5.: i... fugue , E Wflg 6. . 1 am. x r ”I. I Prammyflm . 31.. I gumsfimfl 2w.” (“g 2 .1 M43“. 14:. 3'? 'x 1'. ‘I ‘1‘ t ...u... . RuM.J%a r. .3 5......“ k .m»? £qu )3 .13.: n “a." I: A...“ ts} . 1'16 . 3.35%....9»... 11’ 5.. ..»fl..:... . 3.51%.“... t. 2..... 3... . ._ «1 fi. 5...... p. :2 2.2:: r)... . it. in. IES'B ooé LIBRARY Michigan State University This is to certify that the dissertation entitled ISOSPIN DYNAMICS AND THE ISOSPIN DEPENDENT EOS presented by TIANXIAO LIU has been accepted towards fulfillment of the requirements for the Ph.D. degree in PHYSICS Major Profeéor’s Signature 43/2 V/Of’ Date MSU is an Affinnative Action/Equal Opportunity Institution PLACE IN RETURN Box to remove this checkout from your record. TO AVOID FINES return on or before date due. MAY BE RECALLED with earlier due date if requested. DATE DUE DATE DUE DATE DUE 2105' cfifiC/Daie'Due—indci—pns' fl. ISOSPIN DYNAMICS AND THE ISOSPIN DEPENDENT EOS By Tianxiao Liu A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the Degree of DOCTOR OF PHILOSOPHY Department of Physics and Astronomy 2005 ABSTRACT ISOSPIN DYNAMICS AND THE ISOSPIN DEPENDENT EOS By Tianxiao Liu Isotopic yields for light particles and intermediate mass fragments have been mea- sured for 112Sn+1mSn, 112Sn+124Sn, 124Sn+1128n and 124Sn+124Sn with the Large Area Silicon Strip Array (LASSA). LASSA consists of nine individual telescopes with each telescope comprised of two Si strip silicon detector layers backed by four CsI(Tl) crystals. LASSA provides high angular, energy and isotopic resolution over a large detection area and energy range. The total solid angle covered by LASSA is about 540msr with polar angle 7°-58° and an angular resolution of :l:0.43°. The energy range covered is E /A = 2.4-140 MeV for protons and E/A = 4.8-335 MeV for 16O. Isotopic yields from central collisions were compared with predictions of stochastic mean field calculations. These calculations predict a sensitivity of the isotopic dis- tributions to the density dependence of the asymmetry term of the nuclear equation of state. However, the secondary decay of the excited fragments modifies signifi- cantly the isotopic distributions produced by the stochastic mean field model. The predicted final isotOpe distributions after decay are narrower and more neutron-rich than the experimental data and the sensitivity of the predicted yields to the density dependence of the asymmetry term is reduced. The distributions calculated using the stiff asymmetry term are more neutron-rich and are closer to the measured values. However the less than satisfactory level of agreement between theory and experiment precludes definitive statements about the density dependence of the asymmetry term of the EOS. Tianxiao Liu Energy spectra and average energy of light particles and fragments from central collisions were obtained and compared with statistic models calculations. We found that neutron deficient isotopes are significantly more energetic than those of neutron rich isotopes of the same element. This trend is well beyond what can be expected for the bulk multi-fragmentation of an equilibrated system. It can be explained, however, if many of these fragments are evaporated from the surface of the system while it is expanding and cooling. Measurements of particles and fragments from peripheral collisions were used to study the dynamics in nuclear reactions. Isoscaling analyses imply that the quasi- projectile and quasi-target in these collisions do not achieve isospin equilibrium, per- mitting an assessment of the isospin transport rates. We find that comparisons be— tween isospin sensitive experimental and theoretical observables, using suitably chosen scaled ratios, permit investigation of the density dependence of the asymmetry term of the nuclear equation of state. This observable appears to be one of those promis- ing ones for providing much needed constraints on the asymmetry at sub-saturation density. ACKNOWLEDGEMENTS First and foremost, I would like to express my most sincere thanks to my disser- tation advisor, Dr. William G. Lynch. He is a unique combination of great intellects and genuine caring for those persons for whom he has some responsibility. I was very fortunate to be one of these people. To me, he is not only a world class advisor, but also a dear brother and an influential mentor. He is a source of endless inspiration even during the most stressful times. His guidance, support, vision and friendship will be a long-lasting influence in my professional and personal life. I would also like to thank Dr. Pawel Danielewicz, Dr. Chien-Peng Yuan, Dr. Betty Tsang and Dr. Edward Brown for serving on my guidance committee. Special thanks go to Dr. Betty Tsang for her enthusiasm. She is an inexhaustible source of intellectual and creative energy. Her good sense in finding general physics laws to explain complicated experimental data was an excellent lesson to me. I am extremely impressed by her ability to get an intuitive feel of the most difficult ideas. Special thanks also go to Dr. A. Wagner, Dr. H. S. Xu, Dr. H. F. Xi, Dr. G. Verde and Dr. MJ van Goethem who helped me with software writing, experiment set-up and data analysis. My work on my dissertation and the research benefited from the successful collab- oration of the groups at Michigan State University (MSU), Indiana University (IU) and Washington University (WU). I would like to thank the MSU group of C. K. Gelbke, W. P. Tan, X. D. Liu, W. G. Lynch, M. B. Tsang, A. Vander Molen, G. Verde, A. Wagner, H. F. Xi, and H. S. Xu, the IU group of L. Beaulieu, B. Davin, Y. Larochelle, T. Lefort, R. T. De Souza, R. Yanez and V. E. Viola and the WU group of R. J. Charity and L. G. Sobotka. A number of people have contributed to the work which culminates in this disser- tation. Among them are: Drs. M. Colonna, M. Di toro, M. Zielinska-Pfabe, W. H. iv Woolter, W. A. Friedman, S. R. Souza, R. Donangelo, L. J. Shi and P. Danielewicz. I gratefully acknowledge their assistance on model calculations. This dissertation and research that it describes could not be accomplished had it not been for the excellent support of the staff at the National Superconducting Cyclotron Laboratory of Michigan State University. I owe my deepest thanks to John Yurkon of the detector lab, Len Morris of the design group, Jim Vincent of the electronics group, Andrew Vander Molen and Ron Fox of the computer group, who helped with data acquisition. CONTENTS LIST OF TABLES viii LIST OF FIGURES ix 1 Introduction 1.1 1.2 1 Background and Motivation ....................... 1 Theoretical Models ............................ 5 1.2.1 Boltzmann-Uehling-Uhlenbeck (BUU) Model .......... 5 1.2.2 Stochastic Mean Field (SMF) model .............. 8 1.2.3 Statistical Multifragmentation Model (SMM) Model and Macro- canonical Approach ....................... 10 1.2.4 Expanding Emitting Source (EES) Model ........... 14 1.3 Studies Prior to This Dissertation .................... 16 1.4 Thesis Organization ............................ 18 2 Experimental Setup 20 2.1 Experimental Layout ........................... 20 2.2 LASSA Telescopes ............................ 24 2.3 The Miniball/Miniwall Array ...................... 33 2.4 The Ring Counter Foward Array .................... 41 2.5 Data Acquisition Electronics ....................... 47 3 Data Analysis 52 3.1 Data Reduction Algorithms ....................... 52 3.2 Data Analysis of LASSA ......................... 53 3.2.1 Overview ............................. 53 3.2.2 Energy Calibration of Si ..................... 54 3.2.3 Multiple-hit pattern Decoding .................. 59 3.2.4 Construction of PID function and Non-uniform Thickness Cor- rection .............................. 61 3.2.5 CsI calibration .......................... 69 3.2.6 Background Subtraction ..................... 73 3.2.7 Efficiency Correction ....................... 76 3.3 Data Analysis of the Ring Counter ................... 80 3.3.1 Energy Calibration ........................ 84 vi 3.3.2 PID function ........................... 88 4 Experimental Data and Model Comparisons 90 4.1 Impact Parameter Selection ....................... 90 4.2 Isotope Multiplicities, Energy Spectra from Central Collisions and Model Calculation ................................ 95 4.2.1 Experimental Data ........................ 95 4.2.2 Models for the Multiplicity Calculation ............. 104 4.2.3 Overall Behavior Predicted by the SMF Calculations ..... 110 4.2.4 Isoscaling Analyses ........................ 115 4.2.5 Sensitivity of the SMF Calculations to the Asymmetry Term . 119 4.2.6 Discussion of the Isotope Multiplicity and SMF Calculations . 123 4.2.7 Energy Spectra and Equilibrium Models Calculation ..... 126 4.3 Isospin Diffusion in the Peripheral Collisions and Model Calculation 135 4.3.1 Data from the Peripheral Collision ............... 135 4.3.2 BUU Calculations ......................... 144 5 Summary 148 LIST OF REFERENCES 152 vii LIST OF TABLES 2.1 3.1 3.2 4.1 Table of the Miniball/Miniwall detectors used for the impact parameter selection, their corresponding polar angles and solid angle coverage . 38 Table of the fitting parameter used in the energy calibration functions of LASSA CsI(Tl) crystals ........................ 72 Table of fragmentation products used in the energy calibration of LASSA CsI(Tl) crystals .............................. 74 Values for a and 5 obtained from fitting the isotope ratios R21 . . . . 103 viii LIST OF FIGURES 1.1 The symmetry potential for neutrons and protons with different pa- rameterization ............................... 1.2 Mean-field enhancement factor of the diffusion coefficient in symmetric nuclear matter as a function of T at a fixed density .......... 1.3 Experimental isoscaling behavior exhibited by the central 124Sn + 124Sn and 112Sn + 112Sn collisions . . . . . . . . . . . ............ 1.4 The ratio of pre-equilibrium neutrons to protons as a function of nu- cleon kinetic energy with both the Coulomb and symmetry potentials 1.5 Isoscaling parameter from central Sn+Sn collisions, experimental data and calculations from the hybrid models ................ 2.1 Major detectors used in the experiment, viewing in the downstream direction of the beam ........................... 2.2 Mechanical Layout of the experiment, viewing from the side ..... 2.3 The relative position of LASSA and the Ring Counter ......... 2.4 Photo of a single LASSA telescope ................... 2.5 A schematic drawing of the LASSA detector .............. 2.6 A two dimensional histogram of the raw signal outputs of DE(first silicon) vs. EF(second silicon) ...................... 2.7 A two dimensional histogram of the raw output signals of DE+EF (silicon) vs. E(CsI) ............................ 2.8 LASSA geometric acceptance ...................... 2.9 A schematic of the elctronics for the LASSA telescope ......... 17 19 21 23 23 25 26 28 29 31 32 2.10 Schematic of phoswich assembly of individual Miniball/Miniball detector 34 2.11 Output of the phoswich of Miniball/Miniwall .............. 2.12 A two dimensional histograms of the raw signal from Slow vs. Fast and PDT vs. SLOW ........................... 2.13 Schematic diagram of the electronics of the Miniball/Miniwall . . . . 2.14 A schematic drawing of the Ring Counter ................ 2.15 A schematic figure of the side view of Ring Counter .......... 2.16 A two dimensional histogram of the raw signal from Pie side and C31 2.17 A block chart of the electronics of the Ring Counter .......... 2.18 Layout of the data acquisition ...................... 2.19 Schematic drawing of the principle of the ECL readouts ........ ix 36 37 40 42 43 44 46 50 51 3.1 3.2 3.3 3.4 3.5 3.6 3.7 3.8 3.9 3.10 3.11 3.12 3.13 3.14 3.15 3.16 3.17 3.18 3.19 3.20 4.1 4.2 4.3 4.4 4.5 4.6 4.7 4.8 4.9 4.10 4.11 Major detectors used in the experiment, viewing in the downstream direction of the beam ........................... The PC-DAQ system used for the silicon calibration .......... Calibration lines of energy(MeV) vs. Dial value(Volt) ......... ADC output vs. precise pulser dial value (Volt) ............ Comparison of isotope resolution before and after thickness correction A histogram of counts vs. nominal DE energy and nominal EF energy A histogram of nominal counts vs. nominal DE+EF energy and matched CsI light output .............................. 1D PID histogram for particles stops in the second silicon detector . . 1D PID histogram for particles stops in the CsI crystals ........ Calibration curves of Carbon for LASSA CsI(Tl) ............ PID line of 3He shows the contamination of 4He ............ Background subtraction for the heavy fragments ............ The raw yield and the efficiency function for 4He produced in the pe- ripheral collision for the 112Sn + 112Sn reaction ............. The corrected yields of 4He produced in the peripheral collisions for different reactions ............................. LASSA 4He yield in peripheral collisions ................ Ring Counter analysis flow chart .................... Ring Counter calibration for silicon detector(Pie side) ......... Ring Counter calibration for the CsI detector ............. The 2D PID lines of the Ring Counter ................. Ring Counter PID 1D .......................... The probability distribution of the charged particle multiplicity and impact parameter as a function of the charged particle multiplicity . . Nc vs. szf ................................. LASSA 4He yield in central collisions .................. LASSA 12C yield in central collisions .................. Energy spectra of H, He, Li, Be from the central collisions of the reac- tion 112Sn—l—112Sn ............................. Energy spectra of B, C, N, O from the central collisions of the reaction 112Sn+112Sn ................................ Isotope yields in central collisions .................... Behavior of supper stiff and soft symmetry terms at low densities Differential multiplicities for 124Sn+124Sn collisions and 112Sn+1128n collisions as a function of A and Z .................... Primary and secondary calculations of Carbon and asymmetries, as well as the comparison with data .................... 55 56 58 59 63 65 66 67 68 71 77 78 81 82 83 85 86 87 88 89 93 94 97 98 100 102 106 111 114 Primary and secondary calculations of ratio of the yields from 124Sn+124Sn and 112Sn+ImSn .............................. 117 4.12 Differential multiplicities for Carbon and Oxygen, data and calculation 120 4.13 Scaled function S (N) = R21 (N, Z) . 6‘”, data and calculation 122 4.14 4.15 4.16 4.17 4.18 4.19 4.20 4.21 4.22 11C and 12C energy spectra from the central collisions, data and calcu- lations ................................... 128 Mean kinetic energies from central collisions, data and calculation . . 129 Mean kinetic energy and emission rates for 12C (11C) from the EES calculation ................................. 133 LASSA 12C yield in central collisions .................. 137 Differential multiplicities for isotopes with 5 > 0.8 cut ......... 138 R21 using 124Sn + 124Sn collisions as reaction 2 and fitted values for isoscaling parameter a .......................... 140 Rn from ln(Yz'eld(7Be)/Yz'eld(7Li)) ................... 142 Calculated R,- and comparison with data ................ 143 Calculated R,- and comparison with data ................ 146 xi Chapter 1 Introduction 1.1 Background and Motivation Heavy nucleus-nucleus collisions provide a unique way of exploring several interesting topics, among which are the nuclear liquid-gas phase transition and nuclear equation of state (EOS). The liquid-gas phase transition has been investigated for a long time. It bears some relevance to the phenomenon of multi-fragmentation but its detailed connection to that process is still open to questions. The nuclear equation of state (EOS) describes the relation between pressure, den- sity, temperature and isospin asymmetry for nuclear matter. Understanding EOS is very crucial in astrophysical contexts. For example, the EOS governs the dynamics of dense matter in supernovae [7]. It also influences strongly the internal structure, radii, maximum masses and moments of inertia of neutron stars [59]. In addition, the EOS influences the time scale for cooling of proto—neutron stars. In particular, the density dependence of the asymmetry term of the EOS can determine whether or not the direct Urca process can occur[57]. Finally, it is relevant to remember that the EOS affects significantly the binding energy and rms radii of neutron-rich nuclei [100] as well as the difference between neutron and proton radii in nuclei [17]. These parameters may influence the formation of the heavy elements via the r-process. Significant constraints have already been placed on the symmetric matter EOS at high densities [31]. However, the constraints on isospin-asymmetry term of the EOS, which describes the sensitivity of EOS to the difference between neutron and proton densities, is poor. Better knowledge of this term is essential to understand both the central density of neutron-rich isotopes [105] and the internal structure, radii and moment of inertia for neutron stars [81, 58]. Theoretical studies have shown that the EOS of asymmetric nuclear matter can be approximately expressed as [63] E(p.5) =E(p,5=0)+5(p)52, (M) where p = pn + )0p is baryon density, 6 = (pn - pp)Pn + pp) is the relative neutron excess, and E(p, 6 = 0) is the energy per particle in symmetric nuclear matter. The bulk symmetry energy is denoted by S (p) Its value at normal density, SOES(po), is known to be in the range of 27-36 MeV [46]. Different density dependences of S (p) have been found, depending on the nuclear forces used in the calculation. One simple parameterization can be written as [85] 3 S(P) = (22/3 -1)(5E2~)[u2/3 — 1701)] + Sofia) , (12) where E} is the Fermi energy and u E p/po is the reduced baryon density, and F(u) denotes the potential energy contribution. To understand the role the density depen- dence of F(u) has on various quantities, it has been approximated by the following parameterizations: F ( ) 2212 u = 1 1 + u’ F2(u) = u, (1.3) F3(U) : u1/2, which provide a range of reasonable density dependencies[63]. 2 The symmetry potential for neutrons and protons due to the interaction compo- nent of the symmetry energy can be obtained, 3(pF (1052) V"(p)(p, 5) = ea apub») asy (1.4) where an E [So —— (22/3 — 1)(g-E2«)] is the contribution of nuclear interactions to the bulk symmetry energy at normal nuclear matter density. Fig. 1.1 shows the symmetry potential as a function of u, using two forms of F (u), i.e. F1 and F3 and So = 30MeV. The symmetry potential from F2, which is not shown, always falls between those for symmetry potentials from F1 and F3. From Fig. 1.1 we can see that the symmetry potential is quite sensitive to the form of F (u), the neutron excess 6 and the baryon density p. In collisions of neutron-rich nuclei at intermediate energies, both 6 and p can be appreciable in a large space- time region where the isospin-dependent mean fields, which are opposite in sign for neutrons and protons, are strong. This will affect differently the reaction dynamics of neutrons and protons, leading to possible differences in their yields and energy spectra. This suggested that it is possible to extract information on the asymmetry term of the nuclear EOS in regions under laboratory controlled conditions. Although in heavy- ion collisions at intermediate energies we certainly cannot reach very high-density regions, we should be able to obtain information on the slope of the asymmetry term below the saturation density. Qualitatively different effects arise in central collisions, with bulk fragmentation, and peripheral collisions with neck fragmentation. In this way we hope to put some experimental constraints on the effective interactions used in astrophysical contexts [9]. 100 T Y I I I T T‘TWI Y I, I I I I T T r I T T ‘1? I C I 1 I l j 80 — /6'—0.8 _ r Neutron , : r- , .. l J L 2 C 1 A ; j > _ _ o i- - E L v i- 5‘ > i i .- a. - . \ \ -40 :- g \\ “- \ \ -» i ’ ' " f \ j - Fl \ 5:0.8 J -60 L 3 \ \ _ Z \\ (Hi: -80 :_ Proton . _ ’ 5—08 I \ -100 I . I I I 1 . . . r 1 . . .\ . J . . . . 1 . . I . 0.0 0.5 1.0 1.5 2.0 2.5 p/p0 Figure 1.1: The symmetry potential for neutrons and protons corresponding to the two forms of F(u) and $0 = 30MeV with 6 = 0.2 and 0.8. Images in this dissertation are presented in color. 1.2 Theoretical Models The extraction of information about the EOS or the liquid-gas transition of nuclear matter are usually done by comparing experimental data to a theory model calcula- tion. In this dissertation, several models were used. Two of them are the Boltzmann- Uehling-Uhlenbeck (BUU) and Stochastic Mean Field (SMF) models, which make predictions for the reaction dynamics without any assumption about the degree of thermal equilibrium achieved. The other two models, Expanding Emitting Source (EES) and Statistical Multifragmentation Model (SMM), assume some degree of ther- malization. In the following, we will discuss and contrast these different approaches. 1.2.1 Boltzmann-Uehling—Uhlenbeck (BUU) Model The exact theoretical assumption of the collision of two complex nuclei at the incident energies described in the dissertation can be in principle found in the many body Schrddinger equation. This task, however, is fraught with difficulties, both formal and technical, which motivate the introduction of approximate methods that focus on specific aspects of the collisions. Here, we describe an approach formulated to calculate observables that are related to the average dynamics of a collision. The Boltzmann-Uehling-Uhlenbeck (BUU) [113, 74, 51] equation describes time evolution of the single particle phase space distribution function f1 (p, r, t) and reads as follows: (11.1) at R'Vrfl—vv'val=Icolla (1-5) where, [can accounts for collisions and can be written as 1 Icon = —/W0V12If1f2(1 — fl')(1— fz') — fl’f2'(1 — f1)(1— f2)l X (2r)353(p1 + P2 - p’1 - p’z)d3p2d3p’1d3p’2 - (1-6) Here, a is the differential scattering cross section for a certain change of momentum, -'~'---c-_-A. H-“w g o (p1,p2) -> (p’1,p’2), and v12 is the relative velocity for the colliding nucleons. Icon differs from the classical collision by the Pauli-blocking factor (1 — f1(1:))(1 — f2(2I)). In the BUU equation, V is the mean-field potential, which will be a function of the local density. It can be parameterized as an arbitrary function of density, making it possible to model a variety of equations of state. Typically, it can be written as the sum of three terms: vzw+m+nw, up where VC, V" and Vasy represent the Coulomb potential, the isoscalar nuclear potential and the symmetry energy, respectively. With the right-hand side set to zero, Eq. 1.5 becomes the Vlasov equation, which describes the evolution of a single-particle distribution under the influence of a self- consistent mean field. Interpreting f1 (p, r, t) as the Wigner transform of the quantum mechanical density matrix, the Vlasov equation can be derived from time-dependent Hartree-Fock (TDHF) in the limit of smoothly varying potentials. To solve BUU equation, given in Eq. 1.5, one often uses the so called Test Particle method [5, 30], in which the continuous variables p and r are restricted to a set of discrete values, and the phase-space density is represented in terms of a set of 6-functions, or test particles, f(p, r, t) = (33") X 6(p — p.