JWMWJ ”5.3.55.3: a 9; .44.: 3.1; Q 3m: .1; .Q...«..&. .5. . 5., . . . 1.” MN». 12. r .. xiwzl... : 253 v... 1.. s... H yen .. . . {H.C. .x #7 :..n:.. :( X . . .3 . .. . . :75 v 1:92.; fismlt: 1.? « . . I‘ll-“HI. ..o‘hVY.:.’ . AN 4.! 1 abuh.’ 3 LIBRARY 7 0'7; MiChigan State University This is to certify that the dissertation entitled PROBING THE FREEZEOUT MECHANISMS AND ISOSPIN EFFECTS IN MULTIFRAGMENTATION presented by WANPENG TAN has been accepted towards fulfillment of the requirements for the Ph.D. degree in Physics \ MW Major Prfiessor’s Signature ,1 Zf/o 37/0 L Date MSU is an Affirmative Action/Equal Opportunity institution —-—--—a—--.—.-A-.-.-.-o—-c—o-c.---.-.-.-.-.-I-.-.-O-O-.- 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 6/01 c:/CIRC/DateDue.p65-p.15 PROBING THE FREEZEOUT MECHANISMS AND ISOSPIN EFFECTS IN MULTIFRAGMENTATION By Wanpeng Tan 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 2002 ABSTRACT PROBING THE FREEZEOUT MECHANISMS AND ISOSPIN EFFECTS IN MULTIFRAGMENTATION Bv ‘/ “’an pen g Tan IV'Iultifragmentation processes have been studied by measuring 129Xe+197Au col- lisions at 50A MeV with a large area silicon—strip/Csl detector array (LASSA). The LASSA was designed and constructed to provide excellent energy, angular and iso- tope resolution for the detection of charged particles. Each of nine LASSA telescopes consists of two silicon strip detectors (65/1111 and 500mm thick, respectively) and four CsI(Tl) crystals. A non-uniformity in light output of the CsI(Tl) crystals better than 100 was obtained via the crystal selection and quality control procedure. Isotopi- cally resolved particles from hydrogen to oxygen were detected in polar angles of 12" < 6 < 62° with an angular resolution of $0.43". Experimental correlation functions for particles detected in LASSA are analyzed using the Koonin-Pratt formalism and by making an assumption of thermal equilib- rium. A method of probing the breakup density is demonstrated and the results from the equilibrium approach are compared with those from the Koonin-Pratt formalism for light particle correlatitms. The extracted values of the free volume are about 2.5 (10% uncertainty) times the volume of the total system at saturation density, which corresponds to a freezeout density of 1/5-1/3 times the nuclear saturation density. A new technique of spin determination from particle correlation functions is pro- posed utilizing the same equilibrium assumption. A few examples of correlation functions are studied to demonstrate the promising sensitivity of this new teclmique. In particular, the spin of the first excited state of 8I3 at 0.774 MeV is determined as J21 via the p—7Be correlation function. To better understand the isospin effects and to further confront calculations with experimental data. an empirically modified statistical multifragmentation model (SMM) was developed. This model treats the properties and decay of nuclear excited states Self-consistently and calculates the secondary decay using the empirical infor- mation about the excited states. The importance of using empirical binding energies and experimental level densities is shown in detailed comparisons. Nuclear ther- mornetry is studied and con‘ipared with data to demonstrate that the corrections for secondary decay can be modelled accurately within this approach. From the isotopic composition of particles emitted during an energetic nucleus- nucleus collision, we can test the isospin dependence of the nuclear equation of state (EOS). Hybrid calculations were performed using isospin dependent BUU transport model and the improved SMM. Comparisons between measured isotopic yield ratios and theoretical predictions in the equilibrium limit are used to assess the sensitivity to the density dependence of the asymmetry term of the EOS. This analysis suggests preference for a stiff density dependence and indicates that such comparisons may provide an opportunity to constrain this important property of nuclear matter. To my wife - LII/wt Zhai and my parents - G'uoyang Tan 8; Fengqm Wang. iv ACKNOWLEDGEMENTS First of all, I would like to express my deepest gratitude to my thesis adviser, Dr. William G. Lynch. His invaluable support and guidance are indispensable for the completion of my thesis. In the past years, I have learnet‘l nmch from his deep under- standing of nuclear physics on both experimental and theoretical aspects, his attitude and patience of doing physics research without omitting any details, and his strong leadership and enthusiastic stamina of completing research projects successfully. I would also like to thank Dr. S. Billinge, Dr. P. Danielewicz, Dr. C. Schmidt, and Dr. B. Sherrill for reading my dissertation and also serving on my guidance committee. I am deeply indebted to the LASSA collaboration. W'ithout the productive collab- oration of the groups at I\-Iichigan State University (MSU), Indiana University (IU) and W’ashington University (WU), the LASSA project would not have been completed in such a great. success. I would like to thank the MSU group of CK. Gelbke, T.X. Liu, X.D. Liu, WC. Lynch, I\I.B. Tsang, A. Vander Molen, G. Verde, A. Wagner, H.F. Xi and HS. Xu, the IU group of L. Beaulieu, B. Davin, Y. Larochelle, T. Lefort, RT. de Souza, R. Yanez and V.E. Viola, and the WU group of R.J. Charity and LC. Sobotka. I am also very grateful to my collaljmrators on model calculations. I own my sincere thanks to Dr. Sergio Souza and Dr. Raul Donangelo from Brazil for their great help on SMM calculations. Dr. Bao—An Li from Arkansas State University offered valuable discussions and BUU calculations for the collaboration on studying the isospin dependence of the equation of state. I would also greatly appreciate the help of Dr. Bob Charity from Washington University for incorporating his GEMINI model into our statistical calculations. In the later stage of my thesis, Dr. G. Verde kindly offered a great amount of help on the imaging analysis and model simulations V of collective effects and I owe great thanks to him. Special thanks go to Dr. Betty Tsang who has been very helpful throughout my thesis experiment and later data analysis. I have also benefited nmch from various discussions with Dr. Pawel Dai'iielewicz, Dr. Konrad Gelbke and Dr. Scott Pratt. I want to thank Dr. Aaron Calonsky, Dr. Pawel Danielewicz and Dr. Betty Tsang for their kindness of providing me letters of references. I will cherish the friendship and various helps during my stay at NSCL from l\v‘Iichael Famiano, Marc-Jan van Goethem, Paul Hosmer, Patrick Lofy, Michal Mocho. Lijun Shi, Richard Shomin, Mark Wallace, besides people 111eutioned above. I would like to acknowledge the excellent support of the staff at the National Superconducting Cyclotron Laboratory of Michigan State University. I appreciate the help from John Yurkon of the detector lab, Len Morris of the design group, Jim Vincent of electronics group, Ron Fox of the computer group and Mathias Steiner of the operations group. Last but not the least, I would thank the love and patience of my wife and my parents, and the understanding and support of my sisters and brothers. Contents LISI OF TABLES LIST OF FIGURES 1 Introduction 1.1 Freeze-out Conditions of l\Iultifragmentation .............. 1.2 Fragmentation and Isospin Effects .................... 1.3 Organization of the Thesis ........................ 2 Experimental Setup 2.1 Mechanical Setup ............................. 2.2 Electronic Scheme ............................. 2.3 Miniball/Miniwall ............................. 3 LASSA Telescopes 3.1 Silicon Detectors ............................. 3.1.1 Specifications ........................... 3.1.2 Energy Calibration ........................ 3.2 CsI(Tl) Detectors ............................. 3.2.1 Pre—selecting and Scanning CsI(Tl) Crystals .......... 3.2.2 Testing Wrapping Materials ................... 3.2.3 Other Effects and Assembly ................... 3.2.4 Energy Calibration ........................ 4 Data Reduction and Analysis 4.1 Overview of the Analysis ......................... 4.2 Pixelation Technique ........................... 4.3 CsI Crosstalk ............................... 4.4 Uniformity Correction .......................... vii xii 10 12 12 14 18 20 20 22 23 24 28 31 34 39 39 41 44 4.5 Isotope Resolutions and PID ....................... 52 5 Particle Correlations 55 5.1 Selecting the Impact Parameters ..................... 56 5.2 Two Particle Correlations and the Koonin-Pratt Formalism ...... 59 5.3 Influence of Collective Motion ...................... 69 5.4 Equilibrium Correlation Functions .................... 84 5.5 Interpretations of Correlation Functions Using the Equilibrium Corre- lation Approximation ........................... 88 5.6 Spin Determination of Particle Unstable States ............ 103 5.7 Multiple Particle Correlations ...................... 114 6 Statistical Multifragmentation Model with Empirical ModificationleO 6.1 Microcanonical Statistical Multifragmentation Model (SMM) ..... 121 6.1.1 Underlying Formalism ...................... 121 6.1.2 Temperature Distributions .................... 125 6.1.3 Effects of Temperature Variations ................ 130 6.1.4 Chemical Potentials ........................ 135 6.2 Empirically Improved Model ....................... 140 6.2.1 Ground State Energies ...................... 140 6.2.2 Internal free energy ........................ 144 6.2.3 Empirical Sequential Decay ................... 155 6.3 Model Predictions and Comparisons 158 6.3.1 Caloric Curve ........................... 158 6.3.2 Elemental and Mass Distributions ................ 162 6.3.3 Isotope Thermometry ....................... 167 7 Isospin Dependence of the EOS 177 7.1 Density Dependence of Asymmetric EOS ................ 178 7.2 Hybrid Model Calculations ........................ 181 7.3 Isotopic Composition and Isospin Dependence ............. 184 7.3.1 Relative Fme n/ p Densities and Mirror Nuclei Ratios ..... 186 7.3.2 Mirror Nuclei Ratios ....................... 190 7.4 Remarks .................................. 190 8 Summary 193 viii LIST OF REFERENCES 200 ix List of Tables 2.1 3.1 5.1 U" N) 5.3 5.4 5.5 List of the original number of detectors in a ring of I\Iinil)all/.\Iiniwall, the solid angle of a detector in that ring. and the numbers of mounted and removed detectors in that ring, respectively. ........... List of fragmentation 1’)ro(.lucts used in the energy calibration of the LASSA CsI(T 1) crystals. ......................... List of quantities are shown for the p—p correlatimis (Figs. 5.3-5.6) for inclusive and three different c.m. energy gates. The A that relevant to the fraction of protons contributed to the fitted source and the rms radius rms of the two proton source are extracted from the fitted source distribution. The X values are calculated from the integral of the source functions for r l 7 fm in Eq. 5.5. The density p/po is estimated by taking into account collective effects in Eq. 5.9 ...... List of quantities are shown for the (LG correlations (Figs. 5.12-5.13) for inclusive data and three different c.m. energy gates. The A that rel- evant to the fraction of the involved particles contributed to the fitted source and the rms radius rm, of the two proton source are extracted from the fitted source distribution. The density p/po is estimated by taking into account collective effects in Eq. 5.9. ............ Relevant spectroscopic information of 6Li, 7Li. 7Be and 8Be which is adopted from ref. [98] is listed for the correlation functions discussed in this section ................................ List of quantities are shown for the d-(i, t-(t', 3He-oz and 0-0 corre— lations (Figs. 5.16-5.19). The values of Vf/VU are obtained from the corresponding experimental correlation functions. Price is the correction factor in Eq. 5.26 calculated from secondary decay contributions in the SMM. The density p/po is estimated in Eq. 5.28 using the secondary decay correction factor Fdec from the SMM while p/p0(KP) is obtained using Fdec(K P) (see text). The corresponding pf/po and pf/ p0(K P) are calculated in Eq. 5.29 without considering excluded volumes. . . . Spectroscopic information of IOB" —+ a+6Li adopted from ref. [98]. The state at the 6.56 MeV is assigned tentatively with J=4 and only alpha decay is confirmed. The last three states contribute very little to the fitting. For details see text ..................... 17 35 66 78 93 102 112 6.1 6.2 7.1 Best fit parameters to the LD;\»I formula. of ref. [105]—[106]. All the values are given in MeV, except for k which is unitless. ........ List of isotopic thermometers with AB > 10 MeV. The left column shows the IMF thermometers involving isotopes of 3 3 Z S 8. The right column lists the He t.hermometers involving the isotope pair of SJHQ The first two columns provide the N / Z ratio and number of nucleons in the prefragments produced in the calculations for an elapsed time of 100 fm/c and density cutoff of po/ 8. The next two columns provide corresponding information for the same cutoff density but a shorter elapsed time of 80 fin/c. All calculations were performed at an impact parameter of 1 fm. ............................ xi 170 List of Figures 2.1 2.2 2.3 2.4 3.1 3.2 3.3 3.4 3.5 3.6 3.7 3.8 3.9 3.10 4.1 4.2 4.3 4.4 4.5 4.6 Schematic drawing of l\-‘Ii11ibtrll/Miniwall + LASSA setup in the vertical plane. ................................... Miniball/h'finiwall + LASSA setup housed in the 92” chan‘iber. . . . . Miniball/L‘Iiniwall + LASSA detector setup viewed inside from the beam direction. .............................. Schematic diagram of electronics setup for LASSA. .......... One of the nine identical LASSA telescopes. .............. Picture of one 500 nm thick double-sided silicon strip detector. . . . . Pulser calibration for three dynamic ranges. .............. Energy calibration for a typical silicon strip. .............. CSI crystal scanning setup with alpha source. ............. Uniformity results of accepted and rejected CsI crystals. ....... Comparisons of different wrapping materials used for CsI crystals. . . CsI preamplifiers and their housing .................... CSI packing in one telescope ........................ Calibration curves for 11C, 12C, 13C and 14C for the CsI (Tl) crystals obtained using direct fragmentation beams listed in Table 3.1. The curves are the best fit according to Eq. 3.1. .............. Flow chart of data analysis for the LASSA ................ EB-EF fine-tuning calibration for pixelation. On the left panel dis- crepancy between EB and EF calibrated from the precision pulser is shown. After the EB-EF flattening routine is applied, the good agree- ment between EB and EF is shown on the right panel. ........ layout of CsI crystals within one telescope and typical crosstalk shown for light leakage from CsI X to C51 Y in Telescope 2. Similar results for CsI Y ——i CsI X are not shown. Units are in raw channel numbers. Non-uniformity of one typical 65 pm thick silicon detector. ...... Non-uniformity of one typical 500 pm thick silicon detector. ..... PID lines of Li and C isotopes before and after uniformity correction. xii 37 40 45 47 48 49 4.7 Isotopically resolved PID lines from H to O are shmvn for 1_)articles stopped in Silicon Detectors ....................... 50 4.8 Isotopically resolved PID lines from H to O are shown for particles stopped in CsI crystals .......................... 51 4.9 One-dimensional PID lines from H to O are shown for particles stopped in Si detectors ............................... 52 4.10 One—dimensional PID lines from H to O are shown for particles stopped in C31 crystals ............................... 53 4.11 Mass resolution is illustrated from one-dimensional PID plots of Car— bon isotopes. Left panel is for Carbons stopped in the second layer of Silicon detectors. Right panel is for Carbons stopped in the CsI crystals. 54 CI! )—‘ reduced impact parameter as function of total charged particle multi— plicity. ................................... 57 5.2 Differential multiplicites of hydrogen and helium isotopes at angles of 17°, 27°, 37°, 470 and 57". ........................ 58 5.3 Inclusive p—p correlation function in central 129Xe-l—lg-fAu collisions at E/ A = 50 MeV is shown. The black points are experimental data and the grey line is the best fit. by the imaging technique. See table 5.1 for specific parameters ............................. 61 5.4 The p-p correlation functions are shown for three c.m. energy gates of 0 < Em, < 15 MeV, 15 < Em < 30 MeV, and ECm > 30 MeV. For the middle and high energy gates, the image technique is used. The simple Gaussian source parametrization in Eq. 5.6 is applied for the low energy gate. See table 5.1 for specific parameters .......... 62 5.5 The imaged source function is shown by inverting the pp correlation in Fig. 5.3. See table 5.1 for the extracted source radius. ....... 63 5.6 The source functions are shown for the p—p correlation functions in Fig. 5.4 for three c.m. energy gates of 0 < Em, < 15 MeV, 15 < Em < 30 MeV, and Em > 30 MeV. For the middle and high energy gates, the image technique is used. The simple Gaussian source parametrization in Eq. 5.6 is applied for the low energy gate. See table 5.1 for extracted source radii. ................................ 65 5.7 The inclusive d—a correlation function for the central collision gate is shown as a function of relative momentum by the black points. The solid line is the fit by the simple Gaussian source parametrization in Eq. 5.6 without corrections for collective motion. The blown-up in the top right shows the poor quality of the fit to the second peak of the d—a correlation. .............................. 68 5.8 A source with only thermal motion (open arrows) is shown on the left. The collective velocity field (solid arrows) is drawn on the right and results in a grey area of the source where emission into the right half plane is improbable. ........................... 