of selected aspects of Phase transitions in finite systems applied in particular to the liquid-gas phase transition in nuclei.
1
INTRODUCTION
Phase transitions are universal properties of matter in interaction. They have been widely studied in the thermodynamical limit of infinite systems. However, in many physical situations this limit cannot be accessed and so phase transitions should be reconsidered from a more general point of view. This is for example the case of matter under long range forces like gravitation. Even if these self gravitating systems are very large they cannot be considered as infinite because of the non saturating nature of the force. Other cases are provided by microscopic or mesoscopic systems Metallic clusters can melt before being vaporized. Quantum fluid may undergo Bose condensation or superfluid phase transition. Dense hadronic matter should merge in a quark and gluon plasma phase while nuclei are expected to exhibit a liquid -gas phase transition. For all these systems the theoretical and experimental issue is how to sign a possible phase transition in a finite system. In this contribution we show that the problem of the non existence of boundary conditions can be solved by introducing a statistical ensemble with an averaged constrained volume. In such an ensemble the microcanonical heat capacity becomes negative in the transition region. We show that the caloric curve explicitly depends on the considered transformation of the volume with the excitation energy and so does not bear direct informations on the characteristics of the phase transition. Conversely, partial energy fluctuations are demonstrated to be a direct measure of the equation of state. Since the heat capacity has a negative branch in the phase transition region, the presence of abnormally large kinetic energy fluctuations is a signal of the liquid gas phase transition. 155
156
2 2.1
MODELS FOR P H A S E T R A N S I T I O N IN FINITE SYSTEMS Lattice-gas model
Let us first consider an exactly solvable model for second and first order phase transitions, the Lattice Gas Model of Lee and Yang 1 . This is a simplified model which can be interpreted as a schematic representation of a classical fluid with a Van der Waals type of equation of state. In our numerical implementation the TV sites of a lattice are characterized by an occupation number r = 0 or 1 (see Fig. 1). Particles occupying nearest neighboring sites interact with an energy e. A kinetic energy term is also included so that the Hamiltonian is given by:
i=i
ij
where the second sum runs only over neighboring sites. In the liquid-gas phase transition, since the order parameter is the density difference between the two phases, the volume is essential in determining thermodynamical properties. Many studies have been performed considering periodic boundary conditions in order to avoid the effects of the surface 2>3>4. Systems in a fixed cubic volume have also been investigated 5,6 . In the experimental case however the volume is not defined through boundary conditions because we are dealing with an open system. However, a typical average size of the fragmenting system might be deduced from experimental observables. For example the (average) radius of a hot source can be defined through interferometry or through comparisons with statistical models. From a theoretical point of view this implies that at equilibrium the entropy of the system should be maximized under the constraint of a specific value for the average volume. In the absence of a preferred direction, an average volume can be defined through the one-body observable 7,8 :
? = &£*'
<2>
i=i
where r; is the distance to the center of the lattice. Introducing first a canonical description in which the energy observable H as well as the volume V are known in average we have to introduce the associated Lagrange multipliers /?, A so that the partition function reads:
ZPiX = J2 ex P ( - / ^ ( n ) - A^ (n) ) (»)
(3)
157
Here, U^ and lA") are the expectation values of the operators H and V in the nth event, (3 is the inverse of the canonical t e m p e r a t u r e (3 = l/Tcan, and the quantity P — —A//? has the dimension of a pressure. In other words, the experimental fact t h a t the volume is known only in average means t h a t the pressure, interpreted as the Lagrange multiplier associated with the volume observable, can be considered as the relevant state variable. A statistical ensemble of events associated with an average volume can easily be generated through a canonical sampling using a constrained Hamiltonian E— H—P V where P V can be considered as a constraining one-body external field. T h e averaged constrained energy E — U — PV can be interpreted as an enthalpy. The actual value of the pressure parameter must be defined to get the desired average volume. For finite systems the various ensembles are not equivalent because fluctuations cannot be neglected. Since energy is a directly accessible observable in each event, t h e correct statistical framework is the microcanonical ensemble. An easy way to access microcanonical quantities is to sort the canonical partitions according to their total energy. It is i m p o r t a n t to notice t h a t this procedure m a y be in some cases numerically time consuming but it is an exact method to generate constant energy events with the correct microcanonical weight (see eq.(4) below). W h e n b o t h energy and volume are considered as thermodynamical observables one should in principle sort the canonical events sampled for a given t e m p e r a t u r e and pressure as a function of the energy and the volume. However, if the pressure is fixed it is sufficient to sort the events as a function of the constrained energy E. At a given t e m p e r a t u r e (3 the canonical distribution reads: Pp(E)
= ^±exp(-/3E)
(4)
where W is the degeneracy of the state. In a sampling of NQ events the probability Pp can be estimated from the number Np of events falling in the enthalpy bin of size AE around E, Pp [E) AE « Np {E) /N0. Equation (4) can be inverted leading to the entropy S (E) = log(W (E)). This allows a direct estimation of the microcanonical caloric curve
r^tm-t+'JsgiS.
(5)
which is valid for every /3. W i t h a single microcanonical sampling at an arbitrary (3 it is in principle possible to directly compute the whole microcanonical caloric curve (5) without any approximation. In the calculations shown below a number A — 216 of particles is fixed; to illustrate a first order phase transition the pressure P is chosen in such a
158 way t h a t the isobar crosses the canonical coexistence line at about the half of the critical t e m p e r a t u r e . Canonical statistical averages are taken over events obtained with a s t a n d a r d Metropolis sampling of the lattice occupations according to the partition function (3). It is interesting to compare the microcanonical and canonical caloric curves as shown in Fig. 1. In finite systems two canonical caloric curves can be defined, corresponding to the average and most probable energy associated to a given j3. In infinite systems these two energies are equal because fluctuations can be neglected. Far from the phase transition the canonical and microcanonical curves agree. Indeed, from equation (5) we can see t h a t the most probable canonical energy is characterized by the equality of the microcanonical and canonical temperatures. In the coexistence region however the predictions of the two ensembles differ in an noticeable manner. T h e canonical caloric curves are by definition monovaluated while this restriction does not apply to the microcanonical case. T h e microcanonical caloric curve presents a back bending while in the back-bending region the canonical caloric curve associated with the most probable energy presents a discontinuity equivalent to the Maxwell construction. T h e observed energy j u m p is directly related to the latent heat of the first order phase transition. Allowing a fluctuating volume is essential
Figure 1: Right: schematic drawing of the lattice-gas model. Left: Microcanonical caloric curve (full line) compared with the most probable (dots) and average (line) canonical energy for each /3. to obtain the caloric curves of Fig. 1: constant volume lattice gas calculations produce smoothly increasing caloric curves 9 even within the microcanonical ensemble as we will discuss later. It should be noticed t h a t the partitions which fall in the energy region corresponding to the canonical t e m p e r a t u r e j u m p are hardly sampled by the canonical ensemble, b u t are accessible in the micro-
159
canonical ensemble. Therefore, in a finite system the microcanonical sorting of events allows to study regions of the phase diagram which are forbidden in the canonical formalism. These regions are characterized by specific properties such as negative heat capacities which we will later on study in more detail. In particular it is i m p o r t a n t to identify experimental observables which can directly inform us about these peculiar properties. 2.2
Analytical
quantum
models
In the lattice-gas model the connection between a phase transition and a back bending in the caloric curve appears evident. However, one may worry about the generality of such a statement. Is it a general property found in m a n y systems? Do such anomalies also exist in q u a n t u m systems?
40 20
a^fe
10
•K.
V J\ £.AA.
•Iff
» f i 6&&4AA&.&
20
30
40
H=e,(iS|-A12+A2)+e,N2
SO
Energy Figure 2: Right: schematic drawing of the double quantum oscillator model with the associated Hamiltonian. Left: entropy, temperature and heat capacity as a function of the excitation energy. In order t o address this question we have investigated a model of A particles which can j u m p from one harmonic oscillator to an other. In the first one all particles strongly interact while in the second one they are free. T h e curvature
160
of the second well plays the role of a confining potential i. e. of a pressure. The corresponding Hamiltonian reads: H = s1 (jVi -A\+
A2) + e2Ni
(6)
with the operators A
A
s a a
Ni = J2 L i n
Ai = Y,5l
n=l
(7)
n=l
where in is the harmonic well occupied by the particle i. Using this Hamiltonian we can compute the level density and so the entropy. To simplify the calculation we have chosen S\ and 62 to be commensurable. Then, we can compute the temperature and the associated heat capacity (see Fig. 2). We observe that the system indeed presents an anomaly in the curvature of the entropy. Back-bending and negative heat capacities automatically follow. 3
ROLE OF T H E V O L U M E
As far as the liquid-gas phase transition is concerned it is essential to discussed the role of the volume since it is the order parameter and since we know that the divergency of the heat capacity depends upon the isobar or isochore character of the considered transformation. 3.1
Caloric curves and abnormal fluctuations
The normalized fluctuations < T | / 7 \ 2 obtained in the microcanonical ensemble with a constrained average volume,
161 t e m p e r a t u r e as a function of energy at a constant pressure or a constant average volume in the subcritical region is displayed in the upper part of Fig. 3. At
*-WSM
l(A,HeV)
°
^iA-MeV).
Figure 3: Left: Isotherms and contour plot of the kinetic energy fluctuations in the X — E plane. The level corresponding to the canonical expectation
162
sent a strong m a x i m u m which exceeds the canonical value: an anomalously large fluctuation signal will be always seen if the system undergoes a first order phase transition, independent of the p a t h . As an example the lower part of Fig. 3 shows a constant P\ or < V >\ cut of the bidimensional fluctuation surface. The quantitative behavior of the heat capacity as a function of energy depends on the specific transformation, but at each point the heat capacity extracted from fluctuations is a direct measure of the underlying equation of state. This is clearly demonstrated in the m e d i u m part of Fig. 3 which compares the exact heat capacity C\ with the fluctuation approximation. The agreement between the two results illustrates the accuracy of the estimation through fluctuations11. 3.2
Constant
volume transformation
and negative
compressibility
We now turn to the constant volume ensembles i.e. ensembles defined through the presence of sharp boundary conditions. As expected, for small boxes, this ensemble does not presents anomalies in the equation of states. Indeed, the heat capacity at constant volume is not expected to diverge in the thermodynamical limit. Only the infinite volume solution which can also be thought as a zero pressure system presents the back bending already discussed in the context of open systems (see Fig. 4 ) 1 2 .
Figure 4: Left: caloric curve and associated fluctuations in a constant volume ensemble at two different values of the the box volume (higher for the solid line). Right: isotherms of a canonical ensemble at constant volume in the pressure versus density plane. T h e absence of a back bending in the caloric curve does not mean t h a t the phase transition is not present. Indeed, varying the volume one can see
163
t h a t the thermodynamical potential has a convexity anomaly. In such a case the compressibility is negative. This clearly shows t h a t the volume is the order parameter. Therefore controlling the order parameter allows to explore the phase transition region and to spot the characteristic back bending in the associated conjugated variable, here the pressure. This is illustrated in the case of the canonical ensemble at constant volume in the right part of Fig. 4. One can clearly see t h a t the isotherms present a strong back bending below the critical t e m p e r a t u r e . It should be noticed t h a t this demonstrates t h a t the existence of back-bending's is not limited t o the microcanonical ensemble. In fact the best ensemble to spot a first order phase transition is the one in which the order parameter is directly constrained. Then a back bending should be seen in the associated EOS i.e. the equation giving the conjugated variable (the second derivative of the thermodynamical potential) as a function of the order parameter no m a t t e r the variables chosen to control the other t h e r m o d y n a m i c a l degrees of freedom. Conversely, if the order parameter is controlled only in average its conjugated variable appears as a Lagrange multiplier controlling the statistical ensemble. T h e n the order parameter EOS does not back-bends any more since only one average order parameter is allowed for a given Lagrange multiplier. However, if another intensive variable is controlled and happens to be different in the two phases (for example here the energy) then it m a y play the role of an alternative order parameter and the associated E O S (here the caloric curve) m a y present an anomaly.
4 4-1
A D D I T I O N A L SIGNALS OF A P H A S E T R A N S I T I O N Critical mass
partitions
In order to better characterize t h e back-bending region it is interesting t o examine the properties of the associated partitions. It should be noticed t h a t since the back bending region of a microcanonical calculation corresponds t o events with a very small probability in a canonical ensemble one may expect t h a t the microcanonical sampling can be different from the typical partitions of the corresponding canonical distribution. Fig. 5 presents canonical and microcanonical fragment mass distributions in the phase transition region 8 . Let us first start with the microcanonical ensemble, at the lowest energy, in the first uprising branch of the caloric curve, the fragment distribution shows an exponential fall off of light fragments associated with the gas phase and a big liquid drop, characteristic of a subcritical system. At the highest energy corresponding t o t h e beginning of the gas branch of the caloric curve, the distribution resembles to a typical supercritical vaporized system. In the middle of the back-bending region a critical distribution is
164
observed. This observation of a critical behavior of the fragment size distribution in the coexistence region has also been reported in the canonical lattice-gas context 2 (shown in the Fig. 5). Critical behaviors are observed irrespectively of the considered density. Therefore, a line of critical points can be identified in the whole phase diagram, from a percolation type at high density toward the thermodynamical critical point and down to the interior of the coexistence region. The microcanonical study additionally suggests that anomalous fluctuations coming from the negative heat capacity characteristic of a microcanonical first order phase transition, are to be expected in the region where the fragment size distribution becomes critical. The conjunction of these two observations signs a first order phase transition.
_ Ui|Ulil
| 2
So.t
3e ? 2J.3JJB
GZib-
V -U |
10 100 I
,\
10 1UU1 Mass
2: s
\1G,
It]
Vv
4": %
•Is* l«»
'«
fee
\ '»
jf
A
e 3,s •g3.49 £. 3.4•-
«NS»
\vutt
1^^
0.5 ^
3.3 3.25 3.2
l4//?(A.^eV)
ox
0.4 0.4, OS
Figure 5: Left: microcanonical fragment size distribution at three different energies presented on top of the associated caloric curve. Right: canonical fragment distribution for three density regions. Subcritical, critical and supercritical distributions are shown. The corresponding critical temperature for each density is reported in the phase diagram below.
4-2
Chemical signals
Since we are dealing with a liquid gas phase transition in a two fluid system one also expects the presence of a distillation (or fractionation) of the isospin asymmetric nuclear matter. This was stressed first in Ref. 13 and then illustrated in isospin dependent lattice-gas model 4 . This result showing a strong neutron enrichment of the gas phase (light fragments) is reported in Fig. 6. This may provide an additional signal of the phase transition. Finally we would like to
165 stress t h a t a back-bending should also be expected in the chemical potential curve as a function of the number of constituents. A reflection about chemical fluctuation in presence of a negative curvature of the thermodynamical potentials when the number of constituents is modified should also be considered. Indeed, one m a y think t h a t if two parts of the system exchange particles the anomalous curvature will lead to anomalously large fluctuations. This should be further studied but m a y help to spot first order phase transition in many situations where it is also associated with chemical equilibria.
.2 JO2F 0
Figure 6: Left: t/He isotopic ratio calculated for an asymmetric matter at various temperature (full line) compared to the combinatorial expectation (thin line). Right: evolution of the canonical chemical potential as a function of the number of particles.
CONCLUSIONS In this contribution we have shown t h a t convexity anomalies of the thermodynamical potentials can be used to signal and to define phase transitions. We have presented t h e first experimental evidences of such a phenomenon. We have discussed in details the role of t h e volume in the liquid-gas phase transitions and explained why caloric curves are not unique since they depend upon the volume of the system at the various considered energies. Conversely we have stressed t h a t fluctuations are state properties which can be used to infer the t h e r m o d y n a m i c s of the considered ensemble of events. Finally we have stress t h a t t h e fragmentation p a t t e r n s as well as the chemical properties can be used as complementary information to control t h e existence of a phase transition.
166
References 1. 2. 3. 4. 5. 6. 7. 8. 9. 10. 11. 12. 13.
C. N. Yang and T . D. Lee, Phys.Rev. 87 (1952) 410. F . Gulminelli and P. Chomaz, Phys. Rev. Lett. 82 (1999) 1402. J. M. C a r m o n a e t al., Nucl. Phys. A 643 (1998) 115. P. Chomaz and F. Gulminelli, Phys. Lett. B447 (1999) 221. X. C a m p i and H. Krivine, Nucl. Phys. A 620 (1997) 46 and Physica A, in press. J. P a n and S. Das G u p t a , Phys.Rev. C 51 (1995) 1384; Phys.Rev.Lett. 80 (1998)1182. V. Duflot et al., Phys.Lett.B 476 (2000) 279. F . Gulminelli et al., Europhys. Lett., 50 (2000)434. J. M. C a r m o n a e t al., nucl-th/9910061 and J. Borg et al., Phys. Lett. B 470 (1999) 13. P. Chomaz, F . Gulminelli and V. Duflot, Phys. Rev. Lett. (2000) in press. P. Chomaz, contribution to this conference, eq.(6). F . Gulminelli et al, cont. to the XXXVIII Winter meeting on nucl. Phys. Bormio (Italy), Ed. I. Iori. H. Mueller and B. D. Serot, Phys.Rev. C 52 (1995) 2072.
PRESENT STATUS AND FUTURE PROSPECTS OF INVESTIGATIONS OF THE LIQUID-GAS PHASE TRANSITION
CLAUS-KONRAD GELBKE National Superconducting Cyclotron Laboratory and Department of Physics and Michigan State University, East Lansing, MI 48824-1321, USA
Astronomy,
Nuclear systems formed in central heavy-ion collisions at intermediate energy (20 MeV < E/A < 400 MeV) expand and disintegrate into a mixture of nucleons and light nuclei. Qualitatively, such decays are expected when nuclear matter at subnormal density undergoes a liquid-gas phase transition. Selected experimental results about multi-fragment decays will be discussed, and problems that need further theoretical and experimental work will be pointed out.
1
Introduction
The atomic nucleus is a mesoscopic system that can be viewed as a finite drop of a multi-component quantum Fermi liquid with strong interactions between the constituents. Bulk nuclear matter is likely to exist in neutron stars. Experimental attempts of determining the equation of state of nuclear matter and studying its predicted phase transitions can only be made by means of nucleusnucleus collisions where finite systems of nuclear matter may be heated up and compressed temporarily. Low densities, pertinent to an expected liquid gas phase transition, may be reached during a subsequent expansion and decompression stage. The extremely high densities and temperatures, pertinent to a possible deconfinement transition to a quark-gluon plasma, can only be reached in collisions at ultra-relativistic energies. Since nucleus-nucleus collisions produce transient systems of very short lifetime, full statistical equilibrium of all degree of freedom is unlikely to be reached. Hence, the extraction of thermodynamic properties of bulk nuclear matter presents a significant challenge. Since the limit of a thermodynamic freeze-out is well defined and understood, it is important to establish, for which (if any!) degrees of freedom statistical equilibrium is attained, where (local) equilibrium concepts become inappropriate, and which observables are sensitive to temperature, pressure, density, size, and shape of the reaction zone formed during the collisions. In this talk, I will discuss a number of experimental approaches aimed at elucidating the thermodynamics of the liquid-gas phase transition in nuclear matter and at extracting information about its equation of state (EOS). I will restrict
167
168
myself to intermediate energies (20 MeV < E/A < 400 MeV) where multifragment disintegration, pertinent to the liquid-gas phase transition, is well-established [1], 2 Measuring Temperatures Extractions of nuclear temperature are generally based upon comparisons of experimentally observed distributions to those predicted by statistical theory [2]. In the most naive approach, one can determine the slopes of the kinetic energy spectra of emitted particles and extract "kinetic temperatures", T ^ , from fits with Maxwell distributions. These temperatures typically characterize the early phase of the reaction before the system expands and cools. Complications arise from preequilibrium emission and collective flow effects, which distort the shapes of the energy spectra. Alternative approaches, based upon the assumption of chemical equilibrium at freeze-out, extract apparent temperatures, TAE and T^, from relative populations of states or ratios of isotopic yields, respectively [2,3]. These temperature determinations can be complicated by feeding from the sequential decay 100
°
T
JP> • TJ^ 1 )
T
5
Ne+Pb
from i n t e r m e d i a t e / ^ I™
s^
it CentralAu+Au collisions T conected for radial flow
10 -
10
100
1000
(E-VcyA (MeV) of excited states. Figure 1: Energy dependence of kinetic temperatures extracted in various reactions indicated in the figure. The open points show mid-rapidity data from inclusive experiments; solid squares and diamonds show data from central Au+Au collisions that are corrected for collective radial flow, filled circles are temperatures, T« E, extracted from the relative populations of widely separated states in 5Li.
Figure 1 illustrates the long-standing "temperature dilemma". For intermediate energy nucleus-nucleus collisions, measured values of T ^ are typically higher than
169
measured values of TAE. Qualitatively, such a difference can be attributed to the expansion and cooling of the initially compressed nuclear system, but a quantitative analysis of the collective expansion effects could not resolve this dilemma in a consistent thermodynamic picture [4]. Much excitement has been generated by the extraction of a "caloric curve" suggestive of a phase transition similar to that of boiling water [5]. Unfortunately, the analogy to boiling water is superficial at best. For boiling water, the pressure stays constant while heat is added to the system; but for an expanding nuclear system, the pressure varies while no heat is added, i.e. the nuclear expansion is nearly isentropic rather than isobaric. In addition, significant inconsistencies have emerged [6,7] in the cross-calibration of the helium-lithium isotope ratio thermometer (THe-ii) employed in ref. [5]. Temperatures extracted from isotope ratios involving 3He and 4He nuclei differ from those extracted from other isotope ratios or from relative populations of widely separated states, see Figure 2. These differences are not yet understood. Some, but not all differences may arise from sequential decays, but additional complications could arise from a sequential freezeout of different degrees of freedom, from in-medium modifications, or from a strong pre-equilibrium component in the 3He emission [7]. At this time, the reliable determination of the caloric curve for hot nuclear matter remains a significant 10
" 1
Kr + N b : b / b 9.0
1 OTJH^Li) : O TJC-Li)
J I ' I I
|
I I
<0.45 max
{
8.0
> •
o
7.0
:
• T 4 /Be)
-
* T4/He) ;
6.0 5.0
4.0 3.0 20
40
60
80
100
E/A (MeV)
120
140
170
challenge, and the validity of the caloric curve of ref. [5] is questionable. Figure populations of states and from isotope ratios for central re 2: Temperatures extracted from the relative r< Kr+Nb Vb collisions collisions at at various various beam beam energies. energies. From Fr ref. [7].
3
Observing Radial Expansion
Evidence for an expansion of the reaction zone comes from two observations: the detailed shapes of energy spectra [8], especially those of light nuclei, and the multiplicity of intermediate mass fragments [9]. Early work [9] showed that standard compound nucleus evaporation models significantly underpredict the observed IMF multiplicities and that better agreement with observations was be obtained by assuming statistical emission from a source of subnormal nuclear density. Both the rate equation approach of the expanding evaporating source (EES) model [11] and microcanonical treatments [12,13] of a low-density freeze-out scenario have met with considerable success in reproducing many features observed for multifragmenting systems. Illuminating discussions of the similarities and differences of the various models can be found in refs. [14,15]. 0.4
A u + Au, central collisions v
(r) = r v It :
[ J, J.
0.3
%\ A
o
f
\
£ 0.2 >
ft
V 0.1
4+ 20
\
100
200
a o • •
EOSTPC FOPI Miniball Kr+Au Miniball 1000
E/A (Me V) Figure 3: Systematics of collective radial expansion velocities observed in Au+Au collisions.
The energy spectra of emitted particles are influenced by two major effects: a mass-independent thermal energy component (E,n «: kT) and a mass-dependent collective component (Ecoll <*= mv2cou). The superposition of these two effects produces characteristic shoulder-arm shaped energy spectra [8,10] with a strong mass-dependence. Many detailed investigations [4] have established the systematics
171
of radial collective flow that characterizes the expansion of the hot reaction zone, see Fig. 3. Yet, an appropriate quantum treatment of nucleon and cluster emission during the full compression-decompression cycle in a nuclear collision remains a challenge to theory. 4 Measuring Space-Time Characteristics Information about the space-time characteristics of the emitting source can be obtained from two-particle correlation functions, l+R(q,P), defined as the ratio of the two-particle coincidence yield and an appropriately chosen "background" yield. Here, q and P denote the relative and total momentum vector of the particle pair. The shape of the correlation function depends on the final state interactions and quantum statistics of the coincident particles and is thus sensitive to the spatial separation of the emitted particles, which in turn depends on the source size and the — • — i
;
•
i
- • — i —
36
"
i
•
I
'—r
'
i
P
= 200 - 400 MeV/c .
tot,cm
b/b
™ = °- 3 "
:
l\
+ 1
'•
45 Ar + Sc/ E / A = 160 MeV ;
^H&r30%
I'd
:
0.5
i
V \
1.5
DC
•
BUU / ~ X
^*>
* BUU + delayed component t = 160 fm/c
7 f
:*/ ./
i
.
i
20
.
i
.
i
40 q (MeV/c)
60
80
emission time scale [16-19]. Figure 4: Two-proton correlation functions for central Ar+ Sc collisions at E/A = 160 MeV. The points represent data, the solid curve shows BUU prediction, and the dashed curves show BUU calculations in which 20% and 30% of the proton emission is artificially delayed by 160 fm/c. From ref. [25].
Longitudinal (q 11 P) and transverse (q J. P) two-proton correlation functions measured for central 36Ar+45Sc collisions at E/A = 80 MeV [20] provided evidence for a fast emission time scale (T = 25 ± 15 fm/c) and were quantitatively reproduced by calculations with the Boltzmann-Uhling-Uhlenbeck (BUU) transport model [21,22]. Despite this impressive success, problems emerged for collisions at higher
172
energy [23-25], see, e.g. Fig. 4. The measured two-proton correlation functions are suppressed as compared to BUU predictions, indicating that the BUU model predicts proton emission from too small a reaction zone, on too short a time scale, or both, see Fig. 4. At the quantitative level, this failure of BUU is not understood. Qualitatively, it has been suggested that some fraction of the emitted protons may result from delayed secondary decays of highly excited fragments. The dashed curves in Fig. 4 show that a 20-30% sequential decay contribution, delayed by some 160 fm/c, might account for the discrepancy between BUU prediction and experiment. While there is firm experimental evidence that sequential decays play an important role for nucleus-nucleus collisions in this energy domain, both magnitude and time scale of sequential decays are not yet sufficiently well understood to allow reliable predictions. Note that sequential decays also affect the emission temperatures extracted from ratios of isotopes and/or widely separated states [26,27]. The emission of highly excited clusters and their sequential decay remain "I # 2
|—•—|—•—'—I—•—•—*—•—|—•—|—|—•—i—•—•—•—•—i—|—•—*—•—I—•—•—•—••
rr +
0
10
20
30
40
50
60
important problems. Figure 5: Two-fragment correlation fiinctions for central Ar + Au collisions at E/A = 50 MeV without (solid points, thin curves) and with (open points, thick curves) directional cuts. The reduced velocity is defined as v,* = v«,(Z, +Z2)~m [30]. From ref. [28].
173
An example of two-fragment correlation functions, measured [28] for central Ar + 197Au collisions at E/A = 50 MeV, is shown in Figure 5. The data are quantitatively reproduced by many-body Coulomb trajectory calculations indicating fragments emission during a short time interval, T < 100 fm/c. As fragments are emitted with low initial velocity, they will not move far during this time. Hence, the difference between instantaneous and sequential multi-fragment emission becomes rather blurred and difficult to resolve experimentally [29]. 5 Measuring Pressure 36
Reaction zones formed in energetic nucleus-nucleus collisions, E/A > 100 MeV, may reach densities significantly higher than the saturation density of nuclear matter in its ground state. This is the only terrestrial situation in which such densities can be achieved and investigated experimentally. During compression and heating, the pressure increases in the reaction zone. Eventually, an expansion to sub-normal density sets in. The ejection of nucleons and clusters is influenced both by the repulsive pressure and by the attractive mean field. At low energies, the attractive mean field dominates, and particles are preferentially emitted to negative deflection angles. At higher energies, the pressure dominates, and deflection is largely to positive deflection angles. The expected change of sign has been measured by detecting the circular polarization of gamma rays from target residues in coincidence with charged particles emitted from the reaction zone [31]. The most commonly investigated pressure-related observable is the sidewarddirected flow, F, of nuclear matter in the reaction plane [32], defined as the derivative of the average transverse momentum per nucleon in the reaction plane with respect to the normalized cm. rapidity y„: F = d
174
3001
1
1
1
0
2
4
6
r
8
10
13
A Figure 6: Mass dependence or sideward-directed flow for two cuts on impact parameter b. Here, flow was determined only for particles of fixed mass number A. From ref. [33],
This flow depends strongly on impact parameter and incident energy. Due to the interplay of the collective flow velocity (which is independent of particle mass) and the randomly oriented thermal velocity (which decreases with particle mass), flow is typically larger for heavier than for lighter emitted particles [33,34], see Fig. 6. This strong mass dependence of flow complicates quantitative comparisons with models that do not treat fragment formation. As a remedy, one typically constructs the "effective proton flow" by performing a charge-weighted average of the flow measured for all detected particles. The impact parameter dependence is extracted by using an empirical impact parameter filter. Figure 7 gives an example of such a comparison between experiment and theory [33]. The data clearly favor a momentum dependent mean field, consistent with the nonlocality effects observed in nucleon-nucleus potential scattering [35], and in-medium cross sections smaller than the free nucleon-nucleon cross sections, but there is little sensitivity to the stiffness of the EOS. Similar conclusions were drawn from investigations [36] of the "balance energy", defined as the energy for which the average sideward directed flow vanishes due to the cancellation of the average mean-field attraction by the average repulsion from nucleon-nucleon scattering.
175
,
0
•
,
,
I
2
,
•
.
•
I
4
•
•
.
.
I
.
•
6
b (fm) Figure 7: Impact-parameter dependence of sideward-directed flow. The crosshatched rectangles depict the effective proton flow. BUU transport calculations are shown for the following parameter sets: H - hard EOS without momentum dependence (open circles) ; S - soft EOS without momentum dependence (open squares); HM - hard EOS with momentum dependence (solid diamonds); SM - soft EOS with momentum dependence (open diamonds); SM (0.8cf) - soft EOS with momentum dependence and a 20% reduction in the nucleonnucleon cross section (solid circles). Fromref. [33].
As the quality of experimental data of increases, more detailed investigations of flow observables become possible. Of particular interest are investigations of azimuthal distributions of particles emitted at mid-rapidity [37,38] - or, more specifically, of the elliptic flow variables vj =
176
quantitative description of the initial population of high-lying particle-unbound states and their sequential decay. With regard to the liquid-gas phase transition, the major challenge remains the development of a tractable quantum transport model that allows a reliable treatment of density fluctuations and of fragment formation. Such a dynamic theory might eventually overcome the limitations encountered in thermal freeze-out scenarios. Substantial progress has been made in the understanding of collective flow and its sensitivity to the EOS of excited nuclear matter. At lower energies, the need for a momentum dependent mean field has been credibly established, and there are indications that the in-medium nucleon-nucleon cross-sections are smaller than the free nucleon-nucleon cross sections. There is great hope that accurate measurements of the energy dependence of elliptic flow will have an increased sensitivity to the EOS. Additional progress, relevant for the extrapolation to the highly asymmetric nuclear matter encountered in neutron stars, can be expected by investigating reactions with nuclei of very different N/Z ratios, but fixed total mass. Such reactions can provide unique information about the symmetry-energy term of the EOS about which little is known at this time. The needed investigations can be performed with energetic beams of rare isotopes produced by projectile fragmentation at the next generation RIA (Rare Isotope Accelerator) facility currently discussed by the nuclear science community. This work was supported by the National Science Foundation under Grant No. PHY-95-28844. References: 1. G.F. Peaslee et al., Phys. Rev. C49, R2271 (1994). 2. D. Morrissey et al., Annu. Rev. Nucl. Part. Sci. 44, 27 (1994) and references given there. 3. S. Albergo et al., Nuovo Cimento 89,1 (1985). 4. W. Reisdorf and H.G. Ritter, Annu. Rev. Nucl. Part. Sci. 47, 663 (1997) and references given there. 5. Pochodzalla et al., Phys. Rev. Lett. 75, 1040 (1995) 6. V. Serfling et al., Phys. Rev. Lett. 80, 3928 (1998). 7. H. Xi et al., Phys. Rev. C58, R2636 (1998). 8. P.J. Siemens and J.O. Rasmussen, Phys. Rev. Lett. 42, 880 (1979). 9. D.R. Bowman et al., Phys. Rev. Lett. 67, 1527 (1991). 10. W.C. Hsi, Phys. Rev. Lett. 73, 3367 (1994).
177
11. W.A. Friedman, Phys. Rev. Lett. 69, 2135 (1988). 12. D.H.E. Gross and Xiao-ze Zhang, Phys. Lett. B161, 47 (1985); D.H.E. Gross, Rep. Prog. Phys. 53, 605 (1990); D.H.E. Gross et al., Ann. Phys. 1, (1992) 467. 13. J.P. Bondorf et al., Nucl. Phys. A443, 321 (1985); Nucl. Phys. A444, 460 (1985); Phys. Reports 257, 133 (1995). 14. W.A. Friedman, Phys. Rev. C40, 2055 (1989). 15. D.H.E. Gross and K. Sneppen, Nucl. Phys. A567, 317 (1994). 16. S.E. Koonin, Phys. Lett. B70,43 (1977). 17. S. Pratt and M.B. Tsang, Phys. Rev. C36, 2390 (1987). 18. D.H. Boal et al., Rev. Mod. Phys. 62, 553 (1990). 19. W. Bauer et al., Annu. Rev. Nucl. Part. Sci. 42, 77 (1992). 20. M.A. Lisa et al., Phys. Rev. Lett. 71, 2863 (1993). 21. M.A. Lisa et al., Phys. Rev. Lett. 70, 3709 (1993). 22. D.O. Handzy et al., Phys. Rev. C50, 2858 (1994). 23. G.J. Kunde et al., Phys. Rev. Lett. 70, 2545 (1993). 24. S.J. Gaff et al., Phys. Rev. C52, 2782 (1995). 25. D.O. Handzy et al., Phys. Rev. Lett. 75, 2916 (1995). 26. Z. Chen and C.K. Gelbke, Phys. Rev. C38, 2630 (1988). 27. M.J. Huang et al., Phys. Rev. Lett. 78, 1648 (1997), and references given there. 28. T. Glasmacher et al., Phys. Rev. C50, 952 (1994). 29. R. Popescu et al., Phys. Rev. C58, 270 (1998). 30. Y.D. Kim et al., Phys. Rev. C45, 387 (1992). 31. M.B. Tsang et al., Phys. Rev. Lett. 57, 559 (1986); Phys. Rev. Lett. 60 (1988) 1479; R.C. Lemmon et al., Phys. Lett. B446, 197 (1999). 32. P. Danielewicz and G. Odyniec. Phys. Lett. B157, 146 (1985). 33. M.J. Huang et al., Phys. Rev. Lett. 77, 3739 (1996). 34. F. Rami et al., Nucl. Phys. A646, 367 (1999). 35. C. Gale et al., Phys. Rev. C35, 1545 (1990). 36. R. Pak et al., Phys. Rev. C54, 2457 (1996); Phys. Rev. Lett. 78, 1026 (1997). 37. P. Danielewicz et al., Phys. Rev. Lett. 81, 2438 (1998). 38. Bao-An Li and A.T. Sustich, Phys. Rev. Lett. 82, 5004 (1999); Yu-Ming Zheng et al., Phys. Rev. Lett. 83, 2534 (1999).
CRITICAL P H E N O M E N A IN F I N I T E SYSTEMS
1
A. BONASERA 1 , T. MARUYAMA 1 - 2 and S. CHIBA 2 Laboratorio Nazionale del Sud, Istituto Nazionale di Fisica Nucleare, V.S.Sofia 44, 95123 Catania-Italy. 2 Advanced Science Research Center, Japan Atomic Energy Research Institute, Tokai, Ibaraki, 319-1195, Japan. We discuss the dynamics of finite systems within molecular dynamics models. Signatures of a critical behavior are analyzed and compared to experimental data both in nucleus-nucleus and metallic cluster collisions. We suggest the possibility to explore the instability region via tunneling. In this way we can obtain fragments at very low temperatures and densities. We call these fragments quantum drops.
1
Introduction
The dynamics of a fragmenting finite system can be well modeled by Classical Molecular Dynamics (CMD). Of course in such studies we are not interested in reproducing the data coming from nucleus-nucleus, cluster-cluster or fullerenefullerene collisions, which are strongly influenced by quantum features, but simply in the possibility that a finite (un)charged system "remembers" of a liquid to gas phase transition which occurs in the infinite case limit J . Classical particles interacting through a short range repulsive interaction and a longer range attractive one have an Equation of State (EOS) which resembles a Van Der Waals (VDW) l ' 2 . It is also well known that the Nuclear EOS resembles a VDW as well 3 , thus classical studies of the instability region are quite justified in order to understand the finite size plus Coulomb effects. Starting from a two body force, for instance of the Yukawa type, the EOS can be easily calculated within CMD using the virial theorem. This was done in ref.1 and the resulting EOS spinodal line is plotted in Fig. 1 (full line) in the temperature T density p plane. These quantities have been normalized using typical values of the system. When dealing with phase transitions it is usual to normalize the various quantities by their values at the critical point of a second order phase transitions. Since later on, we will compare with some results on finite systems, where the values at the critical point (if it exist) are not known, we normalize by typical values in the ground state of the system such as density po, or the absolute value of the binding energy E\>. In such a way we can compare results coming from different systems such as nuclei or clusters 4 . Within CMD the time evolution of a finite and excited systems can be numerically calculated. In ref.1 the time evolution of 100 particles was followed. 179
180
P/Po Figure 1: Time evolution of the biggest fragment in the density temperature plane.
The particles were initially given a Maxewell Boltzmann distribution at temperature T. The density and temperature estimated for the biggest fragment are plotted in Fig. 1 (full lines). Each dot is printed at regular time intervals of 1 fm/c. We have used a gs density for the finite system of 0.12fm -3 (0.15fm -3 for the infinite case limit) and a B.E. of - 1 0 MeV (-16 MeV in the infinite case) to normalize the curves. Without such normalizations the lowest curve (T = 2 MeV) for instance will enter deeply inside the instability region which is not the case since this is a typical case of evaporation. "Critical" events occur for the initial T = 5 MeV 1 , which we see that enters into the instability region near the critical point for a second order phase transition. Note that the value of the excitation energy E* for this case is E*/Eb = 0.71, which coincides with the value of the E*/Eb obtained in the infinite case limit at the critical density and temperature. Some features in this figure are of interest. The first one is that the expansion is isentropic (dashed lines in Fig. 1) until the system enters the instability region; events at high T do not enter the
181
instability region, i.e. the system dissolves quickly because of the very large excitation energy. Already from these numerical experiments we understand that finite systems are quite different from the infinite limit case, in that there is no confining volume, thus T and p are time dependent quantities. Their values cannot be fixed from the outside but, to some extent the initial excitation energy and density can be fixed. The concept of temperature becomes questionable as well especially at high E*. In fact the system expands rather quickly, and the very energetic particles leave the system without interacting with other particles. Thus fluctuations are small and decrease as we will show later with increasing E*. Looking more carefully at Fig. 1, we notice that it is quite difficult to explore the instability region at small T. In fact in the case T = 2 MeV the system is not able to enter the instability region. We need to give more excitation energy to it, but if we do so the T when the system enters the instability region is larger. We can imagine to have a barrier in a collective space where the radial coordinate R and its conjugate momentum are the relevant degrees of freedom. Analogously to the process of spontaneous fission (SF) we can imagine that in presence of the long range Coulomb force, the barrier as a finite width, thus the quantum mechanical process of tunneling might be possible and the system enters the spinodal region at low T. We call the fragments thus formed quantum drops and discuss the process in a later section.
2
Observables
During a heavy ion collision many fragments are formed and finally detected. The first observable that was tested was the mass yield 5 . In fact from the Fisher model of phase transition we expect that if there is a phase transition, a power law should appear in the spectrum, see ref.6. In order to illustrate the Fisher model and to have a look at the modifications needed when dealing with a finite system we have repeated the calculations described above, each time changing randomly the momenta of the particles. In the simulation we can generate thousands of events and produce mass distributions at each T. The mass distribution so obtained is in good agreement with the Fisher law. The fits are rather good especially near and above the T = 5 MeV case. The latter gives a very good power law distribution with r = 2.23. Such a value of r is in good agreement with what observed in liquid-gas phase transitions of finite systems. At very high T the yield is practically exponentially decreasing, while low T give one big fragment accompanied of some small ones. Other variables like moments of the mass distributions, Campi plots 7 , give indications for a possible second order phase transition. In ref.8 we have tested if these findings
182
based on CMD results have some resemblance with reality. The experiment Au+Au at 35 MeV/A performed at MSU using the Multics-Miniball detectors 8 has been analyzed in terms of moments of mass distributions. First the collision was simulated in the framework of CMD and the effects of the experimental device were studied in detail 9 . It was found that a good reconstruction of the PLF can be obtained with this device. Thus peripheral collisions are rather well detected by the Multics-Miniball apparatus. For such collisions a critical behavior had been predicted within CMD in ref.10 and confirmed for this system and at this energy as well 9 ' 11 . 3
Chaotic Dynamics
The observation of large fluctuations in fragmentation hints to the occurrence of chaos. In order to address this problem quantitatively we calculate the maximal Lyapunov Exponents (MLE) as a function of the initial excitation energy. An important property of chaotic motion is the high sensibility to changes in the initial conditions. Closely neighboring trajectories diverge exponentially in time. For regular trajectories, on the other hand , they are found to diverge only linearly. The quantity that properly quantifies the rate of exponential divergence are the LE 1 2 . The MLE have been calculated in ref.12 for the system of Fig. 1 and analogous calculations have been performed in ref.13 within the Boltzmann Nordheim Vlasov (BNV) framework, i.e. a mean field description of a disassembling nucleus. In Fig. 2 the LE is plotted vs. excitation energy in the CMD (diamonds) and BNV case (squares). The qualitative behavior is the same, i.e. both calculations display a maximum at the normalized E* = 0.5. Such maximum corresponds in the CMD to the value where the mass distribution is a power law. At low excitation energy the LE calculated in BNV are larger than the CMD case because the gs of the nucleus is a liquid while the classical gs is a solid. Very important is the decrease of the LE at high E*. This clearly demonstrates that the degree of chaoticity, i.e. thermalization is not increasing and the initial excitation energy is partially thermal but a large amount is in the form of collective expansion. This is consistent with the picture of a limiting temperature that the nucleus can sustain 1 4 . In fact we can have a properly thermalized system when the self consistent field is able to bind the particles in some volume for some time. This field acts as a confining volume where the particles stay to boil. But, when the excitation energy is large, particles have enough kinetic energy to leave the system promptly. This picture greatly clarifies the dynamics of fragmentation. At low excitation energy we can have a liquid at a temperature T which evaporates particles. At high E* energetic fragments are quickly emitted and a small liquid at a
183
0.06
0.04
0.02
Figure 2: Maximal Lyapunov exponents in CMD (diamonds) and BNV (squares) calculations.
limiting T remains. Thus the transition is from liquid to free particles thus somewhat different from a liquid to gas phase transition that would occur if the system was confined in a box. 4
Quantum Drops
Imagine that in some way we have been able to prepare the system at density p and temperature T. Because of the compression and/or thermal pressure, the system will expand. If the excitation energy is too low the expansion will come to an halt and the system will shrink back. This is some kind of monopole oscillation. On the other hand if the excitation is very large it will quickly expand and reach a region where the system is unstable and many fragments are formed, cf. Fig. 1. It is clear that in the expansion process the initial temperature will also decrease. We could roughly describe this process with a collective coordinate R(t), the radius of the system at time t and its conjugate coordinate. Here we are simply assuming that the expansion
184
t =
o = 300 fin/c
0fta/c
20 -
20
0
-20 -
-20 H
>—I
>—I
H—'
1—t—
(-
/=80fon/c
H—i h ??! = 600 fin/c
-+-
20
20
0
-20
-20 -\—,
1
.-
•< 1— ^ = 1 0 0 fin/c
20
-+-
-+* = 800 ftn/c 20
a 0
-20
-20 -^10
-20
20
40
4 )
-20
0
20
40
Figure 3: Time evolution of 23SU at 6.8 MeV/nucleon excitation energy. At time t = 100 fm/c the system enters imaginary time and exits at t = 600 fm/c. The real- and imaginarytimes are indicated by t and t/.
is spherical. These coordinates are somewhat the counterpart of the relative distance between fragments in the fission process. Similarly to the fission process we can imagine that connected to the collective variable R{t) there is a collective potential V(R)15. When the excitation energy is too low it means that we are below the maximum of the potential. That such a maximum exists comes, exactly as in the fission process, from the short range nature of the nucleon-nucleon force and the long range nature of Coulomb. Thus, similarly to SF, we can imagine to reach fragmentation by tunneling through the collective potential V(R). When this happens, fragments will be formed at very low T not reachable otherwise than through the tunneling effect. The
185
price to pay as in all the subbarrier phenomena is the very low cross sections. Since the expansion of the system can be parametrized in one-dimensional collective coordinate, we can, in principle, apply the imaginary-time method to its study similarly to ref.15. Here we combine the imaginary time prescription with Quantum Molecular Dynamics 16 model to simulate fragmentation of finite system with relatively low excitation. We simulate the expansion of 230 U system. First we prepare the ground state of a nucleus and then compress uniformly to give an excitation energy E* from 5 to 8 MeV/nucleon. Due to the fluctuations between events caused by the different initial configuration of nucleus, the potential energy during the expansion is different for different events. Therefore the tunneling fragmentation occurs in some events where the potential energy is eventually high, while there is no tunneling for events with lower potential energy. The number of events with tunneling fragmentation is larger for lower excitation energy. In Fig. 3 we display snapshots of a typical tunneling event. The collective coordinate P(t) becomes zero at t = 100 fm/c. At this stage we turn to imaginary times and the tunneling begins. Notice that the system indeed expands and its shape can be rather well approximated to a sphere at the beginning. But already at 300 fm/c (in imaginary time) due to the molecular dynamics nature of the simulation, the spherical approximation is quite bad. This shortcoming of our approach should be kept in mind because the calculated action will be quite unrealistic due to the approximation used. This is similar to the use of one collective coordinate (the relative distance between centers) in SF process. In that case many calculated features are quantitatively wrong but qualitatively acceptable. Since our Quantum Drops is a proposed novel mechanism we can only give qualitative features and the model assumption must be refined when experimental data will start to be available 17 .
5
Conclusions
Some theoretical indications for the behavior of a finite excited system are clear. Both CMD and BNV calculations of the MLE show that the system cannot hold more than a certain temperature. This seems intuitively reasonable because it is quite clear that the finite nuclear system is not able to stay bound when an excitation energy larger than 1.5~2 times the BE is given to it. It seems that at such large excitation energies, the systems quickly disassembles and consequently the biggest fragment left will have a lower excitation energy. The concept of a limiting temperature that the system can sustain seems reasonable, and such a limiting temperature is strictly related to the
186
maximum value of the Lyapunov exponent . We propose also to search for fragmentation at very low excitation energies where quantum effects play a dominant role. These rare events should give valuable informations about the instability region. References 1. V. Latora, M. Belkacem and A. Bonasera, Phys. Rev. Lett. 73,1765 (1994); M. Belkacem, V. Latora and A. Bonasera, Phys. Rev. C 52 (1995) 271; P. Finocchiaro, M. Belkacem, T. Kubo, V. Latora and A. Bonasera, Nucl. Phys. A600, 236 (1996). 2. L. Landau and E. Lifshits, Statistical Physics, Pergamon, New York, 1980; K.Huang, Statistical Mechanics, J.Wiley , New York, 1987, 2nd ed. 3. A.L.Goodman, J. I. Kapusta and A. Z. Mekjian, Phys. Rev. C 30, 851 (1984). 4. A.Bonasera, Phys. World (Feb. 1999) 20. A. Bonasera and J. Schulte, contr. to the Proceedings Similarities and Differences between Atomic Nuclei and Clusters, et al. Abe et al. editors, AIP (1998). 5. M. L. Gilkes et al. Phys. Rev. Lett. 73, 1590 (1994). 6. M. E. Fisher, Proc. International School of Physics, Enrico Fermi Course LI, Critical Phenomena, ed. M. S. Green (Academic, New York, 1971); 1967255. 7. X. Campi, J. of Phys. A19, 917 (1986); Phys. Lett. B208, 351 (1988). 8. P. F. Mastinu et al. Phys.Rev.Lett. 76, 2646 (1996), and references therein. 9. M. Belkacem et al. Phys.Rev. C54, 2435 (1996). 10. V. Latora, A. Del Zoppo and A. Bonasera, Nucl. Phys. A572, 477 (1994). 11. A. Bonasera, M. Bruno, C. Dorso and P. F. Mastinu, Riv. Nuovo Cimento 23, 1-101 (2000). 12. A. Bonasera, V. Latora and A. Rapisarda, Phys. Rev. Lett. 75, 3434 (1995). 13. G. F. Burgio and A. Bonasera, in preparation. 14. K. Hagel et al., Phys. Rev. 62, 34607 (2000). 15. A. Bonasera and A. Iwamoto Phys. Rev. Lett. 78, 187 (1997). 16. T. Maruyama et al., Phys. Rev. 57, 655 (1998). 17. T. Maruyama, S. Chiba and A. Bonasera, in preparation.
E X P E R I M E N T A L SIGNALS OF THE FIRST P H A S E T R A N S I T I O N OF N U C L E A R M A T T E R
B. BORDERIE Institut
de Physique
Nucleaire,
IN2P3-CNRS,
P-91406
Orsay Cedex,
France.
Vaporized and multifragmenting sources produced in heavy ion collisions at intermediate energies are good candidates to investigate the phase diagram of nuclear matter. The properties of highly excited nuclear sources which undergo a simultaneous disassembly into particles are found to sign the presence of a gas phase. For heavy nuclear sources produced in the Fermi energy domain, which undergo a simultaneous disassembly into particles and fragments, a fossil signal (fragment size correlations) reveals the origin of multifragnsentation:spinodal instabilities which develop in the unstable coexistence region of the phase diagram of nuclear matter. Studies of fluctuations give a direct signature of a first order phase transition through measurements of a negative microcanonical heat capacity.
1
Introduction
The decay of highly excited nuclear systems through a simultaneous disassembly into fragments and particles, what we call multifragmentation, is a subject of great interest in nuclear physics. Indeed multifragmentation should be related to subcritical and/or critical phenomena. Thus it is fully connected to the nature of the phase transition which is expected of the liquid-gas type due to the specific form of the nuclcon-nuclcon interaction; as van dcr Waals forces, the nuclcon-nuclcon interaction is characterized by attraction at long and intermediate range and repulsion at short range. Although multifragmentation has been observed for many years, its experimental knowledge was strongly improved only recently with the advent of powerful devices built in the last decade. Selecting the ''simplest'' experimental situations, well defined systems or subsystems which undergo vaporization (simultaneous disassembly into particles) or multifragmentation can be thus identified and studied. It is a difficult task to deduce information on the phase diagram and the related equation of state of nuclear matter from nucleus-nucleus collisions at intermediate energies. But it is also a very exciting novel physics in relation with thermodynamics of finite systems (connection to other fields) without external constraints (pressure,volume) 1 ' 2 .
187
188
From multifragmentation to vaporization : identification of the gas phase
^ o
5
0 4.5
• CI+Au35MeV/u O Ge+Ti 35 MeV/u • Au+X 600 MeV/u (Aladin) • Xe+Sn 3 2 - 5 0 MeV/u (Indra)
-?3.5
+^
i 3 E
2 1.5
T5 a>
1
|
oxr
^0.5 E
-
T
T-
"S o,
6
8
10
12
14
16
Figure 1. Fragment multiplicity (normalised to the number of incident nucleons) as a function of the excitation energy. (from 3 ).
Excitation energy (MeV/u)
Let us first locate in which excitation energy domain multifragmentation takes place. Fig 1 indicates the evolution of the reduced (normalised to the size of the multifragmenting system) fragment multiplicity as a function of the excitation energy per nuclcon of the system. A universal behavior characterized by a bell shape curve is observed. The onset of multifragmentation is observed for excitation energies around 3 McV per nuclcon. the maximum for fragment production is found around 9 McV per nuclcon. i.e. close to the binding energy of nuclei. At higher excitation energy, the opening of the vaporization channel reduces fragment production. Temperature (MeV) S 10 1214 16 18 20 22 24 ~^45
. 1 1 1 1 1 1 l/l i / | i i J i jj i i i if |
&
J*
I / I
a *
r * n 25
0
* 1 -
L • P
• • 15 !" 10 15 L 20
26
i / | i i i 1.
• A A O D
d 1 'lie 4 He "He(x5)
m^~* • i••••i••
*
•
•--•
ftTfl'Vn-rn
: •:
15 20 25 30 Excitation energy (AMeV)
Figure 2. Composition of vaporized quasiprojectiles, formed in 3e 95 AMeV Ar+SBNi collisions, as a function of their excitation energy per nucleon. Symbols are for data while the lines (dashed for He isotopes) are the results of the model. The temperature values used in the model are also given, (from 4 ) .
The gas phase has been identified by studying the dcexcitation properties
189
of vaporized quasi-pro jectilcs with A around 36 4 . Chemical composition (first and second moments) and average kinetic energies of the different particles arc well described by a gas of fcrmions and bosons in thermal and chemical equilibrium. Inclusion of a van dcr Waals-likc behavior (final state excluded volume) was found decisive to obtain the observed agreement (sec for example figure 2). In the model, the experimental range in excitation energy per nuclcon of the source was covered by varying the temperature from 10 to 25 McV and the free volume was fixed at 3Vo, which corresponds to an average inter-distance between particles of about 2 fm. close to the range of the nuclear force (freeze-out configuration). 3
Thermometry and calorimetry : caloric curves and first-order phase transition ?
The plateau observed in the shape of the caloric, curve (determined from calorimetry and nuclear thermometry) was proposed by the ALADIN collaboration a few years ago as a signature of a first-order phase transition 5 . Since this observation works from different collaborations, covering a large range in mass of nuclei, have been published 6-7-8'9. Many caloric curves have been obtained which can roughly be classified in two groups depending on the nuclear thermometer chosen (isotopic double ratios using (6Li/7Li)/(3Hc/iHc) i 4 or (d/t)/C Hc/ Hc)). Moreover the presence of a plateau was not confirmed in these studies, even by the ALADIN collaboration when looking at properties of target-like spectators in Au+Au collisions at lOOOAMcV 10 . All these studies clearly indicate that no decisive signal can be extracted. We do not have an absolute nuclear thermometer and above all, experimentally one docs not explore the caloric curve at constant pressure nor at constant volume. In fact measured caloric curves arc sampling a mono dimensional curve on the microcanonical equation of state surface (T versus energy and volume) 2 ' n ; for each energy of the system a different average volume at freeze-out (no more nuclear interaction) is obtained, depending on the observed partition. 4
Statistical and dynamical descriptions of multifragmentation
Many theories have been developed to explain multifragmentation (sec for example rcf. 12 for a general review of models). Among the models some arc related to statistical approaches 13 ' 14 . valid at and after freeze-out, whereas others try to describe the dynamical evolution of systems, from the beginning of the collision between two nuclei to the fragment formation 15 ' 16 . I shall very briefly focus here on two models which will be compared later on to ex-
190
pcrimcntal d a t a . Firstly a statistical description of multifragmentation (SMM model 1 4 ) , in which an equilibration of a system a t low density is assumed. T h e n t h e statistical weight of a given break-up channel / , i.e. t h e number of microscopic states leading to this partition, is determined by its entropy AFf=cxp Sf within t h e microcanonical framework. In such an approach the initial p a r a m e t e r s as t h e mass a n d charge of t h e multifragmenting system, its excitation energy, its volume (or density) and t h e eventual added radial expansion have to b e backtraccd to experimental d a t a . Secondly, dynamical stochastic mean-field simulations which arc obtained by restoring fluctuations in deterministic one-body kinetic simulations. In particular in such simulations, relative to t h e s t a n d a r d nuclear Boltzmann t r e a t m e n t , an approximate tool is provided by introducing a noise by means of a brownian force in t h e m e a n field (Brownian One-Body (BOB) dynamics 1 7 , 1 8 ) . T h e m a g n i t u d e of t h e force is adjusted to produce t h e same growth r a t e of fluctuations as t h e full Boltzmann-Langcvin theory 1 9 . Such simulations completely describe t h e time evolution of t h e collision and thus help in learning a b o u t nuclear m a t t e r and its phase diagram whereas statistical models start from t h e phase diagram a n d have more t o do with the t h c r m o d y n a m i c a l description of finite nuclear systems. B o t h descriptions have been successful in reproducing average static and kincmatical properties of fragments (sec for example rcf. 2 0 ) . They will b e compared in what follows to more constrained obscrvablcs which arc expected t o bring decisive information on the origin and properties of multifragmentation.
5
5.1
Correlations in events: spinodal instabilities and equilibration Fragment
size. correlations:
a fossil
signature
Dynamical simulations predict t h a t during a central collision between heavy nuclei in t h e Fermi energy domain (30-40 McV per nucleoli incident energies) a wide zone of t h e phase diagram is explored (gentle compression-expansion cycle) a n d t h e fused system enters t h e liquid-gas coexistence region (at low density) and even more precisely t h e unstable spinodal region (domain of negative incomprcssibility). Thus a possible origin of multifragmentation may b e found t h r o u g h t h e growth of density fluctuations in this unstable region. W i t h i n this theoretical scenario a breakup into nearly equal-sized "primitive" fragments should b e favored in relation with the wave-lengths of t h e most unstable modes present in t h e spinodal region 2 1 . However this picture is
191
expected to be blurred by several effects: the beating of diffcrcnts modes, the presence of large wave-length instabilities and eventual coalescence of nascent fragments. Then how to search for a possible "fossil" signature of spinodal decomposition? A few years ago a new method called higher order charge correlations was proposed in 2 2 . All fragments in one event (average fragment charge < Z > and the standard deviation per event AZ) arc used to build the charge correlation for each fragment multiplicity. Experiment
BoB
Figure 3. Fragment charge correlations from fused events produced in central collisions between Xe and Sn at 32 MeV per nucleon incident energy: comparison between experiment (left) and BOB calculations (right) for fragment multiplicities equal to 4 and 6. (from 2 0 ) .
Figure 3 shows results from 20 for such correlation functions in experimental fusion events and BOB simulated events (Xc+Sn system at 32 MeV per nuclcon). For all fragment multiplicities the charge correlation has a peak in the bin AZ — 0-1, indicating an enhancement of partitions with equalsized fragments. This weak but non ambiguous enhancement (0.1% of events if wc restrict to the bin 0-1 and about 1% if we enlarge to bin 1-2 to take into account secondary decays of fragments) is interpreted as a signature of spinodal instabilities as the origin of multifragmentation in the Fermi energy
192
domain. Moreover the occurrence of spinodal decomposition signs the presence of a liquid-gas coexistence region and consequently, although indirectly, a first order phase transition. 5.2
Fragment-particle correlations: equilibrium at freeze-out
Fragment-particle velocity correlations in events have been proposed to experimentally measure excitation energy of hot primary fragments produced in multifragmentation 23 . By means of this technique multiplicities and relative kinetic energy distributions between fragments and light charged particles that they evaporate arc determined to reconstruct the excitation energies of fragments. For the Xc+Sn reaction, the INDRA collaboration has measured the evolution of the average excitation energy per nuclcon of primary fragments produced in multifragmentation of fused systems at different incident energies (from 32 to 50 McV per nuclcon) 24 . Within the error bars a constant value around 3.0-3.5 McV per nuclcon was measured in good agreement with the approach at equilibrium (SMM). This suggests that equilibrium is reached at freeze-out. Note that dynamical simulations (BOB) performed at 32 McV per nuclcon also predict the same excitation of fragments at freeze-out 20 . How to reconcile the dynamical (spinodal instabilities) and statistical (equilibrium at freeze-out) aspects which have b e e n extracted from correlations ? The following scenario can be proposed: spinodal instabilities cause multifragmentation but when the system reaches the freeze-out stage, it has explored enough of the phase space in order to be describable through an equilibrium approach. 6
Kinetic energy fluctuations and negative microcanonical heat capacity
Within the microcanonical equilibrium framework, it was recently shown 25 ' 2 that for a given total energy of a system, the average partial energy stored in a part of the system is a good microcanonical thermometer, while the associated fluctuations can be used to construct the heat capacity (sec 2 ). In presence of a phase transition large fluctuations arc expected to appear as a consequence of the divergence and of the possible negative branch of the microcanonical heat capacity. From experiments the most simple decomposition of the total energy E* is in a kinetic part E\ and a potential part E2 (Coulomb energy + total mass excess). However these quantities have to be determined at freeze-out and consequently it is necessary to trace back this configuration on an event by event basis. The true configuration needs the knowledge of all the charged
193
j . . . . . . . . . . .
1
•£" 40 "O JJ a. 3 30
m
, . . . , . . . ,
i i i j i
i | i i
^ Xe + Sn * : 32A.MeV|M
s B
^20 -40
*
-60
*
*%
20
10
* %
0
a
"
r -
-10
EVA, (A MeV>
-20
Figure 4. Measurements of microcanonical heat capacity per nucleon (symbols) as a function of the excitation energy per nucleon for quasi-projectiles produced in A u + A u collisions. The two panels refer to different freezeout hypotheses. The grey contour indicates the confidence region for C*. (from ).
-30 -40
:
ri\
:•
ill
:
1 IL **%, i"* 1% \
F
/
•
1
\
'
1
\
i . . . i jJ. , ...,...,.
-
~ SM. . . . I
:
14 „(AJNfcV)
Figure 5. Same as figure 4 for fused nuclei produced in central collisions between Xe and Sn at 32 MeV per nucleon incident energy, (from 2 ' ) .
particles evaporated from primary hot fragments and of the undetected neutrons; consequently some reasonable hypotheses have to be done. Note also that fragment-particle correlations discussed just before can help to obtain a better knowledge of freeze-out configurations (sec 2 8 ) . Then the experimental correlation between the kinetic energy per nucleon E\j AQ and the total excitation energy per nucleon E* /AQ of the considered system can be obtained as well as the reduced variance of the kinetic energy a\j <E\>. Finally the microcanonical temperature of the system can be obtained by inverting the kinetic equation of state and the total microcanonical heat capacity C± is extracted from the following equations: Ci =
S < EI/AQ
8T
>
and
C{
Ct = Ci
¥
Figures 4 and 5 show results obtained by M. D'Agostino ct al. and the INDRA collaboration for hot nuclei with mass number around 200 formed in different experimental conditions. In figure 4 the micocanonical heat capacity is calculated over a large excitation energy range for quasi-projectiles (formed in Au+Au collisions at 35 MeV per nucleon incident energy) assuming two different hypotheses to trace back freeze-out configurations; figure 5 refers to
194
fusion events produced in central Xc+Sn collisions at 32 McV per nuclcon; in this latter case, a narrower excitation energy distribution (bell shape curve on the figure) is observed. A distinct negative branch is observed, revealing a first order phase transition. The distances between the poles arc associated with the latent heat. Note that the same location of the pole at high excitation energy is found when similar hypotheses are made for f rcczc-out reconstruction (left part of figure 4 and figure 5). 7
Signatures of critical behavior
For finite systems, related to the correlation length, a critical region instead of a critical point is expected. The following signatures of critical behavior were reported: power laws have been observed within selected conditions. critical exponents have been measured in agreement with with those of a liquid-gas model 29 ' 26 assuming that fragment multiplicity or thermal excitation energy is the control ("order") parameter and a nuclear scaling function has been evidenced by the EOS collaboration 2 9 . However, as compared to infinite systems, potential divergences are smoothened over finite regions of the chosen control parameter and the choice of the fit regions where the data are assumed to reflect the critical behavior is crucial. Clearly one needs a precise and objective procedure in order to determine order parameters and critical signals. Such a methodology was recently proposed for second order phase transitions 30 and we can expect in the future to dispose of a similar methodology for first order phase transitions. 8
Conclusions
A set of coherent results showing the existence of a first order phase transition in nuclear matter has been obtained and the two signals observed related to correlations and fluctuations constitute a strong starting point for systematic investigations. Clearly caloric curves do not and can not give a decisive signal of a first order phase transition. Selected data have properties compatible with the equilibrium hypothesis at freeze-out and this framework is up to now chosen for progressing in the experiment-theory interaction. Experimentally an effort has to be made to better define configurations at freeze-out. which arc key points to bring more quantitative information (latent h e a t . . . ) . On the theoretical side, concerning thermodynamics of finite systems, some improvements arc needed in lattice gas models for example to take into account the specificities of nuclei (quantal aspects and Coulomb interaction). Concerning the signatures of critical behavior and the definition of the critical
195
region, experimentalists need a methodology dedicated to first order phase transition for finite systems. I am highly indebted to my colleagues of the INDRA collaboration and to R. Botet, X. Campi, Ph. Chom,az, M. Colonna, M. D Agostino and F. Gulminclli for valuable discussions. References 1. 2. 3. 4. 5. 6. 7. 8. 9. 10. 11. 12. 13. 14. 15. 16. 17. 18. 19. 20. 21. 22. 23. 24. 25. 26. 27. 28. 29. 30.
D.H. E. Gross, this conf. Ph. Chomaz, this conf. D. Durand, Nucl. Phys. A 630, 52c (1997). B. Bordcric ct al., (INDRA coll.), Eur. Phys. J. A 6, 197 (1999). J. Pochodzalla ct al., (ALADIN coll.) Phys. Rev. Lett. 75, 1040 (1995). Y. G. Ma ct al., (INDRA coll.) Phys. Lett. B 390, 41 (1997). J. A. Haugcr ct al., (EOS coll.) Phys. Rev. C 57, 764 (1998). K. Kwiatkowski ct al., Phys. Lett. B 423, 21 (1998). J. Cibor ct al., Phys. Lett. B 473, 29 (2000). W. Trautmann for the ALADIN coll., Advances in Nuclear Dynamics A (1998) p.349. Ph. Chomaz, V. Dufiot and F. Gulminclli Ganil report P 00 01. L. G. Moretto and G. J. Wozniak, Ann. Rev. of Nuclear and Particle Science 43, 379 (1993) and references therein. D.H.E. Gross, Rep. Prog. Phys. 53, 605 (1990) and references therein. J. Bondorf ct al., Phys. Rep. 257, 133 (1995) and references therein. A. Guarncra ct al., Phys. Lett. B 373, 267 (1996). A.Ono an H. Horiuchi, Phys. Rev. C 53, 2958 (1996). Ph. Chomaz ct al., Phys. Rev. Lett. 73, 3512 (1994). A. Guarncra ct al., Phys. Lett. B 403, 191 (1997). Ph. Chomaz, Ann. Phys. Fr. 21, 669 (1996). M. F. Rivet ct al., (INDRA coll.) this conf. and references therein. S. Ayik ct al., Phys. Lett. B 353, 417 (1995). L. G. Moretto ct aL.Phys. Rev. Lett. 77, 2634 (1996). N. Marie ct al., Phys. Rev. C 58, 256 (1998). S. Hudan ct al., (INDRA coll.) Proc. Bormio (2000), p. 443. Ph. Chomaz and F. Gulminclli, Nucl. Phys. A 647, 153 (1999). M. D'Agostino ct al., Phys. Lett. B 473, 219 (2000). N. Lc Ncindrc ct al., (INDRA coll.) this conf. M. D'Agostino ct al., this conf. J.B. Elliot ct al., (EOS coll.) Phys. Lett. B 418, 34 (1998). R. Botet and M. Ploszajczak, Ganil report P 00 17.
NUCLEAR FRAGMENTATION, PHASE TRANSITIONS AND THEIR CHARACTERIZATION IN FINITE SYSTEMS OF INTERACTING PARTICLES
Dipartimento
J. M. CARMONA di Fisica, Via Buonarotti 2, Ed. B, 56127 Pisa (Italy) E-mail: [email protected]
N. MICHEL Grand Accelerateur National d'lons Lourds (GANIL), CEA/DSM-CNRS/IN2P3, BP 5027, F-14021 Caen Cedex (France) J. RICHERT Laboratoire de Physique Theorique, Universite Louis Pasteur, 3, rue de I'Universite, 67084 Strasbourg Cedex (France) E-mail: [email protected] P. WAGNER Institut de Recherches Subatomiques (IReS) BP 28, 67037 Strasbourg Cedex 2 (France) E-mail:pierre. [email protected] Since nuclei are finite, the study of critical properties of nuclear matter needs methods adapted to finite systems of particles. Tests which are aimed to characterize these transitions in finite sytems are presented and used on a simple lattice model. The Coulomb interaction is added and its influence on the thermodynamic properties of finite systems is examined.