(t))6 p0, than the linear dependence, while the opposite is true at p < p0. Thus one would expect the stiffer (u = 2) symmetry energy term to have a large diffusion at high density and the soft (V = 1) symmetry energy term to have a larger diffusion at lower density as shown by Fig. 1.2. In fact, the difl’usion coefficient depends monotonously on the symmetry energy and is large when the symmetry energy is large and small when the symmetry energy is small. At low temperature and moderate to high densities the amplification is very strong suggesting that the measurements sensitive to the diffusion can be used to probe the asymmetry energy. The same calculation also predicts the diffusion coefficient to be reduced by proton-neutron scattering by the residual interaction [93]. 1.2.2 Stochastic Mean Field (SMF) model The BUU equation provides an accurate description of the time dependence of the one body distribution function. Accurate solutions of the BUU equation average away fluctuations in the density that might lead to the formation of fragments in \ — Linear \ - - - - Quadratic Figure 1.2: Mean-field enhancement factor of the diffusion coefficient in symmetric nuclear matter, RED5(ei’;fin)/D5(e§';fin = 0), at a fixed density n, as a function of temperature T. The solid and dashed lines, respectively, represent the factors for the assumed linear and quadratic dependence of the interaction symmetry energy on density. The lines from top to bottom are for densities p = 2po, p0, 0.5p0, and 0.1po, respectively. At normal density the results for the two dependencies coincide. [93] an individual collision. This is usually achieved by solving the BUU equations with a large number of test particles per nucleon Nmt. Because the density fluctuations that lead to fragment formation are suppressed in the BUU equation, the calculation of fragment yields directly via the BUU model is not feasible. Therefore, alternative Models, such as Stochastic Mean Field (SMF) model [23] has been developed to address the density fluctuation. SMF, like BUU, describes the time evolution of the collision using a self-consistent mean field. The application of the SMF model to the unstable situations relies on the knowledge of the most important unstable modes, which maybe difficult to identify in cases where the modes are not known apriori. To account for the stochastic fluctuation in SMF model, two approaches are used. In the first approach, a fluctuating term is added to the standard Boltzmann-Nordeim— Vlasov (BNV) equation to account for the stochastic force, the strength of which is adjusted to reproduce the growth of the most important unstable mode in the system. In practical term, the stochastic force is a type of stochastic noise. The second approach uses this fact by approximating it with numeric noise caused by solving BUU with a small number of test particles (i. e. N¢m=50) per nucleon. 1.2.3 Statistical Multifragmentation Model (SMM) Model and Macrocanonical Approach Statistical models assume the existence of statistical (equilibrated) distributions of states during the fragmentation. Equilibrium statistical theories provide one approach to calculate the properties of such fragments. Such models can provide precise pre- dictions. The accuracy of such prediction is directly related to the accuracy of the equilibrium approximation itself. The copious production of intermediate mass fragments (IMFs), with mass in- termediate between that characteristic of light particle evaporation and fission, is a 10 distinguishing feature of intermediate-energy heavy-ion reactions [73]. The emission of IMFs at low emission rate is a prediction of statistical evaporation theories [37, 69]. The observation of very high IMF multiplicities, however, lies outside of the realm of prediction of evaporation theories [15, 37]. Such an observation suggests a different mode of decay via a bulk disintegration [10, 37, 42]. Microcanonical Statistical Multifragmentation Model (SMM) uses a Monte Carlo method to calculate physics observables in approximate microcanonical equilibrium approach [102]. If one assumes that equilibrium is achieved for a source with excita- tion energy E5, volume V, mass A0 and charge Z0 specified. Possible decay modes are weighted by the entropy of that decay mode within this volume. All decay modes that conserve the total mass, total charge and total energy are constructed. The entropy of each decay mode is assumed to be additive and is obtained from the entropy of the individual fragments. Both energies and entropies for individual fragments can be calculated from the free energy of the individual fragments [14] T2 T2 — T2 _ A 1 c 5/4A2/3 .0) + 8(——,.> 713+ (A — 2Z)2 + 32262 A 5 RAZ mnAT — T1n[g.znvo(—2;m—.)3” + T FAz(T) = (—Wo - [1 — (1 — n)_l/3] (1.11) ln(NAzl) NAZ , +25 where, W0 = 16MeV is the binding energy of nuclear matter; so is the inverse nuclear level density (50 z 16MeV for the Fermi gas model); TC = 16MeV is the critical temperature; R A z = 1.1714” 3 f m is the radius of a fragment corresponding to normal nuclear matter density p0; m N is the nucleon mass; g},Z is the spin degeneracy factory; K. is a model parameter, which is introduced since the individual fragments are as- sumed to move freely in a volume of [Cl/0; V0 is the volume of the system corresponding to normal nuclear matter density p0. The first two terms in the r.h.s. are the contributions from volume and surface; 11 the third term is from symmetry energy; the fourth term is from Coulomb interaction and the last two terms are from the translational motion. In this approximate mi- crocanonical approach the internal degree of freedom of the fragments act somewhat as a heat reservoir. The temperature T of the fragments are then determined by the equation for overall energy conservation. The equilibrium can also be more approximately studied within macrocanonical frame work, which provides some insightful approximate expressions. Under the Macrocanonical approach, the fragment yield for 2"" fragment in its k‘h state can be written as [1] 1.3/2 n -31 16,],(14, Z) = VIA; (2.12-,1C +1)eNfl 1‘2” (:8 NT’Z e—TL , (1.12) where, Y,,k(A, Z) is the yield of a given fragment with mass A and charge Z; B (A, Z) is the binding energy of this fragment; and J”. is the ground state spin of a nucleus with charge Z and mass A; pp and p" are the proton and neutron chemical potentials; A = (/27rh2 / m; V is the free(unoccupied) volume of the system and T is the temperature. Eq. 1.12 describes the primary fragment yields. The fragment yields from exper- iment are the yields after secondary decay. To minimize the effects of the secondary decay, in some cases, one can construct observables that are approximately insensitive to the secondary decay. One can construct the ratio of the fragment yields for the same isotope from two different reactions under similar conditions. By taking this ratio in the macrocanonical approximation Eq. 1.12, one finds that the volume and binding energy factor cancel out. If the feeding from the decay of particles of stable or unstable states are similar in the two reactions, it may cancel approximately as well. In this case, if we define R21 as the ratio of isotope yields from two different reactions 12 1 and 2, R21 = Y2(N, Z) / Y1(N, Z), the following isoscaling relationship [110] is found to be quite robust in nuclear reactions R21 or eanZ, (1.13) Where: a : (Alum/T) = (#722 - Mud/T and .8 : (App/T) : (”p2 “ [Jpn/T- Fig- 1.3 illustrates the isoscaling property observed with the fragments produced in the mid-rapidity region of 124Sn + 124Sn and 112Sn + 112Sn reactions [123]. Here, 124Sn + 124Sn is the reaction 2 and 112Sn + 112Sn is the reaction 1. I“ o 9.0!" OIQO I“ o Y(‘2‘Sn+ ‘z‘sm /Y(1128n+ 112Sn) .09!" cnszo Figure 1.3: Experimental isoscaling behavior exhibited by the central 124Sn + 124Sn and 11"’Sn + 112Sn collisions. The data are the nuclide yield ratios, R21(N, Z) from the two reactions plotted as function of N (top panel) and Z (bottom panel) 13 a and 6 can be shown to directly relate to the neutron excess 6, i. e. 0 oc (62 — 61)(1 — (61 + 62)/2) [111]. We are going to use a as an observable to study the density dependence of the asymmetry term and explore the isospin diffusion in this dissertation. 1.2.4 Expanding Emitting Source (EES) Model The achievement of equilibrium assumed in Sec. 1.2.3 can be achieved and maintained only if the emission of particles and their absorption by collisions with other particles are in detailed balance. Statistical surface emission provides a way to calculate the rates and to describe how an excited system might proceed towards equilibrium or away from it. In this section, we discuss the Expanding Emitting Source (EES) [37, 38] model which calculates the binary decay of excited nuclei via emission rate equations within a statistical frame work. This formalism, which allows the description of time dependent phenomena, integrates the contributions for each type of particle as a function of time over the de-excitation process. The formulae that describe the surface emission mechanism are similar to those developed by Weisskopf’s seminal paper on compound nuclear decay [119]. They are also conceptually similar to the ideas of Richardson’s thermionic emission. Let us consider the decay of a compound system labeled G into a daughter B, capable of further decay, and a particle labeled b, C —> B + b Let Nb be the number of emitted particles of type b. Then fi—fg expresses the rate of emission (at kinetic energy E in to a given interval dE) of that type of particle. The energy spectra can be obtained from the double differential quantity «in- We, dE " o dEdt dt. (1.14) 14 The emission rate is obtained by an energy integeration dNb_ °°d2Nb cfi_odwt dE , (1.15) and the multiplicity by a further time integeration °° dNb N=/-—fi. 11 b 0 dt ( 6) The rate of emission using the approach of Weisskopf [119] can be written as dN 2s 1 m = (72;. >EMa.....le/wci , (1.17) where M, E, and 3 refer to the mass, kinetic energy and spin of the emitted particle b; and w3(Eg) is the density of states for the daughter system B at excitation energy, E5, which remains after particle b is emitted with kinetic energy E from the compound system C, originally having excitation energy E5. The quantity 012+ B _) C is the cross section for the formation of system C which we take to be geometrical. For the basic EES calculation, initial temperature and excitation energy need to be specified for the EES calculation. The evaporation is complicated by other factors, i.e. the isospin of the excited system, the recoil momentum of the target, the reduction of the Coulomb barrier which accompanies charge emission as well as the range of temperatures traversed. A discussion can be found in [37]. The EES model allows expansion of the nucleus subject to its internal pressure and predicts a rapid emission of fragments once the density of the residue decreases below about 0.4p0. This suggests that the multifragmentation principally occurs after the system expands to subnuclear density. 15 1.3 Studies Prior to This Dissertation The isospin-dependent Boltzmann-Uehling—Uhlenbeck (BUU) transport model has successfully explained several isospin-dependent phenomena in heavy-ion collisions at intermediate energies [61]. In isospin-dependent Boltzmann-Uehling-Uhlenbeck (BUU) transport model, the isospin dependence was included in the dynamics through nucleon-nucleon collisions by using isospin-dependent cross sections and Pauli block- ing factors, the symmetry potential V023”) (p, 6) and the Coulomb potential. The isospin-dependent BUU model was used to calculate the ratio of yield of neutron and protons in pre-equilibrium emission [63]. It is found that if one drops the Coulomb and asymmetry term Vaggpr, 6) in the BUU model, the ratio of pre—equilibrium neu- trons to protons is almost a constant as a function of nucleon kinetic energy. However, when one includes the Coulomb and the asymmetric term of the EOS, the ratio of pre—equilibrium neutrons to protons changes monotonously as a function of nucleon kinetic energy. Fig. 1.4 has shown this calculation. One can then study the effects of the symmetry energy S (p) since the Coulomb potential is well understood. Since the symmetry potential Vajg") (p, 6) tends to make more unbound neutrons than unbound protons, one expects that a stronger symmetry potential leads to a larger ratio of free neutrons to protons. One can see from the figure that the effects of symmetry poten— tial show up at higher energy for central collisions and at lower energy for peripheral collisions. A naive picture of nuclear reactions at intermediate energies undergoes three im- portant stages: compression and expansion stage when the preequilibrium light par- ticles are emitted, after which a prefragmetation source is formed; the disassembly stage (evaporation) after the prefragmetation source reaches equilibrium; a secondary decay process of various emitted fragments. While the dynamic model such as BUU and SMF give a good description of preequilibrium emission during the compres- 16 4 Vvfi WW ,, v -- : i T f' i : .ooo”.‘+ [I 1 1 3 .' '3ZSn-i-‘328n '1 ’ i 2 ' 1 1 [ 4 4 - ; '24Sn+'24Sn O. a; 1 '4 z : 1 ill 2 ; .‘?00’-'..**** + j , z: 3i i "D f 1 ' . 4 ................... — - ] “28n+“28n —— F 1(U): * “2811-1-“2Sn 3,- l Fast Slow Tail Feta gate Logic for Feras ::> Figure 2.13: Schematic diagram of the electronics of the Miniball/Miniwall 40 2.4 The Ring Counter Foward Array Structure of the Ring Counter The Ring Counter Forward Array is an array of Silicon-CsI(Tl) telescopes. The sili- con AE detector elements are segments of a large annular strip silicon detector, the CsI(Tl) E detectors are elements of an annular array of 16 CsI(Tl) detectors. The active area of the Ring Counter is defined by the ring-shaped 280 pm(67.05 mg/cm2) thick silicon detector which forms the first element [34]. The inner diameter of the detector is 47.88mm, while the outer diameter is 96.10mm. The front side, called the Strip side, of the silicon detector is divided into four quadrants, and each of these quadrant is subdivided into 16 strips, each at a constant polar angle. So there are total 4x 16 2.8mm-wide strips. The back side, called Pie side, of the Ring Counter is segmented into 16 azimuthal pie slices. The silicon detector is mounted directly on a copper plate, which serves as a cooling bar for the silicon detector. Following the silicon, there are 16, 2-cm thick, CsI(Tl) scintillators closely packed inside of a plastic ring. The 16 CsI(Tl) crystals are aligned so that each CsI crystal is directly behind one of the pie slices. Finally, a thin 5.0 mg/cm2 Sn—Pb (60% Sn, 40% Pb) foil was placed in front of the array to absorb electrons and x-rays emitted from the target. Fig. 2.14 is a schematics drawing showing the various components of Ring Counter. Fig. 2.15 gives a side view of the Ring Counter. Principle of Operation The Ring Counter Foward Array identifies the charged particles using the same prin- ciples as LASSA. It uses the signal from the silicon as AE, the signal from the CsI(crystals) as E. Ordinarily, the signal from the pie side of the silicon detector is used for AE, and the signals from the Strip side gives the polar angle and those from the Pie side gives the anzimuthal angle. Fig. 2.16 shows the correlation between the 41 k \\ Strip side of Si Pie-sector side of Si \,\ / \l‘f VK CSI ”1 (T1) Crystals Figure 2.14: A schematic drawing of the Ring Counter 42 AE UV Lucite l CsI(Tl) Pin Diode Annular Si Sn-Pb foil Figure 2.15: A schematic figure of the side view of Ring Counter 43 energy loss signals measured in the pie side of the silicon detector verses the light output signals measured in the corresponding CsI(Tl) detector. 4000 - DE (Raw ADC Output) 3 N 5.” "’ R 8 § 8 s s §. ‘— 3 ”2000 L 3500 4000 0 ‘1 1 . 1 L I 0 500 1000 1500 20002500 3000 J. CsI (Raw ADC Output) Figure 2.16: A two dimensional histogram of the raw signal from Pie side and CsI Electronics of Ring Counter The pie side is the p-type side and the strip side is the n-type side. The diode de- tector is reversely biased. Each strip or pie slice is read out independently. Similar electronics were used to process the signals from the Ring Counter Forward Array as 44 for the LASSA telesc0pes. Lower gain LBL preamplifiers were used for Ring Counter silicon signals, however, because the energy loss of the projectile-like fragments is very large. A schematic of the electronics is found in Fig. 2.17. The electronics used to read out the 16 channels of the azimuthal pie-slice segments and the 64 channels of the polar-angle strips are similar, the only differences are the polarity of the signals and the additional requirement that the pie sector provides the timing information. Although the preamplifiers of the signals from the 16 azimuthal pie sectors generate both a Fast and a Slow outputs, we injected the Slow output into a Washington University (WU) shaper-discriminator module, which contains a second stage slow- shaping amplifier and a fast timing-filter amplifier. The fast timing—filter amplifier output was used as the trigger to the experiment. The outputs of the fast amplifiers of the 64 polar-arc segments and the 16 CsI were amplified in a second stage slow- shaping amplifier in a single-width WU shaper. All the outputs from the slow-shaping amplifiers for the strips, Pie’s and CsI’s were sent to the Phillips 7164B 4096 channel peak-sensing ADCs. Only those signals above the low-threshold of the peak-sensing ADC were digitized and read by the DAQ. 45 _— Slow Shaping _4 Peak sensing Strip __ Preamplifier Pie-Sector _S Preamplifier CsI H Preamplifier Amp ADC Trig for DAQ Timing Filter Amp __ DGG _, Slow shaping Peak sensing Amp '— ADC Slow shaping Peak sensing Amp b ADC Figure 2.17: A block chart of the electronics of the Ring Counter 46 2.5 Data Acquisition Electronics The Data Acquisition(DAQ) for this experiment was a big challenge. Strip-silicon detectors usually require a lot of electronics. In this experiment, there are 9 x 16 DE signals, 9 x 16 EF signals, 9 x 16 EB signals, 64 Strip signals, 16 Pie signals, for a total of 512 energy signals from the strip detectors. In addition, there were 36 energy signals from the CsI(Tl) detectors as well as 144 timing signals from the EB strips. We also need to incorporate Fast, Slow, Tail and Time signals from the roughly 200 Miniball/Miniwall phoswich detectors, which corresponds to 800 electron channels. To reduce the dead time and incorporate LASSA and the Ring Counter electronics as well as Miniball/Miniwall electronics, all channels were read out via the F ERA ECL readout scheme. The Miniball/Miniwall electronics used FERA 4300 ADCs which could be read out more quickly via the ECL (Emitter Coupled Logic) port. This is a double-bus readout scheme, with one bus for command and another for ECL port data. A rate of up to 100nsec/word can be reached, comparing with the typical CAMAC hardware cycle of 1 ,us. To make use of the ECL port readout, there are some problems needed to be solved: First, Phillips ADC does not have an ECL readout port like FERA. To solve this, an auxiliary CAMAC module, called 8832 is used. This CAMAC module is designed by J. Toke of University of Rochester. 3832 works as a inter-media between Phillips ADC and Lecroy FERA Driver 4301. It acts like a very fast CAMAC controller to the Phillips ADC, and outputs the data via the ECL port like a standard Lecroy 4300 FERA. All the data from the FERAs and 8832 were sent via the ECL port to a FERA Driver 4301. In more detail, the 8832 collects the data from Phillips ADC, handshakes with FERA driver 4301, then puts the data to the data bus. Every converter executed a hit-pattern—based sparse readout of up to 24 Phillips model 7164 ADCs via the standard CAMAC protocol. This non-standard readout effectively moves the data 47 and control cords from the back plane of the CAMAC crate, to the front, ECL ports of the 8832. The non-standard readout, which required no hardware modifications to the 7164H ADCs, or the CAMAC crates, resulted in a 7.5-fold reduction in the readout time of the LASSA electronics as compared to a standard CAMAC readout. The old Miniball electronic design is tree-structured; one problem that needed to be solved is that FERAs were not side by side as needed for a bus-structured readout. Special cables and connectors are made to solve this problem. A schematic drawing of the topology of the electronicsis shown in Fig. 2.18. The readout for the Miniwall was analogous to the Miniball and the readout for the Ring Counter was analogous to the LASSA, so the readout for the Miniwall and the Ring Counter were not shown in this figure. In Fig. 2.18, the readout has a two-level scheme. The lower level is the FERA modules and 8832 modules which are controlled by FERA drivers; the FERA drivers, in turn, are controlled by a VME module Fera Faucet Maier (FFM). In Fig. 2.18, those rectangles with thick black box lines are the CAMAC crates. Those lines connecting the modules in the CAMAC crates denote the bus lines for control signals and data. The signals from the Miniball detectors come from the upper-right corner and go directly to the splitter. The trigger signals from the splitter go through the discriminator and the multi-box to the main trigger control module OCF 8000. The outputs from OCF 8000 are sent to FFM, different ADC and TFC modules. The multi-box is used to determine the coincidence multiplicity of Miniball trigger, which we set to 3 in this experiment. The Fast, Slow and Tail signals from the splitter are sent directly to the corresponding FERAs in the CAMAC crates. The signals from the LASSA are from the left side of Fig. 2.18. They first go to a double- width WU shaper and discriminator/TFC module. The outputs of WU shaper are sent to the Phillips ADCs while the outputs of TFCs are sent to FERAs for timing signals. The data in those Phillips ADCs are collected and sent to FERA drivers by 48 S832. The principle of ECL port readout can be understood from Fig. 2.19. For the same reason as Fig. 2.18, only the parts for LASSA and the Miniball are shown. Fig. 2.19, one sees a microprocessor, the “transputer”, developed by Inmos company, was used as the front-end computer. CES High Speed Memory (HSM) 8170 is a standard VME memory module. FFM integrates the CES 8170, FERA drivers and ECL readout modules. The usual trigger used for this experiment was the observation of a multiplicity in the Miniball/Miniwall > 3 or with the Ring Counter trigger, downscaled by a factor of 25. Zero compression is used for all ADCs to suppress all data words which are consistent with the experimental pedestal in the individual ADC channels. In upper part of Fig. 2.19, one can see that the electronics for the Miniball array are basically the same as what is shown in Sec. 2.3. The main difference is in the lower part, where one sees that the output signals from delay gate generator (DGG) are sent to the gate input (GAI) of F ERA driver, instead of the gate input of FERA. The LASSA elctronics, as described in Sec. 2.2 are incorporated with the Miniball electronics. Also, the busy outputs from CBS 8170 and FFM are added to the busy logic boxes. 