69 xiii 5.9 5.10 5.11 5.12 5.14 5.15 5.16 The source reduction factor, fwu. obtained from the ratio of the rms radius to the system radius for the two particle sources with collective motion, are shown by the solid line as a function of the mass number. The ratio without collective motion is denoted by the dashed line. The solid line is calculated for a uniform spherical system with a radius of 11 fm while the three symbols denote the corresponczling calculations for a system with a radius of 7 fm. ................... The solid line shows a simulation of the resonant distribution as a function of relative energy for the d-alpha correlation. The dashed line denotes the simulation for non-resonant background while the dotted line demonstrates the relative energy distribution for mixed event back— ground. All the distributions are normalized to one for comparisons. . The ratios of the resonant yield over the mixed event distribution for the d-a correlation are shown in three panels for maximum, 80% and zero of the collective motion, respectively. The corresponding values of Te” are extracted from the fits (dashed lines). ........... The inclusive d-a correlation function for the central collision gate is shown as a function of relative momentum by the black points. The solid line is the best fit by the simple Gaussian source parametrization in Eq. 5.6 with Teff corrections for collective motion. The blown-up in the top right shows the good reproduction of the second peak of the d-a correlation after Te" corrections. The extracted source radius is To = 3.3 fm (see Table 5.2 for the rms value). The dashed and dotted lines are the fits assuming that the source radius is 1‘0 2 2.0 fin and r0 = 5.0 fm, respectively. ......................... The d-a correlation functions are shown for three c.m. energy gates of 0 < Ed,cm < 20 MeV and 0 < Ema" < 25 MeV, 20 < Edm, < 40 MeV and 25 < Emcm < 45 MeV, and Ed”, > 40 MeV and Emcm > 45 MeV. Fits by the simple Gaussian source parametrization in Eq. 5.6 are shown by the solid lines after Teff corrections for collective motion. See table 5.1 for extracted source radii and /\ values. ......... The ratios of the resonant yield over the mixed event distribution for the p—p correlation are shown in three panels for maximum, 50% and zero of the collective motion, respectively. The corresponding values of T8” are extracted from the fits (dashed lines). ........... The experimental d—a correlation function is shown. The smooth solid line is the empirical background used for the analysis. The dashed and dotted lines are the Coulomb correlation calculated from Eq. 5.21 for sharp sphere radii of R=12.4 and 15.6 fm, respectively. ........ The left panel shows the experimental d—a correlation function. The solid, dashed and dotted lines are the different backgrounds used for the estimation of uncertainty. The right panel exhibits the fit of the first resonance peak after subtracting the background (solid line in the left panel) .................................. xiv 77 79 80 89 92 5.17 5.18 5.19 5.21 5.22 5.23 5.24 5.25 5.26 5.27 5.28 5.29 The experimental t—a correlation function is shown in the left panel. The solid, dashed and dotted lines are the different backgrounds used for the estimation of uncertainty. The right panel exhibits the fit of the first resonance peak after subtracting the background (solid line in the left panel). .............................. The left panel shows the experimental 3He-Oz correlation function. The solid, dashed and dotted lines are the different backgrounds used for the estimation of uncertainty. The right panel exhibits the fit of the first resonance peak after subtracting the background (solid line in the left panel) ................................. The experimental a-a correlation function is shown in the left panel. The solid, dashed and dotted lines are the different backgrounds used for the estimation of uncertainty. The right panel exhibits the fit of the first resonance peak after subtracting the background (solid line in the left panel). .............................. The densities of the resonance states from d—a, t-a, 3Han and a- a correlations are fitted after subtracting the backgrounds carefully selected by the solid lines in Figures 5.16-5.19 .............. The proton—7Li correlation function is plotted. The dashed line is the selected background. See text for details ................. The density profile of the resonances of 8Be is shown in the p-7Li cor- relation after subtracting the background selected in Figure 5.21. . . . The best fit (solid line) is performed for the p—7Li correlation by vary- ing the background and the spin value of the 17.64 MeV state. The dashed line is the fitted background. Two calculations are shown as the dotted lines assuming that the spin of the 17.64 MeV state is 0 and 2, respectively and keeping the other parameters the same. ...... The p—7Be correlation function is fitted by the solid line assuming only two states at 0.774 MeV and 2.32 MeV. The dashed line is the fitted background. ................................ The p—7Be correlation function is fitted by the solid line assuming the existence of an additional state at 1.4 MeV. The dashed line is the fitted background .............................. The a-6Li correlation function is fitted by taking into account all the spectroscopic information shown in Table 5.5 except for the 6.56 MeV state. By varying the spin of the 6.56 MeV state from 1 to 4 and assuming the decay branching ratio is 100%, one obtains the dashed, solid, dotted and dot-dashed fitting lines, respectively. If a branching ratio of 55% is assumed, the solid line represents a fit of J =4. The a-a—a correlation function is shown. Resonances from the excited states of 12C are labelled with the first peak seen more clearly in the inner upright panel ............................. The p—p—oz correlation function is shown. ................ The p—a-a correlation function is plotted. The resonances from the decay of 9B are shown. .......................... XV 94 95 97 104 106 107 109 110 111 116 117 118 5.30 The d-a-a correlation function is shown. The resonance states of 10B 6.1 6.2 6.3 6.4 ' 6.5 6.6 6.7 6.8 that contribute this decay are listed. See (‘letailed discussions in text. The points denote distributions of temperatures calculated with the SMM approach for the decay of a ”2511 nucleus at three different excitation energies. The lines denote gaussian fits to the calculated distributions. ............................... The points denote temperature distributions calculated with the SMM approach for the different isotopes considered in the carbon thermome- ter for an excitation energy of E5 / A = 6MeV. The lines denote gaussian fits to the calculated distributions. ................... Comparisons of various primary temperatures Tat/C, T1,.” p and "mm from the SIV’IIVI and 72:3} from the analytical calculation in the grand canonical limit. For details see the text. One point is missing for 7103‘: with or = 0.8‘MeV because the calculated value for p for the correction term in Eq.(6.17) becomes negative at EJ/A = 3MeV, i.e. the expansion breaks down in this case .................. The solid squares and circles denote the free proton and neutron yields, respectively, calculated via the SMIVI approach. The solid and dashed lines denote fits to the calculated yields following Eq. (6.23). ..... The squares, circles and triangles denote neutron chemical potentials derived from Eq. (6.24) using SMM predictions for Carbon and Lithium isotopic yields at various initial excitation energies for the decay of the nucleus 1128a. The stars and the dot-dashed line denote approximate values calculated from Eq.(6.28) for T20 and 4.58 MeV, respectively. The error bars denote the statistical errors in the calculation, which in many cases are too small to be observed in the figure .......... Difference between the total binding energies predicted by the LDM and those recomended in ref. [104]. Plot A corresponds to the param- eter set adopted in standard SIV'IM [103], whereas Plot B is obtained using the parameters presented in this work. .............. Total binding energies for different nuclei. The full lines correspond to the corrected LDM formula, whereas the symbols represent the exper- imental data of ref. [104]. The dashed lines correspond to the predic- tions given by Eq. (6.30). For details see text. ............. Internal free energies for A = 20 (upper panel) and A = 200 (lower panel). The standard SMM expression [Eq. (6.34)] is represented by the full line whereas the dashed lines stand for the results obtained with the Taylor expansion [Eq. (6.36)]. The Free energy calculated 119 127 129 134 137 142 143 through the level density given by Eq. (6.38) is depicted by the symbols. 147 xvi 6.9 6.10 6.11 6.12 6.13 6.14 6.15 6.16 6.17 6.18 Level densities as a function of excitation energy for 2”Ne and '31P. Two energy ranges are plotted to show the behaviors of level densities at both low and high energy ends. The density of experimentally known levels is shown as bars in the low energy region. The dashed lines are the extrapolations of the empirical values according to Eq. 6.43. The dotted lines are the level density (Eq. 6.38) parametrized from the standard SMM. The solid lines are the level density adopted in this work (Eqs. 6.49-6.52). .......................... Comparison between F *(E ) calculated through Eqs. (6.35) and (6.49)- (6.52), symbols, and the approximation given by Eq. (6.53), full line. To illustrate the influence of quantum effects at low temperatures, the dashed line represents the free energy used in standard Sl\-’IM calcula- tions Eq. (6.34). For details see text. .................. Best fit values of T0 for different nuclei (symbols). The dashed line corresponding to Eq. 6.54 is used for Z > 15. ............. Caloric curves are shown for calculations of the system of A=168 and 2:75 at fixed breakup density and multiplicity-dependent density. The dotted lines are calculated from the standard Sl\-'IM. The dashed lines are calculated as empirical binding energies are taken into account. The solid lines are obtained from the improved model with empirical modifications of both binding energies and free energies. ....... Pressure curves due to kinetic motion and Coulomb interaction (see Eq. 6.58) are plotted for the system of A=168 and 2:75 at fixed breakup density and multiplicity-dependent density. The dotted lines are cal— culated from the standard SMM while the improved SMM presents the solid lines .................................. Average breakup multiplicities are shown for the system of A=168 and 2:75 at fixed breakup density and multiplicity-dependent density. The dotted lines are calculated from the standard SMM while the improved SMM presents the solid lines. ...................... Dependences of temperature on excitation energy and breakup density are shown for the system of A=168 and Z=75. Calculations as function of excitation energy at fixed density of 1 / 6 normal density are shown as solid circles. Calculations as function of density at fixed excitation energy are shown as open squares ..................... Mass and charge distributions are shown for the system of A2186 and Z=75. The dashed lines are the calculations from the standard SMM. The solid lines are calculated using the improved model. ....... Final mass and charge distributions after applying the empirical sec- ondary decay procedure discussed in Sect. 6.2.3. The dashed lines are calculated from the primary results of the standard SMM while the solid lines are from the improved model. ................ Final mass and charge distributions from the present model (solid lines) and the Botvina version (dashed lines) are shown. For reference, some measured data is plotted as solid circles. ................ xvii 155 160 161 162 163 164 166 6.19 6.20 6.21 6.22 6.23 6.24 6.25 7.1 7.2 Primary isotopic distributions are shown for Be, C. O and Ne nuclei. The dashed lines correspond to the calculations of the Botvina code while the solid lines represent the results of the improved model. . . . Isotopic distributions are shown for isotopes from Li to O. Experimen~ tal data is shown as the solid circles. The dashed lines denote the Botvina calculations and the solid lines are the final distributions after decaying the hot primary fragment via the empirical secondary decay procedure discussed in Sect. 6.2.3 ..................... Isotopic temperatures are extracted from 18 IMF thermometers (see table 6.2) with 3 3 Z S 8 and AB > 10 MeV. Experimental data is shown as the solid circles. The open squares are the calculations from the improved model. For reference, the primary temperature calculated from the present model is shown as the dashed line. .......... Isotopic temperatures are extracted from 18 IMF thermometers (see table 6.2) with 3 S Z S 8 and AB > 10 MeV. Experimental data is shown as the solid circles. The open triangles are the calculations from the Botvina model. For reference, the primary temperature calculated from the present model is shown as the dashed line. .......... Isotopic temperatures are extracted from 12 light thermometers (see table 6.2) satisfying AB > 10 MeV and involving the isotope pair of 3“‘He. Experimental data is shown as the solid circles. The open circles are the calculations from the present model without corrections of nonequilibrium emissions. For reference, the primary temperature calculated from the present model is shown as the dashed line. . . . . Isotopic temperatures are extracted from 12 light thermometers (see table 6.2) satisfying AB > 10 MeV and involving the isotope pair of 3"1He. Experimental data is shown as the solid circles. The open squares are the calculations from the present model with corrections of nonequilibrium emissions of 3He. For reference, the primary tem— perature calculated from the present model is shown as the dashed line. .................................... Isotopic temperatures are extracted from 12 light thermometers (see table 6.2) satisfying AB > 10 MeV and involving the isotope pair of 3"lHe. Experimental data is shown as the solid circles. The open circles are the calculations from the Botvina code. For reference, the primary temperature calculated from the present model is shown as the dashed line. .................................... The symmetry potential for neutrons and protons is shown for two different density dependences of asymmetry term: asy—stiff F1 and asy- soft F3 (see Eqs. (7.4)) ........................... Isotopic distributions from Li to O are shown for central collisions of 124Sn-l-mSn. The full circles are experimental data while the solid (dashed) lines denote the final (primary) calculations from the hybrid model using the density dependence asy-stiff F1 (Eq. 7.4) for the asym- metry term of the EOS. ......................... 168 169 171 172 174 175 176 180 7.3 7.4 7.5 7.6 Isotopic distributions from Li to O are shown for central collisions of 1“SIM—124811. The full circles are experimental data while the solid (dashed) lines denote the final (primary) calculations from the hybrid model using the density dependence asy-soft F3 (Eq. 7.4) for the asym- metry term of the EOS. . . . - 0. x ‘) 9 ‘ f ‘ Relative isotope ratios, R21, of two reactions l"‘SIHJASn and ll“SnJrIUSn are shown as a function of neutron number. The upper panel presents the primary calculations using the hybrid model while the final isotope ratios after secondary decay are plotted in the lower panel. The lines denote the best fits through the symbols with the same slope. Both panels: The solid circles and solid squares show values for [3,, and {3", respectively; measured in central “28114—112811, 112Sn+124$n and 124Sn+12ilSn collisions at E/A=50 MeV. Left panel: the open and cross-hatched rectangles show corresponding hybrid calculations for R21 calculated from the primary and final fragment yields, respec- tively, predicted by the hybrid calculations using the Asy-stiff EOS. Right panel: the open and cross-hatched rectangles show correspond— ing hybrid calculations for B21 calculated from the primary and final fragment yields, respectively. predicted by the hybrid calculations us- ing the Asy—soft EOS. The solid and open points in the upper and lower panels show the mir- ror nuclei ratios measured for 124Sn+mSn and 112Sn+UZSn collisions, respectively. Left panels: The open and cross-hatched rectangles show corresponding hybrid calculations of the mirror nuclei ratios calculated from the primary and final fragment yields, respectively, predicted by the hybrid calculations using the Asy-stiff EOS. Right panel: The open and cross-hatched rectangles show corresponding hybrid calculations of the mirror nuclei ratios calculated from the primary and final fragment yields, respectively, predicted by the hybrid calculations using the Asy- soft EOS. xix 185 187 189 191 Chapter 1 Introduction Nuclear matter is a strtmgly—interacting Fermi liquid [1, 2] at low temperature and is expected to undergo a. phase transititm to a nucleonic gas within a mixed phase region bounded by a critical temperature of order 15 MeV [3]-[6]. This liquid-gas phase transition is one of two bulk phase transitions in strongly interacting matter. The other is the transition between hadronic matter and the quark—gluon plasma, which is being studied at the Relativistic Heavy Ion Collider (RHIC) and elsewhere [7, 8]. The properties of mixed phase of the liquid-gas phase transition have been calcu- lated in a number of equilibrium theories [9, 10, 11]. In general, these theories predict the equilibrium distribution will be characterized by a mixture of light particles with Z s 2 and intermediate mass fragments (IMF’s) with 3 3 Z s 30. Such fragments are rarely emitted in the decay of compound nuclei at low excitation energies E*/A g 2 MeV, but become emitted with increasing multiplicity as the excitation energy is in- creased significantly beyond that approximate threshold [12, 13, 14]. These excitation energies can be easily achieved at the NSCL as well as at other intermediate energy heavy ion facilities. Experimentally there are basically three ways to generate n'iultifragmentation. One is by central collisions of two heavy ions of comparable masses at incident en- ergies of 40-100 MeV per nucleon. This is the type of experiment perforn’ied in this thesis. Dynamical calculatitms indicate that such collisions lead to dilute systems in central collisions at energies greater than E/A=35 MeV [15, 16]. From a purely dynamical point of view, multiple fragment production from the bulk system is very likely in such collisions when the density drops below and the system cross into the region of adiabatic instability, density fluctuations grow exponentially and spinodal decomposition n‘iay occur [17]. In addition to central collisions, one can also investi- gate multifragrnentation in peripheral collisions (large impact parameters) of heavy ions at incident energies of a few hundred MeV per nucleon. Such studies have been pursued by groups at GSI and Berkeley [18, 19, 20]. The third approach is by light ion induced collisions with much higher bombarding energies (> 1 GeV per nucleon) [21]. All these reactions can form sources with excitation energies that exceed the threshold predicted for statistical multifragmentation and multifragmentation has been observed for the highly excited systems produced in all three reaction scenarios. Studies of nmltifragmentation invoked the application of statistical techniques to nuclear systems with a finite number of particles. Such applications have a long tra- dition in nuclear physics that. stated with the description of highly excited compound nuclei. The properties of nuclei or nuclear matter have been well studied at their low- excited states, where the nuclear shell and collective models have been successfully used to predict the behaviors of these states. The beta, gamma, alpha and fission decay modes of these states have been well studied. Formalisms exist to calculate these decay rates, and to provide information about the parent and daughter nuclear states. As the system becomes more excited, the decay properties of individual states become more difficult to isolate. Then the average properties of groups of states or levels become the quantity that is measured. The natural choices for model descrip- tions of such average properties become the statistical formalisms by VVeisskopf or 2 Hauser and Feshbach[22. 