1
Introduction
Fragmentation is the regime in which one expects to find t h e signs for t h e existence of a phase transition in nuclear matter. Since ordinary nuclear m a t t e r behaves like a Fermi liquid it is expected t h a t a change of phase shows the properties of a first order liquid-gas transition, by analogy with the behaviour of macroscopic systems. In the nuclear case, however, one is confronted with two problems. T h e first is the fact t h a t whatever heavy nuclei can be, they remain small fermionic q u a n t u m systems, contrary to other small q u a n t u m systems like Bose-Einstein condensates, q u a n t u m dots or aggregates in which there exists a much larger flexibility in the number of constituents. T h e second point concerns the conditions under which a critical transition can be seen. Highly excited nuclei
197
198
are generated in violent collisions in which nuclei may or may not reach thermodynamic equilibrium. These two points have received considerable attention in the past. Up to now, there exists no satisfactory answer to the second point which concerns also other fields of physics dealing with critical off-equilibrium systems. In the present contribution we concentrate on the first point. We make the implicit assumption of thermodynamic equilibrium. It is our aim to show the problems raised by the determination of the order of a transition when dealing with finite systems. This is done with the help of a lattice model which is expected to provide a classical but sensible description of a finite excited nucleus. 2
Characterization of phase transitions in finite systems
Several tests have been devised in order to characterize transitions and possibly their order in finite systems. The most popular one concerns the determination of critical exponents which fixes the universality class to which the transition belongs. It relies on the finite size scaling (FSS) assumption x . This assumption presupposes that observables scale with the size of the system in the neighbourhood of values of quantities like temperature, bond probability, . . . which correspond to a critical point in the thermodynamic limit. Recently there has been a revival of interest for the Lee and Yang theory in which critical behaviour corresponds to zeroes of the partition function, i.e. a singular behaviour of the free energy, for real values of the temperature 2 . It has been shown recently that information may be gained about the existence of a transition, at least for simple systems, by studying the location and evolution of zeroes of the partition function in the complex temperature plane, in the vicinity of the real axis 3 . The importance of the behaviour of fluctuations has been demonstrated by Botet and Ploszajczak 4 . Using scaling arguments they showed that order parameters characterizing finite systems which undergo second order transitions follow distribution functions which are different whether the system is at, close, or far from the critical point. These fluctuations have universality properties at the critical point. Last, finite systems behave differently when they are described in the framework of different statistical ensembles. The most striking effect is seen on the caloric curve. In a finite system which undergoes a first order transition in the thermodynamic limit the temperature gets a multivalued function of the energy, related to the convexity of the entropy in the region where liquid and
199
gas coexist and surface energy is stored in the domain walls which separate the two phases 5 . Here we show an application of the FSS method and the microcanonical test to a lattice model in an attempt to fix the order of the transition whose existence is detected in corresponding finite systems. 3
The model and thermodynamic properties
Since it has already been described in previous work briefly its essential features. The canonical distribution function reads
Z(J3) = $>xp[/3(K ~V0 Y,
6
n n
we sketch here very
* j)]
(!)
where /? = T _ 1 is the inverse temperature, K the kinetic energy, Vo the strength of a potential which acts between nearest neighbours. Particles are located on the TV sites of a 3d cubic lattice, nj = 0 (1) if site i is empty (occupied). The total number of particles A is rigorously fixed, which fixes also the density p = A/N. We concentrate the study on the potential energy U and the specific heat Cv- If, for fixed p, the transition which occurs at some temperature T is second order then one expects that
Cv ~ 00 - 0c)-a where (5C is the inverse critical temperature in the infinite system and a a real positive exponent. < U > is a monotonously increasing function of T showing an inflection point and Cv presents an enhancement which could be the sign for the existence of a phase transition. Using FSS arguments, one may look for a parametrization Cv{L) = A + (3C(L) - &(oo) =
BLa'v DL^'V
where L is the linear size of the system, A, B, D are constants. Fits to numerical simulations of systems up to L = 48 indicate that FSS works, allowing to determine a/v and v and the values of the exponent a are compatible with the universality class of the Ising model 6 . Since however, the numerical fits which lead to the determination of a showed some very small deviations from a straight line in a representation of InCV as a function of InL, further tests have been implemented 7 . The
200
most efficient one concerns the microcanonical behaviour of the caloric curve mentioned above which shows a typical backbending corresponding to the multivaluedness of T when represented as a function of E if the transition is of first order 8 . An explicit microcanonical determination of the caloric curve of the system in the framework of the model described above does not show the expected backbending expected for a first order transition. The transition can nevertheless be expected to be first order if the constant volume constraint is removed, allowing for the volume to fluctuate for a fixed pressure 9 . These results lead to the following comments. Constraints like a fixed number of particles in a fixed volume as discussed here may produce effects which are apparently able to change the behaviour of thermodynamic quantities. An example of constraints applied to the Potts model has been presented in ref. 9 . This observation should work as an incentive to caution in the interpretation of experimental results. The present analysis shows that the extrapolation to the infinite limit and hence the characterization of the behaviour of observables in the infinite system where they may become truly singular may be a delicate task.
4
Long range interactions
Nuclei experience the Coulomb interaction. In order to study its effect on the thermodynamic properties of the system described by the lattice model presented above we introduce the Hamiltonian
where H0 is the Hamiltonian used in (1), e is here the electron charge and Zi = 1 (0) if the particle on site is a proton (a neutron). The Metropolis Monte Carlo simulations are performed in the framework of the canonical ensemble with open boundary conditions. Z protons chosen such that the ratio A/Z corresponds to nuclei lying in the valley of stability are randomly distributed over the lattice. The typical behaviour of the caloric curve and Cv without and with the Coulomb interaction is shown in Fig. 1 for p = 0.3 and L = 10. As it can be seen the Coulomb interaction produces a sizable shift in the energy and reduces the temperature in the peak of CV by about 1 MeV, reducing the area of the coexistence zone in the (p,T) phase diagram. It does however not qualitatively affect the thermodynamic properties of the system.
201
with Coulomb -0.2
with Coulomb
without Coulomb
-0.4
-
-0.6
'-
_ » • *
without Coulomb
jr
r5 £ °-4
a
-0.8
<" 0.3 -1
'
a
& -1.2
-1.4
0.2
»»" TEMPERATURE (MeV)
Figure 1. Caloric curve and specific heat without and with Coulomb interaction, for L = 10 and p = 0.3. See comment in the text.
5
Summary and conclusions
The determination and the characterization of phase transitions through the study of finite systems has been examined by means of a lattice model. We applied two different tests in order to try to determine the order of the transition and showed that in finite systems constraints can play an essential role. Recently new methods which are aimed to detect phase transitions through the examination of finite systems have been proposed. It would certainly be worthwhile to test them on specific models. The presence of the long range Coulomb interaction has a quantitative but no qualitative impact on the thermodynamic properties of the system. References 1. Jae-Kwon Kim, J.F. Adauto de Souza and D.P. Landau, Phys. Rev. E
202
54, 2291 (1996). 2. T.D. Lee and C.N. Yang, Phys. Rev. 87, 410 (1952). 3. P. Borrmann, 0 . Miilken and J. Harting, Phys. Rev. Lett. 84, 3511 (2000). 4. R. Botet and M. Ploszajczak, cond-mat/0004003. 5. D.H.E. Gross and M.E. Madjet, Z. Phys. B 104, 541 (1997). 6. J.M. Carmona, J. Richert and A. Tarancon, Nucl. Phys. A 643, 115 (1998). 7. J.M. Carmona, N. Michel, J. Richert and P. Wagner, Phys. Rev. C 61, 37304 (2000). 8. D.H.E. Gross and E.V. Votyakov, Eur. Phys. J. B 15, 115 (2000). 9. Ph. Chomaz, V. Duflot and F. Gulminelli, preprint GANIL P 00 01, January 2000, submitted to Physical Review Letters.
TOPOLOGY A N D PHASE TRANSITIONS: TOWARDS A P R O P E R MATHEMATICAL DEFINITION OF FINITE N TRANSITIONS MARCO PETTINI, ROBERTO FRANZOSI", LIONEL SPINELLI* Osservatorio Astrofisico di Arcetri, Largo E. Fermi 5, 50125 Firenze, Italy E-mail: [email protected] Abstract A new point of view about the deep origin of thermodynamic phase transitions is sketched. The main idea is to link the appearance of phase transitions to some major topology change of suitable submanifolds of phase space instead of linking them to non-analyticity, as is usual in the Yang-Lee and in the Dobrushin-Ruelle-Sinai theories. In the new framework a new possibility appears to properly define a mathematical counterpart of phase transitions also at finite number of degrees of freedom. This is of prospective interest to the study of those systems that challenge the conventional approaches, as is the case of phase transitions in nuclear clusters. 1
Introduction
The current mathematical definitions of thermodynamical phase transitions are based on the loss of analyticity of thermodynamical observables. This is to some extent suggested, though not implied, by the experimental relations among macroscopic observables. This conflicts with the analytic character of the statistical weights that are attributed by any ensemble in statistical mechanics to the microscopic configurations. Thus, as it has been proved by the Onsager solution of the 2D Ising model and by the Yang-Lee theorem, the only way to eliminate this conflict is to consider the limit N —f oo (thermodynamic limit) 1 . Obviously phase transitions in Nature occur at finite N, but it is commonly argued that - for macroscopic objects - N is so large that it can be considered "infinite" from a physical point of view. However, by looking at a small and embroidery-like snowflake that melts into a drop of water one can wonder why such a phenomenon should ever be explained only in terms of the infinite JV limit. Moreover, there is a growing experimental evidence that phase transitions can occur also at finite and small N, i.e. with JV
204
number. Some examples of small N systems undergoing phase transitions are: i) nuclear clusters as well as atomic and molecular clusters; it) nano and mesoscopic systems; tit) polymers and proteins; iv) small drops of quantum fluids (BEC, superfluids and superconductors). Whence the prospective interest of a new mathematical characterization of thermodynamical phase transitions which, instead of resorting to the loss of analyticity of macroscopic observables, might naturally encompass also finite N systems. The search for a broader mathematical definition of phase transitions is also of potential interest to the treatment of other important topics in statistical physics, as is the case of amorphous and disordered systems (like glasses and spin-glasses), or for a better understanding of phase transitions in the microcanonical ensemble, for first-order phase transitions, and so on. 1.1
Heuristic
arguments
Everything here refers to classical Hamiltonian systems with continuous variables and described by a standard Hamiltonian H\p = {pi,...,pN),q=(qi,...,qN)]
= Y^2P*+V^
'
(1)
As a consequence of a systematic study of the dynamical counterpart of thermodynamical phase transitions, performed by numerically solving the Hamilton equations of motion of (1), it has been found that Lyapunov exponents display "singular" energy and temperature patterns at the transition point 2 . Thus, since in a differential-geometric description of Hamiltonian chaos Lyapunov exponents are seen as probes of the geometry of certain submanifolds of configuration space, it has been conjectured that a phase transition could be due to some major geometrical, and possibly topological, change in the support of the statistical measures 2 ' 3 ' 4 . This is to say that it is on the basis of the just mentioned work that we have been led to formulate the following argument. Let us consider the canonical configurational partition function
where a co-area formula has been used to unfold the structure integrals
fi Ws
" t-.ii£iF
(3)
i.e an infinite collection of purely geometric integrals on E ^ _ 1 , the equipotential hypersurfaces of configuration space defined by S ^ _ 1 =
{qElRN\V(q)
=
v}ClR.N.
205 If we consider the microcanonical ensemble, the basic mathematical object is the phase space volume
il{E) =jdr,
tf-\E
-n)
J dNp S(J2 \P< ~ V)
where
rt-\E-T,)
= JdNqe[V(q)-{E-T,)}
=J
re-v r \v J
a,, p ^ j
(4)
whence
also here, as in the above decomposition of Zc (/3, TV), the only non-trivial objects are the structure integrals (3). Once the microscopic interaction potential V(q) is given, the configuration space of the system is automatically foliated into the family {E„}r627e of equipotential hypersurfaces independently of any statistical measure we may wish to use. Now, from standard statistical mechanical arguments we know that the larger is the number TV of particles the closer to some E„ are the microstates that significantly contribute to the statistical averages of thermodynamic observables. At large TV, and at any given value of the inverse temperature /3, the effective support of the canonical measure is narrowed very close to a single E„ = Et,(/3c); similarly, in the microcanonical ensemble, the fluctuations of potential and kinetic energies tend t o vanish at increasing TV so t h a t the effective contributions to il(E) come from a close neighborhood of a E„ = E„( E c ). Now, the "topological conjecture" consists in assuming that some suitable change of the topology of the {E„}, occurring at some vc = wc(/3c) (or vc = Vc(Ec)), is the deep origin of the singular behavior of thermodynamic observables at a phase transition; (by change of topology we mean that {£„}„<„,. are not diffeomorphic to the {!*,}*>«,.)• In other words, the claim is that the canonical and microcanonical measures must "feel" a big and sudden change - if any - of the topology of the equipotential hypersurfaces of their underlying supports, the consequence being the appearance of the typical signals of a phase transition, i.e. almost singular energy or temperature dependences of the averages of appropriate observables. The larger is TV the narrower is the effective support of the measure and hence the sharper can be the mentioned signals. Eventually, in the TV —t oo limit this sharpening will lead to non-analyticity.
206 2
A theorem about topology and phase transitions
We have recently proved a theorem stating that topological changes of the hypersurfaces £„ are necessarily at the origin of phase transitions 5 6 ' . It applies to physical systems described by short-range potentials V, bounded below, of the general form:
V({q}) = Yl V°(H& " &H) + I>(H&H) " The theorem is proved in the reversed formulation: if the surfaces £» with -» = V/N € / = f» m ,*M] are diffeomorphic, then no phase transition-will occur in the corresponding temperature interval [/J(wTO), /3(*M)]The proof is lenghty and rather complicated but it proceeds along a logically simple path. Diffeomorphicity of the £ „ , after the "non-critical neck theorem" in Morse theory, implies the absence of critical points of V, i.e. VV ^ 0. In the absence of Morse critical points:
HL^=LDk{wn)da
dvk where
D'oiwir 1 ) = 2||v^ir2Mi - nvvir3AF
and where Afi is the trace of the shape operator of the hypersurface £ „ . There is an operator algebra to generate the powers Dk so that, being SN(V) = -^lnfijv (v(v)) the microcanonical configurational entropy, it is possible to show that sup SW(*) < oo ; N,v£I
sup
_.
N,v£I
£**
(*>) < oo , k = 1 , . . . , 4
that is: if the £„ are diffeomorphic then S;v(-») is uniformely convergent in C3(I) as N —> oo, and, by using the Legendre transform relationship SN(V) = /JV(/3) + /3 • v-+ o(N) between microcanonical configurational entropy and the canonical free energy /JV(/9) = jf InZ C (/3,N), this implies that - in the JV —¥ oo limit - the canonical configurational free energy is /oo(/3) € C2(I), i.e. at least twice differentiable, thus there are neither first nor second order phase transitions according to their standard definition. There is not a one-to-one correspondence between phase transitions and topology changes of the £ „ , the latters are necessary but not sufficient. Sufficiency conditions, to point out the special class of topology changes that give rise to phase transitions, are based on relations like 6 ' 6
i^*k¥mwn~c{v)lVoKs[
)]
5>(S,) t=0
207 that bridge thermodynamics and topology; here &i(E„) are the Betti numbers (cohomologicaJ invariants) of £ „ . It turns out that a "second order" topology change, i.e. a sudden change in the way of changing of topology as a function of v, is sufficient to entail a first or a second order phase transition. Such topological discontinuities can exist and can be detected at any finite N, though they yield non-analyticity of thermodynamic observables only when the support of the statistical measure indefinitely shrinks with N —¥ oo. In other words, we have here the possibility of properly defining phase transitions also at finite N, whose detection, direct or indirect, can be performed through quantities that probe the topology of the E„. It is an interesting and open question how to make the link with other approaches tackling the finite N transitions in a macroscopic phase space of thermodynamic variables 7 . 3
A direct numerical confirmation
The family of {Ev}v£2i associated with a (p4 model on a d-dimensional lattices Zd with d = 1,2 has been used 8 for a numerical check of the scenario sketched in the previous section. The model is described by
i€Z*
^
'
(ik)$.Zd
(ik) stands for nearest-neighbor sites, and in d = 2 it undergoes a second order phase transition. In order to directly probe if and how the topology change - in the sense of a breaking of diffeomorphicity of the surfaces £„ - is actually the counterpart of a phase transition, a diffeomorphism invariant has to be computed. This is a very challenging task because of the high dimensionality of the manifolds involved. One possibility is afforded by the Gauss-Bonnet-Hopf theorem that relates the Euler characteristic x(^v) with the total Gauss-Kronecker curvature KG of the manifold X
(S„)=7/
KG da
(7)
which is valid in general for even dimensional hypersurfaces of euclidean spaces 1R.N, and where 7 = 2/Vo/(S") is twice the inverse of the volume of an n-dimensional sphere of unit radius, and da is the invariant volume measure induced from TZ.N. In Fig. 1 x(^«) *s reported vs v: the Id case gives a "smooth" pattern of x{v), whereas the 2d case yields a cusplike shaped x(v) a t *h e phase transition point. There is here a direct evidence of a major and very sharp "second order" topological transition that underlies the phase transition, it is also remarkable that with a very small lattice of N = 7 x 7 sites such a sharp signal would never be obtained through standard thermodynamic observables 9 .
208
v/N
Figure 1: Euler characteristic *(£«) for X-d and 2-d tp* lattice models. Open circles: 1-d case, N = 49; full circles: 2-d case, N = 7x7. The vertical dotted line, computed separately for larger N, accurately locates the phase transition. Data are from Ref.[8].
References 1. C.N. Yang and T.D. Lee, Phys. Rev. 87, 404 (1952); D. Ruelle, Thermodynamic formalism, Encyclopaedia of Mathematics and its Applications, (Addison-Wesley, New York, 1978). 2. L. Caiani, L. Casetti, C. Clementi, and M. Pettini, Phys. Rev. Lett. 79, 4361 (1997); L. Caiani, L. Casetti, C. Clementi, G. Pettini, M. Pettini, and R. Gatto, Phys. Rev. E 57, 3886 (1998); M. CerrutiSola, C. Clementi and M. Pettini, Phys. Rev. E 6 1 , 5171 (2000). 3. R. Franzosi, L. Casetti, L.Spinelli and M. Pettini, Phys. Rev. E 60, R5009 (1999); L. Casetti, E.G.D. Cohen and M. Pettini, Phys. Rev. Lett. 82, 4160 (1999). 4. L. Casetti, E.G.D. Cohen and M. Pettini, Phys. Rep. 337, 237-342 (2000). 5. R. Franzosi, (PhD thesis, Universita di Firenze, 1998). 6. L. Spinelli, (PhD thesis, Universite de Provence, 1999). 7. D. H. E. Gross, Phys. Rep. 279, 119 (1997). 8. R. Franzosi, M. Pettini and L. Spinelli, Phys. Rev. Lett. 84, 2774 (2000). 9. L. Caiani, L. Casetti, and M. Pettini, J. Phys. A: Math. Gen. 3 1 , 3357 (1998).
Introduction to the Phase Transition Discussion Philippe CHOMAZ G.A.N.I.L.(CEA-DSM/IN2PS-CNRS),
1
BP 5027, 14076 Caen cedex 5, France
INTRODUCTION
Since nuclear forces resemble to Van der Waals interactions the nuclear phase diagram is expected to present a liquid gas phase transition. Our present knowledge of the nuclear equation of state is limited. The main reason is the difficulty to treat the nuclear many-body problem and to define a reliable in medium interaction. The saturation energy and density, i.e. the ground state of nuclear matter, are well established but the compressibility, i.e. the variation of the energy as a function of the density around the saturation point, is still under discussion because of the recent results of relativistic approaches. As far as the temperature dependence of nuclear properties is concerned very little is also know in an absolute way. Only the entropy variation, i.e. the level density parameter o = S/T, of a finite nucleus as an open system has been clearly established through evaporation studies. A huge research activity is now devoted to the extraction of reliable information of the nuclear equation of states and the associated phase diagram. Heavy ion reactions are routinely used to test mechanical and thermodynamical properties of nuclei. In the recent years the multifragmentation regime has been tentatively associated with the occurrence of a liquid-gas phase transition. The main problem is to link the dynamics of a collision with the extraction of meaningful thermodynamical quantities. It is rather surprising that a huge amount of experimental data can be explained by many different models. For example the experimental data obtained in the Xe+Sn reaction at 32 MeV/nucleon can be explained by a dynamical simulation of the phase transition as well as by a simple statistical model 1 . This may indicate that the dynamics is sufficiently chaotic to populate the whole phase space. As a consequence a thermodynamical approach might be justified. In the following we will concentrate on the properties of a nuclear system in a statistical equilibrium. 2 2.1
FROM INFINITE TO FINITE SYSTEMS Phase transition in infinite systems
Phase transitions are routinely encountered in the everyday life. Everybody knows that a boiling water keeps a constant temperature during the whole boil209
210 ing process (since the pressure is usually constant). It should be noticed t h a t this "kitchen"-experiment is in fact a microcanonical type of thermodynamics since we usually control the amount of energy (the fire) given to the system (the water in the kettle). This amazing constancy of the t e m p e r a t u r e can easily be understood at the thermodynamical limit. Indeed, at this limit the state
Figure 1: Left: typical evolution of the entropy and the associated temperature of an infinite system undergoing a phase transition. Right: same as left but for a finite system. Schematic pictures of the system at the phase transition are also shown. of the system is the one which maximizes the entropy because the fluctuations around this m a x i m u m are neglected. Moreover, if we split the system in two p a r t s containing a fraction A and (1 — A) respectively of the total mass of the system, again because we assume t h a t in the t h e r m o d y n a m i c a l limit we can neglect the role of interfaces on b o t h energy and entropy, we can construct a macro-state giving the energy e\ = XEy to the first part and e2 = (\ — X) E2 to the second part. T h e total energy and entropy of this mixed events are: E=e1+e2
= XE1 + (l-X)E2
(1)
211
S = s1 + s2 = \S1+(l-X)S2
(2)
Indeed, the number of mixed states is simply the number of states in the first part W\ — exp s\ — exp \S\ times the number of states in the second part W2 - exp s2 = exp (1 - A) S 2 • If in some region the introduction of such mixed events leads to an entropy which is greater t h a t the original one the mixed event will be infinitely more probable t h a t the original one. Since S and E are in the mixed region simple linear functions of the mixing proportion A, the entropy is a straight line so t h a t its derivative, the microcanonical t e m p e r a t u r e , remains constant. This is nothing but the Maxwell construction. This is the reason why at the thermodynamical limit the entropy should always be concave. This thermodynamical scenario is shown in the left part of Fig. 1.
2.2
Channel opening in
finite
systems
In a finite system the simple argument about the mixed events does not hold anymore. Indeed, gluing two parts in proportion A and (1 — A) of two classes of events associated with the energy E\ and E2 does not generate a class of events at energy E = AJEI + (1 — A) E2 because we cannot neglect the energy cost paid in terms of interfaces. Therefore neither the energy nor the entropy are expected to be additive and mixed events m a y not be the most i m p o r t a n t ones. T h e Maxwell construction is therefore not possible any more and the construction of mixed event is not the way we may think about a phase transition. Nothing then prevents the entropy to present concave regions. It has been proposed t h a t a concave anomaly can be considered as a general definition of phase transitions in finite s y s t e m s 2 . This idea has been generalized to any convexity anomaly of any generalized thermodynamical p o t e n t i a l 3 . If mixing is no more possible, sorting in categories remains a way to understand the transition from one type of event to another. In general m a n y different types of events have to be considered in a phase transition region. In the infinite system one m a y think to classify events according to their proportion A of one of the two phases, in finite systems fewer categories can be defined. For simplicity let us introduce only two types (labelled 1 and 2) of events on the basis of a large difference in an observable which can be thought as the order parameter. These two categories can be thought as two different channels. Far from the transition region only one of the two channels dominates. As soon as we pass the energy threshold for the opening of the second channel the relative probability of the two channels will be fixed by the respective degeneracies W\ = exp Si and W2 = exp S2. Since we are only sorting
212
events in categories the total degeneracy of the considered energy is simply: W = Wx + W2
(3)
If the entropy increase of the second channel is fast then it may rapidly overcome the first one leading to a concave anomaly in the total entropy S = log W. This demonstrates that the idea of channel opening is intimately linked to the idea of phase transition. 3 3.1
EXPERIMENTAL SIGNALS Melting of clusters^
In the year 2000 the first experimental signature of a back bending caloric curve has been reported in the melting of metallic clusters. The experiment is rather simple. The clusters are first produced and selected. Then the clusters get thermalized in the melting region in an helium heat bath. After thermalization they are further excited by a laser beam absorbing several photons, thanks to the plasmon vibration. The energy is then such that the cluster has time to evaporate atoms within the experiment time scale. The number of evaporated atoms provides a measure of the cluster energy distribution. Changing the temperature the thermal excitation changes and the distribution of evaporated atoms is shifted. The obtained bidimensional pictures of the number of evaporated atoms as a function of the oven temperature clearly show an anomaly corresponding to the melting point. If now we consider a back bending or a monotonous caloric curve the energy distribution goes from a bi-modal to a mono-modal shape. This induces a modification of the fragmentation pattern. The observed pattern is only compatible with a negative heat capacity system. 3.2
Negative heat capacities and abnormal fluctuations
Recently many progresses have been performed in order to extract the nuclear thermodynamics from experimental data. Vaporization threshold have been measured by the INDRA collaboration 5 . Among the most famous attempts stands the ALADIN caloric curve which shows a saturation in the temperature (plateau) in the phase transition region 6 . More recently the possibility to signal this transition using the fluctuation of the energy partition has been investigated 7 ' 8 , 9 and the presence of a negative heat capacity have been reported 1 0 > n . The investigation method can be easily explained for a classical fluid and tested in the framework of the lattice-gas model. The total energy E of the
213
r?
WkOEidWp^
wore*
S
£ ' * * - •<Efc>2/3 .52
h
.60 .
t>
rl-
1 S
«
o
4*
Carti&niLoaX.
_4j
•
- . OK*A/T a BI»
0.
,_£l
C
QX a t 0.6 Emtft^ (O
Figure 2: Left; schematic distribution of the partial energy for a fixed total energy. Right: comparison of the various measurements (dots) with the exact results of the lattice-gas model (lines). considered system can be decomposed into two independent components, its kinetic and potential energy: E = Ek + Ep. In a microcanonical ensemble with a t o t a l energy E the t o t a l degeneracy factor W (E) — exp (S (E)) is thus simply given by the folding product of the individual degeneracy factors Wi (Ei) = exp (Si (E()) of the two subsystems i = k,p. One can then define for t h e t o t a l system as well as for the two subsystems the microcanonical t e m p e r a t u r e s 7} and the associated heat capacities Ct-. If we now look at the kinetic energy distribution when the t o t a l energy is E we get PkB (Ek) = exp (Sk (Ek) + SP(E~
Ek) - S (E))
(4)
Using Eq. (4) we directly get t h a t the most probable partitioning of the t o t a l energy E between the potential and kinetic components is characterized by a unique microcanonical t e m p e r a t u r e T = Tk (Ek ) = Tp (E — Ef?). Therefore the most probable kinetic energy Ef? can be used as a microcanonical therm o m e t e r as shown in Fig. 2. Using a Gaussian approximation for Pf? (Ek) t h e kinetic energy variance can be calculated a s 1 2 : 2
^ k -— IT (T
2
CkCp Ck 4- Cp
(5)
where Ck and Cp are the microcanonical heat capacities calculated for the most probable energy partition.
214
As shown in Fig. 2 when Cp diverges and then becomes negative, cr\ remains positive but overcomes the canonical expectation
^
+C
r=T^I
<6>
Fig. 2 shows that the heat capacity extracted from the kinetic energy fluctuations is in very good agreement with the exact one. This means that kinetic energy fluctuations are an experimentally accessible measure of the heat capacity which allows to sign divergences and negative branches characteristic of the phase transition. Examples of experimental use of the proposed signal of the phase transition are given in Refs. u>10. The important point to notice is that now the study of the nuclear phase transition is becoming quantitative. References 1. 2. 3. 4. 5. 6. 7. 8. 9. 10. 11. 12.
A. Guarnera, PhD thesis GANIL Caen (1996). D. H. E. Gross, Phys. Rep. 279 (1997) 119; cond-mat/9911257. F. Gulminelli and P. Chomaz, Phys. Rev. Lett. 82 (1999) 1402. M. Schmidt et al, Nature 393 (1998) 238 and to be published. M. F. Rivet et al., Phys. Lett. B 388 (1996) 219; B. Borderie et al.,Phys. Lett. B 388 (1996) 224. J. Pochodzalla et al., Phys.Rev.Lett. 75 (1995) 1040. F. Gulminelli et al., Europhys. Lett., 50 (2000)434. P. Chomaz and F. Gulminelli, Nucl. Phys. A647, 153 (1999). P. Chomaz, F. Gulminelli and V. Duflot, Phys. Rev. Lett. (2000) in press. M. D'Agostino et al., Nucl.Phys. A 650 (1999) 329 and Phys.Lett.B 473 (2000) 219. R. Bougault et al, cont. to the XXXVIII Winter Meeting on Nucl. Phys. Bormio (Italy), Ed. I. Iori. J. L. Lebowitz et al., Phys.Rev. 153 (1967) 250.
Finite nuclear fragmenting systems: an experimental evidence of a first order liquid gas phase transition M. D'AGOSTINO ( 1 ) , F. GULMINELLI<2>, PH. CHOMAZ ( 3 \ M. BRUNO' 1 ', F. CANNATA^ 1 ', N. LE NEINDRE' 2 ' 1 ', R. BOUGAULT (2) , M. L. FIANDRI ( 1 ) , E. FUSCHINI (1) , F. GRAMEGNA' 4 ', I. IORI ( 5 ) , G. V. MARGAGLIOTTI< 6 ', A. MORONI<5>, G. VANNINI (1) , E. VERONDINI (1) (1) Dipartimento di Fisica and INFN, Bologna, Italy (2) IPC Caen (IN2P3-CNRS/ISMRA et Universite), F-14050 Caen Cedex, France (3) GANIL (DSM-CEA/IN2P3-CNRS), B.P.5027, F-14021 Caen Cedex, France (4) INFN Laboratorio Nazionale di Legnaro, Italy (5) Dipartimento di Fisica and INFN, Milano, Italy (6) Dipartimento di Fisica and INFN, Trieste, Italy An experimental observation of negative heat capacity is inferred from the event by event study of partial energy fluctuations in Quasi-Projectile sources, formed in Au + Au peripheral collisions at 35 A.MeV. Between about 2.5 and 5 A.MeV excitation energy, the system undergoes a first order phase transition.
1
Introduction
Since the first a t t e m p t s 1 , 2 to use Heavy Ion reactions to construct the caloric curve T(E), which links the temperature to the excitation energy of the nuclear fragmenting systems, many interpretations have been given about the order of the phase transition of nuclear m a t t e r , by looking for the presence or the absence of a plateau. Moreover the backbending (S-Shape) of the caloric curve, p r e d i c t e d 3 for a finite system undergoing a first-order phase transition, has never been experimentally observed, due in part to the experimental uncertainties of the t e m p e r a t u r e measurements, b u t also to the dependence of the caloric curve T(E) upon the specific thermodynamical transformation which the system undergoes from one energy state to another. Experimentally we do not know this transformation (isochore, isobar, isonothing, ...) and thus the observation of a plateau or a monotonous increase of the caloric curve is not significant for the identification of the transition and of its order. It has recently been shown 5 t h a t for a given total energy the average partial energy stored in a subsystem of the microcanonical ensemble is a good thermometer while the fluctuations associated to the partial energy can be used to evaluate the heat capacity. An example of such a decomposition is given by the kinetic E^ and the potential Ep energies 6 . In particular first order phase 215
216
transitions are marked by poles and negative heat capacities , corresponding to fluctuations anomalously large than the canonical expectation of. > CkT2. Indeed 5 :
c Ck + c
^
^ = c^k^
A single measurement of anomalous fluctuations at a given energy E is then an unambiguous proof of a first order phase transition.
2
Experimental details
The experiment was performed at the K1200-NSCL Cyclotron of the Michigan State University. The MULTICS and MINIBALL arrays were coupled to measure charged products, with a geometric acceptance greater than 87% of 4ir. Peripheral collisions of a predominantly binary character were selected for the reaction Au + Au 35 A.MeV, by requiring the velocity of the largest fragment in each event to be at least 75% of the beam velocity. The source was reconstructed event by event. For a detailed description of the data selection see Ref.4. Only events with a total source charge within 10% of the Au charge were analyzed. The excitation energy of the source, reconstructed from calorimetry, on an event by event basis 4 , ranges from about 1 to 8 A.MeV. This allows a sorting of the events as a function of energy as in the microcanonical statistical ensemble. With our selections, only « 30% of the total reaction cross section is studied. It corresponds to peripheral collisions showing predominantly a binary character. The contamination from other sources with respect to the QP resulted to be very small 4 . Indeed for heavy symmetric systems at a relatively low beam energy the important Coulomb repulsion in the entrance channel seems to preserve an essentially binary character for peripheral collisions, better than in the case of lighter systems at higher beam energies 8 . The isotropic emission of the source was verified by looking at the fragment angular distributions 6 . The source equilibration is also confirmed 2 ' 4 ) by the very good reproduction of the experimental charge partitions by the statistical model SMM 9 . Possible non thermal contributions 4 to the experimental excitation energies (removed in the thermostatistical analysis) were found limited to about 0.6 A.MeV at the most central reduced impact parameter here considered (0.5), where the average calorimetric excitation energy is 6.2 A.MeV.
(1)
217
3
Extraction of the heat capacity
In order to apply eq.(l), one has to decompose the total energy into the kinetic and potential terms Ek and Ep. However, in the nuclear case fragments at the freeze-out can be hot, while the detected fragments are collected at infinity after secondary de-excitation, i.e. with smaller size. In addition the asymptotic kinetic energies have to be corrected from the Coulomb boost. Primary partitions were therefore reconstructed 6 by following two extreme freeze-out (f.o.) hypotheses 3,9 . In the first one [cold fragments hypothesis) all charged products are considered as primary. Primary fragments (Z > 3) de-excite only through neutron evaporation. In the second assumption (hot fragments hypothesis) also light charged particles are emitted by primary fragments. The Coulomb energy has been calculated by randomly positioning non overlapping spherical primary products in the f.o. volume 1 0 . To apply eq.(l) we estimated the microcanonical temperature T of the system from the reconstructed average kinetic energy, through a kinetic equation of state 6 , where the internal excitation energy of primary products, as well as the thermal motion are accounted for. ^60
J
' 1 ' ' ' 1
' ' 1 ' i 1'
1 | l,l
\
•
z± 40 o
l |
it
]
| ' ' '-
\
Q.20
o
"
o
**Mf"
\h -
£-20
-40 -60
*
:
»' ~
7, i . , , i
• . 1 . . . 1
2
4
6
2
f
~
i , , 7-
4 , 6 C/Ao (A.MeV)
Figure 1: Ct per nucleon of the QP source 6 . The grey contour indicates the confidence region. Left panel: cold fragments hypothesis, right panel: hot fragments one.
By applying these f.o. reconstructions and eq.(l), we observed 6 the heat capacity of Fig. 1, showing two divergences and a negative branch in between. This points to a 1-st order liquid-gas phase transition, the distance between the poles being associated with the latent heat. This specific feature of 1-st order phase transitions in finite microcanonical systems 7 was reported for the first time in Ref. 6 and has been observed almost at the same time in the melting transition of sodium clusters 11 . The persistence of the negative heat capacity against the change of the f.o. hypothesis may look surprising, if one considers that the caloric curves T(E*) are appreciably different in the two f.o. hypotheses (Fig. 2). The fluctuation observable seems in this sense more robust than the caloric curve. This is due to the fact that, even if the (model dependent) temperature in our analysis acts
218
as a normalization factor for the fluctuations (eq.(l)), the reference fluctuation scale (Ck) and the fluctuation itself (
Improved freeze-out reconstruction
In Ref. 1 3 a sophisticated velocity correlation analysis was performed on central fragmentation events. For a source with size similar to the one here considered and with thermal excitation energy from 5 to 7 A.MeV, the percentage of light charged particles, emitted after the f.o. stage, resulted 40% and 30%, respectively. Correspondingly the excitation energy of primary fragments increases from 2.8 to about 3.8 A.MeV. 1
1
1
1
1
>0) 44
< 2 --
O o#° 0 oooO^°
- o°° 1 i 1 1 1 1 1 0 w 0 2 4 6 8 E*/Ao (A.MeV)
2 4 6 .8 . E'/Ao (A.MeV)
Figure 2: Left panel: Excitation energy of primary fragments (open symbols) for the QP source and values of Ref. (stars). Right panel: Caloric curve for the QP source (open symbols) and HeLi temperature of Ref.2 (stars). Dashed and solid lines represent T(E) for the hot and cold fragments hypotheses, respectively.
This i m p o r t a n t experimental information allows us to fix the parameters governing the reconstruction of the freeze-out. Indeed, once we establish the primary product multiplicity, only the volume is still a free parameter. Because
219
of the microcanonical constraint, larger volumes (lower Coulomb energies) lead to higher excitation energy of primary fragments than smaller volumes. By requiring that e*IMF for our QP events fall in top of the values of Ref.13 (Fig. 2 left panel) we come to a f.o. volume about 3 — 4 Vo 10- Another result of this refined f.o. reconstruction, is that the value of the microcanonical temperature T turns out in very good agreement with the experimental isotope temperature of Ref.2 (Fig. 2 right panel).
1
" n i h
0 2.5 5 7.5
0
2
i
i
1
i
4
E*/Ao (A.MeV) Figure 3: Left panel: Ct/Ao for QP events (improved f.o. reconstruction). Right panel: Ct/Ao for QP events (full points) and for central collisions (open point Au + C 25 A.MeV, open square Au + Cu 25 A.MeV, open triangle Au + Cu 35 A.MeV).
We show in the left side of Fig. 3 the heat capacity per nucleon of the QP source, evaluated with the improved f.o. reconstruction, for a f.o. volume 3Vb- With respect to Fig. 1 we observe that, while the position of the first divergence is nearly unchanged, the second one falls at about 6 A.Mev, close to other experimental results 1 4 . In the right panel of Fig. 3 we show Ct/Ao, obtained for the most central 10% of the events in the reactions Au + C 25 A.MeV, Au + Cu 25 A.MeV and Au + Cu 35 A.MeV, measured with the same apparatus. For these reactions we have analyzed only events where the charge of the source is within 10% the total available charge. This analysis is a strong confirmation of the results we have obtained for the QP source. 5
Conclusions
Our experiment presents the first evidence of a negative heat capacity in finite physical systems undergoing a first order phase transition. Fragment partitions of the QP source 4 show signals of a critical behavior at about 4.5 A.MeV excitation energy, which falls in the energy region where the heat capacity
220 shows the negative branch. The appearance of a negative branch in the heat capacity is very robust and the divergences are not artificially induced by experimental inefficiencies or reconstruction hypotheses. T h e transition point from the liquid side for this system is quantitatively established; further work and d a t a are needed to reach the same conclusion on the vapor side 1 4 . T h e only parameter t h a t can affect the distance between the two divergences is the freeze-out volume. We have shown here how to solve this ambiguity, by using new experimental information 1 3 . The authors would like to acknowledge the Multics-Miniball collaboration, which has performed the experiments. This work has been partially supported by grants of the Italian Ministry of University and Scientific and Technological Research (contract MURST-COFIN99). References 1. J. Pochodzalla et al, Phys. Rev. Lett. 75 (1995) 1040; J. A. Hauger et al, Phys. Rev. C 5 7 (1998) 764; Y. G. M a et al, Phys. Lett. C 5 B 3 9 9 (1997) 4 1 . 2. P. M. Milazzo et al., Phys. Rev. C 5 8 (1998) 953. 3. D. H. E. Gross, Rep. Prog. Phys. 5 3 (1990) 605; Phys. Rep. 2 7 9 (1997) 119. 4. M. D'Agostino et al, Nucl. Phys. A 6 5 0 (1999) 329. 5. P h . Chomaz and F . Gulminelli, Nucl. Phys. A 6 4 7 (1999) 153. 6. M. D'Agostino et al, Phys. L e t t . B 4 7 3 (2000) 219. 7. D.H.E. Gross et al., Zeit. fur Phys. D 3 9 (1997) 75. 8. C. P. Montoya et al., Phys. Rev. Lett. 7 3 (1994) 3070; J. F . Dempsey et al, Phys. Rev. C 5 4 (1996) 710; J. Lukasik et al., Phys. Rev. C 5 5 (1997) 1906; T. Lefort et al., Nucl. Phys. A 6 6 2 (2000) 397. 9. J. P. Bondorf, A. S. Botvina, A. S. Iljinov, I. N. Mishustin, K. Sneppen, Phys. Rep. 2 5 7 (1995) 133. 10. Spherical fragments are randomly positioned in the f.o. volume in such a way t h a t their centers fall into the freeze-out sphere. 11. H. Haberland et al., to be published in Nature and contribution to this conference. 12. M. D'Agostino et al, XXXVIII Int. Winter Meeting on Nucl. Phys., Bormio, J a n u a r y 2000, ed. by I. Iori, Ric. Sc. and E. P. 1 1 6 (2000) 386. 13. N. Marie et al., Phys. Rev. C 5 8 (1998) 256; S. Hudan et al, XXXVIII Int. Winter Meeting on Nucl. Phys., Bormio, J a n u a r y 2000, ed. by I. Iori, Ric. Sc. and E. P. 116 (2000) 443. 14. N. Le Neindre, contribution to this conference.
P H A S E T R A N S I T I O N IN X E + SN C E N T R A L E V E N T S B E T W E E N 32 A N D 50 A . M E V N. LE N E I N D R E COLLABORATION4
3|1
A N D R. B O U G A U L T 1 A N D T H E INDRA AND F. GULMINELLI \ PH. CHOMAZ 2
1
2 LPC (Caen-France), GANIL (Caen-France) Dipartimento di Fisica and INFN (Bologna-Italy) See the M.F. Rivet's proceeding to have the list of members 3
4
A study based on energy fluctuations was performed to calculate the total heat capacity of hot nuclei. The observation of a divergence of the total heat capacity and a negative branch provides the signature of a first order liquid gas phase transition in nuclear matter.
1
Introduction
This work is focussed on the thermodynamical aspect of multifragmentation and more specifically on the observation of a phase transition in finite nuclei. For this goal a set of Xe + Sn central events obtained with a beam energy between 32 and 50 A.MeV has been selected using the 47r INDRA array. These events lead to equilibrated sources of hot and dense nuclear matter at various excitation energies in the center of mass. The equilibration was checked by comparison with the SMM 1 model by performing a backtracing analysis 11 . An example of this analysis at 50 A.MeV for different static (Fig. la-b-c) and dynamic (Fig. Id) observables is shown on the Fig. 1. The good agreement between model and data allows us to speak about a thermalised source of nuclear matter composed of the greatest part of the two colliding nuclei. To get the agreement of the Fig. Id it's necessary to add a radial collective energy into the SMM sources. This collective energy is interpreted as the result of compressional effects occuring during the collision 2 ' 11 > 12 . The mean value of the collective energy is 0.8 A.MeV at 32 A.MeV up to 2.1 A.MeV at 50 A.MeV beam energy. As far as equilibrium is reached we can now proceed to the study of a signal based on energy fluctuations 3 to calculate the total heat capacity of the system and to observe its dependence on the beam energy. 2
Heat capacity
The method used to calculate the total heat capacity consists in dividing the energy of the system into two ensembles, a kinetic part and a potential part 221
222
Figure 1. data.
0
10
20
0
25
50
30 40 Charge Z
O S
75 100 Zbound
10 IS 20 Multiplicity Z>3
20 Charge Z
Results of a backtracing analysis comparing SMM calculations with
experimental
and then calculating the fluctuations of kinetic energy a^ event by event. This method is described in detail in differents works 3 - 7 ' 11 . In the coexistence region of a liquid-gas phase transition fluctuations are expected to be maximum. In particular the kinetic energy fluctuations Ok are related to the total heat capacity ct by the formula : 2 ct = — % r Cfc &
(1)
In order to calculate the fluctuations at the break up of the source we need to know the primary partitions, at the freeze-out. In fact in the detectors we detect the products only after secondary de-excitation. Moreover a part of the (potential) Coulomb energy is transformed into kinetic energy. In order to trace time back to the break up stage, two extreme hypothesis have been assumed the "Cold fragment" and "Hot fragment" hypothesis. In the first one we suppose that all light charge particles are already present at the "freeze-out" and hot primary fragments emit only neutrons. In the second hypothesis we put all neutrons and light charge particles in the prefragments. Moreover we place all the hot prefragments in a volume to calculate the
223
coulomb energy. To properly account for the finite range of the nuclear interaction we have used a standard 1 definition for the freeze-out volume which monotonously increases with the multiplicity of charged products. An additional difficulty counts in the fact that in these reactions we are in presence of collective energy. To calculate the thermal and only thermal kinetic energy fluctuations we need to correct for these effects. The importance of this correction is demonstrated in Fig. 2 which shows a simulation generated with the SMM model 1 in a range of excitation energy between 4 and 13 A.MeV and with a flow E c o t = l , 2, 3 and 4 A.MeV. If a collective component is included in the kinetic energy, not only the total excitation energy is overestimated but the partial energy fluctuations are analitically suppressed. This keep being true if fluctuations are normalised by the estimation of the temperature as required by equation 1 (upper part of the Fig. 2). Since in the middle of the coexistence region fluctuations are abnormally large 3 , even the inclusion of a fluctuating flow has the effect of decreasing the height of the peak of Fig. 2 11 . On the other side, if an average flow value is subtracted event by event the genuine value of the fluctuations is approximatively restored (lower part of Fig. 2). 3
Experimental results
The experimental results are presented in the Fig. 3. The columns refer to the four incident beam energies. The two first rows concern the "Cold fragment" hypothesis with and without corrections and the two last rows the "Hot fragment" hypothesis. The full points refer to the kinetic heat capacity Cfcand the open one to the kinetic energy fluctuations. The crossing of the two curves indicates a divergence for the total heat capacity of the system at the first order phase transition point (see Eq. 1). The transition signal, from coexistence at low energy (negative heat capacity) to a pure vapor phase at high energy (positive heat capacity) is already visible without correcting for the radial motion, however the correction is necessary to quantitative location of the transition point. Figure 4 represents, for the "Hot fragment" hypothesis, the experimental results concerning the value of the total heat capacity from Eq. 1 after correction on the mean value of collective energy E0O(. The curves indicated by the stars represent the total excitation energy distribution of the source after correction on EC0(. The open circles refer to the total heat capacity and the shadowed area represents a confidence interval. A divergence in the heat capacity and a clear negative branch is present at 32 and 39 A.MeV indicating
224 1
S, 9
SMM without E, SMM witti E,
s-
' I ' ' ' ' I • ' ' M SMM without r ^ , ^ OO SMM with E_
10 IS E*/A0 (A.MeV) A
3-
7T-I
|
; •
SMM without E ^ , ^ ^
~ O
SMM corrected E ^ true
1 1 I
1 |
1 P 1 1 .
• • SMM without ECQBKtfn CO SMM corrected E_, true •
•§•20
*» 10 5
- \
i
^
H
^
\
VJ
: . . ,5 . . . 7T**».: 10 (E*/A0 thermal),
Figure 2.
5 10 (E*/A0 thermal),,
Influence of collective energy of the fluctuation observable (see text)
the transition between the coexistence zone and the gas phase. We remember that we are dealing with highly energetic centrals events and we are therefore not able to see the first divergence between the liquid zone and the coexistence region 5 . 4
Conclusion
The INDRA ATT multidetector has allowed us to select among all the Xe + Sn collisions between 32 and 50 MeV per nucleon events leading to the formation of unique source of hot nuclear matter. These events are compatible with the hypothesis of thermodynamical equilibrium. For this reason they are a good basis for studying phase transition in nuclei. We have used a method based on thermal energy fluctuations to sign the phase transition. We have measured the effects due to the presence of a collective radial energy and we have been able to correct them. We have seen a clear signal of a first order phase transition in experimental Xe + Sn central events. For more details about this analysis and others related works see: 2 . 3 . 4 . 5 . 6 . 7 . 8 . 9 . 1 °. 1 i
225
10
0
E* (A.MeV)
10 E' (A.MeV)
Figure 3. Kinetic heat capacity Cfc (open circles) and kinetic energy fluctuations cr^/T^ (full circles). Comparison of the two methods of freeze-out reconstruction, "Hot fragment" and "Cold fragment", with and without correction on Ecoi and for four incident beam energies. N. Le Neindre thanks the Istituto Italy) for hospitality and support.
Nazionale
di Fisica Nucleare
(Bologna
section,
References 1. J . P Bondorf a n d A . S . B o t v i n a a n d A . S . Iljinov a n d I.N. M i s h u s t i n P h y s . R e p 257 (1995) 2. R. B o u g a u l t et al. X X X V I I I i n t e r n a t i o n a l winter m e e t i n g B o r m i o 2000 (Italy) Ed. by I. Iori 3. P. C h o m a z et al. Nuclear Physics A 647 (1999)
226
E"-E^,(AJVleV)
Figure 4. Total heat capacity and excitation events for four beam energy
E -Em„ (AJMeV)
energy distribution for the Xe + Sn central
4. P. Chomaz et al. XXXVIII international winter meeting Bormio (Italy) Ed. by I. Iori 5. P. Chomaz and M. D'Agostino contribution to this conference 6. M. D'Agostino et al. Nuclear Physics A650 (1999) 7. M. D'Agostino et al. Physics Letters B473 (2000) 8. M. D'Agostino et al. XXXVIII international winter meeting Bormio 2000 (Italy) Ed. by I. Iori 9. F. Gulminelli et al. XXVII International Worshop on Gross Properties of Nuclei and Nuclear Exitations Hirschegg 1999 (Austria) 10. F. Gulminelli et al. XXXVIII international winter meeting Bormio 2000 11. N. Le Neindre These de Doctorat de l'Universite de Caen (1999) 12. N. Marie et al. Phys. Letters B391 (1997)
N U C L E A R CALORIC CURVE: I N F L U E N C E OF T H E S E C O N D A R Y DECAYS ON T H E ISOTOPIC THERMOMETERS AL. H. RADUTA AND AD. R. RADUTA National Institute of Physics and Nuclear Engineering, Bucharest POB-MG 6, Romania The sharp microcanonical multifragmentation model from [Al. H. Raduta and Ad. R. Raduta, Phys. Rev. C 55, 1344 (1997); ibid. 56, 2059 (1997)] is refined and improved by taking into account the experimental discrete levels for fragments with A < 6 and by including the stage of secondary decays. The improved version of the model is employed for evaluating the nuclear caloric curve predictions of 17 isotopic thermometers for three representative nuclei in both primary and asymptotic stages. Effects of the secondary decays on the primary decay caloric curves are evidenced and discussed. A procedure for calibrating the isotopic thermometers on the microcanonical predictions for both primary decay and asymptotic stages is proposed. A set of calibrating parameters corresponding to each of the considered isotopic thermometers, independent on the source size, on its excitation energy and, in the case of the primary decay, on the freeze-out radius assumption is obtained.
Nuclear caloric curve is of central interest in the nuclear multifragmentation studies, one of the main reasons being that it can serve to identify a possible liquid-gas phase transition in nuclear matter. While statistical multifragmentation models have predicted transition regions in the caloric curve since 1985, the first experimental nuclear caloric curve evaluation was reported by the ALADIN collaboration ten years latter 1 . A wide plateau-like region interrupting the caloric curve suggesting the existence of a liquid-gas phase transition was evidenced with that occasion. Various experiments followed obtaining more or less different results. The Albergo garandcanonical isotopic thermometers have been employed in these experiments. Which is the reason for such different experimental results is a naturally raising question. Besides the experimental uncertainties in separating the pre-equilibrium particles one has to deal with two inherent drawbacks of the Albergo thermometers. The first drawback is fundamental: since they are deduced on grandcanonical grounds the isotopic thermometers do not account for finite size effects (appearing in nuclear systems). The second one is again of experimental nature: secondary decays modify the isotopic yields entering the Albergo formula leading to results deviated from the primary break-up ones. A remedy for these problems would be a sharp microcanonical calibration of the isotopic thermometers in both primary break-up and asymptotic stages. 227
228
(70,31)
J i
''il'' $
$
$
jjiiiiit* 1 ^ 1
i'g I i i 3 s i 10
• . * ° . . »ff! *»f
rftfiiriifofr
15
0
10
is
kitilf111111 i i:
'^I'Tt t i iiiSissai 10
(190,79)
i-f i s 5 • • • E„(MoV/nuclBon)
15
f
,1
i "•uC/"He 0 •*"C/ M H* * "•"B/"He
?
15
r«i"l Ti iilli=is
•Aftfltti*
5 10 E „ (MeV/nucleon)
E.^MeV/nuctaon)
15
.iitnTi i n ^ ^ S 10 E „ (MeV/nucteon)
15
Figure 1. Caloric curves corresponding to 17 isotopic thermometers, evaluated for 3 nuclear sources with the freeze-out radius 2.5 A 1 / 3 fm in both primary decay (columns 1 and 3) and asymptotic stages (columns 2 and 4). Microcanonical caloric curves are represented by dashed lines.
In the present work we use the sharp microcanonical model from Ref.2 (improved by taking into account in the primary break-up the experimental discrete levels for fragments with A < 6 and by including the stage of secondary decays 3 ) in order to perform a calibration of 17 isotopic thermometers ( 6 ' 7 Li/ 3 ' 4 He, 7 ' 8 Li/ 3 ' 4 He, 8 - 9 Li/ 3 - 4 He, 12 - 13 C/ 3 ' 4 He, n - 1 2 B/ 3 ' 4 He, 12,13B/3,4He) l,2H/3,4He) 2 , 3 ^ 3 , 4 ^ 13,14C/3.4Hei 6 , 7 ^ 1 1 , 1 2 ^ 8,9Li/ll,12C) 7,8Li/ll,12Cj 9,10Be/ll,12C; 12,13C/H,12C; 13,14(^11,12^ ll,12B/ll,12Cj 12 13
' B/ U ' 1 2 C) in both primary decay and asymptotic stages. This model has already proven the ability of reproducing the experimental HeLi caloric curve reported by the ALADIN collaboration in 19974. In the microcanonical model nuclear temperature can be evaluated starting from its statistical definition:
T =
((3N/2-5/2)/K)~1
(1)
The well known Albergo isotopic temperature formula, widely used in experimental measurements of nuclear temperature, writes: r1234 = A B / l n [ S
(YJY2)/(Y3/Y4)]
(2)
229
^ , 1 " '
.*"
..,'111
«** , * • > • '
10
IS
u
0
10
..M
,.!^
...^
15
r 10
15
"0
10
*"C/"He
; "u/"-"c
l§
..M1
"«'"»• . . - i - »
-.**•* I E.^MoWnuolaon)
"U/"'"C J Li/ I M , C1 '•"Be/"-' C ,, 1 - u * C/" "C *"C/"-"C
-t-«
,.*•••
" *****
if
J S 10 E „ (MeV/nucleon)
15
w
*
s
15 "C
* ^* (190, 79)
0
Figure 2. The isotopic caloric curves from Fig. 1 adjusted as in eq. (3) using the parameters from Table 1.
where AB — (Bi — B2) — (B3 — _B4), Bt and Y{ are respectively the binding energy and the yield of the isotope i (= 1, 2, 3, 4), s is a statistical factor deduced from the ground state spin and masses of the isotopes and the isotopes (1,2) and (3,4) must differ by the same number of neutrons and/or protons. Isotopic caloric curves evaluations are presented in Fig. 1 for 17 isotopic thermometers in both primary decay (columns 1 and 3) and asymptotic (columns 2 and 4) stages together with the microcanonical curve, for the case of three representative nuclei (70,32), (130,54) and (190,79) and a freeze-out radius parameter 2.5A 1 / 3 fm. In all cases the isotopic caloric curves present a dispersive character increasing with the increase of the source excitation energy. Remarkably, in the asymptotic stage, this character is drastically reduced. In order to calibrate eq. (2) on the microcanonical results we fit the ratios Tmicro/Tiso(Eex) with first (thermometers from Fig. 1, column 1) and second (thermometers from Fig. 1, column 3) order polynomials: fT(Eex) = aEex + b ; / T ( - E C X ) = aE\x + bEex + c. The fitting parameters are given in Table 1. In order to verify the obtained parametrization we represent the caloric curves from Fig. 1 adjusted using the parameters from Table 1 according to
230 therm. /3'4He 6 7 ' Li 7 8 ' Li 8 9 ' Li 12,13/-i 11,12 B 12,13g 1,2 H 2,3H 13,14 C
/U'12C 6 7 ' Li 8 9 - Li 7 8 - Li 9 10 ' Be 12,13/-i 13,14 C 11,12 B 12,13g
primary decay stage a b c 0.0165 1.3284 0.0025 1.0040 -0.0117 0.9164 0.0168 1.0615 -0.0130 0.8803 0.0000 1.0232 0.0111 1.0049 0.0134 1.0048 -6.7680e-4 -2.3336e-2 6.4431e-l a b c 2.0414e-3 3.0558e-2 1.9313 5.3155e-4 1.2332e-2 1.1823 2.0480e-3 -1.3545e-3 1.4271 2.1179e-4 -2.7342e-2 9.6059e-l 1.7500e-3 2.3999e-2 1.5584 6.0614e-4 -2.6735e-3 1.1527 1.0410e-3 -9.4635e-4 1.2724 2.1379e-3 2.7530e-3 1.5116
asymptotic stage a b c -0.0044 1.4923 0.0430 1.2332 -0.0050 1.0815 0.0248 1.3335 0.0164 1.1330 0.0180 1.2546 -0.0023 1.3180 0.0006 1.2351 -2.5222e-3 2.3087e-2 8.1780e-l a b c -1.3887e-4 4.3266e-2 2.0042 -1.1145e-3 4.5556e-2 1.2936 1.8077e-3 4.7329e-2 1.6656 -1.4423e-3 -3.7369e-4 1.2043 -2.3857e-4 6.9886e-2 1.7521 -1.5319e-3 5.2393e-2 1.3460 1.3799e-3 2.7188e-2 1.5801 -8.4445e-5 5.2134e-2 1.7578
Table 1. Calibrating parameters evaluated for both primary decay and asymptotic stages corresponding to 17 isotopic thermometers. The corresponding units are: a [u/MeV], b [1] - for f(Eex) = aEex + b and a [u/MeV 2 ], b [u/MeV], c [1] - for f(Eex) = aE^x + bEex + c
the relation: Ti°or = TiSo fr(Eex)-
(3)
The result is presented in Fig. 2. As one can see, the corrected caloric curves overlap very well each other and also the microcanonical curves. Remarkably the obtained parametrization is independent on the dimension of the source, on the excitation energy and, as shown in Ref.5, on the freeze-out radius assumption - in the primary decay stage. References 1. J. Pochodzalla et al., Phys. Rev. Lett. 75, 1040 (1995). 2. Al. H. Raduta and Ad. R. Raduta, Phys. Rev. C 55, 1344 (1997); ibid. 56, 2059 (1997). 3. Al. H. Raduta and Ad. R. Raduta, Phys. Rev. C 6 1 , 034611 (2000). 4. Hongfei Xi et al, Z. Phys. A 359, 397 (1997). 5. Al. H. Raduta and Ad. R. Raduta, Nucl. Phys.A671, 609 (2000).
P H A S E T R A N S I T I O N SIGNALS IN THERMALLY EXCITED NUCLEI
Department
V . E . V I O L A for t h e E900 Collaboration" of Chemistry and IUCF, Indiana University, Bloomington, E-mail: [email protected]
IN
47405
5-15 GeV proton-, pion- and antiproton-induced reactions with 1 9 7 A u have been studied with the Indiana Silicon Sphere (ISiS) charged-particle detector array at the Brookhaven AGS accelerator. Using reconstructed values of E*/A as as sorting variable, the IMF kinetic energy spectra suggest a weak extra thermal expansion energy above E*/A W 5 MeV. IMF-IMF correlation functions show an order-ofmagnitude decrease in the relative emission time in the interval E*/A X 2-5 MeV. Above this value, a constant limiting time of 20-50 fm/c is found.
1
Introduction
Two key questions relevant to the possible link between multifragmentation reactions and a nuclear liquid-gas phase transition are those of expansion and time scale. For this purpose, GeV hadron projectiles provide the unique advantage of producing only a single target-like source that on average moves slowly in the lab system (v S 0 1 i r c e ^ 0.01c) and emits multiple particles nearly isotropically. Transport codes1'2 indicate that compressional and rotational effects are minimal and that energy deposition occurs very rapidly (r = 30 fm/c). Measurements were performed with the ISiS charged-particle detector array3 at the Brookhaven National Laboratory AGS accelerator (E900). Beams of 515 GeV/c protons, TT~ and antiprotons were incident on a 197 Au target. Measurements at 8 GeV/c were performed with a tagged negative beam to identify p and n~. For each beam the total number of events that met the trigger level of three or more charged particles was greater than N > 106, except for antiprotons, where 24,000 events were obtained. Further details are found in4'5. 2 Distributions of E * / A In Fig. 1 the E/A* distributions are shown for several systems. The qualitative systematics of the results are consistent with the prediction of transport codes; i.e., antiprotons produce the highest excitation energies, protons and pions behave nearly identically, and above beam momenta of about 6 GeV/c one obtains little additional excitation energy5. The trend in E*/A between 5.0 a T . Lefort, L. Beaulieu, K. Kwiatkowski, R.T. de Souza, W.-c. Hsi, Indiana Univ.; R. Laforest, E. Martin, E. Ramakrishnan, D. Rowland, A. Ruangma, E. Winchester, S.J. Yennello, Texas A&M Univ.; L. Pienkowski, Warsaw Univ.; R.G. Korteling, Simon Fraser Univ.; A. Botvina, Inst, for Nucl. Research; H. Breuer, Univ. of Maryland; D. Durand, LPC de Caen
231
232
Figure 1: Probability distributions plotted as a function of E*/A of the residue. Systems are defined on figure. 2
4
6
8
10
12
E'/A (MeV)
G e V / c TT" a n d 14.6 G e V / c protons is accounted for primarily by a constant E*, accompanied by a decrease in the size of the thermal-like source with increasing b e a m m o m e n t u m , an artifact of the fast cascade stage of the reaction. Because the results for the 8.0 and 9.2 G e V / c n~ and 10.2 G e V / c protons are nearly identical, these three systems have been s u m m e d to produce a set of 7.4 x 10 6 events with at least one thermal Z > 2 fragment. Examination of experimental observables as a function of E * / A reveals several features t h a t meet the requirements of a phase-transition interpretation. First, for Z > 6 fragments the invariant cross sections in vy versus vx space are nearly isotropic in the laboratory frame, illustrating t h a t the system is at least randomized prior to breakup 6 . Second, below E * / A w 4 MeV the Coulomb-like peaks and spectral shapes of the I M F spectra strongly resemble those obtained in lower energy studies 7 . At higher E * / A , however, there is a distinct broadening of the peak towards lower energy, accompanied by flatter spectral slopes, consistent with the breakup of an e x p a n d e d / d i l u t e system. Third, the probability for observing three or more IMFs in one event?, increases rapidly above E * / A ~ 4 MeV and is the dominant decay mode above E * / A > 6 MeV (see below). Finally, the charge distributions evolve toward the formation of increasingly heavy fragments u p t o E * / A = 5-6 MeV, but above this value the trend reverses as the multiplicity grows. Related to the fragment size distribution, in Fig. 2 we show the E * / A dependence of the average charge for the three heaviest fragments, with the reconstructed heavy residue taken as the heaviest fragment, (Zmaxl). One observes t h a t at low E * / A the heaviest residue accounts for most of the charge (fission/evaporation), whereas the charges of the heaviest fragments all correspond to IMFs at high E * / A (multifragmentation).
233
E*/A(AMeV)
E*/A(AMeV)
Figure 2: Size of the three largest fragments per event (Z m oa:l = residue; Z m o a ; 2,3 = heaviest IMFs) as a function of E*/A. Lines are SMM prediction. Vertical line: 1% probability level.
3
Expansion and T i m e Scales
Studies of I M F kinetic energy spectra in heavy-ion-induced reactions have suggested the existence of a collective radial expansion energy, a t t r i b u t e d to the compression/decompression stage of the collision 9 . T h e SMM model 1 0 has served to account for the Coulomb and thermal contributions to the spectra in many such analyses. Since the decay of residues produced in hadron-induced reaction is dominated by thermal and Coulomb effects, the E900 d a t a provide an i m p o r t a n t calibration of this assumption. In the left panel of Fig. 3 the average fragment kinetic energies {EK} are plotted as a function of Z for bins of E * / A = 4-6 and 6-9 MeV. Little sensitivity of (EK) on E * / A is found. This behavior can be understood in terms of the excitation-energy-gated spectra for fragments in the right panel of Fig. 3. As E * / A increases, the Coulomb peak decreases, consistent with the expansion/dilution scenario, but the spectral slopes become flatter, as expected for increasing t e m p e r a t u r e , the effects offsetting one another. Adopting the same approach as in the heavy-ion work, the (EK) values, IMF multiplicities and charge distributions have been compared with several statistical models. Above E * / A > 4 MeV, the sequential model SIMON evaporation 1 2 is unable to reproduce the I M F multiplicities and charge distributions. In contrast the simultaneous SMM model with a breakup volume of V = 2-3 Vo, describes all three observables satisfactorily at the multifragmentation threshold, E * / A ~ 5 MeV. Similar results are found for SIMON-explosion. Comparison with the (EK) d a t a at E * / A = 4-6 MeV is shown by the dashed line in Fig. 3(left). Thus, these results show t h a t statistical models involving a low density source 1 0 ' 1 1 ' 1 2 ' 1 3 give a satisfactory account of thermal and Coulomb effects near the multifragmentation threshold. At higher E * / A , however, SMM underpredicts the averages (solid line in Fig. 3) although the multiplicities and
234 Figure 3: Left: {EK) compared with SMM model for E*/A > 4-6 MeV (open points, dashed line) and E*/A = 69 MeV (closed points, solid line). Right: Angle-integrated spectra of oxygen fragments for three E*/A bins.
: SMM <3V0) 92E7A>6
- - 62E*/A>4
Fragment charge
Energy (MeV)
charge distributions are still reasonable u p to E * / A = 8-9 MeV. Adding an extra radial expansion energy CR = 0.5A MeV improves the comparison with the (EK) d a t a for the highest E * / A bins for all Z > 5 fragments, consistent with the mass dependence expected for collective radial expansion. T h e extra expansion energy deduced from comparisons with SMM has been extracted as a function of E * / A , shown in the third frame of the final figure. We a t t r i b u t e this extra radial energy to thermal expansion effects associated with increasing t h e r m a l pressure in the multifragmenting system. However, the magnitude of this effect is small in comparison with radial expansion energies derived from heavy-ion studies, consistent with the interpretation t h a t the increased (EK) values observed in heavy-ion reactions are due to collective radial expansion, e.g., compression/decompression. A critical question in associating multifragmentation events with a nuclear phase transition is t h a t of time scale. To investigate this issue, I M F - I M F relative velocity correlations have been constructed 1 4 . This analysis has included 5 x 10 5 thermal-like events with ZJMF = 4-9 and observed multiplicity N / M F > 2. In order to estimate the relative emission times, the d a t a have been compared with the Coulomb-trajectory model of Glasmacher 1 5 . T h e initial system was defined by the reconstructed E*, A and Z values, as discussed in Section 2. Best-fit values required t h a t the model reproduce not only the correlation functions, but also the experimental I M F kinetic-energy spectra and charge distributions. These constraints imposed a charge separation distance consistent with a breakup volume to V pa 2-3 VoIn Fig. 4 the experimental correlation functions are compared with simulations. In the b o t t o m panel of Fig. 5 the relative emission times plotted as a function of E * / A . T h e width of the shaded area indicates space-time constraints imposed by the fitting procedure, the upper points responding to V = 2 Vo and the lower values to V = 3 Vo-
the are the cor-
Initially, the relative emission time is long, r = 400-600 f m / c at E * / A = 2-2.5 MeV, consistent with sequential decay of the system. W i t h increasing
235 Source charge 75
70
65
60
55
2 « H, S 10 U#/u EVA.-
I SGeV/cJT + ^ A l i
3
f •f
2
< ~^:... 20
40
V«(10-3c) E
Figure 4: Correlation functions for Z = 4-9 IMFs as a function of reduced velocity (open circles). D a t a gated on source E * / A = 2.0 - 2.5 MeV (top), 4.5 -5.5 MeV (center)and 8.5 - 8.5 MeV (bottom). Solid and dashed lines are results of Coulomb trajectory calculation for fit parameters indicated on figure.