49 ECL Readout “PM MN? for ADC; DE Matti-box DISC 7 EF EB CsI Time EF :umuadunj Miniball Lassa Figure 2.18: Layout of the data acquisition (FD means FERA Driver) 50 a Miniball ll Driver Terminator Figure 2.19: Schematic drawing of the principle of the ECL readouts 51 Chapter 3 Data Analysis 3.1 Data Reduction Algorithms As discussed in the last chapter, three types of devices were involved in the exper- iment: LASSA, Ring Counter and Miniball/Miniwall detectors. Miniball/Miniwall detectors are only used for impact parameter selection, so the energy calibration was not done for Miniball/Miniwall detectors. However, a rough PID function was gener- ated to separate LPs (light charged-particles with Z <3) and IMFs (heavier fragments with Z Z 3). For LASSA, the calibration was done for isotopically resolved particles. For the Ring Counter, the calibration was done for atomically resolved particles. Since both LASSA and Ring Counter are composed of silicon-strip/CSI telescopes array, the data analysis for these two devices are quite similar. This data analysis is basically divided into two parts. One consists of the analysis of the signals in the Silicon detectors while the other one is the corresponding analysis for the signals in the CsI detector. For the Silicon detector part, the most difl‘icult issue is the multiple hit correction. In multiple hit events, two or more particles hit the same telescope at the same time. When a multiple hit happens, there will be more than one way to pair up the signals from the vertical and the horizontal strips, the task is to determine the 52 correct pairing. For the CsI detector part, the main issue is the energy calibration. Since the lightput out of CsI(Tl) depends on the mass number A, atomic number Z and energy of the charged particles, different calibrations must be done for each isotope (LASSA) or each element (Ring Counter). We will talk about details of the data analysis of LASSA and the Ring Counter in the following sections. 3.2 Data Analysis of LASSA 3.2.1 Overview The energy calibration and the generation of the PID function were the first two steps of analysis for the data analysis of the LASSA telescopes. The LASSA array was calibrated in three stages. The first stage involved cal- ibration of the device in the running configuration with a precision pulser and a charge terminator. In the second stage, a series of fragmentation beams was used and each detector was repositioned to allow the beam to be put directly into the telescope. Beams of 20 MeV/u 16O and 40 MeV / u 36Ar were accelerated by the K1200 cyclotron at Michigan State University and used to bombard a production target at the exit of the cyclotron. Reaction products from projectile fragmentation were subsequently selected according to magnetic rigidity by the A1200 magnetic channel. These low intensity secondary beams (~1300 p/s) were then used to directly scan the detector face. The beams used are listed in Sec. 3.2.5. These beams were useful in assessing the thickness non-uniformity of the 65 pm Si detector, as well as in calibrating both of the Si detectors and the CsI(Tl) detector. Finally, proton recoils used for proton calibration was measured by bombarding a Polyethylene target (CH2) with a 112Sn beam. Two sets of PID functions were generated for LASSA: A Si-Si PID function is 53 generated for particles stopping in the second silicon detector, a Si-CsI PID func- tion is generated for particles stopping in the C81. To generate added 2D PID his- tograms without losing resolution, one needs to do the correction for the thickness non-uniformity of silicon detectors as well as the gain-matching of the CsI detectors. The thickness non-uniformity correction is done for each “pixel” defined by the inter- section of a given front and back strip in the 500pm detector. In order to properly assign particles to their respective pixel, a multiple-hit decoding routine is needed to determine which pixel the charged-particle hit. This routine will also give the posi- tions and energies for each particle when more than one particle hit the same detector at the same time. After the energy calibration is done and the PID functions are generated, the full information for each charged particles detected by LASSA telescopes can be obtained and written to a “physics tape”. This information includes: the Particle ID, the total kinetic energy and the emission angle. In order to obtain the final yields and energy spectra, background from noise and random coincidence should be subtracted. The LASSA array also has a limited solid angle coverage. Thus, the detection efficiency should be calculated and corrections for the efficiency should be made. The final yields are then obtained by applying various cuts to these efficiency corrected data. Fig. 3.1 gives an overview of the scheme of data reduction of LASSA. It also shows the sequence following which the analysis is done. We will discuss the data analysis details according to this sequence. 3.2.2 Energy Calibration of Si For the range of charged particles we analyze here, the signal generated depends linearly on the energy deposited by the charged particle, independent of the type of 54 Raw Data on Tapes a? \ \{iort Raw Data onjafisr/ [ DE Raw Output EF Raw Outpud [ EF Raw Output CsI Raw Output 1 . 1 1 r Subtract Pedestal Subtract Pedestal Subtract Pedestal Subtract Pedestal V Y ' DE Calibration® EF Calibration @ EF Calibration ® 1 f 1 fi T fi N F a Dead Layer Dead Layer Dead Layer [ Correction Correction J k Correction 6 1 Match EF & EB, Get rid of 1% difference 1 l Multiple-hit Decoding, get hitting position, final EF/EB Energy i i DE Non-unifonn EF & EB Non-unifonn C I G . M t h thickness Correction thickness Correction 5 am a c Silicon PID(PID1) CsI PID (PID2) CsI Calibration® @3, EF & EB Dead Layer CorrectiB Physics Data(Energy, angle PID) o© ® Ortec Pulser + Monte Carlo Simulation (Background Subtraction J charge terminator Generate Efficiency ® Fragmentation Beam F unction Efficiency :Correction) And Proton Recoil Final Yield, Energy, Angle ) Figure 3.1: Major detectors used in the experiment, viewing in the downstream direction of the beam 55 the charged particle detected. Indeed, we expect an electron-hole pair to be collected from the detector for each 3.1 eV of energy deposited. Based on these facts, we can simply calibrate the energies of particles detected in the silicon. Specifically, we cross- calibrate a pulser system by a source detected in the silicon detector at low energy. Then we extrapolate the calibration to the whole energy range of Silicon with a linear pulser system. This pulser system consists of two parts: an Ortec precision pulser generator and a precision capacitor that is placed between the pulser and the cable that connects the detector to is preamplifier. The Ortec precision pulser generates precise pulser signals, with amplitude that is linearly dependent on its dial setting. As the input to the preamp is a “virtual ground”, the capacitor value sets the scale of the injected charge. In our experiment, we used the following: 1) A PC based data acquisition system was used to calibrate pulser system relative to known energies of 228Th and 241Am (1 sources. Fig. 3.2 is a schematic drawing of this PC-DAQ set. The energy of the 0 particles from these 0 sources are relatively a Source 228 Th 01' 24lAm PC DAQ I 7 J; 0 l" PA Shaper Si detector Vacuum chamber Figure 3.2: The PC-DAQ system used for the silicon calibration low, comparing with the energy range of particles detected in the experiment. To 56 calibrate the pulser dial setting, we set the amplifier gain to put 8.784 MeV into Channel 2220 of a 4k channel ADC. Then, we measured the pulse height spectrum for the pulser for a variety of dial settings on the pulser. 2) This relation between the ADC spectrum and the pulser dial setting can be expressed as: ADC readout = a X (Dial + b1) + b0 (3.1) where, a is a pulser attenuation factor, be is the channel number corresponding to zero pulser height from the pulser, bl is a constant describes how the zero may change with pulser setting. 3) Using 1) and 2) and several attenuation and dial settings we obtained the relationship: Energy vs. Dial. Since the gain settings for DE, EF and EB are different, calibration using different attenuation factors were needed when the dynamic energy ranges for various detectors differ. Three different dynamic ranges for the pulser calibration, 140MeV, 200MeV and 500MeV, were used to calibrate the DE, EB and EF signals. Fig. 3.3 shows the relationship between the Dial and the Energy for the DE, EF and EB pulser calibrations. 4) At the end of the experiment, we coupled the pulser capacitor to the cable between the Silicon and Preamplifier by using a T—connector. Everything includ- ing detectors and DAQ were maintained under the same conditions as during the experiment. 5) Then we changed the dial of the pulser, to get a relationship: Energy vs. the channel in the ADC, used in experiment, until the full response over the full energy range of the detector is obtained. This full energy range depends on the gain settings of preamplifier and shaper. Normally, the linearity is good for energies in the middle of the energy range, but gets worse at low energies, when the ADC output is close to zero channel or at high energies where the amplifier become non-linear. To address 57 this effect, more calibration points are done for ADC output below 100 and above 2500 channel in the 4095 channel Phillips ADC. Then we fitted the calibration with a linear function for ADC output channels 0~100, a second linear function for ADC output channels of 100~2500, and a polynomial of order 4 for ADC output channels of 2500~4095. Between those fitting functions, a spline interpolation is used to make a smooth connection. Fig. 3.4 shows the calibration line for a typical Silicon strip. Step 4) and 5) of this procedure was repeated for all 432 strips of AE, EF and EB one by one. After the calibration is done, the calibration is checked by plotting the energy of EF vs. EB. If the energy calibration is correct, with no noise, the energy from EF and EB should be exactly the same for each particle. However, we had about a 1% difference in the energy between EF and EB after our calibration. The reason for this is not clear. Possible reasons are: a difference in the efficiency of charge collection between EF and EB or an systematic error in the attenuation setting of the Ortec pulser. To correct for the discrepancy, we forced the energy from Different Dynamic Range Settings for DE, EF and EB 16 . ._. . ._ _. _m_ ##### ,"‘ jfifih ,- -, _.. ....._. ., 14 :SOOMeV Dynamic Range l y] = 49.70216x + 0.0285 12 e10 s g . H W 1 E 8 il40MeV Dynamic Range ] , I g 6 y3= l.93486x + 0.0l483 ] 4 - - ~- —~~ »—- ~ [200 MeV Dynamic Range 7 2 y2= 20.7257lx + 0.01723 0 O 0.2 0.4 0.6 0.8 Pulser Dial (Volt) Figure 3.3: Calibration lines of energy(MeV) vs. Dial value(Volt) 58 Energy Calibration for One Silicon Strip _a N _ 4 3 2 10 y-a3x +b3x +C3X +d3x+e3 Ea :: y=a2x+b2 36 G '6 +34 Y=31X+b1 a 2 0 0 1000 2000 3000 4000 ADC Readout Figure 3.4: ADC output vs. precise pulser dial value(Volt) the EF channels to be equal to the energy from EB channels, which we took to be more accurately matched to the a calibration points. 3.2.3 Multiple-hit pattern Decoding Before one can get the PID function, some preparations need to be done. Most important preparations are the Pixelation (Multiple-hit Decoding) and Non-uniform Thickness Correction. In LASSA, strips of EF and strips of EB are perpendicular to each other. The scattering angle of particle is determined from the X, Y coordinates where the particle hits the telescope. Here, X and Y are basically determined by which EF and EB strip is fired. (We use X to denote position information obtained from EF and Y to denote position information obtained from EB.) When only one particle hit the telescope, one can easily tell X and Y values from the energies detected in the strips. However, when two or more particles hit the telescope at the same time, several EF and EB strips will have data. For example, when two particles hit one telescope at the same time, there will usually be four 59 signals X1, X2, Y1, Y2. Sometimes, there would be fewer if one of the strips in the detector is not working properly or the two particles happen to hit the same strip. If four signals are observed, one need to know whether the two particles hit (X 1, Y1) and (X2, Y2), or in coordinates (X1, Y2) and (X2, Y1) instead. One needs also to know the amplitude of the signals (corresponding to energy loss) induced by these two particles. Most of the times, the position and energy information can be retrieved by requiring the signals from EF and EB to be equal. But sometimes, this comparison does not yield a unique answer. For example, if the signals (energy loss) induced by the two particles happen to be the same then both parings (X1, Y1) and (X2, Y2), or (X1, Y2) and (X2, Y1) are equally favored. However, when the signals from AE and CsI are included in the paring, it is often possible to determine which paring is the correct one. Thus in practice, various hitting patterns must be tested to find the most probable one. Occasionally, more than two particles hit the telescope at one time. We have not found a reliable algorithm to analyze such events. Instead we reject them and count them as lost. When there are bad strips in a telesc0pe and particles hit these strips, the re- construction fails and the information from that particle can not be retrieved. When either position or energy information for a particle are irretrievable in a strip detector, this particle is also lost. Other losses occur: when two or more particle go into the same CsI crystal, then, the energy information is lost because the signals induced by multiple particles superimpose on each other in the CsI and cannot be separated. Also, particles are lost when they hit at the gaps between the C31, and no valid signals for CsI can be constructed. Another output of the Multiple-hit decoding program is a record of the fraction of particles which are lost. This information is then used to constrain the efficiency 60 correction. More details about efficiency correction can be found in Sec. 3.2.7. 3.2.4 Construction of PID function and Non-uniform Thick- ness Correction When constructing the PID function, it is desirable to make it applicable to all two dimensional (2D) AE vs. E histograms. However, when one does this, the isotopic resolution of the 2D spectra becomes rather poor. This occurs, because the measured values for AE and E depends on the position where the particle hits the telescope. Part of this position dependence comes from the thickness variation of the silicon detector across its surface. For AE, the thickness variation is of the order of 10 pm, causing a position dependent variation in the energy loss of the particles and smearing the relationship between AE and E values. Only if the thickness variation is of the order of 1%, can one calculate the PID without addressing this position dependence. Second, when the 2D PID histogram is constructed by using silicon as AE vs. CsI, it is easiest to use the uncalibrated output of the CsI detector. However, the uncalibrated output from the CsI depends on the gain setting of the electronics, so this signal will be crystal related. To obtain a PID function that is applicable to all telescopes without losing reso- lution, we should correct for these effects. For the silicon detector, we assumed that the energy loss a particle experiences in passing through the detector is proportional to the detector thickness, which is good for relatively small energy losses. Then we determined the relative thickness of different AE by using particles of fixed energy and measuring the energy losses in the silicon. This can be done if you have a beam of defined energy or if you measure particles that stop in the second silicon. In the latter case, you can gate on the total energy deposited in the silicon. Such gates applied to 01 particles were quite important for establishing the variation of AE thickness with 61 position. In general, the thickness correction is done for each “pixel”, defined by the inter- section of a given front and back strip in the 500 pm detector. In this fashion, we obtained thickness corrections for both the 65pm silicon detector and the 500nm sili- con detectors. (Note, the non-uniformity is less crucial for the 500pm silicon detector, which is of the order of 1%). After the non-uniform thickness correction, we define an effective “energy” so that particles with a given energy have the same “effective” energy loss in the silicon detector regardless where it hits. Then we applied the PID function to all telescopes. The effect of making this non-uniform thickness correction is shown in Fig. 3.5. On the Left panel of Fig. 3.5, the PID of Lithium and Carbon isotopes are obtained without making the thickness correction. One can see that the resolution is bad. On the right hand panel, the thickness correction is performed. A much better resolution is achieved. A comparison of the left panel and right panel of Fig. 3.5 illustrates the crucial importance of correcting for the thickness variation in the 65 pm detector in achieving optimal particle identification. Details of the construction of the 1D PID function will be discussed later in this section. For the CsI detector, the gain match is done by matching 36 CsI crystals of LASSA to one selected crystal, using Helium and Carbon calibration points from a beam calibration run. After the silicon detector calibrations, the non-uniform thickness correction for the silicon detector and the gain matching of the CSI crystals. 2D AE vs. E PID histograms were constructed. These PID histograms are now independent of the po- sition where the particles hit the detectors. In practice, different measured quantities were used for AE and E in order to construct the more accurate PID functions. For example, the DE detector is used as AE and the value of EF is used as E for particles 62 500 — 7Li 7Li Before Correction Li After Correction 0~6789 6789 12c 130 Before Correction 12c 13c Amer Correction 100 111213141516A 111213141516 Figure 3.5: Comparison of isotope resolution before and after thickness correction that stopped in the second silicon. The low EF gain allowed higher Z fragments to be properly identified. For p, d, t, 3He, 4He and 6He that stopped in the CsI crys- tals, EB is used as AE and the matched raw output from CsI is used as E. The EB signals, which used higher gain shaping amplifiers, improved the identification of the isotopes of Hydrogen and Helium at the expense of dynamic range. The identification of higher Z fragments required the use of the lower gain EF signal. For isotopes with Z > 3 that stopped in the CsI crystals, DE+EF is used as AE and the matched raw output from CsI is used as E. A typical DE—EF two-dimensional spectrum for a telescope is shown in Fig. 3.6. In this figure, the measured energies in DE and EF detectors are matched first, which we call them nominal energies, and then are used to construct the 2D histogram. PID gates are drawn from this histogram. Compared with the raw energy spectra shown in Fig. 2.6 in the previous chapter, CsI signals 63 are used to veto the punch through lines. A typical DE+EF-E(CsI) two-dimensional spectrum for a telescope is shown in Fig. 3.7. The isotopes of oxygen can be clearly distinguished in this figure. PID gates are drawn from this histogram too. The 2D PID histogram is not convenient. To make various gates for different isotopes as well as to estimate the background level, it is desirable to work on a 1D histogram. One cannot simply project the 2D histogram to either X or Y since the PID lines are curved in the 2D histogram. However, one can straighten the 2D PID histogram by drawing curved line along each isotope and projecting the 2D histogram according to these lines. Because the real isotope yields are not symmetric about these lines, this projection can cause some problems in the normalization, making it difficult to fit the curve of isotopic yield. Fig.3.8 show the 1D PID histogram for particles stopped in the second silicon detector. Fig. 3.9 show the 1D PID histogram for particles stopped in the CsI crystals. 