23] and such models describe well the emission of nucleons or light clusters. In such models, the concept of nuclear temperature is introduced as a parameter that describes the ensemble average over the various collisions with slightly different incident energies, which contribute to the measured data. The validity of a thermal approximation for such finite systems is sometimes questioned by scientists that are more familiar with the thermal properties of much larger systems. In the nuclear case thern‘ial approximations are justified because of the large number of initial states that one is averagii‘ig over when one collides projectile and target nuclei under well controlled conditions. To illustrate this point. consider the formation of a compound nucleus with 150 nucleons at. an excitation energy of 2 MeV per nucleon (E* = 300 MeV). Like all Fermionic systems at moderately low temperature, the excitation energy of such a system in contact with a. thermal bath at temperature T is given by a quadratic relationship: E*/A = Tz/co. where T is the temperature in units of MeV and 60 is a parameter which for nuclei is of order 8 MeV. It is straightforward exercise in thermodynaruics to show that implies that the number of states per unit energy p(E*) is given by p(E*) oc exp (2.4(/E*/ (8A)) z exp(200). (1.1) Using the K1200 cyclotron. one can study such a system using a beam with a typical energy resolution of about 0.3% (AE* = 0.9.1! cV). Taking a standard nuclear level density formula [24] one can estimate that there are about 1083 levels in this energy window. Even the population of a small fraction of these states in the entrance channel leads to a vast ensemble of quantum states over which one is averaging. One therefore does not need the existence of an external thermal reservoir to justify the thermal limit. At the higher excitation energies characteristic of 1'1mltifragnmutation processes, the thermal limit is therefore justified provided the systems reach a state where the approximation of local thermal equilibrium may be applied. Et‘ptilibrium models simply assume that equilibrium is achieved at a given freezeout time and characterized by a freezeout density (probably 0.1-0.4,00) and temperature (typically 4-7 MeV) or excitation energy (typically E*/A 2 3 MeV). Phase transition arguments may be relevant when equilibrated system is dilute. \Vhether such dilute systems can be described as an equilibrated phase mixture of fragments (liquid) and light particles (gas) is one of the important questions that need to be addressed [13, 14]. Although some problems related to the finite size effects have been studied [25, 26], the question of how the system goes out of equilibrium and equilibrium observable are modified as fragments and nucleons decouple from each other and prepagate to the detectors is usually not addressed. For the interpretation of such fragment production in terms of a bulk disintegra- tion, it was important to assess how quickly the system disintegrates and exclude the possibility of a slow sequential series of binary fission-like decays, each of which increases the fragment multiplicity by one. Two observations rule out such decays occurring over the long time scales normally associated with fission: 1) Correlations between pairs of fragments emitted in such breakups reveal an anticorrelation in the fragment emission at low fragment-fragment relative energies due to their mutual Coulomb repulsion [27]. Strong anticorrelations are observed between every pair of fragments which implies short breakup timescales that are inconsistent with a slow sequence of binary decays, but within the range expected for a bulk multifragment disintegration [28]. 2) A11 approximately linear mass dependence of the mean frag- ment energies in the center of mass has been observed indicating a collective expansion that at incident energies of E / A=50 MeV and above exceeds values consistent with 4 a purely Coulomb induced expansion. Both observations imply lneakup time scales of the order of 100 fm/c or less. Temperatures have been obtained for such events from measurements of fragment isotopic yields [29, 30. 3'1, 32] and excited state populations [33, 34, 35]. These mea- surements are consistent with there being a plateau of the nuclear caloric curve at temperatures of about T N 5 MeV and E*/.4 a: 3-10 MeV [29]. In equilibrium sce- narios, such a plateau is theorr-étically expected [36] for multifragmentation process, reflecting the latent heat for transforn‘iing the nuclear liquid to the nucleonic vapor. Scaling laws have also been observed in the fragment elemental [37, 38] and isotopic [39] distributions. The former observation has been loosely interpreted by many au- thors using Fisher liquid drop theory[40] to extract critical parameters and the critical temperature for nuclear multifragmentation [38]. The ol,)servation of isotopic scaling laws, discussed in this thesis, provides a powerful simplification of the dependence of the fragment isotopic. distributions on the overall isospin of the system [41, 42]. Despite these interesting developments, there are many missing elements in the interpretation of fragmentation observables in terms of phase transitions. The early stages of the reaction during which energy is deposited and may or may not be equi- librated are not described by such models. Thus, practitioners have the freedom to adjust the thermal energy (or temperature) and the breakup density to improve the agreement with data. This has (.lelayed the determination of whether equilibrium bulk fragmentation models are justified and should be used instead of time dependent rate equation [43] or dynamical fragmentation [44] approaches. There have also been very few real tests of equilibrium multifragmentation models other than comparisons to fragment charge distributions and their related multiplicities [45]. Comparisons to energy spectra within equilibrium models are not very quantitative because such require the manual inclusion of collective motion induced by the earlier dynamical 5 evolution. More and better tests are possible if one can compare equilibrium calcu- lations to isotopically resolved data, excited state populations and other observables. The predictions for such observables. however, require an accurate descriptirm of the secondary decay of the excited fragments that had not been available prior to this thesis work [26, 46]. Even though most theoretical studies of multifragmentation have involved equi- librium models. it is relevant to consider the alternative approaches. More generally. theoretical models for the multifragmentation process can be divided into two cate- gories: static and time dependent. Examples of static models include multi—particle phase space models. such as the Statistical Multifragmentation Model (SMM) [36. 9] and the Berlin multifragmentation model [47]. which can incorporate specific nuclear properties more directly. Accordingly, a semi—microcanoiiical version of SMM [36] that incorporates detailed miclear structure information relevant to the population and secondary decay of the excited fragments was (levelt')1‘)ed [26. 46] and utilized as part of this thesis work. Additional static models include percolation [48] and lattice gas [49] approaches; these approach have the virtue of providing relatively simple schematic algorithms suitable for the exploration of critical phenomena in finite sys— tems. Statistical rate equatitms, which allow the descrij‘)tion of time dependent. phe- nomena within a statistical frz‘unework, have long been the standard aj")proach for calculating the decay of equilibrated compound nuclei and also provide an alternative description of nmltifragmentation [43]. An extension of these approaches to the case of fragment emission from an expanding residue has been developed by Friedman [10]. This Expanding Emitting Source (EES) model predicts a rapid emission of fragments once the density of the residue decreases below about 0.4/20, and is consistent with many basic multifragment observables [50, 51]. In its present form, however, the 6 description of multifragmentation is still rather schematic, leaving out many specific nuclear properties that influence fragu’ieut isotopic distributions and excited state populations. More complete dynamical armroaches. such as the Boltzn'iann-Uehling—Uhlenbeck transport model (BUU)[44], solve equations of motion that involve a self-consistent mean field. Solutions of the BUU equation provide the Wigner transform of the one body density matrix. Hybrid calculations that use the BUU equation to define the input excitation energies and source size for various statistical models can then be explored (e.g. [42]). The calculation of fragment yield directly via the BUU model without the hybrid approach is not feasible, however, because density fluctuations that lead to fragment formation are suppressed in the BUU equation. Therefore, alternative formalisms, such as the Stochastic Mean Field model [52] and Antisym- metrized Molecular Dynamics model (AMD)[53] have been developed to address this deficiency. These approaches, however, have not progressed to the point that accurate predictions of isotopic observables are currently possible. In the absence of a better description of dynamical fragment prmluction, we therefore test the equilibrium as- sumptions via comparisons of data to equilibrium calculations or hybrid BUU-SMM calculations, using the improved SMM we developed, which can provide quantitative comparisons to fragmentation observables. 1.1 Freeze-out Conditions of Multifragmentation Every equilibrium multifragmentation approach requires external input that. specifies two independent intensive parameters of the equilibrated system at freezeout. These two parameters could be the excitation energy per nucleon and temperature. Excita- tion energy be somewhat constrained by n'ieasurements are sensitive to the thermal energy of the system. Examples of such n'ieasurements are the energies and mass distributions of particles emitted from projectile- or target-like residues in peripheral heavy ion collisions [29]. Such information is not readily available in central colli— sions. Isotope temperatm'es may be extracted from measurements of double ratios of isotopic yields [29, 30, 31, 32] of IMF ’s using the Albergo formula[54] and excited state temperatures may be obtained from measurements of the relative populations of excited states of intern’iediate mass fragments [33, 34, 35]. Systematic measurements indicate that the values of T = 53:05 MeV may be expected in central nucleus-mtcleus collisions ten‘iperatures at incident energies of about. E / A=50 MeV. The relationship between the excitation energy and the temperature is only weakly density dependent. Both excitation energy and temperature dependent observables are sensitive to preequilibrium emission prior to freezeout. Thus, it would be ex— tremely useful to have an independent measure of the density or volume of the sys- tem, even in systems for which n'ieasurements of both quantities can be attempted. Compared to the large numbers of ten'ijj)erature measurements that have been per- formed, there are comparatively few measurements aimed at detern'iining the density. The main attempts have involved classical analyses of fragment-fragment correlation functions [55, 56] and a recent attempt involving light particles [57]. In this thesis, we explore alternative density determinations, which utilize two particle correlation functions involving light 1.)a.rticles and much lighter fragments. This work is more closely related to the intensity interferometry techniques, applied originally to the problem of measuring stellar radii by Hanbury-Brown and Twiss [58], adapted to strongly interacting particles by Koonin and others [59, 60] and utilized extensively in heavy ion reactions over a wide range of energies [61, 62]. Recently, Verde et al. [63] have revealed the importance of describing the detailed shape of such correlation functions; this work shows that the source size is related mainly 8 to the width of the correlation peak instead of the height [63]. Verde et al. also indicate how one can assess the degree to which the height of correlation functions are diminished by secondary decay effects and determine the importance of secondary decay processes [63]. Fruitful results are obtained in this dissertation by applying these ideas to mul- tifragmentation events. In addition, such particle correlation functions are extended to much heavier particles by utilizing thermal correlation function techniques. These thermal correlation function techniques allow the extraction of additional information relevant to the freezeout assumption and suggests a new technique for spin determi- nation of particle unstable states using correlation functions. 1.2 Fragmentation and Isospin Effects The study of isospin effects in heavy ion reactions is expected to be a major theme for research at the CCF and later at RIA. Recently, im’estigations reveal the existence of isoscaling laws that govern the dependence of fragment isospin distributions on the total isospin of the system [41, 39]. This isoscaling behavior is manifested in multi— fragmentation, compound nuclear decay and in strongly damped heavy ion collisions [64]. The isoscaling parameters extracted for nmltifragmeritation processes reveal the occurrence of isospin fractionation [41] whereby the gas phase manifested by the properties of the primary light particles prior to secondary decay is more enriched in neutrons than the liquid phase represented by the fragments. The degree of fractionation appears to be sensitive to the density dependence of the asymmetry term of the nuclear equation of state (EOS). The nuclear EOS is a key property of nuclear matter that. is very relevant to supernova explosions and to the structure and stability of neutron stars [65, 66, 67]. Experiments have succeeded in constraining the EOS for symmetric matter at a variety of densities. The nuclear n'ionopole and isoscalar dipole resonances[68], for example, have been performed to sample the curvature of the symmetric matter EOS near the saturaticm density ,0”. The collective flow measurements[69] place constraints on the symmt—ttric matter EOS at densities as high as 4 -— 5/20. The properties of very neutron rich systems such as neutron stars, however, can be dominated by the asymmetry term. Unfortunately, the difficulties of finding ob— servables sensitive only to the isospin effects and of creating very asymn'ietric matter under laboratory—cont.rolled conditions have left the density dependence of the asym— metry term largely tmconstrained. 111 this dissertation, we develop an improved SMM model and use it to investigate the sensitivity of the isoscaling parameters to the asymmetry term of the EOS [42]. These investigatitms suggest that the availability of high-intensity radioactive beams at the newly—upgracled coupled cyclotron facility (CCF) of Michigan State University and possibly at the proposed rare isotope accel- erator (RIA), will make the exploration of the isospin dependence of the EOS one of the more lIlt(-‘.I'(‘.Stlllg issues to be explored in coming years. 1.3 Organization of the Thesis The follmving chapters are organized as follows. In Chapter 2 experimental setup for 129Xe+m7Au at 50A MeV is described. Detailed specifications of the LASSA telescoj'x-s, the preparation of CsI scintillation detectors and energy calil')rations of the LASSA telescopes are described in Chapter 3. In Chapter 4 the data reduction and analysis procedures are described in detail for the LASSA array. Two and three particle correlations obtained in this experiments are described in Chapter 5. There, a. new technique for spin determination from particle correlations is proposed and a 10 probe to freeze-out density of lgireakup is discussed. In Chapter 6 an modified statis- tical multifragmentation model (SMM) which incorporates self-consistelitly nuclear structure information and calculates self—consist.ently the secondary decay of the ex- cited fragments is described and predictions of the model and comparisons to data are discussed. In Chapter 7 possible constraints on the isospin (l(,>pe11cle11ce of the nuclear equation of state are presented by use of the model described in Chapter 6. Chapter 8 summarizes the findings of this dissertation. 11 Chapter 2 Experimental Setup The experin‘ient of l29Xe+m7Au at 50A MeV was performed at the National Super— conducting Cyclotron Laboratory (NSCL) with the K1200 accelerator. To measure the fragmentation of heavy ion collisions with particles emitted from all directions, a large detection system with nearly full solid angle coverage is required. In this experiment, the Miniball/hIiniwall array[70] was utilized to cover most of the solid angle. However, the. Mi1iiball/Miniwall array alone can not fulfill the demanding measurements for this study due to its relatively coarse granularity and limited mass resolution. Therefore, a large area. silicon-strip/Csl detector array (LASSA)[71, 72] was constructed to provide excellent energy, angular and isotope resolution for charged particles. This experiment was one of the four consecutive experiments conducted in the first campaign with the complete nine LASSA telescopes. More details about the LASSA will be discussed in the next chapter. 2. 1 Mechanical Setup The 92” scattering chamber at the NSCL was used to house the wli(‘)le com flex (1) detector system (Miniball/l\=‘Iiniwall + LASSA) in 'acuum. The 129Xe beam with an intensity of ~ 108 particles per second was produced from the K1200 cyclotron to 12 bombard the 3 mg/cm2 thick Au target. Figure 2.1: Schematic drawing of Miniball/h’liniwall + LASSA setup in the vertical plane. A schematic drawing of the detector setup in the vertical plane is shown in Figure 2.1. The 3 mg/cm2 thick Au target was situated at the center of the Miniball array. The Miniwall array covered the forward angles 0 g 25" and provided a better granu- larity over this angular domain. Some of the forward elements of Miniball/Miniwall array were removed to insert the LASSA, which was centered at 35° and at a distance of 20 cm away from the target. The geometric acceptance of the combined apparatus was about 80% of 47r. To reduce the noise level, the preamplifier system for the LASSA was mounted along with the detectors in the vacuum chamber. Photo 2.2 shows the sideview of the detector setup with these preamplifiers mounted above the array. A cooling system was run inside the chamber to remove the heat generated by the detectors and 13 Figure 2.2: Miniball/Miniwall + LASSA setup housed in the 92" chamber. preamplifiers. A temperature monitoring system was also installed with temperature sensors attached to critical elements and read out by a FERA module to ensure that the system remained near room temperature during the experiment. Figure 2.3 shows a. photograph viewing from the beam direction. This picture allows one an inside view of the detector setup after the experiments were finished. Some Miniball detectors in the upper and backward quadrant were removed in order to Show all nine LASSA telescopes in the photo. 