250
0
1
2
3
4
5
6
7
8
9
10
11
Excitation energy (A MeV)
Figure 5: Dependence on E * / A for the following quantities, from bottom up: relative IMF emission time r, extra radial expansion energy E e i p , charge distribution power-law exponent tau, and probability for N IMFs.
excitation energy the time scale decreases rapidly to a value of r = 50-100 f m / c at E * / A = 4 MeV. At higher excitation energies a limiting time of 20-50 fm/c is found, comparable to the fluctuation time of the system 2 . 4 Summary Fig. 5 summarizes the evidence for a possible nuclear phase transition based on the present results. Up t o E * / A ~ 5 MeV (E* ~ 800 MeV), I M F multiplicities are low. Correspondingly, the emission t i m e scales are relatively long, the charge distributions evolve toward an increasingly heavy IMFs (power-law parameter tau in panel two), and no evidence for extra expansion is present. These features conform t o a sequential binary decay mechanism, t h a t accounts for 80-90% of all observed events. At E * / A = 4-6 MeV, there is a distinct change in all quantities. At these excitation energies, multifragmentation events become the dominant source of fragments, while the time scale reaches
236
its minimum value, as does the power law parameter (maximum fragment size). This transition region coincides with the threshold energy predicted by statistical models of a nuclear nuclear phase transition. As the highest excitation energies are approached, the signals of additional heat input to the system become apparent in the onset of extra thermal expansion energy, the dominant yield of MJMF > 3 events, and the decreasing size of the fragments (larger tau). In conclusion, the broad picture that emerges from the present studies is consistent with a phenomenon that involves the near-simultaneous breakup of a hot, expanded/dilute system above E*/A s=a 5 MeV. Such behavior is suggestive (but not proof) of a nuclear liquid-gas phase transition. This work was supported by the U.S. Department of Energy and National Science Foundation, the NSERC of Canada, the Polish State Comm. for Scientific Research and the Robert A. Welch Foundation. References 1. J. Cugnon et al. Nucl. Phys. A352, 505 (1981); J. Cugnon, Nucl. Phys. 462, 751 (1987). 2. G. Wang et al, Phys. Rev. C 57, R2786 (1998). 3. K. Kwiatkowski et al. Nucl. Instr. Meth. A 360, 571 (1995). 4. T. Lefort et aZ.Phys. Rev. Lett. 83, 4033 (1999). 5. L. Beaulieu et al. Phys. Lett. B 163, 159 (1999). 6. V.E. Viola et al., CRIS 2000, Catania, Italy, May 22, 2000 (ed. S. Costa). 7. J. Wile et al., Phys. Rev. C 45 2300 (1992). 8. J. P. Bondorf et al., Nucl. Phys. A443, 221 (1985). 9. W. Reisdorf et al., Nucl. Phys. A612, 493 (1997), R. Bougault et al., Proc. XXVII Intl. Wrksp. on Gross Prop, of Nucl. Matter (ed. H. Feldmeier, J. Knoll, W. Norenberg and J. Wambach)(1999); W.-c. Hsi, et al., Phys. Rev. Lett. 73, 3367 (1994). 10. A. Botvina, et al., Nucl. Phys. A507, 649 (1990). 11. D. Durand, Nucl. Phys. A541, 266 (1992). 12. W.A. Friedman, Phys. Rev. C 42, 667 (1990). 13. D.H.E. Gross, Rep. Prog. Phys. 53, 605 (1990). 14. L. Beaulieu et al., Phys. Rev. Lett. 84 5971 (2000). 15. T. Glasmacher et al., Phys. Rev. C 50, 952 (1994).
HYDROGEN CLUSTER MULTIFRAGMENTATION AND PERCOLATION MODELS F. GOBET, B. FARIZON, M. FARIZON, M.J. GAILLARD Institut de Physique Nucleaire de Lyon, IN2P3-CNRS UMR et Universite Claude Bernard, 43 boulevard du 11 Novembre 1918, F-69622 Villeurbanne Cedex, France J.R BUCHET AND M. CARRE Laboratoire de Spectrometrie Ionique et Moleculaire, CNRS UMR 5579 et Universit Claude Bernard, 43 boulevard du 11 Novembre 1918, F-69622 Villeurbanne Cedex, France P.SCHEIER AND T.D.MARK Institut fur Ionenphysik, Leopold Franzens Universitat, Technikerstr.25, Innsbruck, Austria
A-6020
Using a recently developed multi-coincidence technique, the cluster fragmentation resulting from a collision between a high energy H^~7 ion with a He target is investigated on an event by event basis. The data obtained are analysed in terms of fluctuations in the fragment size distribution. A comparison with a 3D lattice bond percolation with the same size (27 constituents) shows the presence of a critical behaviour in these finite systems. The importance of the hydrogen cluster's speficity is studied with a second lattice bond percolation of same size with two kinds of bond.
Fragmentation covers a wide range of phenomena in science and technology, including polymers, colloids, droplets, and rocks. Therefore fragmentation can be observed on different scales ranging from the femtometer scale in the case of nuclear decay reactions to astronomical dimensions for collisions between galaxies 1 . It has been recognized recently that some features feature of this phenomen are rather independent of the decaing systems and the underlying interaction forces2. Thus it is highly desirable to identify those features which are common to all kind of fragmentations independent of the size and interaction forces. This can be achieved by studying either experimentally or theoretically (using appropriate models such as percolation models) a specific decaying system in as much detail as possible and then comparing the various features obtained with other differently sized and bonded system. It has been shown recently that a particular useful system for this purpose are clusters of atoms or molecules 3 . The hydrogen cluster ions, as the simplest ionized molecular complexes, have recently attracted great interest and it was thus possible to clarify their 237
238
structure and other properties 4 ' 5 ' 6,7 . Their structure has been calculated to consist of a central H j core which is solvated by shells of unperturbed H2 molecules. One important feature of these cluster ions is the large difference between the strong intramolecular interaction (covalent bond with an energy between 4 to 5 eV) and the rather weak intermolecular interaction (dipole - dipole interaction with an energy close to 0.1 eV 6 ' 7 ) . Fragmentation of hydrogen cluster ions induced by collisions with either atomic helium or fullerenes has been studied using a recently developed multi-coincidence technique. With this technique it was possible to detect and to identify in terms of mass to charge ratio all charged fragments produced in a collision on an event by event basis (for more experimental details, see Refs.8'9 and references therein). Under various experimental conditions a power law fall-off in the fragment mass distribution, ie, p~ T , with p the fragment mass and r an exponent has been observed with r close to (i) the critical exponant 2.6 found in nuclear fragmentation experiments and close to (ii) predicted values (~2.23) using Fisher's droplet model 9 ' 10 ' 11 ' 12 ' 13 ' 14 . Indeed, whereas fragmentation induced by low energy deposition has been usually interpreted in the frame of the evaporative ensemble model 15 ' 16 , recently reported fragmentation patterns obtained in high energy cluster atom-collisions 17 ' 18 ' 19 and cluster multicharged ion collisions20 showing bimodal (U-shape) decay patterns (see Fig.l, curve 2) have been interpreted by the occurence of both evaporative cooling reactions and multifragmentation processes. Moreover, beam foil experiments with hydrogen cluster ions 21 have lead to the complete disintegration of the projectile yielding only atomic fragments (see Fig.l, curve 4); an intermediate regime characterized by a a fragment distribution consisting solely by a decreasing function in p _ r (dashed curve in Fig.l) between the bimodal one and the complete disintegration has been observed in the cluster-cluster collisions9. In addition the question about the existence of a critical behaviour in the fragmentation of hydrogen cluster ions has been addressed recently 22 . We report on a comparison between (i) experimental results obtained by fragmentation of H^7 cluster ions induced by collisions with a He target and (ii) the results obtained by a calculation using a lattice bond percolation model to simulate critical behaviour. We are considering here the dependence of the average size Pmax of the largest fragment produced in a single event on the total number of fragments, usually termed multiplicity m. Campi already showed that the shape of this dependence of Pmax on multiplicity m will be different in the presence or absence of a critical behaviour 23 . As the fluctuations in the fragment size distribution are largest near the critical point, it is interesting to also plot as a measure for these fluctuations the
239
Figure 1. Schematic representation of normalized fragmentation yield versus normalized fragment size p / n for different fragmentation regimes. (1) Evaporation from low energy collisions between H„ cluster ions and atoms after Reuss and Van Lumig 1 6 ;(2) bimodal regime (evaporation and multifragmentation) 1 7 ;(3) the multifragmentation regime observed in cluster-cluster collisons 9 ; and (4) complete disintegration from high energy beam foil collision experiment with cluster ions 2 1 .
standard deviation of Pmax versus the multiplicity and thus obtain a further characteristic fingerprint of the presence or absence of a critical behavior type of situation. In Fig.2 we have plotted these two types of correlations (i.e., Pmax versus m and the standard deviation of Pmaw versus m) for 29041 analyzed events resulting from the interaction between HJV and the He target (designated by open squares). It can be seen that in general the average size of the largest fragment Pmax decreases with increasing multiplicity m. The data show a distinct change of slope at around m =' 13. In contrast the standard deviation of Pmax versus multiplicity curve has a bell shaped form with a maximum peak value of 1.5 at m = 6. In order to simulate a critical behaviour in a finite system 12 we have used a 3-dimensional percolation model with the same number of sites in the lattice as the number of constituents in the hydrogen cluster ion studied, i.e., a cubic system with 27 sites (3*3*3). We have carried out these calculations for differing probabilities p between 0 and 1 in steps of 0.01 and by creating 10.000 "fragmentation events" for each step. The corresponding results in terms of Pmax versus multiplicity m and the standard deviation of Pmax versus multiplicity for this model calculations are plotted in Fig.2 and designated
240
30 25 20-
Jl5J Bo go
10 5 .1 . . . . I . .
0-
. .1 . . . . I . . . . I
.9.
ax 3.5-
I 3-
a M
0°
Or,
0.5M»»t + * l * l
0
5
10 15 20 Multiplicity, m
25
30
Figure 2. Comparison of the fragmentation behaviour for systems of same size 27 a)hydrogen cluster (open square) b)3D bond percolation lattice (open circle) c) 3D bond percolation lattice with two kinds of bond (open triangle). Upper part: Average size P m a x of the largest fragment produced in a single event versus the multiplicity m. Lower part: Standard deviation (fluctuation) of Pmax given in the upper part versus m. The normalized standard deviation is defined as \ / ( < Pmax 2 > — < Pmax > 2 ) .
as open circles. On a first sight there exists quite a good agreement in the general shape between the experimental data obtained in the H^-He collision experiment for Pmax versus m and the standard deviation of Pmax versus m and the calculated data from this percolation model. As mentioned above from this good agreement we can conclude and confirm (thereby extending the earlier conclusions based on the existence of power laws) the existence and presence of a critical behaviour in the fragmentation of hydrogen cluster ions. This is the more convincing as calculations performed with a simple one dimensional percolation model yield quite different results, e.g., not showing the resonance-like peak in the standard deviation function23 so characteristic for the presence of critical behavior. Nevertheless, as can be seen in Fig.2 there exists slight deviations of
241
the experimental data from the theoretical predictions by the 3-dimensional standard percolation model used. First of all, Pmax of the experimental data decreases faster with increasing m than the calculated points and in addition the calculated points do not show the clear change in slope exhibited by the experimental data. Moreover, the absolute magnitude of the standard deviation of Pmax is much smaller in the case of the experimental data (close to 1.5) than in the percolation model (close to 3.6). Thus we have developed a novel type of percolation model (including two types of bonds) in order to take into account the fact that there are these two kinds of bonds in the hydrogen cluster ions, i.e., modelling the decay of the H^7 cluster ion (which is equivalent to H3~(H2)i2) by a percolation lattice where the two differing bond strengths are characterized by two different probabilities Ps and Pw representing the two bond strengths, strong and weak. In Fig.2 we have plot the correlation between Pmax and multiplicity for the events generated from this two bonds lattice percolation model (open triangle). Firstly, a very good agreement for the decrease of Pmax with multiplicity is observed between H^-He collision experiment and this new percolation model. We can clearly see presence of a change of slope for multiplicity m=13. Secondly, a resonance-like peak is also observed in the standard deviation of Pmax with a maximum in the same order of magnitude as the one observed in the case of one H^7 experimental data (close to 1 for the new lattice bond percolation model and close to 1.5 for H^7 fragmentation). Nevertheless, the maximum of the peak of the standard deviation versus multiplicity is localized at different multiplicity (m=5 in the case of Hj 7 whereas m=13, m=14 in the case of two bonds percolation model). This difference is due to the existence of very minority events which appear in the experimental set of data but are completly absent in the percolation results. The correlations between multiplicity and average size of largest fragment in single event for a given multiplicity are good way to examine the presence or not of critical behaviour. The comparison with the available experimental data obtained from collision between Hj 7 and He target and a 3D bond lattice percolation model of same size shows remarkable similarity in these correlations. The observation of a critical behaviour has been interpreted for a long time as an evidence for a second order phase transition 14 ' 24 ' 25 . Recently, the critical behaviour experimentally observed in the fragmentation of small systems has been demonstrated to be compatible with a first order phase transition because of finite size effects26 but also with a "percolation like" transition (transition without a discontinunity for the thermodynamical quantities) which has been also reported inside the single fluid part of phase diagramme 27 ' 28 . From the comparison presented in this paper, strong simi-
242
larities are evidenced but some differences are also observed between the data and the percolation model. The critical behaviour experimentally observed may not result from a percolation transition. Research of the caloric curve of this fragmentation phenomena may give some informations about the nature of this critical behaviour. References 1. R. P. Binzel, Nature 388, 516 (1997) 2. Fragmentation Phenomena D.Beyson, X.Campi, E.Pefferkorn(eds), (World Scientific, Singapure, 1995) 3. B. Farizon et al, Eur. Phys. J. D 5, 5 (1999) 4. I. Stich et al, J. Chem. Phys. 107, 9482 (1997) 5. I. Stich et al, Phys. Rev. Lett. 78, 3669 (1997) 6. M. Farizon et al, J. Chem. Phys. 96, 1325 (1992) 7. B. Farizon et al, Phys. Rev. B 60, 3821 (1999) 8. B. Farizon et al, Chem. Phys. Lett. 252, 147 (1996) 9. B. Farizon et al, Int. J. Mass Spectrom. Ion. Proc. 164, 225 (1997) 10. J. E. Finn et al, Phys. Rev. Lett. 49, 1321 (1982) 11. A. Hirsch et al, Phys. Rev. C 29, 508 (1984) 12. X. Campi et al, Nucl. Phys. A 495, 259c (1989) 13. V. Latora et al, Phys. Rev. Lett. 73, 1765 (1994) 14. M. E. Fisher et al, Rep. Progr. Phys. 30, 615 (1967) 15. A. Van Lumig et al, Int. J. Mass Spectrom. Ion. Phys. 27, 197 (1978) 16. C. E. Klots, J. Phys. Chem., 92, 5864 (1988) 17. B. Farizon et al, Int. J. Mass Spectrom. Ion. Proc. 164, 225 (1997) 18. B. Farizon et al, Nucl. Instr. and Meth. in Phys. Res. B88, 86 (1994) 19. R. Ehlich et al, J. Chem. Phys., 104, 1900 (1996) 20. T. LeBrun et al, Phys. Rev. Lett. 72, 3965 (1994) 21. B. Farizonei al, Nucl. Instr. and Meth. in Phys. Res. B28, 497 (1987) 22. B. Farizon et al, Phys. Rev. Lett. 81, 4108 (1998) 23. X. Campi et al, Phys. Lett. B208,351(1988) 24. J. A. Hauger et al, Phys. Rev. Lett. 77, 235 (1996) 25. J. B. Elliott et al, Phys. Lett. B 418, 34 (1998) 26. F. Gulminelli et al, Phys. Rev. Lett. 82, 1402 (1999) 27. J. Kertesz, Physica (Amsterdam) 161A, 58 (1989) 28. X. Campi et al, Nucl. Phys. A620, 46 (1997)
SINGLE QUASIPARTICLE E N T R O P Y IN EXCITED NUCLEI WITH T < 1 MEV
M. G U T T O R M S E N , M. H J O R T H - J E N S E N , E. M E L B Y , J. R E K S T A D , A. S C H I L L E R A N D S. S I E M Department
of Physics,
University
E-mail:
of Oslo, Box 1048 Blindern, Norway [email protected]
N-0316
Oslo,
Entropies in rare earth nuclei are extracted from measured level densities. The entropy difference between odd and even-even mass nuclei are almost independent of excitation energy in the 1 . 5 - 6 MeV region. A global study of level densities from so-called anchor points, reveals that this is a common property for mid-shell nuclei. The findings indicate thermodynamical extensivity, suggesting that each thermal quasiparticle is carrying an entropy of AS ~ 2.
1
Introduction and Anchor Points
The concept of thermal quasiparticles is based on the idea that thermal properties at low excitation energy are governed by a few quasiparticles coupled to the cold core of Cooper pairs. These quasiparticles are thermally scattered on available single particle states. Thus, they are not well-defined by one single (Nilsson) orbital with parity, spin-alignment and signature, but exhibit an average of the spectroscopical properties of the neighbouring orbitals. At higher temperatures the pair correlations are quenched 1,2 and the core of Cooper pairs is no longer well-defined. In a recent paper 3 , we measured the microcanonical entropy in hot i6i,i62 Dy a n d i7i,i72 Y b nuclei. Figure 1 shows that 161 . 162 Dy displays an approximately constant odd-mass entropy difference of AS ~ 1-8(1) (in units of the Boltzmann constant ks) in the 1.5-6 MeV excitation region. The purpose of this talk is to show that the single quasiparticle entropy is a global property of mid-shell nuclei. The entropy is calculated from the nuclear level density by S(E) = So + In p(E), where So is a constant. Level density information for several nuclei are obtained by determining two anchor points. The first point (pi,Ei) is based on counting known discrete levels at low excitation energy, 0 < E < 2 MeV. Here, we have analyzed levels of ~ 1600 nuclei from the database of Ref. 4 . The second anchor point (p2, £2) is estimated from the average neutron resonance spacing at the neutron binding energy E = Bn, where we use the compilation of Iljinov et al. 5 . 243
244
-
J
:
•
16,
Dy_,-
***
.*'
Dy
w', , ,
I.'..'.. I 0 1
2
3 4 5 Excitation energy E [MeV]
Figure 1. Experimental microcanonical entropy for ergy E.
161 162
'
6
D y as function of excitation en-
In Fig. 2 the anchor points extracted for nine isotopes of hafnium shown. In between the anchor points we approximate the level density cording to the constant temperature formula p{E) = Cexp(E/T), where normalization constant C and the temperature T is determined from the chor points. 2
are acthe an-
Single Quasiparticle Entropy
In the following, we will compare the entropy for odd and even-even mass nuclei. In order to have a common play ground, we extra- or interpolate the first and second level density anchor points to excitation energies 1 and 7 MeV, using the constant temperature formula. Figure 3 displays the entropy S — So evaluated at E — 1 MeV. The data are plotted as function of the mass number, where nuclei along the /?-stability line are chosen. Figure 3 reveals that in the mid-shell regions the odd-mass nuclei have about A 5 ~ 2 higher entropy compared to the even-even system. This feature holds for odd-even as well as for even-odd nuclei. Thus, a single proton or neutron quasiparticle outside the core of Cooper pairs carries approximately the same amount of entropy. Similar plots can be made at 7 MeV, which are most sensitive to the second anchor point, namely neutron resonance spacings.
245
,7!
10"
Hf
"*Hf
'"Hf
10' 10! D
•^UU,
, ' ' iTlfc ; 10* '"Hf MO'
; io ! •10'
10'
nC\ ,77
""Hf
/
0.56 MeV
Hf
.
^ ,7B
10'
.
/
/
179
0.56 MeV
Aitt.,
0 2 4 6 8
,BD
Hf
/
^S.
0.55 MeV
r\
lf\i
10 !
Hf
0.54 MeV
Hf
/
0.58 MeV
nHii, ,
0 2 4 6 8 0 2 4 6 8 Excitation energy [MeV]
Figure 2. The level density anchor points extracted for 2— Hf. The lower point is based on known discrete levels. In cases where the second anchor point is known, the temperature T has been extracted.
We find the same conclusions here. It is amazing that AS is approximately the same at both 1 and 7 MeV of excitation energy. This means that even with 4 - 6 quasiparticles excited at 7 MeV, the last thermal quasiparticle still carries A.S ~ 2. Thus, the experimental entropy scales with the number of quasiparticles. This property is also accounted for in a simple model 6 . The regions of quasiparticle extensivity can be identified in Fig. 3 as entropy-gaps between the odd-mass and even-even nuclei. 3
Conclusions
We have investigated the global systematics of nuclear entropy as function of nuclear mass, excitation energy (or temperature) and number of excited quasiparticles. The level densities are detemined by counting discrete levels and from neutron scattering data. The single quasiparticle entropy is determined to be AS ~ 2, which is relatively independent of the presence of other quasiparticles. This appearent extensivity is valid for mid-shell nuclei with mass number A > 40.
246
:
E = 1 MeV
7 N=126, Z=82
'
25
I
50
i
75
100
125
150
175
200
225
250
Mass n u m b e r A
Figure 3. Entropy as function of mass number at 1 MeV of excitation energy. The oddeven (filled triangles) and even-odd (filled squares) nuclei exhibit higher entropy than the even-even (open circles) nuclei. The appoximate locations of closed shells for neutrons and protons are indicated.
Acknowledgments We acknowledge the support from the Norwegian Research Council. References 1. 2. 3. 4. 5. 6.
E. Melby et al., Phys. Rev. Lett. 83, 3150 (1999). A. Schiller et al., preprint, nucl-ex/9909011. M. Guttormsen et al., Phys. Rev. C, (2000), in press NNDC On-Line Data Service from the ENSDF database, January 2000. A.S. Iljinov et al., Nucl. Phys. A543, 517 (1992) M. Guttormsen et al., preprint, June 2000.
Section III Nuclear Caloric Curve and Thermodynamics of Heavy Ion Collisions
S U R V E Y I N G T E M P E R A T U R E A N D D E N S I T Y M E A S U R E M E N T S IN N U C L E A R CALORBVIETRY
G.RACITI 1 . R.BASSINI 2 , M.BEGEMANN-BLAICH 3 , S.FRITZ 3 , C.GROB 3 , G.IMME 1 , I.IORI 2 , U.LYNEN 3 , M.MAHI 3 , T.MOHLENKAMP 3 , W.F.J .MULLER 3 , B.OCKER 3 , T.ODEH 3 , J.POCHODZALLA 3 , G.RICCOBENE 1 , F.P.ROMANO 1 , A.SAIJA 1 , C.SCHWARZ 3 , V.SERFLING 3 , M.SCHNITTKER 3 , A.SCHUTTAUF 3 , W.SEIDEL 4 , C.SFIENTI 1 , W.TRAUTMANN 3 , A.TRZCLNSKl'.G.VERDE 1 , HONGFEIXI 3 , B.ZWIEGLINSKI 5 . INFN Laboratori Nazionali del Sud and Sezione di Catania and Dipartimento di Fisica ed Astronomia dell' Universita di Catania, Catania, Italy 2
Dipartimento
di Fisica, Universita di Milano and I.N.F.N Milano, Italy
2
GesellschaftfurSchwerionenforschung, Forschungszentrum
D-64291 Darmstadt,
Rossendorf, D-01314 Dresden,
Germany
Germany
Soltan Institute for Nuclear Studies, 00-681 Warsaw, Hoza69, E-mail: [email protected]
Poland
An experimental investigation on theimodynamical observables characterizing the conditions of multifragmenting systems is reported. High granularity hodoscopes allowed simultaneous measurements of isotopic and emission temperatures. HBT interferometry with light charged particles allowed radii measurements. The disagreement between the two temperature measurements could be related to the space-time evolution of the fragmentation process as confirmed by density measurements. The slope temperatures derived from the target spectator decay fragment energy spectra suggest a dependence on the Fermi motion within the initial system. The dependence of the Nuclear Caloric Curve on the mass of the systems was probed.
1 Introduction One of the most intriguing phenomena observed in macroscopic systems is that of a phase transition. The exploration of the phase diagram of a substance is possible via calorimetric studies [1]. Indeed, from the correlation of the measured isotopic temperature [2] and excitation energy of projectile spectator in 197Au+197Au collisions at 600 AMeV, a caloric curve of nuclei has been reported [3]. The similarity of this curve to first-order liquid-gas phase transition in macroscopic systems has initiated a widespread activity, both theoretical and experimental in particular on the temperature measurements. In order to permit a cross comparison of different thermometers we measured simultaneously both isotope yields and particle-particle correlations from the decay of unstable states in the emitted fragments [4-7]. In a series of experiments we studied 197Au + 197Au collisions at 1 249
250
AGeV to investigate the spectator decay [4,8,9], and mid-central collisions at 50, 100, 150, and 200 AMeV [5] to study the decay of the participant region which is known to exhibit a considerable collective flow[10]. To investigate on system size dependence, we studied also central collisions 84Kr+93Nb at NSCL of MSU [6] at incident energies from 35 to 120 AMeV and 40Ca+45Sc at 40 AMeV at the LNSCatania [7]. HBT interferometry technique was applied to charged particle relative momentum correlation functions in order to measure the size of the emitting systems [11] and densities were deduced once the system mass was determined. Moreover some discrepancies between the caloric curve of [3] with the ones obtained later by other groups [12-14] were suggested [15] to be related to the different masses of the studied systems. Recently a size dependence in the melting of sodium clusters has been reported [16]. 2 Experimental Apparata The experiments were performed at the SIS accelerator of the GSI, using the ALADiN facility, at the NSCL of MSU and at the LNS-Catania. In the GSI experiments [4,5,8,9] three modular hodoscopes each consisting of Si-CsI(Tl) telescopes(Si 3x3 cm2 300um thick, Csl scintillators 3x3 cm2 6 cm long with photodiode readout), were set up with an angular resolution and granularity optimized to identify excited particle-unstable resonances in light fragments from the correlated detection of their decay products. A set of seven telescopes (Si 50 um, 300 um, and 1000 um, Csl 4 cm long) at selected angles, measured the yields of isotopically resolved light fragments. At 1 AGeV the impact-parameter selection was achieved by measuring the quantity Zbound (the sum of all Z>1 belonging to the projectile) through the forward angles TOF wall of the ALADiN setup [17]. At lower bombarding energies the associated charged particle multiplicity was sampled, at forward angles, with an array of 36 CaF2-plastic phoswich telescopes (ZDO) and with a set of 6 Si-strip 16-fold detectors 300 um thickness, 5x5 cm2 surface. At the LNS laboratory we studied the 40Ca + 45Sc reaction at 40AMeV with a similar setup [7,15]. Two high granularity hodoscopes were used to measure simultaneously particle-particle correlations and the yields of isotopically resolved fragments. One, the HODO-CT was the same 96 element hodoscope used at the GSI, improved with a 50 um Si-detector in front of each telescope, in order to reduce the detection threshold. The other, HODO-FWD, covered the angular range 9iab = 5° -=-11° around zero degree in order to detect projectile fragmentation. It was composed by 80 two-folds telescopes (Si 300 um lxl cm2 followed by a 10 cm long CsI(Tl)) and 40 three-folds telescopes of the same kind as the HODO-CT's ones. Charged particle multiplicity was sampled by a set of 12 Si strip detectors 16-folds set up around the target. At the NSCL the 84Kr+93Nb reaction , was studied [6] by coupling the HODO-CT with the 4JI detector (Soccer Ball) that served as impact parameter filter.
251
3
Nuclear Thermometers Cross Comparison
The methods currently in use to measure temperature of nuclear systems are summarized in Fig. 1-a [1]. In particular the Caloric Curve of Ref.[3] was built by correlating excitation energy with "isotope" temperature, i.e. the one deduced from the double ratio of isotopes differing by one mass unit[2]. Indication for equilibrium in the spectator system is related to the universality with respect to the colliding systems and the incident energy [18] and to a good agreement with the statistical multifragmentation models[19].However, several both theoretical and experimental works appeared on the problems related to the influence of sequential decay [20,21] and the different values measured depending on the selected isotope pairs[22,23]. Nevertheless, "isotope" temperatures seem to well reflect the breakup conditions. In Fig. 1-b the results for THeu in 197Au on 197Au spectator fragmentation, obtained in three different measurements, are shown as a function of Zb0Und- Integrated yields of 3,4 He and 6'7Li isotopes were measured, for the projectile-spectator breakup at forward angles by the ALADIN spectrometer, and for the target spectator, by a standard moving source fit procedure of the energy spectra measured in the backward region with the four-element telescopes.In both cases Zb0Und was determined from the coincident projectile fragments detected by the ALADINsetup.
0 10 20 50 40 50 60 70 ~Umi c) Figure 1 : a) Nuclear thermometers, b) Temperatures THCU for target (E=1000 A MeV) and projectile spectators (E=600 and 1000 A MeV) as a function of Zbound- c) QSM-corrected isotope temperatures for 157 Au target spectator at E=1000 A MeV as a function of Zbound •
In Fig.l-c results are given for a variety of isotope thermometers that are all characterized by a double difference of the binding energies AB>10MeV. All values have been corrected [1] for the effect of secondary decays by using the predictions of the quantum statistical model (QSM) [24].The rise at small Zb0Und, ie. high excitation energy, already observed for THeu ,is well reproduced by both TH,He and Tfieu (7,9Be and 6'8Li isotope ratios). This demonstrates that the rise is not necessarily related to a particular behavior of either 3He or 4He and is not very much affected by the sequential feeding for the quoted thermometers.
252
3.1
Emission vs. Isotope Temperatures.
In order to permit a cross comparison of two different thermometers we measured, for the same class of events of the same fragmenting system, both "isotope" and "emission" temperatures. The latter can be calculated assuming that the states of the decaying fragments are populated according to a Boltzmann law. Therefore from peaks in relative momentum correlation functions due to resonant decay of two coincident particles, acceptance corrected yields of fragments excited states have been determined [5]. Figure 2 shows the comparison of the isotope and emission temperatures measured in central collisions in the Au+Au and Kr+Nb reactions at different incident energies. Differences up to 8 MeV between the two thermometers are reached at the highest bombarding energy. We notice that temperatures related with Carbon isotopes ( TCLi, TCc) show the same behavior of the emission ones (Fig.s. l-c,2-b) o
s
u
D *He A "Be
01
-
+ /•
. > 8
Kr
• Heli • Hadt
*/
•
5
_ •
l '
+
0
T
•
T
N b : b/b JHe-Li)
I
,PU) s
•
T c(
•
T r(*He)
Bej
ll
? 6
\
<
(i
i I
. . r . . . .
a)
100 150 EA (MeV)
60
to
E/A
80
100
120
140
(MeV)
Figure 2: a) Emission temperatures extracted from the excited states of 4 He, ! Li, 8 Be and isotope temperature are shown as a function of the bombarding energy for central Au+Au collisions [5]. b) Same for Kr+Nb central collisions [6].
Emission temperatures measured by fragments excited states populations should be related to a later time when the adiabatic expanding system has cooled down[5,15]. A different freeze-out time of the chemical degree of freedom and the internal ones of the single constituents could explain the observed different behavior. 3.2
Kinetic Energy Spectra.
Energy spectra of charged particles and fragments measured in the I97 Au + 197Au target spectator fragmentation at 1 AGeV exhibit Maxwellian shapes to a good approximation [8,9]. The slope parameters, shown in Fig. 3-a, apart for protons, are scattered closely around a mean value of 17 MeV. A constant temperature and the corresponding mass invariance of the mean kinetic energies is a very general indication of equilibrium.
253
Since the mutual Coulomb repulsion after breakup and the small but finite motion of the target spectator have relatively small contributions and, at backward angles, should cancel each other, the measured slope parameters of T=17 MeV may be taken as representative of the thermal component of the fragment kinetic energies [9]. However, as it was suggested by Goldhaber [25], the large kinetic energies of fast-fragmentation products may have their origin in the nucleonic Fermi motion and the resulting behavior is indistinguishable from that of a thermalized system with rather high temperature. Goldhaber's idea has been extended by Bauer [26] to the case of expanded fermionic systems at finite temperatures. The slope temperatures, represented by lines in Fig. 3-b, calculated assuming the measured breakup temperatures THeLi as the system temperature and the two values p/p0= 10 and 0.3 ,for the breakup density, are in the range of the experimental values. They reproduce the rise with decreasing Zb0und which follows as a consequence of the rising breakup temperature THeu- It also seems that a better agreement favor the prediction for p/p0 = 1.0. — 35
> |
30
|_™20 15
I
I I I ! IT T 20
^4 L.J* -: •
I i 111111111
11 11 11 I 11 r i 1 1 1 1 I 11 11 i r
-j 0
t
•far T* j
2-1 _: 2=2 2=^ T 5 — : A 2-4 0 - I I I I I i i i i I 0 1 2 3 4 5 6 7 3 9 1011
10
a)
^n
-
^
*
•
A
•
0
10
*
^
«L
*
*
*
I....I
20
30
jf
* • +-. I...,I,.,.I,
40
SO
60
7D BO Ground
Figure 3: a) Slope temperatures measured in the 197Au + ,97 Au reaction at E ^ 1000 A MeV, as a function of the fragment mass number A. b) Slope temperatures for light charged particles and the isotope temperature THd.i as a function of Zbomd.
This result suggests a fast abrasion stage that will induce fragmentation of the remaining spectators but do not affect much the Fermi motion of the constituent nucleons in the equilibrated system before its decay. This is consistent with a partitioning according to phase space but contradicts the equilibration of the kinetic degrees of freedom at the breakup that is assumed in the statistical multifragmentation models [24]. 4
Density measurements
The density value p/po , used to study slope temperatures in the previous section, was measured in Au + I97Au target spectator fragmentation at 1 A GeV [4] by means of particle-particle HBT interferometry [11]. Details on the analysis of the
254 particle-particle correlation functions, sorted into bins of Zbound in order to select excitation energy and impact parameter, are given in [27]. Radii were deduced in the zero-lifetime limit for the source, and densities were calculated by dividing the number of spectator constituents evaluated from Zbound [3,4] by the source volume. The derived densities (Fig.4-a) vary considerably with impact parameter, roughly in proportion to the variation of the spectator mass. In the p-p case, the mean relative densities decrease from p/p 0 = 0.4 for the near-peripheral to below p/p 0 = 0.2 for the most central collisions [4]. These values compare well with the densities assumed in the statistical multifragmentation models [24] in contrast with the p/po=l value suggested by the slope temperatures analysis. The lower values related to the p-a and d-cc correlations may be partially due to the different formalisms that were used but they could indicate a moderate expansion of the system. 14 13
j= 12
" 11 8 10 9 8
-
s §s \ i •
P«
T
100 150 200 E BMm (AMeV)
b)
Figure 4: a) Breakup density p/po as deduced from p-p, p-a, and d-ot correlation functions for " 7 Au + 197 Au collisions at 1000 A MeV per nucleon. The error bars represent the statistical uncertainty. b) Radii extracted from p-p, p a and d a correlation functions [15,28].
Similar differences are evidenced in the analysis of the proton-proton and unlike particle correlations measured in Au+Au central colhsions. The collective expansion, that represents, in central collisions, a significant fraction (about 40%) of the energy available in the center-of-mass frame [10], was taken into account in the radii determination from the experimental correlation functions. In fig.4-b the constant values measured from p a and d a contrast with the ones from p-p that drop with increasing beam energy. These differences, enhanced by the increasing of the collective radial flow, could be related to the lower density reached by the expanding system at the freeze-out time, when alpha correlations sample the emitting source. This scenario agrees well with the EES [29] prediction that the emission time for a ' s and larger clusters is delayed with respect to p, d, t, and 3 He. Moreover, since an adiabatic expansion imply the cooling of the system, it may not be too surprising that the discrepancy between the isotope and emission thermometers arises at beam energies where collective radial flow becomes important. The time evolution of the system is emphasized by the saturation of the
255 excited-state temperatures, radii and fragments emission time measured in Ref.[30]. The saturation temperature T«5 MeV is similar to the asymptotic temperature obtained in dynamical models describing the reaction process [31].
5
Caloric curve as an image of the finite size
An intriguing problem is related to the different behavior of the ALADiN Caloric Curve [3] and the one obtained by the INDRA Collaboration studying the 40Ar projectile fragmentation at 95 A MeV incident energy[12]. Fig. 5-a shows a comparison between these caloric curves and the isotope temperature THeLj measured in central collisions 84 Kr+ 93 Nb at different energies [6] and in 40Ca+45Sc at 40 A MeV [7], The excitation energy for the 84Kr+93Nb central collisions were evaluated roughly from the C M . available kinetic energy reduced by the collective radial expansion term deduced from the systematics [10]. The intermediate source excitation energy m 40Ca+45Sc was evaluated considering the complete transfer of the available kinetic energy to an incomplete fused system of 82 nucleons, as suggested from the C M . velocity spectra fit [7]. We notice that the same excitation energy (E/A=8MeV) determines higher temperature in the system produced in the 40 Ca+45Sc (A=82) than in the A-170 system selected in '"Kr+^Nb at 35 AMeV incident energy. '"Au+'^Au 600 AMeV *Ca+"Sc:. 40 AMeV (LM5) "Kr+"Nb (NSCL-MSU) „
/*
^ *
310
+
1. A
«T
V A V
»
T
* 2
"Ar+"Ni 95AMeV (INDRA) " • " 1
5 10 15 < E 0 > / < A „ > (MeV)
«
nihil
ao M i n i » no i«n MO an
b)
Figure 5: a) Cross comparison between calorie curves of nuclei and b) the related latent heat/nucleon.[7]
The comparison suggests a size dependence of the temperature-excitation energy correlation. The same indication is given by the latent heat reported in Fig 5-b versus the mass of the investigated systems. The latent heat values were deduced from the caloric curves of [3] [12] and [ 13] by integrating the gaussian shapes of the 8E/8T functions around the transition temperature[7]. The increasing values with the system mass are inconsistent with a volume dependence (curve in fig 5-b). Recently [16] surprisingly large variations in the melting point and latent heat have been observed with changing cluster size of ionized sodium atoms. These variations
256 .
cannot be yet fully explained theoretically, and refer to different microscopic structures and phase transitions, but we want to emphasize that a size dependence of the thermodynamical properties of different systems opens the possibility of precise studies on finite systems far away from the infinite size limit. References 1. 2. 3. 4. 5. 6. 7. 8. 9. 10. 11. 12. 13. 14. 15. 16. 17. 18. 19. 20. 21. 22. 23. 24. 25. 26. 27. 28. 29. 30. 31.
JPochodzalla, Prog.Part.Nucl.Phys. 37(1997)443 and references therein S.Albergo et al., // Nuovo Cimento 89A (1985) 1. JPochodzalla et al., Phys. Rev. Lett. 75 (1995) 1040. S. Fritz et al., Phys. Lett. B 461(1999)315. V. Serfling et al., Phys. Rev. Lett. 80 (1998) 3928. H. Xi et al., Phys. Rev. C 58 (1998)2636. C.Sfienti, PHD Thesis -Univ Catania 2000 Hongfei Xi et al., Z.Phys. A359 (1997)397. T. Odeh et al., Phys. Rev. Lett. 84 (2000)4557. W.C.Hsi et al., Phys. Rev. Lett. 73(1994)3367 and M.A. Lisa et al., Phys. Rev. Lett. 75(1995)2662 and G. Poggi et al., Nucl. Phys. A586(1995)755. S.E. Koonin, Phys. Lett. B70(1977)43, D.H.Boal et al. Rev. Mod. Phys. 62 (1990) 553 Y.G..Ma et al., Phys. Lett. B390(1997)41. G.Auger et al. Proceedings of CRIS 96 Ed.World Scientific (1996), 133 LA.Hauger et al. Phys Rev Lett. 77(1996) 235 G.Raciti et al. Nuovo Cimento Alll(1998)987 M. Schmidt et al., Nature 393 (1998)238. J. Hubele et al., Z.Phys. A340 (1991)263. A.Schuttauf et al., Nucl. Phys. A607 (1996)457. A.Botvina et al., Nucl Phys. A584 (1995) 737 M.B. Tsang et al., Phys.Rev.Lett 78(1997)3836. H.Xi et al. Phys. Rev. C59 (1999) 1567 A.Siwek,et al Phys. Rev. C57 (1998) 2507 H.Raduta, R.Raduta Proceedings of this Conference D.H.E.Gross, Rep. Prog. Phys. 53 (1990)605, J.P.Bondorf et al., Nucl. Phys. A444 (1985)460. J.P.Bondorf et al., Phys Rep. 257(1995)133. A.S. Goldhaber, Phys.Lett. B53 (1974)306. W. Bauer, Phys.Rev. C51(1995)803. C.Schwarz et al. Proceedings of this Conference G.Imme et al. Proceedings of CRIS 98 Ed.World Scientific (1999) 289 W.A. Friedmann, Phys. Rev. C42 (1990)667 E.Bauge et al. Phys Rev Lett. 70 (1993) 3705 C. Fuchs et al., Nucl. Phys. A 626 (1997)987.
CALORIC CURVE OF F R A G M E N T I N G SYSTEMS C. O. DORSO AND A. CHERNOMORETZ Depto. Fisica, Facultad de Ciencias Exactas y Naturales, Pab. 1 Ciudad Universitaria, I4S8 Buenos Aires, Argentina E-mail: [email protected] We analyze the Caloric Curve (CC) of small fragmenting aggregates in two and three dimensions, excited isotropically and anisotropically. It is shown that the CC displays a rise plateau pattern in both cases.
1
Introduction
In this paper we study the fragmentation of excited classical Lennard-Jones (LJ) drops in 3-D (disks in 2-D), formed by JV = 147 (TV = 100) particles, simulated via molecular dynamics. We use this simple interaction potential because its equation of state displays Van der Waals behavior as is expected for nuclear systems. The main results can be summarized as : i) The time of fragment formation can be precisely determined, ii) the CC displays a "rise-plateau "shape (no vapor branch present), iii) the temperature of the system at fragmentation time can be obtained from the temperature of the asymptotic fragments, and v) the shape of the CC is the same for 2 and 3 dimensional systems and for isotropic and anisotropic excitation mechanisms.
2
3D explosive numerical experiments
We study fragmentation of excited drops formed by particles interacting via a 12-6 LJ potential truncated at rco — 3
257
258
2.1
Time scales of fragment formation and emission
We now focus on the determination of the times of fragment formation and emission. In previous papers we have fully analyzed the main fragment recognition algorithms currently in use 2 . In the simplest one, MST, a cluster is defined as: a set of particles i,j,k,... belong to a cluster C if: V i € C , 3 j £ C / |r; - r,-1 < Rct, where r, and r^ denote the positions of the particles and Rci is the clusterization radius, Rci < rc. In our calculations we took Rci = 3a. On the other hand, the early cluster formation model (ECFM) 3 , is based on the following definition: clusters are those that define the most bound partition of the system, i.e. the partition ({C,}) that minimizes the sum of the energies of each fragment: -E{d} = J2i \12jeCi Kjm + 52j,ked ViM > w n e r e t h e n r s t s u m i s o v e r t h e clusters of the partition, and Kjm is the kinetic energy of particle j measured in the center of mass frame of the element of the partition which contains particle j . The algorithm ( ECRA) devised to achieve this goal is based on an optimization procedure in the spirit of simulated annealing. Moreover the elements of a partition are refered as fragments when they attain microscopic stability, 1 . We say that an ECRA partition is microscopically stable when it differs from the asymptotic fragments in a simple evaporation process. The time at which this happens is called "fragment formation time" T/f(E). In the same way "fragment emission time" Tfe (E) is that time at which the MST partition does. For all the energies considered it is found that Tff(E)
2.2
Caloric curve - Effective temperatures
We define Caloric Curve as the temperature of the system at fragmentation time Tff(E). Being the radial collective velocity position dependent, we divide our drops in concentric spherical regions, centered in the c m of the system, of width Sr =2a. We define the mean radial velocity of region i as: —v^d(t) — IvrTtT J2ev S i e i Tr (til > w n e r e t n e n r s t s u m r u n s o v e r t n e different events for a given energy, the second over the particles j that belong, at time t, to region i; Vj and r^ are the velocity and position of particle j . Ni(t) is total number of particles belonging to region i in all the events. We define the local
259
1
• 1
III
1
1
>
0.8 _ 1
1
•
a
1
•*
§ 0.6
•
ro
1
2.0.4 E
1
*->
1
0.2 -6.0
1
•
-2.0 0.0 energy (e)
-4.0
0.0
4.0
2.0
Figure 1. Caloric Curve for 3D finite system
(i)
temperature Tf0'c: r(0 L
_
loc ~~
2J_^p 1
m
V
(0
rad
(1)
where Ni is the total number of particles in cell i in all events. At this time velocity fluctuations in radial and tangential directions are of the same order, so, the conjecture that the fragmenting system is in local equilibrium is consistent. In Fig. 1 we show, the "extended caloric curve", which encompasses the solid-like phase (region I), the liquid-like phase (region III), the associated phase transition 5 (region II) , and also the higher energy process of evaporation and multifragmentation (region IV). In region IV we plot the local temperature averaged over the three innermost regions from our fragmentation computer experiments. This represents the break-up temperature of the system. We can see from Fig.l that the local temperature at break-up is quite independent of the total energy of the fragmenting system. We now ask ourselves if there is a way of measuring this temperatures. In Fig. 2 we show the temperature of the clusters (related to the fluctuations of the velocity of the cluster constituents with respect to the velocity of the center of mass of the cluster) 1 . In 2a) we show the temperatures of the fragments for
260 i
.
•
i
'
i
'
0.8 0.6 tA
0.4
(a) 0.2 0.0
i
0.0
'
I
.
I
.
I
,
25.0
50.0 75.0 mass number
100.0
25.0
50.0 75.0 mass number
100.0
Figure 2. Cluster temperatures at fragmentation time (a) and at asymptotic times (b)
all excitation energies analyzed in this work at fragmentation time, and in 2b) the same quantity but calculated at asymptotic time. Comparing 2a) with 2b) and 1) it is immediate that the effective temperature at fragmentatioin time Tff{E) can be obtained from the temperature of the clusters at asymptotic times. 3
Analysis of 2D collisional numerical experiments
In the previous calculation the excitation of the system was isotropic. We now ask ourselves how the determination of the caloric curve of a system is influenced by the anisotropy of the excitation. For this purpose we analyze 2 dimensional numerical experiments in which a Lennard Jones Disk composed of N=100 particles is impinged by an energetic 3 particle projectile.
261
An initial highly-collisional stage starts when the projectile hits a localized area of the drop surface. Then, the momentum is transferred in two different ways: shock waves that travel along preferred directions through the ordered ground state configuration, and a chaotic collision pattern that thermalizes the excitation. The first shock-wave mode is responsible for the rapid emission of light clusters (prompt emitted clusters), mainly composed of particles from the target surface. We calculate the mean velocity transfer in j direction: Mj(t) = < ]Ci=o l(^«'(* + <&) ~ ^t(*))-&/1 >etientswhere j denotes the x or y directions, dt = 0.25, ex = x corresponds to the incident direction, and ey=y corresponds to the perpendicular one. We call the time at which Mx ~ My, tpre- We then define the promptly emitted particles (PEP) as those unbound light clusters (mass < 4) that are detected at time tpre, and that remain unbound at any later time . The isotropic collisions (disordered collisional mode), are responsible for the momentum redistribution among the non-promptly-emitted particles. The energy involved is used to heat up the system and to build a collective expansive motion for the 'remaining' particles. In order to determine this energy we first calculate the energy carried by the PEPs as a function of the beam energy. Once the number of PEP, as a function of the beam energy, is known, the energy per particle that remains in the system after tpre as a function of the beam energy can be obtained and the system excitation energy can be calculated. From it we calculate the effective temperature of the systems at fragmentation time. We have used the same methodology as in the case of 3D fragmenting drops, described above. In this case the effective temperature is calculated on a 2D grid (no obvious symmetry is present due to the anisotropic excitation mechanism) The resulting caloric curve for excited 2D LJ disks is shown in Fig. 3. (triangles) together with the corresponding cluster temperatures (squares) is shown in Fig. 3. Once again the rise-plateau pattern is obtained 4 4
Conclusion
The fragmentation process can be divided in three stages. From t = 0 to t — Tff(E) the radial flux and density fluctuations develop, which determine the cluster partition according to the ECFM model. By the end of this stage the asymptotic fragments are already formed although most of the mass of the system is still interacting and forming a big MST cluster. The second stage of the process, goes from Tff(E) to Tfe(E). During it, the already formed fragments separate in configurational space. The third stage, t > Tfe, is the
262
0.00
'
0.0
•
T
>
'
•
1.0
1
'
'
•
2.0
r
'
3.0
•
'
4.0
excitation energy
Figure 3. Caloric Curve for 2D finite system, For low excitation energies the temperature of the biggest fragment (circles) is shown. For high energies we show the system temperature (triangles) and the cluster temperature (squares), both at fragmentation time
free expansion stage. We have calculated the extended caloric curve that describes the thermal behavior of our L-J drop from the solid-like regime all the way to the fragmentation regime. The resulting picture shows the standard behavior up to the liquid-like state. At higher energies the break-up temperature displays a plateau followed by a monotonous, slow, decrease as the system fragments into smaller and smaller clusters. We do not find any evidence of the "gas phase" behavior in the caloric curve. Acknowledgments: This work was done under partial financial support from the University of Buenos Aires via grant TW98. References 1. 2. 3. 4. 5.
A. Strachan and C. O. Dorso, Phys. Rev. C 59, 285 (1999) A. Strachan and C. O. Dorso, Phys. Rev. C 56, 995 (1997). C. O. Dorso and J. Randrup Phys. Lett B 301, 328 (1993). A. Strachan and C. O. Dorso, Phys. Rev. C 58, R632 (1998) P. Labastie and R. L. Whetten, Phys. Rev. Lett. 65, 1567 (1990).
T H E R M O D Y N A M I C S OF EXPLOSIONS G. NEERGAARD Niels Bohr Institute, Blegdamsvej 17, DK - 2100 Copenhagen, i
Denmark
and
Institute of Physics and Astronomy, University of Aarhus, DK - 8000 Aarhus E-mail: [email protected] J. P. BONDORF Niels Bohr Institute, Blegdamsvej 17, DK - 2100 Copenhagen, E-mail: [email protected]
Denmark
I. N. MISHUSTIN Frankfurt University, D-60054, Germany and The Kurchatov Institute, Russian Research Center, 123182 Moscow, Russia E-mail: [email protected] We present our first attempts to formulate a thermodynamics-like description of explosions. The motivation is partly a fundamental interest in non-equilibrium statistical physics, partly the resemblance of an explosion to the late stages of a heavy-ion collision. We perform numerical simulations on a microscopic model of interacting billiard-ball like particles, and we analyse the results of such simulations trying to identify collective variables describing the degree of equilibrium during the explosion.
1
Introduction
The assumption of thermodynamic equilibrium at an intermediate stage of a heavy-ion collision is often incorporated in models of the colliding nuclear matter. These models range from statistical models of nuclear multifragmentation to the fluid dynamical models of the quark gluon plasma. In contrast, microscopic models of molecular dynamics type (e.g. RQMD, FMD and NMD), which are based upon constituent interactions, do not contain this assumption. Such models are appropriate for testing to what extent thermodynamic equilibrium is actually achieved. And if it is not, the application of thermostatic concepts such as temperature and entropy becomes questionable. In this study we employ a very simple model, and focus on the thermodynamic or "overall" description of the system.
263
264
2
The model
Our model consists of a number A of identical balls of radius The having mass m. They perform classical non-relativistic hard-sphere scatterings, conserving energy, momentum and angular momentum. Initially the A balls are placed randomly within a sphere of radius R — RQA1/3, and the initial velocities are chosen as a superposition of thermal (Maxwell-Boltzmann) and collective motion. We use a spherically symmetric Hubble-like flow field for the initial collective motion: v(f) = -v0f r/R
(1)
where Vof is a model parameter, vof > 0 for ingoing flow and t>0/ < 0 for outgoing flow. We fix the total energy E = Efi + Eth, and vary the fraction 2
?? of the flow energy, r\ = Efi/E, where Efi = -^jf X).=i *? anc ^ &th aie t n e flow energy and the thermal energy, respectively. Because of the way in which the system is built up, these energies will fluctuate from event to event with a relative uncertainty of the order of A~ 2. In our simulations we have chosen nuclear-scale parameters: m = 940 MeV, rhc = 0.5 fm, RQ = 1.2 fm, 0 < v0f < 0.5 (in units of the velocity of light, c = 1), but since the behavior of the model only depends on the two combined parameters mv^f and rhc/R, the choice of nuclear scale is not crucial. We choose A = 50, so with these parameters the initial radius of the system is 4.2 fm. We focus on four different types of event: • 'th20': The particles are started in 100% thermal motion inside a spherical container of radius 4.2 fm, at t = 20 fm/c the container walls are removed. • 'in': 100% ingoing flow. After interacting, the particles will move out again. This implosionexplosion process is intended to simulate some features of a heavy-ion collision. • '50/50': 50% thermal motion + 50% outgoing flow, simulating an explosion from a non-thermalized state. • '100out7': 100% outgoing flow inside a spherical container of radius 7 fm. The results are averaged over an ensemble of 20 events of each kind. 3
Thermodynamic considerations
It is clear that we cannot use ordinary thermodynamics (or its well-known extensions to small systems 1 or to small deviations from equilibrium 2 ) for the description of the overall behavior of our model. First, it is not clear that equilibrium prevails, even locally. Indeed we wish to investigate to what extent equilibrium is reached in the course of an implosion-explosion process.
265
Second, our system has no fixed volume, it expands freely into the vacuum. It is the combination of these two facts, no temperature and no volume, that makes our approach different from much previous work on the subject 3 . Equilibrium thermodynamics is linked to the motion of the individual constituents making up the macroscopic system via the entropy 4 ' 5 . A natural starting point for the investigation of the overall, i.e. the "thermodynamic", behavior of our system is therefore to apply an expression similar to the entropy, but in a way that makes sense in this highly non-equilibrium system. To study one-body observables, we reduce the 6A dimensional phase-space of the A particles to 6 dimensions in the standard way 4 . Then we introduce a finite grid in the reduced phase-space, dividing each of the 6 axes into D segments. Instead of working with a fixed grid in phase-space, which would give us the usual entropy 0 , we let the entire grid expand or contract along with the swarm of points in phase-space in a uniform way: The outer grid edge follows the outermost point, the boxes are of equal size, and the number of boxes is kept fixed, thus the physical size (e.g. in units of 7i3) of each box in phase-space varies with time. This is to deal with the no volume problem, we mentioned above. We then introduce the pseudo-entropy as
S = -|E^lo6W
(2)
i
where _
number of points in box i . . total number of points in phase space and £ is a normalization constant. We choose £ as the theoretical maximum value of — YliPi l°g(P«) s o that E € [0,1]. In the current set-up, for the case of N points in a reduced phase-space divided into D6 boxes, £ is the smaller number of log(iV) and 61og(D). We refer to the quantity E as the pseudoentropy instead of entropy, since some important features of the entropy, e.g. that it increases, are not retained in this formulation. Nevertheless, we shall see that E has some nice properties, including that of characterizing the degree of equilibrium. For a system of fixed volume in equilibrium, E is the usual one-body entropy (apart from normalization). 4
Results
In the calculations presented here, we have chosen D = 7, so the 6 dimensional phase-space is divided into 7 s boxes. Because we are dealing with a small "in the limit D —• oo in an ensemble of infinitely many events
266
number of events (typically 20) in the ensembles, the precise values of £ depend on the choice of grid. We have, however, verified that £ behaves qualitatively similar to what is shown here over a range of grid-sizes, varying D between 2 and 9. To give an idea of the dynamics and the timescales, we show in Fig. 1 the scattering rate.
I 0
5
10 15 20 25 30 35 40 45 50 time (fm/c)
Figure 1. The scattering rate (number of scatterings per particle per time unit) for the four cases mentioned in the text. In the implosion-explosion event ('in') practically all particles scatter around the time t ~ 6 fm/c. This is the time when the system is maximally compressed. Then, as the expansion begins, the scattering rate decreases until t ~ 15 fm/c, when interactions have essentially ceased. In the '100out7' case the particles start to hit the container wall at t ~ 5 fm/c (the scatterings against the container walls are not counted here), and the peak in the scattering rate at t ~ 20 fm/c results from particles moving back after hitting the wall and scattering against other particles still on their way out.
Fig. 2 shows how the pseudo-entropy behaves in each of the four cases. In the 'th20' case, the pseudo-entropy £ ~ 1 as long as the particles are in equilibrium at fixed volume inside the container. Then at t = 20 fm/c, when the container is removed and the system starts to expand, the pseudo-entropy decreases, reflecting the fact that the system goes out of equilibrium1'. In the case '100out7', where the particles are started in an extreme nonequilibrium situation, the pseudo-entropy is low (£ = 0.8 is a low value in this context), but increases towards £ = 1 as the scatterings equilibrize the system. By comparison with Fig. 1 one can see that the first jump in £ at 6
The particles stay almost thermalized, though, in the sense that they retain their MaxwellBoltzmann velocity distribution. But they are certainly not in equilibrium, since this means that the phase-space distribution is independent of time.
267
0
5
10 15 20 25 30 35 40 45 50 time (fm/c)
Figure 2. The pseudo-entropy for the four cases described in the text. This variable seems to quantify the degree of equilibrium in the system, S = 1 characterising an equilibrium state.
t ~ 5 fm/c is due to particles scattering against the container wall (when particles hit the wall their velocity is reversed, so in this process many new states in phase-space are being populated), and the second "jump" around t ~ 20 fm/c is due to the many particle-particle scatterings around this time. The interesting case is the implosion-explosion ('in') scenario, since here we do not know in advance if the system reaches a state of equilibrium or not. Prom the fact that the pseudo-entropy in Fig. 2 reaches a value £ ~ 1, the same value as the known equilibrium case 'th20', we infer that the system is in a state of equilibrium around * ~ 6 fm/c (which is also the time of maximum compression). We have checked that the speed distribution of the particles becomes nearly Maxwellian from t ~ 6 fm/c with a temperature in the compressed state of 47 MeV, which is also the theoretical value of the temperature assuming that all of the initial flow energy is converted to thermal energy. Another interesting feature of the pseudo-entropy is that it seems to decay in a characteristic fashion when the system expands from a state of equilibrium, see Fig. 3. 5
Conclusions
In this note we address the problem of thermodynamic equilibration in the context of heavy-ion collisions. We have defined a variable inspired by the entropy which, at least for the cases we have considered, seems to characterize
268
W
f c o
T3 3
0.9
0.8
0.7
in 50/50 th20
', y> \ 10
100
time (fmJc) Figure 3. The pseudo-entropy in a log-log plot, together with the functions: 1.521 - 0 - 2 and 1 . 8 7 1 ~ 0 2 . The decrease of E during the initial expansion of the system seems to follow a power law when a state of equilibrium was present, in contrast to the case '50/50' (intended to simulate the expansion from a not-fully thermalized state) which does not show this behavior. At later times E decreases less rapidly and turns over to approach a finite limiting value, one sees the beginning of this behavior at the curve 'in'.
the degree of equilibrium in an a priori highly non-equilibrium process such as an explosion. Now, more theoretical work needs to be done in order to understand why S behaves in this seemingly interesting way. Acknowledgements This work was in part supported by the Danish Natural Science Research Council. GN thanks the Niels Bohr Institute for kind hospitality and The Leon Rosenfeld Foundation for financial support during the preparation of this work. J P B thanks the Nuclear Theory Group at GSI, where part of this work was done. INM thanks The Humbold Foundation for financial support. References 1. See e.g. T.L. Hill: Thermodynamics of Small Systems I-II, W.A. Benjamin, NY, 1963-64. 2. See e.g. N.D. Zubarev: Nonequilibrium Statistical Thermodynamics, Consultants Bureau, NY, 1974. 3. To mention a few: J. Randrup, Nucl. Phys. A314 (1979) 429-453; J. D. Bjorken, Phys. Rev. D 27 (1983) 140-151; G. Baym, Phys. Lett. 138B (1984) 18-22; H. Heiselberg and X.-N. Wang, Phys. Rev. C 53 (1996) 1892-1902 and references herein. 4. E.H. Kennard: Kinetic Theory of Gases, McGraw-Hill, 1938. 5. Landau and Lifshitz: Statistical Physics I, Pergamon Press, 1968.
M I C R O C A N O N I C A L INVESTIGATION OF T H E R E C E N T N U C L E A R CALORIC CURVE E X P E R I M E N T A L EVALUATIONS AL. H. RADUTA AND AD. R. RADUTA National
Institute of Physics and Nuclear Engineering, Bucharest POB-MG 6, Romania
Nuclear caloric curve is evaluated using the recently improved version of the microcanonical multifragmentation model [Al. H. Raduta and Ad. ft. Raduta, Phys. Rev. C 5 5 , 1344 (1997); 56, 2056 (1997); 6 1 , 034611 (2000)]. The sequence of equilibrated sources formed in the reactions studied by the ALADIN group ( 1 9 7 A u + 1 9 7 A u at 600, 800 and 1000 MeV/nucleon bombarding energy) is deduced by fitting simultaneously the model predicted mean multiplicity of intermediate mass fragments {MJMF) a n d charge asymmetry of the two largest fragments (012) versus bound charge {Z\,oun^) on the corresponding experimental data. This allows for a direct comparison between the model predicted HeLi isotopic temperature curves as a function of the bound charge with the experimentally deduced ones. A good agreement is obtained.
One of the most investigated nuclear physics problem from the last decade is the possible occurrence of a liquid-gas phase transition in nuclear matter. The first experimental caloric curve evaluation was reported in 1995 by the ALADIN collaboration 1 . Three distinct regions were evidenced on the curve: a first increasing region which was connected with a liquid phase of the nuclear matter, a plateau-like region associated with a liquid-gas coexistence and a third linearly increasing region associated with the nuclear gas phase. The result was impressing indicating the possible existence of a liquid gas phase transition in nuclear matter. Two year latter, ALADIN reported a reevaluated version 4 ' 5 of the 1995 caloric curve motivated by the improvements in the detection conditions. With that occasion the sequences of "equilibrated sources" were proven to depend strongly on the bombarding energy suggesting quantities of nonequilibrium energy included in the experimental evaluations. The Albergo HeLi isotopic thermometers were used in these evaluations. However, apart from the experimental uncertainties the grancanonically deduced Albergo thermometers are affected by both finite size effects appearing in nuclear systems and secondary particle emission from the primary decay excited fragments. One therefore needs accurate microcanonical predictions in order to elucidate the real shape of the nuclear caloric curve. The present contribution presents the results obtained with a pure microcanonical multifragmentation model 2 ' 3 and compares them with the most 269
270
-
O
micro.
•
Au+Au 1000 MeV/nudeon
(a) -
•»*•» •o
e •
•b *0
•
•°
o
- rf>s
•q .
.
••• ..•• 9 0 4* *i 0 o •• o..« -
L
.»
0
•
<*
0
10
20
30
40
50
,
00
70
;
80
Figure 1. (MJMF) {(Zbound)) and (012) ((-Zbound)) evaluated by means of the microcanonical model in comparison with the experimental evaluations corresponding to the reaction 197 A u + 1 9 7 A u at 1000 MeV/nucleon bombarding energy. (The deviation of calculated (012) from experimental data for Z),ound > 65 is, as explained in 4 ' 3 , due to some detection problems.)
recent experimental data 4 ' 5 of the ALADIN collaboration. Let us shortly remind the main ingredients of the model. The microcanonical multifragmentation model 2 ' 3 concerns the break-up of statistically equilibrated systems formed in heavy ion collisions. The conservation of mass, charge, center of mass energy and momentum are strictly satisfied. The statistical weight (Wc) of a configuration C : {Ai,Zi,ei,Ti,i = l,...,Nc} can be analytically exactly evaluated 2 . The average value of any arbitrary observable X can be formally expressed as: (X) — ^2C WcX/ ^2C Wc • Since the above formulae are not numerically tractable, we evaluate them using a Metropolis-type simulating procedure 2 . While for fragments with A < 6 the experimental discrete energy levels with T < 1 MeV are employed, for those with A > 6 a Fermi gas type level density formula adjusted with the cut-off factor exp(—e/r) (as to account for the level density limitation at high excitation energy) is used. Primary break-up excited fragments are allowed to undergo sequential secondary decays. Here we consider two processes. For primary fragments with A < 6 and e < B/3 we consider the secondary break-up process. Otherwise, the standard Weisskopf evaporation scheme, extended as to account
271 ...,...,..
| i i i 1
I
""''""
...,.-.,...
A
Au+Au, 1000 MeV/nucleon •
•
Au+Au, 800 MeV/nucleon -
•
Au+Au, 600 MeV/nucleon •
0
micro.
-'•*x: . . i . . . i . . . i . . . i . . . i
0
0
20
40
60
80
100 120 A
140
160
W. i-i.i
180 200
Figure 2. The sequence of equilibrated sources evaluated by means of the microcanonical model (open circles) following the criterion of simultaneous fitting of the calculated MjMF{Zbound) and a\2(Zbound) on the corresponding experimental data. The close symbols represent the sequence of excitation energy experimentally measured for the reaction 197 A u + 1 9 7 A u at 600, 800, 1000 MeV/nucleon bombarding energy 4 ' 5 .
for emitted fragments with mass number up to A = 16 is considered. For each primary decay excited fragment the entire chain of secondary particle emission is simulated using standard Monte Carlo. In order to make a direct comparison between the isotopic HeLi temperature obtained with the microcanonical model and the experimental data of the ALADIN group we first evaluate the sequence of the equilibrated sources obtained in the peripheral collision 1 9 7 Au+ 1 9 7 Au at 1000 MeV/nucleon bombarding energy. In this respect the criterion of fitting simultaneously the average values of the multiplicity of intermediate mass fragments (MiMF) and the charge asymmetry of the two largest fragments (012 = {Zmax - Z2)/(Zmax + Z2), where Zmax and Z2 are the largest fragment charge and, respectively, the second largest charge in one fragmentation event) versus the bound charge (Zbound = E z < 2 ^ < ) ^s u s e c i - Fig. 1 presents the calculated and experimental distributions of MiMF{Zbound) and a\2{Z\,ound). The agreement between the calculated and experimental data is fairly good. The sequence of equilibrated sources is presented in Fig. 2 together with the experimental data obtained for 1 9 7 Au+ 1 9 7 Au peripheral collision at 600, 800 and 1000 MeV/nucleon bombarding energy. It can be observed that the theoretical curve Eex (A) is relatively close to the experimental data corresponding to 600 MeV/nucleon. The deviations are due to quantities of nonequilibrium type energy contained in the experimental evaluations increasing with the increase of the bombarding energy. The origin of the nonequilibrium energy may lay in the pre-equilibrium and pre-break-up stages of the reaction 4 ' 5 . Once the sequence of equilibrated sources is determined, the compari-
272 i
i
1 i •
1
''
1
•
|
1
•
:ul
'
•
O
Y
-
•
1 ll
^
•
•
•
,
.
.
.
.
,
.
.
.
.
HeLi uncorr. micro.
•
Au+Au 1000 MeV/nucleon
-
D
Au+Au
•
600 MeV/nucleon
ffooi
-
° o: .
0
10
1
20
.
30
40
50
60
70
80
Figure 3. HeLi caloric curves evaluated with the microcanonical model (open circles) in comparison with the experimental HeLi temperatures 4 ' 5 corresponding to the reactions 197 A u + 1 9 7 A u at 600 and 1000 MeV/nucleon bombarding energy.
son between the isotopic HeLi temperature experimentally obtained by the ALADIN group in 19974'5 and the prediction of the microcanonical model is possible. In this sense we used the uncorrected Albergo temperature formula: THeLi = 13.33/In [2.18 (YeLi/Y?Li) / (YsHe/Y4He)}. This comparison is presented in Fig. 3 as a function of Zboun
C A N W E D E T E R M I N E T H E N U C L E A R E Q U A T I O N OF STATE F R O M HEAVY ION COLLISIONS? T. GAITANOS, H.H. WOLTER Sektion Physik, Universit&t Munchen, D-85748 Garching,
Germany
C. FUCHS, A. FAESSLER Institut fiir Theor. Physik, Universitat Tubingen, D-72076 Tubingen,
Germany
We discuss the problems involved in extracting the nuclear equation-of-state from heavy-ion collisions. We demonstrate that the equation of state becomes effectively softer in non-equilibrium and this effect is observable in terms of collective flow effects. Thus, non-equilibrium effects must be included in transport descriptions on the level of the effective mean fields. A comparison with transverse momentum, rapidity, and centrality selected flow data show the reliability and limitations of the underlying interaction which was derived from microscopic Dirac-Brueckner (DB) results.