64 120 DE (Matched Energy in MeV) ON c: .Hiliiiiliililiiiiliiiiliii 0 50 100 150 200 250 300 EF (Matched Energy in MeV) 1l1111 350 400 Figure 3.6: A histogram of counts vs. nominal DE energy and nominal EF energy 65 DE+EF (Matched Energy in MeV) ‘ ~-——'_ _. __ a, . _._____-::a-—--- 00 500 1000 1500 2000 2500 3000 3500 CSI (Matched ADC Output) Figure 3.7: A histogram of nominal counts vs. nominal DE+EF energy and matched CsI light output. The isotopes of selected elements are indicated by labels that are immediately above the corresponding data. 66 Counts(a.u.) lljllll llllllllllJllllllllllllllJlllLl 0 500 1000 1500 2000 2500 3000 3500 PID (a.u.) Figure 3.8: 1D PID histogram for particles stops in the second silicon detector 67 Counts(a.u.) j—d O O\ 1 l r l O 500 100015 I I 1 l 00 2000 2500 1 l 1 1 3000 3500 4000 PID (a.u.) l Figure 3.9: 1D PID histogram for particles stops in the CsI crystals 68 3.2.5 CsI calibration The energy calibration of the CsI crystal is more difficult than that of Si detectors, because the light output of CsI(Tl) crystals depend on the charge, mass and energy of the detected particle. That is why one must construct the PID function before one performs the energy calibration of a CsI crystal. The light output for a particle with a certain charge and mass is a non-linear function of the energy loss for heavy ions and low energies. To address this, a non-linear energy-to-light relation was established. That depended both on the charge Z and mass A of the particle. Many different empirical functions have been employed to construct energy cal- ibrations of CsI(Tl) [47, 22, 72]. These parameterization typical are approximately linear at high energies but become very non-linear for low energies. Some of these parameterizations have both a charge and a mass dependence while others are simply charge dependent. The best choice of calibration function depends on the T1 doping of the crystals. For the crystals used in LASSA in this experiment, four calibration functions were used depending on the particle type [118]. These functions are similar to that proposed by Y. Larochelle et al. [56]. Larochelle’s calibration function has the at- tractive feature that the deposited energy E is expressed as an analytical function of the light output, which is good for calibration purposes. It has an explicit mass dependence, necessary for the calibration of heavy ions. We discuss the other details in the following. Following Larochelle, we parameterized the incident particle energy E for heavy ions of Z>3 as a function of the light output L, the charge Z, and the mass A of the particle, as follows: E(L, A, Z) = aAZzL + b(1 + cAz2)L1-d~/7*Z . (3.2) This expression has a linear part, that dominates at high energies and an exponential 69 part that dominates at low energies. In Eq. 3.2.5 (1, b, c, d are the fitting parameters which can be determined by fitting experimental points. The parameters a, b, c depend on the electronic gain and scintillation efficiency. The parameter a (affecting the linear part) is important for small Z and high energies. The parameters b and c (weighting the exponential part) describe the non-linearity at low energies and higher Z ’8. All parameters in Eq. 3.2.5 are positive. The parameter d is responsible for the transition from non-linearity to linearity in the energy response. Larochelle identifies L with the slow component (7 [us time constant) of the light output. In our parameterization, L is the total light output instead. In Fig. 3.10 the fitting curves as well as the calibration points are shown for Carbon. In Fig. 3.10, the solid and dot-dashed lines represent the best fit of Eq. 3.2.5 to the experimental energy calibration data corresponding to different carbon isotopes (A=11-14). The need for a mass dependence can be demonstrated by examining the light output of the higher energy carbon isotopes. At high energy, the light response is expected to be linear. Both the 11C points should lie in the linear domain. However, a straight line joining the two 11C isotopes does not pass through the high-energy 12C, 13C, and 14C isotopes. A curve going through all points for the 11‘14C would lead to a very large and unreasonable curvature compared to calibration procedures adopted elsewhere in the literature. This mass-dependent calibration curve appears to be a more reasonable solution. Since several fragmentation beams would be required to have the full calibration curve for each isotope, we adopt the mass-dependent ansatz (closely related to the quenching effect) of Larochelle [56]. For light-charged particles with Z33, the parameterization described in Eq. 3.2.5 did not accurately describe the detected energies. Compared to the observation of Larochelle et al. [56], a less pronounced isotopic effect was observed for light ions. This may be the result of the increased concentration of the activator element, T1, 70 600 400 ECSI(MeV) ZOO l1. O 1000 2000 Light(a.u.) l l l 1 Figure 3.10: Calibration curves of Carbon for LASSA CsI(Tl) in the CsI-crystals used in the present study compared to those studied in Ref. [56]. We find the mass dependence to be overestimated by the AZ 2 factor in Eq. 3.2.5, and employ a modified function of Eq. 3.2.5 with a weaker dependence on A to fit Z 53 particles. This will give us the following three modified calibration functions for light particles: For lithium (Z = 3) particles, we changed the first term of Eq. and used the follows: E(L, A, Z) = ax/AZzL + b(1 + cAZz)L1-d\/7‘Z . (3.3) 71 For helium (Z = 2) isotopes, the following calibration equation was used: EXL,A,Z)==aL-+bA%1——d”). (34) The variables a, b, c, and d in Eq. 3.2.5, 3.3 and 3.4 are fit parameters. There are sufficient data to reproduce with good accuracy the light-output response for all the isotopes of the same element using Eq. 3.2.5 - 3.4. As we have only two calibration points for each isotope, p, d and t, we adopt the simple linear function for Z = 1 particles. E=aL+b, (am where a, and b are fitting parameters. A linear CsI(Tl) response is consistent with that observed for hydrogen isotopes by Handzy [45]. Table 3.1 show the fitting parameters for all the isotopes. Z A a b c d 1 1 0.2010 -0.9587 1 2 0.1916 -l.214 1 3 0.1784 0.8212 2 0.1696 4.575 0.3380 -0.05772 3 0.01783 0.2456 0.09743 0.06358 4 0.0006680 0.4493 0.01015 0.02616 5 0.0006354 2.944 0.002940 0.03713 6 0.0004102 2.362 0.003566 0.02869 7 0.0003101 4.834 0.002213 0.02734 8 0.0002289 4.701 0.002042 0.02301 Table 3.1: Table of the fitting parameter used in the energy calibration functions of LASSA CsI(Tl) crystals The calibration functions used above have been used to calibrate 36 CsI (Tl) crys- tals of LASSA. To provide the calibration points, fragmentation products ranging from hydrogen to oxygen isotopes produced in fragmentation beams and direct a and proton beam particles were used. The detectors were directly exposed to low- intensity (1000 particles / s) beams of different isotopes and energies. These ions were 72 obtained by fragmenting 2160 MeV 36Ar and 960 MeV 160 primary beams from the NSCL K1200 cyclotron in the A1200 fragment separator [90]. The main advantage of this method is the availability of a large number of particles that could be detected simultaneously (up to 52 isotopes were identified in the case of the 36Ar fragmen- tation). Since particles are selected only by their magnetic rigidity (Bp=l.84l Tm for the 36Ar beam and Bp=1.295 Tm for the 16O beam) one obtains a broad range of different isotopes and energies. The FWHM of the momentum widths for these particles were selected to be 0.5%. The atomic and mass numbers as well as energies of the particles used to calibrate the CsI crystals in the present work are listed in Table 3.2. Hydrogen and helium isotopes were also calibrated by elastic scattering of E /A=30 MeV p-4He molecular beams on a Au target and by 240 MeV direct 4He beam particles. Our fitting procedure resulted in a precision of the energy calibration better than 2% for isotopes from 4He to 200. 3.2.6 Background Subtraction The spectrum shown in Fig. 3.6 is for all particles that do not have a CsI (Tl) signal; i. e. particles that either stop in the 500nm Si or “punch-through” 500nm Si but somehow escape the detection by the CsI(Tl) crystals behind the silicon detector. The low intensity background haze visible in Fig. 3.6 between the element groups is due to particles passing through inactive regions (edges, inter-strips, etc.) of the telescope, resulting in inconsistent DE—EF combinations. Also visible in Fig. 3.6 is a faint band of particles with a positive slope at the high energy event of the isotope lines between He and C. Particles in this band “punch-through” the 500pm Si detector but are not rejected because no signal is observed in the C31. These particles are present due to a slight misalignment of the Si detectors relative to the CsI(Tl) crystals, or due to 73 16O fragmentation products E(MeV) 36Ar fragmentation products E(MeV) p 77.17 (1 39.78 d 79.57 1: 26.72 t 53.75 3H6 105.00 3H6 210.00 4H6 79.99 4H6 160.00 6H6 53.64 6H8 107.90 6L1 119.90 6L1 240.00 7L1 103.10 7L1 206.80 8L1 90.40 8L1 181.60 7Be 182.20 7Be 363.40 9311 142.50 9311 285.60 loBe 128.40 loBe 257.90 lOB 199.90 1013 400.00 11B 182.10 11B 364.90 12B 335.40 11C 261.20 “C 521.60 12C 239.90 12C 480.00 13C 221.80 13C 444.40 ”C 413.70 14N 279.90 14N 560.00 15N 524.00 l6N 492.40 150 340.80 150 680.70 160 640.00 170 603.70 180 571.30 Table 3.2: Table of fragmentation products used in the energy calibration of LASSA CsI(Tl) crystals the dead area between adjacent CsI(Tl) crystals (due to wrapping for the purpose of optical isolation). One is able to eliminate this contribution, at the expense of geometric efficiency, by gating selectively on the position in the Si detectors. The background haze visible in Fig. 3.7 has the following origins. First, particles passing through the inactive regions of the telescope give rise to spurious EF-E(CSI) values. Also, coincidence summing of two or more particles detected within a sigle CsI(Tl) crystal is particularly evident in the low EF region. One source of these double hits may be that high energy light particles in combination with a second particle 74 in a single CsI(Tl) crystal distort the E(CsI) signal. A second source of coincidence summing occurs when neutrons hit the CsI at the same time as a charged particle. This produces signals which superimpose on the signals of the charged particles and shift the PID lines. Higher order coincidence summing as well as noise from either the detector or the electronics can shift the PID lines as well. It is very hard to make correction for the signals causing the background, i.e. to separate the noise from the real signals. A more practical way is to make an esti- mation of the background level and correct for the background accordingly. There are two types of corrections. One is the subtraction of background under the PID curve of each isotOpe and the other is the correct for losses due to coincidence sum- ming that remove particles from a given PID line and move them elsewhere. We will discuss the background subtraction first. We studied the background and observed that the background level depend on three factors. First, the coincidence summing probabilities are detector related: charged particles have different coincident sum- ming probabilities when they stop in the Silicon detector than when they stop in the CsI crystals. Second, the background level is energy related: charged particles with different energies have different background levels. Finally, the background level is also isotope related: the background level is somewhat different for different isotopes. For most of the isotopes, we estimated the background as follows. For central colli- sions (the event selection will be discussed in the next chapter), we generated 1D PID histograms for different detectors and in ten different energy bins, then background levels were estimated for each elements. For peripheral collisions, since the statistics is not high, we generated 1D PID histograms for different detectors, then background levels were estimated for each elements. We treated 7Be, 3He, 11C, 15O separately, because their background levels are not the same as other isotopes of the same charge number. For example, 7Be has no neighboring isotopes, so the background level of 75 7Be was estimated separately. For 3He, 11C, 150, there were significant leakages into these PID lines because their yields are weak compared with their neighboring 4He, 12C and 16O isotopes. Fig. 3.11 show the contamination of 3He by 4He. To obtain the yields of 3He, 11C, 150, we only used the halves of the PID lines which are lo— cated at the far sides of their neighboring isotopes, which are believed to be free of contamination. Fig. 3.11 also shows that the PID line of 3He is asymmetric. The reasons for that are: a) The PID algorithm does not requires isotope yields to be strictly symmetric about the middles of PID lines. b) There is a contamination of 3He by 4He. Fig. 3.12 shows that the background levels for IMFs as estimated by those fitted straight lines in the figure. 3.2.7 Efficiency Correction Ideally, if the detectors have perfect 47r coverage and can detect charged particles with any energies, there is no need for an efficiency correction. However, LASSA only has a limited geometric coverage and can only be placed at certain angles. For example, we can not detect those particles emitted in the direction very close to the beam line, since we can not put the detectors too close to the beam line without damaging the detector. We also cannot detect particles with energies that are below the detector thresholds or above the dynamic range of the electronics. This type of efficiency was easily noted but not easily corrected because we cannot definitely correct for data loss at angles where the efficiency is zero. Even if the particles are emitted in the directions covered by the LASSA array, they cannot be fully detected. The LASSA array can only detect and identify charged particles that stop in either the second silicon detector or the CsI crystal. The particles can not be detected and identified if they stop in other places: such as the Sn-Pb foils that covered the telescopes, the first silicon detector, the dead layers of the silicon detectors, the mylar foil that covers 76 2000 1800 1600 Counts (a.u.) 1400 1200 I I I I I I I I I I I I I I I I I I I 1000 I 800 I I T I I I I 600 I 400 I I I I l I I I 200 lllllllllllllllllllllllllllllllLLl114LilJlll 0 275 300 325 350 375 400 425 450 PID (a.u.) Figure 3.11: PID line of 3He shows the contamination of 4He. It also shows that the PID line of 3He is asymmetric. O N UIfiii 77 é Counts(a.u.) S O Q fl ‘ LIE—— csééggél mammmimm [C l 1 11111 I 1 l l 1000 1500 2000 2500 3000 3500 4000 4500 PID (a.u.) Figure 3.12: Background subtraction for the heavy fragments, the straight dashed lines are the estimated background levels. 78 the crystals and the gap between Csi crystals. Multiple hits also cause losses in the charged particle detection. At polar angles where the detection-efficiency is not zero, the efficiency can be corrected for such effects. In our analysis, efficiency functions were generated for each isotope and for differ- ent reactions. We construct the efficiency functions for central collisions and periph- eral collisions in different ways for different purposes of study. For central collisions, we calculated 1D efficiency as a function of EC", within each degree of 6m. For pe- ripheral collisions, we calculated 2D efficiency as a function of p 1 /A and 3],. The eficiency functions are calculated as follows, isotopes were first generated from the even phase space in the center of mass frame, and then the kinetic energy and an- gle in the lab frame are calculated. If the isotope survives the double hits loss and passes the detection threshold and was successfully detected by a good detector, then a weight of l/N will be added to the corresponding point in either the EC". space or the y, vs. p 1 /A space for this isotope. Where, N is the number of total particles generated. The raw yields are then corrected by these efficiency functions. On the top panel of Fig. 3.13, we show the raw yield of 4He produced in the peripheral collisions of the 112Sn + 112Sn reaction as a function of vs. parallel relative rapidity, called y, vs. the momentum per nucleon perpendicular to the beam direction, called p 1 /A. The rapidity is defined by %ln%[—:' The parallel relative rapidity is calculated by normalizing the rapidity parallel to the beam direction of 4He particles by the beam rapidity. In the figure of the raw yield of 4He, we see that there is an annular ring facing the origin with low counts. This discontinuity of the 4He was caused by either 4He particles stop in the dead layer of the second silicon or in the Mylar foils of C31. This discontinuity was also caused by 4He particles that stop in the CsI, but cause a very low light output due to the non-linear response of the C31, which is so low that we can not detect it. This loss in the efficiencies affects the yields for 79 central collisions at the order of 2%. We also see edges of enhanced yields for certain polar angles 0. On the bottom panel of Fig. 3.13, we show the efficiency function, with double hit effect taken into account, of 4He for the 112811 + 112Sn reaction. We see the same enhanced angles of emission because we take the solid angle coverage into account in the efficiency simulation. On the other hand, we make no attempt to simulate the loss of counts due to the discontinuity where 4He punches into the CsI because this is difficult to simulate accurately enough. In the first panel of Fig. 3.14, we divided the measured yield by the efficiency function to obtain the corrected yields vs. y, and p 1 /A for 4He produced in the peripheral collisions of the 112Sn + 112Sn reaction. We see that the angle dependent of the efficiency is well corrected. In the other 3 panels of Fig. 3.14, we show the corrected yields for all the other reactions. We see from this figure that after correction, the yield is smoother comparing with the raw yield. The “Coulomb ring”, which centered at 3;, ~ 0.95 is more prominent in the corrected yield comparing with the raw yield. In Fig. 3.15, we reflected the corrected yield of 4He with respect to the line of y, = 0.5 and the line of p 1 = 0, since the yields should be symmetric with respect to these lines for symmetric reactions. The corrected y, vs. p1 /A plot provides our best global picture for 4H8 emission in this reaction. 3.3 Data Analysis of the Ring Counter The Ring Counter has similar structure to the LASSA array. The data analysis for the Ring Counter is basically the same as the LASSA array. One needs to do a similar energy calibration and generate a Si-CsI PID function as what are done for the LASSA array. We will discuss about the energy calibration and PID function of the Ring Counter in detail in the following sections. Since the Silicon detectors and the CsI crystals are not tightly mounted in a box as was LASSA, there could be 80 300_ ERaineld 4 250? A :0.8<fi<1 He 0 L S 200: 0 : @1507 $ : £21100: 50;— 300m 250 éEfficiency Function A :0.8<6<1 Q L 3 200: 0 : @150.— s 5 £100: 50;— 0—111111.1...1...1...11111...1. 0 0.2 0.4 0.6 0.8 1 1.2 1.4 yr Figure 3.13: The raw yield and the efficiency fimction for 4He produced in the pe- ripheral collision for the 112Sn + 112Sn reaction 81 300 _ ECorrected Yield 4 ECorrected Yield 4 250 50.8 < b < 1 He 50.8 < 0 < 1 He ;1 12Sn+l 12Sn " ”2Sn+12.4Sn I N O O p i/A(MeV/c) 5 5 IIIIIIIII'IIIIIIIIII 50 - O _ J l 1 1 l l l l I l I l l l l l l l l 1 I l l l 1 l l l ECorrected Yield 4 ECorrected Yield 4 A 250 50.8<6<1 He :_0.8<fi<1 He g 200 Er124Sn +1:12Sn 21248 n +1248n O : : 2 150 r 7 z I I 100 I- '- 21 . e 50 I— I- 0 : l I I 41—]; J L 1 l l l l l l : l l l J l 1 l l 1 l 1 l l l 0 0.5 l 0 0.5 1 yr yr Figure 3.14: The corrected yields of 4He produced in the peripheral collisions for different reactions. l 200 : A 4 _08 0000 4» O o 000 0.0 0 0.0 O 4r 00 6 :- O‘O\O T'- 00 d 6 A ‘O 1’ 0‘0, l v.0 P 0‘ 4 . O 0 1' 4 E 4 f Ob 4? \O‘O\o 4 W 1- \ '0' '1 4 b 0‘0 4b b\o 4 \ \ 4 _Q' ; 112 112 0.0 :1 112S 1248 50 ‘ 1 Sn+ Sn 1., . n o . r- O 41- 0‘ -4 . b 45 0‘ 4 b b L 0‘ 4 0'0 1» DU '0 b . . . . L J 11111 1 ..... 1 0900 1 ..... 1 ..... 1 ..... 1 OOQoom ''''' I V V V V V I r f '7‘, V V V V V I V V V V V I f r I 1 ....oooooooooo.... E .gooooooooooo... 3 . 0 f. .~ 4 P ‘. ... 4p . 5" 4 0 01 " 0 01 ‘ O U . :F .\. 4F ‘3 14 . D \ : z 1 . 3. . . \ ,. \ . 1E'3 E 124 112 '\ 1:? 124 °\. 1, 1E'3 b n 1’ 4 r E 1.1:. \115-4 1: . b4) 4 11 \1 44 e: :A 1: :¢ :: :‘ ‘cie44%47:‘ ‘: ‘ L‘ 4 #;,;44;A }44;¢ -1 #. .:*::—' I) 4 00 > o _ 000 I 000 I o 00 > O 4} 4 o 0 6 I- 0‘0 di- 00 -1 6 , ‘ ‘O 4 A r 0.0.0 1: 0.0. ‘ ”O 4» 00 4 O\ 4» ”O 4 (4.4 Al" 0‘0 1’ b‘ '41 ‘0 4» 0‘ 4 ‘-' b 0 Ob 1 _Q 1248 112S 5 1 124s 1248 ‘5‘ 1 4» o 4 2 n+ n ,1 , n+ n ,2 0‘ 4 O ‘ 4 4 F p 1 ‘0 . 1- O. 4 L K ‘O O '0 O 1 .OQOQnmn: 1 9000913 0”71‘0”2‘0”‘30 0102050 40 Nc Nc Figure 4.1: The probability distribution of the charged particle multiplicity, NC, and the derived impact parameter as a function of the charged particle multiplicity. 93 Zplf 0 llllllllLlilIllllllllllllllliJllll 0 5 10 15 20 25 30 35 Figure 4.2: The charge of project-like fragments, ZW, as a function of the charge multiplicity Nc 94 Other observables can also be used to construct impact parameter filter. Since there is a strong correlation between charged particle multiplicity and impact param- eter. We can test the correlations between other observables and charged particle multiplicity to check their correlations with impact parameter. We explored the cor- relation between the charge of PLF(Project-Like Fragments), detected in the Ring Counter, and charged particle multiplicity for example. In Fig. 4.2, we see that there is a strong correlation between the Zpgf, the charge of PLF detected in the Ring Counter, and NC. In the region of Zplf larger than 20, Z1,” decreases monotonously with NC. This can be understood since the more charge left in the PLF, the less charge left for the overlap regions of the reactions thus less charge particles were detected by Miniball/Miniwall and LASSA. In the region of Zplf smaller than 20, Zplf smears out, suggesting that the correlation does not exist anymore. We do not always detect the PLF because it sometimes goes to smaller angles outside of our coverage. Thus, we chose the reduced impact parameter constructed from Nc as the primary impact parameter filter for this study and not the PLF charges. A two-dimensional cut on two different impact parameter gates has not been used to place more restrictive gates on the event selection, even though that might lead to greater precision in the impact parameter selection for some studies. 