2.2 Electronic Scheme The LASSA detectors required the development of a high density electronic system ca- pable of processing the signals. Special CAMAC modules that contained 16 channels of shaping amplifiers and 16 channels of discriminators were developed by Washing- 14 Figure 2.3: Miniball/Miniwall + LASSA detector setup viewed inside from the beam direction. ton university and used for the LASSA readout. A schematic of our electronic setup for processing the LASSA is shown in Figure 2.4. 2.3 Miniball/Miniwall The original Miniball is a portable 47r phoswich detector array for the detection of charged particles. The array of 187 phoswich detectors covers the angular range 9° 3 91“], 3 160° with a solid angle of 89% of 47r. Each phoswich detector consists of a 0.08 mm (8mg/cm2) thick plastic scintillator foil backed by a two cm thick CsI(T1) crystal. Particles which penetrate the plastic scintillator foil and stop in the CsI(T1) crys- tal are identified by atomic number up to Z=18. Pulse shape discrimination in the 15 Trigger Fast ECL R8232C Readout Interface ’ D' T' /A\ Control 3 18C am to OR FERA g\> FERA P 1 user L.E. Disc 1: . Lin FIFO nspection : ‘ L"'Scope Pre-Amp 3‘ T.F. Amp E I 3 PH7164H ? ADC Dt t ‘ Slow Shaper 3 . 6 CC or 1 ..... , ....... , ................................... {“Specn-on Lm FIFO / L“’ Scope \ GPIB Bus for Slow Control > / Figure 2.4: Schematic diagram of electronics setup for LASSA. Csl(Tl) crystals provides isotopic resolution for H and He isotopes as well. Approx- imate energy thresholds are of the order of 1 MeV for the detection of a charged particle ”hit” in the plastic scintillator. The thresholds to particle identifications are about E,h/A=2 MeV for Z=3, Eth/Az3 MeV for Z=10, and EU,/A=4 MeV for Z=18 fragments. However, these thresholds are not very relevant to this thesis work because the h'liniball/h'liniwall array were used only for providing the multiplicity of hits. The hit multiplicity was used to gate centrality of collisions as discussed in Chapter 4. More details about the detector can be found in the reference by de Souza et al., Nucl. Instr. and Meth. A295, 109 (1990). To increase the granularity of the l\r’liniball in the forward direction, the Wash- ington University group (L. Sobotka and R. Charity) constructed a forward array of 16 128 fast plastic - CSI phoswich detectors that mates exactly to the MSU Miniball at 25". Similar to the Miniball detectors. each I\=Iiniwall phoswich detector consists of a ~ 8mg / cm2 plastic foil and a three centimeter long CsI(Tl) crystal. In the experiment 130 Miniball phoswich detectors and 80 Miniwall detectors were mounted while the other MB/ MW detectors were removed to make space for LASSA (for details see Table 2.1). The total solid angle coverage of the detectors of the combined MB/ MW apparatus that were used in the experiment corresponds to about 80% of 47r. To reduce the counting rates of electrons and gamma rays, all Miniball/Miniwall detectors were covered by a layer of 4mg/cm2 Sn—Pb foil. MB / MW Ring Detectors AQ(msr) Mounted Removed MB3 28 11.02 15 13 M134 24 22.9 11 13 M85 24 30.8 20 4 MB6 20 64.8 14 6 MB7 20 74.0 19 1 MB8 17 113.3 17 0 MBQ 14 135.1 14 0 MBIO 12 128.3 12 0 MEN 8 125.7 8 0 MW2 16 2.57 16 0 MW3 22 2.59 17 5 MW4 26 2.85 19 7 MW5 24 5.56 15 9 MVVB 24 10.64 13 11 Table 2.1: List of the original number of detectors in a ring of Miniball/Miniwall, the solid angle of a detector in that ring, and the numbers of mounted and removed detectors in that ring, respectively. 17 Chapter 3 LASSA Telescopes Since the initial measurements that demonstrated the phenomenon of multifragmen- tation, there has been a growing need to explore this phenomenon with higher reso- lution devices. h-Ieasurements of isotopic distributions, excited state populations and correlation functions are much needed to extract the relevant temperatures and den- sities of such processes. However, existing charged particle detection arrays such as the Miniball/Miniwall[70] lack the energy, angular and isotopic resolutions for such studies. The Large Area Silicon—Strip/Csl detector Array (LASSA)[71, 72] was designed and constructed to fulfill the highly demanding requirements of such studies. The LASSA consists of nine identical telescopes, each of which is composed of two silicon- strip detectors of 48 channels and four CsI crystals as shown in Figure 3.1 for one telescope. The whole array was centered at the forward angle of 35° in the experiment setup and positioned 20 cm from the target. The 3 mm pitch of the silicon-strip detectors corresponds to an angular resolution of :l:0.43° for this setup. The total array covered a solid angle of about 540 msr with the polar angle 6 ranging from 12° to 62° and the azimuthal angle (5) ranging from 24° to 156°. Figure 2.1 indicates the distances and angular coverage of LASSA. 18 Figure 3.1: One of the nine identical LASSA telescopes. 19 3. 1 Silicon Detectors Silicon detectors are widely used in nuclear experiments because of their excellent energy resolution and linear response for charged particles. Extremely thin (~ several tens of am) to relatively thick (~ 1mm) silicon strip detectors of large areas have been manufactured to provide position sensitivity for the detection of charged particles. Both layers of silicon strip detectors used in the LASSA are variants of the Micron design~W[73] and have an active area of about 5x5 cm2 ( Figure 3.2 ). / Hi _l am Figure 3.2: Picture of one 500 pm thick double-sided silicon strip detector. 3.1.1 Specifications In each LASSA telescope, the front (labelled here as 7’ DE”) silicon detector is about 65 pm thick and 50x50 mm2 in area; it has 16 strips on the front side that are 3 mm wide each and has a single electrode (no strips) on the back. The second layer of silicon is 20 about 500 pm thick and is also 50x50 mm'“) in area. It is double-sided: Each side is divided into 16 strips with 3 mm pitch. For convenience, we label 16 strips of the front side by ”EF” and the strips on the back side, which are. perpendicular to the front strips, by ”EB”. For both types of silicon detectors, there is a 0.1 mm wide interstrip region without electrodes and its effects will be discussed in the data analysis later. The thicknesses of the front layer of silicon can vary up to 10% and the back layer can vary by up to 2%. These variations will affect the particle identifications in ways that will be elucidated in the next chapter. The closely packed design required the development of a highly flexible fiat printed circuit board cable connecting the silicon strip detectors with the pre-amplifier hous- ings. These cables were wire—bonded directly to the silicon wafer (see Figure 3.2). All 432 preamplifiers were housed in 9 boxes, each one housing 16 preamplifiers for the DE, 16 preamplifiers for the EF and 16 preamplifiers for the EB of one telescope. Each one of the preamplifiers is driven by :l:12V supply voltage and has a power consumption of approximately 300 111W. The gains of the DE preamplifiers are 15 mV/MeV and the gains of the EF and EB preamplifiers are 5 mV/MeV. The silicon detectors need to be fully depleted to achieve the best energy reso- lution. Interestingly we found that the full depletion voltage listed by Micron was typically a factor of two smaller. The DE and EF strips were biased in the experiment to be fully depleted. The stopping energy for alpha particles is 10 MeV in the first layer of silicon and 36 MeV in the second layer after passing through the first. layer. Because any particle stopped in the first layer of silicon can not. be identified by this device through the AER identification technique, the stopping energy of the first silicon sets the lower threshold for the corresponding particle we can analyze. Software thresholds were set 21 on the analysis of the experimental data, which were a little higher than the physical stopping energies. For example. the low energy cutoff for alpha particles is about 10 MeV for the LASSA. 3.1.2 Energy Calibration One advantagemis feature of silicon ('letectors is their linear and largely particle— independent energy response. \Vhile fission fragments have in general, rather different energy calibrations than alpha particles due to space charge effects near the stopped fragments, only one energy calibration should be needed in principle for any species of charged particles with relatively low charge number Z s 8. As few as two points should be able to define the energy calibration curve for such species. However, due to the nonlinearity of electronics system including preamplifiers, shapers and ADCS at. low and high energies, the calibrations of both low and high energy particles must be treated more carefully than those of particles at the center of the energy spectrum. An Ortec precision pulser generator was used to calibrate the silicon detectors. The pulser has a group of attenuation switches to calibrate detectors with different dynamic ranges. Three different dynamic ranges of 140 MeV, 200 MeV, and 500 MeV corresponding to three different combinations of pulser attenuation switch settings were used to calibrate LASSA silicon detectors. An absolute calibration was obtained from the measurements of 2‘”Am and 2”Th alpha sources for these three settings. Figure 3.3 shows the linear calibration relations of energy in MeV versus pulser dial value in Volt for the three dynamic ranges. Right after the data—taking experiment was finished, all the electronics setup was retained the same except that the pulser calibration system was plugged in each input channel of the silicon detector preamps while the detectors were still attached. Then the one-by-one pulser calibration was carried out for all 432 silicon channels. For each 22 161-“ -———-- y = 49.702x + 0.0286 If y = 20.726x + 0.0172 12- / y =13.935x + 0.0148 Energy (MeV) 03 \ O 140 MeV Range I 200 MeV Range A 500 MeV Range — Linear (140 MeV Range) Linear (200 MeV Range) --- Linear (500 MeV Range) 0 I I I r I 0 0.2 0.4 0.6 0.8 1 1.2 Dial Value (Volt) .1“--- _. __ T. _ _ Figure 3.3: Pulser calibration for three dynamic. ranges. channel, more calibration points were applied to the low energies and high energies to measure the possible nonlinearities on those regions as shown in Figure 3.4. In the low and middle energy regions labelled as 1 and 2, respectively, in Figure 3.4, linear fitting procedures were employed while in the high energy end (labelled as 3) a 4th order polynomial fit was adopted. Spline interpolaticm was used to join the three pieces together smoothly. 3.2 CsI(Tl) Detectors Csl(Tl) scintillation crystals are cost effective for detecting charged particles with energies of E/Az30—200 MeV. Less expensive than silicon detectors, less hygroscopic than Nal crystals, and easily machined into different. shapes, CsI(Tl) detectors have been used to stop higher energy particles in many AE-E type detection arrays with large solid angle coverage. 23 .3 O ,: 4 3 2 ) 03x +b3x +c3x +d3jc+e3 fill -.'l‘ V M Spline fit I = PM + bra i 1 1 l l Pulser Dial Value (Volt) 0 A N 00 £> 01 O) \l 03 CD it 0 500 1000 1500 2000 2500 3000 3500 4000 Channel Number F io'ure 3.4: Energy calilnation for a tv )ical silicon stri ). b C). v The Csl crystals had the original dimensions of 3.5x3.5x6.0 c1113. The final shape of Csl crystals is defined by tapering the two outer sides and making the front surface reduced to dimensions of 2.5x2.5 cm2 while keeping the dimensions of the back surface the same. Four crystals are grouped together and placed behind the two layers of silicon detectors. With a thickness of 6 cm, all CsI(Tl) can stop alpha particles with energies up to 580 MeV and protons with energies up to 145 MeV. 3.2.1 Pre-selecting and Scanning CsI(Tl) Crystals To ensure that. Csl crystals in LASSA have good energy resolution. only the crystals with light output uniformity better than 100 were adopted. A careful alpha—scanning procedure was applied to control the quality selection. 24 L Amp MCA COMPUTER CsI Box with Preamp Vacuum chamber Sliding Scales Figure 3.5: Csl crystal scanning setup with alpha source. 25 The CsI(Tl) crystals uianufacturetl by Scionix[74] were originally rectangular in shape with dimmisions of 3.5x3.5x6.0 curl. The front square surface was polished and the other sides were sanded when obtained commercially. After a careful inspection for visual cracks or imperfections, the back square surface was sanded down and polished. It was then optically coupled to a clear acrylic light guide that was in turn optically connected to a 2x2 cm2 photo—diode. The front surface of the crystal was covered with an aluminized mylar foil and the other sides were wrapped with two layers of teflon tape to ensure uniform light collection. Details of discussion on wrapping materials will be shown in the next section. The whole detector set was contained in a small aluminum box along with a charge sensitive preamplifier attached in the back of the box to maximize the signal to noise ratio. The 5.486 MeV alpha. particles from a collimated 1 uCi Q'HAm alpha source were used to scan the crystals. Figure 3.5 shows the scanning setup inside a vacuum chamber. A two dimensional sliding scale supported the alpha source and the crystal box on two perpendicularly crossing arms, res1')ectively, where the scanned surface position of the crystal was measured by the scale. For each crystal nine regions that uniformly divided the front surface were in turn irradiated by the alpha source as shown in Figure 3.5. The first point was scanned again after finishing all nine points to make sure that the scanning system was stable and not affected by temperature changes and other drifting effects. The alpha spectra corresponding to the centers of the nine sub-squares were recordI—xl with a. multi—channel analyzer equipped with a. peak sensing ADC and then stored in computer for offline analysis as shown in the electronics scheme of Figure. 3.5. 26 -0.03 +0.03 +0.14 -0.09 -0.01 +0.05 0.0 0.11 -0.04 +0.05 Crystal 652 -1.0 +1.0 +0.95 -0.05 -0.92 +0.88 -0.01 -0.87 0.0 +0.80 +0.14 091 Crystal 291 -1.0 Figure 3.6: Uniformity results of accepted and rejected CsI crystals. 27 The different grey levels in Figure 3.6 represent the percentage deviations of the nine alpha peaks from the median value. The specific numbers of these deviations are shown in the corresponding table next. to the shaded crystal surface. Only the crystals with light output varyii‘ig within i051: were accepted. such as the one shown in the top panel with deviations ranging from -0.11% to 0.05% of the mean. On the other hand, crystals with light output uniformity beyond i0.5% were rejected and returned to the manufacturer. The bottom panel demonstrates one of the rejected samples, which clearly shows the existence of a light output gradient that is caused by the non-uniformity of T1 doping inside the crystal. 3.2.2 Testing Wrapping Materials A reflective entrance foil is needed to cover the front surface of a Csl crystal for optimal light collection efficiency that improves the energy resolution. In order to minimize the dead layer loss due to the Csl front foil, a very thin layer of aluminized mylar foil (0.15 mg/cm2 mylar + 0.02 ing/cm2 Al) was applied to the C81 front surface. As for the other sides of the crystals. opaque materials are needed for diffuse light reflection. This side wrapljnng is very critical not only for maximizing the light collection efficiency but also for minimizing the possible light crosstalk between four bundled crystals within one telescope. To achieve the best resolution and the thinnest wrapping, we tested different materials for wrapping up crystals. Two different materials, 0.1 mm thick white teflon tape and 0.14 mm thick white cellulose nitrate membrane with pore size of 0.2 nm are shown in fig 3.7 for different numbers of wrapping layers. 28 900 WWHW m Ot SOUI'CG If 0. I l l O ”5800: / 13700 ;— ....... O _ ........... 9"" CL. : /® 600 T ________ we” . per r , bt+®f l++k+l+s++lafi 500 _ q I OTeflonITape I t - 9 Cellulose Membrane .\7 Q ' l 0 FWHM(%) 0 O I I Y \ l \ \ \ \ \ \ \ \ \ Number of Layers Figure 3.7: Comparisons of different wrapping materials used for Csl crystals. 29 Using the same 2"”Am alpha source as used in the scanning 1.)rocedure, we found that the light output, i.e., the peak channel of alpha spectrum, increases as more layers are wrapped around the crystal for both materials. This is shown in the upper panel of Figure 3.7. The improvement of the percentage of energy resolution as shown in the lower panel is directly related to the enhanced light output since the FW’HM of the alpha peaks dictated by the electronic noise is constant at about 40 channels (corresponding to 250 keV for the largest light output). When the number of layers is large enough, for example, two for cellulose membrane and five for teflon tape, however, the collected light output saturates and nothing can be gained by increasing the thickness further. The overall performance of cellulose nitrate membrane is much better than that of teflon tape. \Vhen the same number of layers are used, the cellulose nitrate membrane shows about 40 percent increase of light collection over the teflon tape, making it certainly the choice. Since four crystals are grouped together behind the two silicon detectors in each telescope, a. large gap between CsI(Tl) crystals leads to a loss in solid angle. This is also argues for the cellulose membrane. Based on the saturation shown in Fig. 3.7, two layers of the cellulose membrane is the optimal choice. In the final wrapping, each crystal was wrapped with two layers of cellulose nitrate Inen‘ibrane around the two outer sides and one. layer around the two inner sides which touches two other adjacent crystals. Counting both shared layers of the adjacent crystals, each crystal had effectively two layers of cellulose nitrate membrane on all four sides. This was enough for measuring the light collection and resolution but not, enough to remove all cross—talk as we discovered later. 