Heavy-ion collisions open the possibility to explore the nuclear equationof-state (EOS) far away from saturation. In particular, the collective flow is connected to the dynamics during the high density phase of such reactions and thus yields information on the nuclear EOS 1. The different components of collective flow, in particular, the in-plane and out-of-plane flow 2 has recently attracted great interest 2 ' 3 , 4 , since also the energy, centrality and transverse momentum dependence has been measured. From the theoretical point of view, however, the determination of the EOS, which is an equilibrium concept, is not straightforward in heavy ion collisions, which are highly non-equilibrium processes. An appropriate starting point are the Kardanoff-Baym equations 5 for the non-equilibrium realtime Green functions. To arrive at transport descriptions several approximations are usually introduced: the semi-classical approximation neglecting non-locality and memory effects, the quasi-particle approximation neglecting the finite width of the spectral functions, and the T-Matrix approximation, specifying the self-energies in the Brueckner ladder approximation. Several attemps have recently been made to improve the first two approximations 6 . They are expected to be important e.g. for subthreshold particle production, but it is not known whether they are also relevant for flow observables. With the above approximations one arrives at a BUU-type transport equation with DB self-energies which should be specified for the general nonequilibrium configuration. Since these cannot be calculated one further introduces a local nuclear matter approximation neglecting spatial variations 273
274
:
DBHF, v, u -0.0 DBHF, v ra -0.8 — - DBHF, v^-0.99 soft/hard Skyrme
*z£**Z 2#*
%fe_ i
10
20
30
-
i
40
2
time [fm/c] Figure 1. Left: Time evolution of the quadrupole moment of the energy-momentum tensor Qzz (bold lines) and of the central density (in units of pSat, dashed lines) at the center in central Au + Au collisions at 0.6 and 6.0 A.GeV. Right: Energy per particle of a twoFermi-sphere configuration (CNM) (minus the relative kinetc energy) as a function of total density. This "reduced" EOS is shown for DB self energies for several relative velocities and for two (equilibrium) Skyrme parametrizations.
and adopting a model for the local nuclear matter phase space distribution. Two models have mainly been used: One is the Local Density Approximation (LDA) assuming equilibrated nuclear matter at the local total density, usually supplemented with an empirical momentum dependence taken from the real part of the optical model. The other is the Local phase space Configuration Approximation (LCA) which assumes a configuration of two interpenetrating streams of nuclear matter with a given relative velocity (also called colliding nuclear matter, CNM), i.e. a two-Fermi-sphere configuration with Pauli-effects taken into account. In the first case this is directly the nuclear EOS. For the second case the self energies have not been calculated in the DB approach (but for the non-relativistic case 7 ) . However, we have developed a self-consistent extrapolation starting from nuclear matter DB calculations as discussed in detail in ref. 8 . This then gives the model non-equilibrium self energies as a function of the configuration parameters, namely the two Fermi momenta and the relative velocity. The relevance of such non-equilibrium effects in the dynamical situation of heavy ion collisions is demonstrated in Fig. 1 (left side). Here the quadrupole _
_
_^
iyrri33
/rill
rp22
m snown moment of the energy-momentum tensor Qzz = TM+T^+TM together with the baryon density as a function of time. It is clearly seen that the local phase space is significantly anisotropic during times which are comparable with the compression phase of the collision independent of beam energy. Hence, non-equilibrium effects are present during a large part of the compres-
275
sion phase where one wants to study the EOS at high densities. Therefore non-equilibrium effects should be considered at least on the level of the effective in-medium interaction, i.e. the mean fields in transport calculations should be taken for non-equilibrium configurations instead for equilibrated nuclear matter. Actual transport calculations have verified that the anisotropic phase space in heavy ion reactions can be parametrized reliably by CNM configurations 9 . In order to have a qualitative feeling for the non-equilibrium effects we define an effective, "reduced" EOS for the CNM configuration. This is given by subtracting from the binding energy per particle for the CNM configuration with a given relative velocity vrei the irrelevant kinetic energy of relative motion between the two streams. This quantity is shown on the right side of Fig. 1 for symmetric CNM configurations (fc^ = fcp2) for relative velocities Vrei = 0.8(0.99) corresponding to beam energies of 0.6(6.0) GeV per nucleon as a function of the total density. vrei = 0 corresponds to the isotropic case (equilibrated nuclear matter, i.e. one Fermisphere). As a reference we also show the widely used Skyrme parameterisations with a soft/hard EOS with K = 200/380 MeV 1. It is seen that the effective EOS becomes more attractive, i.e. softer, compared to the equilibrium EOS and saturates at a higher total density. This effect is understood qualitativelyin the following way: If there was no interaction between the two nuclear matter streams then the reduced energy per particle of the two-Fermi-sphere configuration were just the energy per particle of one Fermi sphere at half the total denisity. The minimum of the reduced EOS would then be shifted to twice paat- The interaction energy per particle can be estimated as twice the real part of the optical potential at the relative momentum which can be estimated as twice the Fermi momentum at the corresponding density. For saturation density this is attractive but decreases with increasing momentum. Thus we expect a minimum of the reduced EOS between one and twice psat and below the saturation energy of equilibrium nuclear matter. This is what is seen qualitatively in Fig. 1. The question of interest is now whether these non-equilibrium effects are observable in terms of collective flow. We have therefore performed transport calculations using CNM self energies derived from DB mean fields calculated by the Tubingen-group (DBF) 10 in the two approximations: the LDA which refers to equilibrated nuclear matter (DBF/LDA), and the local configuration approximation (LCA) where the non-equilibrium DBF mean fields are used in the transport calculations (DBF/LCA). We discuss azimuthal distributions as function of transverse momentum, rapidity, centrality and beam energy. Such distributions have been
276 60 50 40
s. <
PN
30 20 10 0.05
£
0.00
O
tic
53
-0.05
Q*
^V
-0.10 -0.15
1
10
Figure 2. Flow per particle Fy/A and elliptic flow V2 at mid-rapidity (|AV(°'| < 0.1) in semi-central Au + Au reactions from SIS to AGS energies. The transport calculations n with microscopic DB mean fields in the LCA (solid) and LDA (dashed) approaches are compared to the data (symbols) 2>4.
parametrized in terms of a Fourier series dN/d
quantity Fy =
d
as the slope of the mean in-plane transy(»)=o verse momentum at mid rapidity, vi is often referred to as the elliptic flow. Its sign describes the transition from out-of-plane (yi < 0) to in-plane flow (v2 > 0) 4 . The squeeze-out ratio RN is connected to the elliptic flow through RN = (1 — 2a 2 )/(l + 2v2). Fig. 2 shows the observables Fy/A and v2 for Au + Au collisions from SIS to AGS energies. Rather complete data are available from the EOS and FOPI collaborations, which are, however, not completely consistent for w2- It is seen that the calculations n reproduce the flow Fy only for SIS energies, while at AGS energies the calculations oversti-
277
-
1
0
1
Y /Y . cnr
proj
2
-
1
0
1
2
Y /Y . cm
proj
Figure 3. Rapidity distributions for Au + Au collisions at 400 AMeV selected according to bins of ERAT which is a measure of centrality (increasing with increasing centrality). Calculations are shown for the LCA (solid) and LDA (dashed) approximations for the self energies and compared with the data (symbols) 13 .
mate the data. This is to be expected because the DB self energies do not reproduce the empirical saturation of the real part of the optical potential at energies above about 1 GeV, but rather lead to linearly increasing repulsion. This situation also appears in non-relativistic descriptions using momentum dependent Skyrme-parametrizations (see 1 2 ) . The difference between the LDA and LCA descriptions is not significant because the DBF self energies are not strongly momentum dependent at saturation density. The description of the elliptic flow is generally much better because as a ratio it is not so strongly dependent on the absolute transverse flow. We see that both models generally reproduce the qualitative trend of the data, i.e. the transition from small v2 at 0.2 A.GeV to a preferential out-of plane flow (v2 < 0) which is maximal at 0.4 A.GeV with the subsequent transition back to in-plane flow around 4 A.GeV. There is now, however, a pronounced difference between the two approximations. The calculation without non-equilibrium effects (LDA) shows a larger squeeze-out effect and also has a larger transition energy from a out-of-plane to an in-plane emission of participant matter. This effect is reduced for the LCA calculation which can be understood by the effective softening of the equation of state during the initial non-equilibrium phase of the collision as discussed in connection with Fig. 1. We also mention that the observed difference is of the same magnitude as the difference between a soft and hard EOS (see Fig. 2 of Ref. 1 2 ) . Furthermore the inclusion of the non-equilibrium effects considerably improves the agreement with the data
278
which means that the underlying effective interaction (DB) is able to describe the dynamics of heavy ion collisions. In Fig. 3 we show rapidity distributions for Au + Au collisions at 400 AMeV selected according to a quantity ERAT, the ratio between transverse and longitudinal total momentum, which has been shown to be a measure of centrality 1 3 . The rapidity distributions are reasonably well reproduced, in particular for central collisions (large ERAT). A general tendency is that the spectator particles observed at y ( ° ' ~ 1 in the experiment are more stopped in the calculations. This effect is larger in LDA than LCA, because the field is more repulsive in the former case (see Fig. 1). To summarise the effective equation-of-state probed by the compression phase in energetic heavy ion reactions is in a significant way governed by local non-equilibrium. The anisotropy of the phase space lowers the binding energy per particle and makes the effective EOS seen in heavy ion reactions softer. This fact is reflected in the behaviour of collective flow phenomena. We conclude that "geometric" phase space effects should be taken into account on the level of the effective interaction when conclusions on the equilibrium EOS are drawn from heavy ion collisions. The comparison with experiments shows that microscopic many-body methods methods that are successfull for nuclear structure also explain heavy ion reactions quantitatively at SIS energies 14 . References 1. H. Stocker and W. Greiner, Phys. Rep. 137 (1986) 277. 2. N. Bastid et al., contributions to this conference; A. Andronic et al., Nucl. Phys. A661 (1999) 333c. 3. P. Danielewicz, Nucl. Phys. A673 (2000) 375. 4. C. Pinkenburg et al., Phys. Rev. Lett. 83 (1999) 1295. 5. W. Botermans, R. Malfliet, Phys. Rep. 198 (1990) 115. 6. W. Cassing et al., Nucl. Phys. A672 (2000) 417; M. Effenberger et al., Phys. Rev. C60 (1999) 051901; R. Ionescu et al., nucl-th/0009022. 7. H. Elsenhans, L. Sehn, A. Faessler, H. Muther, N. Othsuka, H.H. Wolter, Nucl. Phys. A536 (1992) 750. 8. C. Fuchs, L. Sehn, H.H. Wolter, Nucl. Phys. A601 (1996) 505. 9. T. Gaitanos, C. Fuchs, H. H. Wolter, Nucl. Phys. A650 (1999) 97. 10. T. Gross-Boelting, C. Fuchs, and A. Faessler, Nucl. Phys. A648 (1999) 105. 11. C. Fuchs, T. Gaitanos, A. Faessler, H.H. Wolter, in preparation. 12. P. Danielewicz, Roy A. Lacey, et al., Phys. Rev. Lett. 81 (1998) 2438. 13. W. Ramillien et al., Nucl. Phys. A587 (1995) 802. 14. C. Fuchs, E. Lehmann, R.K. Puri, L. Sehn, A. Faessler, H.H. Wolter, J. Phys. G22 (1996) 131.
T H E R M O D Y N A M I C A L D E S C R I P T I O N OF HEAVY ION COLLISIONS T. GAITANOS, H. H. WOLTER Sektion Physik, Universitat Munchen, Am Coulombwall 1 D-85748 Garching, Germany E-mail: [email protected] C. FUCHS Institut fur Theoretische Physik, Universitat Tubingen, Auf der Morgenstelle 14 D-72076 Tubingen, Germany We analyze the thermodynamical state of nuclear matter in transport descriptions of heavy ion reactions. We determine thermodynamical variables from an analysis of local momentum space distributions and compare to blast model parameters from an analysis of fragment energy spectra. These descriptions are applied to spectator and fireball matter in semi-central and central Au+Au collisions at SISenergies, respectively.
A topic of great interest in the study of intermediate energy heavy-ion collisions is the investigation of the nuclear equation of state away from saturation at lower densities and at finite temperature in connection to possible signals of phase transitions 1 ' 2 ' 3 , 4 . Theoretical multifragmentation models 5 have addressed this question , but their application in the dynamical situation of heavy-ion collisions is difficult due to non-equilibrium 6 and finite size effects 7 . The question is whether the nuclear matter in heavy-ion collisions is both in thermodynamical equilibrium and instable, i.e. whether the fragment emitting source can be characterized by a thermodynamically well defined freeze-out configuration. In this work we investigate this question by studying the results of transport calculations described in detail in Refs. 8 ' 9 . We have applied thermodynamical analyses to spectator and participant matter using either the information of the local momentum distribution to define a local thermodynamical temperature (Tjoc) or from a blast model analysis 10 of fragment energy spectra to obtain a slope parameter (Tgjope) of the fragment spectra. We compare these "theoretical data" with experiments. Fig. 1 on the left shows the temporal evolution of the local temperature T| oc in spectator matter for semicentral (6 = 4.5 fm) Au+Au reactions at different energies, as obtained from fits to the local momentum space distribution of transport calculations. After the spectators are well defined TJ0C approaches rather constant values and remains fairly stable for several fm/c. The local
279
280
, . , . , . is
•
A
•
gio
\
\
• — • 0.25 AGeV •—»f)A AGeV • - • 0 . 6 AGeV A--*0.8 AGeV
25 •
I V OAIADIN
£» S.10
i H s
: *
30
:
^20
40
50
60
•
i
70
SO
| •Ttoc
|
0 90
0
2
4
«
8
10
time [ftn/c] fragment mass Af Figure 1. Left: local temperatures 7 / o c as function of time in spectators for semi-central Au+Au reactions. Right: slope temperatures Tsiope for spectator fragments as function of Af with comparison to ALADIN data 3 . The local temperature of nucleons Tj o c is also shown.
temperature is independent on the incident energy indicating the existence of an intermediate equilibrium state. By considering an effective compressibility K ~ dP/dp 8 and instability by K < 0 we find that after t ~ 40 - 45 fm/c the spectator enterns into an instability region 8 and therefore may break up into fragments. The breakup conditions of Tj oc ~ 5 — 6 MeV and p ~ {\ — \)psat are consistent with experimental isotope temperatures and densities extracted from 2-particle correlations 3 , n . The generation of fragments by a phase space coalescence in the spectator matter and an analysis of the fragment energy spectra in the blast model 10 leads to the results of Fig. 1 on the right. The independence of Tsiope on Af indicates that the fragments are emitted from an equilibrated source. However, the slope temperature of the fragments is higher compared to Tsiope of the nucleons, which in turn is equal to Tj oc . This is understood by the Fermi-motion of the nucleons in the fragmenting source in the spirit of the Goldhaber model, as discussed in Refs. 3>9>12>13. The slope parameter of the fragments is thus consistent with the nucleon Tsiope and also with the local temperature 7j o c . We thus conclude the existence of a common freeze-out configuration for nucleons and fragments in spectator decay. This interpretation is also supported by the experiments 3 shown in the Fig. 1. The results of the same analysis for participant matter in central heavyion reactions are summarized in Fig. 2 where a radial flow velocity /?/ appears as an additional parameter in the blast model fits due to the radial expansion of the fireball. We observe that slope temperatures rise and flow velocities fall with increasing fragment mass in contrast to the behavior for spectator fragments in Fig. 1. A similar behavior has been seen experimentally at 1 AGeV in 14 and theoretically in 15 - 16 . This behavior cannot be interpreted as fragments originating from a common freeze-out state, i.e. from a unique fragmenting
281 100 80 l>
-1
1
1
1
100
r-
fragments blast model
nucleons/local
80
60 "J
60
S ,
1
401-
40
H " 20
H
1
I
1
1
H
11111111111111 [ 111111111
0.4
0/4
u.
0J
0J^ 0.2 _ l
I
n i
l_
0.1 1
2
3
4
5
6
0
5
10
IS
20
fragment mass tfnra-out-' P™^] Figure 2. (left) Slope temperatures and radial flow as function of fragment mass, (right) Local values from the momentum distributions at times before the freeze-out for a central Au+Au reaction at 0.6 AGeV.
source. To arrive at an interpretation we have compared these results with the local temperatures and flow velocities for different times before the nucleon freeze-out, i.e. for t' = tfreeze-out — t with tfreeze-out ~ 35 fm/c. It is seen that for Af = 1 the values at freeze-out are close to the blast model ones, as required. However, for fragment masses Af > 1 the slope temperatures and velocities behave qualitatively very similar to the local temperatures and flow velocities at earlier times. This would suggest to interpret the fragment temperatures and velocities as signifying that heavier fragments originate at times earlier than the nucleon freeze-out. This may not be unreasonable since in order to make a heavier fragment one needs higher densities which occur at earlier times and hence higher temperatures. However, this does not neccessarily imply that the fragments are really formed at this time, since fragments could hardly survive such high temperatures, as also discussed in refs. 10 ' 17 . But it could mean that these fragments carry information about this stage of the collision. In any case it means that in the participant region fragments are not formed in a common equlibrated freeze-out configuration, and that in such a situation slope temperatures have to be interpreted with great caution. In summary fragmentation phenomena in heavy ion collisions are studied as a means to explore the phase diagram of hadronic matter. For this it is neccessary to determine the thermodynamical properties of the fragmenting source. One way to do this experimentally is to investigate fragment kinetic
282
energy spectra. In theoretical simulations the thermodynamical state can be obtained locally in space and time from the phase space distribution. In this work we have compared this with the information obtained from the generated fragment spectra. We apply this method to the spectator and participant regions of relativistic Au+Au-collisions. We find that the spectator represents a well developed, equilibrated and instable fragmenting source. In the participant region the local temperature at the nucleon freeze-out and the slope temperatures from fragment spectra behave differently from those of the spectator. The slope temperatures rise with fragment mass which might indicate that the fragments are not formed in a common, equilibrated source. References 1. 2. 3. 4. 5.
6.
7.
8. 9. 10. 11. 12. 13. 14. 15. 16. 17.
J. Richert, P. Wagner, nucl-th/0009023 (to appear in Phys. Rep.). J. Pochodzalla, Prog. Part. Nucl. Phys. 39, 443 (1997). T. Odeh et al., Phys. Rev. Lett. 84, 4557 (2000). Proc. of the International Workshop XXVII on Gross Properties of Nuclei and Nuclear Excitations, Hirschegg, Austria, 1999. J.P. Bondorf et al., Nucl. Phys. A 443, 321 (1985); A. S. Botvina, D.H.E. Gross, Nucl. Phys. A 592, 257 (1995) and contribution to this conference. H.H. Wolter, C. Fuchs, T. Gaitanos, Prog. Part. Nucl. Phys. 42, 137 (1999); T. Gaitanos, C. Fuchs, H.H. Wolter, Nucl. Phys. A 650, 97 (1999). PH. Chomaz, F. Gulminelli, Nucl. Phys. A 647, 153 (1999); M. D'Agostino et al., Phys. Lett. B 473, 219 (2000) and contribution to this conference. C. Fuchs, P. Essler, T. Gaitanos, H.H. Wolter, Nucl. Phys. A 626, 987 (1997). T. Gaitanos, H.H. Wolter, C. Fuchs Phys. Lett. B 478, 79 (2000). P.J. Siemens, J.O. Rasmussen, Phys. Rev. Lett. 42, 880 (1979); W. Reisdorf et al., Nucl. Phys. A 612, 493 (1997). S. Fritz et al., Phys. Lett. B 461, 315 (1999). A. S. Goldhaber, Phys. Lett. B 53, 306 (1974); Phys. Rev. C 17, 2243 (1978). W. Bauer, Phys. Rev. C 51, 803 (1995). M. Lisa et al., Phys. Rev. Lett. 75, 2662 (1995). F. Daffin, K. Haglin, W. Bauer, Phys. Rev. C 54, 1375 (1996). A. Hombach et al., Eur. Phys. J. A 5, 157 (1999). W. Neubert, A.S. Botvina, Eur. Phys. J. A 7, 101 (2000).
Section IV Statistical and Dynamical Aspects of Fragmentation
E X A C T SOLUTION OF T H E STATISTICAL MULTIFRAGMENTATION MODEL A N D T H E LIQUID-GAS M I X E D P H A S E K. A. BUGAEV1'2, M. I. GORENSTEIN1'2,1. N. MISHUSTIN1'3'4 and W. GREINER1 1
Institut fur Theoretische Physik, Universitat Frankfurt, Germany 2 Bogolyubov Institute for Theoretical Physics, Kyiv, Ukraine 3 Kurchatov Institute, Russian Research Center, Moscow, Russia Niels Bohr Institute, University of Copenhagen, Denmark
An exact analytical solution of the statistical multifragmentation model is found in thermodynamic limit. The system of nuclear fragments exhibits a 1-st order liquidgas phase transition. The peculiar thermodynamic properties of the model near the boundary between the mixed phase and the pure gaseous phase are studied. The results for the caloric curve and specific heat are presented and a physical picture of the nuclear liquid-gas phase transition is formulated. Nuclear multifragmentation is one of the most interesting and widely discussed phenomena in intermediate energy nuclear reactions. The statistical multifragmentation model (SMM) (see 1 ' 2 and references therein) was recently applied to study the relationship of this phenomenon to the liquid-gas phase transition in nuclear matter 3 ' 4 ' 5 . Numerical calculations within the canonical ensemble exhibited many intriguing peculiarities of the finite multifragment systems. However, the investigation of the system's behavior in the thermodynamic limit was still missing. Therefore, there was no rigorous proof of the phase transition existence, and the phase diagram structure of the SMM was unknown. Previous numerical studies for the finite nuclear systems (the canonical and microcanonical calculations) lead to the speculative (and sometimes wrong) picture of the nuclear liquid-gas phase transition in the thermodynamic limit. In our recent paper 6 an exact analytical solution of the SMM is found within the grand canonical ensemble which naturally allows to study the thermodynamic limit. The self-consistent treatment of the excluded volume effects is an important part of this study. In this letter we investigate the peculiar thermodynamic properties near the boundary between the mixed phase and the pure gaseous phase. New results for the caloric curve and specific heat are presented and a physical picture of the nuclear liquid-gas phase transition in SMM is formulated. This physical picture differs from that expected in the previous numerical studies. The system states in the SMM are specified by the multiplicity sets {n*} 285
286
(rik = 0,1,2,...) of fc-nucleon fragments. The partition function of a single fragment with k nucleons i s 1 : ujk — V (mTk/2Tr) ' zk , where k = 1,2, ...,A (A is the total number of nucleons in the system), V and T are, respectively, the volume and the temperature of the system, m is the nucleon mass. The first two factors in w* originate from the non-relativistic thermal motion and the last factor, Zk, represents the intrinsic partition function of the ^-fragment. For k — 1 (nucleon) we take z\ = 4 (4 internal spin-isospin states) and for fragments with k > 1 we use the expression motivated by the liquid drop model (see details in Ref. 1): Zk = exp(—fk/T), with fragment free energy fk =
-[W0
+ T2/eQ]k
+ a{T)03
+ rTlnfc.
(1)
Here W0 = 16 MeV is the bulk binding energy per nucleon, T 2 /e 0 is the contribution of the excited states taken in the Fermi-gas approximation (e0 = 16 MeV) and cr(T) is the surface tension which is parameterized in the following form: a{T) = <70[(T2 - T 2 )/(T C 2 + T 2 )] 5 / 4 , with a0 = 18 MeV and Tc = 18 MeV (CT = 0 at T > Tc). The last Fisher's term in Eq. (1) with dimensionless parameter r is introduced for generality. The canonical partition function (CPF) of nuclear fragments in the SMM has the following form:
Z%{V,T) = £ {nh}
I I ^jSiA-^knk). *=1
*-
(2)
k
The SMM defined by Eqs.(l,2) with r = 0 was studied numerically in Refs. 3,4 ' 5 . This is a simplified version of the SMM, e.g. the symmetry and Coulomb contributions are neglected. However, its investigation appears to be of a principal importance for studies of the liquid-gas phase transitions. In Eq. (2) the nuclear fragments are treated as point-like objects. However, these fragments have non-zero proper volumes and they should not overlap in the coordinate space. In the excluded volume (Van der Waals) approximation this is achieved by substituting the total volume V in Eq. (2) by the free (available) volume V) = V - fc£k knk, where b = l/p0 {p0 = 0.16 f m - 3 is the normal nuclear density). Therefore, the corrected CPF becomes: ZA{V,T) =
Z%(V-bA,T). The calculation of ZA(V, T) is difficult because of the constraint ^ f c krik = A. This difficulty can be partly avoided by calculating the grand canonical partition function: 00
Z(V,T, A*) =
£ A=0
exp (fiA/T)
ZA(V,T)
Q(V - bA) ,
(3)
287
where chemical potential \i is introduced. The calculation of Z is still rather difficult. The summation over {n*,} sets in ZA cannot be performed analytically because of additional A-dependence in the free volume V/ and the restriction V) > 0. The problem can be solved by introducing the so-called isobaric partition function (IPF) which is calculated in a straightforward way (see details in Refs. 6,? ): /•OO
Z(s,T,p)
=
/ Jo
dV exp(-sV) Z(V,T,p)
=
s -
J-(s,7>)
(4)
where
*-™-(£r
z\ exp
(s^) (5)
fc=2
^
'
with v = n + WQ + T2/ea- In the thermodynamic limit V —¥ oo the pressure of the system is defined by the farthest-right singularity, s*(T,(i), of the IPF £(*,T,M) p{T,p)
= T
lim
l n Z( T ]
l' ^
= T s*(7» .
(6)
The study of the system's behavior in the thermodynamic limit is therefore reduced to the investigation of the singularities of Z. The IPF (4) has two types of singularities: 1) the simple pole singularity defined by the equation sg(T, p) = .F(s g , T, /x) ; 2) the singularity of the function T itself at the point sj(T,/i) = v/Tb where the coefficient in linear over fc terms of the exponent in Eq. (5) is equal to zero. The simple pole singularity corresponds to the gaseous phase where pressure pg = Tsg is determined by the following transcendental equation: pg (T, /x) = TF{pglT,T,n). The singularity «j(T,/x) of the function T defines the liquid pressure: pi{T, /x) = Tsi(T,/x) = v/b. Here the liquid is represented by an infinite fragment with fc — oo. In the region of the (T, jx)-plane where v < bpg(T,fj,) the gaseous phase is realized [pg > pi), while the liquid phase dominates at v > bpg(T,p). The liquid-gas phase transition occurs when the two singularities coincide, i.e. sg(T,p) = si(T,fi). As T in Eq. (5) is a monotonously decreasing function of s the necessary condition for the phase transition is that the function T is finite in its singular point sj. At r = 0 this condition requires a(T) > 0
288
and, therefore, T
«
+f>w<*-'*«>*-''tn fc=9
^
(7)
' -
A similar expression for pg within the excluded volume model for the pure nucleon gas was obtained in Ref. 8 . The phase transition line p*(T) in the (T, ^)-plane corresponds to the mixed liquid and gas states. This line is transformed into the finite mixed-phase region in the (T,p)-plane shown in Fig. 1. The baryonic density in the mixed phase is a superposition of the liquid and gas baryonic densities: p = Xpi + (1 — \)pg , where A ( 0 < A < l ) i s a fraction of the system's volume occupied by the liquid inside the mixed phase. Similar linear combinations are also valid for the entropy density * and the energy density e with (i = I, g) st = (dpi/dT)^, 6i = T (dpi/dT)^ + p (dpi/dp,)T - p,-. Inside the mixed phase at constant density p the parameter A has a specific temperature dependence shown in Fig. 2: from an approximately constant value plp0 at small T the function A(T) drops to zero in a narrow vicinity of the boundary separating the mixed phase and the pure gaseous phase. This corresponds to a fast change of cluster configurations from the state which is dominated by one infinite liquid cluster to the gaseous multicluster configurations. It happens without discontinuities of the thermodynamical functions inside the mixed phase. An abrupt decrease of A(T) near this boundary causes a strong increase of the energy density as a function of temperature. This is evident from Fig. 3 which shows the caloric curves at different baryonic densities. One can clearly see a leveling of temperature at energies per nucleon between 10 and 20 MeV. As a consequence this leads to a sharp peak in the specific heat per nucleon at constant density, cp{T) = (de/dT)p/p , presented in Fig. 4. A finite discontinuity of cp (T) arisen at the boundary of the mixed and gaseous phases
289
0
S
10 T (MeV)
15
Tc 20
Figure 1: Phase diagram in the (T, p)-plane. The mixed phase and pure gaseous phase boundary is shown by the solid line. The pure liquid phase (shown by crosses) corresponds to the fixed density p — pQ. Point C is the critical point, at T > T c only the pure gaseous phase exists.
is shown in Fig. 4. This finite discontinuity is due to the fact that A(T) = 0, but (d\/dT)p ^ 0 at this boundary inside the mixed phase (see Fig. 2). It should be emphasized that the energy density is continuous at the boundary of the mixed phase and the gaseous phase, hence the sharpness of the peak in cp is entirely due to the strong temperature dependence of X(T) near this boundary. Therefore, in contrast to the expectation in Refs. 4 ' 5 , the maximum value of cp remains finite and the peak width in cp(T) is not zero in the thermodynamic limit considered in our study.
290
Figure 2: Volume fraction A(T) of the liquid inside the mixed phase is shown as a function of temperature for fixed nucleon densities p/po = 1/6,1/3,1/2,2/3,5/6 (from bottom to top).
5
10
15
20
25
30
e/p + W0 (MeV) Figure 3: Temperature as a function of energy density per nucleon (caloric curve) is shown for fixed nucleon densities pjpQ = 1/6,1/3,1/2,2/3.
291
5
10
T (MeV)
Figure 4: Specific heat per nucleon as a function of temperature at fixed nucleon density pjpQ •=. 1/3. The dashed line shows the finite discontinuity of cp(T) at the boundary of the mixed and gaseous phases.
292
In conclusion, the simplified version of the SMM is solved analytically in the grand canonical ensemble. The progress is achieved by reducing the description of phase transitions to the investigation of the isobaric partition function singularities. The model clearly demonstrates a 1-st order phase transition of the liquid-gas type. The considered system has peculiar properties near the boundary of the mixed and gaseous phases. They reveal themselves in a fast change of cluster configurations from the state which is dominated by one infinite liquid cluster to the gaseous multicluster configurations. It happens without discontinuities of the thermodynamical functions inside the mixed phase near its boundary with a gaseous phase. In the caloric curves shown in Fig. 3 a leveling of temperature is observed at energies per nucleon between 10 and 20 MeV. Acknowledgments The authors are thankful to A.S. Botvina, Ph. Chomaz, D.H.E. Gross, A.D. Jackson and J. Randrup for useful discussions. We are thankful to the Alexander von Humboldt Foundation and DAAD, Germany for the financial support. The research described in this publication was made possible in part by Award No. UP1-2119 of the U.S. Civilian Research & Development Foundation for the Independent States of the Former Soviet Union (CRDF). 1. J.P. Bondorf, A.S. Botvina, A.S. Iljinov, I.N. Mishustin, K.S. Sneppen, Phys. Rep. 257, 131 (1995). 2. D.H.E. Gross, Phys. Rep. 279, 119 (1997). 3. K.C. Chase and A.Z. Mekjian, Phys. Rev. C 52, R2339 (1995). 4. S. Das Gupta and A.Z. Mekjian, Phys. Rev. C 57, 1361 (1998). 5. S. Das Gupta, A. Majumder, S. Pratt and A. Mekjian, nucl-th/9903007 (1999). 6. K.A. Bugaev, M.I. Gorenstein, I.N. Mishustin and W. Greiner, Phys. Rev. C62, 044320 (2000). 7. M.I. Gorenstein, V.K. Petrov and G.M. Zinovjev, Phys. Lett. B 106, 327 (1981); M.I. Gorenstein, W. Greiner and S.N. Yang, J. Phys. G 24, 725 (1998). 8. D. H. Rischke, M. I. Gorenstein, H. Stocker and W. Greiner, Z. Phys. C 51, 485 (1991).
WHAT CAN W E LEARN FROM NUCLEAR INSTABILITIES?
Laboratorio
MATTER
V. B A R A N ° , M. C O L O N N A , M. DI T O R O Nazionale del Sud, Via S. Sofia 44, 1-95123 Catania, NIPNE-HH, Bucharest, Romania
Smith
Sektion
Italy
and
M. Z I E L I N S K A - P F A B E College, Northampton, USA
Physik,
H.H. W O L T E R University of Munich,
Germany
We discuss the features of instabilities in binary systems, in particular, for asymmetric nuclear matter. We show its relevance for the interpretation of results obtained in experiments and in "ab initio" simulations of the reaction between 124 Sn + 1 2 4 Sn at 50AMeV.
1
Instabilities in asymmetric nuclear matter
T h e process of multifragmentation following the collision of heavy nuclei in the region of m e d i u m energies displays several features analogous to usual liquidgas phase transitions of water. However in this analogy one should by aware of differences due to Coulomb, finite size and q u a n t u m effects as well as to the binary, i.e. two-component, character of nuclear m a t t e r . Moreover, the time scales of the process are of the order of, or shorter t h a n , the relaxation times of the relevant degrees of freedom. Therefore we have to consider not only equilibrium phase-transition in binary systems, b u t in an equally i m p o r t a n t way the dynamical evolution driving such phase transition and its dependence on the symmetry t e r m of the equation-of-state. Indeed after a fast compression and expansion we expect the nuclear system to be quenched into an unstable state either inside the coexistence curve in the metastability region (where the phase is unstable against short wave length but large amplitude fluctuations) or in the instability (spinodal) region (were the system becomes unstable against long wave length b u t small amplitude fluctuations). T h e n the system will evolve toward a stable thermodynamical state of two coexisting phases either through nucleation (in former case) or through spinodal decomposition (in the latter case). These aspects were discussed repeatedly in the p a s t , 1 ' 2 , 3 ' 4 , 5 ' 6 but the binary character will induce new features for b o t h scenarios t h a t are absent in one-component systems. Therefore studying the nature of instabilities which "on leave from NIPNE-HH, Bucharest, Romania
293
294 develop in nuclear systems, their relaxation times and the influence of the isospin degree of freedom will provide information on the mechanisms involved in the fragment formation processes and will set limits for the applicability of fully equilibrium approaches. In the framework of Landau theory for two component Fermi liquids the spinodal border was determined by studying the stability of collective modes described by two coupled Landau-Vlasov equations for protons and neutrons. In terms of the appropriate Landau parameters the stability condition can be expressed a s 7 ,
(l + C ) ( l + O - ^ T > 0 '
(!)
It is possible to show t h a t this condition is equivalent to the following t h e r m o dynamical condition
'?P\ d
PJr,y
(dtA V dy
>0
(2)
JTP
discussed i n 8 , 9 , where y is the proton fraction. In Fig. 1 we show the spinodal lines obtained from eq. (1) (continuous line with dots) which for asymmetric nuclear m a t t e r is seen to contain the lines corresponding to "mechanical instability", ( f r H < 0 (crosses). Therefore eqs. (1,2) describe the "chemical instability" of nuclear m a t t e r .
Q f l l ) I I I I I I I I I I I I II I I I I I I I T I
0.00 0.02 0.04 0.06 0.08 0.10 0.00 0.08 0.0* 0.06 0.06 0.10 0.00 0.02 0.04 0.06 0.0B 0.10
p{tm-)
/>(fcO
p{tm-*)
Figure 1: Spinodal lines corresponding to chemical (joined points) and mechanical (crosses) instability for three asymmetries, I=:(N-Z)/(N+Z)=0.0,0.5,0.8 of nuclear matter.
We want to stress, however, t h a t by just looking at the above stability conditions we cannot determine the n a t u r e of the fluctuations against which a binary system becomes chemically unstable. Indeed, the thermodynamical condition in eq. (2) cannot distinguish between two very different situations
295 which can be encountered in nature: an attractive interaction between the two components of the mixture (FQP , FQ" < 0), as is the case of nuclear m a t t e r , or a repulsive interaction between the two species. We define as isoscalar-like density fluctuations the case when proton and neutron Fermi spheres (or equival e n t ^ the proton and neutron densities) fluctuate in phase and as isovector-like density fluctuations when the two Fermi sphere fluctuate out of phase. Then it is possible to prove, based on a thermodynamical approach of asymmetric Fermi liquid mixtures 1 0 , t h a t chemical instabilities are triggered by isoscalar fluctuations in the first, i.e. attractive, situation and by isovector fluctuations in the second one. For the asymmetric nuclear m a t t e r case because of the attractive interaction between protons and neutrons the phase transition is thus due to isoscalar fluctuations t h a t induce chemical instabilities while the system is never unstable against isovector fluctuations. Of course the same attractive interaction is also at the origin of phase transitions in symmetric nuclear m a t t e r . However, in the asymmetric case isoscalar fluctuations lead to a more symmetric high density phase everywhere under the instability line defined by eq.(l) 7 .
Figure 2: 1 2 4 S n -f124 Sn 50AMeV: time evolution of the nucleon density projected on the reaction plane. First two columns: b = 2Jm collision, approaching, compression and separation phases. Third and fourth columns: 6 = 4 / m and 6 = 6 / m , separation phase up to the freeze-out.
296
2
"Ab initio" simulations of Sn + Sn reactions
A new code for the solution of microscopic transport equations, the Stochastic Iso-BNV, has been written where asymmetry effects are suitably accounted for and the dynamics of fluctuations is included 11,12 . We have studied the bOAMeV collisions of the systems 124Sn +124 Sn and 112 Sn + 1 1 2 Sn, where new data have been measured at NSCL — MSU 13 . We will discuss averages over 100 events generated in semi-central (6 = 2/m) and semi-peripheral (6 = 6/m) reactions. In Fig.2 we show the evolution with impact parameter of the density plot (projected on the reaction plane) for one event each (neutron rich case 12ASn + 1 2 4 Sn, EOS with stiff asymmetry term) 14
We remark: i) In the cluster formation we see a quite clear transition with impact parameter from bulk4'5,6 to neck15'16,17 instabilities, ii) The "freezeout times", i.e. when the nuclear interaction among clusters disappears, are decreasing with impact parameter. These two dynamical effects will influence the isospin content of the produced primary fragments, as shown below.
t IP*/*)
Figure 3: 124 Sn -)- 124 Sn 50AMeV erties. See text.
tfra/e)
I (to/,)
b = 2/ro collisions: time evolution and freeze-out prop-
A detailed analysis of the results for the same system is shown in Fig. 3 (central, 6 = 2/m) and Fig.4 (peripheral, 6 = 6/m). Each figure is organized in the following way: Top row, time evolution of: (a) Mass in the liquid (clusters with Z > 3, upper curve) and the gas (lower curve) phase; (b) Asymmetry I = (N — Z)/(N + Z) in gas "central" (solid line with squares), gas total (dashed+squares), liquid "central" (solid+circles) and IMF's (3 < Z < 12, stars). "Central" means in a cubic box of side 20fm around the c m . . The horizontal line shows the initial average asymmetry; (c) Mean Fragment Multi-
297
Figure 4: 124Sn +124 Sn 50AMeV erties. See text.
b = 6 / m collisions: time evolution and freeze-out prop-
plicity Z > 3. The saturation of this curve defines the freeze-out configuration, as we can also check from the density plots in Fig.2. Bottom jow, properties of the "primary" fragments in the freeze-out configuration: (A) Charge Distribution, (e) Asymmetry Distribution and (f) Fragment Multiplicity Distribution (normalized to 1). ' We see a neutron dominated prompt particle emission and a second neutron burst at the time of fragment formation in the "central region". The latter is consistent with the dynamical spinodal mechanism in dilute asymmetric nuclear matter, as discussed before. The effect is quite reduced for semiperipheral collisions (compare Figs. 3b and 4b) and the IMF's produced in the neck are more neutron rich (Figs. 3e and 4e). This seems to indicate a different nature of the fragmentation mechanism in central and neck regions, i.e. a transition from volume to shape instabilities with different isospin dynamics. In more peripheral collisions the interaction time scale is also very reduced (Fig.4c) and this will quench the isospin migration. 3
Conclusions
Starting from the thermodynamical features of instabilities and the dynamics of phase-transitions in binary systems we obtain a consistent description of the multifragmentation process in heavy ion collisions and, in particular, of isospin effects. Isospin proves to be a useful probe in signaling a change in the fragment formation mechanism passing from central to semipheripheral collisions. Moreover this "isospin dynamics" is found to be quite sensitive to the symmetry term of the EOS, opening new stimulating perspectives on such
298
studies, which are of great astrophysical interest, under laboratory controlled conditions. Acknowledgments One of authors (V.B.) gratefully acknowledges for the warm hospitality at LNS Catania during his Postdoctoral fellowship. References 1. C.J.Pethick and D.G.Ravenhall,M/e/. Phys. A 471, 19c (1987). 2. C.J.Pethick and D.G.Ravenhall,,4rm. Phys. (NY.) 183, 131 (1988). 3. H.Heiselberg, C.J.Pethick and D.G.Ravenhall, Ann. Phys. (N.Y.) 223, 37 (1993). 4. M.Colonna and Ph.Chomaz, Phys. Rev. C 49, 1908 (1994) 5. M. Colonna, M. Di Toro and A. Guarnera, Nucl. Phys. A 580, 312 (1994). 6. A. Guarnera, M. Colonna and Ph. Chomaz, Phys. Lett. B 373, 267 (1996); M. Colonna, Nucl. Phys. A 630, 136c (1998). 7. V.Baran, A.Larionov, M.Colonna and M.Di Toro, Nucl. Phys. A 632, 287 (1998). 8. L.D. Landau and E.M. Lifshitz, Statistical Physics, pag. 208. 9. H.Mueller and B.D.Serot, Phys. Rev. C 52, 2072 (1995). 10. V.Baran, M. Di Toro, M. Colonna, Nuclear Fragmentation: sampling the nature of instabilities in binary systems, (submitted for publication). 11. M.Colonna, M.DiToro, G.Fabbri and S.Maccarone, Phys. Rev. C 57, 1410 (1998). 12. M.Colonna, M.Di Toro, A.Guarnera, S.Maccarone, M.Zielinska- Pfabe', H.H.Wolter, Nucl. Phys. A 642, 449 (1998) 13. H.S.Xu et al, Phys. Rev. Lett. 85, 716 (2000). 14. M.Di Toro et al, CRIS2000 Conference, Nucl. Phys. A (in press). 15. M.Colonna, M.Di Toro and A.Guarnera, Nucl. Phys. A 589, 160 (1995). 16. M.Colonna et al., Perspectives in heavy ion physics: 3rd Italy-Japan joint meeting, Ed.s C.Signorini et al., World.Sci,1999 pp.172-185. 17. D. Durand, Nucl. Phys. A 630, 52c (1998).
ENERGETIC PROTON EMISSION AND REACTION DYNAMICS IN HEAVY ION REACTIONS CLOSE TO T H E FERMI ENERGY
R. CONIGIiONE, P. SAPIENZA, E. MIGNECO, C. AGODI, R. ALBA, G. BELLIA, M. COLONNA, A. DEL ZOPPO, P. FINOCCHIARO, V. GRECO, K. LOUKACHINE, C. MAIOUNO, P. PIATTELLI, D. SANTONOCITO AND P.G. VENTURA INFN Laboratori Nazionali del Sud and Dipartimeto di Fisica, Via S. Sofia 44, 95124 Catania, Italy E-mail [email protected] N. COLONNA INFN Sezione di Bari, Bari, Italy M. BRUNO, M. D'AGOSTINO, M.L.FIANDRI AND G. VANNINI Dipartimento di Fisica, Bologna, Italy P.F.MASTINU AND F. GRAMEGNA INFN Laboratorio Nazionale di Legnaro, Padova, Italy I.IORL L.FABBIETTI AND AMORONI INFN Sezione di Milano and Dipartimento di Fisica, Milano, Italy G.V.MARGAGLIOTTL P.M.MILAZZO, R.RUIAND F.TONETTO, INFN Sezione di Trieste and Dipartimento di Fisica, Trieste, Italy Y. BLUMENFELD AND J.A. SCARPACI Institute de Physique Nucleaire, IN2P3-CNRS-F-91406 Orsay, France The energetic proton emission has been investigated in the 58Ni + 58Ni at 30 MeV/u and 40Ar + V at 44 MeV/u reactions. Information on the origin of the energetic protons and on basic ingredients of the BNV models such as the mean field interaction and the elementary nucleonnucleon cross section was extracted. Extremely energetic protons were measured and their impact parameter dependence indicates the presence of cooperative processes.
51
1
Introduction
Heavy ion collisions at intermediate energy have been shown to be a good tool to study the properties of nuclear matter out of equilibrium. Dynamical calculations show that in the early non equilibrated stage of the reaction, where the nuclear matter is under extreme conditions of pressure and/or temperature, particles such as subthreshold mesons or energetic photons and nucleons can be emitted. On the 299
300
other hand, the experimental observation of extremely energetic nucleons or deep subthreshold particles, not yet reproduced by calculations, addresses the question of what mechanisms, beyond the NN collisions, could enable to concentrate a relevant fraction of the available energy in the production of a single energetic or massive particle. Moreover, since the reaction dynamics in heavy ion collisions at intermediate energy are ruled by the interplay of mean field and two body collisions, experimental observable sensitive to basic ingredients of the calculations such as the nucleon-nucleon (NN) cross section and the mean field potential are eagerly awaited. In this paper we present some results concerning the study of energetic and extremely energetic proton emission in heavy ion reactions around the Fermi energy which can supply information on the dynamics at pre-equilibrium and on the emission of particles with energy much larger than that provided by the coupling of the nucleon Fermi motion with the relative motion of the colliding nuclei.
2
Experimental set-up and results
We show results from two different sets of data, 58Ni+58Ni at 30 MeV/u and 40 Ar+51V at 44 MeV/u. The 44 MeV/u data were collected at GANIL with the MEDEA detector [1], while the Ni+Ni experiment was performed at Laboratori Nazionali del Sud with the MEDEA and MULTICS [2] apparatus. MEDEA consists of a ball built up with 180 BaF2 detectors (30°<9<170°) and a phoswich wall of 120 plastic detectors (1O°<0<3O°). The BaF2 permits to detect and identify light charged particles (LCP) (Z=l,2) (Ep<300 MeV) and photons up to EY = 200 MeV. The phoswich wall permits the measurement of charged particles. The Ni+Ni experiment was performed replacing the phoswich wall with the MULTICS array which is built up with 55 telescopes, each one consisting of an Ionization Chamber, a Silicon detector and a Csl crystal. MULTICS allows the identification of charged particles up to Z=83. The total geometric acceptance was greater than 90% of 4K. High statistic experimental proton energy spectra at different laboratory angles have been measured for Ar+V at 44 MeV/u and Ni+Ni at 30 MeV/u [3,4]. The analysis in terms of three Maxwellian distributions puts in evidence the presence of a source, called intermediate source, with velocity close to the half beam velocity and high inverse slope parameter that accounts for the energetic proton emission at large polar angles. This source is usually ascribed to the pre-equilibrium emission due to quasi-free NN collisions [5]. In this hypothesis, the high inverse slope parameter originates from the random composition of the beam velocity with the Fermi motion velocity. Moreover, the observed linearity of the energetic proton
301
multiplicity as a function of the number of participant nucleons [5] and the y-p anti-correlation [6] indicates that a relevant fraction of the energetic protons and photons is emitted from first chance NN collisions in the interaction zone. According to the moving source analysis, it is possible to put constraints in the phase space (Ep >40 MeV and 9 >42°) which allow to sort out the protons emitted from the intermediate source. The selected proton spectra were then transformed in the NN reference frame (VNN = 0.5vbeam) [4,7]. In these spectra the observation of protons with energy more than twice the kinematical limit (calculated in the hypothesis of first chance NN collisions and sharp Fermi momentum distribution) is a new interesting experimental information. In heavy ion reactions at intermediate energy, the emission of extremely energetic protons is a matter of peculiar interest since they can give direct information on their production mechanism and can also provide a clue for the comprehension of deep subthreshold particle production [8]. In order to achieve a deeper understanding of the energetic proton emission, we have compared our data with BNV calculations, which provide a complete description of the reaction dynamics [9]. In particular, a detailed study of the protons with energy near the kinematical limit has been performed studying their angular distributions for peripheral and central collisions and various systems with different mass asymmetry confirming that energetic protons are mainly due to first chance NN collisions [4].
•81
50
•
•
•
J
I
•
100
•
•
•
I
•
•
150
EOT (MeV)
•
.TTT
200
1
50
,
1
.
100
,
,
,
150
I
,
200
E ^ (MeV)
Figure 1. Experimental proton energy spectra (full squares) for central collisions (see text) and relative BNV spectra with a Skyrme interaction (open squares) and with a momentum dependent Skyrme interaction (open circles) calculated at b=lfm. The arrows indicate the kinematical limits (see text)
In fig. 1 experimental and calculated proton energy spectra are reported for central reactions for Ar+V at 44 MeV/u and Ni+Ni at 30 MeV/u systems. For the Ar+V reaction the LCP multiplicity was used as a selector of the reaction centrality [10], while, for the reaction Ni+Ni at 30 MeV/u the most central events have been selected requiring a coincidence with evaporation residues (Z>28 and 0.92 vcm< v
302
<1.05 vcm). These criteria will be briefly described later on in this paper. In fig. 1 two different BNV calculations are shown: one using a local Skyrme interaction (open squares) for the mean field and another using a momentum dependent Skyrme interaction (Gale - Bertsh - Das Gupta) (GBD) (open circles). It is worth mentioning that the absolute comparison must be viewed with caution since there are uncertainties due to the impact parameter assignment in the experimental data and to the criteria used for sorting free protons in the BNV calculations. The calculations with the momentum dependent mean field [11] appear in better agreement with the data. Due to the less attractive mean field, we observe in the GBD calculations a larger fraction of escaped particles and more energetic particles. However, the calculations undershoot the experimental proton data at high energy, thus indicating that the emission of extremely energetic protons calls for the introduction of other processes. To shed some light on the emission process of extremely energetic protons we decided to investigate their impact parameter dependence in the Ni+Ni experiment. At energy as low as 30 MeV/u the fluctuations on light charge particle multiplicity can affect the determination of the impact parameter especially for the most central collisions [10].
T3
Figure 2: Angular distributions of heavy residues for different parallel velocity and different charge (full squares 4.6 < vres < 5.5 cm/nsec, open squares 3.0 < vres < 4.5 cm/nsec, full circles 2.0 < v„s < 2.9 cm/nsec). The solid line is the result of BNV + GEMINI calculations (see text).
So we decided to perform, besides the selection in terms of charged particle multiplicity, an event selection based on the detection of heavy fragments and the measure of the associated hard photon multiplicity. Indeed, the detection of heavy fragments spanning from projectile-like fragments to evaporation residues allows the selection of events with different centrality. In particular, evaporation residues with velocity close to the centre of mass velocity and charge higher than the projectile charge (Z>28) can be produced in the most central collisions. The
303
measured angular distributions, for different velocity and charge of the heavy residues are reported in fig. 2. The comparison of the experimental evaporation residues angular distributions with the results obtained coupling the evaporation code GEMINI with the outcome of the BNV calculations at b=lfm, is shown in fig. 2a (solid line) and a good qualitative agreement is observed. The agreement with the calculations indicates that this hybrid approach provides a complete and consistent description of the nuclear dynamics from the pre-equilibrium phase to the formation and decay of a fusion like fragment. In fig. 2b, the angular distributions for residues with 20
0
10
20
30
rt
part
40
50
0
10
20 A
30
40
50
part
Figure 3: (a) High energy proton multiplicity (full diamond 60<Ep<80 MeV, full circles 100<Ep<120 MeV and full squares 140<Ep<160 MeV) versus the number of participating nucleons (see text), (b) Ratio between the multiplicity of protons with 140<Ep<160 MeV energy and 60<Ep<80 MeV (full squares) and between 100<Ep<120 MeV and 60<Ep<80 MeV (full circles). The open symbols refer to the corresponding BNV calculations.
This procedure emphasises the deviations from a linear dependence and allows a better comparison with the BNV calculations. The most striking feature of fig. 3b is an almost linear increase of Mp (140) / Mp (60) versus Apa,, (full squares), which
304
corresponds to an almost quadratic dependence of the extremely energetic proton multiplicity, in contrast with the results of the BNV calculations which exhibit a rather flat trend (open squares). On the other hand, the Mp(100)/Mp(60) data (full circles) exhibit a slight dependence on Ap^, which is well reproduced by the BNV results (open circles). The high energy data point out the presence of a mechanism beyond the two body nucleon-nucleon collisions, which can reflect the onset of cooperative processes. Three body collisions can be implemented in the BNV code [9]. Calculations to evaluate the contribution of this effect are still in progress. A similar behaviour of the multiplicity as a function of the participant nucleons has been observed, at much higher incident energy, in the deep subthreshold production of r| and K+ [13] and interpreted in terms of multi-step processes involving nucleon resonances. 3
Conclusions
In summary, we have investigated the energetic proton emission in the reactions Ni+Ni at 30 MeV/u and Ar+V at 44 MeV/u. Wimin the BNV dynamical approach, a better agreement with data is found with a momentum dependent mean field. Moreover, in the Ni+Ni at 30 MeV/u reaction, the evaporation residues angular distributions are well reproduced by a hybrid model made coupling the evaporation code GEMINI to the BNV results, thus indicating that a satisfying description of the reaction dynamics is achieved. Finally the multiplicity of the extremely energetic protons versus the number of participant nucleons was investigated and an almost quadratic increase was observed, thus indicating the onset of cooperative processes. Other effects, such as the presence of high momentum tails in the nucleon momentum distribution, can play a role and should be investigated. References 1. 2. 3. 4. 5. 6. 7. 8. 9. 10. 11. 12. 13.
Migneco E. et al, Nucl. Instr. Methods A 314 (1992) 31. Iori I. et al, Nucl. Instr. Methods A 325 (1993) 458. Sapienza P. et al, XXXVIII Int. Winter Meeting on Nucl. Phys., Bormio 2000. Coniglione R. et al, Phys. Lett. B 471 (2000) 339. Alba R. et al, Phys. Lett. B 322 (1994) 38. Sapienza P. et al, Phys. Rev. Lett. 73 (1994) 1769. Sapienza P. et al, Nuovo Cimento. Vol. I l l (1999) 999. Martinez G. et al, Phys. Rev. Lett 83 (1999) 1538. Bonasera A. et al, Phys. Rep. 243 (1994) 1. Del Zoppo A. et al, Phys. Rev. C 50 (1994) 2930. Greco E. et al., Phys. Rev. C 59, Vol.2 (1999). Piattelli P. et al, Phys. Lett. B 442 (1998) 48 and reference therein. Wolf A. R. et al, Phys. Rev. Lett. 80 (1998) 5281 and reference therein.
NEW RESULTS ON PREEQUILD3RIUM y-RAY EMISSION AND GDR SATURATION ON REACTIONS AT 25A MeV G.CARDELLA 2 , F.AMORINI 1 , A.DI PIETRO 1 , P.FIGUERA 1 , G.LANZALONE 2 , J.LU 1 , A.MUSTJMARRA 4 , M.PAPA 1 , G.PAPPALARDO 1 3 , S.PIRRONE 1 , F.RIZZO 1 3 , S.TUDISCO 1 ' 3 1 - INFN - Laboratorio Nazionale del Sud, Catania, Italy 2 - INFN, Sezione di Catania, Italy 3 - Dipartimento di FisiCa, Universita di Catania, Italy 4 - Dipartimento di Metodologie Fisiche e Chimiche per 1'Ingegneria, Universita di Catania, Italy Y-rays in coincidence with heavy evaporation residues have been measured for the reactions '">Ca+4M0Ca, 46Ti at 25A MeV beam energy. A large enhancement in the Y-ray multiplicity has been found in the 40Ca+48Ca data with respect to the 40Ca+40Ca ones. Different contributions to this enhancement have been identified. One of these contributions is centred, in the spectra ratio, around 1^=10 MeV. This contribution is associated to the dinuclear dipole generated by the charge asymmetry between projectile and target in the first instants of the collision. At higher Tf-ray energy, our data show, in the Ca target case, a strong suppression of the statistical GDR y-ray decay. Whereas the data in the GDR region relative to 48Ca and Ti targets are well reproduced by statistical model calculations once the deuteron decay channel is included. An increase of the GDR width with the excitation energy is observed.
1
Introduction
In fusion reactions, an increase in yield has been observed in the y-ray spectra in the whole statistical GDR energy range [1,2], whereas in more peripheral reactions, like incomplete fusion and deep inelastic collisions, the extra-yield has been observed around 8-10 MeV[3-5]. The explanation suggested for the extra-yield observed in peripheral reactions is that the asymmetry of the (N-Z)/A ratio in the two reaction partners initiates a nucleon exchange in the first stage of the collision. This generates a pre-equilibrium dipole y-emission [5]. In this contribution we report on the evidence of such pre-equilibrium y-ray emission obtained by comparing y-ray energy spectra produced in the reactions ^Ca+^'^Ca, 46Ti at 25 A MeV. The statistical part of the y-ray spectra has been reproduced assuming a slowly increasing GDR width and including in the calculations the most important open decay channels.
2 DETECTION SYSTEM AND EXPERIMENTAL RESULTS The experiment was performed at the Superconducting Cyclotron of the Laboratori NazionaU del Sud (LNS) in Catania. We used the multidetector system TRASMA[6]. Fragments stopped in the strip detector were charge identified by 305
306
using a pulse shape analysis of the signals [7]. In fig.la,b we show the multiplicity y-ray energy spectra measured in coincidence with heavy residues.
1
:
?o i o
o w Ca+*"Ca j. • "Ca+^Ca # - m a) W% &O
j, fm
_1
a
r *• 8
e s
2
-3
-o 10 io"
:
_
o *°Ca+wCa • w Ca+**Ti b
)
•
8O
*O0 o
-*
•So-
r
V
4
,.. i , , . . i . . . , i , , , . rt^fc . , , , I , . , . I , , . , I . . . , F T * ; 10 20 30 40 10 20 30 40 Energy (MeV) Figure 1 a) Multiplicity y-ray spectra as a function of energy for the reactions Ca+ Ca (open dots), 40 Ca+40Ca (full dots)for fusion and incomplete fusion reactions, b) Same as a) but with 40Ca+46Ti data as full squares.
We compare the yields obtained for the '"'Ca+^Ca reaction (open dots) with those relative to the reactions '"'Ca+^Ca (full dots fig.2a) and Ca+46Ti (full squares fig. lb). We note in fig. la the rather low y-ray multiplicity obtained in the GDR region for the ""Ca+^Ca reaction. Although from the comparison of the reactions with Ca isotope targets it seems that the charge asymmetry between projectile and target [(N-Z)/ACCa)=0.2, (N-Z)/A("°Ca)=0.] produces a quite large enhancement from Ey~8 MeV to Ey-30 MeV, this is not observed in the comparison between the ^Ca and 46Ti targets data [(N-ZyACTi^M]. In this case, we still have a quite large projectile-target charge asymmetry but a rather different behavior is observed. The yield in the GDR region for the two reactions is quite similar even if somewhat smaller (-10-20%) for the 40Ca+46Ti reaction. An enhancement is observed in the reaction ^Ca+^Ca around Ey=8-10 MeV. The ratios between the measured y-ray spectra, shown in fig.2a,b,c evidence better the previous observations. Enhancements in different energy regions can be observed by comparing the two ratios 40Ca+48Ca/40Ca+46Ti (first panel fig.2 peak around E^IO MeV) and 40 Ca+46Ti/40Ca.+40Ca (second panel fig.2 step rise in the spectrum at energy higher than 13 MeV). These two enhancement regions are mixed in the 40 Ca+48Ca/40Ca+ Ca ratio. Fromfig.2,by comparing the 48Ca target data with both 46 Ti and '"Ca targets data an extra-yield is observed around 10 MeV. The extrayield is thus related to the (N-Z)/A asymmetry of projectile and target It is also similar to what has been observed at lower beam energy in ICF and deep inelastic reactions[3-5]. BNV calculations qualitatively reproduce this behavior[8].
307
Figure 2 Spectra ratios for the different reactions. Full dots are the experimental y-ray spectra ratios for the reactions indicated. (N/Z) values are also indicated in the figure.
3
Statistical model calculations and discussion
We have preformed statistical model calculations to evaluate the GDR gamma decay. The characteristics of the sources used in the calculations, extrapolated from ref.[9], are reported in table 1. We used a modified CASCADE isospin dependent version.[10]. In previous analysis at very high excitation energy[ll] the data were reproduced Table 1. Summary of the estimated properties of the statistical y-ray sources.
ICF
Average excitation Energy(MeV)
Si+zbSi Si+J4Si ^Si+J4P
312 354 335
Reaction 4U
Ca+ 4 U Ca "Ca+"Ca 4U Ca+"Ti
2a
2B
Spin 26 29 28
h
NZ/A 14 15.35 15.43
assuming a cut-off in the GDR y-ray decay and a saturation of the GDR width. In our case a betterreproductionof the spectra is obtained by assuming a large GDR width, 26 MeV. This result, shown in fig.3, is obtained by adding to the standard CASCADE decay channels the deuteron emission, quite important at this high excitation energy. A strength of 1W.U. can be extracted from the fit without any GDR cut-off. A similar result is obtained also for the ^Ti target ( strength of 0.8 W.U.). However the result obtained for the "°Ca target is quite different A strength of only 0.2 W.U. is obtained. In this case a large GDR cut-off should be used. The
308
fact that this cut-off is not necessary for the other targets suggests that other effects could be responsible of this behavior. ^™ •
"Co+*Co
30
1
_,....--"t 20
u-
a, 1_
•
^
+• 0 ..
I...
•
*N^ref-Z3
O A-J-S3
10
10O
r*f-20
*
"S+^AI fisf.24'
|
*Ca+**Ca this work
200 C (MeV)
300
figure 3 a) Data reproductions relative to the ' Ca+ Ca reaction b) Measured values of the GDR width for systems around mass 60
The width used is not too large when compared to the available data at lower excitation energy. As shown infig.3bthe GDR width (T) increase is steeper at low excitation energy where also the spin increase must be considered, whereas at excitation energy higher than 100 MeV, probably due to the spin saturation, the increase is slower. References 1. S.Flibotte et al; Phys.Rev.Lett. 77(1996). 2. M.Cinausero et al; Nuovo Cimento 111(1998)613. 3. L.Campajola et al; Z.Phys.A352(1995)421.; MSandoli et al Eur.physJ. A6(1999)275. 4. F.Amorini et al; Phys.Rev.C 58(1998)987. 5. M.Papa et al; Eur.Phys.Journ A4(1999)69. 6. AJVIusumarra et al; NIM A370(1996)558. 7. CPaush et al; NTM A365(1995)176 andreferencestherein. 8. M.Papa et al; Int. Work, on Nuclear Reaction and Beyond August 24-27, 1999-Lanzhou, China to be published. 9. T.M.V. Bootsma et al; Z. Phys. A359( 1997)391. 10. RPuhlhofer; NucLPhys. A280 (1977),267; M.N.Harakeh private communication. 11. T.Suomijarvi et al; Phys.Rev.C 53(1996)2258.
T H E O N S E T OF MID-VELOCITY EMISSIONS IN S Y M M E T R I C HEAVY ION R E A C T I O N S E. PLAGNOL", J. LUKASIK al> ' c and the INDRA COLLABORATION* Institut de Physique Nucleaire, IN2P3-CNRS, F-91406 Orsay Cedex,Prance. b GSI, Planckstrasse 1, D-64&91 Darmstadt, Germany. c Institute of Nuclear Physics, ul. Radzikowskiego 152, 31-342 Krakow, Poland. a
Experimental data obtained with the 4-n multi-detector system INDRA1 are used to study mid-velocity emissions for peripheral and semi-central collisions of Xe and Sn at energies between 25 and 50 MeV/nucleon. The analysis is performed as a function of incident energy and of impact parameter. The onset of mid-velocity emissions is found to be close to 25 MeV/nucleon. Evaporative processes are also identified and are found to be sensitive to the impact parameter but show, for a given impact parameter, little dependence on the incident energy. Results are compared with the predictions of a hybrid CHIMERA+GEMINI model.
1
Introduction
It is only quite recently that experimental data 2 _ 1 0 have given indications of the transition from low energy regime (E < 20 MeV/nucleon) to the relativistic collisions (E > 300 MeV/nucleon) of heavy ions. This transition can be investigated by looking at the reaction products with parallel velocities intermediate between those of the projectile and of the target. These emissions appear to be strongly influenced by dynamical effects and are thought to proceed on a relatively short time scale. They are called mid-velocity (-rapidity) emissions, dynamical emissions, or intermediate velocity products. An important source of particle and fragment production remains of course the statistical evaporative process from the excited quasi-projectile and quasi-target. For larger impact parameters their life-times are long enough that these processes take place once the different sources are well separated. The mid-velocity emissions can therefore be defined by opposition to these processes. They include a variety of mechanisms: fast pre-equilibrium particles, neck emitted particles and fragments, as well as light fission fragments preferentially aligned in between the two main reaction partners. Due to limited space, we present here only the main points of the study of mid-velocity emissions in the Xe+Sn reaction, and refer the reader to more comprehensive presentations 8,10 . *See contribution of M.F. Rivet for the full list of collaborators.
309
310
2
Experimental observations
The analysis is performed as a function of incident energy and of impact parameter (b) defined through the total transverse energy of light charged particles - Etransl2. This allows an instructive comparison of the data and hence of different production mechanisms at different incident energies for equivalent impact parameter ranges. The mid-velocity component is found using the following subtraction method: knowing the projectile-like fragment (PLF) velocity and assuming forward-backward symmetric statistical emission pattern in this PLF frame, the mid-velocity component is extracted by doubling the forward yield with respect to the source, and subtracting it from the total yield in the forward hemisphere in the frame of the principal axis of the momentum tensor (see 8 for more details). It should be stressed, that this simple subtraction method does not ensure a sharp separation between dynamical and "pure statistical" emissions, since Coulomb and proximity effects may induce some forward-backward asymmetry in the statistical emission pattern in the emitting source frame u . In the following, it is assumed that the most probable velocity of the heaviest fragments detected in a sample of events for a given impact parameter bin is a reasonable estimate for the mean source velocity of that sample. 1
'
.
' a)
'
i
•
'
.
•
i
'
'
•
•
i
.-•
20
.,;r-J''
10
'
*
'
'
MID-VELOCITY
30
©
» o
'' Eneigy (MeV/DUcIeou)
40
b)
EVAPORATION
o ® '. ©
a 20
s, 10
o
Figure 1. Percentage of charge emitted in the forward hemi-ellipsoid of momentum tensor as a function of incident energy for the four most peripheral E t r a n s l 2 bins (bin 1 being the most peripheral) for: a) midvelocity emissions and b) evaporation processes.
30 40 50 Eneigy (MeV/nucleon)
Figure 1 presents the size of the mid-velocity and evaporative components for all incident energies in percent of the mean total detected charge in the forward hemisphere for the 4 most peripheral Etransl2 bins (b > 4 fm). Two important conclusions can be drawn from this figure: First, panel a) shows that the mid-velocity component evolves from very
311
small values at 25 MeV/nucleon incident energy (around 12% of the total charge) to significant proportions at 50 MeV/nucleon (up to 30%). The onset of mid-velocity emissions is therefore clearly observed in this energy range. Secondly, a surprising observation can be made concerning the evolution of the evaporative component (panel b)): For a given impact parameter, this component is observed to be insensitive to the incident energy. Coupled to the previous observation of the onset of mid-velocity emissions, the implication of this feature on the excitation energy of the evaporative source has to be studied in more detail. 3
Model predictions
Emission of intermediate velocity products as well as statistical emissions are also studied with the molecular dynamics model CHIMERA 12 coupled to the statistical sequential decay code GEMINI 13 used as an afterburner.
50
50
b)
100 150 200 Energy (MeV/nucleon)
250
EVAPORATION
40
1 30-•-.— 10
<
®
i
i
i
100 150 200 Energy (MeV/nucleon)
i
Figure 2. Model equivalent of Fig. 1 but for extended energy range (25-250 MeV/nucleon Xe+Sn). a) percentage of charge in mid-velocity component as a function of incident energy for the four most peripheral impact parameter ranges, b) same as a) but for statistical emission. White circles with numbers denote impact parameter bins corresponding to those from Fig. 1.
250
Figure 2 presents the results of the model, for the soft EOS parameters, using the same impact parameter bins and analysis method as for the experimental data in Fig. 1, but for extended energy range (25-250 MeV/nucleon). The qualitative agreement up to 50 MeV/nucleon is clear and in particular the onset at low energies. The quantitative comparison shows that mid-velocity emissions are however underestimated by about 30-40%. Another interesting observation follows from the panel b) of Fig. 2. The model reproduces the experimentally observed constancy of the evaporated charge as a function of incident energy, at least for the 2 most peripheral impact parameter bins, even above 50 MeV/nucleon. This observation might at first glance suggest the invariance of the excitation energy per nucleon of
312
the primary source, on the incident energy for a given impact parameter bin. In fact, the model predicts such an invariance. The model predicts saturation of the mid-velocity component between 50 and 150 MeV/nucleon depending on the impact parameter. Does it indicate that the "pre-participant-spectator" scenario could be applied at these surprisingly low energies? This prediction should be verified soon with the use of the recent INDRA@GSI campaign experimental data. Nevertheless, it should be stressed that the origin of the mid-velocity component is far more complex than that of the simple fireball model. In summary, the analysis of the Xe+Sn reaction between 25 and 50 MeV/nucleon performed as a function of impact parameter shows the following features : First, the mid-velocity component has an onset around 25 MeV/nucleon. Secondly, investigation of the evaporative component shows that the amount of evaporated charge for a given impact parameter range almost does not depend on the incident energy, for peripheral and mid-central collisions. Thus, mid-velocity emission constitutes the main difference in emission pattern for the incident energies studied. The CHIMERA+GEMINI calculations predict saturation of the midvelocity component at surprisingly low energies and confirm the invariance of the evaporative component on the incident energy for the same impact parameters. References 1. 2. 3. 4. 5. 6. 7. 8. 9. 10. 11. 12.
J. Pouthas et al., Nucl. Instr. Meth. in Phys. Res. A 357, 418 (1995) L. Stuttge et al., Nucl. Phys. A539, 511 (1992). G. Casini et al., Phys. Rev. Lett. 7 1 , 2567 (1993). C.P. Montoya, et al., Phys. Rev. Lett. 73, 3070 (1994). J.F. Lecolley et al., Phys. Lett. B 354, 202 (1995). J.Peter et al., Nucl. Phys. A593, 95 (1995). J.F. Dempsey et al., Phys. Rev. C 5 4 , 1710 (1996). J. Lukasik et al., Phys. Rev. C 5 5 , 1906 (1997). P. Pawlowski et al., Phys. Rev. C57, 1771 (1998). E. Plagnol et al., Phys. Rev. C 61, 014606 (2000). A.S. Botvina, contribution to these Proceedings. J. Lukasik and Z. Majka, Acta Physica Polonica B 24, 1959 (1993) J. Lukasik, PHD thesis, Krakow 1993, unpublished. 13. R.J. Charity et al., Nucl. Phys. A483, 371 (1988), computer code GEMINI obtained from WUNMR.WUSTL.EDU via anonymous FTP.
CONTRIBUTION OF P R O M P T EMISSIONS TO T H E P R O D U C T I O N OF I N T E R M E D I A T E VELOCITY LIGHT P A R T I C L E S I N T H E 3 6 A r + 5 8 N i R E A C T I O N A T 95 MeV/NUCLEON
P. P A W L O W S K F . D. B O R D E R I E , A N D T H E I N D R A C O L L A B O R A T I O N ' ' Institut de Physique Nucleaire, IN2P3-CNRS, F-91406 Orsay Cedex, France. The prompt component at intermediate velocity of light charged particles is investigated. An improved coalescence model coupled to the intra-nuclear cascade code ISABEL is used to obtain light complex particle energy spectra and multiplicities as a function of impact parameter. The results are compared with experimental data from the 3 6 Ar + 5 S Ni experiment at 95 MeV/nucleon, performed with the INDRA detector. The calculated prompt component is found to rather well reproduce proton spectra. For complex light charged particles the calculated components well populate the high energy part of spectra. Prompt emission can therefore explain the large transverse energies experimentally observed at mid-rapidity.