4.2 Isotope Multiplicities, Energy Spectra from Cen- tral Collisions and Model Calculation 4.2.1 Experimental Data The differential multiplicities of different isotopes were obtained by first selecting the central collision events with (3 < 0.2, as explained in the section 4.1, then within each event selecting the fragments emitted at center of mass angles of 70° 3 0cm 3 110°. 95 The acceptance of LASSA and the impact of this acceptance cut on the data are illustrated in Fig. 4.3. In this figure, we show the corrected yield in the center of mass frame as a function of ECm and polar angle 06", for 4He produced in the central collisions of the 112Sn + 112Sn reaction. The cut 70° 3 0cm 3 110° is indicated by the dashed lines in the figure. At these angles, the coverage of the LASSA array is excellent and the efficiency is well behaved. The only losses occurred for particles emitted at very low energies E /A < 0.2 MeV in the center of mass, corresponding to small laboratory angles of 61.1,, < 7°, as well as at high energies, corresponding to laboratory angles of 010;, > 58°. These losses (total < 2%) were estimated by fitting the energy spectra for each isotope and corrected by the efficiency function. The acceptance of LASSA is also shown in Fig. 4.4. In this figure, the efficiency corrected differential multiplicity dM/dydpi for 12C fragments is shown. The angle cut is indicated by the dashed lines in the figure as well. The measured differential multiplicities smoothly depend on the transverse momentum pi and rapidity y and are centered about the rapidity of the center of mass, consistent with emission from the participant source. They display no characteristic “Coulomb holes” near the projectile rapidity, expected from the Coulomb repulsion of 12C fragments from projectile-like residues moving with velocities somewhat less than that of the projectile. The final differential multiplicities were obtained after background subtraction and efficiency correction. Background level was approximated by fitting each element on the 1D PID histogram with different energy bins cut, then subtract the background corresponding to each energy bin. Totally ten bins were chosen according to the statistics of the yields of the fragments. The efficiency correction was done for each degree from 70° to 110° in the center of mass frame. Fig. 4.3 shows the 2D Emu-0cm histogram for 4He after the background subtraction and efficiency correction. The 2D Emu-06m histograms are then projected to 1D to generate the energy spectra and the 96 o 70° @110° 4 S a I I 00 § 1 I l l l _ . m-__‘-__ _--lfl- - 0 20 40 60 80 100 120 140 160 180 (*9 cm Figure 4.3: Corrected yield of ‘He produced in the central collisions for the 112Sn + 112Sn reaction 97 O O mSn+mSn E B< 0.2 12C p J_/A(MeV/c) '5 171 N 95’ U] Q .,... -150} goo—0126' '012’014'016'618' '1' ”112 yr Figure 4.4: Efficiency corrected differential multiplicity dM/dydpi for 12C fragments. The cut 70° 3 0cm 5 110° is indicated by the dashed lines in the figure. I N p-a O dM/(dEdQ) 3 ES 1. 1 {.11 10 1'1 111111111 W 11_11 . 1 11111111 71111 11111111111111111 I'l;'li: i; ‘31‘ 11 11. 11. o Proton 1 Deutron ’ Triton f 1 O 25 50 75 100 (3111 O ‘5‘. AAAAAAAA r. ..... 1‘1zsn+112Sn g ;;.1»<0.2 I AA“ .‘A‘ 2‘1: 1566‘ ::.!!!!6 l<_ 1~; E— 1 - . °‘ ‘113'. , : 11111" via? _ 4‘} .. 1‘ 1‘ ‘0 ,__ . § .4 1 '6L1 11 ' : ‘7 - I? I l _ 8L1 1, ‘ 1 1? » L1 ‘ . E 9 . : A L1 1 1 1 1 l 1 l 1 1 1 1 1 l 1 1 1 l 0 50 100 150 Ecm(MeV) 10 -5 w“. 10 10_ 11:21} 10 ,1. 5111111 1 1 1 1 1 ‘1 1 1 l 050 100 150 Ecm(MeV) A - C.‘ 4 “ a 10 E' a 1!. llen+mSn .0 11' =11, S Z for 124Sn+mSn and located at N = Z for 112Sn+”28n. In general, the drop from the peak toward more proton-rich isotopes is rather steep especially for elements with even values for Z. The main differences between the isotope yields for the four different systems are observed in the tails of the iso- tope distributions, where it is greater than a factor of 4 for 200. Larger differences may be expected for even more exotic isotopes, but the background in the present 101 10_1:L,f...I....1,24...12.,-...,...., ~ —° Sn+ Sn; E/A=5OMeV - ----- 0 mixed ___,. 1128n+1128n Figure 4.7: Average differential multiplicities at 70°§ 60M 5 110° for Li, Be, B, C, N and O isotopes as a function of neutron excess (N — Z) of the isotope. The solid points drawn by solid lines to guide the eyes are data for the 124Sn+1248n system with N /Z =l.48. The open circles are data for the lightest system 124Sn+12‘13n with N / Z =1.24. The open diamonds are the averaged values from the two mixed systems, 124Sn+1128n and 112Sn+mSn. The dashed and dot-dashed lines are predictions from Eq. 4.2. See text for more details. 102 measurement due to multiple hits in the LASSA telescopes does not allow for their accurate determination. For multi-fragmentation which can be described by statistical fragment emission mechanisms, it was found that ratios of isotopic yields R21(N , Z )=Y2(N , Z ) / Y1(N , Z) for a specific pair of reactions with different total isotopic composition, follow an isoscaling relationship [123, 110, 112] Y N,Z a R21(N, z) = 7372—; = 0d ”+le . (4.2) Here, C is an overall normalization factor, a and ,8 are the isoscaling parameters for the chosen pair of reactions. This parameterization is discussed in greater detail within the isosaling section below. If we adopt the convention that reaction 2 is more neutron-rich than reaction 1, one expects a to be positive and ,8 to be negative. We have adOpted that convention here and have fitted the ratios of the isotopic yields for these four systems to extract the corresponding values for a and B. These values for a and [3 are given in Table 4.1. Reaction 2 Reaction 1 a fi 112Sn+124sn 112Sn+1128n 0.18:t0.01 -0.193L_0.01 124Sn+124Sn 112Sn+u2Sn 0.36i0.02 -0.39:t0.01 Table 4.1: Values for a and 3 obtained from fitting the isotope ratios R21. Eq. 4.2, with only three parameters, C, a and [3 can be used to predict the isotope yields of 112Sn + 1”Sn as well as the mixed system, 112Sn + 124Sn or 124Sn + 112Sn systems using the measured yields of one system. To illustrate how well this parameterization relates the yields of these four systems, we take the yields of the 124Sn + 124Sn system as a reference and use those yields and the fitted values of a 103 and B to predict the yields for the other four systems. The dash and dot-dashed lines in Fig. 4.7 are the calculated yields for 1123n + 112Sn and 112Sn + 124Sn, respectively. These isoscaling parameters can, for this limited range of asymmetry, be described by a linear dependence on either the initial N /Z or the asymmetry parameter, 6 = fig of the reactions [112]. The excellent agreement between the predicted yields and the data suggests that such scaling law extrapolations may have useful predictive power. For example, we expect that these scaling predictions can be accurately extrapolated to other mass-symmetric systems of A=200-250 nucleons at the same incident energy per nucleon but with very different isospin asymmetry. 4.2.2 Models for the Multiplicity Calculation Now we turn to the theoretical interpretation of the central collision data. To study the density dependence of the asymmetry term of the EOS, we adopt the viewpoint of the Stochastic Mean Field (SMF) approach described in refs. [2, 23]. In this approach, the time evolution of the nuclear density is calculated by taking into account both the average phase-space trajectory predicted by the Boltzmann-Northeim-Vlasov equa- tion and the fluctuations of the individual collision trajectories about this average that can be predicted by equations of the Boltzmann-Langevin type. Both the intro- duction of fluctuating forces and the inclusion of numerical noise have alternatively been utilized in modeling the fluctuations in practical solutions to such equations; the simulations presented here utilize the latter technique. The virtue of such a dynamical approach for. the study of isotopic effects lies in its self-consistency. The flow of neutrons and protons is calculated under the influence of Coulomb and asymmetry terms, which reflect self-consistently the motion of these nucleons. Several different density dependences of the asymmetry term were explored. Here, we present results for two different asymmetry terms. In both cases, 104 the asymmetry term is approximated by the form Esym (p, 5) = S (p) 62 , (43) where for the asymmetry term with a stronger density dependence :a _P_ 2/3 200/Po)2 5“” (p0) ”Hm/p0) ' (4'4) Here, p is the physical and p0 is the saturation density, a=13.4 MeV, b=19 MeV [58, 64, 85]. In the following, we refer to this as the “super stiff” asymmetry term. For the asymmetry term with the weaker density dependence, S(p) = (4%)?” + 240.9p — 819.1,)2 , (4.5) where a=12.7 MeV [23]. In the following, we refer to this as the “soft” asymmetry term. In Fig. 4.8, it can be seen that the two expressions are nearly equal at satura- tion density but differ at densities that are either much larger or smaller that p0. In addition to the asymmetry term, the nuclear mean field has a Skyrme type isoscaler mean field with a soft equation of state for symmetric matter characterized by an incompressibility constant K 2201 MeV. The isoscaler mean field and the asymmetry term of these equations of state are used for the construction of the initial ground state and for the time evolution of the collision. The nucleon-nucleon collisions by the residual interaction are calculated from an energy and angle dependent parame- terization of the free nucleon-nucleon interactions and the isospin dependence of the Pauli-blocking is considered during these collisions. The calculation solves the transport equations by evolving test particles of finite width. As mentioned above, we use a reduced number of test particles (50 test particles per nucleon) in the present calculations to inject numerical noise into the evolution. In test calculations, we alternatively employed the fluctuation mechanism discussed in ref. [23], which involves damping the numerical noise by utilizing a large 105 A O _] l—‘ [\3 [\D W ()3 U] :3 U1 O U1 I I I I I 3(9) (MeV) —super stiff _ ' " soft )—l O I CU! 0 0:2 0:4 0:6 03; 1 1.2 P/PO Figure 4.8: The solid curve and dashed curves indicate the density dependencies for the super stifl and soft asymmetry terms, respectively. 106 number of test particles and introducing explicitly physical noise according to thermal fluctuations. It was checked that both methods lead to similar results. In contrast to the Brownian one-body (BOB) method of ref. [44], these methods of inserting fluctuations are well suited to reactions at finite impact parameter because they do not presuppose knowledge of the most unstable modes. When the system expands and reaches the spinodal instability (after about 1 10- 120 fm/c), the most unstable modes are amplified and initiate the formation of fragments via spinodal decomposition. The evolution of the system is continued after spinodal decomposition until freeze-out where the number of dynamically produced fragments and their properties are finally determined. The system is decomposed into fragments using essentially a coalescence mechanism in coordinate space; specifically, fragments are defined by regions of density in the final distributions that are above a “cut- off” density of 1/8po. By definition, the freeze-out time occurs when the average calculated number of fragments saturates. This occurs about 260 fm/c after initial contact of projectile and target nuclei in the present simulations. The excitation energy of the fragments is determined by calculating the thermal excitation energy in a local density approximation. The procedure is rather rough and will overestimate the excitation energy particularly for light fragments. Some of the important features of these calculations and of the prior BUU-SMM [101] and EES [1 10] calculations can be understood simply by considering the influence of the density dependence of the asymmetry term on the relative emission rates of neutrons and protons. In all Sn+Sn collisions, the symmetry energy in the liquid drop model is positive, i.e. repulsive. The interaction contribution to the symmetry energy gives rise to a repulsive contribution to the mean field potential for neutrons and an attractive contribution to the mean field potential for protons. The mean field potential for an asymmetry term with stronger density dependence is larger at 107 high density and weaker at low density than that for an asymmetry term with weaker density dependence. It is the low-density behavior that dominates the predictions for the isoscaling parameter. As the system expands and eventually multi-fragments, the prefragment remains at subnuclear densities for a long time while it is emitting nucleons. The asymmetry term with weaker density dependence around p0 increases the difference between the neutron and proton emission rates leading to a more symmetric prefrag- ment than is produced by calculations with the asymmetry term which has a stronger density dependence. The SMF calculations are interesting because they are free, in principle, of arbi- trary assumptions about whether the fragments are formed at the surfaces or from the bulk disintegration of the system. Comparisons between the fragmentation dynamics for different asymmetry terms were reported in ref. [107]. The trends of these calcu- lations are consistent with the prefragment isospin dependences discussed above. In particular, fragments produced in calculations with an asymmetry term with strong density dependence tend to be more neutron-rich than the fragments produced in calculations with an asymmetry term with weak density dependence. In this respect, these predictions are similar to the results of the BUU-SMM calculations of ref. [101] and opposite to the results of the EES calculations of ref. [110]. However, the SMF fully dynamical formation of fragments should actually be more sensitive than the hybrid BUU—SMM to the interplay of the EOS, i.e. to the density dependence of the asymmetry term, with the fragmentation process. In the hybrid BUU-SMM calculations, the EOS is entering only in the “pre-equilibrium” nu- cleon emission described above. In the SMF approach, we have not only this isospin eflect on fast particle emission but also the full dynamics of the isospin fractiona— tion/distillation mechanism during the cluster formation. In a neutron-rich system, 108 this leads to a different N / Z “concentration” in the liquid phase (the fragments are more symmetric) and in the gas phase (nucleons and light ions, are more neutron- rich) [75, 3]. This effect is associated to the unstable behavior of dilute asymmetric nuclear matter and so in this way we have the chance of testing the EOS also at very low density. Asymmetry terms with weaker density dependence around p0 must show a faster increase at low densities and so a larger isospin fractionation / distillation during the fragment formation [4]. Therefore in a fully dynamical picture of fragmentation events a “sof ” behavior of the asymmetry term around saturation density will enhance the formation of more symmetric fragments for two converging reasons: i) A larger “preequilibrium” neutron emission rate as discussed before; ii) A stronger isospin fractionation / distillation during the bulk disintegration. Opposite effects are of course predicted for a rapidly increasing (“stiff”) asymmetry term around p0. In this sense we can expect the SMF results to be more sensitive to the isospin dependences of the EOS at sub-saturation density. At freezeout, the fragments are highly excited. For simplicity, we assume that the de—excitation of these fragments can be calculated as if the fragments are isolated. For this de-excitation stage, we have tabulated the known masses, states, spins, isospins and branching ratios for nuclei with Z 315. Where experimental information is complete, it is used. Alternatively, empirical level density expressions are used for the discrete levels. These discrete levels are matched to continuum level density expressions as described in ref. [103]. The decay of primary fragments with Z 215 are calculated, following ref. [103], using known branching ratios, when available, and using the Hauser-Feshbach formalism when the information is lacking. The decays of heavier nuclei are calculated using the Gemini statistical decay code [103, 20]. While the SMF calculations predict the numbers and properties of the hot frag- 109 ments that are produced at breakup, the predictions for the relative abundances of light clusters such as the isotopes with Z =1-2 that are emitted before the system expands to sub-nuclear density are not very realistic. This prevents a precise model- ing of complete events including their detection efficiency and means that the impact parameter selection based on multiplicity cannot be imposed straightforwardly on the calculated events as on the data. This and the considerable numerical effort it requires have persuaded us to limit our comparisons to calculations composed of 600 events for each of the 112Sn+“28n and 124Sn-l-mSn reactions at a fixed impact pa- rameter of b=2fm. We note, however, that the widths in the multiplicity distributions at fixed impact parameter are large enough that a range of impact parameters may contribute significantly to the experimental data. Future calculations are necessary to assess quantitatively the importance of this impact parameter smearing. 4.2.3 Overall Behavior Predicted by the SMF Calculations In Fig. 4.9, the solid circles and open squares in the left panel show the measured ele- mental multiplicities for 23 Z S8 averaged over 70°§ 9cm $110° for 124Sn+mSn and 112Sn +1123n collisions, respectively. The right panels show the corresponding mea- sured multiplicities as a function of the fragment mass. These averaged multiplicities were obtained by summing the isotopic multiplicities for 2_<_ Z 38. The dashed lines denote the corresponding distributions of hot primary fragments calculated by the SMF model using the super stiff EOS. Due to the low total number of events, we averaged these calculations over a slightly larger angular interval of 60° 5 0cm 3 120°. The solid lines show the multiplicities of cold fragments after secondary decay. The statistical uncertainties in these calculations are shown in the figure as vertical bars. The corresponding uncertainties in the data are smaller than the data points. If the angular integration was performed over the entire solid 110 t V I V I f l I v V I V V l V V 124 124 Sn+ Sn E! . Data _ . §§ O ---- Primary , E I . —— Secondary: 1r I -b - II I- d. . p {b F i. ‘F ii I .1:- nt- -P p dM/do (sr'l) (3’): .O p—I r1 0.01 Figure 4.9: Differential multiplicities for 124Sn+”"Sn collisions (upper panel) and 112Sn+“28n collisions (lower panel) as a function of the fragment charge (left panels) and the fragment mass (right panels). The points are the data. The dashed and solid lines are the calculated primary and final fragment differential multiplicities, re- spectively. Statistical uncertainties are shown for the calculations; the corresponding uncertainties in the data are smaller than the data points. 111 angle, the averaged calculated multiplicities are about 20-30% larger. This difference reflects an anisotropy in the calculated primary angular distributions for the heavier fragments. In the present calculations, however, we do not have the capability to accurately calculate the modifications of the angular distribution due to secondary decay so we presently cannot explore this issue more quantitatively. As we will show later, this anisotropy has no impact on the shapes of the isotopic distributions for Z =3-8. In general, the calculated primary and secondary fragment multiplicities are smaller than the measured values for the lighter fragments Z =3,4 and are somewhat closer to the measured values for Z =6—8. The lighter fragments with Z < 4 are mainly produced in secondary decay stage of the theoretical calculations; the primary yields of these light fragments are much smaller relative to the final yields than are the values for the heavier fragments. Because the fragment multiplicities and angular distributions depend on impact parameter, the comparison shown in Fig. 4.9 may be sensitive to the impact parameter ranges included in both calculation and data. Future calculations over a wide range of impact parameters are needed to address this issue. Concerning the greater discrepancy for Z =2-4 fragments, we have already noted that the formation of light clusters in the dynamical stage before breakup is not well described in BUU- and SMF-type simulations, because the unique structural properties of these fragments are not therein well treated. (Treatments of the emission of light clusters in coupled transport equations for nucleons and light clusters can be found in refs. [109, 29] and in the framework of FMD [35] or AMD [78] simulations.) On the other hand, there is a considerable emission of protons and neutrons during this stage; the total emission and consequently the asymmetry of the remaining source may still be realistic. Now we turn to an examination of calculated isotopic yields. The upper left panel 112 of Fig. 4.10 shows the isotopes of carbon nuclei predicted by the SMF calculations over the entire angular range for 124Sn+12‘18n (solid line) and 112Sn+m’Sn (dashed line); the dotted-dashed and dotted lines show the corresponding calculations over the 60° 3 0cm 3 120° gate. Not surprisingly, the more neutron rich 124Sn-l—liMSn system preferentially produces the more neutron rich isotopes. The peak of the carbon primary distribution for the 124Sn+12‘13n system occurs at about 15C while the peak for 112Sn+1128n system occurs at lower mass, i.e. somewhere between 13C and 14C. The differences between the angle gated and total primary yields are small, and these difference translate into negligible differences in the shape of the isotopic yield distribution after secondary decay; we therefore do not plot the gated data because the two curves are indistinguishable when normalized to each other. As the statistics of the present calculation make it difficult to perform comparisons to isotopic yields with a 60° 5 6a,, 5 120° gate imposed on the calculation, the remaining calculated multiplicities are integrated over the entire solid angle. After sequential decays, one obtains the secondary distributions shown in the lower left panel. No longer is there a noticeable difference between the peak locations (at 12C in both systems); instead, the main differences are found in the shape of the distribution, which is higher in the neutron-rich isotopes and lower in the neutron- deficient isotopes for the 124Sn+”“Sn system than it is for the 112Sn+“2Sn system. Such trends are also qualitatively observed in the experimental data shown for the 124Sn+1248n system by the solid circles and for the 11"’Sn+1128n system by the open squares in the lower left panel. However, the experimental distributions are consider- ably wider and more neutron rich than the model predictions. This trend is replicated in the elemental distributions for all of the other measured elements. Another way to quantify the differences in the isotope distributions is by the asymmetry parameter 6 = (N — Z) / (N + Z). The average asymmetry < 6 > of the 113 . . v v g . r - I I I I F I I 0.30 Carbon primary -- mSn + “2811 all e 1 " 124 124 if d 0'25 — Sn + Sn all 6 --- "23n+"23n 60° .. 