30 3.2.3 Other Effects and Assembly The light. output of Csl crystals is also affected by temperature changes. During the preselection scanning, we also tested the temperature effect which typically gives a 1% decrease in light output for every three degrees centigrade increase at room tempera- ture. In Xe+Au experiment, a cooling plate was attached right behind each telescope to stabilize the temperature within one degree centigrade at room temperature. So the variation of temperature would not be the limiting factor of energy resolution of the Csl detectors. Optical clear front and back surfaces are essential for collecting the light and moni- toring the gluing process. A step by step sanding and polishing procedure was applied by using 400 grit sanding paper and 30 ,um, 15 pm, 5 am, and 1 am polishing paper with the aid of pure ethanol (any water-containing resolvent could easily damage the Csl crystals). The final step with silk cloth and polishing compound would make the surfaces crystal clear. In addition, four other sides were slightly sanded with 400 grit sanding paper with the sanding grooves parallel to the long axis of each crystal for a better uniformity and light diffusion. For each crystal, a 1x3.5x3.5 cm3 acrylic light guide was glued to the back surface with optical epoxy BC600[75] and a photodiode with active area of 2x2 cm2 was glued with clear silicone rubber compound RTV615[76]. The gluing process was monitored to ensure that no bubbles were left in the layer of glue. To prevent the light leak and crosstalk between adjacent crystals, the outer sides of light guide and the photodiode were painted with a reflective white paint BC620[75]. Four crystals were bundled in the back of silicon detectors for stopping high energy particles. Csl preamplifiers are attached right behind the crystals within the telescope box. Figure 3.8 shows the set of four Csl preamps and their housing which is divided 31 Figure 3.8: CsI preamplifiers and their housing. and shielded with copper into four small cells to eliminate the crosstalk between the preamps. The closely packed four crystal setup in one telescope box is shown in Figure 3.9. The two layers of 5x5 cm2 silicon detectors and their frame sit on top of the box in the final assembly with the flexible signal cables running inside the metal box next to the wall. The silicon preamps are placed in separate boxes because there are 48 silicon channels per telescope. 32 ,\.i.um. walnut—u. but.» Figure 3.9: Csl packing in one telescope. 33 3.2.4 Energy Calibration The fluorescent light emitted by the CsI(Tl) scintillator has two major components of a fast (500 ns) and a slow (7 ,us) decay time ccmstants. Both components have a relationship of light output and energy that is mass and charge dependent. This property has been exploited to provide mass identification for light ions using pulse- shape discrimination. This pulse-shape discrimination capability is not needed in the AE-E type LASSA telescopes where silicon detectors are used as AE detectors and Csl crystals as the stopping detectors. However, the mass dependence of pulse shape remains important because of its influence on the energy calibration. At low energy, the light output of a Csl crystal shows a non—linear response to the deposited energy, especially for heavy ions. It also depends on the T1 doping of CsI crystals. To determine the energy calibration for different ions, the detectors were directly exposed to low intensity (~ 103 particles per second) beams of different isotopes and energies. These ions were obtained by fragmenting 2160 MeV 3“Ar and 960 MeV 160 primary beams from the NSCL K1200 cyclotron in the A1200 fragment separator. The main advantage of this method is the a failability of a large number species of particles that could be detected simultaneously (up to 52 isotopes were identified in the case of the 36Ar fragmentation). Since particles are selected only by their magnetic rigidity (8,0 = 1.841 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 FVVHM 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 Csl crystals in the present work are listed in Table 3.1. Hydrogen and helium isotopes were also calibrated by elastic scattering of E/A=3O MeV p-4He molecular beams on a Au 34 target and by 240 MeV direct 4He beam particles. 16O fragmentation E (MeV) 36Ar fragmentation E (MeV) p 77.17 (I 39.78 (1 79.57 t 26.72 t 53.75 3He 105.00 3He 210.00 4He 79.99 4He 160.00 6He 53.64 6He 107.90 6Li 119.90 6Li 240.00 7Li 103.10 7Li 206.80 8Li 90.40 8Li 181.60 7Be 182.20 7Be 363.40 9Be 142.50 9Be 285.60 10Be 128.40 lOBe 257.90 10B 199.90 10B 400.00 11B 182.10 11B 364.90 128 335.40 11C 261.20 11C 521.60 12C 239.90 12C 480.00 13C 221.80 13C 444.40 1“C 413.70 14N 279.90 14N 560.00 15N 524.00 16N 492.40 150 340.80 150 680.70 16C) 640.00 170 603.70 180 571.30 Table 3.1: List of fragmentation products used in the energy calibration of the LASSA CsI(Tl) crystals. The energy calibration for each isotope was done following the mass and charge dependence of the light output described in ref. [77], which in turn was based on previous studies of the light emission of Csl-crystals and on semi-empirical model proposed by Birks [78]. In this approach, the incident particle energy E is parameter- ized as a function of the light output L, the charge Z, and the mass A of the particle, 35 ‘dS E(L, A, Z) = aAZ‘ZL + b(1+ (-AZ'~’)L1—"V’EZ (3.1) where a, b, c and d are the fitting parameters with values greater than zero. This expression describes a linear term, dominating at high energies and a non—linear term dominating at low energies. In Fig. 3.10, the solid and dashed lines represent the best fit of Eq. 3.1 to the experin'iental energy calibraticm 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 llC isotopes does not pass through the l'1igh-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 other Csl(Tl) calibrations adopted elsewhere in the literature. Instead we adopt another solution which assumes a mass dependent calibratitm (closely related to the quenching effect)[77]. To confirm this mass dependent ansatz that could allow the constructioin of the full calibration curve for each isotope. several fragn'ientation beams of different incident energies would be required. For light charged particles with 233, the parameterization described in Eq. 3.1 did not accurately describe the detected energies. Compared to the observation of Ref. [77], a less pronounced isotopic effect was observed for light ions. This may be the result of the increased concentration of the activator element, T1, in the LASSA Csl crystals compared to those studied in ref. [77]. We find that the AZ 2 factor in Eq. 3.1 overestimates the mass dependence. we therefore employ a modified function of Eq. 3.1 with a weaker dependence on A for the particles with 233. The expression is modified for each element. For Lithium (Z=3) particles, we change the first term 36 600 , , ... - . . , l ' , ’4; _ g f //.. _ I/ /.~///// _ 12 C (7:). ,// ’ _ [3” / 1 N f~4oo—- 14 cvy/ C a l?) P C /'/‘* / (Jig/I 1:) i \2/ {27?1/ )C 200 — 12 / — : ‘C/Tl / C ’ I _/ 113 ,s. I ‘\ l l l O 1 "i I i l i i i 1 O 1000 2000 Light(a.u.) Figure 3.10: Calil‘n‘ation curves for 11C, 12C, 13C and MC for the Csl (Tl) crystals obtained using direct fragmentation beams listed in Table 3.1. The curves are the best fit according to Eq. 3.1. of Eq. 3.1 and used, E(L. A, Z) = (MEL + b(1+ (:A22)L1~dm (3.2) For Helium (2:2) isotopes, we use E(L, He) = 0L + bAC(1— a”) (3.3) The variables a, b, c, and d in Eq. 3.1-3.3 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 Equations 3.1-3.3. Our fitting procedure resulted in a precision of the energy calibration better than 2% for isotopes from He to O. 37 As we have only limited calibration points for p, d and t. two calibration points from each isotoi‘w, we adopt the simple linear function for 221 particles. E(L,H) = 0L + b (3.4) where a. and b are fit 1')arameters. A linear CsI(Tl) response is consistent with that observed for hydrogen isotopes by Handzy [79]. 38 Chapter 4 Data Reduction and Analysis 4.1 Overview of the Analysis The detection system is very complex with 432 silicon and 36 Csl channels from LASSA plus hundreds of detectors from Miniball/l\-'Ii1‘1iwall. To reduce the complexity in the analysis, we decided to use the MB/MW array to determine the charged particle multiplicity as the impact parameter filter so that detailed energy calibration of the MB / MW system could be avoided for this experiment. N evertl'ieless, we did separate the light particles (LP, ZS?) from the heavy fragments (223) in the MB/ MW in order to obtain separately the IMF multiplicity in addition to the total multiplicity. Detailed calibrations and analysis on the LASSA, 011 the other hand, were con- ducted for isotopically resolved charged particles. Figure 4.1 shows the flow chart of data analysis for the LASSA. After the raw data from the ADC was decoded and all the zero’s and pulser events were suppressed, presorted data tapes were generated for further analysis. The whole analysis procedure depends on the pixelation subroutine which associates specific x and y strips (BF and EB) with specific particles to locate where they hit. This procedure also allows one to associate the particle with specific data in the proper DE strip and the Csl crystal. The silicon energy calibration, one of the two main analysis tasks, was done using the information obtained with the 39 LASSA Calibration Flow Chart Raw Data Tapes: [ Pedestal offset? for Csl [ Csl crosstalk ] corrections [Gain-Match for Csl] Presort Tapes I 7 Csl PID] / \ Final PID I \ Pedestal Offset for Si [ Si Pulser Calib (DE, EF, EB) ’ V Flatten EB-EF] [PfielatiogbLow channel calib \ [Si uniformity] correction v Si PID EB associatiog ._/ DE, E energy, angle l [Csl Energy Calib] \ Physics Tapes [ (Energy, angle, PID, marker) ] Figure 4.1: Flow chart of data analysis for the LASSA. 40 precision pulser (for more details see the previous chapter). The energy calibration of Csl crystals, on the other hand. was conducted using the inforn'iation from a series of beam calibration runs and from measurements of proton recoils during the clam- taking runs. After making the corrections for the thickness variations in the silicon detectors, particle identification (PID) gates were constructed. In the final physics tapes, the particle ID value, the total kinetic energy, and the emission angle for each particle were recorded as well as a marker to label the circumstances when the parti- cle \, ’as identified. In the following sections, several aspects of this analysis procedure are discussed in greater detail. 4.2 Pixelation Technique One of the most important advantages of Silicon-strip detectors is their position sensitivity. Since the second layer of Silicon in the LASSA array is double—sided, we can use this to obtain (x,y) pixelwise position resolution that tells us which pixel(3x3 mn12) a particle strikes. This is simple, in principle, if only one single particle hits the silicon in one telescope during one event. When more than one particle hits the silicon, i.e., in multiple-hit cases, however, one must use care in pairing up the signals from the vertical and horizontal silicon strips in order to obtain the correct position information for each particle. This pairing is done in the ”pixelation procedure” described below. \Vhen there are two particles or more hitting into one telescope, there can be ambiguities in assigning the position of each particle, which need to be addressed. First, one need to determine whether the particles can be identified or not. For example, if two charged particles punch through two layers of silicon detectors and stop in the same Csl crystal, then it is not feasible to obtain the correct particle ID 41 and energies for these particles. Second, the signals from DE, EF, EB and C‘sI need to be associated properly with specific particles for the analyses to be correct. When a particle is stopped in the first layer of single-sided silicon detector its proper ID (mass and charge) and position can not be determined; such particles have to be disregarded. For the other cases (eg. particles stopped in the second silicon or the C31), signals from the second double-sided silicon detector can provide the best choice for the determination of particle multiplicity and position since the perpendicular EF and EB strips granulate the detector into fine pixels and have basically the same energy signal for the same particle when the calibrations for both EF and EB are accurate. We use this to pair up the signals of EF and EB and identify the particles. 5.0 9 2-5 7.: (l) 2. LL 0 ”a w: 53 -2.5 _5' 7. ‘ I L l . l 1 l 0 40 80 I20 40 80 120 EB (MeV) EB (MeV) Figure 4.2: EB-EF fine-tuning calibration for pixelation. On the left panel discrep- ancy between EB and EF calibrated from the precision pulser is shown. After the EB-EF flattening routine is applied. the good agreement between EB and EF is shown on the right panel. In order to make the best use of the pairing process. the energies of EF and EB should be calibrated in the same way. However, the independent silicon calibration of EF and EB strips can only offer a precision of 1% which results in larger deviations 42 of EF from EB signals at higher energies and consccpiently a difficulty in resolving two hits with similar energies. A EF-EB flattening routine was developed to enforce that the energy signals from both EF and EB for the same particle are the same. Figure 4.2 shows how the spread of EB-EF difference changes before and after the flattening routine. On the left panel the difference between EF and EB signals before the flattening procedure is plotted. One can see one intense, nearly horizontal contour, a line that slopes downward, and a grey background. The grey background results when the EB and EF signals are from different 1')articles. The crooked, nearly 11(1)rizontal line corresponds to the correlation between EB and EF when both are calibrated properly with the alpha-sources and pulser as described in Sect. 3.1.2. The steeper line corresponds to the case that particles pass through a EB or EF strip that has problems with its ralibration. After performing this procedure, one straightens out the correlatitm between EF and EB as shown in the right panel of Figure 4.2 and recovers these ill—calibrated silicon strips. A gate of $0.5 MeV (shown by the dashed lines in Figure 4.2) is applied to select the correct pairing of EF and EB signals. For signals that can not be paired up with any EB or EF strip within the gate of :t().5 MeV, we need to distinguish cases where two I.)articles hit in the same silicon strip (double—hit, which occurs about 3% of the time) and cases where the charge of one particle is split between two neighboring strips (split-hit, which occurs about 5% of the time). Specifically speaking, if the magnitude of one EB signal is close to the sum of two neighboring EF signals, then with a high probability this case can be identified as a single particle hitting into the 0.1mm wide interstrip gap with its signal split into two adjacent channels; if those two EF channels are separated, then one knows that one has the double-hit situation where two particles have gone into the same EB strip. 43 Due to limited dynamic range of the electronics the signals of some particles, esl')ecially heavy particles, are saturated in a small fraction (< 1%) of all the signals, which need to be excluded in the data analysis since the pairing technique and energy determination become obscure for these particles. Such is the case for the particles hitting into strips that give no output signals because either their electronics has failed or the detectors or the wire bounds to the strips are had. 4.3 CsI Crosstalk During this experiment with the LASSA device, it was discovered that a small frac- tion (typically < 2%) of the light emitted in one crystal could leak into an adjacent crystal. After this experiment, a thin layer of alumiuized mylar (0.15 ing/cm2 mylar + 0.02 ing/cm2 aluminum) was inserted between adjacent. crystals. This additional foil decoupled the detectors completely. Since this decoupling was achieved after this dissertation experiment was performed, the present analysis must deal with and remove the effects of the light leak as discussed below. Four Csl crystals are closely packed together within one telescope as indicated by the sketch at the top of Figure 4.3. Figure 4.3 also shows the correlations between the light output observed in pairs of CsI crystals within one typical telescope (the reversed crosstalk is similar and not shown). Note that the Y scale is blown up for showing the light output correlation between CsI X and Y when a signal is observed in Csl X. It is clearly seen that the crosstalk is about the level of 1-2% between two directly neighboring crystals and basically vanishes between diagonal crystals. In the cases of two diagonal crystals (two middle plots in the figure), small slightly tilted lines close to Y axis are just the correlation of two crosstalk signals when one of their common neighbors has been fired. 44 Csl2 Csl4 Csl] Csl3 Csl] —>Csl2 Csl] —>Csl3 Csl] —>Csl4 Csl2—>Csl3 Figure 4.3: layout of CsI crystals within one telescope and typical crosstalk shown for light leakage from 031 X to Csl Y in Telescope 2. Similar results for CsI Y —> GSI X are not shown. Units are in raw channel numbers. 45 During the experiment. it was determined that a linear crosstalk correction could be applied. Therefore the decoupling foil was not inserted between Csl crystals in an effort to conserve. data taking time. This linear assumption is demonstrated by Figure 4.3. In order to correct the light output in each crystal a conversion matrix is constructml as follows: L, 2 142+ 2((10-‘(10-(112— L?J)) — 2((11,(L3 — L3») (4.1) J?“ J¢i where the L' is the channel number of apparent light output; L is the channel number of reconstructed light. output after correcting the crosstalk between CsI crystals; (1, is defined as the crosstalk matrix; Q is the gain factor due to the gain differences between crystals; L0 is the channel number of the offset. The second term is the light gain from other crystals while the last term stands for the light loss to adjacent crystals. The matrix elements of a and L0 can be easily obtained from the slope and offset parameters of the crosstalk lines in Figure 4.3. After the reconstruction of Csl light output, calibration of individual Csl crystals were carried out according to the procedure described in Sect.3.4 4.4 Uniformity Correction Before starting the particle identification (PID) through the DE-E technique, one has to correct the non-uniformity of Silicon thicknesses. As mentioned in Sect. 3.1.1 Large area thin silicon detectors have a disadvantage of non—uniform thickness up to 10%. Different energy losses are in general recorded for identical particles of the same energy passing through different pixels. To correct for this thickness variation, it becomes essential to apply the pixelation procedure described in Sect. 4.2 and use it to map the thickness variation. 46 Figure 4.4: Non-uniformity of one typical 65 pm thick silicon detector. 47 m:_m> cmoE 9 mmmcxoé @259 Figure 4.5: Non-uniformity of one typical 500 pm thick silicon detector. 48 500 Li Before Correction Li After Correction 250 *- 12 c: :3 O 8 , 6 7 8 9 6 7 8 9 10 i C Before Correction 0 After Correction l IOU— l 0 l l l l l l l l 4 i ll 12 13 14 15 lb ll l2 I3 14 15 It: A Figure 4.6: PID lines of Li and C isotopes before and after uniformity correction. The relative thickness of silicon detectors was determined from the direct beam calibration experiment ( which is also part of calibration for CSI crystals, see Sect. 3.2.4). The direct beam with fixed high energy was applied to punch through the silicon detectors. The correction matrix can then be constructed from the measured energy loss in each pixel. Assuming a linear relation between energy loss and thickness the thickness variation can be obtained according to the energy loss in each pixel. Figure 4.4 shows an example of the non-uniformity of the 65 um silicon detectors. On the other hand, the 500 um silicon detectors have much better uniformity as illustrated in Figure 4.5, suggesting that the absolute thickness variations of both types of detectors are comparable and the uniformity scales merely with the overall thickness. 