T h e precise characterization of t h e hot nuclei produced in heavy ion reactions at intermediate energies calls for a detailed comprehension of t h e mechanisms since dynamical or out of equilibrium effects are known t o play a n i m p o r t a n t role in this energy regime. For energies below 10-15 MeV/nucleon dissipativc collisions arc dominated by mean field effects giving rise t o complete fusion or binary deep inelastic collisions (DIC), whereas for much higher energies, above 200 MeV/nucleon, nuclcon-nuclcon (N-N) collisions take over a n d t h e reactions can b e interpreted in a participant-spectator framework. T h e situation in t h e intermediate energy domain is more complicated. This region exhibits a transitional regime where two-body collisions come into play a n d compete more and more with the m e a n field as t h e incident energy increases. T h e i m p o r t a n c e of two-body collisions is particularly revealed by some direct experimental facts such as prc-cquilibrium e m i t t e d particles a n d high energy 7r a y s 1 , 2 , 3 . But it also appears, in an indirect way, t h r o u g h dynamical properties of collisions showing deviations from the pure binary pictures (DIC) observed at lower incident e n e r g y 4 , 5 , 6 . In this work we a t t e m p t to bring some information a b o u t t h e role of twob o d y collisions in p o p u l a t i n g t h e mid-rapidity region (MR) in t h e 3 6 Ar + 5 8 Ni reaction a t 95 MeV/nucleon. Our goal is t o interpret t h e transverse energy rise for light charged particles ( L C P ) observed experimentally at m i d - r a p i d i t y 5 , 6 . "E-mail: [email protected] 'See contribution of M.F. Rivet for full list of collaborators 313
314
To obtain t h e direct emission component we use t h e intra-nuclcar cascade code I S A B E L 7 , producing nucleoli densities in m o m e n t u m space. These densities arc t h e n used by a coalescence model to construct complex light particle energy spectra. In comparison with t h e s t a n d a r d version of t h e coalescence model. we have developed a more precise formalism, which can b e properly applied in t h e incident energy domain discussed here. As coalescence radius po we have taken two extreme limits drawn from previous experimental measurements 8 ' 0 . For details of this calculation t h e reader can refer t o 1 0 . Figure 1 presents energy distributions obtained in t h e angular region between 60° and 120° in t h e center of mass for p . d, t. 3 H c . and 4 H c in four i m p a c t p a r a m e t e r windows. For dcuterons, which arc t h e most produced coalescence particles, a good agreement can be observed, except in t h e region of smaller energies, where a contribution of evaporative components from Q P and Q T , and possible neck-like emissions arc expected. Increasing t h e mass of complex particles, contributions from p r o m p t emissions decrease a n d only p o p u l a t e correctly t h e high tails of energy spectra. These results clearly reveal t h e importance of p r o m p t processes for populating t h e high energy p a r t s of complex particle spectra. T h e p r o t o n spectra arc not so well reproduced, because of an excess of protons in the lower energy region and slopes somewhat too steep at large impact p a r a m e t e r . This last effect reflects in t h e small disagreement, observed a t large energies, for complex particles in t h e range of peripheral collisions. However, without any normalization the agreement is remarkable, and we can infer t h a t t h e major p a r t of protons in M R comes from p r o m p t processes. We studied also m e a n multiplicities obtained by integrating spectra presented in Fig. 1. We observe here an ovcrestimation of proton multiplicities in calculation, particularly for t h e lower impact p a r a m e t e r . T h e proportion of p r o m p t complex particles decreases when t h e particle mass increases b u t increases with t h e violence of collisions. Multiplicities of dcuterons, tritons a n d 3 H c seem to be reasonable, if one remembers t h a t contributions of evaporative component and possible neck-like emissions are present in experimental d a t a . Similar results arc obtained by using t h e INC model of Cugnon, followed by a percolation procedure and evaporation code of D r e s n e r 1 1 . For alpha particles the proportion of calculated p r o m p t emission is quasinegligible. T h e number of p r o m p t 4 H c can b e underestimated, as it can also originate from prc-formed alpha structures in nuclei 1 ' 1 2 . It should b e remarked t h a t t h e excess of protons corresponds approximately t o t h e deficit of alphas, multiplied by two. It suggests t h a t t h e number of cascade protons produced by t h e ISABEL code is quite correct, a n d t h a t t h e calculation is not able t o create enough alpha particles.
315
-
5
N. ,
-..Y
\
\
,
\ 100
„
200
100
•
.
.
i i i i i i i4ll
\
100
6
J
100
.
200
200
KW
200
E (AMeV)
Deuterons
0
x
x
^V
\ •
o Experiment INC + Coalescence
6
7
0
!0
60'^^-a:
Protons
200
100
1
I__l
I
5
I
I
r
iVt
200
I
100
i_i
|
|__1
4
I
I
LJBJLJ.
200
100
200
200
100
200
E(AMeV) Tntons
100
0
200
100
^e r
200
100
Jri(Ai vievj
7
33
\ \\. . . i . . . , 100
6
1 K
200
100
200
.\V, , l . , , .
5 < b< 6 S
100
\
4 < b< 5
200• • » - 100
200
E(AMeV)
Figure 1: LCP differential multiplicities dM/dE. Comparison of simulated energy spectra for p, d, t, 3 He, and 4 He (double solid lines, corresponding to two extreme limits for po) with experimental ones (circles). Impact parameter windows are determined with the use of light charged particle multiplicity.
316 In summary, this calculation is found t o describe r a t h e r well proton energy spectra and well t h e high energy p a r t of spectra for emitted light complex particles like deutcron. triton and hclium-3. For alpha particle energy spectra. only t h e very high energy tail is well reproduced. Thus production of coalcsccnt p r o m p t particles with r a t h e r high energies can b e found responsible for large transverse energies experimentally observed in the intermediate velocity region at incident energies around 100 McV/nuclcon. References 1. H. Fuchs a n d K. Mohring, Rep. Prog. Phys. 5 7 , 231 (1994). 2. H. Nifcncckcr and J. A. Pinston, Ann. Rev. Nucl. Part. Sc. 4 0 , 113 (1990). 3. P. Sapicnza et al., Nucl. Phys. A 6 3 0 , 215c (1998). 4. L. S t u t t g c ct al.. Nucl. Phys. A 5 3 9 , 511 (1992); J. F . Lccollcy ct al., Phys. Lett. B 3 5 4 , 202 (1995); J. Tokc ct al., Nucl. Phys. A 5 8 3 , 519 (1995); J. Tokc ct al., Phys. Rev. Lett. 7 5 , 2920 (1995); J. P e t e r ct al., Nucl. Phys. A 5 9 3 , 95 (1995); J. F. Dcmpscy ct al., Phys. Rev. C 5 4 , 1710 (1996); S. L. Chen ct al., Phys. Rev. C 5 4 , R2214 (1996); J. Lukasik ct al., Phys. Rev. C 5 5 , 1906 (1997); E. Plagnol ct al., Phys. Rev. C61, 014606 (1999); Y. Larochcllc ct al., Phys. Rev. C 5 5 , 1869 (1997); P. Pawlowski ct al., Phys. Rev. C 5 7 , 1771 (1998); F . Bocagc ct al., Nucl. Phys. A . (in press); 5. J. C. Angcliquc ct al., Nucl. Phys. A 6 1 4 , 261 (1997); 6. T. Lefort ct al., Nucl. Phys. A 6 6 2 , 397 (2000). 7. Y. Yariv a n d Z. Fracnkcl, Phys. Rev. C 2 0 , 2227 (1979); Phys. Rev. C 24, 488 (1981). 8. H. H. G u t b r o d at al., Phys. Rev. Lett. 3 7 , 667 (1976). 9. R. L. Aublc a t al., Phys. Rev. C 28, 1552 (1983). 10. P. Pawlowski ct al., s u b m i t t e d to Phys. Rev. C; sec also web site: http://www.cyf-kr.edu.pl/~uupawlow/coalescence.ps.gz 11. D. Dorc ct al., P r o c . of Varcnna 2000 Conference on Nuclear Reaction Mechanisms, to b e published in '"Ricerca Scicntifica cd Educazionc Pcrmancntc". 12. T. Ncff at al., P r o c . of t h e X X V I I - t h International Workshop on Gross Properties of Nuclei and Nuclear Excitations, Hirschcgg, Austria, edited by H. Fcldmcier, J. Knoll, W. NSrcnbcrg a n d J. Wambach, p . 283 (1999).
P R O T O N EMISSION TIMES IN SPECTATOR FRAGMENTATION C. S C H W A R Z F O R T H E A L A D I N C O L L A B O R A T I O N Gesellschaft
fur Schwerionenforschung,
D-64291
Darmstadt,
Germany
Proton-proton correlations from spectator decays following 1 9 7 A u + 1 9 7 A u collisions at 1000 AMeV have been measured with an highly efficient detector hodoscope. The constructed correlation functions indicate a moderate expansion and low breakup densities similar to assumptions made in statistical multifragmentation models. In agreement with a volume breakup rather short time scales were deduced employing directional cuts in proton-proton correlations.
1
Introduction
Densities lower than ground state density of nuclei are a prerequisite of statistical models describing multifragmentation li2. While static statistical models assume fragment formation in an expanded volume breakup, the dynamic statistical model 3,4 combines surface emission during expansion with volume breakup of the remaining source. Interferometry-type methods are widely considered as valuable tools in determining the space-time extent of such sources and, recently, spectator remnants of the reaction Au 4- Au at 1 AGeV incident energy were found to break up at densities considerably lower than ground state density 5 . In that analysis a instantaneous volume breakup was assumed. In this article we report on results of a directional analysis of the measured proton-proton correlations of protons from the target spectator following the collisions of Au + Au at 1 AGeV incident energy. The results are found to be consistent with low breakup densities with values close to those assumed in the statistical multifragmentation models and short emission time differences with values close to those anticipated for volume breakup. 2
Experiment
Targets, consisting of 25 mg/cm 2 of 197 Au were irradiated by an 1 AGeV Au beam delivered by the Schwerionen-Synchrotron (SIS) at GSI in Darmstadt. For the results presented here, we employed one multi-detector hodoscope, consisting of a total of 96 Si-CsI(Tl) telescopes in closely packed geometry. The hodoscope covered an angular range @iab from 122° to 156° with the aim of selectively detecting the products of the target-spectator decay. Each 317
318
telescope consisted of a 300 fim Si detector with 30 x 30 mm 2 active area, followed by a 6 cm long CsI(Tl) scintillator with photodiode readout. The distance to the target was 60 cm. The products of the projectile decay were measured with the time-of-flight wall of the ALADIN spectrometer 6 and the quantity ZBOUND was determined event-by-event. ZBOUND is denned as the sum of the atomic numbers Zi of all fragments with Zi> 2. It reflects the variation of the charge of the primary spectator system and serves as a measure of the impact parameter. 3
Data analysis
The correlation functions were constructed dividing the spectrum of relative momenta of two coincident particles by the spectrum of pairs from different events. At a relative momentum of q « 20 MeV/c, one observes a maximum of the correlation function the height of which is inversely related to the diameter of the source for simultaneous emission 7 ' 8 . Directional cuts on the angle between sum momentum and relative momentum allow the determination of the spatial and temporal separation of the two protons at emission 7 ' 9 . Instead of using sharp cuts on the angle between sum momentum and relative momentum we employed harmonic weights (sin2 and cos2) for the generation of transversal and longitudinal correlation functions 10 . The analysis of the
HnKkftfmfc)
timetfm/c)
tims(fm/c)
tinie(fm/c)
Figure 1. %2 distributions as a function of radius and emission time of the emitting source for four impact parameter ranges indicated by the ZBOUND ranges in the panels. The lines in the left panel are explained in the text. p-p correlation functions was performed with the Koonin-Pratt formalism 7 , s . Particles were chosen to be randomly emitted from the volume of an uniform sphere and their velocities were sampled according to a Maxwell-Boltzmann distribution. An additional velocity component was added in order to simulate the Coulomb repulsion corresponding to the location of the particle within the
319
source. An experimentally observed anisotropy in the proton-energy spectra was modeled by assuming a sideward movement (bounce) of the source perpendicular to the beam. This causes a reaction plane which was included in the Monte-Carlo calculations. 4
Results and Discussions
We varied the radius of an uniform density distribution and the Gaussian emission time of the protons. The simulated correlation functions were used to perform a x 2 -test in the region of relative momentum region 10 < q < 35 MeV/c. The results are presented in Fig. 1, where the shadings denote the deviation from the data. The minima of %2 yield approximately constant source radii of w 8 fm and short emission times of « 10 — 15 fm/c. For the most peripheral bin a radius and emission time could not be deduced due to the low statistics of the correlation function. The experimental correlation functions (symbols) and the simulations (lines) for the parameters corresponding to the minimum of x 2 a r e shown in Fig. 2, left panel. They agree well
Figure 2. The left panel shows experimental longitudinal (open symbols) and transversal (closed symbols) correlation functions and results of MC-simulations (lines). The right panel compares the their experimental ratios (symbols) with the results of the MC-simulation (lines).
with each other. The minimum x 2 values per degree of freedom are within the range of 1.3 < x 2 /d.o./. < 2.4. The right panel in Fig. 2 presents the ratios between longitudinal and transversal correlation functions. One recognizes the weak enhancement of about 5% above unity (dashed line) of data (symbols) and simulations (solid line) for relative momenta q < 50 MeV/c. For a source size of « 8 fm the quantum suppression is expected at q = v/q2" < \/3 * h/r « 40 MeV/c.
320
5
Conclusions
We constructed correlation functions for pairs of protons detected at backward angles in the reaction Au+Au at 1000 AMeV incident beam energy. Using high energy cuts of E > 20 MeV we selected protons which are only little affected by sequential feeding n . Comparing the results with Monte-Carlo simulations within the Koonin-Pratt formalism fairly constant freeze-out radii of R « 8 fm are deduced and emission times of r = 10 — 15 fm/c are surprisingly short of the order of the passing time of the projectile through the target. The extracted radii are larger than the ground state radii of target spectators and show expansion. Because of the short emission times in the order of the passing time of both spectators we cannot exclude that the protons come from first stage scattering of the nuclear cascade. Acknowledgments This work was supported by the Deutsche Forschungsgemeinschaft under contract Schw510/2-l and by the European Community under contract ERBFMGECT-950083. References 1. J. P. Bondorf, A. S. Botvina, A. S. Iljinov, I. N. Mishustin, and K. Sneppen. Phys. Rep., 257:133, 1995. 2. D. H. E. Gross. Phys. Rep., 279:119, 1997. 3. W. A. Friedman and W. G. Lynch. Phys. Rev. C, 28(1):16, 1983. 4. W. A. Friedman. Phys. Rev., C42:667, 1990. 5. S. Fritz et al., Phys. Lett. B, 461:315, 1999. 6. A. Schuttauf et al., Nuc. Phys, A607:457, 1996. 7. S. E. Koonin. Phys. Lett. B, 70(1):43, 1977. 8. S. Pratt and M.B. Tsang. Phys. Rev., C36:2390, 1987. 9. M. A. Lisaet al., . Phys. Rev. Lett, 71(18):2863,1993. 10. C. Schwarz et al., Proceedings of CMS '2000, 3rd Catania Relativistic Ion Studies, Acicastello, Italy, May 22-26, 2000, to be published in Nucl. Phys. A. 11. C. Schwarz and S. Fritz. In H. Feldmeier, J. Knoll, W. Norenberg, and J. Wambach, editors, Proceedings of the International Workshop XXVII on Gross Properties of Nuclei and Nuclear Excitations, page 168, Hirschegg, Austria, January, 17-23 1999. Gesellschaft fur Schwerionenforschung mbH.
EXPERIMENTAL EVIDENCE FOR SPINODAL D E C O M P O S I T I O N IN M U L T I F R A G M E N T A T I O N OF H E A V Y SYSTEMS G. T A B A C A R U 1 ' 2 , B . B O R D E R I E 1 , M . F . R I V E T 1 , P. C H O M A Z 3 , M . C 0 L 0 N N A 4 , J . D . F R A N K L A N D 3 , A. G U A R N E R A 4 , M. P A R L O G 2 , B . B O U R I Q U E T 3 , A. C H B I H I 3 , S. S A L O U 3 , J . P . W I E L E C Z K O 3 , G. A U G E R 3 , C H . O . B A C R I 1 , N. B E L L A I Z E 5 , R. B O U G A U L T 5 , R. B R O U 5 , P. B U C H E T 6 , J.L. C H A R V E T 6 , J. C O L I N 5 ,D. C U S S O L 5 , R. D A Y R A S 6 , A. D E M E Y E R 7 , D . D O R E 6 , D. D U R A N D 5 , E. G A L I C H E T 1 ' 8 , E . G E R L I C 7 , B . G U I O T 3 , S. H U D A N 3 , D . G U I N E T 7 , P. L A U T E S S E 7 , F . L A V A U D 1 , J.L. L A V I L L E 3 , J . F . L E C O L L E Y 5 , C. L E D U C 7 , R. L E G R A I N 6 , N. L E N E I N D R E 5 , Q. L O P E Z 5 , M. L O U V E L 5 , J. L U K A S I K 1 , A . M . M A S K A Y 7 , L. N A L P A S 6 , J . N O R M A N D 5 , P. P A W L O W S K I 1 , E . P L A G N O L 1 , E . R O S A T O 9 , F . S A I N T - L A U R E N T 3 ! J . C . S T E C K M E Y E R 5 , B . T A M A I N 5 , L. T A S S A N - G O T 1 , E . V I E N T 5 , C. V O L A N T 6 (INDRA COLLABORATION) 1 Institut de Physique Nucleaire, IN2PS-CNRS, F-91406 Orsay Cedex, France. 2 National Institute for Physics and Nuclear Engineering, RO-76900 Bucharest- Magurele, Romania. 3 GANIL, CEA et IN2PS-CNRS, B.P. 5027, F-14076 Caen Cedex, France. 4 Laboratorio Nazionale del Sud, Viale Andrea Doria, 1-95129 Catania, Italy. 5 IPC, IN2P3-CNRS, ISMRA et Universiti, F-14050 Caen Cedex, France. 6 DAPNIA/SPhN, CEA/Saclay, F-91191 Gif sur Yvette Cedex, France. 7 Institut de Physique Nucleaire, IN2P3-CNRS et Universiti, F-69622 Villeurbanne Cedex, France. 8 Conservatoire National des Arts et Metiers, F-75H1 Paris Cedex OS. 9 Dipartimento di Scienze Fisiche e Sezione INFN, Universiti Napoli 'Federico II", 1-80126 Napoli, Italy.
Multifragmentation of fused systems was observed for central collisions between 32 AMeV 1 2 9 Xe and Sn, and 36 AMeV 1 5 5 G d and U. Previous extensive comparisons between the two systems led to the hypothesis of spinodal decomposition of finite systems as the origin of multifragmentation for incident energies around 30 AMeV. New results on velocity and charge correlations of fragments bring strong arguments in favor of this interpretation.
•present address: cedex, France.
DR.FC/STEP, CEA/Cadarache, F-13018 Saint-Paul-lez-Durance
321
322
1
Introduction
The decay of highly excited nuclear systems through multifragmentation is currently a subject of great interest in nucleus-nucleus collisions. The recent advent of powerful Ait devices brought high quality experimental data which allow a careful selection of well defined fused systems undergoing multifragmentation. We present in this contribution new results on the multifragmentation of heavy fused systems, formed in central 32 AMeV 129 Xe+™ a 'Sn, and 36 AMeV 1 5 5 Gd+U collisions. Details on the experiment, performed at GANIL with the INDRA 4ir array, and on event selection can be found in ref1-2-3.
Zbound
Charge Z
Figure 1. Comparison of experimental data with SMM (a-d) and BOB (e-g) simulations. a-d concerns the 32 AMeV Xe+Sn system, the lines are for data and the symbols for SMM (all fragments, except open circles and dashed lines which refer to the largest fragment of each partition). Z(, oun( j represents the sum of the charges of all fragments. In e-g the symbols are for data and the lines for BOB simulations. Light grey lines and triangles stand for G d + U and black lines and circles for Xe+Sn.
First evidence for a bulk effect was found in the comparison of the two multifragmenting systems, which have the same available energy ~ 7 AMeV: the charge distributions are identical while the number of fragments scales as the charge of the systems 3 ' 2 . This scaling could simply be the sign of the breaking of statistically equilibrated systems, but could also be envisaged as the occurrence of spinodal decomposition of very heavy composite systems (A>250). Indeed if systems break in the spinodal region of the (T,p) plane ,
323
a "primitive" break-up into fragments with a favoured size is predicted 4 , in connexion with the wave lengths of the most unstable modes in nuclear matter. The enhancement of equal-sized fragment partitions is however washed out by several effects (beating of different modes, coalescence of nuclear-interacting fragments, finite size of the systems, and secondary decays). Stochastic meanfield simulations (BOB) 5 which describe the entire collision process up to the final fragment de-excitation, well account for fragment multiplicity and charge distributions as well as for the average fragment kinetic energies (fig. leg). The SMM statistical model 6 however also well reproduces the same distributions, provided that the mass, charge and excitation energy of the system are traced back to the "freeze-out" time, where fragments cease to feel the nuclear interaction 7 ' 8 (fig. la-d). Therefore the variables considered above do not furnish a stringent enough test of a spinodal decomposition, and one has to turn to correlation studies. In the following sections the samples of experimental and simulated (BOB, SMM) events are exactly the same as those from which fig la-g were built. 2
Velocity correlations
o
10
40
SO
60
70
1000xf,rel/(Z,+Z2)"2
80
Figure 2. Correlation functions for the reduced fragment relative velocities. The fragment charges are in the range 5-20. Experimental data (points) are compared with the results of dynamical simulations and of the statistical model SMM, for the G d + U (top) and Xe+Sn (bottom) multifragmenting events.
Reduced fragment-fragment velocity correlations were built, mixing pairs of fragments with different charges to increase statistics. The same procedure was used for experimental data and for simulated events. These correlation
324
functions should reflect the topology of the events at freeze-out, and especially the strength of the Coulomb repulsion. Experimental and calculated (BOB and SMM) velocity correlation functions are shown in fig 2 for the two systems under study. In this figure the fragment are limited to the range Z=5-20, but similar pictures are obtained when the heavier fragments are included (for both systems < Zmax > ~ 25). In both cases the BOB simulations perfectly match the experimental data, indicating that the configuration at freeze-out is correctly described. As this topology is not of bubble-type,it rules out surface instabilities as being the cause of multifragmentation. Conversely, the SMM calculations, which in this version simply randomly place the fragment in a volume equal to 3 times the normal nuclear volume, do not reproduce the velocity correlations, in particular the Coulomb depletion is too broad, and a bump appears just after. In ref 9 such a bump marks the presence of one very big fragment in the partitions. It is not the case here as it was shown that the details of the partitions are well reproduced by this SMM calculation. 3
C h a r g e correlations
Data
BoB
Xe+Sn 32 MeV/A
A compact way of describing event partitions is to build higher order charge correlation functions, denned as the ratio of yields Y(
325
correlated and decorrelated fragments; < Z > is the average fragment charge and AZ the s t a n d a r d deviation per event 1 0 . In such a representation, any privileged partition (down to the level of 0.1%) will a p p e a r as a peak at the corresponding < Z > , A Z bin. Charge correlation functions built for experimental a n d B O B simulated events are shown in fig. 3; only for t h e X e + S n system was the statistics high enough for such a study. For all fragment multiplicities (we show here only M = 4 a n d 5) the experimental charge correlation has a peak in the bin AZ = 0-1, indicating an enhancement of partitions with fragments of equal size. This peak corresponds to a constant value of the p r o d u c t M x < Z > . Its existence is specific of the fused events selected, it does not a p p e a r for other very dissipative types of collisions 1 1 . T h e number of events in t h e peaks at AZ = 0-1 s u m m e d over all multiplicities corresponds to 0 . 1 % of the fused events. T h e same features are observed in events simulated with B O B (fig. 3 right). Although in B O B we know t h a t all events multifragment through spinodal decomposition, only 0 . 1 % of t h e m survive t h e blurring of the favored initial fragment size. T h u s we s t a t e t h a t all the experimentally selected fused systems also underwent spinodal decomposition. Finally the events simulated with the version of SMM considered here do not present any partition with equal-sized fragments n .
4
Conclusions
New detailed analyses of fragment velocity and charge correlation functions for very heavy multifragmenting systems strongly support the previous ass u m p t i o n t h a t spinodal decomposition is a t t h e origin of multifragmentation, for very heavy systems at incident energies a r o u n d 30-35 AMeV. References 1. 2. 3. 4. 5. 6. 7. 8. 9. 10. 11.
J. P o u t h a s et a l , Nucl. Instrum. Methods A 357, 418 (1995). J.D. Frankland et al., (INDRA coll.) s u b m i t t e d to Nucl. Phys. A. M . F . Rivet et al. (INDRA coll.), Phys. Lett. B 4 3 0 , 217 (1998). A. G u a r n e r a et al., Phys. Lett. B 3 7 3 , 267 (1996), A. G u a r n e r a et a l , Phys. Lett. B 4 0 3 , 191 (1997). J. Bondorf et al., Phys. Reports 257, 133 (1995) N. Le Neindre, these, Universite de Caen, L P C C T 99 02. R. Bougault et al., (INDRA coll.) P r o c . Bormio (2000), p. 404. O. Schapiro and D.H.E. Gross, Nucl. Phys. A 5 7 6 , 428 (1994). L.G. M o r e t t o et al., Phys. Rev. Lett. 77, 2634 (1996). G. T a b a c a r u et al., (INDRA coll.) P r o c . Bormio (2000), p. 433.
N o n - e q u i l i b r i u m effects o n a s e c o n d - o r d e r p h a s e t r a n s i t i o n
V. Latora and A. Rapisarda Dipartimento di Fisica e Astronomia, Universitd di Catania, and INFN sezione di Catania, Corso Italia 57, 1-95129 Catania, Italy E-mail: [email protected], [email protected] We present a model of interacting spins that can be solved analytically in the canonical ensemble and has a second-order phase transition. Dynamical microcanonical simulations indicate that, due to non-equilibrium effects, the model can show a backbending in the caloric curve and a negative specific heat. The potential relevance for nuclear multifragmentation is discussed.
Recent experimental d a t a on multifragmentation reactions 1 ' 2 ' 3 ' 4 have raised an increasing interest in the study of liquid-gas phase transition in finite nuclear s y s t e m s 5 ' 6 ' 7 . From a theoretical point of view, the thermodynamics of a finite system has to be treated in the microcanonical ensemble8. A second important aspect, often neglected, is t h a t a nuclear collision is a dynamical process and the assumption of equilibrium before the multifragmentation process is a strong assumption. In this paper we present to the nuclear physics community the Hamiltonian Mean Field (HMF) model, a simple model of interacting spins allowing a microcanonical and dynamical approach t o explore the dynamics of a phase transitions in a finite system 9 . In particular we focus on the problem of relaxation to equilibrium and on the characteristics of the out-of-equilibrium regime. As a main results we show t h a t , if the system is not perfectly equilibrated, misleading information on the order of a phase transition can be obtained from a specific heat analysis 1 0 . T h e H M F model was introduced by Antoni and Ruffo 1 1 , and since then it has been extensively studied b o t h analitically and n u m e r i c a l l y 1 2 ' 1 3 ' 1 4 . T h e model describes a linear chain of N classical two-dimensional spins m j = (cos8i,sin6i) similarly to the XY model. Each spin i is characterized by the angle 0 < 6i < 2TT and the conjugate m o m e n t u m pi (a rotational velocity), and is fully coupled t o all the others. T h e Hamiltonian is:
H(0,p) = K + V = f ^
+ ^E[l-coS(^-^)]
. (1)
Since the interaction does not depend on the physical distance between two spins, but only on the relative angle (an extended version of the model taking into account such a dependence has also been p r o p o s e d 1 5 ) , the H M F model 327
328
can also be interpreded as a system of N fully-coupled rotators moving on the unitary circle. The model has an analitycal solution in the canonical ensemble 11 . The magnetization M = -^ YL%=i m i i s the order parameter of the phase transition. The system behaves as a ferromagnet at low energy and has a second-order phase transition, with mean field critical exponents, at the critical energy density Uc — 0.75 (U — E/N where E is the total energy), corresponding to a critical temperature Tc — 0.5. The behavior of M and the caloric curve obtained from the analytical solution in the canonical ensemble are reported as full lines in fig.la and fig.lb. At variance with percolation and other statistical models, HMF allows also a microcanonical dynamical approach. In fact the equations of motion: 6i =Pi
,
Pi = -sindiMx
+ cosOiMy ,
i = 1,..., N
(2)
12
can be integrated numerically for different values of N, and the dynamics of the system can be explored by starting from a given set of initial conditions {0i,pi}. The results of the microcanonical numerical simulations reproduce the analytical equilibrium solution (full lines in fig.l), provided that we wait enough time so that the system relaxes to equilibrium 9 (the equivalence of different ensembles in the thermodynamic limit is also supported by recent analytical microcanonical results 1 4 ). Often a very long integration time is necessary, in particular close to the critical point where the presence of nonequilibrium long living states, the so called Quasi-Stationary States (QSS), has been found 12 . In this paper we investigate the effects of non-equilibrium and of the presence of QQS on the caloric curve and on the specific heat. We consider a system o£ N = 500 and different energies U = E/N. We simulate the strong off-equilibrium conditions present in a hot and compressed nuclear system before multifragmentation by starting HMF in a "water bag", i.e. by putting all rotators at g» = 0 and giving an initial velocity according to a constant probability distribution function of finite width centered around zero 1 2 . We follow the simulations for a time t = 3000 (in natural units of the system) and in fig.l we report respectively total magnetization, temperature, kinetic fluctuations and specific heat per particle (filled circles), and we compare to the theoretical predictions (full curves). The temperature is computed from the average kinetic energy per particle T = 2 < K > /N, where the symbol < > stands for time averages. The kinetic energy fluctuations are obtained from S = y/
cv-\
E>2
l-»l
?
(3)
329
u
u
Figure 1: Numerical simulations at time t = 3000 for a system with N = 500 (filled circles), in comparison with the equilibrium analytical predictions. The dashed straight line indicates the critical point.
Average variables such as M and T have almost reached the canonical equilibrium in a large range of energy U, see panel (a) and (b). Due to the presence of the QSS 12 there is still an important difference with respect to the equilibrium in the region 0.5 < U < Uc: the caloric curve T(U) in fig.lb shows a backbending typical of a first-order phase transition, though the system has a second order phase transition in the thermodynamical limit. The fluctuations, on the other side, need longer times to relax to equilibrium and at t = 3000 are still affected by non-equilibrium effects in the whole spectrum of energy. In fig.lc the kinetic energy fluctuations are bigger than the equilibrium ones. As a consequence the specific heat in panel (d) assumes negative values (see filled circles), while the microcanonical specific heat at equilibrium is always positive (full line). The behavior of Cy versus U we have obtained is similar to the one found in nuclear multifragmentation data, see fig.4 of ref.10. In particular the authors of ref.10 claim that such a negative branch in the heat capacity is a direct evidence of a first-order phase transition. Here we have an example of a system with a second-order phase transition that shows, due to non perfect equilibration, a negative specific heat and a backbending in the caloric curve. This dynamical effect can be explained by the presence of superdiffusion and Levy walks in the out-of-equilibrium regime 13 , that implies the coexistence of a
330
liquid (clustered rotators) and a gas (free rotators), and simulates a first-order phase transition 1T . To conclude we need to say that the non-equilibrium features of our model are stricly linked to the long-range nature of the interaction. In particular there is a great similarity between the scenario indicated by our simulations and the conjecture of Tsallis of a different equilibrium for system with longrange forces 18 . However it is not clear how general our results are, and further work in this direction is certainly needed 1T . We thank X. Campi, M. Pettini, J. Richert and C. Tsallis for their comments. This work is part of a long-term collaboration with S. Ruffo. References 1. 2. 3. 4. 5. 6. 7. 8. 9. 10. 11. 12. 13. 14. 15. 16. 17. 18.
Elliott et al., Phys. Rev. C 49, 3185 (1994), and refs. therein. J.Pochodzalla et al., Phys. Rev. Lett. 75, 1040 (1995). Y.-G Ma et al., Phys. Lett. B 390, 41 (1997). M. D'Agostino et al., Nucl. Phys. A 650, 329 (1999). V. Latora, M. Belkacem, A. Bonasera, Phys. Rev. Lett. 73, 1765 (1994), Phys. Rev. C 52, 271 (1995). X. Campi and H. Krivine, Nucl. Phys. A 620, 46 (1997); X. Campi, H. Krivine, N. Sator, Conference Proceedings, CRIS2000. J.M. Carmona et al., Phys. Rev. C 61, 037304 (2000) and refs therein. D.H.E. Gross, Phys. Reports 279, 119 (1997) and refs therein. V. Latora and A. Rapisarda, nucl-th/0007038. M. D'Agostino et al.,Phys. Lett. B 473, 219 (2000). M. Antoni and S. Ruffo, Phys. Rev. E 52, 2361 (1995). V. Latora, A. Rapisarda and S. Ruffo, Phys. Rev. Lett. 80, 692 (1998), and Physica D 131, 38 (1999). V. Latora, A. Rapisarda and S. Ruffo, Phys. Rev. Lett. 83, 2104 (1999). M. Antoni, H. Hinrichsen, and S. Ruffo, to be published in Chaos solitons and fractals, cond-mat/9810048. C. Anteneodo and C. Tsallis, Phys. Rev. Lett. 80, 5313 (1998). J.L. Lebowitz, J.K. Percus and L. Verlet Phys. Rev. A 153 (1967) 250. V. Latora and A. Rapisarda, in preparation C. Tsallis, in "Nonextensive Statistical Mechanics", eds. S.R.A. Salinas and C. Tsallis, Braz. J. Phys. 29, 1 (1999) cond-mat/9903356; V. Latora and A. Rapisarda, to be published in Chaos solitons and fractals, cond-mat/0006112.
MULTIFRAGMENTATION OF E X P A N D I N G MATTER
NUCLEAR
Shinpei Chikazumi 1,2 , Toshiki Maruyama 3 ' 4 , Koji Niita 5 , Akira Iwamoto 2 , Satoshi Chiba 3 Institute of Physics, University of Tsukuba, Tsukuba, Ibaraki 305-0006, Japan1, Department of Materials Science, Japan Atomic Energy Research Institute, Tokai, Ibaraki, 319-1195 Japan2, Advanced Science Research Center, Japan Atomic Energy Research Institute, Tokai, Ibaraki, 319-1195 Japan3, Laboratorio Nazionale del Sud, Istituto Nazionale di Fisica Nucleare, V.S.Sofia 44, Catania, 95123 Italy4, Research Organization for Information Science & Technology, Tokai, Ibaraki, 319-1106 Japan5 We study the properties of dynamically expanding nuclear matter. For this purpose, we apply quantum molecular dynamics (QMD) model to homogeneous expanding three-dimensional system with periodic boundary condition. Simulation is performed for given initial temperature and expanding velocities. The calculated fragment mass distribution for slow expansion obeys the power law predicted by Fisher's droplet model, while that for rapid expansion exhibits the exponential shape.
1
Introduction
Multi-fragmentation has been one of most important subjects in nuclear physics. There are two possible scenarios of its mechanism. One is t h e liquid-gas phase transition where a fire ball made by t h e nuclear collision cools down into liquid fragments as the system expands. In fact, this has been a t t r a c t i n g m a n y authors in connection with the nuclear equation of state and nuclear phase transition. T h e other is t h a t the crack of nuclear m a t t e r gives rise to t h e fragments. Our purpose is to discuss which scenario is the reality by investigation of the fragment mass distribution. So far there are two characteristic fragment mass distributions observed in heavy ion collision experiments. T h e difference between two types is t h o u g h t to be relevant to the difference of the fragment source. In spectator region or in low energy collisions, the power law distribution a p p e a r s 1 . This power law is often regarded as a proof of liquid gas phase transition. I t is because Fisher's droplet m o d e l 2 can reproduce this power law with the assumption of t h e r m a l equilibrium. While, in participant region or high energy collisions, t h e distribution shows exponential shape 3 instead of power law. In such a high energy region, t h e thermal equilibrium cannot be assumed. Hence t h e models 331
332
assuming thermal equilibrium does not work because of non-equilibrium effects. These two types of distribution are often explained independently because the model describing one of them can not deal with the other. The point of this paper is to describe these two types of distribution by a single frame work. We introduce a nuclear matter model based on quantum molecular dynamics model (QMD) simulation 4,5 which is applicable to non-equilibrium (dynamical) phenomenon. In our model, we can deal with the nuclear matter with expansion by using a special type of periodic boundary condition 6 . This model has an advantage that the dynamics in expanding matter is microscopically described without any finite size effect7.
2
Model
The nuclear matter in our model is described by QMD with a constant expansion, i.e., with a radial collective motion. First we prepare nuclear matter with a given density po and initial temperature Tinit by the Metropolis sampling method under a periodic boundary condition. Then we give collective motion of particles as well as the extension of boundaries so that the whole system will expand consistently. Starting from this initial state, we solve the QMD equation of motion. We follow the time evolution till the density p becomes O.OOlpo and the final fragments are identified. We carry out 1000 events for each (h, Tinit) to obtain good statistics. In this model, the consistent expansion is essential. The collective motion of particles and the extension of walls with a fixed rate h are described as follows: P
*°"
=
^ P
MV w a „
=
h^^Pp,
F
,
(1) (2)
where Rj and P£°" are position and collective momentum of i-th particle and Pp is the Fermi momentum at p = p0. That is, i-th particle has a collective momentum proportional to its position and the rate is characterized by h. In the same manner, we deal with the velocity of the boundary V wa n so that it is proportional to its position R wa ii- Notice that the above relation applies for any reference frame we take, i.e., there is no special position of matter in our treatment.
333
3
Fragment mass distribution
Although there are two control parameters h,Tmit in our model, we concentrate on the case for Tjnjt = 30 MeV. In fact, we have found that this initial temperature does not affect the fragment mass distribution. Instead, the value of h is directly related with how the system expands and is more important. Figure 1 shows the fragment mass distribution at p = O.OOlpo averaged over 1000 events with a fixed pair of (h, Xlnit). The difference between the left and the right is only a horizontal scale. In the left hand side, a semi-log scale is used and the fragment from rapid expansion shows linear shape under this scale. This means the fragment yield with rapid expansion has an exponential dependence on the fragment mass. This exponential shape is very similar to that observed experimentally in participant region where the collective motion plays an important role. The slope of each line is also characteristic. As the speed of expansion (h) increases, the corresponding slope becomes steeper. The same feature is also observed in experiments 3 .
1
1
r r
:1
T ^=30 (MeV)
+
•
a
• a •
•"*_
"*
^^*^t
A=0.10(xl00.)1 A=0.50(xl0.) -, »=1.00(xl.0) = ft=2.0O(x0.1) 1
F
1
¥
*
r r
10 mass number A
20
D
,
\ ^
1
•
*
(T=3.10)l i
• A
10
^*>w * ,-
DQ
.
r r
m
•
*
1 I
10
mass number A
Figure 1: Fragment mass distribution for different expansion speed. Both panels show same data in different horizontal scales. Straight lines are eye-guides.
On the other hand, in the right hand side of Fig. 1, only h — 0.1 shows a linear shape. The linear feature in log-log scale is the power law. It is well known that Fisher's droplet model predicts the power law at the critical temperature. Nevertheless, we should keep in mind that the expanding matter is essentially not in thermal equilibrium. In other words, h = 0.1 is not small enough to realize a quasi-static expansion. Therefore, we cannot identify our power law as Fisher's power law because the Fisher's droplet model assumes complete thermal equilibrium 2 . There is a more crucial difference between
334
our power law and Fisher's. As mentioned above, the initial temperature does not affect final fragment mass distribution. In fact, even when the initial temperature is T;nit = 5 MeV, the power law distribution still appears as long as h = 0.1. This result definitely contradicts the mechanism of Fisher's power law. According to Fisher's droplet model, the power law appears when the temperature is equal to a critical temperature. Although the value of this critical temperature is not clear, it must be much higher than 5 MeV at least. Moreover, the effective temperature of matter decreases during expansion and becomes lower than Ti n i t . After all the power law from expanding matter has another mechanism other than the Fisher's liquid gas phase transition scenario. Taking into account that the power law still appears even T\nit = 5 MeV, we can conclude that the cracking scenario is more likely. Acknowledgments The authors thank A. Bonasera, V. N. Kondratyev, T. Kido and H. Takemoto for useful discussions. References 1. J. B. Elliot et al., Phys. Lett. B418, 34 (1998). 2. M. E. Fisher, Rep. Prog. Phys. 30 615 (1967); Proc. International School of Physics, Enrico Fermi Course LI, Critical Phenomena, ed. M. S. Green (Academic, New York, 1971). 3. W. Reisdorf et al., Nucl. Phys. A612 494 (1997). 4. T. Maruyama, K. Niita, K. Oyamatsu, T. Maruyama, S. Chiba, and A. Iwamoto, Phys. Rev. C57, 655 (1998). 5. T. Kido, T. Maruyama,K. Niita, S. Chiba Nucl. Phys. A663&664, 877c (2000). 6. B. L. Holian and D. E. Grady, Phys. Rev. Lett. 60, 1355 (1988). 7. S. Chikazumi, T. Maruyama, K. Niita, A. Iwamoto, Phys. Lett. B476, 273 (2000).
CONTEMPORARY PRESENCE OF DYNAMICAL AND STATISTICAL INTERMEDIATE MASS FRAGMENT PRODUCTION MECHANISMS IN MIDPERIPHERAL N I + N I COLLISIONS AT 30 MeV/nucleon P. M. M I L A Z Z O , M. S I S T O , G. V . M A R G A G L I O T T I , R. R U I Dipartimento
di Fisica and INFN, Trieste, 1-34127 Trieste, Italy E-mail: [email protected]
G. V A N N I N I , M. B R U N O , M. D ' A G O S T I N O Dipartimento
di Fisica and INFN, 1-40126 Bologna, Italy
Bologna,
N. C O L O N N A INFN, Bari, 1-70126 Bari, Italy C. A G O D I , R. A L B A , G. B E L L I A , M . C O L O N N A , R. C O N I G L I O N E , A. D E L Z O P P O , P. F I N O C C H I A R O , C. M A I O L I N O , E. M I G N E C O , P. P I A T T E L L I , D . SANTONOCITO, P.SAPIENZA INFN,
Laboratori Nazionali del Sud, 1-95123 Catania, Italy
Catania,
F . G R A M E G N A , P. F . M A S T I N U INFN,
Laboratori
Nazionali di Legnaro, Legnaro 1-35020 Legnaro, Italy
(Padova),
L. F A B B I E T T I , I. I O R I , A. M O R O N I Dipartimento
di Fisica and INFN, 1-20133 Milano, Italy
Milano,
M. B E L K A C E M University of Minnesota, MN-55455 Minneapolis, USA The 5 8 j V i + 5 8 Ni reaction at 30 MeV/nucleon has been experimentally investigated at the Superconducting Cyclotron of the INFN Laboratori Nazionali del Sud. In midperipheral collisions the production of massive fragments (4
335
336
1
Introduction
The production of intermediate mass fragments (IMF, 3
The experiment Experimental set-up
The experiment was performed at the INFN Laboratori Nazionali del Sud with MEDEA 8 and MULTICS 9 apparata. A beam of 58 Ni at 30 MeV/nucleon bombarded a 2 mg/cm 2 thick nickel target. The angular range 3° < 6iab <28° was covered by the MULTICS array 9 , consisting of 55 identical telescopes, composed of an Ionization Chamber (IC), a Silicon detector (Si) and a Csl crystal. Typical energy resolutions were 2%, 1% and 5% for IC, Si and Csl, respectively. The threshold for charge identification in the MULTICS array was about 1.5
337
MeV/nucleon. A good mass resolution for Z=l-6 isotopes was obtained above 8.5, 10.5, 14 MeV/nucleon for 4 He, 6 Li and 12 C nuclei, respectively 2 . Light charged particles (Z=l,2) and 7-rays were detected at 30° < 8iab <170° by the BaF2 ball of the MEDEA apparatus. The geometric acceptance of the combined array was greater than 90% of 4-zr. 2.2
Data selection
The results presented hereafter will refer to midperipheral events, with the aim of observing the IMF emitting sources. For the impact parameter selection we used the following approach: 1) selecting only the "complete" multifragment events, i.e. events where at least three IMF were produced and at least 80% of the total linear momentum was detected; 2) defining peripheral and midperipheral collisions those for which the heaviest fragment (with charge at least 1/3 of that of the projectile, i.e. Z>9) travels, in the laboratory frame, with a velocity higher than 80% of that of the projectile (vp=7.6 cm/ns). Accordingly, since the energy thresholds make not possible the detection of the QT reaction products, we find that the total detected charge {'LTOT) does not differ from that of the projectile for more than 30% (20
Results
In Fig.la-b we present the parallel component of the velocity (v p a r ), with respect to the beam direction), for carbon and oxygen nuclei. Two distinct distributions are evident: the first is centered at 6.5 cm/ns (the QP velocity) and the second one is at the center of mass velocity (3.8 cm/ns), intermediate between that of the target and of the projectile, because of the system symmetry. This is a common feature for all the detected IMF with Z<12. In midperipheral collisions three different emitting sources are present; there are events in which the IMF can be simultaneously produced by the decay of QP and QT (its fragments are not seen because under energy threshold for identification) sources and from a neck, forming a midvelocity emission source. Thus, disentangling the contributions from the QP and the neck sources becomes a mandatory requirement to improve the understanding of
338
the IMF production mechanism and perform comparisons of the IMF experimental yields with theoretical predictions. To this purpose we first study the process leading to the disassembly of the QP restricting the analysis to the fragments emitted with vpar >6.5 cm/ns. This constraint allows the selection of the decay products forward emitted in the QP decay with negligible contamination due to QT and midrapidity source emissions. We got indications that the QP reaches an equilibration stage before its de-excitation by means of the study of the energy and charge distributions. The energy distributions of the emitted isotopes present a Maxwellian shape with similar values for the slope parameter. Moreover, we compared the experimental charge distribution with the microcanonical Statistical Multifragmentation Model (SMM) predictions 10 , performed for a Ni nucleus at one third of the normal density. The experimental charge distribution is quite well reproduced assuming an excitation energy of 4 MeV/nucleon. Finally, we extracted the temperature T through the method of double ratios of isotope yields n . Since experimental temperature measurements are affected by secondary decays, the values have to be corrected as suggested in Ref. 12 ; their mean value, Tb=3.9±0.2 MeV, can be considered as the break-up temperature of the QP decaying system. In conclusion the present data analysis shows that the QP source has attained thermal equilibrium and that fragmentation is its main de-excitation process, well reproduced by a statistical approach. In coincidence with the statistical fragmentation of the QP, we observed the emission of IMF from a IS. In Fig.lc the ratio between the relative yields (YATBCX/YQP) is presented as a function of the atomic number Z. We observe a bell-like shape, peaked around Z=9. We notice the high probability of IMF emission from the neck zone. The presence of this maximum could be an effect of the breakup geometry; in fact, a rough consistency has been found with percolative calculations that compare emissions from a cylindrical shape neck, joining spherical QT and QP, and from the QP itself5'13. If on one side the QP disassembly is ruled by statistical models after thermal equilibrium has been reached, on the other side the neck emission exhibit quite different features that cannot be reproduced making statistical equilibrium assumptions. We have thus performed BNV calculations using different EOS parameters 3 . We found that with a compressibility term K of 200 MeV there is an evident massive neck formation (after 200 fm/c), that is not reabsorbed by the QP or the QT (this behaviour disappears increasing the K values). These calculations predict that on average we have a Z=8 fragment in the neck zone, and show that the IS fragment are neutron rich. Then, it is interesting to look at the 6He experimental yield. Its emission is quasi negligible from a statistical decay, both because the binding energy favours, for Z=2, a particle emission and because
339
4
v p a r9( c m /P n s )^
Figure 1: (a-b) v p a r distributions for Z=6,8, the arrows refer to the center of mass (CM) and QP velocities; (c) Ratio of the measured yields for neck fragmentation and QP emission; (d) Yields of 4 H e (dot-dashed line) and 6 H e (full line, multiplied by a factor 10) and their ratio as a function of v „ a r .
the N/Z value of 6He is quite different from the corresponding ratio of the system (N/Z=2 for 6 iJe,N/Z=30/28~l for the system). For these reasons the 6 He production should be more abundant in the neck zone with respect to the QP zone. In Fig. Id the experimental 4He and 6He yields are plotted versus vpar: the distribution of the heaviest He isotopes is mainly located where the midvelocity source dominates. This is a further indication of the presence of two different emitting systems following different de-excitation processes. In this sense, in midperipheral collisions we observe an IMF production which is due to two different mechanisms: one of statistical and the other of dynamical nature. 4
Conclusions
In the study of the Ni+Ni 30 MeV/nucleon dissipative midperipheral collisions it has been possible to reveal two different types of IMF emitting sources. We are in presence of a QP (and a QT), with an excitation energy which leads to the multifragmentation regime; its decay can be fully explained in terms of a statistical disassembly of a thermalized system (T=3.9±0.2 MeV, E* ~4 MeV/nucleon). Contemporary to the IMF production from the QP source an intermediate source is formed, emitting both light particles and
340
IMF. These fragments are more neutron rich than the average matter of the overall system and have a quite different charge distribution, with respect to the ones statistically emitted from the QP. These features can be considered as a signature of the dynamical origin of the midvelocity emission. References 1. J. Pochodzalla et al., Phys. Rev. Lett., 75 1040 (1995); J. B. Elliot et al., Phys. Rev. C 49, 3185 (1994); M. L. Gilkes et al., Phys. Rev. Lett., 73, 1590 (1994); J. A. Hauger et al, Phys. Rev. C 57, 764 (1998); P. F. Mastinu et al., Phys. Rev. Lett., 76, 2646 (1996); M. D'Agostino et al., Nucl. Phys. A650, 329 (1999). 2. P. M. Milazzo et al., Phys. Rev. C 58, 953 (1998). 3. W. Bauer et al., Phys. Rev. Lett., 69, 1888 (1992); M. Colonna et al., Nucl. Phys., A583, 525 (1995); M. Colonna et al., Nucl. Phys., A589, 2671 (1995). 4. G. Casini et al, Phys. Rev. Lett., 71, 2567 (1993); J. Toke et al., Phys. Rev. Lett., 75, 2920 (1995); Y. Larochelle et al., Phys. Rev. C 55, 1869 (1997); J. Lukasik et al., Phys. Rev. C 55, 1906 (1997); P. Pawlowski et al., Phys. Rev. C 57, 1711 (1998). 5. C. P. Montoya et al., Phys. Rev. Lett., 73, 3070 (1994). 6. L. G. Sobotka et al., Phys. Rev. C 50, R1272 (1994). 7. J. F. Dempsey et al., Phys. Rev. C 54, 1710 (1996). 8. E. Migneco et al., Nucl. Instr. and Meth. Phys. Res., A314 31 (1992). 9. I. Iori et al., Nucl. Instr. and Meth. Phys. Res., A325 458 (1993). 10. J.P. Bondorf et al., Nucl. Phys. A444 (1986) 460; A. S. Botvina et al., Nucl. Phys. A475 (1987) 663; J.P. Bondorf et al., Phys. Rep. 257 (1995) 133. 11. S. Albergo et al, Nuovo Cimento 89, 1 (1985). 12. M. B. Tsang et al, Phys. Rev. Lett., 78, 3836 (1997); H.Xi et al., Phys. Lett. B431 8 (1998). 13. W. Bauer et al., Phys. Rev. C 38, 1927 (1988).
E X T E R N A L COULOMB A N D A N G U L A R M O M E N T U M I N F L U E N C E O N ISOTOPE COMPOSITION OF N U C L E A R FRAGMENTS
Institute
A.S. B O T V I N A GANIL, 1+076 Caen, France a n d for Nuclear Research, 117312 Moscow,
Russia.
The Markov chain statistical multifragmentation model predicts inhomogeneous distributions of fragments and their isospin in the freeze-out volume caused by an angular momentum and external long-range Coulomb field. These effects can take place in peripheral nucleus-nucleus collisions at intermediate energies and lead to neutron-rich isotopes produced in the midrapidity kinematic region of the reactions.
Studies of multifragmentation phenomenon in heavy-ion reactions at high energies are very promising because of overlapping nuclear physics with universal physical processes taking place in finite particle systems. In particular, nuclear equations of state and phase transitions can be established *. As other complicated many-body processes this phenomenon can be successfully treated in statistical way 2,3 : Fragment production in both peripheral and central collisions has clear statistical features 2 ' 4 , 5 , though a considerable preequilibrium emission and collective energy (radial flow) should be taken into account. In finite-size nuclear systems statistical processes can lead to unusual effects since the fragment formation is governed by both short-range nuclear forces and long-range Coulomb forces. For example, a Coulomb interaction of the target and projectile-like sources leads to a predominant midrapidity ("neck"-like) emission of intermediate mass fragments (IMF, charges Z=3-20) 6 . In this contribution I show that a statistical process can also provide a non-isotropic fragment isospin production in peripheral nucleus-nucleus collisions. The statistical multifragmentation model (SMM) is described in detail in many publications 2 . The model is based upon the assumption of statistical equilibrium at a low-density freeze-out stage. All possible break-up channels (partitions into fragments) are considered with weights defined by the entropies of the channels which depend on excitation energy E*, mass number As, charge Za and other parameters of the source. After break-up of the nuclear source the fragments propagate independently in their mutual Coulomb fields and undergo secondary decays. The new version of SMM version is based on producing the Markov chain of partitions which exactly characterize the whole partition ensemble 7 . In a special way individual partitions are generated and selected into the chain by applying the Metropolis receipt 7 , s . Within this 341
342
method primary hot fragments can be placed directly into the freeze-out volume to calculate their Coulomb interaction and moment of inertia. In this way one can take into account the correlations between positions of the primary fragments and their Coulomb energy that influences the partition probabilities. Angular momentum conservation can be included within this method similar to Refs.3,9. The full analysis of the Markov chain SMM appears somewhere 8 .
Figure 1: The neutron-to-proton ratio N / Z and relative yield of hot primary fragments produced in the freeze-out after break-up of Au nucleus. Solid lines: Markov chain SMM calculations for a thermal source with excitation energy 3 MeV/nucleon, dashed lines: the same source with angular momentum 150ft. 50
100 150 A, fragment mass number
200
An angular momentum influence on isospin of fragments emitted from a single source is instructive. In Fig. 1 I show yields and neutron-to^proton (N/Z) ratios of hot primary fragments produced in the freeze-out volume (the density is ps=Po/6, Po is normal nuclear density) after break-up of Au source ( J 4 S =197, Zs=79) at E*=3 MeV/nucleon. It is seen that an angular momentum favors fission-like fragment partitions with two large equal-size fragments (see also Refs.3,9). That is different from a normal fragmentation pattern dominated by partitions with different-size fragments. An angular momentum leads to increasing N/Z ratio of IMF also. The last effect is important and has a simple qualitative explanation: An angular momentum favors emission of IMF with larger mass numbers since the system in the freeze-out needs to have a large moment of inertia in oder to minimize rotational energy and maximize the entropy. From another side a Coulomb interaction prevents to emit IMF with large charge Z. As a result of interplay of these two factors we obtain the increasing of the N/Z ratio. In peripheral nucleus-nucleus collisions at the projectile energies of 10100 MeV/nucleon a break-up of highly excited projectiles-like nuclei is fast (the characteristic time is around 100 frn/c) and happens in the vicinity of the target-like nuclei. The influence of the Coulomb field of the target nucleus on fragmentation of the projectile source increases charge asymmetry of produced fragments and leads to non-isotropic fragments emission: small fragments are
343
preferably emitted to the side of the target 6 . Within the Markov chain SMM one can study how this effect influences the isotope composition of fragments. Calculations were performed for the same Au source as in Fig. 1. The source was placed at a fixed distance (20 frn) from another Au source. This distance was obtained under assumption that the break-up happens in ~100 fm/c after a peripheral collision of 35 A-MeV projectile Au with target Au. It is naturally to expect the decays happen at different distances, excitation energies and angular momenta. In statistical approach we can take into account a distribution of the sources in distances and other characteristics by considering an ensemble of the sources. Parameters of this ensemble can be found by global comparison with the experiment 2 ' 10 . However, the present approximation of a fixed distance is sufficient for qualitative identification of new statistical effects.
Figure 2: Freeze-out volume coordinate distributions of neutron-to-proton ratio N / Z of primary fragments with Z=8 (top panel) and relative yields of the primary Z=8 and biggest fragments (middle and bottom panels) produced at break-up of Au source at excitation energy 3 MeV/nucleon. The second Au nucleus is placed at -20 fm from the center of the freeze-out. Dotted lines: the isolated Au source, dashed lines: Coulomb influence of the second Au is included, solid lines: angular momentum 150 h is included additionally.
-io
-
s
o
s
io
fragment coordinates in freeze-out, fm
In Fig. 2 I show distributions of yields and N/Z ratio for hot primary IMF with Z=8 and the biggest fragments in the freeze-out volume along the axis connecting the projectile and target sources. It is seen that in case of a single isolated source all distributions in the freeze-out are symmetric respective to the center mass of the source. In case of the target Coulomb influence the IMF are mainly produced closer to the target while the biggest fragments are shifted to the opposite direction. These locations of fragments provide minimum of Coulomb energy in the target-projectile system. However, the external Coulomb alone influences hardly the fragment isospin distribution. In case of angular momentum the N/Z ratio of the IMF increases considerably
344
and becomes larger when the IMF are closer to the target. The reason is again an interplay of the Coulomb and rotational energy: The system needs more heavy IMF to have a large moment of inertia while the Coulomb energy of the system depends also on IMF distance from the target and this energy is lower when the IMF charge is small. This asymmetry of the IMF isospin distribution survives after secondary deexcitation of hot fragments. The following Coulomb propagation push the IMF in the direction of the target providing predominant population of the midrapidity kinematic region by neutron-rich fragments. The Coulomb repulsion may be not sufficient to accelerate fragments up to high energy, however, it can fill with the fragments a considerable part of the midrapidity region 6 . Within the statistical picture a slight radial flow can supply the IMF with high velocities to populate the center of the midrapidity zone. In conclusion, it was shown that in peripheral nucleus-nucleus collisions characteristics of statistically produced fragments depend on Coulomb interaction between the target- and projectile-like sources and an angular momentum transferred to the sources. In particular, it leads to space asymmetry of both fragment emission and their isotope composition respective to the sources. Previously the symmetry violation was considered as a sign of a dynamical "neck" emission. However, there is an alternative statistical explanation: the symmetry of the phase space is deformed under interaction of the two sources. Theoretically such a process gives an example of a new kind of statistical phenomenon influenced by an inhomogeneous external long-range field6. References 1. See for instance: "Multifragmentation", Proceedings of the International Workshop 27 on Gross Properties of Nuclei and Nuclear Excitations, Hirschegg, Austria, January 17-23, 1999. GSI, Darmstadt, 1999. 2. J.P.Bondorf, A.S.Botvina, A.S.Iljinov, I.N.Mishustin, K.Sneppen. Phys. Rep. 257, 133(1995); A.S.Botvina et al., Nucl.Phys. A475, 663(1987). 3. D.H.E.Gross, Phys.Rep. 279, 119 (1997). 4. M.D'Agostino et al, Nucl. Phys. A650, 329 (1999). 5. M.D'Agostino et al, Phys. Lett. B371, 175 (1996). 6. A.S.Botvina, M.Bruno, M.D'Agostino and D.H.E.Gross, Phys. Rev. C59, 3444 (1999). 7. A.S.Botvina, A.D.Jackson and I.N.Mishustin, P%s.JRew.E62,R64(2000). 8. A.S.Botvina, I.N.Mishustin et al, in preparation. 9. A.S.Botvina and D.H.E.Gross, Nucl Phys. A592, 257 (1995). 10. P.Desesquelles et al, Nucl. Phys. A633, 547 (1998).
Section V Intermediate Energy Heavy-Ion Reactions
ISOSPIN FRACTIONATION IN EXCITED N U C L E A R SYSTEMS
S.J. YENNELLO, M. VESLESKY, E. MARTIN, R. LAFOREST, D. ROWLAND, E. RAMAKRISHNAN, A. R U A N G M A E. M. WINCHESTER Cyclotron Institute, Texas A&M University, College Station, Texas 77845-3366,
USA
It has been suggested that a non-homogeneous distribution of isospin could develop during the fragmentation process of excited nuclear material. Originally this was theorized as an outgrowth of a distillation-like process as the nucleus went though a liquid-gas phase transition [1,2]. More recently the phenomenon has been predicted with a lattice gas calculation [3] as well as dynamical models [4,5]. There is some indication in inclusive data that such a separation into a neutron-rich gas and a more symmetric liquid does in fact take place [6]. Further evidence can be gathered by looking at selected events from heavy-ion collisions [7-9]. And very recently there has been an experiment where the reconstruction of the whole system in mass and charge shows this effect [10]. This paper will briefly review the groundwork for fractionation and then elaborate on recent completely reconstructed event from the recent experiment of 2SSi + U2 ' 124Sn.
1
Introduction
There has been significant interest in multifragmentation of excited nuclear matter for many years. While there has been some success in understanding the process of multifragmentation and describing this phenomenon in terms of a liquid gas phase transition [11], those efforts have often treated the nucleus as a single component nuclear liquid. In fact the nucleus is a two component nuclear liquid. Early work by Lamb [1] laid the foundations for treating the nucleus as a two-component system although the results were not connected to multifragmentation. Statistical calculations describing multifragment disassembly predicted that much of the neutron excess would be observed as free neutrons [12]. Thermodynamic calculations by Muller and Serot [2] lead us to the idea that, for the very neutron rich systems, there may exist a distribution of the excited nuclear matter into a neutron rich gas and a more symmetric liquid. It has also been predicted by both lattice gas and mean field calculations [4,5] that fragmentation of a system of asymmetric nuclear matter will express the characteristics of a two component system. This would result in a second order phase transition. Also within a dynamical code, constructed to study influence of charge asymmetry on spinoidal decomposition of nuclear matter at sub-saturation densities, differences in the fragmentation are seen as a function of isospin asymmetry [5]. The difference of the mean N/Z ratio of the light charged particles (LCP) and of the intermediate mass fragments (IMF) may be a possible experimental signature of a separation into a gas (resulting mostly in emitted LCPs) and a liquid (IMFs). An enhanced production of neutron rich H and He isotopes 347
348
from the neck region has been seen experimentally in mid-peripheral collisions [7,13] while a favored emission of more symmetric heavy clusters in the midrapidity region has also been shown [8]. Quite neutron deficient residues have also been seen in intermediate energy reactions [14,15]. Recent results from the reaction 112,124 Sn + U2124 Sn indicate that the relative abundance of free neutrons increases of as the neutron content of the colliding system increases [9], The present work will demonstrate that, in a well defined system that has isotopic identification of all charged fragments, the mean value of N/Z ratio of LCPs is more sensitive to N/Z of quasiprojectile than mean N/Z ratio of IMFs. The more asymmetric the system is the stronger it will favor breaking up into still more neutron rich (deficient) light fragments while the N/Z ratio of heavier fragments remains relatively insensitive. If an inhomogeneous distribution of isospin develops resulting in lower density regions being more neutron rich and higher density regions being more isospin symmetric this may be observable by looking at the neutron content of fragments as a function of cluster size. In figure 1 we show the average neutron to proton ratio as a function of the charge for the reaction of p+Xe. The data shows a decrease in the neutron richness of fragments as one goes from lower to higher Z values.
E
5
A M V
1.40 1.35 1.30 1.25 1.20 1.15 1.10 1.05 1.00 0.95
-
—T—Exp —•—SMM
T ^S
vj
^
4W kX i±x-H
*
>i:
i
1
1
i
w
t-*r
\T/
4-
1
i
•——•—- • — « i
I
I
I
!
Figure 1 Average neutron to proton ratio bound in clusters as a function of cluster charge. Data taken fromref 16.
2
Experimental details
The experiment was done with a beam of 28Si impinging on 1 mg/cm2 self supporting m 124Sn targets. The beam was delivered at 30 and 50 MeV/nucleon by the K500 superconducting cyclotron at the Cyclotron Institute of Texas A&M University. The detector setup [17] was composed of an arrangement of 68 silicon CsI(Tl) telescopes covering polar angles from 2.3° to 33.6° in the laboratory system. Each element is composed of a 300nm silicon detector followed by a 3cm CsI(Tl) crystal. The detectors are arranged in five concentric rings. The geometrical
349
efficiency is approximately 90% for covered angle range. These detectors allow isotopic identification of light charged particles and intermediate-mass fragments up to a charge of Zf = 5. The energy thresholds are determined by the energy needed to punch through the 300|jm silicon detector. These energy thresholds have little effect on the acceptance of particles from the fragmenting projectile, especially at 50 MeV/nucleon, due to the boost from the beam energy. A more detailed description of the experimental procedure and detector calibration can be found in ref. [18]. In the present study we restrict ourselves to the events where all emitted fragments are isotopically identified. The total charge of quasiprojectile (QP) is restricted to the values equal to the charge of projectile Z,ot = 14. This highly exclusive set of data contains information on fragmentation of highly excited projectile-like prefragments over a range of isospin and thus can be used for the study of the effect of the neutron to proton ratio on the fragmentation. Results
(b)
10''
B • •
10'
•
• •
10'
f 0.6
08
1.0
N/Z„„
1.2
0.6
0.8
1.0
1.2
N/Z„„
Figure 2. 3H/*He as a function of the N/Z of the reconstructed quasiprojectile. (a) 30 MeV/nucleon (b) 50 MeV/nucleon solid squares are from the reaction of ^Si + " 2 Sn, open squares are from the reaction of "Si + '^Sn.
In figure 2 is shown the dependence of the isobaric yield ratio 3H/3He on (N/Z)QP . The yield ratio 3H/3He increases as (N/Z)QP increases. Similar effects have been seen in other systems where isotopic or isobaric ratios have been plotted as a function of N/Z of the reacting system [7,20]. In the present data the (N/Z)QP is the reconstructed N/Z of all the detected charged particles. If one fits the data with an exponential function the slope is nearly identical for the reaction with the m Sn target as with the 124Sn target. The magnitude of the slopes show that this increase in the yield ratio 3H/3He as a function of (N/Z)QP is much larger than would be
350
expected from the change in (N/Z)Q P alone. While the integrated values on the two targets are different when the ratios are broken down by the (N/Z)QP the slope is similar. If one looks at the difference with bombarding energy the dependence of yield ratio 3H/3He on (N/Z)QP is lessened at the higher energy. The lattice gas calculation [3] predicted that the 3H/3He ratio would be much more sensitive at lower temperatures. At relatively high temperatures the ratio would asymptotically approach the (N/Z)QP value. The hybrid DIT/SMM calculations reproduce the experimental yield ratio 3H/3He rather well. This shows that statistical model of multifragmentation describes correctly the influence of isospin on the production rates of fragments. The overall 3H/3He ratio for the reaction with the 112Sn target is 0.25 and with the 124Sn target is 0.35 at 30 MeV/nucleon and 0.44 and 0.62 respectively at 50 MeV/nucleon. This increase in the yield of neutron rich fragments with an increase in the N/Z of the reacting system is consistent with other studies of isobaric ratios [9,20]. What figure 2 demonstrates is that this is an integrated feature and mearly shows that the system created with the neutron rich target is more neutron rich. But when one takes the 3H/3He ratio as a function of the (N/Z)QP the behavior may be seen much more clearly.
In figure 3a we show the multiplicity of charged particles versus (N/Z)QP, again for the reaction 28Si + 112Sn at 50 MeV/nucleon. The squares represent the multiplicity of all charged particles. The multiplicity of charged particles increases as (N/Z)QP decreases. This is then broken down into the multiplicity of light charged particles (circles) and the multiplicity of intermediate mass fragments (triangles). Here we can see that the multiplicity of LCPs increases with decreasing (N/Z)QP . Meanwhile the multiplicity of IMFs increases as (N/Z)QP increases. The situation is similar for all cases. The results are identical for different targets at the same projectile energy. The dependence of multiplicity of IMFs on (N/Z)QP is practically the same for all cases and the increase in overall multiplicity of charged particles with increasing beam energy is identical with the increase of the multiplicity of LCPs. Again presented information is consistent with, but much more illustrative than previous works where the difference in isospin of the excited system was only known approximately on average [21,22]. Previous works show that the multiplicity of IMFs as a function of multiplicity of charged particles increases for the more neutron rich system. While this is true the present data shows that this is largely an effect of the decrease in the multiplicity of LCPs at neutron rich system. Our data shows that this effect weakens at higher energies possibly toward the disappearance that was seen by Miller et al. [22] at a much higher energy and would be consistent with the temperature dependence predicted by lattice-gas model calculations [3].
351
0.6
0.8
1.0
N/Z„„
1.2
0.4 0.6 0.8 1.0 1.2 1.4 N/Z„„
FIG. 3: Dependences of emission observables on N/Z of the quasiprojectile. Data are given for reaction 2S Si(50MeV/nucleon)+ 112Sn. (a) - Multiplicity of all charged particles (squares), light charged particles (circles)and intermediate mass fragments (triangles) versus the N/Z of the quasiprojectile. (b) Experimental ratio of mean values of N/Z of light charged particles and intermediate mass fragments versus the N/Z of the quasiprojectile.