0.15 ] 0001- «~ ~O.10 1E-4 - D “2311+“st MIX ~~ « 0.05 OmSn+mSn data I ’ ‘, lE-Slo“"‘1‘5"‘345mé° A Z Figure 4.10: Left panels: Calculated primary (upper panel), calculated final (lower panel) and measured (lower panel) carbon isotopic yields for Sn+Sn collisions. Right panels: Calculated primary (upper panel), calculated final (lower panel) and measured (lower panel) mean isospin asymmetries as a function of the fragment charge for Sn+Sn collisions. The lines and data points are further explained in the text. 114 isotopic distribution for each element is shown as a function of Z in the right panels of Fig. 4.10. Following the same convention as in the left panels, the solid and dashed lines show the average asymmetries for 124Sn+1248n and 112Sn+”28n collisions; the upper and lower panels present results for the primary and secondary fragment dis- tributions, respectively. The calculated differences between the two systems are more pronounced prior to secondary decay than afterwards. The corresponding data, shown for the 124Sn+124Sn system by the solid circles and for the 112Sn+uf"Sn system by the open squares in the lower right panel, are larger and display a stronger dependence on the asymmetry of the system than do the final calculated fragment yields after secondary decay. 4.2.4 Isoscaling Analyses A more sensitive way to compare isotopic distributions is to construct the isotopic ratio R21(N, Z)=Y2(N, Z)/Y1(N, Z) from the isotope yields Y,-(N, Z) with neutron number N and proton number Z from two different reactions. As discussed in the experimental section, R21(N, Z) obeys a simple relationship R21(N, Z) = C e°N+BZ where C is an overall normalization factor and a and ,8 are isoscaling parameters [101, 110, 123, 112] . Such an isoscaling relationship can be obtained in statistical theories for two systems that are at the same temperature when they produce fragments. Binding energy factors common to the yields for the fragments in each system are cancelled by the ratio when the temperatures are equal, leaving terms related to the chemical potentials or separation energies [110]. In grand canonical models of multifragmentation, a = App/T and fl = Ayn/T for example, where Ann and Au... are the differences in the chemical potentials for the neutrons and protons in the two systems and T is the temperature [123, 112]. In some calculations [112], the values for the isoscaling parameters extracted from equilibrium multifragmentation models are 115 “‘er «_— similar before and after sequential decays, an observation that has been attributed to a partial cancellation of secondary decay effects [123, 112]. While isoscaling can be expected for many statistical processes [101, 110, 123, 112] , the question of whether it can be expected for specific dynamical calculations remains open. To investigate whether the SMF dynamical model displays isoscaling, we construct the relative isotope ratios, R21, primary, using the primary fragments produced in 124Sn+mSn collisions as reaction 2 (numerator) and in ll""Sn+11r"’Sn collisions as reaction 1 (denominator). The results are shown in the upper panel of Fig. 4.11. The error bars reflect the statistical uncertainties. The predicted isotope ratios for these primary fragments depend very strongly on the neutron number and follow trends that appear consistent with isoscaling relationship defined by Eq. 4.2. The uncertainties are large reflecting the low statistics of the simulations, but the strong dependence on neutron number makes it possible to discern apparent isoscaling trends nonetheless. The lines are best fits using Eq. 4.2 resulting in C = 0.96, a = 1.07 and B = —1.43. These values for a are much larger than values observed in the experiment. The lower panel provides the corresponding SMF predictions for the ratios, R21, final, of the yields of particle stable nuclei after secondary decay. For comparison purposes, the ordinates of the top and bottom panels are chosen to be the same; this demonstrates graphically that the trends of the final isotope ratios are much flatter and the corresponding isoscaling parameters (a = 0.286 and B = —0.288) are much smaller. Clearly, the isoscaling parameters predicted by dynamic SMF calculations are strongly modified by secondary data. This trend is very diflerent from some equilibrium statistical models for multifragmentation where the isoscaling parameters have been predicted to be insensitive to secondary decay [123, 112]. The isoscaling behavior of the dynamically produced fragments arises not from 116 20.0 : e1.0'71N . 10.0 [— BNV primary Z) .oz-.m .01 01000 T | F3 (\3 T Y1“Sn+‘2‘Sn(N:Z)/ Y‘128n+“23n(N .93 R21: [_x 01 o o K' .0 Figure 4.11: Upper panel: R21 values obtained from the ratios of the primary isotopic distributions for 124Sn+1248n collisions divided by those for 112Sn+112$n collisions. Lower panel: Corresponding R21 values obtained from the ratios of the final isotopic distributions. Each line in the two panels corresponds to ratios for a given element. Elements with Z =2—8 (Z =2 1-8) are represented from left to right in the upper (lower) panel. The lines are the result of fitting R21 with Eq. 4.2; the dependencies on neutron number for the best fits are given in each panel. 117 thermal physics but rather from some special characteristics of the SMF primary distributions predicted for these reactions. We find, for example, the SMF primary isotopic and isotonic distributions can be roughly described by Gaussians, see Fig. 4.9. Isotopic distributions, for example, can be described by (N — N(Z))2 Y(N,Z)=f(Z)€Xp _ 20% , (4-6) where N(Z) is the centroid of the distribution and 0% describes the width of the distribution for each for each element of charge Z. This leads to an exponential behavior of the ratio R21, since, neglecting quadratic terms in N, In R21 = i, [37(2), — N(Z),] N . (4.7) 02 Note Eq. (4.7) requires the values for 0% to be approximately the same for both reactions. We have observed this to be the case for our SMF calculations of Sn+Sn collisions (to within the statistical accuracy ~10%). For the ratios for every element, to be optimally described by the same parameter, the ratio [N (Z )2 — N(Z)1] /az must be independent of Z. The statistics of the calculation do not allow a detailed test of this assumption, but it does appear that this ratio increases somewhat with Z, as Fig. 4.11 suggests. The primary distributions therefore do not respect the isoscaling relationship as well the data do. Now, at variance with the statistical fragmentation models, the secondary decays substantially modify the isoscaling parameter. The width 0% decreases due to sec- ondary decay and the difl'erence [N2 — N1] like ways decreases fractionally, but by a larger amount. Moreover, the final shape is no longer Gaussian, but due to secondary decay, it reflects the binding energy as a function of neutron excess more strongly (see Fig. 4.10). These changes combine to decrease the isoscaling parameter as shown in the lower panel of Fig. 4.11. 118 4.2.5 Sensitivity of the SMF Calculations to the Asymmetry Term The density dependence of the asymmetry term has a significant influence on the relative emission rates of the neutrons and protons and, consequently, on the isospin asymmetry of the hot fragments prior to secondary decay. As discussed previously, an asymmetry term with weaker density dependence tends to remain more important at lower densities, driving the fragments closer to isospin symmetry, than does an asymmetry term with stronger density dependence. Consistent with this general consideration, the calculated primary isotope distributions in 124Sn+124Sn collisions, shown in Fig. 4.12 for carbon (upper left panel) and oxygen (upper right panel), are more neutron-rich for the super stiff asymmetry term (solid line) than they are for the soft asymmetry term (dashed line). A similar trend is also predicted for the 112Sn+m’Sn system, but is not shown in the interest of brevity. A similar trend is observed in the corresponding final distributions that are ob- tained after secondary decay and shown in the middle panel with the same convention for the solid and dashed lines as in the upper panel. Both secondary distributions cal- culated for super stiff and soft asymmetry terms, however, are significantly narrower and more proton-rich than the experimental distributions shown by the closed circles in the figure. (The lower panels, which display corresponding calculations when the excitation energy is reduced by 50%, will be discussed in the next section.) Similar trends are also observed for the 112Sn+1128n and for the other elements with 3$Z£8, though we do not for brevity’s sake show those results. In Fig. 4.13, we present the related dependence of the SMF predictions for the isotope ratios R21 upon the density dependence of the asymmetry term. We take advantage of the fact that the results in the Fig. 4.11 can be compactly displayed by the scaled function S (N) = R21 (N, Z) -e"Bz which condenses the isotopic dependence 119 0.1- "'124"2'4""'10'1 1 Carbon S“ TL S“ Oxygen 0.01» anary ‘0.01 1E-3b «IE-3 .7" 0.01[ -0.01 H U) V % 1E-3» -1E-3 % 11341 11134 C data % + : jgfic : }* : i r 4 secondaryE/Z 0.01- 1- , «0.01 133. O, . 1 -lE-3 l: .; 1E-4- ' ‘lE-4 lE—S-A-A-eAA-AA 14‘113-5 10 15 15 20 Figure 4.12: Upper panel: Dependence of the primary distributions for carbon(left panel) and oxygen (right panel) upon the density dependence of the asymmetry term. Middle panel: Dependence of the final distributions for carbon (left panel) and oxygen (right panel) upon the density dependence of the asymmetry term. The data are also shown as the solid points. The various lines in the figure are described in the text. The excitation energies for the fragments are taken directly from the SMF calculations. Lower panel: The data are the same as in the middle panels. The curves are the calculations obtained when the excitation energies of the primary fragments are reduced by a factor of two. 120 for the various elements onto a single line [110]. The left panel in Fig. 4.13 shows the results for the super stiff asymmetry term and the right panel shows the results for the soft asymmetry term. In each panel, the values for S (N) obtained from the primary distribution are shown by the symbols clustered about the solid lines, the results obtained from the secondary distribution are shown by the symbols clustered about the dashed lines and the results from the data are shown by the dotted-dashed lines in each panel to provide a reference. Both the primary and secondary values for S (N) have been fit by exponential functions to obtain corresponding values for the scaling parameter and these values are given in the figure. Generally, the primary distributions for both equations of state display a much stronger dependence on neutron number than do the final isotopic distributions and the data. However, the influence on the isoscaling parameter is statistically not very significant. Indeed, as we pass from a “stiff” asymmetry term to a “soft” one, we do have a stronger isospin fractionation / distillation, as already discussed before. The centroid of the distribution, N2, decreases (see Fig. 4.12) but also the width 0% de- creases; the decrease in width, however, is of the order of 10% and comparable to its statistical uncertainty. The calculated final distributions display a weak sensitivity to the density dependence of the asymmetry term; the values for (a=0.286) obtained for the super stiff asymmetry term are larger than the values for (a=0.254) obtained for the soft asymmetry term. The sensitivity to the asymmetry term is considerably less than that reported for the EES model [110], and for the BUU-SMM hybrid calcula- tions [101]. Unlike these latter two calculations, both super stiff and soft asymmetry terms yield alpha values that are significantly lower than the value extrapolated from the data (020.36). One should note, however, that the excitation energies of these latter calculations could be more freely varied to achieve better agreement with the experimental observations. 121 5,...-,....,g,....,...fi,. 10 E" ‘ ‘ Primary ;, “a? '3 E—final 1E 1 . 1. 52 data A3 ,3 ,. $ / . A104 t" N0 ‘ 1'.- ng ’. 1 N 5 ° ’ 25 ' ’ a 3 5 super stlff (0g 1; sofi e93 : a. . ’_ 1' :4 i I, a) 3 _._ _ ,3 10 g a? l 2 z 10 ‘2 U) 1 , 101 _. 100 Figure 4.13: Dependence of the scaled function S(N) on the density dependence of the asymmetry term. The left panel provides a comparison between values for S(N) computed from the data (dashed line) and the calculated primary (points about solid line) and final (points about dashed-dotted line) distributions obtained for the super stiff asymmetry term. The right panel provides a comparison between values for S(N) computed from the data and the calculated primary and final distributions obtained for the soft asymmetry term. 122 4.2.6 Discussion of the Isotope Multiplicity and SMF Calcu- lations The calculated final isotopic distributions for both asymmetry terms differ from the measured ones in three respects: (1) they are narrower; (2) they are more neutron deficient; and (3) they show a weaker dependence on the isotopic asymmetry of the total system. In these respects, the calculated results for the two different asymmetry terms are more similar to each other than they are to the data. It is probably premature at this stage to focus attention on the sensitivity of the predicted final distributions to the asymmetry term. Instead, let us concentrate upon what may be required to bring the final isotopic distributions into greater concordance with the measurements. The tendencies of the final isotopic distributions to be more neutron deficient and to display a weaker dependence on the isotopic asymmetry of the system are probably closely related. When the final isotopic distributions for both systems are too neutron deficient, there is not very much room for a strong dependence on the isotopic asymmetry of the system. The problem that the final isotopic distributions are too neutron deficient can either be that the primary distributions were too neutron deficient, that the secondary decay removed too many neutrons or that the freezeout assumption is invalid. First let us consider whether initial isotopic distributions may be too neutron deficient. We note that the average isospin asymmetry of the initial distributions is trivially related by charge and mass conservation to the average isospin asymmetry of the nucleons emitted during the SMF calculations before the freezeout (t=260 fm/c) chosen for these calculations. Clearly, it is important that complimentary measurements of the yields and energy spectra of light particles be performed to provide guidance as to whether these missing neutrons are carried away primary 123 during fast pre-equilibrium emission or slower evaporative emission. However, the present simulations neglect the emission of light clusters ((1, t, 3He, 4He, 6Li, 7Li, etc.) during the dynamical evolution prior to the freezeout. Previous studies [29, 94, 92] have noted that the neglect of the emission of 4He emission is particularly problematic because it is abundant and because each 4He particle enhances the isospin asymmetry of the remaining system by removing four nucleons without changing the neutron excess. Indeed, it has been speculated that that 4He emission may have an influence on the isospin asymmetry of the other clusters and fragments that is of the order magnitude of the influence of the mean field [29, 94, 92]. The treatment of light cluster emission clearly needs additional theoretical attention. The number of neutrons removed by secondary decay depends primarily on the fragment excitation energies and the relative branching ratios for neutron and charged particle emission. There are significant uncertainties in the calculation of the exci- tation energies of the fragments, which are related to the difficulty to establishing precise their ground state binding energies. To explore the sensitivity of the results to the excitation energy, we have reduced the excitation energy of each fragment by a multiplicative factor f where 0.5 S f S 1 and recalculated the final fragment isotopic distributions. The solid and dashed lines in Fig. 4.12 for carbon fragments (lower left panel) and oxygen fragments (lower right panel) show the calculated final distributions for f =0.5 using super stiff and soft asymmetry terms, respectively. Clearly, it is possible by reducing the excitation energy to shift the isotope distribution in the direction of the more neutron-rich isotopes, so as to make the mean isospin asymmetry of the calculated final and measured distributions to be the same. However, the widths of the calculated final isotopic distributions will still be narrower than the measured ones . 124 ". This discrepancy between the theoretical and experimental widths would be re— duced if the theoretical primary distributions were wider in their excitation energy distributions or their isotopic distributions or both. It is interesting that the primary distributions of equilibrium statistical model calculations that reproduce the experi- mental final distributions are much wider in excitation energy and neutron number than those predicted by the SMF calculations [97, 112]. Future investigations will be needed to address whether wider primary distributions in excitation energy or neu- tron number can be attained in the SMF model by altering some of the underlying model assumptions such as the manner in which they are defined at freezeout. Increased widths may be achieved by performing calculations for a range of im- pact parameters rather than the single impact parameter b=2fm presented here. We note that the inclusion of larger impact parameter events may broaden the primary distributions at mid-rapidity because it will require the inclusion of fragments emitted from the neck joining projectile- and target-like residues. In ref. [4], it was shown that SMF calculations predict such “neck” fragments to be more neutron rich because the isospin fractionation / distillation effect is somewhat reduced in peripheral events, leaving more neutrons in the fragments, and there can also be an overall neutron en- hancement in the neck region for such events due to the neutron skins of the projectile and target [4]. Moreover, the excitation energies of the primary fragments in more peripheral collisions are also somewhat reduced. Experiments also suggest that neck fragments in very-peripheral collisions are more neutron-rich [33], but there is little detailed experimental information about the impact dependence at smaller impact parameters. Combining more calculated peripheral events with the calculated central collision events presented in this thesis may result in the broader isotopic distributions that appear to be required by the data. 125 4.2.7 Energy Spectra and Equilibrium Models Calculation The central events display many attributes consistent with bulk multifragmentation following expansion and spinodal decomposition at densities of p S 1/3po [27, 70, 36]. The assumption of thermal equilibrium allows useful baseline predictions for the behavior of complex systems, even when equilibrium may not be fully achieved. For example, the yields of particles emitted in intermediate energy [27, 12], relativistic [120, 60] and ultra-relativistic [77] nuclear collisions have been successfully compared to equilibrium statistical ensembles. This success suggests the attainment of chemical equilibrium, allowing interpretations [76, 11, 116] of such data in terms of phase transitions in strongly interacting matter. This interpretation assumes that systems must last long enough for the relevant degrees of freedom to equilibrate, and the observables being described reflect that equilibrium. Similar to the case for the early universe, observables related to these degrees of freedom may reflect a freezeout time after which these observable cease their evolution. Indeed, many observables in the central collisions of heavy ions have been successfully described by equilibrium calculations for an expanding system with an initial mass, charge, excitation energy and a freezeout density of p z 1 /6po — 1/3po [27 , 60, 11]. Unlike the early universe, however, the small size of these systems may allow particles to be emitted and decoupled from the surfaces of the system before the rest of the system has its freezeout. In this section, we show that the energy spectra of isotopically resolved fragments allow one to observe such effects and quantify the breakdown of the freezeout approximation. The expected properties for the energy sprectra of fragments from the disassembly of thermalized freezeout configuration are straightforward. Because nuclear interac- tions between particles are assumed to be negligible after freezeout, the observed mean kinetic energies of particles < E. > reflect the thermal kinetic energy 3/ 2T, 126 the collective velocity Ucou at freezeout and the energy < ECO“; > gained due to the accelerations of these particles by the Coulomb field of the remaining system [70,50,48L 1 2Am~ < ”Sou > + < ECaul > 1 (4-8) 3 where Am N is the mass of the fragment with A nucleons and m N is the nucleon mass. As the Coulomb energy depends nearly linearly on the fragment charge and for light fragments, Z / 2, Eq. 4.8 suggests a roughly monotonic dependence of < E, > upon A; this has been used previously to extract values for the collective expansion velocity ”can after constraining < ECO“, > with assumptions about the breakup density [70, 50, 48]. Because ECO,“ depends on Z while the collective motion term depends on A, Eq. 4.8 may permit experimental distinctions between the two terms. To minimize sensitivity to the Coulomb effects, for example, one can compare the isotope mean energy for each element. Such studies have been performed for light charge particles Z S 2. These studies show mean energies for 3He that are higher than those for 4He, contrary to Eq. 4.8 [122, 115], suggesting the emission of 3He particles from the surface of the emitting source rather than from an equilibrated freezeout. Since many of the light charged particles such as 3He are strongly emitted before the fragments, the question of whether fragments originate in a thermalized freezeout remains Open. To answer this question, one must measure the isotope energy spectra for fragments heavier than He. In Fig. 4.14, we show the energy spectra for 11C (open circles) and 12C (closed circles). The yield of 12C yield is nearly a factor of 10 higher than that for 11C due to its higher binding energy. The peak in the energy spectrum occurs at higher energies for 11C than for 12C and is broader. These two factors dictate a higher mean energy for 11C than for 12C. The measured mean energies of all the isotopes as a function of the mass number 127 /dEdO —5 510 2—-———ISMM O 0 data [Llllll l 10-6 1 1 1 1 I 1 1 1 1 I 0 50 100 M (MGV) Figure 4.14: The solid and open points represent t e measured center of mass energy spectra for 11C and 12C fragments, respectively. The solid lines represent the cor- responding ISMM calculations. The dashed lines represent the corresponding EES model calculations. 