49 DE with thickness correction (MeV) EB (MeV) Figure 4.7: Isotopically resolved PID lines from H to O are shown for particles stopped in Silicon Detectors After both layers of the silicon detectors have their thickness variation determined, the effects of thickness variations on particle identification (PID) can be removed so that PID gates can be imposed independent of hitting position. As showed in Figure 4.6 the PID lines of Lithium and Carbon isotopes are blurred before uniformity correction applied because the same particles don’t fall in the same PID line when they go through different paths. On the right panel well defined sharp PID lines are achieved, indicating the importance of making an accurate uniformity correction. 50 EB with Thickness correction (MeV) 40 " "o 7““? 24 Ecsi (MeV for 4H9) Figure 4.8: Isotopically resolved PID lines from H to O are shown for particles stopped in C31 crystals 51 E He Particles stopped in Si 10‘1 =l-l Li Be 3 3 U 10 C N a l o 102 U 10 | , | | l l l l 500 1000 1500 2000 2500 3000 PID (Arbitrary Units) Figure 4.9: One—dimensional PID lines from H to O are shown for particles stopped in Si detectors 4.5 Isotope Resolutions and PID Two sets of PIDs can be constructed for charged particles stopped either in the second layer of silicon or in the Csl crystal. Figure 4.7 shows the DEcorr (energy loss in DE with uniformity correction) vs. E (energy loss in Si2) for particles stopped in Si2. And similar plot in Figure 4.8 is shown for particles stopped in Csl. Different PID lines are well resolved in both plots for nuclei up to Oxygen and resolution could be achieved for heavier particles if the data were not limited by statistics and detector dynamic range. However, these curved lines are not very convenient for identifying particles and further analysis. One solution is to straighten them out and make one-dimensional PID plot for easy-handling. To preserve the density of particle distribution on the DE—E plots, the flattened PID lines are drawn according to the intervals normalized 52 H He Particles stopped in Csl 105 g) 5W Li H)4 g] Be B C i N % i03r S O l 102 l l to ; l 1iii i ll 1 500 l 000 l 500 2000 2500 3000 PID (Arbitrary Units) Figure 4.10: One—dimensional PID lines from H to O are shown for particles stopped in Csl crystals along the diagonal direction of the DE-E plots. Simple one-dimensional PID spectrum can therefore be extracted as in Figures 4.9 and 4.10. Specifically for carbon isotopes the mass resolution can be seen in Figure 4.11 for carbons stopped in Silicon detectors (left panel) and in Csl crystals (right panel). 53 500 400 *3 300 » 3 O O 200 100 0M1 l l l 10 12 14 16 Mass ‘l l l A . 10 12 14 16 Mass 1 250 1000 750 500 250 Figure 4.11: Mass resolution is illustrated from one-(liniensional PID plots of Carbon isotopes. Left panel is for Carbons stopped in the second layer of Silicon detectors. Right panel is for Carbons stopped in the Csl crystals. 54 Chapter 5 Particle Correlations Historically intensity interferometry via particle correlations was first studied in as- trophysics. In the 1950’s, Hanbury—Brown and Twiss applied this technique (now commonly called HBT) for measuring the diameters of distant astronomical objects by examining photon correlations[58]. This idea has later been generalized to cor- relations in nuclear physics involving various types of particles[61, 62]. The early examples such as pion-pio1'1[80] and proton—proton[59] correlations involve identical bosons and fermions, respectively. These have been widely used for studying the properties of the sources of particles emitted in heavy ion reactions. Subsequently, non-identical particle correlations such as d-alpha correlations and correlations in- volving heavier fragments (up to Carbon) have also been studied; these studies have provided insight regarding the freeze-out conditions for multifragmentation processes after which interactions effectively cease [34, 81, 82]. Among such studies are the determinations of the populations of excited states of emitted fragments [83]. Con- siderable knowledge about the temperatures at the time of fragment emission has been gained and the hypotheses of thermal equilibrium for the emission mechanism has been tested in these studies[34]. In this chapter, we investigate the correlations of various charged particles emit- 55 ted during central 129Xe+m7Au collisions at E/ A = 50 MeV. The. data presented here were measured with the LASSA array. We begin by discussing how to select the impact parameters included in our analyses. Then we introduce the basic corre— lation function obser tables. The correlation functions are interpreted via within the Koonin-Pratt formalism [59, (50] and within the assumption of thermal equilibrium. After discussing some of the strengths and weaknesses of the two approaches and their interrelationship, we explore the data for central l29Xe+197Au collisions more fully within the equilibrium approach. A determination of the breakup density for these multifragmentation events is attempted and the imcertainties of this determi- nation are addressed. A technique for extracting spin information of unbound states from correlation functions is also presented, and applied to the determination of the spin of the 0.774 MeV excited state of 8B (which is J21) and to some other nu- clei as well. Three particle (:(iirrelations are also shown and discussed for the cases of deuteron-alpl1a—alplia, 1')rot.(_)n—alpl121—alpha, alpha-alpha—alpha, and proton-proton- alpha correlations. 5.1 Selecting the Impact Parameters To select the events of central collisions, we utilized charged particle multiplicity detected in both Miniball/l\rIi1'iiwall and LASSA as the impact 1;)arameter filter. This can be achieved from a monotonic relation between total charged particle multiplicity NC and the reduced impact parameter b[84], r, _ Kawdivci/(Nc) ”‘2 i3: ._ . bmar [39¢ d‘NrCY ( ‘IVC ) (5.1) where Y(NC) is the number of events associated with charge particle multiplicity NC. bnm is the mean impact parameter for the collisions with NC = 3. Figure 5.1 shows the above relation for this experiment, where b = 1 cor'res1')(_)nds to the glancing 56 Figure 5.1: reduced impact parameter as function of total charged particle multiplic- ity. collisions and b = 0 represents the head—011 collisions. In the following discussions, we will choose the gate of 0 < b < 0.3 for the selection of central collisions, which amounts to about 7% of total statistics. Before going on to extract source information in the correlation functions, we show the single particle energy spectra. Five panels in Figure 5.2 present the differential multiplicities for p, d, t, 3He and 4He isotopes at angles of 17", 27°, 37", 47" and 57". These nearly equally spaced angular distributions are at angles where the detection efficiencies are large. The spectra at forward angles are more energetic, reflecting that emission occurs from a frame (i.e. the center of mass) that is moving rapidly with respect to the laboratory frame. Within the energy and angular range of the correlation analysis described below, the detection coverage and efficiency is smoothly well behaved. 57 dM/dE/dQ dM/dE/dQ l '3 V. L X‘ ~ -2 -2 10 i . 1O . ’ o C" t 17 U -3 17° % 10 T 270 -3 27° ,3, l 10 r . 370 . Deutron 37° Proton .. 470 10 -41. 470 57° r 570 -4 t 10 l r l l 4 1 l 0 25 50 75 100 125 15 0 100 150 200 Elab (MeV) Emb (MeV) .2 i l 10 L , -3 i; ' ‘0 r i L -3 ’ L 10 g g" E Q 10 '4 :70 i : 7 .4’ S t ‘V J 2 10 f U 370 : ) 47° . -5 o -5 _ 57 10 r 10 : 0 so 100 150 ”200 250 o 50 100 150 200 250 300 Elab (MeV) Em (MeV) -2i i : 17° 10 r 3 .. 37° ° . g 10 r i “ 47° m : . .3 ‘ E ‘ 5 ‘ 2 l ‘He '0 -4 10 g L -5 10 O 50 100 150 200 250 300 Ebb (MeV) Figure 5.2: Differential multiplicites of hydrogen and helium isotopes at angles of 17° 27°, 37°, 47° and 57°. 58 3 5.2 Two Particle Correlations and the Koonin-Pratt Formalism Experimentally the two particle correlation function may be defined as follows, grams.) = C<1+ thiizizmmmii (5.2) where Ylg is the two particle coincidence yield of a given pair of particles with their individual mmnenta $2 and 7?.2 respectively, and the Y,(?,) are the single particle yields for the two particles measured under the same impact parameter selection but not in the same event. The summations on both sides of the equation run over pairs of momenta $2 and 7?2 corresponding the same bin in relative momentum q. The correlation function describes how the correlation between coincidence particles measured in the same event differs from the underlying correlation dictated largely by phase space and modelled by mixing the single particle distributions of particles from two different events (the so—called event-mixed yield). The correlation constant C is typically chosen to ensure that R(q) = 0 at large relative energies where the correlations due to final state interactions and quantum statistics can be neglected. If the yields are normalized to be the appropriate differential multiplicities, C will be of order unity. In the case of proton-proton and pion-pion correlations, most correlation function analyses have compared data to the Koonin-Pratt equation [59, 60]. If the summation in Eq. (5.2) does not involve strong constraints on the emission angles of particles 1 and 2, the appropriate comparison is to the angle-averaged Koonin-Pratt equation [59, 60], C(q) 5 1+ R(q) = 1+ 47r/d7'7‘2K(q,-7')S(r), (5.3) 59 where the source function S (7) is defined as the probability distribution for emit- ting a pair of particles with relative distance 7‘ at the time when the second particle is emitted; it should be normalized to unity if all the emission components are in— cluded. The angle—averaged kernel, K(q, r), is obtained from the radial part of the ant.isyn'm'ietrized two—proton relative wave function as follows [59, 60], Ice, 7) = are»? — 1 (5.4) where the wave function q(r) describes the propagation of the pair from a separation r out to the detector at infinity, where relative mmnentum q is reached. Correlation functions have been analyzed using Eq. (5.3) for a variety of cor- relations involving hydrogen and helium isotopes [85, 61, 62]. Figure 5.3 shows the proton-proton correlation function measured in central 129Xe+w7Au collisions at E / A = 50 MeV. The p—p correlation function, as a function of the relative momentum, is measured over angles of 12° 3 6 _<_ 62" covered by LASSA without c.m. energy cut. The maximum at relative momentum q 2 20 MeV/c is caused by the strongly attrac- tive singlet S-wave proton—proton interaction; the minimum at q o: 0 MeV/c arises from the interplay between antisymn'ietrization and the long range repulsive Coulomb interaction [59] . The p-p correlation function with different c.m. energy cuts is also analyzed as shown in Fig. 5.4. The bottom panel shows the p—p correlation function for the same angular gate and with a gate which selects low energy protons with 0 < Em, < 15 MeV. The middle panel demonstrates the corresponding correlation function with a gate of 15 < Em < 30 MeV and the top panel presents the p-p correlation function with a gate of Em, > 30 MeV. Consistent with previous oliiservations, the correlation function is clearly more enhanced for protons with higher energies, corresponding to the emission of more energetic protons from sources that are smaller or shorter-lived 60 1.4 . 0 Exp Data [:l Imaging 1.2 , lyk“ g 1 r. 5g!“ KW‘_‘M.D;_M;_. 0.8 r 0.6 .. O 20 4O 60 80 100 qre,(MeV/c) Figure 5.3: Inclusive p—p correlation function in central 129Xe+197Au collisions at E / A = 50 MeV is shown. The black points are experimental data and the grey line is the best fit by the imaging technique. See table 5.1 for specific parameters. (or both) than the sources that emit lower energy protons. In studying the properties of the two proton emission sources, the simple assump- tion of a single Gaussian source [85], with unit normalization, has been used but could not reproduce all the features of the correlation functions. Here, we use the less model-dependent imaging technique of refs. [86, 87, 88, 63] where p—p correla- tion functions have been analyzed by numerically inverting the correlation function in Eq. (5.3) to obtain the source. Such an inversion is not completely straightforward because one has to take into account experimental factors (limited statistics, finite widths of momentum bins, etc.) and the intrinsic resolution of the kernel K(q,r). 61 1'5 0 Exp Data 4r" Imaging 1177"" ‘Iit‘m... 1 ,___. ¥ ‘ W..+-A-+- 4 Ecm > 30 MeV l l l l 1'5 0 Exp Data Imaging g 1 _ frfihiw‘g’t—o- HW-o—mfi— -—O-- —-O--—-—-0- O 5 15 < Ecm < 30 MeV 1. 5 l l l l I I ' 0 Exp Data — Gaussian 1 _ o l O < Ecm < 15 MeV O 5 l l L l 0 20 4O 6O 80 1 00 are, (MeV/c) Figure 5.4: The p-p correlation functions are shown for three c.m. energy gates of 0 < Em < 15 MeV, 15 < Em < 30 MeV, and EC", > 30 MeV. For the middle and high energy gates, the image technique is used. The simple Gaussian source parametrization in Eq. 5.6 is applied for the low energy gate. See table 5.1 for specific parameters. 62 Small fluctuations in the data, even if well within statistical or systen‘mtic errors. can lead to large changes in the imaged source function. On the other hand, successful inversion of the correlation function has the virtue of being model independent. The numerical inversion of the correlation in Eq. (5.3) was achieved via. the optimization algorithm of refs. [87, 88, 63]. Using the imaging technique, we analyzed the data shown in Fig. 5.3 by decomposing the source in a superposition of six b—spline polynomials of the 3rd order over the interval 0 g 7' g 20 fm [87, 88, 63]. -3 x10 0.2 0.15 7 Imaging 'E E 0.1 (71’ 0.05 O 1 l i l l L r 72' l 0 2 4 6 8 10 12 14 ram) Figure 5.5: The imaged source function is shown by inverting the p—p correlation in Fig. 5.3. See table 5.1 for the extracted source radius. The thick gray line in Fig. 5.3 is the best fit of the imaging approach to the experimental data (black points) of the two proton correlation. The imaged source function is shown by the thick grey line in Fig. 5.5 where the width of the line 63 indicates the uncertainty in the inversion. Since. the correlatirm function is difficult to measure at low relative momentum (say. (1 < 10 MeV/c), the imaging technique could not reliably produce the tail of the source at large radii (say, 7' > 13 fm) where the dominant sources are delayed emissions of protons, such as from secondary decay. Thus the source is only specified for 7' < 13 fm. Better data of the correlation function at small relative momentm‘n could provide better estimation 011 the source at larger radii where the source shape is mainly affected by the slow emissions of protons. The extracted rms radius of the source is rm, 2 7.97 fm for fast protons that contribute to r < 13 f m (see Table 5.1). However, the fraction of protons from these slow sources can be obtained via the integration of the imaged source [63]. Due to the omission of those long—lived components in the imaging of the source, the integration of the source is no longer unity [63] f2 E .\ = / S('r)d3r, (5.5) where f is the fraction of protons emitted from fast sources while l-f stands for the fraction of protons emitted from slow sources like secondary decay which can not be imaged in the long tail of the source profile. By integrating out the source shown in Fig. 5.5, we obtain f=0.48, which indicates that the emitted protons with a slow time scale (or in other words, 7* > 13 fm in the source profile) amount to about half of all the protons. The two proton emission sources are also analyzed for the p—p correlation function with different c.m. energy gates in Fig. 5.4. For the middle and high energy gates, the extracted source functions are shown in the middle and top panels of Fig. 5.6 using the imaging technique. Unfortunately, the higher energy gate has no data at low relative momentum and the data at lowest point in relative momentum is measured with poor statistics. Thus the inversion provides no reliable information at r > 7 frn. We have 64 x10 0.6 0.4 0.2 Imaging Ecm > 30 MeV 1 ii. i "1' . ' l S(r) (fm'3) 0.4 — 0.2 Imaging 15 < Ecm < 30 MeV 0.4 ~ 0.2 _ Gaussian 0 < Ecm < 15 MeV l l l l Figure 5.6: The source functions are shown for the p-p correlation functions in Fig. 5.4 for three our. energy gates of 0 < Em, < 15 MeV, 15 < EC", < 30 MeV, and EC", > 30 MeV. For the middle and high energy gates, the image technique is used. The simple Gaussian source parametrization in Eq. 5.6 is applied for the low energy r (tm) gate. See table 5.1 for extracted source radii. 4 6 81012 p-p Inclusive 0 < Em" < 15 15 < EM,m < 30 Elm" > 30 Correlation MeV MeV MeV /\ 0.22 0.93 0.50 0.14 X 0.11 0.22 0.16 0.14 r.,.,,,,, 7.97 11.26 10.95 4.69 p/[),, 0.52 0.18 0.20 2.52 Table 5.1: List of quantities are shown for the pp correlations (Figs. 5.3-5.6) for inclusive and three different c.m. energy gates. The A that relevant to the fraction of protons contributed to the fitted source and the rms radius rm, of the two proton source are extracted from the fitted source distribution. The X values are calculated from the integral of the source functions for r i 7 fm in Eq. 5.5. The density p/po is estimated by taking into account collective effects in Eq. 5.9. calculated rm, by integrating the source for the highest energy gate over 1" < 7 fm and the source for the medium energy gate over 7" < 13 fm and obtained sources sizes of rm, :2 10.95, 4.69 fm and corresponding fast. proton fractions A = 050,014 for the middle and high energy bins, respectively (see Table 5.1). The corresponding reconstructed correlation functions are shown by the grey lines in the middle and top panels of Fig. 5.4. Because the correlation data for the high energy gate (top panel in Fig. 5.4) has no points below q < 10 MeV/c one cannot say what are the source contributions at 7‘ > 7 fm. If the source function obtained from a better measurement that includes data below q < 10 MeV/c reveals a tail on the source function at larger radii r > 7 fm similar to that for the middle energy gate, the extracted /\ value for the high energy gate will be much higher. On the other hand, if the tail of the source function for the middle energy gate are cut off, its A value will be much lower, which is shown by column X in Table 5.1 for the integration of the source functions in Eq. 5.5 for r < 7 fm. As for the lowest energy gate, however, we have not succeeded in stably imaging the data because the data do not conform to the shape expected for a correlation function of the Koonin-Pratt type. Instead we have fitted the correlation function 66 data with a non—unity Gaussian source of the form, /\ 7‘2 ‘5"1 :~"" ' . 1‘4 —_ 1 [7.7 (r) (2,),,,.,.8u