On thefigure3b we show the ratio (NIZ)mF/(N/Z)u;V as a function of (N/Z)QP. The ratio (N/ZW/CN/Z^p decreases with increasing (N/Z)QP. As there are less neutrons available the excess protons go into the smaller fragments rather than the larger fragments. The least neutron rich quasiprojectiles with (N/Z)QP =0.5 prefer to breakup into very neutron deficient LCPs and much more isospin symmetric IMFs. The observed inhomogeneous isospin distribution is likely caused by more favorable energy balance of the deposition of proton excess into LCPs, either free protons or light proton rich clusters. When extrapolating this trend towards very neutron rich quasiprojectiles the non-homogeneous isospin distribution would most likely reappear again in the form of neutron rich LCPs and more symmetric IMFs. Such an extrapolation is consistent with the predicted asymmetric liquid gas phase transition [2] for neutron rich nuclei corresponding to proton concentrations 40% or less. The data presented here raises the question if an analogous phase transition may be expected for very neutron deficient fragmenting systems. 4
Summary
There is significant evidence for the development of an inhomogeneous distribution of isospin during fragmentation of excited nuclei. Enhancement in me neutron content of light particles has been observed in this as well as other work [6,7,9,10]. A complementary decrease in the neutron content of heavier fragments has been observed in the current study. This depends more significantly on the fragmenting system than on the reacting system. At higher energies this dependence is lessened. The multiplicity of charged particles depends on the isospin of the fragmenting
352
system. As the (N/Z)QP decreases the overall multiplicity of charged particles increases. The multiplicity of LCPs dramatically increases while the multiplicity of IMFs decreases. This effect is less significant at higher bombarding energy. These new isospin observables hopefully will lead to a greater understanding of reaction mechanisms and the equation of state when coupled to continued advances in theory. 5
Acknowledgements
This work was supported in part by the NSF through grant No. PHY-9457376, the Robert A. Welch Foundation through grant No. A-1266, and the Department of Energy through grant No. DE-FG03-93ER40773. 6
References
1. D.Q. Lamb, J.M. Lattimer, C.J. Pethick, and D.G. Ravenhall, Nucl. Phys. A 360,459 (1981). 2. H. Muller and B.D. Serot, Phys. Rev. C 52, 2072 (1995). 3. Ph. Chomaz, and F. Gulminelli, Phys. Lett. B 447, 221 (1999). 4. B.A. Li and CM. Ko, Nucl. Phys. A 618,498 (1997). 5. V. Baran, M. Colonna, M. Di Toro, and A.B. Larionov, Nucl. Phys. A 632, 287 (1998). 6. E. Martin, et al., Phys. Rev. C. (2000). 7. J.F. Dempsey et al., Phys. Rev. C 54, 1710 (1996). 8. E. Ramakrishnan et al., Phys. Rev. C 57,1803 (1998). 9. H.S. Xu et al., MSU Preprint MSUCL-1137, 1999. 10. M. Veselsky et al., submitted to Phys. Rev. C, nucl-ex/0002007. 11. J.E. Finn et al., Phys. Rev. Lett. 49, 1321 (1982). 12. J. Randrup and S.E. Koonin, A 356, 223 (1981). 13. L.G. Sobotka et al., Phys. Rev. C 50, R1272 (1994); Phys. Rev. C 55, 2109 (1997). 14. K.A. Hanold, et al., Phys. Rev. C 52,1462 (1995). 15. M. Gonin et al., Phys. Rev. C 42, 2125 (1990). 16. A.S. Hirsch, et al., Phys. Rev. C 29, 508 (1984). 17. F. Gimeno-Nogues et al., Nucl. Inst, and Meth. in Phys. Res., A 399, 94 (1997). 18. R. Laforest et al., Phys. Rev. C 59, 2567 (1999). 19. A. S. Hirsch, et al., Phys. Rev. C 29, 508 (1984). 20. S.J. Yennello et al., Phys. Lett. B 321, 14 (1994); H. Johnston et al., Phys. Lett. B 371,186 (1996). 21. G.J. Kunde et al., Phys. Rev. Lett. 77, 2897 (1996). 22. M.L. Miller et al., Phys. Rev. Lett. 82, 1399 (1999).
THE DISAPPEARANCE OF FLOW AND THE NUCLEAR EQUATION OF STATE GARY D. WESTFALL National Superconducting Cyclotron Laboratory and Department of Physics and Astronomy, Michigan State University, East Lansing, Michigan 48824-1321 USA E-mail: [email protected] We present results for transverse flow, the balance energy, and elliptic flow for Ar+Sc, Fe+58Fe, S8Ni+58Ni, Kr+Nb, and Au+Au. Wefindthat a soft equation of state is required to explain these results. We find that a momentum dependent formulation of the nuclear mean field is required. A 30% density dependent reduction of the in-medium nucleon-nucleon cross sections in BUU is required to reproduce the observed balance energies for light systems.
58
1
The Disappearance of Flow
The study of collective transverse flow in nucleus-nucleus collisions can provide information about the nuclear equation of state (EOS) [1-8]. Transport models have been used to describe collective transverse flow in terms of nucleon-nucleon scattering in a nuclear mean field [9-13]. These models predict that collective transverse flow in the reaction plane disappears at an incident energy, termed the balance energy, Eba) [12], where the attractive scattering (dominant at energies around 10 MeV/nucleon) balances the repulsive interactions (dominant at energies around 400 MeV/nucleon) [9,10], The disappearance of directed transverse flow has been well established through many experiments [14-27]. Comparison of the measured impact parameter dependence of the balance energy with predictions from Quantum Molecular Dynamics (QMD) model calculations demonstrated better agreement with a formulation which incorporates momentum dependence in the mean field [28,29]. Flow and Ebai have been shown to depend on the isospin of the system [25,26]. Ebai has been measured for Au+Au and has been used to show that the nuclear equation of state is soft [30]. 1.1
Momentum Dependence of the Mean Field
Dynamical transport model calculations can incorporate soft and stiff descriptions of the nuclear EOS as well as momentum dependence in the mean field. Predictions of Quantum Molecular Dynamics (QMD) model calculations [28] are displayed in Fig. 1 for a stiff equation of state with momentum dependence (open circles) and without momentum dependence (open squares) for 40Ca+40Ca reactions. Also shown in this figure are the measured values of the balance energies for 40Ar+45Sc reactions 353
354
extracted for six reduced impact parameter bins (solid triangles) [29]. These experimental values for Ebai(b) are plotted at the upper limit of each reduced impact parameter bin in Fig. 1. Ebai(b) increases linearly as a function of the impact parameter. The data suggest better agreement with the QMD model calculations which include momentum dependence in the nuclear mean field, in agreement with the results from studies of nucleus-nucleus collisions at higher bombarding energies [3,4]. Here we are able to place this additional constraint on the nuclear EOS by measuring the balance energies for peripheral heavy-ion collisions. 1.2
160 -o
8
140
u A
QMD with momentum dependence QMD without momentum dependence Data
Figure 1. Measured balance energies for 4"Ar+45Sc reactions at six reduced impact parameter bins [29] compared with the predictions of QMD model [28] with and without momentum dependence in the mean field for ""Ca+^'Ca reactions.
Isospin Dependence
Flow and Ebai are thought to arise from a delicate balance between the attractive nuclear mean field and repulsive nucleon-nucleon interactions. Because the n-p scattering cross sections are about 3 times larger than n-n or p-p cross sections in this energy range, varying the relative number of neutrons and protons will alter the resulting flow and Ebai. To carry out this study, three symmetric systems were chosen with different isospin: 58Ni+58Ni, 58Fe+58Fe, and 58Mn+58Fe [25,26]. Keeping the mass number of the projectile and target nuclei the same and varying the N/Z of the nuclei allowed the study of systems with N/Z ranging from 1.07 to 1.2. This choice of systems keeps the excitation energies and kinematics of the three systems constant removing any acceptance differences. 1.2.1
Isospin Dependence of Flow
Experimental evidence for the isospin dependence of directed transverse flow is shown in Fig. 2 [25]. The extracted values of the collective transverse flow in the reaction plane are displayed as a function of the reduced impact parameter for three different fragment types from 58 Fe + 58 Fe and 58 Ni + 58Ni collisions at 55 MeV/nucleon. The extracted values of the flow are plotted at the upper limit of each reduced impact parameter bin.
355 60
(a)Z = 3
. 'WFe
"+-.
o s,Ni+MNi
.-•*-. ^
-+- -+>
30
2
?0
% o
II)
• "W!Fe
(b)Z = 2
o ™Ni+"Ni
(1)
E
-H A 5W*Fe
1
-+-
(c) Z = 1
"Ni+"Ni
'-+ A''•4 0.3
0.4
0.5
0.6
0.7
0.8
0.9
b/b m „,
Figure 2. Directed transverse flow as a function of the reduced impact parameter for three different fragment types from 58Fe+58Fe and ,8Ni+58Ni collisions at 55 MeV/nucleon [25].
1.2.2
The neutron-rich system 58 Fe+ 58 Fe systematically exhibits larger flow values than 58Ni+58Ni for all three particle types at all reduced impact parameter bins displayed (except for Z=3 in the most peripheral bin). The largest difference in the magnitude of the flow between the isotopic entrance channels occurs for heavier mass fragments in semi-central collisions. The impact parameter dependence of the directed transverse flow shown is in qualitative agreement with previous work [2,16,24], because the flow is maximal for semi-central events. The mass dependence of the directed transverse flow shown in Fig. 2 also demonstrates the well-known increase in magnitude for heavier fragments [2,20,29].
Isospin Dependence of Ebai
Experimental evidence for the isospin dependence of the balance energy is presented in Fig. 3 [26]. The solid squares (circles) are the measured values of the balance energies for 58Fe+58Fe (58Ni+58Ni) extracted for four reduced impact parameter bins. The balance energy increases as a function of impact parameter for both isotopic systems in agreement with previous work [16,23,24], and Ebai(b) is systematically higher for the more neutron-rich system at all measured reduced impact parameter bins. The predictions of BUU model [32,33] calculations, which incorporate an isospin dependent potential
Figure 3. The solid squares (circles) represent the data for 58Fe+58Fe (58Ni+58Ni). The open squares (circles) are BUU predictions for 58 Fe+58Fe (58Ni+58Ni) with reduced cross sections [31]. The dashed (solid) lines represent the BUU calculations with vacuum cross sections for 58Fe+,8Fe (58Ni+S8Ni) [26].
356
and isospin-dependent nucleon-nucleon scattering cross sections for Fe+ Fe and 58 Ni+58Ni are shown as lines in Fig. 3. The errors shown on the calculated points are statistical, and the values for 58Ni+58Ni have been slightly offset in the horizontal direction to show the error bars more clearly. That the balance energy is the same value for all fragment types [20,29] facilitates comparison of the measured values of Ebai(b) to predictions of transport models calculations which involve only nucleons. The BUU calculation by Li et al. [26,32,33] reproduces the observed difference between the neutron-rich and neutron deficient systems. However, this calculation does not reproduce the absolute value of Ebai for central collisions. The BUU calculations of Daffin and Bauer [31] including a 30% density dependent reduction of the in-medium nucleon-nucleon scattering cross sections reproduces the absolute value of Ebai in central collisions but does not reproduce the difference between the two systems. 1.3
Ebai for Au+Au
Flow for Au+Au collisions is plotted vs. incident beam energy in Fig. 4 (solid squares) for three particles of interest. For Z=l, the extracted flow is very weak for all of the measured energies. This small flow is partially due to the well-known fragment mass dependence of flow and to the higher energy threshold for Z=l. However, for Z=2 and Z=3 the figure clearly shows that flow goes through a minimum. Because our measurements are unable to distinguish between the negative attractive scattering, which dominates below Ebal, and the positive repulsive scattering, which dominates above Ebai, such a minimum is indicative of the Ebai for the system. The dotted curves are parabolic fits. Also, the flow data below the minimum are reflected about the x axis to represent the attractive regime, and the dashed (solid) linear least squares fits are with the 44.5 MeV/nucleon data point reflected (not reflected).
30
40
50
60 30
40
50
60 30
40
50
60
Beam Energy (MeV/nucleon) Figure 4. Extracted flow vs. incident beam energy in central collisions for a)Z=l, b)Z=2, and c)Z=3 [27]. Solid squares are experimental data, open squares are reflected about the x-axis to represent attractive scattering. The dashed (solid) linear least squares fit is with the 44.5 AMeV data point reflected (not reflected). The dotted curve is a parabolic fit.
In the same figure, we also show the result of a parabolic fit of the type employed previously. We can see that the balance energies for Z=2 and Z=3 are the same to within error bars, in agreement with previous experimental studies that showed no dependence on particle type [20]. Combining the statistical and systematic errors, we obtain for the Au+Au system the balance energy 42 ± 3stat. ± lsys. MeV/nucleon.
357
1.3.1
Toward the Extraction of K i
Because Ebal for Au+Au is nearly independent of the reduction in inmedium cross sections and the impact parameter, BUU predictions can be compared directly to the experimental value of the balance energy to estimate the nuclear compressibility K [30]. This lack of dependence on b and a for Au+Au differs from lighter systems that showed strong dependence on b and a, which made the isolation of K difficult. In Fig. 5, BUU balance energies for Au+Au are plotted vs. a, and a horizontal line represents the experimental value. Only K=200 MeV, which corresponds to a soft equation of state, falls within error bars of the measurement. The approximate value of K is in good agreement with other measurement techniques [4,34,35].
i
i
i
• • K=380 MeV (stiff) 'a ™ - • K=235MeV A K=200 MeV (soft) o g 65 - I Data % 60 >, 55
[
* < 'i
7
a so : V
£ 45 •a « 40 -
. -
1 1 '
Ml 0)
,
-
1
- -
blb
nux = •• 28 ( most sent"')
35
0
0.1
0.2
0.3
-a Figure 5. BUU balance energies plotted as a function of the a„„ reduction parameter -a for three different values of the compressibility K for Au+Au [30], A flat line with error bars represents the experimental measurement.
Elliptic Flow Elliptic flow measurements at energies between 2 and 8 AGeV have been shown to be sensitive to parameters of the nuclear equation of state [36,37]. Elliptic flow measurements around the balance energy have also been shown to be sensitive to the equation of state [38]. In Fig. 6 we present measured scaled elliptic flow data for Ar+Sc [39] compared with IBUU calculations for 48 Ca+48Ca [38]. Clearly the predictions for the soft equation of state agree much better than those or the stiff equation of state with the data.
•K
!
: f •3-0.5
\
.
\ ''
^
1\
-Q-
IBUU Sliff
4ir Duiu
\
FVJ
'v^, 7^ L
"•J---
I
;
i
:
•
60 80 Beam Energy (AMeV)
Figure 6. Scaled elliptic flow measured for Ar+Sc [39] compared with IBUU calculations [38] for 48 Ca+48Ca for a stiff equation of state and a soft equation of state with reduced 0„„.
358
References 1. 2. 3. 4. 5. 6. 7. 8. 9. 10. 11. 12. 13. 14. 15. 16. 17. 18. 19. 20. 21. 22. 23. 24. 25. 26. 27. 28. 29. 30. 31. 32. 33. 34. 35. 36. 37. 38. 39.
H. Stocker and W. Greiner, Phys. Rep. 137, 277 (1986). H.H. Gutbrod, A.M. Poskanzer, and H.G. Ritter, Rep. Prog. 52, 1267 (1989). C. Gale et al., Phys. Rev. C 41, 1545 (1990). Q. Pan and P. Danielewicz, Phys. Rev. Lett. 70, 2062 (1993). G. Peilert, H. Stocker, and W. Greiner, Rep. Prog. Phys. 57, 533 (1994). M.D. Partlan et al., Phys. Rev. Lett. 75, 2100 (1995). N. Herrmann, J.P. Wessels, and T. Wienold, Ann. Rev. Nuc. 49, 581 (1999). H. Liu et al. Phys. Rev. Lett.75, 5488 (2000). J.J. Molitoris and H. Stocker, Phys. Lett. B162, 47 (1985). G.F. Bertsch, W.G. Lynch, and M.B. Tsang, Phys. Lett. B189, 384 (1987). G. Peilert et al., Phys. Rev. C39, 1402 (1989). C.A. Ogilvie et al., Phys. Rev. C42, R10 (1990). V. de la Mota et al., Phys. Rev. C46, 677 (1992). D. Krofcheck et al., Phys. Rev. Lett. 63, 2028 (1989). C.A. Ogilvie et al., Phys. Rev. C40, 2592 (1989). J.P. Sullivan et al., Phys. Lett. B249, 8 (1990). W.M. Zhang et al., Phys. Rev. C42, R491 (1990). D. Krofcheck et al., Phys. Rev. C43, 350 (1991). J. Peter, Nucl. Phys. A545, 173c (1992). G.D. Westfall et al., Phys. Rev. Lett.71, 1986 (1993). W.Q. Shen et al., Nucl. Phys.A551, 333 (1993). J. Lauret et al., Phys. Lett. B339, 22 (1994). Buta et al. Nucl. Phys.A584, 397 (1995). R. Pak et al., Phys. Rev. C53, R1469 (1996). R. Pak et al., Phys. Rev. Lett.78,1022 (1997). R. Pak et al., Phys. Rev. Lett.78,1026 (1997). D.J. Magestro et al., Phys. Rev. C61,021602(R) (2000). S. Soff et al., Phys. Rev. C51, 3320 (1995). R. Pak et al., Phys. Rev. C54, 2457 (1996). D.J. Magestro, W. Bauer, and G.D. Westfall, MSUCL-1160 (2000). Frank Daffin and Wolfgang Bauer, MSU Preprint, MSUCL-1111 (1998). Bao-An Li and S.J. Yennello, Phys. Rev. C52, R1746 (1995). Bao-An Li et al., Phys. Rev. Lett. 76, 4492 (1996). K.C. Chung, C.S. Wang, and A.J. Santiago, Phys. Rev. C59, 714 (1999). W.D. Myers and W.J. Swiatecki, Phys. Rev. C57, 3020 (1998). P. Danielewicz et al., Phys. Rev. Lett. 81, 2438 (1998). Pinkenburg et al., Phys. Rev. Lett.83, 1295 (1999). Y.M. Zheng, CM. Ko, B.A. Li, and B. Zhang, Phys. Rev. Let. 83, 2534 (1999). D.J. Magestro et al., MSU Preprint MSUCL-1164 (2000).
The Backtracing Procedure in Nuclear Physics P. Desesquelles Institut de Physique Nucleaire, 15 rue G. Clemenceau, F91406 Orsay Cedex,
France.
Simulation on large computers (Monte Carlo sampling) occupies an increasingly important place between theory and experiment. This contribution presents a global protocol for the comparison of model simulations with experimental results, which includes a recursive inversion procedure and a multivariate analysis techniques used to condensate the experimental information on a minimum number of variables. This protocol is relevant in all fields of physics dealing with event generators and multiparametric experiments. Applications from the past few years of this technique will be presented
1 1.1
Analysis Protocol Backtracing
The protocol does not restrict the analysis to some selected features of the experimental observables but seeks to compare the global feature of the experimental results with the theoretical predictions. In other words a model will be considered to be compatible with the experiment if the model and the experiment shows the same correlation between all observables. At a first level, this means, that the correlated probability density function (CPDF) is such that, for all values of the observables : Oexp(o) = Omodei{o)
(1)
where o is the vector of all the experimental observables and Oexp and Omodei the CPDFs. It is clear that Omodei depends on the values of the input variables (referred to in the following as source variables) of the particular code. For example, the multiplicity (observable variable) will reach higher values for higher excitation energy (source variable) and larger excitation energy distributions lead to larger multiplicity distributions. In statistical terms, the observable CPDF is connected to S(s), the source CPDF, via the conditional CPDF C(o|s) of the observables given the source variables by : 0 ( o ) = f S(s)C{o\s)
ds
"Permanent address : ISN, 53 av. des Martyrs, F38026 Grenoble Cedex, France.
359
(2)
360 T h e goal is to find t h e C P D F of the source variables (if existing) t h a t p e r m i t s the model code to reproduce t h e experimental observables. This is done automatically using a proper backtracing a l g o r i t h m 1 . 1.2
Optimum
Observables
W h e n the number of observables is too large (for example when the m a t r i x corresponding to t h e C(o|s) t e r m is too large to fit the computer m e m o r y ) , t h e experimental information must be projected onto a smaller number of observables. In this case, t h e o p t i m u m set of observables (called principal variables) is given by t h e principal component analysis technique 2 ' 3 ' 4 as linear combinations of the primary observables. For a complete description of this protocol see reference 1 . It is embodied into two codes : RCN (backtracing) a n d P C A (principal component analysis). 2
Application to Nuclear Experiment Analysis
Many results were obtained in t h e last few years using this m e t h o d . One is presented in t h e following as a n example. In this experiment (Au (600 M e V / n u c l e o n ) + C u ) , performed at GSI using the ALADIN spectrometer, the observables are the variables extracted from t h e charge partitions of 30 000 e v e n t s 5 . All fragments emitted from t h e quasiprojectile nucleus are detected and identified, a p a r t from the lightest ones which are removed from the analysis. T h e experimental uncertainties a n d rejections are taken into account in the simulations. These hot nuclei are described by two source variables : excitation energy (E*) and mass (A). It is intuitively obvious t h a t E* and A are correlated quantities. Very peripheral interactions of the gold projectile with t h e copper target lead to low energy quasi-projectiles with a mass close to AAU = 197 whereas deeper interactions leave a quasi-projectile with a higher excitation energy b u t a lower number of nucleons. T h e two source CPDFs given by the backtracing using the first four principal variables (which contain more t h a n 97% of the total experimental statistical information) for the two models are given in figure 1. In the case of the SMM model, the backtracing procedure has converged toward a n exact solution (upper figure) but no physical solution was found in the case of the GEMINI code (lower figure).
361
9- 225 < 200 175 150 125 100 75 50 25 0
\
~ =
i
\ i
() e
200
< 175
(a)
M
z~
i 1iii1iii1iirliii1ii i
2
4
6 8 10 12 E* (MeV/nucleon) (b), m
150 125 100
i
75 50 25 0
fin i i i i l i i i i l i i i i l i i i i l i i i i l i i i i l i i i i l i i i i l i i n
0 1 2 3 4 5 6 7 8 9 10 E* (MeV/nucleon)
Figure 1: Correlation of the two source variables (A mass of the parent nucleus, E* excitation energy), obtained by backtracing on Au (600 MeV/nucleon)+Cu ALADIN experiment, a) SMM model. The thin line represents the mean correlation assuming a participant/spectator scenario, b) GEMINI model.
The optimum source CPDFs in the latter case can be understood in the following way : the GEMINI code is based on an evaporative model which generates a high mean number of light particles. Thus, to reproduce the same number of intermediate mass fragments as in the experiment, the code must also produce a large number of light particles. Hence, the total sum of the charges is higher than the charge of the experimental quasi-projectile, and, as the total multiplicity is higher, it also requires high excitation energy. This leads to the up and right saturations seen the in figure. In the case of the SMM code, the [A, E*] correlation has freely converged towards the correlation given by the theoretical models 6 (thin line). All experimental observable distribu-
362 tions are perfectly reproduced by the model. Hence the experimental results are in agreement with the statistical theory of simultaneous multifragmentation, and, a fortiori, with the hypothesis of thermal equilibrium of the parent nuclei. Other analyses performed with this protocol include : the determination of the correlation between the excitation energy and the numbers of pre- and postscission neutrons in slow fission7, the correlation between the mass, the excitation energy and the size of the multifragmenting nuclei in AU(35J4 MeV)+Au collisions8, the determination of the bombarding energy threshold for instantaneous multifragmentation 9 and the evolution of the quasi-fusion cross section for the Ni+Ni system 10 . 3
Conclusion
A global protocol for comparing simulation codes with experimental samples of data has been defined. The central step consists in finding the multivariate distribution of source variables which reproduces the multivariate distribution of the observables through a suitable model. This protocol minimizes the number of arbitrary choices necessary for the comparison in the sense that the adjustement procedure and the choice of the variables are fixed. The use of the principal component analysis principal variables permits to condense the maximum experimental information on a minimum number of variables. This protocol has been successfully applied in different fields of Nuclear Physics. References 1. P. Desesquelles, J.P. Bondorf, I.N. Mishustin and A.S. Botvina, NPA 604, 183 (1996) 2. A.M. Kshirsagar, "Multivariate Analysis", Dekker, Inc., New York (1972) 3. E.Lloyd, "Handbook of Applicable Mathematics", Volume VI, "Statistics", part B, Wiley and Sons (1984). 4. P. Desesquelles, Ann. Phys. Fr. 20, 1(1995). 5. J. Hubele et a/., ZPA 340, 263 (1991) 6. A.S. Botvina and I.N. Mishustin, PLB 294, 23 (1992), Bao-An Li et al., PLB 303, 225 (1993), H.W. Barz et al, NPA 561, 466 (1993) 7. L. Donadille et al, NPA 656, 259 (1999) 8. P. Desesquelles et al., Nucl. Phys. A 633, 547 (1998) 9. M. Charvet, PhD thesis, Grenoble (1997). 10. P. Desesquelles et al, Phys. Rev. C 62, 24614 (2000)
N I + N I COLLISIONS AT 32 M E V / U : E X P E R I M E N T A L INSIGHT WITH THE INDRA MULTIDETECTOR A. M. MASKAY, P. LAUTESSE, P. DESESQUELLES, E. GERLIC, J.L. LAVILLE. AND THE INDRA COLLABORATION* 1
Introduction
Reaction processes around the Fermi energy, implying the formation and decay of a hot intermediate subsystem, constitute a non negligible contribution to the cross-section in central collisions between heavy ions 1>2 . The formation of hot, metastable subsystems has recently been evidenced in various experimental data l i 3 , 4 . We present in this contribution data on Ni + Ni collisions at 32 MeV/u recorded with the INDRA 5 multidetector. These data have been analyzed using a Discriminant Analysis(DA) procedure. 2
Evolution of reaction mechanisms as a function of t h e discriminant variable
The new (DA)method used to separate a one-source component from other contributions is explicited in refs l'6. It is performed here for the set of complete events(requesting events detected with at least 80% of the initial total charge and linear momentum) and leads to a discriminant variable - called ^625 - which was shown to provide the best separation between one-source and poly-source contributions. On fig. l(top) the D625 range is divided in four intervals, going from the pure one-source(interval al) pattern until the pure binary-like one(dl) . Intervals b l and cl correspond to an intermediate situation. Below, various plots illustrate these four situations : contour velocity and charge versus parallel velocity plots, both in the center-of-mass(c.m.) system. The velocity plots depart from the binary scenario(d2) until the fusion-like one(a2) : projectile-like and target-like components get closer and closer until they fuse. The charge/velocity diagrams exhibit a similar behaviour evolving until the fusion scenario which is characterized by a maximum charge emission at velocity ~ 0 in the cm.. This confirms the efficiency of the DA method to disentangle various reaction mechanisms and to identify a one-source emission . In the following, we investigate the fusion-like process (fig.lal,a2,a3) and 'see contribution of M.F. Rivet for full list of collaborators
363
364
bi A
a2
b2
c2
d2
c3
d3
f(v par )
MAM
Figure 1. D625 distribution according to four ( a l , b l , c l , d l ) intervals(top) and typical corresponding plots ; middle : velocity plots(a2,b2,c2,d2) ; bottom : charge versus parallel velocity plots (a3,b3,c3,d3)
the formation of a highly excited metastable subsystem at thermal equilibrium followed by its disassembly. 3
One-Source properties
A way to confirm thermal equilibrium is to check the isotropy and maxwellian character of the energy spectra of the decay particles. We see in fig. l-a2 that the velocity plot accounts for an isotropic emission of decay particles and fragments . The energy spectra of these particles are displayed on fig. 2 for medium polar angles in the c.m.(ranging between 60° and 120°). These limits have been chosen to exclude pre-equilibrium contamination. The energy spectra of protons, deuterons, tritons and a particles exhibit similar slopes, very well reproduced by the statistical SIMON code 7 . The excitation energy - estimated by calorimetry after pre-equilibrium subtraction - is found to be about 5 MeV/u 1. These features confort the hypothesis of a thermally equilibrated hot source which is highly excited : it even reaches the multifragmentation threshold which is estimated to lie around this value of 5 MeV/u 8 . We shall now examine its decay modes. We notice (fig. I,a3) the peculiar two component charge partition of the pure one-source decay : a heavy one(15
365
Center-of-mass energy of particles(protons)
Center-of-mass energy of particles(alphas)
Figure 2. Center-of-mass energy spectra of particles emitted at transverse angles. T h e dashed histogram is the SIMON code.
"classical" evaporative mode of a hot nucleus decay which lets in the final state a cold residue plus IMF's and particles. However the situation is more complex if one considers the asymmetry distribution(fig. 3). The asymmetry parameter Asyml23 is defined by : Asymi23
=
YsiZmaxi-
< Z > ) 2 ,/(v / 6<2>)
»=i
where Zmaxi(i=l, 2, 3) are the three heaviest products of an event (Zmaxl>Zmax2>Zmax3) and < Z > their mean value. It quantifies the partition as expressed by the sizes of these decay products. The SIMON code which describes the one-source emission in the framework of sequential decay fails to describe the major part of that distribution. It reproduces only asymmetries close to 1, i.e. those which should correspond to the "evaporative" class of events (see fig. I,a3). The dashed histogram in fig. 3 represents the contribution of an equiprobable mass partition of the 116 nucleons (Ni+Ni system at 32 MeV/u), where momenta and energies are randomly distributed according to the fundamental conservation laws. The experimental partition departs also from this calculation, ruling out a possible trivial peaking simply due to some random combination effect. It then clearly turns out that most events agree(see fig. 3) with the SMM code 1)9 , which supposes simultaneous decay of a hot nucleus.
366
4
Conclusion
The DA procedure is very efficient to disentangle one-source events from the contribution of binary-like dissipative collisions. This one-source fusion-like nucleus seems to decay via two mechanisms : Besides the "classical" evaporative sequential one , the abundant IMF experimental production agrees with the multifragmentation SMM prediction. Such a co-existence is really not surprising in this transition region, close to the onset (~ 5 MeV/u) of multifragment emission 8 .
Figure 3. Charge Assymmetry Asyml23:black points are the data.The model distributions SMM(continuous histogram),SIMON (dash-dotted histogram) and random mass partition (see text) are arbitrarily normalized to the maxima of the experimental points. Asym1,2,3
References 1. 2. 3. 4. 5. 6. 7. 8. 9.
A. M. Maskay thesis ,University of Lyon (France),1999. N Marie and INDRA coll. , Phys. Lett. B391 (1997)15 . E. Colin et al. PRC C57 (1998)1032 , and refs therein . M. F. Rivet and INDRA coll, Phys. Lett. B430 (1998)217. J. Pouthas and INDRA coll, Nucl. Instr. Meth.A357(1995)418. P. Desesquelles et al., to be published in Phys. Rev. C . D. Durand Nucl. Phys. A 5 4 1 (1992)266. L. Beaulieu et al, Phys. Rev. Lett. 84, 5971, 2000, and refs therein. J. P. Bondorf et al.Phys. Rep. , 257 (1995)133.
TWO-FRAGMENT CORRELATION FUNCTIONS FOR QUASI-PROJECTILE A N D MID-RAPIDITY EMISSION Z H I - Y O N G H E , R . R O Y , L. G I N G R A S , Y. L A R O C H E L L E , D . O U E R D A N E , L. B E A U L I E U , P. G A G N E , X I N G Q I A N , C. S T - P I E R R E Laboratoire
de Physique
Nucleaire, Quebec,
Departement Canada GlK
de Physique, 7P4
Universite
Laval,
G. C. B A L L , A N D D . H O R N AECL
Research,
Chalk River Laboratories,
Ontario,
Canada
KOJ
UO
We have measured two-fragment correlation functions of quasi-projectile sources (QP) and mid-rapidity component formed in 5 8 Ni + 1 9 7 Au at 34.5 MeV/nucleon. The correlation functions of mid-rapidity component change very little with the excitation energy of the QP source. The QP source emission time, extracted from correlation function, decreases monotonically with the excitation energy in the range of 2-6 A MeV from 550 fm/c to 150 fm/c. Above excitation energy of 6 A MeV, it becomes shorter and constant, suggesting that prompt multifragmentation occurs in these QP sources.
1
Introduction
A major objective of current intermediate energy heavy ion collisions is to probe nuclear liquid-gas phase transition and multifragmentation . Experiments have been performed t o try t o find some signs and experimental evidence. Recently a signal for transition from surface to bulk emission expected for spinodal decomposition was reported for the equilibrium-like sources formed in hadron-induced collision, by studying the emission t i m e as a function of the excitation energy 1 . In the past decade, fragment emission time has been extracted in heavy ion collisions from two-fragment correlation functions for various projectile-target combinations in a wide bombarding energy range 2>3>4>5>6 . T h e transition from long to short emission times has been deduced as a function of bombarding energy 2 and excitation energy which was estimated from projectile-target combination and bombarding energy 6 . Since fragments in heavy ion collisions m a y come from various sources, the source selection and source identification are i m p o r t a n t . W i t h o u t impact parameter neither source selection, only emission time averaged over all sources were deduced for these measurements. In this contribution, we report on a study of fragment emission time from well defined Q P sources in 5 8 N i + 1 9 7 A u at 34.5 MeV/nucleon.
367
368
J..U
+
1.4 1.2 1 0.8 0.6 0.4 0.2 n
mid-rapidity Figure 1: Two-fragment correlation function, integrated over all fragment pairs with element jTT A ^
• E-Qp/A=7-8 MeV O E'QP/A=5-6 MeV •
number 6>7>3 from mid-rapidity component, for
58
Ni + 197 Au collisions at E/A = 34.5 MeV.
E* QP /A=0-2 MeV
0 5 10 IS 20 25 30 35 40 45 V red d0- 3 c)
2
Correlation function of mid-rapidity emission
T h e experiment has been performed at T A S C C , Chalk River Laboratories, with a b e a m of 5 8 Ni at 34.5 MeV/nucleon on a 1 9 7 Au target. T h e charged particles were detected in t h e CRL-Laval 4TT array constituted by 144 detectors set in ten rings covering polar angles between 3.3° and 140°. Details on detectors and their calibration can be found in Refs. 7 , s . To reconstruct t h e Q P source, t h e heaviest fragment with Z > 8 in the event was used as t h e Q P evaporation residue. A detailed description of reconstruction of the Q P source can be found in ref. 9 . Then the rest of the particles for t h e system were attributed t o t h e mid-rapidity and Q T emission. Because of high threshold of detectors, these particles were mainly from t h e mid-rapidity components. Fig. 1 shows t h e reduced-velocity (Vred = |Vi - V2\/{Z1 + Z2)1/2) correlation functions for fragment pairs selected from mid-rapidity component for EQP/A=7-8, 5-6 a n d 0-2 MeV. Yield suppression at low Vred is due t o the Coulomb interaction between fragments. A compact source t h a t quickly emits fragments results in larger Coulomb interactions between t h e fragments t h a n a larger source t h a t emits particles more slowly. Coulomb suppression at low Vred changes very slightly with E g P / A , indicating t h a t emission time of mid-rapidity component is independent of t h e Q P excitation energy. 3
Emission time of quasi-projectile source
In contrast with t h e mid-rapidity component, yield suppression at low VTed in t h e correlation function of Q P fragments increases obviously with E * / A , suggesting t h a t t h e emission times decrease with t h e increasing excitation energy. To extract t h e emission times of Q P source, t h e experimental correlation functions are compared with t h e simulation of t h e N-body Coulomb trajectory code of Glasmacher et al. 1 , 1 ° . This code considers t h e fragments
369
Figure 2. As in Fig. 1, but for all fragment pairs from QP source, gated on excitation energy ( E Q P ) . The solid, dashed, and dot-dashed curves represent the calculated correlation functions for fit parameters indicated on the figure. to be emitted from the surface of the source. T h e fragment emission times t were assumed to have the probability distribution P(t) ~ e~t'T, where r is the emission t i m e of the source. In Fig. 2 we show fits to the correlation functions of Q P source for three bins in E Q P / A for a range of nuclear density p and emission time r . T h e left panel shows correlation functions for the lowest bin E Q P / A = 2 - 4 MeV. A very long emission time of about 550 f m / c was extracted, indicating t h a t the Q P source emits fragment by sequential binary disassembly. T h e middle and right panels show correlation functions for the higher two bins at 6-7 MeV a n d 7-8 MeV, respectively. At high excitation energy, the emission t i m e of Q P source becomes very short ( ~ 100 f m / c ) . T h e observed short r values for high excitation energies are consistent with the values ( r ~ 100 fm/c) predicted for multifragmentation decay which originates from bulk instabilities of nuclear m a t t e r at low density. T h e full circles in Fig. 3 present the averaged emission times for the Q P source formed in Ni + Au collision at 34.5 MeV/nucleon as a function of the excitation energy. A clear evolution of emission time from long to short values with excitation energy is observed. Above excitation energy of 6 MeV, the emission time becomes very short and constant. Such a short emission t i m e is suggested as the evidence of p r o m p t multifragmentation of Q P source. To compare with previous studies on emission time, t h e solid squares in the Fig.3 show the emission times for the Q T source formed in Ar + Au collision 11 > 12 ) while the open squares summarize the results for the mixed source of Q P and mid-rapidity contribution formed in Ar + Au, Ne + Au,
370
i I !
i l p| fli if! 1*1$;
0 QP source (this work) D mixed sources • QT source for Ar + Au % *-&^ll&'& O hadron-induced » cCisions
900 800 700 600 500 L 400 j300 200 100 n
!
of E /A. The shaded area indicates the range of possible space-time solutions.
|m\ i'^a\
lik-iKKJf----.,..
i i i i i
0
Figure 3: Emission time as a function
1
11,,, .TW^Pj^aSsMiBJOliilliMj, i
2
3
4
5 6 7 8 9 10 E*/A (MeV/nucleon)
and Kr + N b collisions 2>3>4>6. T h e open circles in the Fig.3 come from a recent study on heavy equilibrium-like sources formed in it~ and p + 1 9 7 A u reaction in which the excitation energy was deduced by a calorimetry m e t h o d 1 . T h e emission t i m e extracted for heavy equilibrium-like sources from hadron-induced reactions are systematically lower at low E * / A compared t o the present work. T h e deduced emission times s a t u r a t e around 5-6A MeV in the case of ref. 1 , while, in the present case, the saturation occurs at about 6A MeV for the Q P source. T h e reaction mechanism might affect the emission t i m e since in hadron-induced reactions there is very little rotation, deformation or expansion involved. T h e authors would like to t h a n k T h o m a s Glasmacher for providing his N-body Coulomb trajectory code. This work was supported in part by the N a t u r a l Sciences and Engineering Research Council of C a n a d a and t h e Fonds pour la Formation de Chercheurs et l'Aide a la Recherche du Quebec. References 1. 2. 3. 4. 5. 6. 7. 8. 9.
L. Beaulieu et al., Phys. Rev. Lett. 84, 5971 (2000). E. Bauge et al., Phys. Rev. Lett. 70 (1993) 3705. Y. D. Kim et al., Phys. Rev. Lett. 67 (1991) 14; Phys. Rev. C 45 (1992) 338. D. Fox et al., Phys. Rev. C 47 (1993) R421; Phys. Rev. C 50 (1994) 2424. Z. Y. He et al., Nucl. Phys. A 620 (1997) 214; Phys. Rev. C 57 (1998) 1824. D. Durand, Nucl. Phys. A 630 (1998) 52c, and references therein. Y. Larochelle et al., Nucl. Inst, and Meth. A 348 (1994) 167; L. Gingras, Master's thesis, UniversiteLaval (1998) (unpublished) L. Gingras et al., XXXVI International Winter Meeting on Nuclear Physics, Bormio (Italy), (1998) p. 365. 10. T. Glasmacher et al., Phys. Rev. C 50 (1994) 952. 11. H. - Y. Wu et al., Phys. Rev. C 57 (1998) 3178. 12. M. Louvel et al., Phys. Lett. B 320 (1994) 221.
F E R M I O N I N T E R F E R O M E T R Y IN N I - I N D U C E D R E A C T I O N S AT E / A - 45 M E V ROBERTA GHETTI Department
of Physics, University of Lund, Box 118, S-221 00 Lund, E-mail: [email protected]
Sweden
Neutron-neutron, proton-proton and neutron-proton correlation functions have been measured simultaneously for the E/A = 45 MeV 5 8 Ni + 27 A1 reaction. From data analysis and model comparison, the space-time extent of the emitting source has been deduced and information on the competition between dynamical and statistical emission mechanisms has been gained.
Within the CHIC Collaboration program of fermion interferometry, an experimental investigation of 45 A MeV 58 Ni-induced reactions has been performed at the superconducting cyclotron of Laboratori Nazionali del Sud in Catania, with the aim of studying the space-time evolution of the reaction zone in heavy ion collisions 1 ' 2,3 . With an experimental apparatus consisting of 48 liquid scintillator neutron detectors, 13 CsI(Tl) proton detectors and 36 phoswich detectors for charged fragments identification, neutron-neutron (nn), proton-proton (pp) and neutron-proton (np) correlation functions were measured simultaneously and in coincidence with forward emitted fragments, used to characterize the class of collisions under study. The experimental correlation function, constructed by dividing the coincidence yield by the yield for uncorrelated events (generated from the product of single particle distributions) is: C(q,P;ot)=K."cf_P^\, Nnc(q,Ptot)
(1)
where the relative momentum q = (p[—p2)/2 and the total momentum Ptot = pi + p~2 of the particle pair are introduced. The normalization constant K is determined so that the correlation function goes to unity at large values of q where no correlations are expected (in our case q > 40 MeV/c). The integrated nn (corrected for cross talk between neighboring detectors 6 ), pp and np correlation functions are shown in Fig. 1. The data have been sorted with a "minimum bias" requirement that at least one fragment detected in the forward wall should be in coincidence with the nucleon pair. This condition greatly reduces the amount of background in-scattering events while simulations performed with the Nuclear Molecular Dynamics code of 371
372
-9 o 53 c
2 1
Is . . . .
nn
. . i . . . . i . . . . i
1 .
c 1.2 o
. . . .
i . . . .
y ^ * ^ - ^ ^
a)
i . . . .
t . . .
b
PP
,
)
I 0.8 o o 0.6
-
.
,
/
.
!
•
.
,
.
1 ,
.
.
.
L ,
.
•
.
1 .
.
.
.
1 .
.
.
I
,
np
X
1.5
,
.
,
.
,
,
c)
1 ".
0
, . . 1
.
10
.
.
.
1
.
20
.
.
1 ,
30
40
,
.
,
1 .
50
,
,
.
1 .
60
,
.
.
I
,
70
,
80
Relative Momentum, q (MeV/c) Figure 1. From the E / A = 45 MeV 5 8 Ni + 2 7 A1 reaction, nn, pp and np experimental correlation functions (symbols). The curves represent the theoretical correlation functions predicted by the statistical model.
ref.7 indicate that it does not affect the impact parameter selection in the case of the Ni + Al reverse kinematics reaction. The overall shape of the three correlation functions is qualitatively as expected from the interplay of quantum symmetrization effects and final state Coulomb and nuclear interactions and the correlation functions appear to be consistent with each-other. In particular the strength of the np correlation function is about half the nn strength, as expected from isospin considerations 4 . The three correlation functions can be reproduced by calculations performed with the statistical model of ref.1'5 (curves in Fig. 1). By fitting simultaneously the correlation functions as well as the single particle proton and neutron energy distributions, a good characterization of the emitting source in terms of average Gaussian radius, emission lifetime, initial temperature and source velocity can be achieved.
373
0
10
20
30
40
50
60
70
80
Relative Momentum, q (MeV/c) Figure 2. Comparison of the experimental integrated (black dots) and high-momentumgated (open squares) correlation functions with the theoretical integrated (solid curves) and high-momentum-gated (dashed curves) correlation functions predicted in the framework of the "two-sources" statistical model discussed in the text.
The calculations with the statistical model, however, are not able to reproduce a very important feature observed in the data, namely the increased strength that all three correlation functions exhibit when cuts on the the total momentum of the particle pairs (calculated in the frame of the moving source) are applied. This experimental effect is shown by the open squares in Fig. 2, representing the correlation functions obtained when gating on high values the total momentum of the pairs (Ptot > 270 MeV/c for nn and np, Vtot > 300MeV/cforpp). This shortcoming of the statistical model is not unexpected as the hightotal-momentum enhancement of the correlations is due to highly energetic pre-equilibrium particles emitted on a very short time scale and preequilibrium effects are not included in the statistical model. The disagreement between data and calculations can be cured by adding emission from a "pre-equilibrium" source, characterized by a high initial temperature and a
374
very short lifetime. The same "two-source" parameter set can be utilized to describe both proton and neutron emission if we allow for a different fraction of pre-equilibrium to equilibrium particles for protons and neutrons respectively. The comparison of the "two-source" calculations with the experimental data is shown in Fig. 2. Including the "pre-equilibrium" source enhances the early-time emission, while the long time tail is consistent with exponential emission lifetime values of TP W 400 fm/c for protons and r n « 600 fm/c for neutrons. The "two-source" model predicts an average spatial emission region with Gaussian radius RG = 3±1 fm and an initial temperature To = 11 ±2 MeV cooling in time. The emission time distributions for protons and neutrons appear to be different, suggesting that the ratio of pre-equilibrium to evaporative particles is larger for protons than for neutrons. This result indicates a different shape of the emission time distributions for protons and neutrons but does not say anything on the relative amount of emitted preequilibrium protons to neutrons. Model-independent information on the emission time can be achieved by applying velocity selections on the np correlation function8. Our preliminary results from this method, which consists in comparing the correlation function for pairs where the proton is faster than the neutron with the correlation function for pairs where the neutron is faster than the proton, indicate that neutrons are in average emitted earlier than protons. A similar effect is predicted by Boltzmann-Nordheim-Vlasov calculations with an equation of state for asymmetric nuclear matter 9 . Acknowledgments Financial support from the Swedish Natural Science Research Council (NFR) is gratefully acknowledged. References 1. 2. 3. 4. 5. 6. 7. 8. 9.
R. Ghetti et al, Nucl. Phys. A 660, 20 (1999). R. Ghetti et al, Nucl. Phys. 674, 277 (2000). R. Ghetti et al, Phys. Rev. C , in press (Sept. 2000). G. Bertsch et al, Europhys. Lett. 21, 817 (1993). J. Helgesson et al, Phys. Rev. C 56, 2626 (1997). R. Ghetti et al, Nucl. Instrum. Methods A 421, 542 (1999). J.P. Bondorf et al, Nucl. Phys. A 624, 706 (1997). R. Lednicky et al, PLB 373, 30 (1996). M. DiToro, private communication.
INVESTIGATION OF A N A N G U L A R D I S T R I B U T I O N OF P R O T O N S IN P E R I P H E R A L A N D C E N T R A L N U C L E U S - N U C L E U S COLLISIONS AT THE M O M E N T U M OF 4.2 A G E V / C
M. K. SULEYMANOV,0. B. ABDINOV, Z. YA. SADIGOV Institute of Physics, Academy of Sciences of Azerbaijan Republic, Baku E-mail: [email protected] N. S. ANGELOV,A. C. VODOPYANOV, A. A. KUZNETSOV JINR, Dubna, Moscow Region, 141980, Russia The experimental results on the relation between the number of events, the angular distributions of protons and full number of protons are presented for 12CCinteractions at the momentum of 4.2 A GeV/c. The influence of nuclear fragmentation process on the results is also considered.The obtained results confirm the assumption that there exist the critical phenomena among the central collisions and it is necessary to use a percolation approach for the full description of the central collisions.
1
Introduction
At present there are many papers in which the processes of nuclear fragmenatation 1 and the processes of a total disintegration of nuclei 2 ( or central collisions) are considered as the critical phenomena and for their description a percolation approach was proposed. Some experimental results obtained in the region of high energy 3 - 5 clearly demonstrate the existence of the regime change points in the behaviour of different characteristics of secondary particles and the events depending on centrality degree of collisions (in different papers the number of all protons 4 , the number of multicharged fragments 3 and other parameters were considered for the latter). We believe that if the observed regime change points are connected with the appearance and the decay of percolation claster then with the increase of the centrality degree of collisions the behaviour of the secondary particles and the events characteristics depending on centrality degree of collisions must also depend on the number of fragments in the event, as the percolation claster could be a source of nuclear fragments. The purpose of our investigation is to test this idea experimentally. 375
376
2
Experiment
The experimental data have been obtained from the 2-m propane bubble chamber of LHE, JINR. We used 20407 12CC- interactions at the momentum of 4.2 A GeV/c ( for methodical details see 6 ) . To reach the purpose we investigated a number of the events depending on the variable Q. To determine the values of Q two variants were considered. In the first variant the values of Q were determined as Q = n„+ + Np — nv-. Here nK+,nK- and Np are the number of identified 7r+ -, ir~ - mesons and protons respectively (in figures these points are denoted as empty starlet).In that determination Q is a number of all the protons in an event without taking into account a remainder of nuclei. In the second variant the values of Q were determined as Q = N+ —n„-. Here N+ are charges of all the positively charged particales in an event including nuclear fragments (in figures these points are denoted as full starlet). In that determination Q is a summary charge of an event. 3 3.1
Experimental results. Q-dependence of the events number.
The distributions of the events depending on Q are shown in fig. la,b. It is seen that with the enclusion of fragments number to determine Q the form of distributions sharply changes and has a two-steps structure(fig. la, full starlet). In fig. lb are shown the Q-dependences of the events for the calculation data obtained from the quark-gluon string model 7 (QGSM) without nuclear fragments. The empty starlets correspond to the cases when the stripping protons were not taken into account and the full starlets correspond to the cases when the stripping protons were included. It is seen that the form of the distribution strongly differs from the experimetal one in fig. la. There is no the two-steps structure in this figure. Therefore we can assert that this difference is connected with the existence of fragments in 12CC - interactions. Thus, the results demonstrate that the influence of nuclear fragmentation process in the behaviour of different characteristics of the events depending on Q has a critical character. We suppose this result to be connected with percolation clusters. 3.2
Q-dependence of the protons angular distributions.
To confirm the existence of percolation claster we analysed the angular spectrums of identified protons depending on Q and the number of fragments. We
377
Figure 1: Q-dependence of the number of events.
have obtained Ni = _ j '—- = /(cosfy) (here JV/ are a number of the protons at an emission angle of #,- and J is a total number of the protons in an event) for the following groups of events depending on Q: Q < 5(this is peripheral collisions-ATi); Q = 6 - 7(N2); 8 - 9(AT3); > 10(this is central collisions-AT4). To investigate Q-dependences of ATj-functions we used the following quan; tities: h = %5$ ; h = Tf^k f* = ifc^k- T o investigate the values of /, depending on the number of fragments we also used(as in (3.1)) two variants to determine the variable Q. Fig.2 shows the Q-dependences of / i , f2 and fz as a function of cos#;. As well as in fig. 1 the empty starlet corrisponds to the cases when the nuclear fragments were not taken into account and the full starlet - the cases when the nuclear fragments were included. It is seen that these distributions difffer for the two different variants of Q-determination. We see the increase of / j with the increase of cos 0,- i.e. an additional production of protons in this interval. We suppose the percolation claster to be a source of additional protons in this interval. 4
Summary
For 12 CC-interaction the behaviour of the number of events, depending on Q also depends on the number of fragments and has a two-steps form. This form is not reproduced by the calculated data in the framework of the QGS model which does not take into account nuclear fragments. This result as well as the results obtained from the analisys of angular distributions of protons in peripheral and central collisions could be a confirmation of the existence of percolation clasters. Finally we want to say that at GSI, AGS and Nuclotron energies this result can signal about the existence of the transition of nuclear matter from nucleon states to its mixed ones. At RHIC or LHC energies , a similar result could
378 ,2
0.4 0.2 f.
0 -0.2 -0.4
CC proton o)
pr
It.l
t . . ,•
•
»* * * ;
*1t »
t"
-1
11 -0.75
, 1 , -0.5 -0.25
i. 0.25
0
0.5
. . 1 0.75
1
0.4 0.2 _
0
-0.2 -0.4
b)
! t1 1 1 1 + , • I'lll't'
• * • * • • . • * * * : . *
(
****
•
:
I _, j j , J . U , ,._!__. J_I.J-.LL j . .i J J . i iJ . , . - . - : i. , . . . I . i.i . 1 i -1 -0.75 -0.5 -0.25 0 0.25 0.5 0.75 1
0.4 0.2 f,
H,,,,t,v,i:::""
0 -0.2 -0.4
-1
-0.75
-0.5
-0.25
0
0.25
0.5
0.75
1
COST.,,
Figure 2: Q-dependence of the angular distribution of protons.
help to detect "critical" signals of phasetransition nuclear matter. References 1. J. Desbois, Nucl. Phys. A 466, 724 (1987); J. Nemeth et al., Z. Phys. A 325, 347 (1986); S. Leray et al., Nucl. Phys. A 511, 414 (1990); A.J. Santiago and K.C. Chung, J. Phys. G 16, 1483 (1990). 2. X.Campi, J. Desbois , Proc. 23 Int. Winter Meeting on Nucl.Phys; Bormio ,1985; Bauer W. et al., Nucl. Phys. A 452, 699 (1986) ;A.S. Botvina, L.V. Lanin. Sov.J. Nuc.Phys. 55, 381 (1992). 3. M.I Tretyakova. Proceedings of the Xlth International Seminar on High Energy Physics Problem, Dubna , 616, (1994). 4. M.K. Suleimanov et al., Phys. Rev. C 58, 351 (1998) 5. S.Vokal, M.Sumbera, JINR Preprint, 1-83-389, Dubna, (1983). 6. N. Akhababian et al., JINR Preprint, 1-12114, Dubna, (1979); A.I.Bondorenko et al., JINR Communication, Pl-98-292, Dubna, (1998). 7. N.S. Amelin, L.V.Bravina, Sov.J. Nuc.Phys. 51, 211 (1990); N.S. Amelin et al., Sov.J. Nuc.Phys. 50, 272 (1990)
REVERSE EXPERIMENT AT LABORATORINAZIONALI DEL SUD
E. GERACI for REVERSE collaboration* Laboratori Nazionali del Sud andDipartimento di Fisica Universita'di Catania E-mail: [email protected] *S. AIELLO(l), M. ALDERIGHI(2), A. ANZALONE(3), M. BALDO(l), R. BARNA'(4), M. BARTOLUCCI (5), I. BERCEANU (6), A. B0NASERA(3), B. BORDERIE(7), A. BOTVINA(8), R. BOUGAULT(9), M. BRUNO(8), G. CARDELLA(l), S.CAVALLARO(3), M. COLONNA(3), M. D'AGOSTINO(8), E. D'AMICO(4), R. DAYRAS(10),N. DE CESARE(11) E. DE FILIPP0(1), D. DE PASQUALE(4), M. DI TORO(3), S. FEMINO'(4), E.FUSCHINI(8), M. GERACI(l), F. GIUSTOLISI(3), A. GRZESZCZUK(12), P. GUAZZONI(5), D. GUTNET(13), M. IACONO MANNO(3), A. ITALIANO(4), S. KOLWASKI(12), G. LANZANO'(l), G. LANZALONE(l,7), N. LE NEINDRE(8,9), S. LI(14), U. LOMBARDO(3), S. LO NIGRO(l), C. MAIOLINO(3), D. MAHBOUB(3), G. MANFREDI(5), G. MARGAGLIOTTI(15), D. MOISA(6), T. PADUSZYNSKI(12), A. PAGANO(l), M. PAPA(l), M. PETROVICI(6), E. PIASECKI(16), S. PIRRONE(l), G. POLITI(l), E. POLLACCO(IO), A. POP(6), F. PORTO(3), A. RAPISARDA(l), D. M. F. RTVET(7), E. ROSATO(l 1), S. SAMBATARO(l), G. SECHI(2), V. SIMION(6), M.L. SPERDUTO(3), J. C. STECKMEYER(9), C. SUTERA(l), A. TRIFILO'(4), M.TRIMARCHI(4), G. VANNINI(7), M. VIGILANTE( 11), J. WILCZYNSKI(17), H. WU(14), Z. XIAO(14), L. ZETTA(5), W. ZIPPER(12) I)INFN and Dipartimento di Fisica Universita' di Catania 2)Istituto Fisica Cosmica, CNR and INFN Milano 3)LNS and Dipartimento di Fisica Universita' di Catania 4)INFN and Dipartimento di Fisica Universita' di Messina 5)INFN and Dipartimento di Fisica Universita' di Milano 6)Institute for Physics and Nuclear Engineering, Bucharest, Romania 7)IPN, IN2P3-CNRSand Universita'Paris-sud, Orsay, France 8)INFN and Dipartimento di Fisica Universita' di Bologna 9JLPCJSMRA University ofCaen,France 10)DAPNIA-SPhN,CEA Saclay, France 11JINFN and Dipartimento di Fisica Universita di Napoli 12) University of Silesia, Katowice, Poland 13)IPN; IN2P3-CNRS and Universite Claude Bernard, Lyon .France I4)lnstitute of Modern Physics Lanzhou, China 15)INFN and Dipartimento di Fisica Universita' di Trieste I6)lnstitute of Experimental Physics, University of Warsaw, Poland 17)Institute of Nuclear Studies Otwock-Swierk, Poland
379
380
Thefirstexperimental campaign of REVERSE experiment has been performed at Laboratori Nazionali del Sud using the forward part of CHIMERA multidetector. Tin projectiles impinging on Nickel targets were studied. Data analysis is in progress, some detectors performances are shown.
REVERSE experiment intends to study isospin effects and clusters' production in nuclear multifragmentation in the field of heavy ion collisions at intermediate energies. Features of the experiment, such as isotopic identification of the nuclear clusters and good particle-particle correlations, are important requirements for studying phase transition in finite nuclear system. The experiment has been performed using the forward part of CHIMERA multidetectorfl], recently installed in the Ciclope vacuum chamber at heavy ion facilities INFN Laboratori Nazionali del Sud, assembled with 40 additional silicon strip detectors. Beams of 112Sn and 124Sn impinging on 58'64Ni and 27A1 targets were used. The REVERSE apparatus is shown in figure 1. The main characteristics of CHIMERA multidetector are a systematic measurement of the time of flight, a low multi-hit probability due to the high adopted granularity and a very low energy threshold (E/A<0.5 MeV/nucl) for fragments detection. The forward part of CHIMERA is composed by 688 telescopes, arranged into 9 wheels, covering polar laboratory angles from 1° to 30°. Each telescope is made of a silicon detector (300 um thick) followed by a CsI(Tl) crystal (12 cm thick), coupled with a photodiode. The signals coming from the photodiode were analyzed by a twogate integration method allowing the identification of light charged particles [2],The identification procedure is illustrated in figure 2 where charges and isotopes are identified from protons to berillium. A fast movable timing device, made of a thin NE102A plastic scintillator optically coupled with a photomultiplier tube by an elliptic reflector, was placed at 20 cm from the target in order to intercept beam particles during some particular runs.
381
Figure 1: Reverse experimental apparatus An enhanced data acquisition, control and trigger system has been developed to manage up to 5000 electronic channels of the detector [3]. Moreover, an on-line computational and control system, based on a two DSP-based board, is under development [4]. Several calibration runs have been performed for both Silicon and CsI(Tl) detectors using Tandem and Superconducting Cyclotron during May-June 1999 at LNS, 'For silicon detectors an energy resolution of 0.4% was achieved for 58Ni elastically scattered at 15 MeV/A, while for CsI(Tl) detectors an energy resolution of 1.3% was obtained with the same beam. In Igure 3 a charge identification matrix for a telescope placed at the most forward angle is shown for the reaction 124Sn+64Ni at 35 MeV/A. The region close to Z=50 is dominated by a two body dissipative collision. ^ s i - s ^ i 3SA.MeV Rfcg8 tetesre]3e©M
134
Ssi f **N5 35 AJMbV R?Bg 1
ttiuuptO&E.
i
38S3
2530 3S0G 3506 CU Slow Low Gab
Figure 2. LCP identification matrix
Figure 3. Charge identification matrix
382
In order to speed the data analysis a large effort has been devoted to obtaining one-dimensional charge identification spectra using Particle Identification Function (PIF) based on AE-E method. In figure 4 PIF versus light pulse of CsI(Tl) (E signal) is shown. The projection on the PIF axis (where the contamination to each charge from the adjacent ones results smaller than 2%) illustrates the high quality of the collected data. This work has been partially supported by grants of the Italian Ministry of University and Scientific and Technological Research (MURST).
Figure 4. PIF versus CsI(Tl) signal matrix.
Figure 5. PIF mono-dimensional spectrum.
References 1. S. Aiello et al, Nucl. Phys., A583 (1995) pp 461 2. S.Aiello et al., Nucl. Inst. And Meth., A369 (1996) pp50 3. S.Aiello et al., "The enhanced data acquisition system for the 4n detector CHIMERA" IEEE Transactions on Nuclear Science, Vol. 47,No.2,(2000) pp. 114 4. S. Aiello et al. "The on-line computational and Control System for the 47i-detector CHIMERA" IEEE Transactions on Nuclear Science, Vol. 47,No.2,(2000) pp. 196
REVERSE: THE FIRST EXPERIMENT WITH THE CHIMERA DETECTOR G. POLITI FOR REVERSE COLLABORATION* LNFN andDipartimento diFisica Universita' di Catania email.pollti@ct. infn. it *S.AIELLO\ M.ALDERIGHI2, A.ANZALONE3, M.BALDO1, R.BARNA'4, M.BARTOLUCCIs, I.BERCEANUs, A.BONASERA3, B.BORDERIE7, A.BOTVINA8, R.BOUGAULT9, M.BRUNO8, G.CARDELLA1, S.CAVALLAR03,M.COLONNA3, M.D'AGOSTINO8, E.D'AMICO4, R.DAYRAS10,N.DECESARE11, E.DEFILIPPO1, D. DEPASQUALE4, M.DI TORO3, S.FEMINO,4,E.FUSCHINI8, M.GERACI1, F.GnjSTOLISI3, A.GRZESZCZUK12, P.GUAZZONI5, D.GUINET13, M.IACONOMANNO3, A.ITALIANO4, S.KOLWASKI12, G. LANZANO'1, G.LANZALONE 1 J , N.LENEINDRE8'9, S.LI14, U.LOMBARD03, S. LONIGRO1, C.MAIOLINO3, D.MAHBOUB3, G.MANFREDI5, G.MARGAGLIOTTI15, D MOISA6, T.PADUSZYNSKI12, A.PAGANO1, M.PAPA1, M.PETROVICI6, E.PIASECKI16, S.PIRRONE1, G.POLITI1, E.POLLACCO10, A.POP6, F.PORTO3, A.RAPISARDA1, M.F.RTVET7, E.ROSATO11, S.SAMBATARO1, G.SECHI2, V.SIMION6, M.L.SPERDUTO3, J.C.STECKMEYER9, C.SUTERA1, A.TRIFIRO'4,M.TRIMARCHI4, G.VANNINI7, M. VIGILANTE", J.WILCZYNSKI17, H.WU14, Z.XIAO14, L.ZETTA5, W.ZIPPER12 I) INFN and Dipartimento di Fisica Universita' di Catania 2) Istituto Fisica Cosmica, CNR and INFNMilano 3) LNS and Dipartimento di Fisica Universita' di Catania 4) INFN and Dipartimento di Fisica Universita' di Messina 5) INFN and Dipartimento di Fisica Universita' di Milano 6) Institute for Physics and Nuclear Engineering, Bucharest, Romania 7) IPN, IN2P3-CNRSand Universite'Paris-sud, Orsay, France 8) INFN and Dipartimento di Fisica Universita' di Bologna 9) LPCJSMRA University of Caen, France 10) DAPNIA-SPhN, CEA Saclay, France II) INFN and Dipartimento di Fisica Universita' di Napoli 12) University of Silesia, Katowice, Poland 13) IPN; IN2P3-CNRS and Universite Claude Bernard, Lyon .France 14) Institute of Modern Physics, Lanzhou, China 15) INFN and Dipartimento di Fisica Universita' di Trieste 16) Institute of Experimental Physics, University of Warsaw, Poland 17) Institute of Nuclear Studies Otwock-Swierk, Poland The forward part of the multidetector CHIMERA has been used for the first time at the Laboratori Nazionali del Sud in Catania, to study heavy ions multifragmentation processes in reverse kinematics.
383
384
1
The Physics of the experiment
The REVERSE experiment intends to elucidate on multifragmentation and isospin dependence of the heavy ion reactions at intermediate energy, by using a part of the CHIMERA multidetector [1,2]. The growing interest on isospin degree of freedom, due to the possible influence of the N/Z ratio on the dynamic of the collisions and on the properties of the nuclear matter, suggested the investigation of colliding systems with different isotopic ratios. Consequently, we studied u 2 Sn + 58Ni and li4 Sn + 64Ni at 35 MeV/A. The attention has been focused on both peripheral and central collisions in order to study "neck formation" and intermediate mass fragments (IMF) production. In the next future we plan to study also 124Sn + 64Ni, 2 Al at 25 MeV/A in order to investigate on the energy dependence of the reaction mechanism. 2
The experimental apparatus
All the performed reactions are in reverse kinematics, implying thus a focusing of the reaction products at the very forward angles. The device covers the laboratory angles with high granularity, in order to ensure the correct reconstruction of the events and to allow good particle-particle correlation studies. Moreover, a charge and/or mass identification of the reaction products is an imperative requirement to discriminate the isospin effects and to measure some of fundamental quantities of the nuclear system (temperature, excitation energy, mass). The experiment has been performed at the Laboratori Nazionali del Sud (LNS) in Catania, by using the forward part of the CHIMERA detector, installed in the Ciclope vacuum chamber. StTtilndtm)
ur
%
3r
%
•
^
ISItconMictar « CMlum Iodide CrjJtal
Fig. 1. Schematic section of the REVERSE device.
The device is thus composed by 688 two-step telescopes made by a silicon detector (300 u.m thick) and a CsI(Tl) crystal (12 cm) coupled to a photodiode, arranged in 9 wheels centered around the beam axis and covering the polar angles from 1° to 30°.
385
40 silicon strip detector (SiFr) are placed on a plane to detect target fragments in the angular range 30° - 90°. A schematic section and a picture of the installed device are presented in Ig. 1 and 2 respectively.
Fig. 2. The 9 wheels of the REVERSK device installed in the Cielope chamber.
Energy and time of light from the silicon detector, fast and slow components of the CsI(Tl) light emission are the four parameters held by the acquisition data system. The huge number of detectors and the good geometrical efficiency (95%), together with the isotopic identification of the reaction products achieved with the combination of the AE-E, Time of Flight and Pulse Shape discrimination [3] techniques, Mill the above mentioned requirements for the experimental device. Direct velocity measurements and low energy threshold, obtained with the TOF technique, are the two main characteristics of this apparatus. 3
The results
The device has been mounted starting from February 1999 and the first and second experimental campaigns have been performed in September 1999 and March-April 2000, together with the needed calibrations. We found good overall performances, i. e., energy resolutions on the elastic beam (Ni at 15 MeV/A) of 0.5 % for the silicon detectors and 1.5 % for the CsI(Tl). The total time offlightresolution (beam plus detector) was better than 1 nsec for most of the experiment. With these features a Ml isotopic discrimination of the more energetic Eght particles (Z <=4) has been possible with the pulse shape discrimination in the CsI(Tl), while the charge identification has been clearly achieved up to Z=50 with the AE-E technique [4]. Isotopic identification of light ions was also achieved, as shown infig.3 for the reaction l24Sn + 64Ni at 3 5 MeVA. • The fragments stopped in the silicon detector have been identified in mass with a resolution depending on the base offlight,namely on the detection angle. An example of mass discrimination obtained at 24° (distance = 120 cm) is presented infig.4; the mass identification is clearly obtained up to the carbon isotopes.
386 u
^ n + M Ni35A.MhV RingS
takKqplM
»JpJ-J-J^jjf~1-,j^
^Sto+^m
aSAJIfeV W^S
%%% aSSTawB* M " * H M im*tm
mm®$«m>i
sK^^B^aaaT^
CM Fast Low Gain
Fig. 3. AE-E matrix for light particles
4
t*»
Fig. 4. Energy -Time of Flight matrix
Conclusions
The first experiment with the CHIMERA multldetector has been successfully performed at the LNS in Catania. The forward part of the detector has been used to study multifragmentation and neck formation reactions in reverse kinematics. The performances of the device have been in good agreement with the design of the project and the preliminary results are very good. The analysis of part of the experiment is now running and we expect to contribute with new results in the heavy ion physics at medium energies in the next fiiture. This work has been partially supported by grants of the Italian Ministry of University and Scientific and Technological Research (MURST).
References 1. 2. 3. 4.
S.Aiello et al., NucL Phys. A5S3 (1995) 461. A.Pagano et al., NucL Phys. A, Proc. of 3rd CRIS 2000 Conference, in press. S.Aiello et al., NucL Instr. AndMetk A369 (1996) 50. E.Geraci et al., Proceedings of this conference.