128 A are plotted in the left panel of Fig. 4.15. In this figure, the isotopes of each element have the same symbols. The even Z (Z =2, 4, 6, 8) elements are represented by closed symbols and the odd Z (Z=1, 3, 5, 7) elements by the open symbols. Generally, < ECM > increases with A; however, the lightest isotopes in each element (3He, 6Li, 7Be, 10B, 11C, 13N) display a significantly larger mean energies than the neutron rich isotopes. (Due to insufficient statistics, the uncertainties in the measured mean energy of the neutron deficient oxygen isotope, 15O is very large and not usable.) This striking trend is contrary to the behavior described by Eq. 4.8 for fragment emission from a single equilibrated source. llllllllilllllllll’T LITTITWTTIITIIAW’TT _IITTIYTIIIIIIIIIITI 60— 1‘c . 111:5: 14 1 — ISMM — - ] N1 _ . . ‘ _ ‘1, __ ‘ _ _ 11‘ n‘_ B ‘1 ‘ _ <><3d<> _ 5O — ' . ”9 — ) N ' L U ' 1 AK. ,9 /\ ' ' 5 " 6 . a L __ [213‘ ‘ z ‘ r“ If! * L1 ‘ 19 V :SHe :. . f? .- ._ _ ~ ' - ‘. tiff) £6 - ll . a- ‘1), . a- I a 30 T i —._ o/ -_ GO _ F . ‘ [.. __ __ ~ ~ . 13.1121 - Q/ 20 . ne‘er” 112S11__ *— — bLllllll|lllll|IlllI.‘_lll|lllllllllllllll ~1111l1_111l1111l1111 O 5 10 15 20 5 10 15 20 5 10 15 20 A Figure 4.15: Left panel: experimental fragment mean kinetic energies. Middle panel: mean kinetic energies calculated with the EES model. Right panel: mean kinetic energies calculated with the ISSM model. While Eq. 4.8 describes the trend expected for an equilibrated system of excited 129 fragments, sequential decay of the excited fragments emitted from a single equili- brated source can change this trend. To estimate such effects, we use the improved Statistical Multi-fragmentation Model (ISMM) of ref. [103] to model equilibrium breakup configurations composed of excited fragments and light particles and calcu- lated the sequential decay of the excited fragments with a Monte-Carlo Weisskopf model [119]. To compare with experimental results, we calculated the energy spectra from the decay of an initial source with Ao=168, Zo=75 and E‘/A= 6 MeV (roughly 75% of the total mass, charge and energy of the combined projectile and target [106]). Initially, we assign randomly a thermal velocity to each fragment and light particle according to a Boltzmann distribution characterized by a breakup temperature of T: 5.24 MeV [103] and we position each particle or fragment randomly within a volume of 8.3 fm. We add a collective velocity v = fivcdz,m to the thermal velocity, and vary 1260“,”, = \/2(Ecou/A)/mN (i.e. Emu/A) to describe the mass dependence of the experimental mean energies. We solve the classical equations of motion these particles interacting via Coulomb forces. The optimal choice of collective velocity depends on details of the placement of fragments within the volume. If one excludes initial configuration that place any part of a fragment outside R, the accepted configurations fragments will be more concentrated in the interior than if one excludes configurations that place the center of mass of any fragment outside R. In the former case, the fragments experience larger Coulomb forces on average and the collective energy per nucleon needed to reproduce the < E], > data, ECou/A z 0, is less than that for the latter case, ECou/A z 2MeV. Regardless of the details of the spatial distribution of fragments at breakup, however, these equilibrium calculations do not reproduce the measured enhancements of the mean energies of the most neutron deficient fragments. 130 To demonstrate this, we show results for a calculation in which the centers of mass of fragments are randomly distributed over a spherical volume of radius R. Predicted 11C and 12C spectra (normalized to the data) are shown as blue lines in Fig. 4.14. The calculation reproduces the peak of the 12C spectrum better than it does the peak of the 11C spectrum; neither the 11C nor the 12C calculation reproduces well the high- energy tails of the corresponding experimental spectrum. The right panel in Fig. 4.15 shows that the predicted mean energies does not reproduce enhanced values observed experimentally for the N < Z neutron deficient fragments. The slightly depressed values for < E], > calculated for symmetric N = Z fragments reflect strong secondary decay contributions to these N = Z nuclei from the secondary decay of heavy particle unbound nuclei that have somewhat smaller initial velocities. The strength of these secondary decay contributions reflect the Q-value for secondary decay, which favor decays to well bound N = Z nuclei and suppresses decay to their N < Z neighbors, but the calculated change in < E], > is very much smaller than that observed experimentally. Changing the assumptions about the spatial distribution of fragments or the collective velocity at breakup will not change this conclusion. This failure and the aforementioned evidence for surface emission of helium iso- topes suggest that fragments may also be emitted from the surfaces of the system while it is expanding and cooling. Because the binding energies of the neutron defi- cient isotopes are significantly less than those of their neutron rich neighbors, their surface emission rates will decrease relative to their neutron rich neighbors as the system cools. The Expanding Emitting Source (EES) model [39] provides a means to test this scenario. First the model predicts a sequence of source conditions that reflect the cooling of system, rather than a single one. Second it provides a self-consistent calculation of unique sequences of emission rates for each isotope. To illustrate these ideas, we have performed an EES calculation, which assumes 131 that particles can be radiated from the surface of the expanding system prior to bulk disintegration. Unlike equilibrium models, which assume the system to have already expanded, this time dependent model calculates the expansion and emission of the system beginning at an earlier time as it expands through saturation density. Specifically, the model assumes surface emission for p Z 0.4po, bulk emission for p S 0.3po and a gradual transition from surface to the bulk emission for densities, 0.4po > p > 0.3p0. For our EES calculations, we take saturation density, E’/A=9.5 MeV, A0=224 and Z0=100 as the specific starting conditions. As one example of the EES model results, we show the time evolution of 11C and 12C yields. At the time of emission, the primary fragment of each isotope acquires an average kinetic energy dictated by the instantaneous Coulomb barrier, expansion velocity and temperature of the expanding system. (The early surface emission con- tributions increase the value of the Coulomb and collective contributions above those obtained from the bulk emission alone.) Taking this time dependence into account, we plot the < E], > values for 11C (blue dashed line) and 12C (red solid line) as a function of the time of emission in the left panel of Fig. 4.16. Over the evolution of the source the carbon isotopes are emitted with a range of kinetic energies but there is very little difference between the values for the two carbon isotopes at any given time. - We next examine the emission rates as a function of time for the two isotopes. This is shown in the lower right panel of Fig. 4.16 where the instantaneous rates for each isotope are plotted and in the upper right panel of Fig. 4.16 where the ratio of the rates is plotted. Here we find that the rate for the emission of 11C relative to that of 12C changes drastically with time. The emission of the former is enhanced at earlier time and the latter at later time. Within the EES model, the rates are determined primarily by three factors: the spin degeneracy factors of each channel, 132 IIII[TIIIIIIIIIII lllllllllllllll 100 P l l A _ \ , > 1 X m \ I” 2 r 1 [.1 V l,‘ ,1], Q l "'1 _ A50 ‘1‘ I’ z 2 \ ’1 \ o _ 1.. 1:1 ..__ 3 V r A O L \ 1'15 F 3 _ p I]; \ v +- // 4“ - O1111l1111i1111111LII/1:"1l1111i1111l‘1‘1‘ 0 50 100 1500 50 100 150 t (fm/c) Figure 4.16: Left panel: Mean center of mass kinetic energies for 12C(HC) calculated as a function of time with the EES model. Lower right panel: Emission rates for 12C (“0) calculated as a function of time with the EES model. Upper right panel: Ratio of the emission rate for 11C divided by the emission rate for 12C calculated with the EES model. 133 free energy, exp((Nf,’;(T) + Z f;(T)) /T), and binding energy, erp(—Q/T). Here, f;(p)(T) are the excited free energy per neutron (proton) of the source, and Q is the Q-value associated with the emission. The degeneracy factors favor 11C. The values for f "‘ are negative and hence the isotope with fewer nucleons is also favored by this factor. The magnitude of f *, however, goes to zero like T2 so the relative advantage for 11C arising from this factor deceases as the temperature goes down. The Q values are greatly influenced by the respective binding energy factor, which strongly favors the N = Z isotopes; this preference strengthens at reduced temperature. The net effect is that the preference for the 11C at the highest temperatures shifts to 12C as the temperature falls with the emission and expansion. We tested whether the binding energy is the dominant consideration by forcing the binding energies for 11C and 12C to be equal. In this test calculation, the relative emission rates for the two isotopes changed little with time. In the EES model, 11C is preferentially emitted earlier than 12C, and the shapes of the energy spectra are consequently not the same. In Fig.4.16, we show the cal- culated energy spectra for 11C and 12C (red dashed lines) by taking a convolution of spectra each arising from the instantaneous conditions of the source. The EES model reproduces the shift in the peak of the energy spectra, but somewhat underpredicts the slope of the spectra for both isotopes. We note that it is necessary to multiply the EES model predictions for both 11C and 12C by a factor of 0.3 to match them to the data. We attribute this reduction to the fact that the emission of elements with Z > 10 are not considered in these EES model calculations. The total yield of 12C contains contributions from the neutron decay of excited 13C and the a decay of excited 16O. The yields of 11C are not affected significantly by sequential decays. Integrating over the energy spectra, we find an average kinetic energy of about 56.7 MeV for 11C and 45.2 MeV for 12C. The difference of about 11 134 MeV is in qualitative agreement with the data. This same scenario applies to the other elements, each of which shows similar patterns for relative emission. We show the EES calculations for the mean energies of all isotopes in the middle panel of Fig. 4.15. The EES model reproduces the basic trends of the data. Besides the success of these schematic calculations, other considerations support the hypothesis of an early surface emission of fragments prior to bulk emission from an expanded system. We note that surface evaporation and fission are the basic decay modes of equilibrated compound nuclei formed at lower incident energies. 'II'ansport theory, which describes preequilibrium emission in collisions at Fermi energies and above, predicts the abundant emission of nucleons and clusters as the system expands to low density. The early emission of fragments should not be surprising. Indeed, the preequilibrium emission of fragments have been previously reported in for proton induced reactions [87], and similar isotopic effects were observed. Here we have shown that such effects are relevant even for the multifragmentation process that have been heretofore interpreted widely as an equilibrium process. Now, the task before us is to understand how this early emission and other non-equilibrium effects impact the utilization of multifragmentation as a probe of the liquid-gas phase transition of nuclear matter. 4.3 Isospin Difl'usion in the Peripheral Collisions and Model Calculation 4.3.1 Data from the Peripheral Collision We used peripheral collisions for our isospin dynamics study. As experimental ob- servables, we focus on features of isotopic yields Y,(N, Z) of particles measured for cc '1) 2 reaction . Here, N and Z are the neutron and proton numbers for the detected 135 fragments. The diflerential multiplicities of different isotopes were obtained by first selecting the peripheral collision events with f) > 0.8, then the background level was approximated by fitting each element on the 1D PID histogram. Since the statistics is not high enough to allow us to do the background level approximation with many different energy bins, only two bins are used, for particles stop in the second silicon detector or stop in the CsI crystals, respectively. The efficiency correction was done by dividing the measured 2D yr-pi histogram by the corresponding efficiency function to get a corrected 2D yr-pj histogram. A histogram of corrected yield of 12C after reflection can be found in Fig. 4.17. A “Coulomb hole” can be seen clearly at the rapidity region near that of the projectile. These 2D yr-p 1. histograms can be used to get the isotopic yields in different rapidity bins. To study the isospin diflusion, projectile rapidity fragments were selected by a ra- pidity gate of y, = y/ybeam > 0.7, where y and ybeam are the rapidities of the analyzed particle and the beam, respectively. The contributions from the neck fragments are minimized in this rapidity region. Fig. 4.18 shows the measured average differential multiplicities of Li, Be, B, C, N and O isotopes at y, > 0.7 which can be compared with the corresponding data from central collisions. One can see that the yields from 112Sn+1243n and 124Sn + 112Sn are not the same, especially in the tails of the isotope distributions. This property can be used to study the isospin diffusion. We analyzed the ratios of isotopic yields R21(N, Z) = 13(N, Z) / Y1(N, Z) for a specific pair of reactions with different total isotopic composition and found that they follow the same isoscaling relationship as described in Eq. 4.2 [110]. It suggested this may be a compact and accurate way to quantify the isotopic asymmetry effects caused by isospin diffusion. The left panel of Fig. 4.19 shows measured values for R21(N, Z) assuming 124Sn + 124Sn collisions as reaction 2 and 112Sn + 11"’Sn collisions as reaction 1 in Eq. 4.2. The 136 p i/A(MeV/c) “ E a -50 -100 -150 -2000 . Figure 4.17: Efficiency corrected diflerential multiplicity dM/dydpi for 12C fragments in peripheral collisions. A “Coulomb hole” can be clearly seen at 31, ~ 0.05 and y, ~ 0.95. UIO GOO 0.8 < 0 <1 112Sn+“28n 12C 0.21461 10.204106018‘ ‘1' ”1.2 yr 137 400 350 300 250 200 150 100 50 0 - 1""""""""E/1;5'o'Mev'1 M. 124Sn+124$n 08 <’b< 1 [‘3 a _ <> 112Sn+121Sii 0'7 < Y/Ybeam (f 4 \'\\ _ _ . 121Sn+ fsn . 10—2 0 \’:\ N :/ — LI V\\\. O C». Q 0 \‘9 /;°\\ 0 / ' \\ ,. [iv/1 \\.\\ ‘ B ?§»\\. XO.1 _4 — \ . \// 10 s 1 f../ ' \\ ‘ 3 \ \g x001 I o\, N Q» \ \ \ 9 (,1 \ § 10‘6 — "f I l n l l l l 1i 2 Figure 4.18: Average differential multiplicities as a function of neutron excess (N — Z) of the isotope for peripheral collisions, f) > 0.8, for Li, Be, B, C, N and O isotopes with rapidity near that of the projectile. The solid circles are data for the 124Sn+12"Sn system with N /Z 21.48. The open circles are data for the lightest system 112Sn+”2Sn with N /Z=1.24. The solid diamonds are data for system 124Sn+“2Sn. The open diamonds are data for system 1128n+mSn The solid, dashed and dot-dashed lines are drawn to guide eyes. 138 fit to Eq. 4.2 are shown as the solid and dashed lines. Fits to the other systems (not shown) are of similar quality. In the right panel of Fig. 4.19, we plot the best fit values for the isoscaling parameter, 0, versus the overall isospin asymmetry of the colliding system: 60 = (No — Z0) / (N0 + Z0), where No and Z0 are the corresponding total neu— tron and proton numbers. The solid and open points denote data for 124Sn and 112Sn projectiles, respectively. In general, the isoscaling parameter (1 increases with the overall isospin asymmetry 60. Statistcal calculations, both from multifragmentation and evaporation predict that the isoscaling parameter should depend approximately linearly on the isospin asymmetry 60 of the entire system. If the linear correlation between a and 60 [123] is assumed, the mixed systems at 60 = 0.153 should assume an a value midway between that of 124Sn + 124Sn (60 = 0.193, a=0.57:l:0.02) and 112Sn + 112Sn (60 = 0.107, 0120). However, the measured value for the 124Sn projectile (solid point) is much larger than the value for the 112Sn projectile (open point) at 60 = 0.153 indicating that the emitting source is not consistent with the achievement of isospin equilibrium is not achieved in the asymmetric reaction systems [33]. The upper value of a=0.42:l:0.02, obtained for 124Sn + 112Sn, represents an effective asymmetry of about 0.17. This value corresponds to roughly half way from the projectile value of 0.193 to the “equilibrium” value of 0.153. The lower value of 0.16:1:0.02, obtained for 112Sn + 124Sn, has the same interpretation; except here the projectile is 112Sn and the change in asymmetry is in the opposite direction. The consistency of these results extracted from two independent measurements adds credibility to the approach. Because there are no isospin differences between identical projectiles and targets, we use symmetric 112Sn + 112Sn and 124Sn + 124Sn collisions in the following analyses to establish diffusion-free baseline values for the measured and predicted observables. The asymmetric 124Sn + 112Sn and 112Sn + 124Sn collisions, on the other hand, have the large isospin differences needed to explore isospin diffusion. Following ref. [86], 139 0.6 ”E U) 51’. fi+ C1 0) ,9); 0.4 3: > C5 C1 to S _+_ 1: 1 0.2 U) j» R? . :07: ;— b/bmax>0.8 i, ,’ : : / 05: Y/ybeam>0-7 <~1211128n+1128n — 0.0 : 1 L n 1 1 1 l 1 1 1 ‘ 1 1 1 1 l 1 1 1 1 10 15 0.10 0.15 0.20 A 50=(No‘zo)/(No+zo) Figure 4.19: Left panel: Measured values for R21(N, Z) = K241124(N, Z)/Y112+112(N, Z) (points) and fits with Eq. 4.2 (solid lines). The solid line and points represent even Z =4,6,8 isotopes while the dash lines and open points represent odd Z 23,5,6 isotopes. Right panel: Best fit values for a as function of 60. The lines serve to guide the eye. The reactions are labeled next to the data points. 60 = 0.107 and 0.194 for the symmetric systems of 124Sn+124Sn and 112Sn-l—“zSn respectively. The two mixed systems have the same 60 = 0.153. Solid points denote 124Sn as the projectile and open points denote 112Sn as the projectile. 140 we define the isospin transport ratio, R, as 2113 - $124+124 '— 50112 +112 R,‘ = , (4.9) $124+124 — $112+112 where :1: is an isospin sensitive observable. For the two symmetric systems 124Sn + 124Sn and 117‘Sn + 112Sn, R,- is automatically normalized to +1 and —1, respectively, allowing quantitative comparison of the measured and predicted R.- values even if the model calculations use isospin observables that are different from the experimental ones. The only requirement is that both the experimental and theoretical observables depend linearly upon the isospin asymmetry 60 of the emitting source. To represent the experimental data, we chose the neutron isoscaling parameter (1 because the pro- jectiles and targets for the asymmetric systems differ only in their neutron numbers. Using a for :1: in Eq. 4.9, we obtain the isospin transport ratios of the two asymmetric systems shown as the shaded bands in Fig. 4.21. The observed values, |R,(a)| z 0.5, are consistent with previous discussion that the isospin asymmetry of the projectile remnant is half way between that of the projectile and the “equilibration value”. Changes in a may result from both isospin diffusion and preequilibrium emission. In the absence of isospin diffusion, preequilibrium emission from the projectile should be approximately equal for asymmetric 124Sn + 112Sn (“28n + 124Sn) collisions as for symmetric 124Sn + 124Sn (llzsn + 112Sn) collisions. By focusing on diflerences between mixed and symmetric systems, R,(a) largely removes the sensitivity to preequilibrium emission and enhances the sensitivity to diffusion. We can use other physical observables to construct R,. Since mirror nuclei have similar secondary decay modes, we can reduce the effects of the secondary decay and enhance the effects of diffusion if we take the ratio of the mirror nuclei. Fig. 4.20 shows Rn, constructed from ln[Yz'eld(7Be)/Yz'eld(7Li)], as a function of relative rapidity in the center of mass, (y — ycm)/ybeam. The Ry, from reaction 112Sn + 124Sn is reflected about 0 to give an overall trend. Note, we use ln[Yz'eld(7Be)/Yield(7Li)] instead of I41 the ratio Yield(7Be)/Yield(7Li) because ln[Yz'eld(7Be)/Yield(7Li)] cc ()1, — 11,,)/T just as 0 oc Ayn/T. We expect this to give a corrected ratio R,- which is closer to linear dependence on the isospin asymmetry 60 of the emitting source than does Yield(7Be)/Yield(7Li). We see that R17 agrees with R,(a) for |(y—ycm)/ybeam| 2 0.2. In addition, Rn give us the ability to explore isospin diffusion as a function of y and P1- 1.5 l\ - m ; 1n(Yield(7Be)/Yield(7Li)) 1; os<9<1o I r 0-5 :Iiiifiifiil::11:iiiiiliiiiiiiiiifiliffiIZlIifiiin _ I: 0 l ' : .1 1 _.. _ I ............... I-~ —0.5 .' --------------------------------------------------- 1 i if: 1315“ : S + SI] 1 m1 ln 1 m L m 1 n n l 1 "LS-0.6 1 16.4 —0.2 0 10.21 0.4' 10.6 (y-ycmyybeam Figure 4.20: R17, constructed using ln[Yz'eld(7Be)/Yz'eld(7Li)], as a function of (y — yan)/ybeam- 142 1OLizfsprlfisajaqéiifgsjqn)________§ x . p2 p 101/3 ‘” f 0 - 5 ;_ ' “MKJQZJQWYIYIYIZIYIYIQW' 0100- ....................................... e .............. _ o: 8 ° v i 5 "0'5. DATA —1-0fris————-——————--—-—-—-----} Figure 4.21: Average measured (shaded bars) and calculated (open points) values for R,- . The labels on the calculated values represent the density dependence of E8ym,,-,,t/A with decreasing stifl'ness from left to right. 