20 MeV/c.
In general, the extracted sources for lower energy protons display larger source
radii, which is consistent with emission from a source that has expanded and may be
disassembling over a time frame that is somewhat longer than is the characteristic of
the higher energy protons. With the uncertainties in the present correlation functions,
we can not image the sources for these long-lived decays. We do, however, have
indirect sensitivity to those decays from the fraction of fast protons f = A extracted
from the integral of the source in Eq. 5.5, which indicates there is a significant
contribution from secondary decay to the proton emission.
Before relating the proton source functions to the density of the multifragment-
ing source, we investigate d-alpha correlations, which can also be addressed by the
more conventional correlation function techniques involving the Koonin-Pratt equa-
tion. At present, d-alpha correlations have not been successfully inverted with the
imaging technique. Instead, we illustrate the fit here with the simple Gaussian source
parametrization described previously.
67
3
09 F4..__.J_._‘_ __l*-_l_-4
7 so 90 100
[I
+ 2 -
,_
1 a
O
. 0 Exp Data
. . ' —- Gaussian w/o correction
i l l 4 l l l
0
0 20 40 60 80 100120140
qrel(MeV/c)
Figure 5.7: The inclusive d—a correlation function for the central collision gate is
shown as a function of relative momentum by the black points. The solid line is the
fit by the simple Gaussian source parametrization in Eq. 5.6 without corrections for
collective motion. The blown-up in the top right shows the poor quality of the fit to
the second peak of the d-a correlation.
Figure 5.7 shows the inclusive d-alpha correlation function as a function of relative
momentum. The solid line in Fig. 5.7 shows the best fit to the experimental data with
Gaussian source without making any correction for collective motion. The Coulomb
suppression of the correlation function at small relative momentum (Ire! < 30 MeV/c
can not be fit with one single Gaussian source assumption like the case of the p—p
correlation due to the existence of slow sources like secondary decay. Therefore the
fit is performed at qrer > 30 MeV/c and describe the magnitude and shape of the first
excited state of 6Li at 2.186 MeV well. However, the magnitude of the second peak at
68
4.31 MeV (see the blown-up in Fig. 5.7) is overpredicted and its shape is incorrect, a
problem that has been noted by previous authors. Before interpreting this correlation
data further, some discussion of the possible effects due to collective motion is needed
because collective motion may be responsible for this difficulty. After this discussion,
we will return to the interpretation of both d-a and p—p correlation functions.
5.3 Influence of Collective Motion
Collective motion is characterized by an average velocity field that assumes values
that are defined by time and by the location within the colliding system. Examples
of collective motion include the collective radial expansion of an excited system under
the influence of pressure due to compression and to the repulsive Coulomb interaction,
or collective ”rotation” induced by a collision that is at non-zero impact parameter.
Another form of collective flow, directed transverse flow [89], is small at E/A=50 MeV
due to the balancing of repulsive and attractive momentum transfers at the ”balance
7
energy’ and will not be discussed further here.
Figure 5.8: A source with only thermal motion (open arrows) is shown on the left.
The collective velocity field (solid arrows) is drawn on the right and results in a grey
area of the source where emission into the right half plane is improbable.
Collective motion has several impacts on the behavior of correlation functions.
First, it effectively decreases the ”source size”, i.e., size of the region that can con-
69
tribute to the emission at the measured angles and energies. The origin of this effect.
is simply illustrated in Fig. 5.8. O11 the left side, we consider the scenario where
there is no collective velocity and on the right side, we include a collective velocity
field. In this drawing, the collective velocity field is drawn by the solid arrows and the
random, i.e., thermal velocities are drawn in open arrows. The grey area indicates
the region of the source where emission into the right half plane is improbable. In
the long mean free path scenario, the source without collective motion corresponds
to the total volume of the system. With collective motion, however, emission to the
right is precluded in the grey region where the collective velocity has a component
to the left that exceeds the typical random velocity, which if thermal in origin would
. _ ‘)
decrease With mass as 1) oc (m) 1/ “.
To estimate the influence of collective motion on the source sizes, we simulate the
interplay between the collective and random source velocities as follows. We assume
a spherical source of A,,,,,,.(_.,. = 0.8(Ap,.0jm,rC + Amget) = 260 nucleons, of which 106
are protons and that the source moves with the velocity of the center of mass. Within
this spherical source, the collective velocity field at 7? is given by
r ‘ 1
+ ’l’tJnaxQ ® 1 (5-7)
$201K?) : Unmax
:Ulrl
:Ulri
where vmm is the radial velocity of the surface expansion of the system at freeze-
out, vtmax is the tangential collective velocity at the surface, R is the radius of the
expanded source, and CC} defines the direction of the rotational velocity field. The
orientation of (I) is approximately perpendicular to the beam axis, but the azimuthal
orientation with respect to LASSA is unknown. Therefore, the azimuthal orientation
is varied randomly assuming the relative azimuthal orientation between LASSA and
(D is isotropic.
We consider first the fraction of the source that can contribute to particle emission
70
into LASSA. Emission was simulated via Monte Carlo for the interplay of this collec-
tive velocity field with random or thermal motion [90]. In this simulation, the thermal
velocity was calculated by assuming m, = /3T/m for simplicity where m is the mass
of the particle and the temperature T -_— 4 MeV is roughly ('tonsistent with isotope and
excited state temperatures in this bombarding energy domain. we choose a radial
expansion velocity consistent with a radial kinetic energy of expansion of 2 MeV (and
it is also affected by the additional Coulomb energy gained by continuing to acceler-
ate the particles by the Coulomb field of the source). It is rather approximate but
consistent with the radial expansion velocities reported in the literature. In addition,
we choose an tangential velocity of chum = 0.16 c, which corresponds to a surface
tangential velocity consistent with the initial velocities of the projectile and target.
We use the above collective flow 1.)arameters for the ” maximum” flow discussed below
and reduce them to see the dependence on the collective motion.
To demonstrate the mass dependence of the source size decreased by the collective
motion, two particle sources from p-p, d-d, up to 20;\le-2U1\Te are studied assuming 80%
in velocity (or 64% in energy) of the ”maximum" collective motion discussed above. In
essence, the rms radius for a two partice source is calculated as 7‘3,” = ((72 — 791?),
where 7’, is the location of the i” particle at the time of emission and the average
is over particles that make it into the LASSA array. If the two particles are emitted
from the same source, the rms radius for the two particle source is \/2 times the
rms radius for a single particle source. The velocity field in Eq. 5.7 is constructed
so as to make the ratio of the rms radii to the overall system radius invariant with
respect to the system size. The ratio fwu, of the rms radius with collective motion
divided by the rms radius without collective motion for the two particle sources, is
plotted as the solid line in Fig. 5.9 as a function of the mass number while the ratio
without collective motion is denoted by the dashed line. The solid line was obtained
71
1.25 "
0.75 *-
fcoIl
0.5 '-
0.25 *‘
l l l
0 5 10 15 20
Mass Number
C
Figure 5.9: The source reduction factor, f coir, obtained from the ratio of the rms radius
to the system radius for the two particle sources with collective motion, are shown by
the solid line as a function of the mass number. The ratio without collective motion is
denoted by the dashed line. The solid line is calculated for a uniform spherical system
with a radius of 11 fm while the three symbols denote the corresponding calculations
for a system with a radius of 7 fm.
from calculations for a uniform spherical system with a radius of 11 fm. The three
symbols denote the corresponding calculations for a system with a radius of 7 fm;
this demonstrates the independence of this ratio with respect to the system size.
The corresponding reduction in the fraction of the participating source, fgou, es-
sentially dictates the fraction of the participating source, i.e. the mass fraction,
independent of the actual source volume. It depends on 1.3mm, Laymax, and the ran-
dom velocity v”, = y/3T/ m, but not on the source radius R. Clearly, the collective
velocity field decreases the mass of the source significantly, for example, by about a
factor of 2 for the p-p correlation. In Fig. 5.9 no cuts on relative momentum are
applied since we want to Show the mass dependence of the source reduction factor.
72
In reality, we need to obtain the ff
,,(
,,, factor for a gate of relative momentum where
we study the correlz-ition function. The corresponding reduction of the d-a source
is also calculated for the purposes of density extraction from the d-a correlation to
be discussed later in this section. This calculation yields a mass reduction factor of
about 5 for a relative momentum gate of 0-150 MeV/c.
In addition, the collective motion causes the particles to be more frequently emit-
ted in the plane perpendicular to of). This leads to a preference for the relative
azimuthal angle of the two particles to be 0° or 180” [91] and also modifies the shapes
and magnitudes of broad correlation function peaks. In such cases, the two particles
are emitted from nearly the same point and experience nearly the same collective
velocity field. On the other hand, the mixed event background includes particles that
are from two different events in which the collective velocity fields are completely dif-
ferent. Thus the relative velocity distribution for the resonant contribution is dictated
essentially by thermal motion and consequently narrower than the relative velocity
distribution for the non-resonant and mixed event backgrounds, for which the relative
velocity reflects differences in the collective velocities as well.
Fig. 5.10 shows a simulation of that effect for the d-alpha correlation. The solid
line shows the relative energy distril'nition between two particles that lie within a
distance of 2 fm from each other as they would be if they were both emitted from a
particle unbound fragment formed at freezeout. This relative energy distribution is
essentially consistent with an exponential decline of the form oc \/E,-€[ exp (—-E,.,_,1/T)
given by the thermal motion. The dashed line shows the corresponding relative energy
distril‘mticm for pairs of particles that are chosen randomly throughout the system as
they would be for the non-resonant background. Here the relative energy (.listribution
is much wider consistent with significant contributions from the difference between the
collective velocity field at the two emission points. The dotted curve in the same figure
73
Yield
ll 11 l in!
0 5101520253035
Erel (MeV)
Figure 5.10: The solid line shows a simulation of the resonant distribution as a func-
tion of relative energy for the d—alpha correlation. The dashed line denotes the sim—
ulation for non-resonant background while the dotted line demonstrates the relative
energy distribution for mixed event background. All the distributions are normalized
to one for comparisons.
shows the additional broadening that occurs in the event mixing in the denominator
of the correlation function where the two particles originate from different events
with different orientations of the reaction plane, i.e., different azimuthal orientations
for (2’. For simplicity, we could describe the distribution of relative energies here by
CC Ere! exp (—Erel/Tnii;r)-
The different broadening in these distributions has important consequences for the
correlation function, where the two distributions are essentially divided by each other.
The division of the resonant yield by the mixed event distribution is illustrated in the
74
10 100% collective motion
Teff = 6 MeV
1
-1
1 0 g
l J l l l l l
10 80% collective motion
E Tait = 7 MeV
t i
g -1 E
>7 10
r l l l l l l
10 E no collective motion
1 ; - - -
-1 E
1 0 E‘
f l
0 5101520253035
Ere,(MeV)
Figure 5.11: The ratios of the resonant yield over the mixed event distribution for
the d-a correlation are shown in three panels for maximum, 80% and zero of the
collective motion, respectively. The corresponding values of T6” are extracted from
the fits (dashed lines).
75
three panels of Fig. 5.11 for maximum. 80% and zero of the collective motion discussed
above, respectively. The dashed lines in the figure are the best fits of an exponential
function o< exp(—E,.,.,/T,,ff), which have values of T...” = 6,7 MeV for maximum
and 8000 of collective fiow, res1.)ectively; the line is fiat in the bottom panel without
collective motion. This simulation indicates that one should expect the resonant
peaks in the correlation function to decline exponentially with the relative energy
or relative momentum according to ac exp(—Er€r/T + Era/Tina) = exp(—E,er/Teff).
The non-resonant background is also expected to decline somewhat, but to a much
reduced extent.
Taking the simulations in Fig. 5.11 into account, we return to the d—a corre-
lation in Fig. 5.7. Consistent with the simulations, we find that the non-resonant
background is relatively well represented in the original calculations. However, the
resonant yield can be much better described by multiplying the resonant contribution
by a factor ex1)(—E,.,.1/'T,,ff) with Te” 2 6.5 MeV. This revised correlation function
is given by the solid line in Fig. 5.12, which fits the second peak of the d-or correla-
tion function at gm 2 84 MeV/c (the blown-up) much better than that in Fig. 5.7.
The extracted source radius is To = 3.3 fin and the rms value and A can be seen in
Table 5.2. Using this empirical correction, we have explored the sensitivity of the d-a
correlation function in general for the extraction of information about the size of the
emitting source using a Gaussian source of the form discussed in Eq. 5.6. Consistent
with conclusions recently obtained for d-a correlations on other systems [92], we find
that the width of the second peak in the correlation function at qrei = 84 MeV/c
is very sensitive to the source size. This is illustrated in Fig. 5.12 where two other
fits are shown by the dashed and dotted lines with the assumption of source radii
of To = 2.0 fin and 5.0 frn, respectively. This situation is similar to that. for the p—p
imaging analysis where the source radius information is primarily obtained from the
76
1.1 r
1
3 _
0.9_ l __ I_.____L l .4
70 80 90 100
CC
+ 2 r
T.
; 0 Exp Data
. . —— Gaussian w/ correction
. (Te,,=6.5MeV)
[ I l l L i l
O .
0 20 40 60 80 100120140
are, (MeV/c)
Figure 5.12: The inclusive d-a correlation function for the central collision gate is
shown as a function of relative momentum by the black points. The solid line is the
best fit by the simple Gaussian source parametrization in Eq. 5.6 with Te” corrections
for collective motion. The blown-up in the top right shows the good reproduction of
the second peak of the d—a correlation after Te” corrections. The extracted source
radius is r0 = 3.3 fm (see Table 5.2 for the rms value). The dashed and dotted
lines are the fits assuming that the source radius is To = 2.0 fin and r0 = 5.0 fm,
respectively.
77
width of the p—p correlation function, not its height[63].
d-a Inclusive 0 < Edm, < 20 20 < Eacm < 40 detn, > 40
Correlation 0 < E01,", < 25 25 < Ema” < 45 Ema" > 45
MeV MeV MeV
A 0.11 0.14 0.13 0.13
rm, 5.72 7.35 5.58 3.01
p/po 0.54 0.37 0.59 3.74
Table 5.2: List of quantities are shown for the d-(r correlations (Figs. 5.12-5.13) for
inclusive data and three different c.m. energy gates. The A that relevant to the
fraction of the involved particles contributed to the fitted source and the rms radius
rm, of the two proton source are extracted from the fitted source distribution. The
density p/po is estimated by taking into account collective effects in Eq. 5.9.
Different energy gates for the d-alpha correlation are also studied as shown in Fig.
5.13. By varying the parameter Teff from collective effects, we can fit the correlation
functions for the high, middle, and low energy gates, especially for the second peak at
q = 84 MeV/c which is blown up in the top right. corner of the corresponding panel.