N O N EQUILIBRATED IMF EMISSION IN HEAVY ION COLLISIONS A R O U N D T H E F E R M I E N E R G Y S.PIANTELLI, L.BIDINI, M.BINI, G.CASINI, P.R.MAURENZIG, A.OLMI, G.PASQUALI, G.POGGI, S.POGGI, A.A.STEFANINI, N.TACCETTI Dipartimento di Fisica Universita di Firenze and Sezione INFN di Firenze The first experimental data obtained with the FIASCO setup for the reaction n6 S n + 9 3 N b at 29.5AMeV confirm the existence of a midvelocity emission of LCPs and IMFs that may come from the neck rupture during the first phase of the reaction
The FIASCO experiment (Florentine Initiative After Superconducting Cyclotron Opening) studies the collision of nuclei of mass around A = 100 in the intermediate energy regime (beam energy between 17AMeV and 40AMeV). The main goal of the experiment is the study of the evolution of the reaction mechanism as a function of the beam energy, with particular focusing on the energy sharing 1 and the angular momentum transfer2 and on the investigation of non equilibrium effects. The first measurement were done in December 1998 - January 1999 at the LNS in Catania (Italy). The setup was mounted inside the CICLOPE scattering chamber. The studied reactions were 9 3 Nb+ 1 1 6 Sn at 29.5AMeV and the corresponding inverse kinematics reaction 1 1 6 Sn+ 9 3 Nb at 29.5AMeV. The symmetric reaction 9 3 Nb+ 9 3 Nb at 29.5AMeV was studied too. The experimental setup includes 24 position sensitive Parallel Plate Avalanche Counters (PPADs) for the determination of the time of flight and the impact point of the heavy fragments (A>20), 96 silicon telescopes (the A.E is 200fj,m thick and the E is 500fj,m) for the measurement of the postevaporative charge and energy of the PLF (Projectile Like Fragment) and 158 phoswich scintillators for the detection of the LCPs (Light Charge Particles with full mass resolution) and IMFs (Intermediate Mass Fragments with full charge resolution for 3 < Z ~ 20) and the measurement of their time of flight. The angular coverage of the PPADs is about 70% of the forward hemisphere and the flight path is as long as about 3.5m for the most forward PPADs. The detection threshold is about OAAMeV. The phoswich telescopes consist of an array of two or three scintillators: a 200^m thick fast plastic plus (only in the three layer version) a 5mm thick 387
388
slow plastic plus a 3 or 5cm thick CsI(Tl) crystal, coupled to a single photomultiplier through a light guide. The identification threshold is determined by the thickness of the fast plastic (about 3AMeV for protons and a particles). The event reconstruction is obtained by means of the Kinematic Coincidence Method (KCM) 3 from the PPADs data. This technique gives an estimate of the primary (pre-evaporative) quantities (velocity and mass) for the 2, 3 and 4 body events starting from the velocity vectors of the heavy fragments measured by the PPADs. In the following we report results for only the two body events. The centraJity of the collision is estimated by the Total Kinetic Energy Loss (TKEL) in the center of mass reference frame. The particle emission pattern has been studied by means of the correlation
1 protons
4 protons
2 alphas
5 alphas
II•• 6 IMFZ=3-7
3 IMFZ=3-7
100 so
-100
-50
0
50 -100 vpar [mnVns]
-50
I , i , , I0 50 vpar [mm/ns]
Figure 1. Correlation between vpeT.p and v// for protons, a particles and IMFs (3 < Z < 7) for two body events for the reaction 116Sn + 9 3 Nb at 29.5AMeV for peripheral collisions (TKEL between 200MeV and 400MeV); left side: experimental results; right side: Montecarlo statistical simulation
389 between the parallel and perpendicular velocity components of the LCPs and IMFs with respect to the estimated pre-evaporative separation axis of the heavy fragments. On the left side of fig. 1 the experimental correlations of protons, a particles and IMFs (3
>.10
I1
fio ^10
"PROTONS i
I
i
i
,
I if. I
10
5
I
210
10
5
10
impact param. (fm)
>10
H10 10
•3
1 FTRITONS
DEUTONS ,
-H -2 -3
-0=0 : * : *fi '9 midvel evap
4r
•*SgP
931
o
- J ^ - IMF (Z = 3-7) I
8
10
4
6
8
o 1000 TKEL (MeV)
n
10
impact parameter (fm) Figure 2. Right side: correlation between the TKEL estimated from the KCM and the impact parameter for the data simulated by the CHIMERA code; Left side part: evaporative and midvelocity multiplicities in function of the impact parameter for two body events for the reaction 1 1 6 S n + 9 3 N b at 29.5AMeV for protons, deuterons, tritons, a particles and IMFs
al bin (TKEL between 200AMeV and 400AMeV),
almost all the IMFs are
390
emitted at midvelocity. The observed kinematics of these emissions is compatible with dynamical processes associated to the neck formation and rupture between the colliding nuclei. We estimated the evaporative and midvelocity multiplicities as a function of the impact parameter. We extract the correspondence between the TKEL axis and the impact parameter by applying the KCM technique to simulated events generated by the QMD code CHIMERA 6 for the studied reaction; the obtained correlation is reported on the right side of fig. 2. The evaporative part has been estimated by means of a Montecarlo simulation based on the GEMINI code, which was tuned to reproduce the forward emission of the PLF. The obtained results for protons, deuterons, tritons, a particles and IMFs are reported on the left side of fig. 2. This figure shows that the midvelocity emission for Z = 1 is neutron rich, in agreement with similar findings of Plagnol et al. 6 ; in fact while the proton evaporative intensity is larger than that of deuterons and tritons, the midvelocity component is almost equal for all the three Hydrogen isotopes. As a conclusion, this study confirms the existence of a non equilibrium emission of LCPs and IMFs, that may originate from the presence of a neck like structure formed between the reseparating primary fragments. The midvelocity emission favours the neutron rich Z — 1 isotopes. The next FIASCO production runs with an improved setup will extend the investigation of this phenomenon to a wider energy region (beam energies between YlAMeV and 40AMeV), in order to deeply study the details of the effect, with particular attention to the very peripheral reactions which the apparatus is optimized for. References 1. 2. 3. 4. 5. 6.
G.Casini et al, Phys. Rev. Lett. 78, 828 (1997) G.Casini et al., Phys. Rev. Lett. 83, 2537 (1999) G.Casini et al.,Nucl. Instrum. Methods A 277, 445 (1989) R.Charity.JVucZ. Phys. A 483, 371 (1988) J.Lukasik et al.,Phys. Rev. C 55, 1906 (1997) E.Plagnol et al.,Phys. Rev. C 6 1 , 14606 (1999)
THE COMPLETE FUSION AND THE COMPETITIVE PROCESSES IN THE ^ + nC REACTION AT E(%)=20MEV/A S.PIRRGNE ', G.POLITJ ' \ G 1 A N Z A L O N E ' 4 . S.AIELLQ 1 , N.ARENA H SEB.CAVALLARO . E . G E R A O '~\ F.PORTO ~\ S.SAMBATARO °
Si /.Ji/*(2/a>stfHK« di f m t a . Vnii*Tsit& d>
Velocity, angular, mass ami charge distributions of isotopieally resolved evaporation residues from ihe 32S+12C reaction have been measured at E(32S)=20MeV/A. Complete, incomplete and direct components have been separated in the velocity spectra using kinematical analisys and deconvolution techniques. Partial and total, incomplete and complete fusion cross sections have been extracted from the obtained angular distributions. The complete fusion cross section and the deduced critical angular momentum are compared with other experimental data and the predictions of theoretical models.
1
Introduction
In previous works [1,2,3] we studied the ^Si, Hs, ,:,CI + 12C reactions at various energies around 5MeV/A in order to measure the fusion cross section and to investigate the existence of incomplete fusion component in the evaporation residue cross section. We present in fig. 1 the experimental excitation functions for ?5 C1+ I; C ^Si+'^C and ,2S+12C systems.
r
-
» n »»c
"• /
» ^r*»; t \
. *
? 002
-ISL* '
004
ism*- "s^^c
-Bass -M*Ssase
w X-
~; !!
"V 0.04
OWT')
Figure 1. Fusion cross section as a function of the inverse of the E< ,, for different systems.
The solid circles are our data, the other symbols are from other authors [1-3]. The dashed and solid curves are respectively the prevision of Matsuse [7] and Bass [S] theoretical models. We proposed thus the study of the ,2 S + llC reaction at E(32S)=20MeV/A [4,5], in order to provide for the lack of data in this energy region and to investigate about the still interesting controversial question if the fusion cross section is either limited by the properties of the compound nucleus or by entrance channel effects. 391
392 2
Experimental procedure
The experimenls were performed at the Tandem and Cyclotron accelerator facilities of the Laboratori Nazionali del Sud (LNS), Catania. The experimental apparatus [6], shown in tlg.2, consists of a sliding seal scattering chamber, 45 cm in diameter, which can be rotated around the target axis and is rigidly connected to AE-E multianode ionization chamber, with a 300fim silicon detector at the end.
Figure 2. Schematic view of the experimental appaiatus.
A time of flight telescope, consisting of two mierochanneJ-pJates as start and stop signal detectors, is coupled to the ionization chamber. The flight path is about 1m. Examples of charge and mass identification are shown in figs. 3 and 4.
EtotfWeV}
Figure 3. Charge versus mass for the detected panicles
Figure 4, Specific energy loss AE versus tolal enerw Etot
393 3
Experimental results and analysis
From the analysis of the velocity spectra performed at lower energies, we deduced that the main contributions to the evaporation cross sections originate from a complete fusion (CF) mechanism, but for all the energies incomplete fusion (1CF) components are also present. We interpreted these incomplete fusion events as generated by cluster transfer reactions [3]. At higher energies, the separation of the different reaction mechanism from the inclusive velocity spectra is possible with complex deconvolution techniques. Fig.5 shows the invariant velocity spectra of evaporation residues for the 32S + ,2 C reaction at 20 MeV/A. The grey filled areas centred at VCN cos8 L (vertical arrow) represent the contributions of complete fusion. The dot and the dash-dot curves represent respectively the incomplete fusion and the direct reactions components. The contributions of complete and incomplete fusion discriminated in this way, allow us to obtain the absolute cross sections aa reported in fig. 1) and OICFIn fig. 6 the ratio O (T /(OCT+ <%F) is presented as a function of the light partner velocity for all our data (grey points), together with the data from the Morgenstem systematic [ 12].
Vipsn 5. Invariant velocity spectra of evaporation residues.
Figure 6. Ratio accf{Ocf+ OKF) as a Junction of light partner velocity.
Finally we reported in fig. 7 the critical angular momentum extracted from the complete fusion cross section at different excitation energies, for several reactions leading to the **Ti compound nucleus. Our result are reported as grey points. In this figure the solid curve corresponds to the statistical yrast line [9], while the dashed line indicates the angular momentum at which the fission barrier vanishes as predicated by the Sierk model [ 10,11]. The trend of the experimental points seems to show a saturation at high energy, which could be consistent with the hypothesis that the limitation of fusion cross section is due to compound nucleus effects.
394 250 r * s*Q-T-"*Si aZBG*EElU a " O t 8 ^ RRASCHEB
"o
13
20
30
48
Figure 7. Excitation energy versus critical angular momentum as obtained from different system leadings to the 44Ti compound nucleus.
References 1. 2. 3. 4. 5. 6. 7. 8. 9. 10. 11. 12.
N. Arena et al., Phys. Rev. C44 (1991) 1947 N. Arena et al., Phys. Rev. C50 (1994) SSO S.Piirone et al., Phys. Rev. CSS (1997) 2482 S.Pirrone, Atti del Congresso SIF 98 (1998) 28 G.Poiiti, Amdel Congresso SIF 99 (1999) 74 F.Figuera and S.Pirrone et al., // Nuovo Cimento A104 (1991) 251 T.Matsuse et al., Phys. Rev. C26 (1982) 2338 R.Bass, Phys. Rev. Lett. 39 (1977) 265 S.M.Lee et al., Phys. Rev. Lett. 45 (1980) 165 Arnold J.Sierk, Phys. Rev. C33 (1986) 2039 SJ.Sanders, Phys. Rev. C44 (1991) 2676 H.Morgenstern et al., Phys. Rev. Lett. 52 (1984) 1104
A SIMPLE PULSE SHAPE DISCRIMINATION METHOD APPLIED TO SILICON STRIP DETECTOR J. LU\ F. AMORINP, G. CARDELLAb, A. DI PIETROa, P. FIGUERA", A. MUSUMARRA", M. PAPAb, G. PAPPALARDOac, F. RIZZO^, S. TUDISCOad "INFN-Laboratori Nazionali del Sud, Via S.Sofia 44, 95123 Catania Italy Email: [email protected] b INFN-Sezione di Catania, 95129 Catania Italy c Dipartimento di Fisica ed Astronomia, Universita di Catania, 95129 Catania Italy Centro Siciliano di Fisica Nucleate e di Struttura della Materia, 95129 Catania Italy The pulse shape discrimination technique is applied to TRASMA silicon strip detector to achieve charged particle identification. Excellent charge identification up to nickel and isotope separation to element oxygen has been achieved within an energy range from 1.5 AMeV to 20 AMeV. The most advantage is that there is no additional electronic module employed in our method with respect to the conventional TOF technique. Recently many efforts have been devoted to exploit the pulse shape discrimination (PSD) method [1-5] based on the sensitivity of the pulse shape on charge number Z and mass number A of ions stopped in silicon detectors. Even though the electronics is rather simple, a slight increase of electronic modules will introduce a huge cost for multi detector arrays. In order to extend the TRASMA [6] telescopes (Si-CsP(Tl) AE-E and PA QDC TOF combination) to discriminate \y heavy particles stopped in a 300 um V annular silicon strip detectors without any increase of electronics, a Stop pulse shape charge identification method has been investigated in the reaction of 25.7 AMeV 58 Ni beam Figure 1. Block diagram of the electronics. SSDimpinging on 3 mg/cm V and Sc silicon strip detector; PA-charge preamplifier; AMPcomposite target. The detector was spectroscopy amplifier; CFD-constant fraction irradiated from rear side and polarized with a bias just enough to discriminator; QDC-charge to digital converter; assure a perfect depletion. TDC-time to digital converter; HF-radio frequency. It is well known that operating Si-detectors with charge preamplifiers the rise time, a combination of plasma column [7] erosion time and finite charge collection time [8], becomes longer coherently for particles having large Z with rear side injection. In our case, the measured time is a superposition of rise time and flight time of charged particles from target to detector. The flight time contributes less than 20% of the time
r\
395
396
difference for two adjacent charges. So the power of charge separation is mainly determined by the rise time of the charge signal. The block scheme of the electronics is simplified in figure 1. The energy signal E was converted by a QDC, but a normal ADC can also be used to achieve the charge separation. The HF signal was extracted from the cyclotron operated in a pulse suppression mode corresponding to a time dynamics between two beam bursts of 155 ns. For heavier particles and lower energy particles, T is much longer than the time dynamics. This introduces a "wrap around" effect on E-T spectrum for events having large value of T: a shifted E-T plot superposes on the real coincidence plot and blurs higher energy region of the light charged particles. By referring to the time of prompt gamma rays registered by TRASMA BaF2 crystal arrays surrounding the target chamber, the shifted portion of the plot can be removed completely and relocated to its proper position making a 155 ns time correction.
500
7S0
1000
Energy (MeV)
;|fr:,
Wisi|g| 1§ lit ItSI J ^ ^ ^ *!^f*^fSfti **$•£$. ';o^W
"££;«••:
fjV;:
-.-.• '.•*.. ..Vf...''i . -'X.ipr-''^ • V C
•fte
S^4^;>- '>*&' : ,J W::%";-^
'•fr,v5?i- ; '.'•:•'.:
~V<;H -Ji,\'&¥:*--ty.''
•*• " - J . 'T-: ,
'/•'-
,v,;;;
'"
W' Jl^
*"•=%•
am
% • • ' « - .
•'-•••
w
&
i
-
te,<. ;
.
•:•;'••.
. *
•.'Jv
.:,<% ' * - , ,
»•?*>!
' : • » ; • " •
in«W •
?^.-'fe'
M^Sfi
•
-
'
:
*&
%
.
•
m* c
.
-••••t*L-
*-l*s*».f: «r«-i
uMt^Ste; .'......-. 75
-.-
.-•^vi-
•SSr ~
••.
Sk»-' ,$&%'S:is •^.:\\ffi,
'•*'.'$&££
?%; ^ • 5 • * « »
* i
•;,:.<,--;-v
.,.!.. ...., ..
100
Energy (MeV) Figure 2. a) T h e E - T scatter plot obtained a t
The attention is focused on how the condition of F=0.33 a n d delay time of 100 ns. performance evolves with the setting of b) the same plot with high gain in energy the CFD, namely the fraction (F) and the delay time. In figure 2a), the scatter plot of time versus energy of charged particles is shown. A good charge separation up to the projectile one can be seen in a wide energy range. The clear deterioration as the energy decreases evidences the sensitivity on the inhomogenity of the resistivity as discussed by G. Pausch [5]. In the inset of figure 2a), which is the zoomed detail of the enclosed area inside the box, the heavier element separation at lower energy region is depicted for a better view. In figure 2b) the higher gain spectrum for lower energy region is plotted, clear isotope discrimination up to oxygen can be obtained down to a lower energy limit of about 5 AMeV for l o n B . We focus ourselves on the charge identification in the following due to the statistics. In the measurement, the time characteristic was extracted at F=0.5 and F=0.6 while the delay time of CFD was set to 60 ns and 100 ns respectively. A fit is used
397
to E-T plot in the low energy region for a fixed charge. The extrapolation with a reasonable width to lower energy of two neighboring lines has an intersection region where the energy threshold is extracted and the uncertainty of the threshold can be estimated. The energy thresholds of B, C and N and the resolving powers show a dependence on the delay time as list in table 1.
Table. 1. Energy threshold and resolving power of charge separation obtained at different conditions.
Delay Time 60 ns 100 ns
F
M Factor
Threshold (MeV) N C B
600 MeV
700 MeV
0.5
21±2
25±2
29+2
2.78
2.50
0.6
20+2
23+2
26±2
2.68
2.15
0.5
15+1
21+1
25+2
2.90
2.81
0.6
16+1
20±1
24+1
2.99
2.91
We projected a thin energy bin of the E-T spectrum on the time axis. In the obtained time spectrum, AT is the time interval corresponding to charge differences AZ and A§ is the average FWHM of these two peaks. We define the resolving power M as:
In table 1, resolving powers obtained at energies of 600 MeV and 700 MeV with a width of ±5 MeV from Ca to Cr are listed for F=0.5 and F=0.6. Obviously as the external delay time becomes longer, a significant improvement of resolving power is presented. The energy thresholds for B, C and N decrease beyond the errors with the increase of the delay time. So both parameters suggest coherently that longer delay time is preferable from the point of view of charge identification. The dependence on the fraction is also essential to understand the performance of this technique. The delay time was kept to 100 ns according to the previous results, while the fraction changed from 0.33 to 0.8. In figure 3, the plot shows the variation of the threshold as a function of fraction for B, C, N, Ne and Na. For all particles, there is nearly no change with the fraction values.. So the threshold itself makes no difference on the determination of the optimal fraction. Alternatively, the resolving power analysis is performed to study its dependence on the fraction. The variation on the fraction is shown in figure 4. Here again the resolving powers at energies of 600 and 700 MeV with a width of +5 MeV are calculated from Ca to Cr. Both sets of data show strong dependence behaviors on the fraction. The resolving power increases monotonously from F=0.33 to F=0.6 and reach the maximum at F=0.6. Then it presents a steep slide to F=0.8. From this point of view, F=0.6 is a candidate for the best condition for our method.
398
70
-
60
« so
; i
1
N>
1
N«
1* JC
t 20
t
-
J
t
t t
. t •
•
0.6
0.7
10
0.3
0.4
0.5
s
«
Fraction
Figure 3. The variation of threshold as a function of fraction .
Figure 4. The behavior of resolving power versus fraction. At 600 (•) and 700 MeV ( • ) .
In conclusion, the PSD method gives a good charge separation up to Ni with a large energy dynamics. The energy threshold is down to 1.5 AMeV for B. The isotope discrimination up to oxygen is also obtained. The charge separation depends essentially on the fraction of the CFD, while there is a small dependence of the energy threshold on the fraction. In our case, the longer external delay time is favorable to the performance of the charge identification. The most prominent feature is that there is no increase of electronic modules compared to the conventional TOF method. And all the electronics is commercially available for general purpose. We just change the delay time of CFD from 20 ns to 100 ns. So this method has the great advantage in the application with large quantity of detector units. Reference 1.
England J. B. A., Field G. M. and Ophel T. R., Nucl. Instr. and Meth. A280(1989)291 2. Klein S. S. and Rijken H. A., Nucl. Instr. and Meth. B66(1992)393. 3. Pausch G. et al., Nucl. Instr. and Meth. A322(1992)43. 4. Pausch G. et al., Nucl. Instr. and Meth. A337( 1994)573. 5. Pausch G. et al., Nucl. Instr. and Meth. A349( 1994)281. 6. Musumarra A. et al., Nucl. Instr. and Meth. A370( 1996)558. 7. Seibt W. et al., Nucl. Instr. and Meth. 113(1973)317. 8. Ammerlaan C. A. J. et al., Nucl. Instr. and Meth. 22(1963)189.
Section VI Reaction Mechanisms around the Barrier. Fusion and Fission in Heavy-Ion Reactions
CROSS SECTIONS FOR COULOMB A N D N U C L E A R B R E A K U P OF T H R E E - B O D Y HALO NUCLEI E. GARRIDO Institute) de Estructura de la Materia, Serrano 123, E-28006 Madrid, Spain E-mail: [email protected] D.V. FEDOROV AND A.S. JENSEN Institute of Physics and Astronomy, DK-8000 Aarhus C, Denmark E-mail: [email protected], [email protected] We present a model describing fragmentation reactions of two-neutron halo nuclei as superposition of all possible reactions where one, two, or the three halo constituents interact simultaneously with the target. The model is suitable for light, intermediate and heave targets. Results in good agreement with the experiment are shown for two-neutron cross sections, interaction cross sections and momentum distributions.
Fragmentation reactions are an important source of information on halo nuclei l'2. Different models describing these reactions on light 3 and heavy 4 targets, where nuclear and Coulomb interaction dominate, respectively, are available. We present a model valid for any target, from light to heavy, and present results for total cross sections and momentum distributions. We start by considering that only those constituents in the projectile inside a cylinder with a certain radius and axis along the beam direction interact with the target, while the others are spectators. We describe fragmentation reactions as superposition of all possible reactions with one, two, or the three halo constituents inside their corresponding cylinder 5 . These reactions contain the nuclear interaction between the halo constituents and the target, as well as the low impact parameter part of the Coulomb core-target interaction. The interference between nuclear and Coulomb core-target interaction is also included. On top of this we include the large impact parameter part of the core-Coulomb interaction (low momentum transfer), that is considered as a process in which the three-body halo system is broken in a smooth collision where the three halo constituents survive in the final state. Let us call Pc and Pn to the probability for the core and the halo neutrons being inside the cylinder in which they interact with the target. The probability of a process in which a certain .group of halo constituents survive in the final state is obtained simply by adding the probabilities for each of the reactions leading to those precise particles in the final state. We can in this way compute total cross sections for reactions in which different groups 401
402 0.4
1
I
I
He on C
0.8
0.3
'
0.6 §0.2 0.4
o.i
0.2
o
I 1 ' 1 I I '1 11
0.4
^
0
' 1 ' 1 ' 1 I1
1.5
Li on C
li
—
Li on C
"i
°_,
0.3
I - °'
e> 0.2
EI
0.1 0 50
0.5
*-
nn
•JL X 11: 250 450 650 850 beam (MeV/nucleon)
E
0 50
J_ _L _]_ 250 450 650 850 ^beam (MeV/nucleon)
50
J_ 250
450
650
850
Figure 1. Two-neutron removal, core breakup and interaction cross sections for fragmentation of 6 He and n L i on C. Cross sections according to the particles left in the final state are also shown.
of halo constituents are left in the final state. In particular it can be seen that the probability for a two-neutron removal process (core surviving in the final state) is P(cr 2 „) = PI1 + P„(l - Pc) + Pn(l - P„)(l - Pc)
(1)
where Pf refers to the probability for the core to be elastically scattered by the target. According to Eq. (1) two neutron removal cross sections can be computed as sum of three processes. In each of them only the interaction between one of the halo constituents (participant) and the target enters. This interaction is described by an optical potential. Differential cross sections factorize then in two parts, one of them describing the participant-target cross section, and another part that is the square of the overlap between the projectile wave function and the final wave function of the spectators 6 . Prom these cross sections one can compute either total cross sections or momentum distributions
403
0 |...-.--r."-7.r.-"..r.--
50
ri
250 450 650 850 Earn (MeV/nucleon)
0I
50
, i , i , i , iI 250 450 650 850 E. (MeV/nucleon)
0I
50
. i , i . i , i 250 450 650 850 E. (MeV/nucleon)
Figure 2. As Fig. 1 for P b target.
of the fragments. We apply the method to fragmentation of u L i and 6 He on C and Pb. We obtain the three-body wave functions by solving the Faddeev equations in coordinate space 6 , and the optical potentials are taken from the literature 5 . In Figs. 1 and 2 we show two-neutron removal, core breakup and interaction cross sections after fragmentation of 6 He and n L i on C and Pb. The cross sections er^, aij, annc and oo (the index refers to the particles left in the final state) are also shown. As a general rule, for light targets the reactions where the core is spectator dominate. For heavy targets the Coulomb contribution dominates in c_2n and absorption of the three halo constituents dominates for <7_c. In Fig. 3 transverse core momentum distributions are shown at a beam energy of 300 MeV/nucleon. The agreement with the experimental data is remarkably good for light and heavy targets. As seen in the lower part of the figure the shape of the distribution is very similar for both targets. This is because the momentum transfer to the core is very small, and in both cases the final core momentum is similar to the one inside the halo projectile. In summary, we present a model to describe fragmentation reactions of
404
Figure 3. Transverse core momentum distribution after fragmentation of 6He and u L i on C and Pb. Beam energy of 300 MeV/nucleon.
two-neutron halo nuclei valid for light, intermediate and heavy targets. This model is able to reproduce experimental two-neutron removal and interaction cross sections as well as momentum distribution of the fragments. References 1. I. Tanihata et al, Phys. Lett. B 160, 380 (1985). 2. T. Kobayashi et al, Phys. Rev. Lett. 60, 2599 (1988). 3. G.F. Bertsch, K. Hencken and H. Esbensen, Phys. Rev. C 57, 1366 (1998). 4. C. Bertulani and L.F. Canto, Nucl. Phys. A 549, 163 (1992). 5. E. Garrido, D.V. Fedorov and A.S. Jensen, Phys. Lett. B 480, 32 (2000). 6. E. Garrido, D.V. Fedorov and A.S. Jensen, Phys. Rev. C 59,1272 (1999).
MULTINUCLEON T R A N S F E R R E A C T I O N S S T U D I E D W I T H T H E PISOLO S P E C T R O M E T E R L. CORRADI, A.M. STEFANINI, A.M. VINODKUMAR INFN, Laboratori Nazionali di Legnaro, Legnaro, (Padova), Italy S.BEGHINI, G. MONTAGNOLI, F.SCARLASSARA Dipartimento di Fisica, Universita' di Padova, and INFN, Padova, Italy
Dipartimento
G.POLLAROLO di Fisica Teorica, Universita' di Torino, and INFN, Torino, Italy
The time-of-flight magnetic spectrometer PISOLO has been taking data since 1995. Important effects in weak channels populated in multinucleon transfer reactions have been identified for a variety of reactions. Some of the outstanding results will be briefly discussed.
1
Introduction
Multinucleon transfer reactions (MNT) between heavy ions at energies close to the Coulomb barrier represent an important field of research in low energy nuclear physics 1>2>3>4>5. These measurements offer in fact the possibility to investigate 1) the interplay between single nucleon transfer modes and more complex degrees of freedom2 (e.g. pairing and cluster effects) , 2) the transition from quasi-elastic to deep-inelastic processes 3 , 3) the coupling effects with other reaction channels 4 (e.g. subbarrier fusion). Moreover, MNT may be a valuable alternative mechanism to spallation with energetic proton beams or to asymmetric fission for the production of neutron rich nuclei, and is therefore important for the physics to be done with radioactive nuclear beams 5 . The time-of-flight magnetic spectrometer PISOLO 6 ' 7 , installed at LNL, has been especially designed to combine a high efficiency and high Z and A resolution for heavy ions produced in binary reactions at energies down to ~ 1 MeV/amu. The spectrometer has been taking data since 1995, and various experiments on different systems have been successfully performed since then 6,7,8,9,10 j n Table 1 a list is given of the studied reactions. In the contribution to this proceedings a brief survey on some outstanding results will be presented. 2
Some recent results
One of the main achievements of the studies performed with PISOLO was the unambiguous identification in several systems of channels corresponding to the pick-up of up to six neutrons and the stripping of six protons. The 405
406 Table 1: Systems studied with the PISOLO spectrometer using the beams delivered by the Tandem-ALPI accelerator complex of LNL.
system 40 48
Ca+ 1 2 4 Sn Ca+ 1 2 4 Sn
64Ni+238TJ 40Ca+90,96Zr 62Ni+206pb 58Ni+208pb
EL (MeV)
subject
ref.
170 174 390 137,152 342-317 370-330
indep.part. transfer modes cluster effects transition quasi-el. to deep-inel. sub-barrier fusion neutron pair transfer modes fission of the heavy partner
[6] [8] [9] [10] new,unp. [5]
variety of channels which could be observed allowed to follow in a systematic way the population pattern of the reaction products in the Z-A plane. Taking as a representative example the system 6 4 Ni+ 2 3 8 U 9 , one could observe (see Fig.l) that the cross sections for pure neutron transfer channels drop by an almost constant factor ~ 3.5-4 per each transferred neutron, in agreement with observations done for other systems studied both by our group at LNL 6 , 1 0 and at Argonne 11,12 . This suggests that neutrons behave as independent particles 6-kNi +*MV (E.^390 MeV) 10°
T — i — n (-6p)
10"
S
T
i (-Sp)
|—i
r (-*p)
1 r I"-! I — (-3p) : (~2P)
I-1
1— (-lp)
r~i—r (Op)
10'
10"
Mt
10"
_1_
50 60 70
!f
• •
i
i r i III.I
50 60 70
50 60 70
50 60 70
50 60 70
50 60 70
50 60 70
MASS NUMBER Figure 1: Experimental (dots) and theoretical (lines) total cross sections for the transfer channels observed in 6 4 N i + 2 3 8 U at E ( a 6 = 3 9 0 MeV (see ref. [9] for details).
in the transfer process. However, looking at particular angles and choosing systems whose ground state Q-values are as close as possible to the optimum ones, one can investigate if nucleon correlation effects show-up. One of these
407
effects is the appearance of an odd-even staggering in the mass yield, which would indicate the presence of pairing modes. Studies of this kind in pure neutron transfer channels are being carried out with PISOLO in 6 2 Ni+ 2 0 6 Pb. Wilczynski plots have been produced 9 for each specific A and Z showing how the Q-value distributions evolve as a function of angle and number of transferred nucleons. It has been shown that even at bombarding energies close to the Coulomb barrier one can have very large energy loss even for few nucleon transfer channels. The energy loss generates, in turn, nucleon evaporation from the primary fragments, shifting the mass distribution. In Fig.l the data have been compared with state-of-art coupled channel calculations for MNT performed with the code GRAZING 13 . The theory is fully developed in ref. [2] and we refer to that work for details. Briefly, theory includes independent single nucleon transfer modes and the inelastic excitations to the lowest nuclear levels, and it takes into account nucleon evaporation from the primary fragments. As one sees, a nice agreement with the data exists for the pure neutron channels and for the (-lp) and (-2p) ones, but as the number of protons increases marked discrepancies show-up. Part of these discrepancies are presently attributed to the small impact parameters leading to large energy losses, which call for further refinements in the calculations. 3
Inputs for radioactive ion beam studies
High quality data as the present ones provide important inputs for the physics to be done with radioactive beams 5 , even if experiments are being performed with stable beams. As mentioned before, these inputs are coming from the accurate determination of a) the Z,A and Q-values distributions, b) evaporation effects from the primary fragments, c) drop factors in the yields as a function of the number of transferred nucleons, d) energy loss distributions. For instance, calculations predict 1 4 an important "inversion" of population from proton stripping and neutron pick-up to the opposite modes when moving from a neutron poor to a neutron rich projectile. This trend is experimentally observed e.g. in the case of the 40 ' 48 Ca-|- 124 Sn systems 6,8 (see also the discussion in ref. [5]). Such a population inversion should strongly favour the production of neutron rich heavy nuclei. An important aspect not yet experimentally investigated into detail is what fraction of the yields in the heavy partner survives fission after the transfer of several nucleons. A high resolution kinematic coincidence experiment to study that aspect is being carried out in the system 58 N i + 208p b 5 I n F i g 2 w e s how a Z-spectrum obtained with PISOLO at an angle close to the grazing one. Proton channels have been identified down to the (-8p) channel with traces of (-10p) events. The ratio between the events
408
200
225 Z
250 275 (Channel)
Figure 2: Mass integrated charge distribution for the reaction and 6iab=80°.
300 58
325
N i + 2 0 8 P b at Ej a (,=347 MeV
taken in coincidence with another detector placed at the kinematically correlated angle and the singles allows to derive the fission probability as a function of Z,A and Q-values of the various transfer channels. References
6. 7. 8. 9. 10. 11. 12. 13. 14.
C.Y.Wu, W.von Oertzen, D.Cline and M.Guidry, Annu. Part. Sci. 40, 285 (1990). A. Winther, Nucl. Phys. A 572, 191 (1994); Nucl. Phys. (1995). K.E. Rehm, Annu. Rev. Nucl. Part. Sci. 4 1 , 429 (1991). M.Dasgupta, D.Hinde, N.Rowley and A.M.Stefanini Annu. Part. Sci. 48, 401 (1998). L.Corradi et al in RIB2000, Divonne (France), April 3-8, Phys. A, in press. L.Corradi et al Phys. Rev. C 54, 201 (1996). G.Montagnoli et al Nucl. Instr. and Methods, in press. L.Corradi et al Phys. Rev. C 56, 938 (1997). L.Corradi et al Phys. Rev. C 59, 261 (1999). G.Montagnoli et al J.of Phys. G 23, 1431 (1997). C.L. Jiang et al Phys. Lett. B 337, 59 (1994). C.L. Jiang et al Phys. Rev. C 57, 2393 (1998). A. Winther program GRAZING unpublished. C.H. Dasso et al Phys. Rev. Lett. 73, 1907 (1994).
Rev.
Nucl.
A 594, 203
Rev. Nucl. 2000, Nucl.
STUDY OF CLUSTER EMISSION BARRIER IN 12C + M8PB ELASTIC SCATTERING AND POSSIBLE OBSERVATION OF QUASIMOLECULAR CONFIGURATION A.A.OGLOBLJN, K.P.ARTEMOV, YUA.GLUKHOV , A.S.DEM'YANOVA, V.V.PARAMONOV M.V.ROZKOV AND V.P.RUDAKOV Kurchatov Institute, Mosco E-mail: [email protected] S.AGONCHAROV Skobltsyn Institute of Nuclear Physics, Mosco E-mail: [email protected]
1
Introduction
Cluster radioactivity leads to formation of fragments which are produced in their ground states contrary to what happens in fission Due to this fact the interaction between them (elastic scattering or fusion) can be considered as a some kind of the inverse process, which could give in favorable circumstances independent information about the shape of the barrier. This is important fo understanding die mechanism of cluster radioactivit because different theoretical models giving simila penetrabilities predict completely different shapes of the barrier [1]. Normally near-barrier heavy-ion scattering is not sensitive to the form of the potential due to strong absorption. However, if one measures the exponential falloff of the Fresnel diffraction cross-section with high accuracy and in large angular interval the situation could be not so pessimistic [2]. 2
Results and Discussion
The experiment was fulfilled at Kurchatov cyclotron at the 12C ions energy 75.7 MeV. The detection system contained several telescopes of Si-counters. The thickness of the self-supporting 208Pb- enriched target was 270 mg/cm2. The energy resolution allowed to separate reliably elastic peak from the inelastic ones. In Fig.l the differential cross-section of 12C +208Pb elastic scattering is presented. Almost in the whole angular range its behaviour does not differ from the wellknown pattern of near-barrier heavy-ion elastic scattering: Fresnel (or Coulomb 409
410
rainbow) oscillations at small angles followed by exponential fall-off. However, at the largest angles some deviations from the standard picture are observed. First, it is change of the slope of the angular distribution, and, second, the appearance of small oscillations. The traditional optical model potential with volume absorption reproduces the data quite well, except for the large angles oscillations. The oscillations under discussion can be obtained by diminishing the absorption. In general, the absorption is already rather weak: the cut-off procedure shows that low partial waves with L = 5 -10 contribute to the cross-section at large angles. This corresponds to the distances of penetration of about 6 fm. The fits with W containing both the volume and surface components give slightly better agreement with the data. However, in order to reproduce quantitatively the structure of the large angles cross-section a rather complicated radial dependence of the imaginary part is required. At die present stage of the analysis the main result obtained with the twocomponent imaginary part is shown in Fig.2. All reasonable potentials have a "pocket". The sum of nuclear and Coulomb potentials form a minimum at the distance of about 10 fm. Though the depth of the pocket is not unambiguous only the potential denoted by solid line is consistent with the decay energy of 220Ra -> 12 C + 208Pb. The position of the minimum is close to the sum of the l2C + 208Pb radii. So the touching point lies inside the potential pocket, and the obtained barrier is a "real" barrier whose penetrability should be taken into account in the decay process. One can calculate the position of the first level in this pocket. It occurs to be very close to the indicated ground state energy of 220Ra. So the probability of formation of quasimolecular deformed configuration 12C + 208Pb in 220Ra or, in other words, a shape isomer, becomes strongly enhanced. Naturally, one can expect the appearance of rotational states based on this quasimolecular state. In 220Ra there are well-known two rotational bands: quadrupole and octupole . The distances between the levels correspond well to the moment of inertia equal to ~ 2/3 of the solid state one for the touching configuration 12C + 208Pb. Preliminary data on l s O + 208Pb elastic scattering cross-section analysis is not yet finished. At the moment we can only state that the deviation from exponential falloff takes place similarly to the case of 12C + 208Pb. Another process which can give an independent information about potentials beyond the top of the barrier is subbarrier fusion. Fusion-fission cross-sections for the same combination of nuclei 12C + 208Pb and 1 6 0 + 208Pb were studied recentl [3], The potential reproducing the data [4] has a pocket at about 10 fm, the touching point for 16 0 +208Pb interaction lies inside it similarly to 12C + 208Pb
411
elastic scattering results. The energy of Th ground state is slightly above the minimum, and (he conditions required for formation of corresponding quasimolecular state are fulfilled. 3
Conclusions
To summarize, we state that large angles heavy-ion elastic scattering measured near the Coulomb barrier in favorable circumstances can be sensitive to the inne part of the interaction potential. Simultaneous study of cluster decay, elastic scattering and fusion of the daughter nuclei can provide complimentary information about the mechanism of cluster radioactivity. From this point of vie the study of interactions of 14C + 208Pb and 22Ne + 208Pb being the inverses of the measured decays of 222Ra and 230U correspondingly seems ver important Study of elastic scattering and fusion-fission in the systems 12C, 1 6 0 + 208Pb gave some evidence of the existence of corresponding exotic quasimolecular configurations ( shape isomers ) in the ground states of 220Ra and 224Th. Probably, this phenomenon is not limited by the nuclei mentioned above, and is more widespread, as it was suggested in [5].
100
12Q + 208pb
O / CJRuth
Eiob = 75.7 MeV
10-'
10-S r
10--! r —
V
r
a
Ws
rs
as
(MeV)
(fm)
(fm)
(MeV)
(fm)
(fm)
12.5
1.39
0.520
8.95
1.40
0.349
--28.2
1.39
0.371
6.72
1.40
0.246
•-• 35.4
1.39
0.350
8.80
1.40
0.208
cm 90
Figure 1. Differential cross-section of elastic scattering data
180
C+
Pb and optical model potentials fitting the
412
80
i
i
i
i T—r—!——i—T—t—T—^r—r——i—r——i—i-
12C+208pb
95,7 MeV
220Ra_J2C+208pb
R( 12 C)+R( 208 Pb)
10
12
14
16
18
20
r, fm Figure 2. Optical model potentials with two-component imaginary part. The touching point of ground state energy of ^"Ra are inside the "pocket" in the real part of the potential.
12
C + 208Pb and tl
References 1. 2. 3.
4.
5.
D.N.Poenaru and W.Greiner, in "Clustering Phenomena in Atoms and Nuclei", ed.by M.Brenner, T.Lonnroth, F.B.Malik,, p.235 A.S.Dem'yanova, Yu.A.Glkhov, S.A.Goncharov, A.A.Ogloblin, M.V.Rozkov in ENAM95, p.401 S.P.Tretyakova et al., Int. Conf. on Nucl.Phys. "Nuclear Shells - 50 years", Dubna, Russia, 21-24 Apr. 1999, ed. by Yu.Ts. Oganessian and R. Kalpakchieva, World Scientific, Singapore 2000, p. 154 V.M. Shilov, Proc. of 6th School-Seminar, Dubna, Russia, 22-27 Sept 1997, ed. By Yu.Ts. Oganessian and R. Kalpakchieva, World Scientific, Singapore 1998,p.331 B.Buck and A.C.Merchant, Phys.Rev.C39, 2097 (1989)
VERY S T R O N G R E A C T I O N C H A N N E L S AT B A R R I E R ENERGIES IN THE S Y S T E M 9 B e + 2 0 9 Bi C. SIGNORINI, A. ANDRIGHETTO, J. Y. GUO, L. STROE, A. VITTURI Physics Department and INFN, via Marzolo 8, 1-35131 Padova E-mail: [email protected] M. RUAN, M. TROTTA INFN, Laboratori Nazionali di Legnaro, via Romea 4, 1-35020 Legnaro, Padova F. SORAMEL Physics Department and INFN, via delle Scienze 208, 1-33100 Udine K. E. G. LOBNER, K. RUDOLPH Sektion Physik, Munich University, D-85748 Garching
Physics Department,
I. THOMPSON University of Surrey, Guilford GU2 5XH, U. K.
D. PIERROUTSAKOU, M. ROMOLI Physics Department and INFN, via Cinthia, Monte S. Angelo, 1-80125 Napoli Very strong reaction channels were observed in the system 9 Be + 209 Bi with cross section ranging from 70 mb, well below the Coulomb barrier, up to 700 mb above. These channels were assigned to 9Be breakup + transfer processes. A simultaneous description of scattering, fusion and transfer + breakup for this system in terms of CC channels theory is fairly successful.
1
Experimental Data
In the course of systematic studies of the system 9 Be + 209 Bi at bombarding energies around the Coulomb barrier a very broad and strong peak, with energy between 18 and 21 MeV, was systematically observed at all bombarding energies and all angles 1 ' 2 , 3 . The connection of this bump to breakup phenomenon (BU) was quite straightforward. In fact, several works with 9 Be + 209gji,2,3 an( j 9g e _|_ 208PJJ4 a t barrier energies, strongly indicate that projectile breakup plays a significant role since 9 Be, though stable, is very weakly bound (Sra = 1.67 MeV). A similar effect has already been observed in the system 6 He -I- 2 0 9 Bi 5 . We therefore concluded that it was worthwhile to analyze the experimental data obtained from three previous experiments 1>2'3 to get quantitative results for this broad peak. In the following we will not describe in detail the experimental setups, we will rather summarize the experimental details relevant to the present paper. 413
414
Experimental data consist basically of two sets of charged particles spectra detected with Si detectors. One is an excitation function with one Si detector, 100 fjaa thick, at 135° to the beam, covering a large solid angle (415 mm 2 effective surface at 7 cm from the target, corresponding to a solid angle of 85 msr and an opening cone of 19°). The spectra were recorded at 9 Be laboratory beam energies ranging from 36 to 50 MeV with steps < 1.0 MeV 1'2. The second data set is an angular distribution collected in ~ 3.5° steps from 40° to 156° at energies of 40, 42, 44, 46, and 48 MeV. In this case there were several identical Si detectors, 300 /zm thick, covering very small solid angles (6 mm diameter collimator at 25 cm from the target, solid angle of 0.45 msr and opening cone of 1.4°) 3 . In both cases the absolute normalization was done by comparison with the Rutherford scattering recorded at 30°. From these two sets we obtained absolute excitation functions under the reasonable assumption that the angular distributions taken at energies < 40 (> 48) MeV are identical to the 40 (48) MeV one.
E l a , (MeV)
Figure 1: Charged particles spectra for the system 9 Be + 2 0 9 Bi observed under the same experimental conditions except for the detector angle. The broad peak in the upper spectrum at ~ 34 MeV, assigned to 2a from 8 B e dissociation, does not appear in the lower spectrum because the detector opening is much smaller in this case (1.4° vs. 19°) and the 2a particles, emitted with a maximum opening cone of 8°, have very low probability to be simultaneously detected.
Fig.l shows two typical spectra recorded at 42 MeV and 135°: the top one with large solid angle and the bottom one with small solid angle. Comparing these two spectra the two broad peaks at ~ 18 and ~ 34 MeV were assigned to l a and 2a particles from the 8 Be unbound g.s. breakup. In fact if, in
415
first approximation, we neglect the 9 Be neutron binding energy and assume that 9 Be primarily breaks into 8 Be + In in the proximity of the target nuclear field, keeping the Coulomb like trajectory, the 8 Be energy is ~ 8/9x E<>Be i.e. (8/9 x 36 MeV) ~ 33 MeV. Since 8 Be breaks immediately into two a particles with a maximum opening cone of ~ 8°, a sizeable number of these two fragments can be simultaneously detected in the large area detector as 2a = 8 Be (opening cone 19°) but not in the small area one (opening cone 1.4°). In the 9 Be breakup also the In transfer reaction 209 Bi( 9 Be, 8 Be) 210 Bi could originate 8 Be (i.e. 2a) nuclei. However, in this case, the g.s. transfer is at 39.3 MeV and the a particles at half energy, while transfer to excited states and/or to continuum levels, having lower energy, would end into the broad peak. Other clearly visible peaks in Fig.l are the a emitted from the ground state de-excitation of the various evaporation residues (~ 9 MeV), and the 9 Be elastic peak with energy around 36 MeV according to kinematics. The angular distribution measured in these conditions reflects that of 8 Be nuclei, with some smoothing due to the 8° opening cone of the 2a particles; from its integral it is anyhow easy to get the total BU + transfer cross section. Neither of the two processes can, of course, be excluded.
BU sequential+Transfer
10-il
i
l
35
i
i
i
i
i
i
i
i
i
l
i
i
i
i
i—i
40
i
i—i—I
i
i—i—i
i
i
i—1
45
E o r ,(MeV)
Figure 2: Comparison of experimental and theoretical breakup, transfer and fusion cross sections.
Fig.2 shows the excitation functions for total BU + transfer cross sections as well as for total fusion cross sections. The results are quite astonishing since the BU + transfer cross sections are very large: see e.g. at E;06 = 36 MeV ((ECTO = 34.5 MeV) were afus = 0.6 mb and aBu+tr = 70 mb! At 50 MeV
416
((E c m = 48 MeV) the two cross sections are comparable (600 - 700 mb). 2
Theoretical approach
The experimental data for the system 9 Be + 209 Bi are a rather complete set consisting of fusion (and related barrier distribution), elastic scattering at 5 energies around the Coulomb barrier and BU + transfer cross sections. We have therefore done a considerable effort to describe all these data with onlyone theoretical approach in the frame of a coupled channel description. The code FRESCO was adopted using a nuclear potential of Woods-Saxon form with both real and imaginary volume absorption. In the couplings both target and projectile excitations and In transfer channels were considered. For the various couplings originating from the 209 Bi target we have assumed 209 Bi = 2 0 8 Pb which could be reasonably justified in a weak coupling limit; this implies a very large reduction in angular momentum space since 2 0 8 Pb has 0+ g.s. to be compared with the 9 / 2 " of 209 Bi. This choice limits computing time and allows to make as many trials as necessary to tune the various potential parameters and reproduce the data. The inelastic excitations considered were the 3 " octupole vibration in 2 0 8 Pb, the first three 9 Be levels with J* = 1/2+, 5/2", 7/2", and the re-orientation of the 3/2~ 9 Be g.s. Negative parity 9 Be levels constitute a rotational band whose experimental coupling strengths 6 are listed in table 1. Table 1: Parameters utilized for the target and projectile excitation couplings in the CC calculations. The deformation length is for the nuclear component of the coupling, M(EA) is for the electromagnetic one. All these parameters are deduced from experimental 6 transition strengths. The parameters for the 5 / 2 " level in 9 Be were deduced from a ,82 = 0.755 value obtained from a scattering 7 since the experimental B ( E 2 ) 6 of 24 W.u. gives non-physical 02 = 2.1 and deformation length = 5.3 fm! The 7 / 2 _ level strength was consequently reduced by a factor 4.
1*
a»»Bi (= 2 0 b Pb) 0+ 9 Be 3/23/23/23/2-
x
E* (MeV)
Deformation length (fm)
M(EA) exfm A
3-
2.60
0.80
815 (E3)
1/2+ 5/27/23/2-
1.68 2.43 6.76 0.00
0.37 1.35 0.79 0.25
J
0.350 3.221 1.893 0.627
(El) (E2) (E2) (E2)
417
For the In transfer we have considered the first 7 levels of 2 0 9 Pb. They are those observed in the ( 9 Be, 8 Be) In transfer from 2 0 8 Pb at 50 MeV 8 : 9/2+, 0.0 MeV; 11/2+, 0.78 MeV; 15/2+, 1.42 MeV; 5/2+, 1.57 MeV; 1/2+, 2.03 MeV; 7/2+, 2.49 MeV; 3/2+, 2.54 MeV. For each of these levels we have assumed a transfer amplitude equal to one since they are all of single particle nature. Keeping these coupling strengths fixed, we have varied the six Woods-Saxon potential parameters to reproduce as well as possible the elastic scattering differential cross sections at the five bombarding energies (40 - 48 MeV).
20
60
100
140
180
cm. (dog)
Figure 3: Experimental and calculated elastic scattering angular distributions. The CC code FRESCO with the inelastic and transfer couplings described in the text has been used.
Fig.3 shows how the experimental angular distributions are fairly well reproduced with only one potential. The best fit parameters are Vo (Wj) = 175.5 (1.5) MeV, r 0 (r*) = 1.178 (1.0) fm, a<> (a*) = 0.68 (0.58) fin. Fig.2 shows the total transfer cross sections obtained summing the seven In transfer channels considered, and the BU cross sections (sum of the three 9 Be inelastic excitations considered). All 9 Be levels are unbound therefore these excitations constitute the 9 Be BU process via resonances (sequential BU), the BU to the continuum was not considered. Total BU + transfer cross sections have to be compared with experimental ones. In Fig.2 we also report fusion cross sections calculated as difference between the total reaction cross section and the various excitations considered. We notice that: i) the two channels, BU and transfer, are calculated with comparable strengths, ii) the total BU + transfer calculations underestimate the experimental data, iii) calculated fusion overestimates the data. These observations suggest that 9 Be BU to the continuum might be
418
able to fill the gap between experimental data and calculation results. References 1. 2. 3. 4. 5. 6.
C. Signorini et al, EPJ 2, 227 (1998). C. Signorini et al, EPJ 5, 7 (1999). C. Signorini et al, PRC 6 1 , 061603(R) (2000). M. DasGupta et al, PRL 82, 1235 (1999). E. F. Aguilera et al, PRL 84, 5058 (2000). F. Ajenberg-Selove et al, NPA 490, 1 (1988) and Nucl. Data Sheets 47, 840 1986. 7. R. Subinit et al, PRC 52, 1524 (1995). 8. D. P. Stahel et al, PRC 16, 1436 (1977).
Nuclear Rainbows, Nuclear Matter and the
16
0 + 1 6 0 System
W. von Oertzen 1 ", A. Blazevic 1 , H. G. Bohlen 1 , Dao T.Khoa 2 *, F.Nouffer1, P. Roussel-Chomaz 3 , W. Mittig 3 , J. M. Casandjian 3 1 Hahn-Meitner-Institut-GmbH, Glienicker Strasse 100, D-1^109 Berlin, Germany 2 Institute for Nuclear Science, VAEC, POBox ST-160, Hanoi, Vietnam 3 GANIL, Bd. Henri Becquerel, BP 5027, F-14021 Caen Gedex, France The elastic scattering of strongly bound nuclei at energies of 10-70 MeV/u shows the phenomenon of "Rainbow scattering". In these cases deflection into negative angles involves strong overlapp of nuclear densities, corresponding to twice the saturation density of nuclear matter. The system 1 6 0 + 1 6 0 has been studied over a wide range of angles and energies between 7 MeV/u to 70 MeV/u with high precision in several laboratories in recent years. Pronounced features which are due to the primary and higher order Airy-structures are observed. At all energies excellent fits are obtained with the double folding model, the result confirms the refractive origin of the large angle scattering, where rainbow scattering yields unique information on the properties of cold nuclear matter at higher densities
1
Nuclear Rainbow Scattering
Nuclear refractive (rainbow) elastic scattering has been the subject of increased attention in the last decades, because it became possible to establish that deep potentials are needed to describe the systematics of heavy ion scattering, which were originally fitted with rather shallow potentials. For the system 1 6 0 + 1 6 0 1,2,3,6,10,11 ag w e n ^ for Q;-particle scattering 4 ' 5 , the use of the double folding model for the real part of the potential has revealed a very clear sensitivity of the large angle scattering data on the details of the real potential at small distances. At these small distances, where large density overlaps of the scattered nuclei occur, the double-folding potential is very sensitive to the details of the effective nucleon-nucleon interaction (based of the Paris M3Y-interactior?' 7 ' 8 ), and it has been shown that a consistent description can only be obtained with a distinct, but small density dependence 2,3,7 ' 8 . The same effective interaction gives in Hartree-Fock calculations a soft equation of state of cold nuclear matter 6 ' 7 ' 8 , with an incompressibility parameter in the range of K~220-250 MeV with an accuracy of (±15%). In this brief survey we discuss the experimental results for the elastic scattering of 1 6 0 + 1 6 0 over a wide range of energies. Precise data have now been measured up to large angles at low energies at IreS (Ej0&= 75 - 124 MeV, at 9 energies 10 ) and at Elab = 250, 350, 480, 704 and 1120 MeV at the HMI and "Also Freie Universitat Berlin, Fachbereich Physik, D-1000 Berlin 39, Germany. Corresponding author: [email protected] 419
420
Dark Region
Primary Rainbow
> Secondary Rainbow!
Scattering Angle
•*•
Figure 1: The Airy function describing the rainbow phenomenon. The firts intensity maximum appears beyond the classical shadow border line, and is followed by higher order maxima and minima, which are indeed observed in 1 6 0 + 1 6 0 scattering at lower energy, see Figs. 2
at GANIL n ; further data at two energies have been measured by Sugiyama et al. 1 2 . The phenomenology and the quality of the data are shown in Fig. 2 for the elastic scattering at higher energies (124 MeV up to 1120 MeV), and at the lower energies. The elastic scattering data, which exhibit well pronounced features of refractive scattering at large angles have been fitted with an optical model, where the real part has been obtained by the double folding model or by a functional dependence of Woods-Saxon squared form f?(r), where fiif) = (1 + e x p ( ^ p - ) ) - 1 ; with i = V,W for the real and imaginary parts. The latter parametrisation gives potential shapes which are very close to the double folding potentials. In addition a surface term, with a form factor of the derivative 4os J ; / s W has to be added to the imaginary potential, as has been found already in earlier work 8 . We repeat here the basic facts of rainbow scattering 13>14. The rainbow structure appears if the nuclear potential is strong enough to deflect particles into "negative angles" and a maximum deflection (rainbow) angle occurs. In this case a particular interference structure due to contributions around the maximum deflection angle will appear, which is described by an Airy-function ( this function is shown in Fig. 1); note the higher order maxima appearing inside the lighted region. The most remarkable feature of this complete data
421
set for 1 6 0 + 1 6 0 is the fact that we can follow the evolution of the primary Airy structure from the higher energies of EJ 0 J, = 350 MeV, where it is well pronounced down to 124 MeV and lower, where the higher order Airy structures appear in the angular range of observation.
0
20
40
60
80
100
Q
Figure 2: Differential cross section of 1 6 0 + 1 6 0 elastic scattering over many energies (the fits to these data are discussed in Refs. 7 - 10): a) between (E (oi ,= 124 to 1120 MeV). The primary rainbow maximum at 350MeV is located at an angle of 55°, it moves outside the observation region at lower energies. b)Differential cross sections at energies between Ej0j,= 95 and 124 MeV, with fits obtained with optical model potentials obtained from the double folding model or with the Woods-Saxon squared potentials 10
2
Double folding model and the EOS of Nuclear Matter
In the systematic analysis the folded potentials and the Woods-Saxon squaredpotentials(WS2) have been found to give equivalent overall fits to the elastic scattering data. We must emphasize the most important point concerning the potentials in this analysis: The originally (30 years ago) used Wood-Saxon potentials have the wrong radial shapes. This is in particular also true for the cases of a-particle scattering, where only in the last 15 years the systematics of scattering states and bound states have been established with the same
422
potentials 9 . The potentials can be classified by their volume integrals defined as 47T
Jv,w =
NANB
2 / Vv,w{r)r dr,
(1)
normalised to the number of interacting nucleon pairs (nucleon numbers are given by iV^ and NB for projectile and target). Already in previous studies using the folding model 8 for the nucleus-nucleus potentials, in particular for a-particle-nucleus scattering 4 ' 5 , it has been found that a consistent description is obtained with values of the volume integrals for nucleon-nucleus potentials 16 . Thus a criterion for the choice of the potentials has often been the value of the volume integral of the real potential per interacting nucleon pair. From
400
6
o+ 16 o
350
-WS2 - Folding
^300
«
E > 250
a> ~2(KH > ^ 15010050
300
600
900
1200
Ebb (MeV)
Figure 3: Volume integrals J\r and Jw of the best-fit real WS2 and folded potentials and of the imaginary potentials, for the 1 6 0 + 1 6 0 system at incident energies from 124 -1120 MeV. The lines are only to guide the eye.
the analysis volume integrals of Jy ^300 MeV- fm3 at Eja(, ~30 MeV/u is obtained 10,11 . The volume integrals of composite particles are slightly reduced due to antisymmetrisation effects. We will come back to this question later. The systematics of these volume integrals are shown in Fig.3 over all energies. Other choices of potentials are often discussed which are obtained by inversion proceedures, however, they have serious deficiencies: a) failure to reproduce over many energies the systematics of the refractive rainbow structures in the angular distributions at large angles; b) they are not able to reproduce the
423
i
150 120
i
» i-r
i
I I I |
i
6o
<
|
I I
• • [
: Cold Nuclear EOS / " (M3Y-Paris)
/
BDM3Y3
:
BDM3Y2
~
/
/
/
/
BDM3Y1 /
\
1 1
.'
:
30
/
/ i K = 43SiMeV
90 '.-
S
i
'•
*/-»«"*y /
K
,{10
MoV
nrnusYi / / :
/ / •' / / '
**^ s .
-30 0.0
, /...-••"
0.2
0.4 P [fm 1
i_'i?i. i i
systematics of the transfer reactions. There are also large variations in absolute normalisation of DWBA calculations relative to the data in the 1 5 0 + 1 7 0 channel, expressed via the product of the spectroscopic factors of the individual nuclei, SiS2=SF, the new potentials give a consistent value of SF at all energies 18 . An important aspect of the study of refractive scattering has been the study of the in-medium effective nucleon-nucleon interaction 5>6>7>8. This is acchieved by introducing into the M3Y-interaction a density dependence in such a way that the corresponding Hartree-Fock calculation reproduces the saturation point of nuclear matter 8 ' 1 6 . In the double folding model and in the Hartree-Fock calculation the finite range exchange part must be treated consistently. With different choices of the density dependence, different values of the nuclear incompressibility described by the faktor K can be chosen. Results for the equation of state (EOS) for different K are given in Fig.4.
\ '•
'
__— ———"— '
= 1 5 0 MeV
0.6
0.8
Figure 4: Results of Hatree-Fock calculations using different versions of the density dependent M3Y(Paris) interaction, giving different values for K.
In this approach the systematics of the nucfeon-nucleus potentials as well as the mean field potentials of Jeukenne et al. 1 6 are reproduced 8 . Finally we come back to the question of the Pauli-blocking effects in the double folding model. In both, in heavy ion scattering and in a-particle scattering very deep Woods-Saxon-squared potentials are obtained, so as to comply with the Pauli principle by generating the appropiate number of nodes for the wave function
424 Distance in momentum space Fermi - spheres K - assymptotic v
<
s
icx:%
1/2
K=[2u(E-V(r))/h] K - with mean field in rainbow region and with density overlapp
p
1Q.
radial extensions Figure 5: Illustration of the reduction of the Pauli-blocking (reduced overlapp of the Fermi spheres) due to the very strong attractive potential in the elastic channel created by the mean field effect in the double folding model.
in the interior (the rule that (2N+L) = E(n,- + 1.) is obeyed 1 2 ) . The double folding model with an effective N-N-interaction adjusted to the properties of nuclear matter, gives a potential for the elastic channel as a mean field effect, with a depth and shape, which is consistent with the cited condition. An alternative to include the Pauli-blocking in the double folding in a self-consistent way has been recently formulated 1 7 . The main idea is shown in Fig.5. In conclusion we find, that the complete set of data for the 1 6 0 + 1 6 0 system gives clear criteria for the choice of a particular class of real potentials, which agree well with the results of the double folding model calculations, based on a nucleon-nucleon interaction with a weak density dependence as discussed in Refs. 2>3,5,6,7,8 -phe re i a te
425
4. 5. 6. 7. 8. 9. 10. 11. 12. 13. 14. 15. 16. 17. 18.
S. Ohkubo, Phys. Rev. Lett. 74 (1995) 2176 D. T. Khoa, W. von Oertzen, Phys. Lett. B342 (1995) 6 G. Bartnitzky et al., Phys. Lett. B386 (1996) 7 D. T. Khoa, W. von Oertzen, and H. G. Bohlen, Phys. Rev. C49 (1994) 1652 D. T. Khoa, W. von Oertzen, Phys. Lett. B304 (1993) 8 F. Michel, S. Ohkubo and G. Reidemeister, Progr. Theor. Suppl. 132 (1998) 7 M. P. Nicoli, F.Haas, R.M. Freeman, et al., Phys. Rev. C60 (1999)064608. Dao T. Khoa, W. von Oertzen, H. G. Bohlen, et al. , Nucl. Phys. A672 (2000)387 Y. Kondo,Y. Sugiyama, et al., Phys. Lett. B365 (1996) 17 M. E. Brandan, G. R. Satchler, Phys. Reports, 285 (1997) 143 J.D. Jackson, Phys. Reports, 320 (1999) 27 D. T. Khoa, G. R. Satchler, W. von Oertzen, Phys. Rev. C56 (1997) 954. J. P. Juekenne, A. Lejeune, C. Mahaux, Phys. Rev. C16 (1977) 80. V. Subbotin, W. von Oertzen, et al., submitted for publ.(2000) H.G. Bohlen, D. T. Khoa, et al., to be published (2000).
BREMSSTRAHLUNG BY NONRELATIVISTIC PARTICLES IN MATTER
Moscow
Institute
A. V. K O S H E L K I N for Physics and Engineering, Kashirskoye sh.31, Russia, [email protected]
Moscow
115409,
Bremsstrahlung by nonrelativistic particles undergoing multiple elastic collisions in the matter is studied. The probability of photon production of such particles is calculated. It is shown that the spectrum of emitted photons differs strongly from the spectral distribution of bremsstrahlung by ultrarelativistic particles.
1
Introduction.
As a rule, we deal with photon production by high energy particle in a medium because the conditions in which photons are generated are such that the energy of emitting particles are much more then their masses. This takes place when we study the bremmstrahlung in a quark-gluon plasma or in a hadronic gas, for example. But there are some situations when the photons are produced by sufficiently slow particles but we should take into account of the influence of the matter. This takes place at the collisions of nonrelativistic protons ( the proton energy is about some hundreds of MeV ) with nuclei l . When the energy of the proton is more then the Fermi energy of a nucleus ( it is correct for not too massive nuclei) the proton penetrates into the such nucleus and generates photons due to multiple collisions with the nucleons of the nucleus. 2
The probability of photon production.
Let us consider the proton undergoing multiple elastic collisions in the matter. We assume that the proton energy is much less the its mass so that the proton can be considered as a nonrelativistic particle. In this case we can neglect by the effects of photon recoil. Then the probability of photon production (h=c=l) is given by the following expression :
d £ j = BRe{j0TdtI0T
^reM-i^nnvit)]
•[ n ^ + r)]|.,
(1)
where ui is the photon energy; dfl is the solid angle along the direction of the vector n = k/k; k is the wave vector of the photon ); a is the fine structure constant; ( v(t) and v(t + r) are the velocities of the proton before and just 427
428
after the photon production, respectively, T is the time of proton movement in the matter. To obtain the probability dW of the photon production in the matter which can be observed we should average Eq.(l) over the all possible v(t) and v(t + T) in the matter :
£; = (as;) - £*{/«" C *C "-x^ [n,v(t)} • [n,v(t + T)] -F(i?,t;i?,t
+ T)\.
Function F(v, t;v',t + T) is the so called conditional probability. It means the probability that the particle velocity is v1 at the time t + r provided that the velocity is v at the time t. If the process of particle scattering in the matter is Markovian one and , the function F can be represented in the form : F(v,t;v',t
+ T)=g1(v,t)-f1(if-v,t
+ T),
(3)
where the functions g(v,t) and /(«* - v, t + T) satisfy the Vlasov-UehlingUhlenbeck ( VUU ) kinetic equation. Taking into account of the law of energy conservation the last formula can be rewritten as follows : F{v, t-v1 ,t + T) = g(v,t) • fiv1 - v,t + T) • 6{v - vp) 6(V' - «Q • (1 - eF/E v2 ' v'2
-
(4)
w/E)W))
where 6 - functions take into account the conservation law in the processes of scattering and the photon production; vo is the proton velocity before the photon emission; E is its energy; ep is the Fermi energy of a nuclear. When the energy of proton much more than the energy of nucleons of not too massive nuclei we can go from the VUU equation to the standard Bolzmann equation. Then function /(£* — v, t + r) satisfies the following equation: a/(cosg,r) dr
\ f da(cos(6' - 6>) tl , f( A a, 0 1/ a f(cos6 ,r) - / ( c o s ( 9 , r ) ^ ,
,,, (5)
(2)
429 where a is the cross section of the elastic scattering of nucleons; da is the corresponding differential cross section; v is the collision frequency in the matter. The equation for the function g is analogous one. The last equation can be rewritten in the following form :
<9/(cos6», ,A y _ dr
, ,
• HI:
^ ^ o s ^ / _ 6>))d(cos<9')/(cos6l',x) - /(cos6«,-r) }> , (6)
where
X (cos(0'
- 6)) =
K
—±
'-
(7)
Solving Eq.(6) with the initial condition
/ ( T = O,/*) = < K A I - V )
(8)
by means of the expansion of function / over the Legendre polynomials Pi{n) we obtain :
f(^T)
=E
2Z + 1 - f - f l ( M ) f l ( / * ' ) e x p ( - i / r ( l - 2Xi/2l + 1)),
(9)
1=0
where Xi
21 + 1 f1 j ^ dfiX(»)Pi(n) 2
fi = cos 6
(10)
fj,' = cos 6'
The solution of the equation for the function g is of the analogous form : °° 21 + 1 g{n,t) = X ] - ^ - f i f o > ) W ) e x p ( - W ( l - 2 X ,/2J + 1)), i=o
(11)
Z
where the parameter fio is the cosine of the angle of the flying of the particle into the matter. Substituting Eqs.(3,4,10,ll) into the formula (2) we obtain :
430
UUJT w 0 2 ( l - HP-) -sin2(9w
dW
37T2
dudfl /
eF
w 2 _j_ j,2 . ( l _ 2>Q
/
eF
u/\d/2)
o-s-s)
• >12>
<j\d/2)
where #w is the emission angle of the produced photon, T)(X) is the unit function. Turning from the probability to the cross section day we have : da1 dwdfi
__aujcrvo2A 37r2
(l - ^f 1 ) • sin2 9U u2 + 1/2 • (l- ^ i ) 2
^ ^
eF y \ W 2 ) • (13) E EJ I
where A is the number of nucleons in the matter. The main differences between the results obtained above and the spectral distribution in the case of the Landau-Pomeranchuk ( LP ) 2 ' 3 effect are following. At the first, da1 ~ to at small u> whereas one is proportional to w - 1 / 2 in the LP effect. At the second, the spectral distribution in the LP effect is monotonous function of the photon energy, while in the studied situation da-y has a maximum. References 1. M.J. van Goethem, H. Loehner, H.W. Wilschut, Quenching of bremsstrahlung in proton-nucleus reactions, KVI Annual Report 1999, 10. 2. L.D.Landau, I.Ya.Pomeranchuk. Dokl. Akad. Nauk SSSR, 92 735 (1953). 3. A.B.Migdal. Phys. Rev. 103, 1811 (1956).
PRONOUNCED AIRY-STRUCTURE IN ELASTIC lsO+12C SCATTERING ATE lab=200MEV Yu.A. GLOUKHOV, A.S. DEM'YANOVA, A.A. OGLOBLIN, M.V. ROZHKOV RRC «Kurchatov Institute*, 123098 Kurchatov sq. 1, Moscov, Russia glukhov @ dni.polyn.kiae.su S.A. GONCHAROV Nuclear Physics Institute, Moscow State University, Moscow, Russia [email protected] R. JULIN, W.H. TRZASKA Department of Physics, Jyvaskula University, Finland [email protected] Investigation of refractive effects in nuclear scattering presents opportunity to gain information on heavy ions potential at small internuclear distances. The system 16 0+12C was chosen to observe a developed Airy-structure. The measurements were carried out at energies 8-18 MeV per nucleon with a step 30 MeV. Especially prominent Airy-structure was observed at energy 200 MeV. Optical model analysis has made a goodfittingof angular distribution at 200 MeV. It is shown that an intensity of Airy-structure depends on the mutual action of refractive and absorption effects. The angular distribution of inelastic scattering at the same energ demonstrated a weak Airy -structure. The explanation of this difference is presented. 1 Introduction Previous investigation of refractive (rainbow) phenomena in nuclear scattering were carried out for the symmetric systems 1 6 0+ 1 6 0 and 12C+12C [1,2]. They allowed to observe Airy-structure within 90 cm. deg. The 16Ot-12C system does not have such a restriction and was chosen as a suitable candidate for a detailed study of the refractive effects. Especially prominent Air -structure was observed at energy 200 MeV. Such an expressive Airy-minimum was explained by the fine balance between refraction and absorption effects. 2 Experiment Experiment was carried out at the cyclotron of Jyvaskula University (Finland) at energy 200 MeV. An intensity of the bea was 100 nA, the energy resolution around 0.3%, the diameter of the 2 beam spot 2x3 . The target was a self-supporting carbon foil of 2 0.24 g/cm . The experimental setups of semiconductor detectors were 431
432
described in [3]. The accuracy of the absolute cross measurements was 15%.