143 4.3.2 BUU Calculations We now explore the relationship between isospin diffusion, and the asymmetry term of the EOS within the context of the Boltzmann-Uehling-Uhlenbeck (BUU) [6, 62] formalism, which calculates the time evolution of the colliding system using a self— consistent mean field. The isospin independent part of the mean field in these calcu- lations is momentum independent and described by an incompressibility coemcient of K =210 MeV [93]. The interaction component of asymmetry term in three sets of calculations provides a contribution to the symmetry energy per nucleon of the form Esym,,-m/A=C,ym(p/po)7 where Cay", is set to 12.125 MeV and 7:2, 7:1 and 7:1/3 . Here, smaller values for 7 dictate weaker density dependence for anm,int- The fourth set of calculations, referred to as SKM, uses an interaction asymmetry term providing Esym,,-,,t/A =38.5 (p/po) — 21.0(p/p0)2 [2] and has the weakest density dependence. Calculations were performed for the 124Sn + 124Sn, 124Sn + 112Sn, 112Sn + 124Sn and 112Sn + 112Sn systems at an impact parameter of b=6 fm [93]. We employed ensembles of 800 test particles per calculation and followed each calculation for an elapsed time of 216 fm/c. At this late time, the projectile and residues can be cleanly separated. Nonetheless, we require that all nucleons in the assigned regions to have density less than 0.05po and velocities above half the beam velocity in the center of mass to be consistent with the experimental gates. To reduce statistical fluctuations in the results, we averaged 20 calculations for each system. Using the average asymmetry of the pro jectile-like residue < 6 > in the calculation, as the isospin observable, x, in Eq. 4.9, we plot predictions for R,(6) as a function of time as bands in Fig. 4.22 for the stiffest asymmetry term, (p2 , top panel) and the softest asymmetry term, (SKM, bottom panel). By construction, [2,-(6) describes the evolution of isospin asymmetry for the projectiles (112Sn or 124Sn) in the mixed reactions relative to that for the symmetric 124Sn + 124Sn (with R,- = 1) and 112Sn + 144 112Sn (R,- = —-1) systems. The widths of the bands reflect the statistical uncertainties of the calculated values for R,(6). Initially, these predictions for R,(6) represent the isospin of the projectile (R.(6) = 1 for 124Sn and 1%.-(6) = -—l for 112Sn) nuclei. Subsequent isospin diffusion drives the R;(6) values towards zero. Even though pre- equilibrium emission from the projectile remnants influences < 6 >, R;(6) is not strongly modified because such pre-equilibrium emission is largely target independent and therefore cancelled in R,(6) by construction. The influence of the asymmetry term depends on its magnitude at sub-saturation density [62, 93]. For the top panel, the energy is Ewmmt/A =Csym(p/po)2, which decreases rapidly at low density and becomes very small, leading to little isospin diffusion. For the bottom panel, E,ym,,-,,¢/A =38.5 (p/po) — 21.0(p/p0)2 remains larger at low density, leading to stronger isospin diffusion and bringing the residues to approximately the same isospin asymmetry. In both cases, the asymptotic value for R,(6) is first reached at around 100 fm/c when the two residues separate and cease exchanging nucleons. This timescale is comparable to the collision timescale Tcou, which can be roughly (~20%) estimated by Tcou z (4RN + d)/vbeam ‘3 80fm/c , (4.10) where RN, vbeam and d are the nuclear radius, incident velocity, and the separation d between the two nuclear surfaces at breakup as illustrated by the time evolution images of the collisions for the 124Sn + 124Sn system in Fig. 4.22. Assuming that the experimental isoscaling relationships, shown in Fig. 4.19, re- flect particle emission from the projectile remnants after 100 fm/c and that such emission can be described statistically, the calculated values for R,(6) may be easily related to the measured ones. In doing so, we take advantage of the nearly linear re- lationship between a and the 6 values that has been shown valid for evaporation and for statistical multi—fragmentation of the remnants that emit the observed particles 145 -[o 11an Elab=50MeV, b=6.5fm ‘ 1 l 1 l 1 1 I 1 l 1 0 20 40) 60 80 100120140 time(fm/c) Figure 4.22: R,(6) from BUU calculations are plotted as a function of time for the mixed systems. The symmetric systems are calibrated to +1 and —I automatically by Eq. 4.9. The top and bottom panels show the calculated results with Em,.-,,¢/A =C,,’,,,,(p/p0)2 and Eme-m/A =38.5(p/po)2 - 21.0(p/p0)2, respectively. The bands above (below) the dashed line represent the system with 124Sn (llZSn) as the projec- tile. The time evolution images of the collisions for the 124Sn + 124Sn systems are superimposed on the upper panel suggesting that the projectile and targets separates around 100 fm/c. 146 [112]: 02 + 01 2 aoc (62—61)(1— ), (4.11) where 61 and 62 are the asymmetries for the two systems involved in the isoscaling ratio. Inserting Eq. 4.11 into the expression for R;(oz) , one can show (61' — 6112+112)(6i _ 61244-124) Ri = Bi (5 _ O (a) ( ) (5124+124 _ 5112“”) — (1 — (6124+124 + 5112+112)/2) (4.12) Since the second term is negligible (< 4%), we can obtain < 6 > from the flat region at 216 fm/c, use it as :1: in Eq. 4.9 to obtain R,(6) values for these two systems and we can compare them to the experimental values of R;(a). The calculated values, shown as open points in Fig. 4.21 in the order of an increasing “softness” from left to right indicate an increased isospin equilibrium for successively softer asymmetry terms. The diamond shaped points indicate the R,(a) values obtained as in the experiment by decaying the residues simulated in BUU to fragments using the statistical model of ref. [112]. As expected, there is close agreement between the diamonds and the corresponding open points according to Eq. 4.12. Fig. 4.21 demonstrates the sensitivity of such data to the asymmetry term of the EOS. The experimental R;(a) values are closest to the predicted R.-(6) values derived from the theoretical predictions for < 6 > using the stiflest asymmetry term, (p2, top panel) with Esym,,-,,t/A =C,ym (p/po)2. This conclusion depends, however, on the assumption that the measured particles are produced after 100 fm/c when < 6 > attains its asymptotic values. If the data include emission from earlier stages when 6 is larger, a favorable comparison with calculations using softer asymmetry terms may be possible. Relevant determinations of the emission timescales for the detected particles are being explored [93] and may serve to make the present conclusions regarding the density dependent asymmetry term more definitive. 147 Chapter 5 Summary In this dissertation, we studied the asymmetry term in the EOS and isosipin diffusion using heavy-ion collisions at the intermediate energy. Calculations have shown that experiment observables such as the energy spectra, intermediate mass fragment (IMF) multiplicities, isoscaling parameters and various nuclei yield ratios are sensitive to the asymmetry term in EOS. We show for the first time in this dissertation that isospin diffusion rates measured in our experiment can also provide constraints on the symmetry term in the EOS. We have measured the isotope energy and yields of Z=2-8 particles emitted in four different Sn+Sn reactions with different isospin asymmetry at 50MeV/ A. To make the measurement, nine telescopes of the Large Area Silicon-Strip/CsI detector Array (LASSA) were developed and used in the experiment. The LASSA telescopes provided a high isotopic, energy and angular resolution which are not available in the previous experiments. Impact parameters were selected by the multiplicity of charged particles, measured with LASSA and the Miniball/Miniwall phoswich array. Techniques were developed to enable the subtraction of background in the ex- perimental data and the correction for the experimental detection efficiency. The background levels were mainly estimated at the elemental levels while for those weak 148 isotopes are estimated at the isotopic level. The efficiency correction was done us- ing efficiency functions obtained from Monte Carlo simulations of the experiment. These efficiency simulations take care of the multiple hits, the inefficiencies in the solid angle and loss at low p 1- We examined the accuracy of the efficiency correction by comparing forward angle data and backward angle data obtained in symmetry collisions. From this comparison, we extract a systematic error bar for our data. For the part which can not be corrected by efficiency function, i.e. the inconsistence of the detection in silicon and C31, a systematic error bar is given. The total charged particle multiplicity detected in Miniball/Miniwall and LASSA was used for impact parameter determination. We selected central collision events, corresponding to a reduced impact parameter of 6 S 0.2 and the top 4% of the charged-particle multiplicity distribution. The impact parameter is estimated to be less than 1.5 fm according to total cross section measurements. The peripheral colli- sions are selected corresponding to a reduced impact parameter of 6 > 0.8 and bottom 40% of the charged-particle multiplicity distribution. The parameter is estimated to be greater than 6.0 fm in this case. We used the isotope distributions for Z =2-8 from the central collision to compare with the results from dynamic model calculations which include fluctuations that give rise to fragment production. The experimental data display a strong dependence on the isospin asymmetry that can be accurately described by an isoscaling parameter- ization. The theoretical calculation reproduces the yields for the heavier fragments with Z =6—8, but underpredict the yields of the lighter ones which are not strongly produced by the calculation as primary fragments. The calculated final isotopic dis— tributions display isoscaling, but the calculated isotopic distributions are narrower, more neutron deficient; and show a weaker dependence on the isotopic asymmetry of the system than do the data. The density dependence of the asymmetry term 149 of the EOS has an effect on the calculated final isotopic distributions. The distri- butions calculated using the asymmetry term with stronger density dependence are more neutron—rich and are closer to the measured values. These trends are simi- lar to prior results obtained for a BUU-SMM hybrid model, but different from the trends for evaporated fragments predicted by EES rate equation calculations. The present level of agreement between theory and experiment precludes definitive state- ments about the density dependence of the asymmetry term of the EOS. A number of theoretical issues, such as the pre—equilibrium of bound clusters, the calculations of fragment excitation energies, and the importance of the extreme tails of the primary fragment distributions, may influence the calculated results. Additional theoretical work is required to explore these issues and to determine the role they may play in the resolution of these discrepancies. Complimentary measurements of the yields and energy spectra of light particles can help to determine whether the missing neutrons in the calculations are carried away primarily by pre—equilibrium emission during the compression-expansion stage or during the later evaporative decay of the hot frag- ments. We compared the isotope energy spectra from central collisions with the data. We found that calculations from the bulk multi-fragmentation models, i.e. the im- proved Statistical Multi-fragmentation Model(ISMM), can not reproduces well the high-energy tails of the corresponding experimental spectra for heavy fragments like 11C and 12C. Neither can the bulk multi-fragmentation models reproduce enhanced mean energy values observed experimentally for the N < Z neutron deficient frag- ments. The Expanding Emitting Source(EES) model provides a better description of the measured energy spectra of the IMF isotopes produced in multifragmentation and mean energies of all isotopes. This success and other evidence for surface emissions reveal the dynamics of the emission process. The more energetic neutron deficient 150 isotopes are consistent with the picture that they are emitted earlier from the surface of the system while the source is expanding and cooling. We have observed the influence of isospin diffusion with comparable diffusion and collision timescales. Simulations using the BUU transport model predict that isospin diffusion reflects driving forces arising from the asymmetry term of the EOS. The comparisons of experimental data and results of calculation suggest better agreement with stiff aymmetry term. This result agrees with what we have found by comparing the isotope distributions from the central collisions with SMF calculations. However, more stringent constraints on the emission timescales for the measured particles in the BUU calculations are needed. The version of the BUU used in our comparison does not have the momentum dependence of the mean field symmetry term. Recently new calculations by other groups indicate that the momentum dependence of the asymmetry needs to be carefully addressed. This issue as well as other issues such as the dependence of isospin diffusion on the impact parameter and the influence of the in—medium neutron-proton cross section by the rasidual interaction should be explored in more detailed future work. 151 LIST OF REFERENCES [1] S. Albergo et al. , Nuovo Cimento 89, 1 (1985). [2] V. Baran et al. , Nucl. Phys. A632, 287 (1998) [3] V. Baran et al. , Phys. Rev. Lett. 86, 4492 (2001) [4] V. Baran et al. , Nucl. Phys. A703, 603 (2002) [5] G.F. Bertsch et al. , Phys. Rev. C29, 673 (1984). [6] G. Bertsch et al. , Phys. Rep. 160, 189 (1988) [7] H.A.Bethe, Rev. Mod. Phys. 62, 801 (1990) [8] I. Bombaci, “Equation of State for Dense Isospin Asymmetric Nuclear Matter for AstrOphysical Applications”, “Isospin Physics in Heavy-IonCollisions at Intermediate Energies”, Eds. Bao-An Li and W. Udo Schroeder, NOVA Science Publishers, Inc. (New York), p. 35 (2001) [9] I. Bombaci, in Perspectives in Theoretical Nuclear Physics, edited by S. Rosati et al. , (Edizioni ET S, Pisa, 1996), p. 233 [10] JP. Bondorf et al. , Nucl. Phys. A443, 321 (1985). [11] J .P. Bondorf et al., Phys. Rep. 257, 133 (1995) and refs. therein [12] B. Borderie et al. , Phys. Lett. B388, 224 (1996) [13] B. Borderie et al. , Phys. Rev. Lett. 86, 3252 (2001) [14] J .P. Bondorf et al., Nucl. Phys. A475, 663 (1987). [15] DR. Bowman et al., Phys. Rev. Lett. 67, 1527 (1991). edited by S. Rosati et al. (Edizioni ETS, Pisa, 1996), p. 233 [16] DR. Bowman et al. , Phys. Rev. Lett. 70, 3534 (1993) [17] BA. Brown, Phys. Rev. Lett. 85, 5296 (2000). [18] BA. Brown, Phys. Rev. Lett. 85, 5296 (2001) 152 [19] C. Cavata et al. , Phys. Rev. C42, 1760 (1990). [20] R.J. Charity et al., Nucl. Phys. A483, 371 (1988) [21] GB. Chitwood et al., Phys Lett. B172 , 27 (1986) [22] N. Colonna et al., Nucl. Instr. and Meth. A321 529 (1992) [23] M. Colonna et al., Phys. Rev. C57, 1410 (1998). [24] M. Colonna et al. , Nucl. Phys. A642, 449 (1998) [25] E. Cornell et al., Phys. Rev. Lett. 75, 1475 (1995) [26] C. Csoto and G. M. Hale, Phys. Rev. C55, 2366 (1997). [27] M. D’Agostino et al. , Phys. Lett. B371, 175 (1996) [28] P. Danielewicz et al., Phys. Rev. C38, 120 (1988). [29] P. Danielewicz et al., Nucl. Phys. A533, 712 (1991) [30] P. Danielewicz, Nucl. Phys. A673, 375 (2000). [31] P. Danielewicz et al. , Science 298, 1592 (2002) [32] B. Davin et al., Nucl. Instr. and Meth., A473, 301 (2001) [33] J.F. Dempsey et al., Phys. Rev, C54, 1710 (1996) and refs. therein [34] J .F. Dempsey, Ph.D. Thesis (Washington University, 1997) [35] H. Feldmeier et al. , Nucl. Phys. A586, 493 (1995) [36] JD. Frankland et al. , Nucl. Phys. A689, 940 (2001) [37] W.A. Friedman and WC. Lynch, Phys. Rev. C28, 16 (1983). [38] W.A. Friedman and WC. Lynch, Phys. Rev. C28, 950 (1983). [39] W.A. Friedman et al. , Phys. Rev. C42, 667 (1990). [40] B. Gatty et al. , Nuc. Phys. A253, 511 (1975) [41] AL. Goodman et al. , Phys. Rev. C30, 851 (1984). [42] D.H.E. Gross et al. , Z. Phys. A309, 41 (1982). [43] D.H.E. Gross, Phys. Rep. 279, 119 (1997) [44] A. Guarnera et al. , Phys. Lett. B373, 373 (1996); A. Guarnera, Thesis U. Caen (1996) 153 [45] DO. Handzy, Ph.D. Thesis (Michigan State University, 1995) [46] RE. Haustein et al. , At. Data Nucl. Data Tables 39, 185-395(1988) [47] D. Horn et al. , Nucl. Instr. and Meth. A320 273 (1992) [48] W.G. Hsi et al. , Phys. Rev. Lett. 73, 3367 (1994) [49] M.J. Huang et al., Phys. Rev. Lett. 78, 1648 (1997). [50] SC. Jeong et al. , Phys. Rev. Lett. 72, 3468 (1994) [51] LP. Kandanoff and G.Baym, Quantum Statistical Mechanics (Benjamin, New York, 1962) [52] YD. Kim, et al. , Phys Rev. C67, 14(1991) [53] YD. Kim, et al., Phys Rev. C45, 338(1992) [54] G.J. Kunde et al. , Phys. Rev. Lett. 77, 2897 (1996) [55] AB. Larionov et al., Nucl. Phys. A658, 375 (1999) [56] Y. Larochelle et al., Nucl. Instr. and Meth. A384, 167 (1994) [57] J.M. Lattimer et al., Phys. Rev. Lett. 66, 2701 (1991) [58] J .M. Lattimer and M. Prakash, Astrophys. J 550, 426 (2001); Phys. Rep. 333, 121 (2000) [59] J .M. Lattimer and M. Prakash Science 304, 536 (2004) [60] BA. Li et al., Phys. Lett. 3303, 225 (1993) [61] BA. Li et al. , Phys. Rev. C52, R1746 (1995). [62] BA. Li, et al. , Int. J. Mod. Phys. E7 (1998) 147, and refs. therein [63] BA. Li et al., Phys. Rev. Lett. 78, 1644 (1997). [64] BA. Li, Phys. Rev. Lett. 85, 4221 (2000) [65] BA. Li, Phys. Rev. Lett. 88, 192701 (2002) and refs. therein [66] W.G. Lynch, et al. , Phys Lett. 1683 , 274 (1982) [67] W.G. Lynch, et al. , Phys Rev. Lett. 51 , 1850 (1983) [68] W.G. Lynch, et al. , Phys Rev. Lett. 52 , 2302 (1984) [69] W.G. Lynch, Ann. Rev. Nucl. Part. Sci. 37 , 493 (1987), and refs. therein [70] N. Marie et al., Phys. Lett. B391, 15 (1997) 154 [71] N. Marie et al. , Phys. Rev. C58, 256 (1998) [72] RF. Mastinu et al. , Nucl. Instr. and Meth. A338 419 (1994) [73] LG. Moretto and G.J. Wozniak, Annu. Rev. Nucl. Part. Sci. 43, 379 (1993). [74] H. Mori, Prog. Theor. Phys. 8, 327(1952) [75] H. Miiller and B. D. Serot, Phys. Rev. C52, 2072 (1995) [76] P. Braun-Munzinger etal, Phys. Lett. B465, 15 (1999) [77] P. Braun—Munzinger etal, Phys. Lett. B596, 61 (2004) [78] A. Ono et al., Phys. Rev. Lett. 68 2898 (1992); A. Ono et al. Phys. Rev. C53, 2958 (1996) [79] A. Ono et al. Phys. Rev. C53, 2958 (1996) [80] One might imagine that the fragment isobaric distributions might display - 2 a Gaussian dependence on asymmetry, i.e. Y(N, Z) = f(A) exp [— 9,31] where 6=(N-Z) / A and 6 = (6) but we lack the statistics in our simulations to test this hypothesis [81] C.J. Pethick and DC. Ravenhall Annu. Rev. Nucl. Part. Sci. 45, 429(1995) [82] R. Planeta et al. , Phys. Rev. C38, 195 (1988), R. Planeta et al. , Phys. Rev. C41, 942 (1990). [83] R. Popescu et al., Phys. Rev. 58, 270 (1998) [84] L. Phair et al. , Nucl. Phys. A548, 489 (1992) and references therein [85] M. Prakash et al. , Phys. Rev. Lett. 61, 2518(1988) [86] F. Rami et al., Phys. Rev. Lett. 84, 1120 (2000) [87] E. Renshaw et al., Phys. Rev. C44, 2618 (1991) [88] L. Scalone et al. , Physics Letters B461, 9 (1999) [89] W.U. Schriider and J .R. Huizenga, in Treatise in Heavy-Ion Science, edited by DA. Bromley (Plenum Press, New York and London, 1984), Vol. 2, p. 113 and references therein [90] BM. Sherrill et al., Nucl. Instr. and Meth. B56/ 57 1106 (1991) [91] D.V. Shetty et al., Phys. Rev. C68, 021602 (2003) [92] L. Shi et al., Europhysics Lett. 51, 34 (2000) 155 [93] L. Shi et al., Phys. Rev. C68, 064604 (2003). [94] LG. Sobotka et al. , Phys. Rev. C55, 2109 (1997) [95] G.A. Souliotis, et al. , Phys. Rev. C68, 024605 (2003) [96] RT. de Souza et al., Nucl. Inst. Meth. A295, 109 (1990); D.W. Stracener et al., Nucl. Instr. Meth. A294, (1990) 485 [97] SR. Souza et al., Phys. Rev. C62, 064607 (2000) [98] R. St6cker et al., w.z. Phys. A203,259 (1981) [99] C. Sturm et al., Phys. Rev. Lett. 86, 39 (2001) [100] K. Sumiyoshi et al., Nucl. Phys. A552, 437 (1993) [101] W. P. Tan et al. , Phys. Rev. C64, 051901 (2001). [102] WP. Tan, Ph.D. Thesis (Michigan State University, 2002) [103] WP. Tan et al., Phys. Rev. C68, 034609 (2003) [104] WP. Tan et al., in preparation [105] I. Tanihata, Report No. RIKEN-AF-NP-229 (1996) [106] These values are consistent estimated that take cooling by preequilibrium emis- sion and expansion prior to breakup into account, e.g. P. Desesquelles et al. , Nucl. Phys. A633, 547 (1998) for example [107] M. Di Toro et al. , Nucl. Phys. A681, 426 (2001) [108] M. Di Toro et al., Eur. Phys. Jour. A13, 155 (2002) and refs. therein [109] MB. Tsang et al., Lett. B297, 243 (1992) [110] MB. Tsang et al., Phys. Rev. Lett. 86, 5023 (2001) [111] MB. Tsang et al., Phys. Rev. C64, 041603 (2001) [112] MB. Tsang et al., Phys. Rev. C64, 054615 (2001) [113] EA. Uehling and GE. Uhlenbeck, Phys. Rev. 43, 552 (1933) [114] A.M. Vander Molen et al., Users Guide to the MSU 47r Array 3’rd ed. (1996) [115] V.E. Viola et al., Phys. Rev. (:59, 2660-2669 (1999) [116] V.E. Viola, Nucl. Phys. A734, 487 (2004) [117] R. Wada et al., Phys. Rev. C62, 034601 (2000) 156 [118] A. Wagner et al. , Nucl. Inst. Meth. A456, 290 (2001) [119] V.F. Weisskopf, Phys. Rev. 52, 295 (1937) [120] G.D. Westfall et al., Phys. Rev. Lett. 37 , 1202 (1977) [121] OF. Williams, Ph.D. Thesis (Michigan State University, 1996) [122] H.Xi et al., Phys. Rev. C67, R462 (1998) [123] H.S. Xu et al. , Phys. Rev. Lett. 85, 716 (2000) [124] H.M. Xu et al., Phys. Rev. C40, 186 (1989) [125] S.J. Yennello et al., Phys. Lett. B321, 15 (1994) 157