-section
3 Results The angular distribution of elasti 16 0+ 12 C scattering at energy 200 MeV is displayed in Fig. l.The curve demonstrates Frauenhover diffraction oscillations at forward angles and the deepest minimum at 65 deg. cm. followed by broad maximum ending in approximatel exponential fall -off. The intensity of this Airystructure has no Fig.1 analogues in the world Elastic scattering at 200 MeV (experiment and theory) systematic of 1 6 0+ 1 6 0 12, and 1ZC+1ZC elastic scattering. The result of the optical model analysis calculation is presented in Fig.l as well. Woods-Saxon shape optical potential wit a shallow imaginary part and a deep real one has made a good fitting. The origin this deepest minimum can be understood in the frame o considerations given in [4].In Fig.2 a deflection function is displayed [5]. It is a dependence scattering angle on an impact parameter (angular momentum L) . The positive angles represent deflection caused b Coulomb scattering Fig.2. (reflection). The Deflection function attractive nuclear forces results in deflection to negative angles (refraction). Every value of the negative
433
angle is corresponded 2 values of angular momentum called L < and L>. Airy oscillations arise from interference of contributions from 2 sides of deflection function, at angular "CH-"C momenta L< and L>. E( O)=200MeV elastic The deflection inelastic (experim function is normally inelastic (theory) very asymmetric: its external branch is much steeper than the Fi 3 e de 9 internal one. As semi16
nelastic scattering at 200 MeV (experiment and theory)
classical cross-section is inversely proportional to [L/dO/dL], the equal contribution from both branches of the deflection function takes place at rather small L -values resulting in production of the deepest minimum if the absorption is not taken into account. The absorption dumps the L< component shifting the deepest minimum to the larger impact parameters. Differential cross-section of inelastic scattering o x O at energy 200 MeV leading to the excitation of 4.43 MeV 2C level is presented i Fig. 3 together with the elastic scattering data given for comparison. The phase rule is excellently fulfilled in the domain of Frauenhover diffraction. Inelastic angular distribution demonstrates some change of a slope at the angles corresponding to the primary Airy -minimum and one can speak about the remnant of a rainbow maximum at the larger angles. However, a striking difference between elastic and inelastic cross-section exhibits in the depth of Airy -minimum which is 2 orders of magnitude deeper in the case of elastic scattering. The analysis of the inelastic scattering was performed using both DWBA and coupled channel (CC) calculations with the phenomenological Woods-Saxo potential. The calculation have been done with the code ECIS in the approximation of the first order vibrational mode with complex coupling. The radial shape of the inelastic form-factor was taken as a derivative of the complex potential. The conventional DWBA calculation based on these statements reproduce the data quit well (full line in Fig. 3).
434
So the refractive behaviour of the inelastic cross -section is evident. However, the question remains, why such strong difference in the sharpness of the rainbow minima in the elastic and inelastic angular distributions takes place. A qualitative answer based on considerati of deflection function can be given. The sharpness of the minimu under discussion came as a result of a fine balance between the slope of two branches of the deflection function and different absorption for L< and L>. Any change of this balance will make the minimu shallower. The form-factor can either enhance or suppress the partia waves forming the Airy-minimum, but due to its radial dependence its influence will be different for L< and L> practically in all cases. 4 Conclusion Differential cross-sections of elastic and inelastic 16 0+ 12 C scattering at energy 200 MeV have been measured. The angular distribution of elastic scattering demonstrates the minimum of the first order at 65 deg. cm. much deeper that known analogues in heav -ion nuclear rainbow scattering. Six -parameter Woods-Saxon optical potential has described angular distribution of elastic scattering quite satisfactory. The reason of forming of the deepest minimum is determined by the fine balance between absorption and the interference of the waves from the both sides of the deflection function. The striking difference in a sharpness of the rainbow minima in elastic and inelastic angular distributions is mainly explained the same reason. This work has been supported by the Academy of Finland. References: 1 F. Stiliaris et al., // Phys. Lett. B 223, 91 (1989) 2. S. Kubono et al., // Phys. Lett. B 127,19 (1983) 3 A.A Ogloblin et al.// Phys. Rev. C 57 1797 (1998), 4 K.W.McVoy et al., // Nucl. Phys. A 417,157 (1984) 5 G.R. Satchler, Nucl. Phys. A 409 (1983) 3c
FUSION ENERGY THRESHOLDS PREDICTED WITH ADIABATIC NUCLEUS-NUCLEUS POTENTIAL
AN
J. W I L C Z Y N S K I Institute
for Nuclear Studies, 05-400 Otwock-Swierk, E-mail: [email protected]
Poland
K. S I W E K - W I L C Z Y N S K A Institute
of Experimental
Physics, E-mail:
Warsaw University, 00-681 [email protected]
Warsaw,
Poland
Experimental values of the fusion energy-threshold, denned as the energy at which the fusion cross-section equals the s-wave absorption cross-section, are compared with barrier heights calculated for the adiabatic fusion potential of the WoodsSaxon shape with parameters unambiguously determined by the liquid-drop-model contact force and the energy of the compound nucleus.
1
Introduction
Precise measurements of fusion excitation-functions provide information on the distributions of the fusion barrier 1 which can be interpreted as a manifestation of coupling to various collective states, effectively enhancing the fusion probability at sub-barrier energies (see recent review article 2 and references therein). On the other hand, on the grounds of the macroscopic dynamical models, the fusion-barrier distributions reflect the coexistence of different dynamical p a t h s through diversified shape sequences - associated predominantly with formation of a "neck" in the contact zone. The low-energy end in an experimentally observed fusion-barrier distribution can be related to the adiabatic sequence of shapes of the fusing system, presumably corresponding to a smooth transition of the nuclear potential energy and the interaction force from the initial situation of the two nuclei interacting in contact configuration to the final state of the mononuclear system with the energy of the compound nucleus.
2 2.1
T h e a d i a b a t i c fusion p o t e n t i a l Nuclear
potential
The potential energy of a nucleus-nucleus system undergoing fusion can be represented in multi-dimensional-configuration space, depending of course on the particular choice of shape parametrization. For determination of the adi435
436
abatic nucleus-nucleus fusion potential, we avoid the question of assuming specific shape parametrization. Instead, following the idea of Refs. 3,4 , we use only the known characteristics of the system at the beginning of nuclear interaction (contact force) and in the final state of the equilibrated compound nucleus. Smooth interpolation between these two reference points is done without free parameters assuming that the effective one-dimensional potential has the Saxon-Woods shape, Vn[r)
•V0
l
(1)
e a .p(S=£a)
+
The depth of the nuclear potential, Vo, is determined by the ground-state energy of the compound nucleus (with its intrinsic Coulomb energy, Ccn, and shell correction, Scn, subtracted) taken relative to the sum of the ground-state energies of the two separated nuclei, also with subtracted intrinsic Coulomb energies C\ and C2, but shell corrections included: Vo = (M x + M 2 - Mcn)c2 + Ccn -d-C2
+ Sc.
(2)
where Mi, M2 and Mcn are the ground state masses, and Cc.
Ci - C*2 = Co = 0.7053
(Zi + Z2f
(A, + Atfl*
A\'3
,1/3 1 2 J
MeV.
(3)
Here, the Coulomb energy constant, Ecoui — 0.7053 MeV, is taken from the standard liquid-drop-model fit to nuclear masses 5 . The shell correction Scn in Eq. (2), which we take from Ref. 6 , has to be subtracted from the ground state energy of the compound nucleus because it produces only a local dip (near the equilibrium shape) in the flat landscape of the nuclear potential energy represented by the inner part of the Saxon-Woods potential. The diffuseness parameter a in Eq. (1) is determined by the strength of the nucleus-nucleus attractive force in the contact configuration RQ = Ri+R2, calculated in frame of the liquid-drop model 7 : \dr)r=R0
4
«
^Rl
+ R: /V
(4)
where 7 = 0.9516 1 - 1.7826
W
MeV/fm 2
(5)
is a value of the surface tension coefficient 5 of the combined system with the number of neutrons N — N\ + N2 and the number of protons Z — Z\ + Z2.
437
Calculating the liquid-drop-model contact force, the radii Ri and R? of the two nuclei should be interpreted as "equivalent sharp radii" 8 , scaled approximately as roiA1'3 + A1'3) with r 0 (a 1.15 fm. Having determined the parameters of the nuclear part of the adiabatic fusion-potential, Eq. (1), a comment should be made on the meaning of variable r in this one-dimensional potential. For large values, r > Ro, its interpretation is clear: r can be identified with the distance between the centers of the two nuclei. However, for r < RQ where fusion proceeds through a sequence of shapes in at least 3-dimensional space representing the distance, neck and asymmetry degrees of freedom 9 , the variable r becomes an effective parameter along the trajectory in this multi-dimensional space, leading from the configuration of two nuclei in contact to the equilibrium shape of the compound nucleus. 2.2
Coulomb potential
For typical applications, such as determination of the fusion barrier, it is sufficient to consider the Coulomb potential only in the outer region, r > Ro, where the point-charge approximation gives sufficient accuracy, Z1Z2e
(6)
For completeness, in the inner region, r < Ro, we propose to use a simple one-dimensional parametrization which joins the known value of the Coulomb energy (and its derivative) in the contact configuration with the Coulomb energy in the equilibrium shape, Vc(r = 0) = Ccn — C\ — C2 = CoThe highest point along the fusion trajectory, defined by the condition d[Vn + Vc)/dr = 0, determines the height of the fusion barrier in central collisions, Bful. As it was argued above, the nuclear potential (1) is expected to describe the adiabatic evolution of the system towards fusion, presumably corresponding to the lowest possible barrier. Therefore, we are going to compare our theoretical adiabatic barriers with the lowest barriers in the fusion barrier distributions observed experimentally. 3
Fusion energy t h r e s h o l d s
Fusion-barrier distributions can be deduced from precisely measured fusion excitation functions by taking the double derivative of the product of the crosssection multiplied by energy, d2(aE)/dE2, as proposed by Rowley, Satchler and Stelson 1 . In most of the studied reactions, the deduced fusion-barrier
438
distribution extends over a range of 5 - 15 MeV. As stated above, the theoretical adiabatic barrier, calculated for a given reaction, should be compared with the low-energy end of the distribution. Since, however, the position of this low-energy edge cannot always be determined with sufficient accuracy, we propose to use the following operational definition of the fusion energythreshold: This threshold is the energy E = Ethr at which the measured fusion cross-section is equal to the s-wave absorption cross section: 2
nh2
(7)
where A is the wavelength of the fusing system, and (i its reduced mass. We have checked t h a t for all excitation functions measured with sufficient precision, the fusion energy-threshold, determined according to the above criterion, perfectly coincides with the low-energy edge of the fusion-barrier distribution. Therefore, in order to reduce uncertainties in determination of the experimental fusion thresholds, in our analysis we compare the theoretical adiabatic fusion barriers with the fusion energy thresholds, Ethr, deduced from experimental fusion excitation functions by consistent use of Eq. (7). Fig. 1 shows the comparison of the experimental fusion energy thresholds with barrier heights calculated with the adiabatic potential. D a t a for about 50 systems, taken from R e f s . 1 0 - 2 2 are included in this comparison. For all these systems, the sub-barrier part of the excitation function has been measured with high precision, at least down to the threshold limit given by Eq. (7), which for heavy systems is of the order of 0.1 mb. However, light systems and also some heavier systems involving light projectiles of A < 28, are not included in the comparison because the concept of the liquid-drop-model contact force, Eq. (4), does not apply to such light nuclei. As it is seen from Fig. 1, the calculated adiabatic barrier heights are very well correlated with the experimental fusion thresholds. For comparison, the barriers calculated with the Bass potential 2 3 are also shown in Fig. 1. Obviously, they are much higher t h a n the experimental fusion thresholds because parameters of the Bass potential are chosen to fit the mean, single barrier values. Moreover, it is important to note t h a t the correlation between the experimental fusion thresholds and Bass fusion barriers is worse t h a n in the case of the adiabatic barriers. This is very likely due to nuclear-structure (isotopic) effects which are not accounted for in the Bass potential, while they are present in the adiabatic potential (via the ground-state masses, determining Vb and thus influencing the fusion barrier). It is seen from Fig. 1 t h a t the calculated adiabatic barriers are systematically slightly smaller t h a n the experimental fusion thresholds. T h e difference
439
0
50
100
150
200
250
Experimental fusion threshold (MeV) Figure 1. Fusion barriers calculated with the adiabatic fusion-potential and the Bass potential, compared with the fusion energy-thresholds deduced from measured 1 0 _ 2 2 fusion excitation-functions.
increases with the barrier height and has a clear interpretation: it is just an average dissipative loss of kinetic energy of the fusing system during its approach to the top of the barrier. To verify this hypothesis, we calculated the dissipative loss of kinetic energy for several systems, using the code HICOL of Feldmeier 2 4 . Very good agreement between the calculated dissipative energy losses and the Ethr — Badiab differences was obtained. This result shows t h a t our prescription for calculating the adiabatic fusion barriers is consistent with predictions of the dynamical effects along the fusion trajectories. T h e knowledge of the exact location of the fusion energy thresholds is essential for planning future experiments on synthesis of new super-heavy elements. Predictions based on the adiabatic fusion potential look promissing because the calculated adiabatic barriers are sufficiently low and do not prevent "cold fusion" of very heavy systems. Moreover, the calculations show a clear trend of lowering the adiabatic barrier with respect to the saddle-point energy for very heavy systems, an effect suggested by Myers and Swiatecki 2 5
440
as an explanation of a larger than expected fusion cross section in the recent successful synthesis of the element Z=118 2 6 . Acknowledgments This work was supported by the Poland-USA Maria Sklodowska-Curie Joint Fund II, under Project No. PAA/DOE-98-34. References 1. 2. 3. 4. 5. 6. 7. 8. 9. 10. 11. 12. 13. 14. 15. 16. 17. 18. 19. 20. 21. 22. 23. 24. 25. 26.
N. Rowley, G.R. Satchler and P.H. Stelson, Phys. Lett. B 2 5 4 , 25 (1991). M. Dasgupta et al, Ann. Rev. Nucl. Part. Set. 4 8 , 401 (1998). J. Wilczynski and K. Siwek-Wilczynska, Phys. Lett. B 55, 270 (1975). K. Siwek-Wilczynska and J. Wilczynski, Phys. Lett. B 74, 313 (1978). W.D. Myers and W.J. Swiatecki, Ark. Fys. 36, 343 (1967). Moller et al, At. Data Nucl. Data Tables 5 9 , 185 (1995). J. Wilczynski, Nucl. Phys. A 2 1 6 , 386 (1973). W.D. Myers, Nucl. Phys. A 2 0 4 , 465 (1973). J. Blocki and W.J. Swiatecki, Report LBL-12811, Berkeley, (1982). M. Beckerman et al, Phys. Rev. Lett. 4 5 , 1472 (1980). M. Beckerman et al, Phys. Rev. C 25, 837 (1982). H.A. Aljuwair et al, Phys. Rev. C 30, 1223 (1984). W. Reisdorf et al, Nucl. Phys. A 4 3 8 , 212 (1985). W. Reisdorf et al, Nucl. Phys. A 4 4 4 , 154 (1985). A.M. Stefanini et al, Nucl. Phys. A 4 5 6 , 509 (1986). W.Q. Shen et al, Phys. Rev. C 36, 115 (1987). A.M. Stefanini et al, Phys. Rev. Lett. 7 4 , 864 (1995). A.M. Stefanini et al, Phys. Rev. C 5 2 , R1727 (1995). J.D. Bierman et al, Phys. Rev. C 54, 3068 (1996). H. Timmers et al, Phys. Lett. B 3 9 9 , 35 (1997). A.M. Stefanini, J.Phys. G: Nucl. Part. Phys 2 3 , 1401 (1997). A.A. Sonzogni et al, Phys. Rev. C 57, 722 (1998). R. Bass, Nucl. Phys. A 2 3 1 , 45 (1974). H. Feldmeier, Rep. Prog. Phys. 5 0 , 915 (1987) W.D. Myers and W.J. Swiatecki, Acta Phys. Pol. B 3 1 , 1471 (2000). V. Ninov et al, Phys. Rev. Lett. 8 3 , 1104 (1999).
N E A R - B A R R I E R F U S I O N OF 3 6 S + 9 0 96 Zr: W H A T IS T H E EFFECT OF T H E S T R O N G OCTUPOLE V I B R A T I O N OF 96 Zr ?
Istituto
L. C O R R A D I , A . M . S T E F A N I N I , A . M . V I N O D K U M A R Nazionale di Fisica Nucleare, Laboratori Nazionali di Legnaro, Legnaro, Padova, Italy
1-35020
S . B E G H I N I , G. M O N T A G N O L I , F . S C A R L A S S A R A , M . B I S O G N O Dipartimento di Fisica, Universitd di Padova, and Istituto Nazionale di Fisica Nucleare, Sezione di Padova, 1-35131, Padova, Italy Near-barrier fusion cross sections have been measured for 3 6 S + 90 > 96 Zr to high statistical accuracy. Subbarrier fusion of 3 6 S 4- 9 6 Zr has a large enhancement compared to 3 6 S + 9 0 Zr. The barrier distribution has one main peak for 3 6 S + 9 0 Zr, and a wider and more complex shape for 3 6 S + 9 6 Zr. Coupled-channels calculations indicate t h a t the subbarrier enhancement for 3 6 S + 9 6 Zr is due mainly t o the octupole vibration of 9 6 Zr. However, t h e calculations do not reproduce well the barrier distribution for this system. Comparing with 4 0 C a + 90 > 96 Zr suggests t h a t couplings to transfer channels may still play a role in the very large fusion enhancement for 4 0 C a + 9 6 Zr at low energies.
1
Introduction
The analysis of 4 0 Ca + 90>96Zr fusion excitation functions around the barrier1, showed that while 4 0 Ca + 90 Zr data are well reproduced by CC calculations including inelastic excitations, the large sub-barrier enhancement and broad barrier distribution of 4 0 Ca + 96 Zr are not accounted for by the same model. That was taken as an indication that nucleon transfer effects are important in the latter system where they might be favoured by the very positive Q-values, whereas only negative Q-values are available in 4 0 Ca + 90 Zr. However, a strong and low-lying octupole vibration exists in 96 Zr; hence it is important to clear up experimentally its effect on subbarrier fusion. This may be accomplished by using a projectile ( 36 S) which is as rigid as 40 Ca, with a similar atomic number but unfavourable transfer Q-values with 96 Zr. The further study of 36g _j_ 902 r g i v e s a consistency check since its fusion cross section should be similar to those of 4 0 Ca + ^ Z r . 2
Experiment and results
The experiment was performed at the XTU Tandem accelerator of LNL with the 36 S beam energies in the range 99.2 MeV to 123.0 MeV. The targets were evaporations of 9 0 ZrO 2 (60^g/cm 2 ) and 96 Zr203 (90jUg/cm2) on Carbon back441
442
ings 20/xg/cm2 and 15/xg/cm2 thick, respectively. Four silicon detectors at about 16° around the beam direction were used for monitoring and normalization purposes, while the evaporation residues were detected at 0°, after beam rejection by means of an electrostatic separator 2 , and further discrimination from beam-like particles by a E-TOF telescope. Angular distributions were measured at E(, eam = 107.6 MeV and 118.1 MeV. The extracted cross sections are plotted in Fig. 1 together with those of 4 0 Ca + 90 ' 96 Zr, in a reduced energy scale. It appears that 36 S + ^ Z r behaves much like 4 0 Ca + 90 Zr while 36 S 4- 96 Zr (benefiting from a stronger octupole coupling, but lacking the transfer coupling) is intermediate between 4 0 Ca + 90 Zr and 40 Ca + 96 Zr.
! . . , . ! > ! . .*ȣ$ &*
0.1
s 0£j
«***°*^**" :
l
s /
0.01
•
a
0~ *
0.001
•
B
r . '
00001 r
o T
*
V
. *.
36„
96v
S 36,,
+ 90.-,Zr S + Zr Ca + Zr
4()„
90_
Ca+
i
> 0.95
.
.
i
Zr .
.
_ .
.
1
E/V
FIG. 1.Fusion excitation functions of36S + 9 0 ' 9 6 Zr (open and full circles respectively) and 40 Ca + ^"'^Zr (open and full triangles). The code CCFULL was used for coupled channels calculations3. It uses the isocentrifugal approximation and the incoming-wave boundary condition inside the barrier. The finite excitation energies of the coupled modes are taken into account fully, as well as the effects of inelastic non-linear couplings to all orders. Vibrational couplings are treated by CCFULL in the harmonic oscillator limit. The relevant channels for the CC calculations are the 2 + (3.291 MeV, (32 = 0.16) and 3~ (4.192 MeV, /33 unknown) states of the 36 S projectile and the analogous excitations of the target nucleus. The parameters in 90 Zr are 2.186 MeV, /?2=0.09 (2+ level) and 2.748 MeV, /?3=0.22 ( 3 " level); in 96 Zr they are 1.751 MeV, /?2=0.08 (2+ level) and 1.897 MeV, /?3=0.34 ( 3 " level).
443
Two-phonon channels were also considered, i.e. the 2+
7*
7S
E
78
60
82
(MeV)
FIG. 2.Excitation function (up) and barrier distribution (down) of36S + 90Zr, in comparison with CC calculations. Only statiscal error bars are shown. Fig. 2 shows the CC calculations and the experimental excitation function and barrier distribution of 36 S + 90 Zr. The dashed lines, labelled "one phonon", result from including the one-phonon states and the 2+
444
in this case it is essential for a good fit of the excitation function. However the calculations give only a marginal fit of the barrier distribution; the full line in Fig.3 (two-phonon calculation) reproduces correctly only the overall width of the barrier distribution and the position of the main structure.
-02 r 70
72
> . • .•..»... . * ,\ 74 7S 78
E
60
• l.j 82
(McV)
FIG. 3. Excitation function (up) and barrier distribution (down) of36S +
g6
Zr.
We can conclude that the octupole vibration in 96 Zr has a strong effect on subbarrier fusion, in agreement with the CC calculations. The conclusion of ref.1 that the difference between 4 0 Ca •+• 90 Zr and 4 0 Ca + 96 Zr can be attributed to the coupling of nucleon transfer is strengthened, even though the strong octupole vibration of 96 Zr does explain part of the difference. References 1. H.Timmers, D.Ackermann, S.Beghini, L.Corradi, J.H.He, G.Montagnoli, F.Scarlassara, A.M.Stefanini and N.Rowley, Nucl.Phys.A633 (1998) 421 2. S.Beghini, C.Signorini, S.Lunardi, M.Morando, G.Fortuna, A.M.Stefanini, W.Meczynski and R.Pengo, Nucl. Instr. Meth. A239 (1985) 585 3. K.Hagino, N.Rowley, A.T.Kruppa, Comp.Phys.Comm. 123 (1999) 143
DYNAMICAL MODEL OF FISSION FRAGMENT ANGULAR DISTRIBUTION V.A.DROZDOV, D.O. EREMENKO, O.V. FOTINA, S.YU. PLATONOV ANDO..A. YUMINOV Institute of Nuclear Physics, Moscow State University, Moscow, Russia E-mail: [email protected] The dynamical model of fission fragment angular distributions is suggested. The model allows one to calculate fission fragment angular distributions, prescission light particle multyplicities, evaporation residue cross sections etc. for the cases of decay of hot and rotating heavy nuclei. The experimental data on angular anisotropics of fission fragmenta and neutron multiplicities are analyzed for the l s O + 208Pb, 232Th, M8Cm and 238U reactions at the energies of the incident "O ions ranging from 90 to 160 MeV. This analysis allows us to extract both the nuclear friction coefficient value and the relaxation time for tilting mode. It is also demonstrated that the angular distributions are sensitive to deformation dependence of the nuclear friction coefficient
In the present contribution it is suggested the dynamical model of fission fragment angular distributions which takes into account stochastic aspects of nuclear fission. The model may be useful tools in analysis of the experimental data on fission fragment angular distributions, prescission light particle multyplicities, evaporation residue cross sections etc. for the cases of decay of hot and rotating heavy compound nuclei when a validity of the traditional transition state models is questionable. The dynamics of induced nuclear fission is considered in the stochastic approach [1] by the one dimensional Langeven equations. The potential energy of fissioning nucleus is evaluated within the liquid drop model by means of the fast method suggested in [2]. The initial spin distribution was calculated using the parametrization of which is based on the surface friction model [1]. Light particle emission was simulated by means of the Monte-Carlo method commonly used in the Langeven simulations of induced fission [1]. The calculations of the fission fragment angular distributions were performed assuming that for every Langeven sample completed by fission there exists a transition point and the nuclear properties (temperature and deformation) at this point determines the angular distribution of fission fragments. However, the stochastic nature of fission results in an ensemble of such points by which the averaging has to made. The position of every transition point was determined by the relaxation time of tilting mode r which was treated as a free parameter. In this case for every Langeven sample we
445
446 -,
20 -
/=»,v<*/»-"c ,
J*30
xM'If'c
10 0-
-10 -20
r/R0
1.5 r/R,,
r/R,,
Figure 1. The transition point distributions over deformations and the fission barriers.
fixed the temperature and the inertia moments at the point in time t' which was determined by the relations nz
447
seen that for the O + Pb reaction the experimental data are between two calculations within the model of transition states in the saddle point. It means that l6 o "WR> 0-PTn o 2.8 4.0 ©
2.4
3.0100
2.0 120
120 140 160
140
o o
o
160 E l a b , MeV
2.4 H W O n 2.0 1.6 H 1.2 100 120 140 160 E i a b , MeV
Figure 2. The angular anisotropy of fission fragments. The points are the experimental data, the solid lines are the calculation result within the present model, the.dashed-dotted and dashed lines are the calculation within the model of transition states in the saddle point with and without consideration of presaddle neutron emission, respectively.
the transition state model in principle can describe the experimental angular anisotropy for this reaction in the energy range under consideration. However it seems, that our calculations reproduce the slope of the experimental energy dependence of the angular anisotropy better than the transition state model. On the other hand the 160+232Th, 160 + 248Cm and 160 + 238U reactions lead to compound nuclei which are heaver than one in the case of the ,6O+208Pb reaction. As is seen from Fig. 3 for these cases the angular distributions offissionfragments can not be reproduced in the frames of the standard transition state model. It is connected with the formation of compound nuclei having angular momentum dependent fission barriers are smaller than temperature values for the considerable part of the initial distributions over angular momentum. It is seen that for the 160+232Th, 16 0 16 238 248, Cm and 0 + U reactions our model allows good descriptions of the experimental data. In the frames of the model angular distribution of fission fragments is connected with the distribution of the transition points over deformation. It is obviously that probabilities of realization of each transition point are determined by the time scales for different stages of fission. The last means that angular distributions of fission fragments have to be sensitive to the deformation dependence of nuclear viscosity. In the present work we used the one-body (only
448
wall contribution) and two-body nuclear dissipation model to illustrate this sensitivity. In these calculations for the deformation dependent inertia coefficient 0.08 -i one-body
•
J two-body dissipation
0.5
,
1.5
2.5
rM0 Figure 3. The transition point distribution over deformation for the one-body and two-body dissipation models with models with rK= »xl0~zls and J=40ti for the 248Cf nucleus at E*=110 MeV 2 21 the Werner-Wheeler method. For the relaxation time r„ 7.xlCT ' sand 9x1ff s were obtained for the one- and two-body models respectively. In Fig. 4 it is presented the distributions of transition point positions over deformation for the one- and two-body nuclear dissipation models. As is seen for the two-body viscosity the average transition point deformation is shifted to lower deformation relative to the one-body value. This is a result of that the two-body viscosity coefficient is larger then the one-body coefficient for presaddle deformations but for saddlescission the relation between these coefficients is opposite [1]. All of the preceding allows us to conclude that the angular distributions of fission fragments are sensitive to the deformation dependence of the nuclear friction coefficient and the model suggested in the present work is a useful tool to study this dependence. We conclude that the dynamical model may be useful tools in analysis of fission fragment angular distributions for the cases of decay of hot and rotating heavy compound nuclei when a validity of the traditional transition state models is questionable. Analisys of the experimental data for a set of complete fusionfission reactions allowed one to obtaine the relaxation time for tilting mode. It was demonstrated that the angular distributions are sensitive to deformation dependence of the nuclear friction coefficient.
References Frobrich P., Gontchar I.I. Langevin description of fusion, deep-inelastic collisions and heavy-ion-induced fission. Phys. Rep. 292 (1998) pp.131-237. Lestone J.P. Fast method for obtaining finite range corrected potential energy surfaces. Phys. Rev. C51 (1995) pp.580-585. Back B.B., Berts R.R., Gindler J.E., Wilkins B.D., Saini S., Tsang MB., Gelbke C.K., Lynch W.G., McMahan M.A., Baisden P.A. Angular distribution in heavy-ion inducedfission.Phys. Rev. C32 (1985) pp. 195-213.
FUSION - FISSION REACTION UC, uO + M8PB AT SUBBARRIER ENERGIES S. P. TRETYAKOVA M. G. ITKIS, E. M. KOZULIN, N. A. KONDRATTEV, I. V. POKROVSKTI AND E. V. PROKHOROVA Flerov Laboratory of Nuclear Reaction, JINR, Dubna, Russia E - mail:tsvetl@ sungraph.jinr.ru L. CALABRETTA AND C. .MAIOLINI, Laboratorio Nazionale del Sud, INFN, Catania, Italy A. YA. RUSANOV, Institute of Nuclear Physics, Alma-Ma, Kazakhstan T. YU. TRETYAKOVA Frank Laboratory of Neutron Physics, JINR, Dubna, Russia The cross sections and angular distributions of fission fragments have been measured for the compound nucleus 220Ra and 224Th produced in the reactions 12C, 16O+20SPb at three energy values in the ranges from -90.0 to -54.0 MeV for 12C and from -78 to 68 MeV for l s O.
1
Introduction
A great number of experimental data obtained in recent years on subbarier fusion cross sections of heavy ions with a wide range of A and Z compound nuclei have stimulated an extensive discussion on the mechanism of the subbarrier reaction enhancement. Earlier this type of investigations has been carried out only using reactions with light charged particles [1]. The study of deep subbarrier reactions using Pb and Bi nuclei is of special interest, since they make a basis for the synthesis of superheavy elements in cold fusion reactions, and are daughter nuclei in the case of cluster decay of nuclei from Ra to U. The employed methods allow one to investigate deep subbarrier reactions with the cross section level of up to 10"36, which opens up new perspectives for analysis of the possibility of new elements synthesis as well as for the study of inverse reactions of cluster decay. 2
Experimental procedure
Beams of 12C obtained at the tandem accelerator (NFTN, Catania, Italy) were focused onto a 208Pb target located in the center of a scattering chamber. Scheme of the reaction chamber is presented in [4]. A target of 130 |xg/cm2 of 208Pb (99.1 % 449
450
enrichment) was evaporated onto a -50 |xg / cm2 carbon backing. The target was oriented with a carbon side facing the beam. Single fission fragments were detected in the backward angular ranges of 30° - 120° and 198° - 290° by using mica dielectric detectors with an area of 170 cm2. The detectors were located 13 cm away from the target. The irradiated mica detectors were annealed during 6 hours at 460°C for decreasing the background events which arise in mica when the scattered 12 C ions interact with mica atoms and their compound nuclei give registered tracks. For scanning under optical microscope the mica was etched in 40% HF. The background of the fissile element contamination in the target and mica detectors was less than 1 event for the whole detector area. Table 1. Experimental conditions and results for reaction 12C + 20sPb.
Beam energy, Elab, MeV 89.7 72.8 61.9 58.9 56.7 55.7 53.9
Huence of 12 C, ions/cm2
Cross-section of 220 Ra fission, mb
1013 10" 10'4 10' 4 10" 10" 10"
1158.19 ±0.02 358.79 ±0.89 19.92 ±0.06 (7.68 ±0.07) 10"' (4.51 ±0.09) 10"3 (8.29 ± 0.74) 10"4 (3.09 ± 1.09) 10"5
2.36 2.25 1.42 3.56 9.27 3.26 5.47
Fission fragments anisotropy 2.10 1.70 1.29 1.20 1.20 1.40 1.19
±0.04 ±0.06 ±0.06 ±0.20 ±0.50 ±0.70 ±0.80
Table 2. Experimental conditions and results for reaction s O + 2 Pb.
Beam energy, Elab MeV 77.8 74.8 72.8 67.8 3
Huence of 10O, ions/cm2 1.40 1.35 4.10 5.60
1014 10" 10" 10"
Cross-section of Th fission, mb
224
4.30 ±0.1 (6.2±0.1)10 2 (2.3 ±0.1) 10"3 (3.0±2.0)10" 6
Fission fragments anisotropy 1.20 ±0.10 1.16 + 0.15 1.23 ±0.15 -
Results and discussion
The energies and fluences of the accelerated 12C beams, used for this experiment, the loss energy interval for passing through the target of 12C beam, the obtained cross sections and the angular anisotropy of the fission fragments of 220Ra* compound nuclei are presented in Table 1 and 2. The angle-integrated fusion-fission cross sections obtained into 12C+ 208Pb-> 220Ra and 16CH- 208Pb-> 224Th [2-4] reactions are shown in Fig.l (A, B). In Fig.lA the energy scale is presented as the
451
1
10 20 30 40 50 60 70 80
E*
(MeV)
-20
-10
0
10
E*-E*
20
30
(MeV)
A. Figure 1.
220
Ra and
224
Th fusion - fission excitation function. Points are «-12C; * 16 0 [5]; ° - [4]; A-[2].
I I I I I I I I I I I I I I I I I 505560667075808590
EJMeV)
70
75
80
85
Ec™.(MeV)
B. Figure2. The experimental anisotropy of fission fragment (solid points) as function of E c ra for and 224Th (B). The open points on Fig.2 (B) are results of previous measurements [5].
Ra (A)
452
excitation energy E* and in Fig.lB as a difference between the excitation energy and the Bass-barrier excitation energy E*-E* (Bass). The experimental values of the angular anisotropy of fission fragments presented in Table 1 were derived by extrapolation of the experimental angular distribution W (0) to the angles 9=90° and 9=180°, using expression W (0) = a + b cos20. The error of the value of the angular anisotropy is no more than 10%, the statistical and fitting error included. The experimental fission fragment anisotropics as a function of E cm are shown in Fig.2. 4
Conclusion
For the first time the cross section and angular distributions of fission fragments have been measured for the nucleus 220Ra and 224Th produced in the reaction 12 C, 16O+208Pb at three energy values in the range from 0.5 to 8.0 MeV for 220Ra and 5 to 15 MeV for 224Th below the fusion barrier calculated in terms of the Bass model. The performed measurements allowed one to obtain data for the region of interaction of heavy ions with nuclei characterized by the cross section of 220Ra and 224 Th being by 6 and 8 orders of magnitude lower than the geometric cross sections. In the investigated energy range of' C and 1 6 0 beam there is a plateau characterized by the relation W(180°) / W(90°) = 1.2 in the angular anisotropy of fission fragments. This implies that the available angular packet of harmonics, which govern the production cross section of a compound nucleus, is approximately the same. References 1. Majumdar N., Bhattacharya N., Biswas D.C., Choudhury R.K., Nadkarni D.M. and Saxena A. Reexamination of the nuclear orientation model of quasifission reactions to explain anomalous fragment anisotropics at subbarrier energies. Phys. Rev. C 53 (1996) pp. R544 - R548. 2. Marakami T.,. Sahm C. C , Vandenbosch R., LeachD. D., Ray A and Murphy M. J., Fission probes of subbarrier cross section enhancements and spin distribution broadening. Phys. Rev. C 34 (1986) pp.1353 - 1365. 3. Morton C. R., Hinde D. J., Leigh J. R., Lestone J. P., Dasgupta M., Mein J. C , Newton J. O. and Timmers H., Resolution of the anomalous fission fragment anisotropies for the 16 0 + 208Pb reaction. Phys. Rev. C 52 (1995) pp. 243 - 251. 1. Oganessian Yu.Ts., Itkis M. G., Kosulin E. M., Tretyakova S. P., Calabretta L. and Guzel T, JINR Rapid Communications, Investigation of the fusion-fission reaction 208Pb + 1 6 0, JINR Rapid Communications, • 1 [75] - 96, (1996) pp.123 -132. 2. Pustylnik B. I., Calabretta L., Oganessian Yu. Ts., Itkis M. G., Kosulin E. M., Tretyakova S. P., Tretyakova T. Yu., Anisotropy of fission fragment for the reaction 16 0 + 208Pb. JINR Rapid Communications, • 3 [89]-98 (1998) pp. 57 64.
SUB-BARRIER FUSION WITH A HALO NUCLEUS: THE 6He CASE
Istituto
M. T R O T T A Nazionale di Fisica Nucleare, Laboratori Nazionali di via Romea 4, 1-35020 Legnaro, Padova, Italy E-mail: [email protected]
Legnaro,
J.-L. SIDA, N. A L A M A N O S , F . A U G E R , A D R O U A R T , D . J . C . D U R A N D , A. G I L L I B E R T , C. J O U A N N E , V. L A P O U X , A. L U M B R O S O , F . M A R I E , S. O T T I N I , C. V O L A N T CEA SACLAY DSM/DAPNIA,91191 Gif-sur-Yvette, Cedex, France
A. ANDREYEV,D.L. BALABANSKI, N COULIER, G. GEORGIEV, M. HUYSE, G. NEYENS, R. RAABE, S. TERNIER, P. VAN DUPPEN, K. VIVEY IKS, University
IPNE,
Bucharest
of Leuven,
Leuven,
C. B O R C E A - Magurele, P.O.Box
Belgium
MG6,
Romania
J.D. H I N N E F E L D Centre de Recherche du Cyclotron, UCL Louvain-la-Neuve, Physics Department Indiana University, South Bend, 46615
Istituto
A. M U S U M A R R A Nazionale di Fisica Nucleare, Laboratori Nazionali Via Santa Sofia 44, 1-95129 Catania, Italy
IFU, Sao Paulo,
Institute
A. L E P I N E CP 20156, Sao Paulo,
R. W O L S K I of Nuclear Physics, Cracow,
Belgium Indiana
del
Sud,
Brazil
Poland
A large enhancement of the fusion-fission probability has been observed, at subbarrier energies, for the halo nucleus 6 H e compared to its stable partner 4 H e for reactions on an Uranium target. T h e experiment has been performed at Louvain-laNeuve (Belgium). In order to understand the origin of this enhancement, coupledchannel calculations have been performed by means of the code ECIS. The calculations are in good agreement with the fusion cross section for 4 H e and 6 H e above the barrier. Below the barrier, the fusion cross section for the halo nucleus is much larger than the one expected from the fusion calculation. As the calculations take explicitly into account the diffuse density of the halo nucleus, t h e observed enhancement must be originating from the very peculiar role of the unusual structure of 6 H e in the process.
453
454
1
Introduction
Fusion reactions at energies near and below the Coulomb barrier are very sensitive to the structure of nuclei and to the couplings of the relative motion of the colliding nuclei to several nuclear intrinsic motions 1,a . The process is described into the framework of coupled-channel calculations, which allow for a good understanding of sub-barrier fusion of stable nuclei 3 ' 4 . The main remaining questions concern the role of other processes such as transfer or break-up reactions, which could act as doorway to fusion, as well the effect of unusual structure of nuclei such as neutron skins or halos 5 . In this framework we have started a research program aimed at studying the effet of halos and skins on sub-barrier fusion. A large enhancement of the fusion-fission probability has been observed, at sub-barrier energies, for the halo nucleus 6 He compared to the stable partner 4 He. The fission cross section for the systems 6 ' 4 He+ 2 3 8 U has been recently reported 6 , where a detailed description of the experiment can also be found. We report here on a more detailed analysis of the data, which allows to extract the fusion cross section. 2
Experimental results and data analysis
The experiment has been performed at Cyclone, the Radioactive Nuclear Beam Facility of Louvain-la-Neuve (Belgium), using a halo beam of 6 He and its stable partner 4 He as projectiles impinging on an Uranium target. At energies around the barrier the Pu compound nucleus formed by fusion of the 6 ' 4 He+ 2 3 8 U systems fissions. Anyway, fission could also be triggered by an inelastic or transfer reaction. In such cases, the two fission fragments are accompanied by a third particle, a residue of the projectile. The two fission fragments and the third particle have been measured in coincidence through a set-up 6 with an angular coverage of about 70% 4TT (#=26.6-153.4 deg). The analysis for obtaining the fission and the fusion cross sections is based on multiplicities. The multiplicity is determined via software by giving a threshold value for each detector and counting the number of detectors that have fired. The spectrum so built contains the multiplicity 1, which mainly corresponds to elastic scattering, the multiplicity 2, corresponding to two fission fragments, and the multiplicity 3, corresponding to two fission fragments plus a third particle, the residue of the projectile. For the systems studied, the center of mass frame and laboratory frame are quite equivalent, so the two fission fragments are emitted back to back also in the laboratory. By selecting the multiplicity 2 and by requiring that the two fired detectors be back to back, the fission cross section has been extracted. By selecting the multiplicity 3 and
455
by requiring that, among the three detectors, two of them are back to back, the contribution of transfer and inelastic reactions to fission has been estimated. This contribution, which has been found negligible for the 4 He incident beam, is of the order of the 10% of the total fission cross section at the highest energies in the case of the weakly bound nucleus 6 He. This contribution has been subtracted to the total fission cross section in order to extract the fusion cross section. The fusion cross section for the systems 6 ' 4 He+ 2 3 8 U is shown in Fig. 1 (points). The data for fission of 4 He+ 2 3 8 U at above barrier energies 7 , coincident inside the errors with fusion, are shown for comparison. The cross sections for the two systems are rather similar at above barrier energies, but a large enhancement is observed at sub-barrier energies for the halo nucleus 6 He compared to the stable 4 He. The regular trend of the fusion cross section suggest that no structured hindrance at the barrier due to the break-up effect of the nucleus is observed. 3
Coupled-channel calculations
In order to understand the origin of this enhancement, coupled-channel calculations have been performed by means of the code ECIS 8 . The potential has been calculated by the double-folding method 9 using the density-dependent nucleon-nucleon interaction BDM3Y1 10 . The densities of 4 He and 238TJ are extracted from electron scattering measurements 11 , while for the 6 He nucleus a halo density was used 12 . The excited states 2 + , 4+, 6+ and 8+ of the U target and the first excited level 2+ of the 6 He projectile have been coupled into the calculations. As one can see in Fig. 1, the calculations (lines) are in good agreement with the fusion cross section for 4 He and 6 He above the barrier. Below the barrier, the fusion cross section for the halo nucleus is much larger than the one expected from the fusion calculation. As the calculations take explicitly into account the diffuse density of the halo nucleus, the observed enhancement must be originating from the very peculiar role of the unusual structure of 6 He in the process. 4
Conclusions and Perspectives
In summary, a large enhancement of the fusion probability below the barrier has been observed for the halo nucleus 6 He compared to its stable partner 4 He. Morevover, no hindrance effect due to the break-up of the nucleus has been evidenced. The next steps in our experimental program, aimed at studying the effect of the halo and skin structures on sub-barrier fusion, will be:
456
F~~
10 3
t
-« 2 10
a
10 1
10
r r
-1 F
r
14
f
y /
ft
J»
/ A
• ja- - j ^ ^ S ^ S 9-
m *
4
•
^ e + ^ U
•
4
n
, i
H I
.16 L', , I18I , , 20 ,I
22
24
He + 23*U Viola et al.
, , I , , , I , , , I , , ,
26
28
30
32
Ecm(MeV) Figure 1: Fusion cross section for 4 H e + correspond t o ECIS calculations.
238
u (squares) and 6 H e +
238
U (circles). T h e lines
• a more careful measurement of the contribution of transfer and inelastic reactions to the total fission cross section for the system 6 He+ 2 3 8 U by a 4.7T set-up and an improved electronics at Louvain-la-Neuve; • a measurement of the fusion cross section for the skin nucleus 8 He+ 2 3 8 U at SPIRAL-GANIL (Caen, France). References 1. R.G. Stokstad et al., Phys. Rev. C 21, 2427 (1980). 2. M. Beckerman, Rep. Prog. Phys. 5 1 , 1047 (1988). 3. M.J. Rhoades-Brown and P. Braun-Munzinger, Phys. Lett. B 136, 19 (1984). 4. A.B. Balantekin and N. Takigawa, Rev. Mod. Phys. 70, 77 (1998). 5. I. Tanihata et al., Phys. Rev. Lett. 55, 2676 (1985). 6. M. Trotta et al., Phys. Rev. Lett. 84, 2342 (2000). 7. V.E. Viola, Jr. and T. Sikkeland, Phys. Rev. 128, 767 (1962). 8. J. Raynal, Phys. Rev. C 23, 2571 (1981). 9. G.R. Satchler and W.G. Love, Phys. Rep. 55, 183 (1979). 10. D.T.Khoa et al., Phys. Rev. C 56, 954 (1997). 11. C.A.De Jager et a l , At. Data Nucl. Data Tables 14, 49 (1974). 12. S. Karataglidis et al., nucl-th/9811045; private communication.
Concluding Remarks
IMPACT OF NUCLEAR SCIENCE ON MODERN SOCIETY R.A. RICCI Dept of Physics University of Padova Honorary President of Italian Physical Society
INTRODUCTION It is my pleasure to close the Conference with the address of the Italian Physical Society and to be here with old friends whose contribution to the development of Nuclear Physics has been, and still is, of great significance. By the way, let me ask a question: Are "old friends" friends since a long time or friends getting old? At the end of the XX century, "old" could mean for a number of us, at least more than half of the history of Nuclear Physics. Is our science, as we are, still alive? I would like to mention a number of related questions that other excellent colleagues have raised on many occasions and in publications: "Is Nuclear Physics already too old? Why do we continue to study Nuclear Physics? Quo vadis Nuclear Physic?" and so on. There are, in my opinion, two possible answers to these questions. The first is the "continuity" of the field of Nuclear Science as a whole and the fact that a significant number of young people is always involved and contribute to its development hopefully means that Nuclear Physics is still in a good shape. Therefore, it is good to see here, for instance, not only old friends but also quite a number of representatives of the young generations. The second answer is, in a certain sense, another question: What is the impact of Nuclear Science on the modern society and consequently its perception by the public opinion? In other words: What is the public awareness of Nuclear Science? I shall try to discuss with you, in a very short presentation, these two approaches which are, in my view, strongly related. 1
NUCLEAR PHYSICS AS A FUNDAMENTAL SCIENCE
I will start with Nuclear Physics which is basically a fundamental science. It has covered almost all the last century from the discovery of radioactivity (which is a nuclear phenomenon) until the investigation of the ultimate structure of the matter by the possible identification of the Quark Gluon Plasma with relativistic heavy ions (a typical nuclear tool). I don't need to emphasize here this aspect by mentioning the extension of our field of research. This Conference is, among others, in the year 2000, a good illustrative 459
460
example. All the aspects of nuclear theories, basic and expanding models of nuclear structure and reactions, phenomenological and microscopic approaches to nuclear spectroscopy, high excitations (hot nuclei), isospin phenomena, extension of the nuclear landscape (exotic and superheavy nuclei), hadron interactions, phase transitions in nuclear matter, equation of state, transition to the QGP, nuclear astrophysics. There is a crucial point however. In fact Nuclear Physics is a quite complex science, just because of its "extension". Furthermore, its borderlines with other fundamental fields of physics, i.e. the elementary particle and the condensed matter physics, are very peculiar. Fig. 1 shows the standing position of Nuclear Physics, with its relation to the fundamental problems in physics. It also clearly illustrates the crucial positions where the nucleus stands between the primary constituents (elementary particles) and the complex material systems (condensed matter). 1 am using here the terms "simplicity" and "complexity" (we could also refer to the Weisskopf description of "intensive" and " extensive" physics) for the two extreme cases with the nucleus as a crossing point. This is the reason why nuclear physics (which has been often anagrammatized into "unclear physics"), presents a large variety of phenomena and possible investigations. In fact, the nucleus is a two-faced object with a variety of contradictory properties which make this specific stage of matter too small to be considered as a microscopic piece of a real many-body nuclear system and too big to derive all its properties from the primary constituents. 2
PHYSICS AS A "BASIC RESEARCH SERVING SOCIETY"
I am using the title of a report, prepared already in 1994, by the Nuclear Physics Division of the American Physical Society to be addressed to the U.S. Congress, as reported by Gary Crawley in Nuclear Physics News (vol. 5, N. 1, 1995). I found interesting the questions the report is trying to answer: " What do nuclear physicists do?" and "What do we gain from our investment in nuclear science?" . The document -Crawley says- considers three main aspects: Science, Education and Applications. The main problem was -and we know still is- the illustration to various audiences ".... how previous investment in basic research in nuclear science has led to a vastly increased understanding of the universe... and how ....
461 WHERE NUCLEUS STANDS NUCLEAR PHYSICS • UNCLEAR PHYSICS nuclei
atoms quantum mechanics many body
electron. interactions
/////////////////A condensed matter
complexity extensive
simplicity intensive
Fig. 1 Standing position of Nuclear Physics and the relationship$ with fundamental problems.
this and the techniques developed by nuclear scientists have led to numerous applications which have enhanced the quality of life ". It is, of course, a common idea, among nuclear physicists, that there are clear indications for appreciating the fall-out of our science. First the impact on technology, which is, or should be, more visible than it was before. We all are aware of the strong correlation with nuclear energy (fission and fusion), with the enormous development of accelerators and detectors and the new techniques using radio-frequency superconducting cavities and magnets with superconducting wires, so as with analogue and digital electronics and computing systems. Second, to show the impact on more "social" fields, it is worthwhile to mention that the progress of many such fields is due to new methods and techniques "invented" in the frame of nuclear science. A few examples are: the nuclear (I
462
emphasize this term which is now neglected) magnetic resonance (NMR) in medicine and chemistry, the synchrotron radiation in material and bio-sciences, the neutron scattering in determining the structure and dynamics of condensed matter, the medical diagnostic and therapy (with PET, ionizing radiations , particle beams), nuclear analyzing techniques in environmental science and in archeometry and so on. It is clear that the relationships with scientific culture, the application fields and, of course, the education at school level would give us a gleam of optimism. Is it really so? Here the question comes back to my first considerations. We have to face the more general public awareness of our science.
3
THE PUBLIC AWARENESS OF NUCLEAR SCIENCE
It is not by chance that I am using, also here, such a specific title. PANS (Public Awareness of Nuclear Science) is a kind of task force or working group which was formed recently as a common initiative of the Nuclear Physics Board of the EPS (the European Physical Society) and NuPECC (the Nuclear Physics European Coordination Committee). The scope ".... is to enhance the knowledge of nuclear science to a broad public. This can be achieved by acquainting them with some basic features of the atomic nucleus by showing the many impressive peaceful applications and by increasing their awareness that we all live in a natural radiation environment. This hope will be to have a much more informed public concerning "nuclear" issues. This is a necessary condition for a balanced and meaningful debate on issues such as the future of nuclear energy and on scenarios for clearing the huge stock of Plutonium left over from the cold war....". I have expressely underlined the terms radiation and nuclear energy because at least part of the decreasing consideration of nuclear issues is due to the way these terms have been and are presented to the public opinion and the policy makers through the current information given by the mass-media. Of course such an awareness is connected to at least three main approaches to the problem: the understanding which is not only a public opinion affair but also involves the education in the school and the basic cultural background; the appreciation which is the way the public sensitivity is promoted via a correct scientific information; the acceptance which is the final stage for a possible scientific policy once the society as a whole has adopted the correct behaviour. It is clear that the third stage needs the first two as necessary conditions. However, there is no evidence that this root is followed. On the contrary, the public concern about not only nuclear but almost all scientific issues is increasing more and more and -in my opinion- is strongly related to a clear antiscientific attitude arising, at least for a substantial part, from antinuclear movements.
463
Let me quote an editorial of the Nuclear Physics News in 1998 (N.P.N. Vol. 8. N. 1, 1998): " All of physics and science as a whole have been an enormous progress that has dramatically changed the face of the world The changing world that physics is now facing has been changed by physics (itself) as the paradigm of science...." But".... a mood of pessimism has become prevalent about adequate support of research, adequate career opportunities for our students and respectable social future....". I think that pessimism quite justified. There are several points that have been taken seriously into account, for instance by the European Physical Society: -i) the decreasing attraction of physics in general at the University level (decreasing number of students and teachers). This trend is not only limited to Europe and to Physics; it extends also to the USA and to other fields of science; -ii) the decreasing importance given to scientific issues as compared with others. This is related to the lack of scientific culture which is even worse among the so called "intellectuals" than in the common people; -iii) the ideological and sometimes political opposition to the working criteria of science, specifically of physics, and its methodology which do not favour dogmatic statements. This attitude contains over-simplified ingredients: The paradigm of science is physics ; physics means, after all, nuclear physics, nuclear physics has to do with nuciear energy, nuclear weapons (i.e. Chernobyl, Hiroshima and so on....). Forget all the benefits arising from that science. Therefore a statement: "Only fewer people with a knowledge of nuclear physics are needed" and the final sentence: "Research in nuclear physics needs fewer people and less research". I think we should be seriously aware of that problem. It is a duty, in my opinion, not only of our community of nuclear physicists but of the scientific community at large, to be more involved in this cultural engagement. Let's hope that this message will be taken especially by the young generations. But, for this goal, we need more "open-minded" scientists as good researchers and teachers. During the last year we have celebrated the bicentenary of the Volta's battery. Next year we will celebrate the hundred anniversary of the birth of Enrico Fermi (the inventor of the first nuclear reactor). There are three hundred years between these two great inventions, which represent real milestones in the history both of physics and human society. I don't know if Volta and Fermi have to be considered as "intellectuals" in the current sense as many philosophers, opinion makers or politicians. I certainly can say -and you will agree, I hope- that they have built, in their respective period of time, the bridge between scientific knowledge and the benefit of human culture and civilization.
AUTHOR INDEX
Abdinov O. B. 375 Abreu M. C. 91 Agodi C. 299, 335 Aiello S. 379, 383, 391 Ajitanand N. N. 137 ALADIN Collaboration 249, 317 Alamanos N. 453 Alba R. 299, 335 Alderighi M. 379, 383 Alessandro B. 91 Alexa C. 91 Alexander J. 137 Amorini F. 305, 395 Andreyev A. 453 Andrighetto A. 413 Anderson M. 137 Angelov N. S. 375 Antinori F. 73 Anzalone A. 379, 383 Arena N. 391 Arnaldi R. 91 Artemov K. P. 409 Atayan M. 91 Auger F. 453 Auger G. 321
Bassini R. 249 Bastid N. I l l Beaulieu L. 367 Bedjidian M. 91 Begeman-Blaich M. 249 Beghini S. 405, 441 Belkacem M. 335 Bellaize N. 321 Bellia G. 299, 335 Beole S. 91 Berceanu I. 379, 383 Best D. 137 Beusch W. 73 Bidini L. 387 Bini M. 387 Bisogno M. 441 Blazevic A. 419 Bloodworth I. J. 73 Blumenfeld Y. 299 Bohlen H. G. 419 Boldea V. 91 Bonasera A. 179, 379, 383 BondorfJ. P. 119,263 Borcea C. 453 Bordalo P. 91 Borderie B. 187, 313, 321, 379, 383 Botvina A. S. 341, 379, 383 BougaultR. 215,221,321,379,383 Bouriquet B. 321 Brady F. P. 137 Brou R. 321 Bruno M. 215, 299, 335, 379, 383 Buchet J. P. 321 Buchet P. 237
Bacri Ch. O. 321 Baglin C. 91 Balabanski D. L. 453 Baldit A. 91 Baldo M. 379, 383 Ball G. C. 367 Baran V. 293 Barna R. 379, 383 Bartolucci M. 379, 383
465
466
Bugaev K. A. 285 Bussiere A. 91 Calabretta L. 449 Caliandro R. 73 Cannata F. 215 Capelli L. 91 Cardella G. 305, 379, 383, 395 Carmona J. M. 197 Carre' M. 237 Carrer N. 73 Casagrande L. 91 Case T. 137 Casini G. 387 Caskey W. 137 Casandjian J. M. 419 Castor J. 91 Cavallaro S. 379, 383 Cavallaro Seb. 391 Cebra D. 137 Chambon T. 91 Chance J. 137 Charvet J. L. 321 Chaurand B. 91 Chbihi A. 321 Chernomoretz A. 257 Chevrot I. 91 Cheynis B. 91 Chiavassa E. 91 Chiba S. 179, 331 Chikazumi S. 331 Chomaz Ph. 155, 209, 215, 221, 321 Chung P. 137 Cicalo C. 91 Claudino T. 91 Cole B. 137 Colin J. 321 Colonna M. 293, 299, 321, 335, 379, 383 Colonna N. 299, 335 Comets M. P. 91 Coniglione R. 299, 335
Constans N. 91 Constantinescu S. 91 Corradi L. 405, 441 Coulier N. 453 Crowe K. 137 Cussol D. 321 D'Agostino M. 215, 299, 335, 379, 383 D'Amico E. 379, 383 Das A. 137 Dayras R. 321, 379, 383 De Cesare N. 379, 383 De Falco A. 91 De Filippo E. 379, 383 Dellacasa G. 91 De Marco N. 91 Demeyer A. 321 Dem'yanova A. S. 409, 431 De Pasquale D. 379, 383 Desesquelles P. 359, 363 Del Zoppo A. 299, 335 Devaux A. 91 Di Bari D. 73 Di Liberto S. 73 Di Pietro A. 305, 395 Dita S. 91 Di Toro M. 293, 379, 383 Dore D. 321 Dorso C. 0 . 257 Draper J. 137 Drapier O. 91 Drouart A. 453 Drozdov V. A. 445 Ducroux L. 91 Duflot V. 155 Durand D. 321 Durand D. J. C. 453 D'Yachenko A. T. 133 E895 Coll. E900 Coll.
137 231
467
Elia D. 73 Eremenko D. O. 445 Espagnon B. 91 Evans D. 73 Fabbietti L. 299, 335 Faessler A. 273 Fanebust K. 73 Fargeix J. 91 Farizon B. 237 Farizon M. 237 Fayazzadeh F. 73 Fedorov D. V. 401 Femino S. 379, 383 Fiandri M. L. 215, 299 Figuera P. 305, 395 Fini R. A. 73 Finocchiaro P. 299, 335 FOPI Collaboration 111 Force P. 91 Fotina O. V. 445 Frankland J. P. 321 Franzosi R. 203 Fritz S. 249 Ftanik J. 73 Fuchs C. 273,279 Fuschini E. 215,379,383 Gagne P. 367 Gaillard M. J. 237 Gaitanos T. 273, 279 Galichet E. 321 Gallio M. 91 Garrido E. 401 Gavrilov Y. K. 91 Gelbke C.-K. 167 Georgiev G. 453 Geraci E. 379, 383, 391 Geraci M. 379, 383 Gerlic E. 321, 363 Gerschel C. 91 Ghetti R. 371
Ghidini B. 73 Gilkes M. 137 Gillibert A. 453 Gingras L. 367 Giorgini M. 149 Giubellino P. 91 Giustolini F. 379, 383 Gloukhov Yu. A. 409, 431 Gobet F. 237 Golubeva M. B. 91 Goncharov S. A. 409, 431 Gonin M. 91 Gorenstein M. I. 285 Gramegna F. 215, 299, 335 Greco V. 299 Greiner W. 285 Grella G. 73 Grigorian A. A. 91 GroB C. 249 Gross D. H. E. 53 Grossiord J. Y. 91 Grzeszczuk A. 379, 383 Guarnera A. 321 Guazzoni P. 379, 383 Guber F. F. 91 Guichard A. 91 Guinet D. 321, 379, 383 Guiot B. 321 Gulino M. 73 Gulkanyan H. 91 Gulminelli F. 155, 215, 221 Guo J. Y. 413 Gushue S. 137 Guttormsen M. 243 Hakobyan R. 91 Haroutunian R. 91 He Z. 367 Heffner M. 137 Helstrup H. 73 Henriquez M. 73 Hinnefeld J. D. 453
468
Hirsch A. 137 Hjort E. 137 Hjorth-Jensen M. Holme A. K. 73 Horn D. 367 Hudan S. 321 Huo L. 137 Huss D. 73 Huyse M. 453
243
Iacono Manno M. 379, 383 Imme G. 249 Idzik M. 91 INDRA Coll. 221, 309, 313, 363 Iori I. 215, 249, 299, 335 Italiano A. 379, 383 Itkis M. G. 449 Iwamoto A. 331 Jacholkowski A. 73 Jensen A. S. 401 Jones G. T. 73 Jouan D. 91 Jouanne C. 453 Julin R. 431 Justice M. 137 Kaplan M. 137 Karavitcheva T. L. 91 Keane D. 137 Khoa D. T. 419 Kinson J. B. 73 Kintner J. 137 Klay J. 137 Kluberg L. 91 Knudson K. 73 Kolwaski S. 379, 383 Kondratiev N. A. 449 Koshelkin A. V. 427 Kozulin E. M. 449 Kralik I. 73
Krofcheck D. 137 Kurepin A. B. 91 Kuznetsov A. A. 375 Lacey R. 137 Laforest R. 347 Lanzalone G. 305, 379, 383, 391 Lanzano G. 379, 383 Lapoux V. 453 Larochelle Y. 367 Latora V. 327 Lautesse P. 321, 363 Lavaud F. 321 Laville J. L. 321, 363 Le Bornec Y. 91 Lecolley J. F. 321 Leduc C. 321 Legrain R. 321 Le Neindre N. 215, 221, 321, 379, 383 Lenti V. 73 Lepine A. 453 Li Bao-An 129 Li S. 379, 383 Lietava R. 73 Lisa M. A. 137 Liu H. 137 Liu Y. M. 137 Loconsole R. A. 73 Lobner K. E. G. 413 Lombardo U. 379, 383 Lo Nigro S. 379, 383 Lopez O. 321 Loukachine K. 299 Lourengo C. 91 Louvel M. 321 L0vh0iden G. 73 Lu J. 305, 395 Lukasik J. 309, 321 Lumbroso A. 453 Lynen U. 249
469 Macciotta P. 91 Mac Cormick M. 91 Mahboub D. 379, 383 Mahi M. 249 Maiolini C. 449 Maiolino C. 299, 335, 379, 383 Manfredi G. 379, 383 Manzari V. 73 Manzoor S. 149 Marie F. 453 Margagliotti G. V. 215, 299, 379, 383 Mark T. 237 Martin E. 347 Maruyama T. 179, 331 Marzari-Chiesa A. 91 Masera M. 91 Masetti S. 119 Maskay A. M. 321, 363 Masoni A. 91 Mastinu P. F. 299, 335 Maurenzig P. R. 387 Mazzoni M. A. 73 McGrath R. 137 Meddi F. 73 Mehrabyan S. 91 Melby E. 243 Michalon A. 73 Michalon-Mentzer M. E. 73 Michel N. 197 Migneco E. 299, 335 Milazzo P. M. 299, 335 Milosevic J. 145 Milosevich Z. 137 Mishustin I. N. 119, 263, 285 Mittig W. 419 Moisa D. 379, 383 Mohlenkamp T. 249 Montagnoli G. 405, 441 Monteno M. 91 Morando M. 73 Moroni A. 215, 299, 335
Miiller W. F. J. 249 Musso A. 91 Musumarra A. 305, 395, 453 Na49 Collaboration Na50 Collaboration Nalpas L. 321 Nara Y. 115 Neergaard G. 263 Neyens G. 453 Niita K. 331 Norman P. I. 73 Normand J. 321 Nouffer F. 419
99 91
Ocker B. 249 Odeh T. 249 Odyniec G. 137 Oglobin A. A. 409, 431 Olmi A. 387 Olson D. 137 Ottini S. 453 Ouerdane D. 367 Paduszynski T. 379, 383 Pagano A. 379, 383 Panitkin S. Y. 137 Papa M. 305, 379, 383, 395 Pappalardo G. 305, 395 Paramonov V. V. 409 Parlog M. 321 Pasquali G. 387 Pastirak B. 73 Pawlowski P. 313, 321 Peitzmann T. 123 Petiau P. 91 Petrovici M. 379, 383 Pettini M. 203 Piantelli S. 387 Piasecki E. 379, 383 Piattelli P. 299, 335 Piccotti A. 91
470
Pierroutsakou D. 413 Pinkenburg C. 137 Pirrone S. 305, 379, 383, 391 Pizzi J. R. 91 Plagnol E. 309, 321 Platonov S. Yu. 445 Pochodzalla J. 249 Poggi G. 387 Poggi S. 387 Pokrovskii I. V. 449 Politi G. 379, 383, 391 Pollacco E. 379, 383 Pollarolo G. 405 Pop A. 379, 383 Porile N. 137 Porto F. 379, 383, 391 Prino F. 91 Prokhorova E. V. 449 Pshenichnov I. A. 119 Puddu G. 91 Qian X. 367 Quercigh E. 73 Quintans C. 91 Raabe R. 453 Raciti G. 249 Raduta A. H. 227, 269 Raduta A. R. 227, 269 Rai G. 137 Ramakrishnan E. 347 Ramello L. 91 Ramos S. 91 Randrup J. 105 Rapisarda A. 327, 379, 383 Rato Mendes P. 91 Rekstad J. 243 REVERSE Collaboration 379, 383 Riccati L. 63, 91 Ricci R. A. R. 459 Riccobene G. 249 Richert J. 197
Ritter H. G. 137 Rivet M. F. 321, 379, 383 Rizzo F. 305, 395 Rohrich D. 73 Romana A. 91 Romano F. P. 249 Romano G. 73 Romero J. 137 Romoli M. 413 Ropotar I. 91 Rosato E. 321, 379, 383 Roussel-Chomaz P. 419 Rowland D. 347 Roy R. 367 Rozhkov M. V. 409, 431 Ruan M. 413 Ruangma A. 347 Rudakov V. P. 409 Rudolph K. 413 Rui R. 299, 335 Rusanov A. Ya. 449 Sadigov Z. Y. 375 Safai'k K. 73 Saija A. 249 Saint-Laurent F. 321 Salou S. 321 Sambataro S. 379, 383, 391 Sandor L. 73 Santonocito D. 299, 335 Sapienza P. 299, 335 Saturnini P. 91 Satz H. 71 Scarlassara F. 405, 441 Scarpaci J. A. 299 Scharenberg R. 137 Scheier P. 237 Schiller A. 243 Schnittker M. 249 Schroeder L. S. 137 Schiittauf A. 249 Schwarz C. 249, 317
471
Scomparin E. 63, 91 Sechi G. 379, 383 Seidel W. 249 Segato G. 73 Serci S. 91 Serfling V. 249 Sfienti C. 249 Shahoyan R. 91 Sida J. L. 453 Siem S. 243 Signorini C. 413 Silva S. 91 Simic L. J. 145 Simion V. 379, 383 Sisto M. 335 Sitta M. 91 Siwek-Wilczynska K. 435 Soave C. 91 Sonderegger P. 91 Soramel F. 413 Sperduto M. L. 379, 383 Spinelli L. 203 Srivastava B. 137 Staroba P. 73 Steckmeyer J. C. 321, 379, 383 Stefanini A. A. 387 Stefanini A. M. 405, 441 Stone N. T. B. 137 St-Pierre C. 367 Stroe L. 413 Strobele H. 99 Suleymanov M. K. 375 Sutera C. 379, 383 Symons T. J. M. 137 Tabacaru G. 321 Taccetti N. 387 Tamain B. 321 Tarrago X. 91 Tassan-Got L. 321 Ternier S. 453 Thompson I. 413
Thompson M. 73 Tonetto F. 299 Topilskaya N. S. 91 Trautmann W. 249 Tretyakova S. P. 449 Tretyakova T. Yu. 449 Trifilo A. 379, 383 Trimarchi M. 379, 383 Trotta M. 413, 453 Trzaska W. H. 431 Trzcinski A. 249 Tudisco S. 305, 395 Urban J. 73 Usai G. L. 91 Van Duppen P. 453 Vannini G. 215, 299, 335, 379, 383 Ventura A. 119 Ventura P. G. 299 Vercellin E. 91 Verde G. 249 Verondini E. 215 Veslesky M. 347 Vient E. 321 Vigilante M. 379, 383 Vik T. 73 Villalobos Baillie O. 73 Villatte L. 91 Vinodkumar A. M. 405, 441 Viola V. E. 231 Virgili T. 73 Vitturi A. 413 Vivey K. 453 Vodopyanov A. C. 375 Volant C. 321, 453 Von Oertzen W. 419 Votruba M. F. 73 Wa97 Collaboration Wa98 Collaboration
73 123
472
Wagner P. 197 Wang S. 137 Wells R. 137 Westfall G. D. 353 Whitfield J. 137 Wieleczko J. P. 321 Wienold T. 137 Wilczyriski J. 379, 383, 435 Willis N. 91 Winchester E. M. 347 Wirzba A. 83 Witt R. 137 Wolski R. 453 Wolter H. H. 273, 279, 293 Wood L. 137 Wu H. 379, 383
Xi H. 249 Xiao Z. 379, 383 Yang X. 137 Yennello S. J. 347 Yuminov O. A. 445 Zavada P. 73 Zetta L. 379, 383 Zhang W. N. 137 Zhang Y. 137 Zichichi A. 3 Zielinska-Pfabe M. 293 Zipper W. 379, 383 Zwieglinski B. 249
- • i
•
J
T0\
ISBN 981-02-4731-
www. worldscientitic. com 4795 he