While the first peak can be fitted somewhat better by adjusting Teff in this simple
one Gaussian source parametrization and the Coulomb suppression at low relative
momentum can be fitted better by adding a tail to the gaussian source, we will defer
further discussion of the first peak until the next. section and devote our attentions
here to achieving the best fit to the second peak.
The collective motion effects are more evident for the lower energy gate where the
Ten value is lower than for the highest energy gate where the effects virtually vanish
and no Teff correction is needed. The extracted rms radii and A values for the three
energy gates are in Table 5.2. Same as the case in the p-p correlation, the sources
extracted from the d—a correlation functions have smaller radii for the higher energy
gates.
In principle, sources extracted by inverting p—p correlation functions may also be
influenced by collective motion. Fig. 5.14 shows the corresponding results from simu-
78
lac,cm > 40MeV
4 ”Emu" > 45MeV
(Teflzoo)
O P‘ l l n} i l l
20 < Edm < 40MeV 1.1 '
4 25 < Eum < 45MeV 1 A
[E ne,=6.6MeV) L _ J“ J _.
‘— 2 L-
O
O "P ". l l l l 1 i
0 < Ed.cm < 20MeV 1.1 i
4 #0 ( Emcm < 25MeV 1 W
=4.6MeV
(Ten ) . 0.9] m l 1 l
2 n 70 80 90 100
.0
0 "f i L l l l i
0 20 40 6 80 100120140
qre,(MeV/c)
Figure 5.13: The d-a correlation fimctions are shown for three our. energy gates
of 0 < Edy", < 20 MeV and 0 < EM", < 25 MeV, 20 < Ed,cm < 40 MeV and
25 < Ema", < 45 MeV, and Edm, > 40 MeV and Emcm > 45 MeV. Fits by the simple
Gaussian source parametrization in Eq. 5.6 are shown by the solid lines after Tc”
corrections for collective motion. See table 5.1 for extracted source radii and A values.
79
10 100% collective motion
. T9,, = 8.7 MeV
1 r
-1
1 0
h l L l l
10 6:: 50% collective motion
'5 : A T9,, = 29 MeV
: i " M
g -1 E
>-5 1 O
l l 1 l
10 S no collective motion
1 r v
-1 E
1 0 g
7 1 l l l
0 5 1 0 1 5 20
Erel (MeV)
Figure 5.14: The ratios of the resonant yield over the mixed event distribution for
the p—p correlation are shown in three panels for maximum, 50% and zero of the
collective motion, respectively. The corresponding values of Teff are extracted from
the fits (dashed lines).
80
lations of the infliu—mce of collective motion on the p-p correlation functions assuming
three different sets of collective flow parameters. The corresponding values of the
effective temperature T,“ are much larger than the values extracted from the do
correlation, which indicates this effect is much smaller for the p-p correlation. In
addition, the peak of the p-p correlation function falls within the very low energy
region of EM < 0.5 MeV, unlike the case of the second peak in the do correlation
function which is at EM 2 2.836 MeV. Therefore, the (,listortion effects from Te];
should be much smaller for the p—p correlation than those for the d-a correlation. In
the interest of brevity, we neglect them in the following analyses.
Using these analyses of the p—p and d-a correlations, we attempt to place some
constraints on the densities of the system at the time of particle emission. Because
we are using source radii determined from the widths of the p—p correlation function
and the width of the second peak in the d—a' correlation, we are insensitive to how
collective motion reduces the fraction of deuterons, alphas or protons that come to the
LASSA and contribute to the mixed event background to the correlation function. On
the other hand, we are sensitive to how the imaged region may be effectivr—ély shrunk
by the collective motion and we must make some correction for that effect. For
these estimates, we use the factor f2,” determined earlier using T,” = 7 MeV, which
is close to the value obtained for most inclusive correlation functions, to determine
the fraction of the system that is contributing to these emissions. Take the p—p
correlation function shown in Fig. 5.3, for example, the rms source radius rm, =
7.97. The simulation in Fig. 5.9 suggests that the p—p correlation function, actually
originates from a fraction of the source. This fraction is given by f3,“ where f mu = 0.8.
We assume that the mass contained in the region sampled by the pp correlation
function to be f2,” - A,,,...f,.ag,,,e,,t :2 (0.8)3 - 1419”,",ng where Apnqmgmem is the mass
of the system after the end of pre-equilibrium emission. For this estimate. we take
81
Aprefmgment a: 260 corresponding to 80% of the total system.
To obtain an averaged density. we need to find the volume lr’_,.(,,,,.a. of the breakup
source measured by the two 1;)article source function. It. is impm‘tant to note that the
rms radius of the two particle source function is obtained by adding the rms radii of
single particle source functions in quadrature, i.e.. rm, 2 \/21'1. For an equivalent
sharp uniform spherical source with a radius of R, the rms radius of the two proton
source has a relation of rm, 2 ,/(6/5)R and consequently we can obtain the source
volume,
4 4 3
1/source 2 §R3 : i: ' rrms - (5.8)
Therefore, the density can be written as,
P/PO = 30" ' A’ (5.9)
7r 5 . i
‘3—[ afr'ms] 100
where the saturation (ilensity [)0 = 0.16 nucleons/fin3 is assumed. Densities obtained
for the all energy gates for the p—p correlations are. given in Table 5.1 by using Eq.
5.9. The extracted value of the density for the low energy gate is not as reliable as
those obtained by the imaging technique since the Gaussian source approach does not
reproduce the correlation for the low energy gate in Fig 5.4. However, one does see
a trend that the system is at a higher density when higher energy protons are being
used to construct the correlaticm. The density value from the high energy gate is so
high (/)/[)0 = 2.52) that it seems the correlation either ’sees’ the very early source
(compressed and before being expanded) since fast protons tend to emit earlier, or it.
just ’sees’ a smaller fraction of the source. The latter case would apply if the source
in the early stage is limited to the surface due to the short mean free path in the
dense interior.
82
For the d-a correlation, we l'l‘dV’C a different collective correction factor fmu = 0.58,
which is obtained by Simulating d—a correlation sources as discussed previously at
low relative momentum where we fit the data. Values for the density obtained from
the d—a correlation functions are given in Table 5.2 by using Eq. 5.9. Since the
simple Gaussian source paranretrization is used in this case, it can not reproduce all
the features of the d-a correlation fimctions and the presumed Gaussian shape for
the sources may be rather schematic. Consequently. the extracted source radii and
densities may not. be as reliable, as those obtained from the p-p correlation functions
by the imaging technique. After taking into account the effects of collective motion
properly, however, we obtain similar values of density (0.52 and 0.54, respectively)
from both the inclusive p-p and do correlaticm functions. Moreover, a similar trend of
the energy dependence of the density is observed. which indicates that higher energy
particles are emitted at denser regions or earlier times. we note that it can be argued
that more accurate values for the density could be obtained if fa,” were calculated
for each of these energy gates via simulations tuned to reproduce the values for T6”
observed for each gate. Since lower values of Toff will result in smaller fwu, the values
for density in this case woulcl be somewhat. lower.
Clearly, there are large uncertainties in this approach. These uncertainties stem
from several different sources. First concerns whether the corrections for reduction of
the source sizes from the Monte Carlo simulations are of the correct magnitude. We
believe that while they are qualitatixr'ely correct, they could be off in their magnitudes
somewhat and we are considering ways to try to estimate this uncertainty more quan-
titatively. Secondly, we have some concerns about the assumption of an infinite mean
free path that this density estimate employs. Clearly, this assumption is more correct
if the estimated density is small. When it is large, however, transport theoretical cal-
culations indicate a. sensitivity of two particle correlation fimctions to the iii-medium
83
cross section. Thus the estimated values for density for the p-p correlation functions
in Table 5.1 and the d-a correlation function in Table 5.2 are somewhat uncertain
on this account and if there are corrections needed, it. would be in the direction of
reducing the density.
5.4 Equilibrium Correlation Functions
One of the factors limiting the extension of the Koonin—Pratt equation to heavier
particles is the care needed to construct the Kernel K(q,r). Essentially, one must
search for a set of attractix-e nuclear potentials that can reproduce the experimental
phase shifts. Right now, we have only the necessary 1‘)otentials for the p—p and d-a
correlation functions. To rapidly extend the correlation function to heavier particles
and to facilitate the comparison to statistical models, we develop here a. f(,)rn'ialism
for calculating the correlation function within equilibrium theory.
The starting point for this development is the coi’isideration of elements needed
for the equilibrium description. First, one needs to have a con'ipact method for
incorporating both the long range Coulomb and short range nuclear interactions.
Second, one must address the volume that is occupied by other particles. We choose
to address the second issue by invoking the excluded volume approximation. This
essentially amounts to counting as particles only those that are isolated, a procedure
that is consistent with most equilibrium multifragmentation approaches [9, 10].
Equilibrium correlation function expressions are derived by considering how the
two particle phase space is modified by interactions. For simrflicity, we consider the
simplified geometry wherein the pair of spinless particles with charges 21 and Zg is
with its center of mass at the center of a volume V. To calculate how the phase space
of relative motion is modified by the Coulomb interaction, we follow semi-classical
84
theory which states that the phase space density is given as a function of the relative
spatial separation 7’ and relative momentum 77 by
(In. _ 1 (L, 10)
JWWWL—W’ 0~
where TKL is the local momentum given in terms of q, the reduced mass it 2: Ml Alg/(il[1+
ill-3) and the Coulomb potential by
2/I.Zl 2202
T
‘11. = (.12 (5.11)
Reexpressing Eq. 5.10 in terms of the momentum 7 at. large distances where the
Coulomb interaction can be neglected, we have
(In (In, (19% 1 2 [Z 2 e73
.. .nz—e LHL;%< an)
(1371!qu (Pg [1“ H]-
m lCi'oulomb :
If the above equation is integrated over a volume V and divided by the corresponding
integral of the relative phase space density of two free particles, (In/(1371" = V/h3, an
expression for the Coulomb correlation function 1 + Ram, may be obtained as follows,
2/1-212282
5.13
"I, < )
1 .
1 + R0011] ((1) = V f‘ (137‘ 1
The extension of this Coulomb correlation scenario to include the influence of
short ranged nuclear interactions can be accommodated using a formalism for calcu-
lating the second virial coefficient[93]. One begins by imagining that two interacting
particles are placed in a spherical container centered about their common center of
mass. Assume that the appropriate boundary condition is for the wavefunction of
relative motion to vanish at the container walls. In the asymptotic region, the radial
wavefunction is of the form
sin [qr/h — 7; ln (2qr/h) — 57V? + 5t((1)l
o<
(17‘
Kw an)
85
The lfioundary condition therefore requires
qR/fz. — I]lll (‘ZqR/h) — [7r/‘2 + (55((1) 2 77m,
A
v.31
p—A
C,“
v
r r) , . . .
where 17 = Z1Z2c“/hv is the Coulomb parameter. Separating the phase shift into
0 o I u a
Coulomb m- and strong interaction (5; components, the density of states 1s
a : (2(+1)d—"i
dq
dq (5.16)
(I
= %:(2€+ 1) {7})? + % [nln(2qR/fi.) +O'g(Q)]} + —71;Z(2€+ 1) {$51}
In Eq. 5.16, the first term represents the density of states for the pure Coulomb
problem and the second term is the density of states due to the strong interacticm.
Since the first term is difficult and unwieldy to use, one can use the seu’iiclassical
expression in Eq. 5.13 or some similar shape for the Coulomb density of states. We
will use the second term for the strong interaction effects. Taking the spin of the
particles and resonances into account, the two particle phase space of relative motion
becomes
(in 25 1 23 1 V . 2 .Z Z‘ :22
,"13 = ( ‘+ )l 2+ l—f (137*1———”‘2‘ (5.17)
(1'3 q h~‘ V . v 1712
+47r1q2J :(2€ )dq ’
where V f and V are the free (unoccupied) and total (including occupied) volumes
of the system, respectively. Given this relaticmship, the correlation function as a
function of relative momentum becomes
1+ R((1) : 1+ RCioul(q )+ Rim-C(q) (518)
_ v/‘d3__,\/1_2;_.___1Z1Z2e2
_ rq2
86
I13 do"
_. + .H.’
+ .. . . , , 2/ 1 .
‘7r2q2f‘f'(2.51+1)(252 + 1) ;( (Iq
and as a. function of relative energy Em, becomes
1 + R (Erel) : 1 + R(‘()111(Erel) + Rnuc (Erel) (519)
1 . 212-262
v /v f ' 7-E,.,.,
11.3 , dd,”
2f+1)
+ , ' f 1 , .
47M», - (251 + 1) (252 + 1) lz.\/2;1E.e, g; “Err!
It is useful to reflect on how the equilibrium correlation function both resembles
and differs from the correlation functions calculated via the Koonin-Pratt equation.
\Vhile the general case of a finite emission time lies outside the equilibrium cm‘relation
formalism, there is a close connection in the limit of instantaneous emission between
ec‘niilibriun'i and Koonin-Pratt predictions for the second term in Eq. 5.18, stemming
from the influence of strong interactions. This term in the equilibrium correlation
pictui‘e remesents static iiimlifications in the phase space distributions due to reso-
nances and other strong interaction effects. In this respect, the resonances can be
regarded as ”pie—existing". ()n the other hand, the correlations within the Koonin—
Pratt formalism have been examined in the "prompt” or instantaneous emission limit
for a. source that is much larger than intrinsic ”size” of the resonance itself by Boal
et al., and can be expressed by [94, 95]
3 . ~'
It do,”
. 2z+1 ,
4W2(12-(281+1)(252+1);( ) dq
1?"... (q) = S (0) (5.20)
for the strong interaction correlation function. If the source is taken to be uniformly
spherical, the 5(0) 2 UV}, and the equilibrium and Koonin—Pratt forn‘ialisms yield
identical expressions. Thus, the assumption of instantaneous emission implies that the
resonance structures predicted by the Koonin—Pratt formalism are those of unstable
particles that are ”pro-existing” at breakup.
87
Unlike the general situation for correlation functions calculated within the Koonin-
Pratt formalism (see section 5.2), the width of the resonance structure within this
large source approximation is given by the line shape of the resonance state. The
source size inforn'iation is therefore only contained in the magnitude of the correlation
structure through its dependence on S (0) = 1 / Vf. For states that are much narrower
than the intrinsic resolution of the detection apparatus, such as the case for the
2.186 MeV resonance in the d-a correlation, this does not. lead to a significant loss of
inforn'iation. Comparisons between analyses performed in equilibrium and Koonin-
Pratt limits by Jennings et al. indicate that the correlation function predictions for
the magnitude of this resonance peak by both approaches are about the same for
gaussian sources with tyrfical source radii [95].
5.5 Interpretations of Correlation Functions Using
the Equilibrium Correlation Approximation
To extend correlation function analyses to heavier particles that are more identified
with multifragmentation and the liquid-gas phase transition, we can at the present
time only apply the equilibrium correlation approximation. As mentioned in the pre-
vious section, this approximation is equivalent in many respects to the Koonin-Pratt
formalism. However, it does limit one to examine only the magnitude of the two par-
ticle correlation function peaks and there remains a. sensitivity to the secondary decay
corrections that cannot be determined from the technique itself, unlike the case for
the imaging procedure. This sensitivity means that one must employ statistical mod-
els to estimate the secondary decay and make the necessary corrections as discussed
below.
We begin our discussion of equilibrium correlation functions by returning to the
d-a correlation function, as shown again in figure 5.15 as a function of relative energy.
88
2.186 MeV 3+
4 d-alpha correlation
3 e .
[I
+
'— 4.31 MeV 2"
2 a
Ere, (MeV)
Figure 5.15: The experimental d-a correlation function is shown. The smooth solid
line is the empirical background used for the analysis. The dashed and dotted lines
are the Coulomb correlation calculated from Eq. 5.21 for sharp sphere radii of R2124
and 15.6 fm, respectively.
Here one can see clearly the peaks from the resonance states of 6Li: the 3+ state at
2.186 MeV and the 2+ state at 4.31 MeV. Also shown by the dashed and dotted lines
in the figure are Coulomb correlation functions calculated for a spherical volume of
radius R as follows,
1 . Z Z 1’2
1+ RCoul (Emil) : V /V dif1‘\/1 — TIEZ: (521)
_ [1 7‘min:| ”2 1 1 Tmin 3 (Tum)?
_ R 4 R. 8 R
89
I?
7 min
—1
3 (rmin > 31 . +
8 R rmin
where r,,,_,,, = Z1 ch’Q/Erd and we have used the spherical source radii of R12877
and 11.05 fm (corresponding ap1_)r(‘)ximately to the two-particle source radii of R =
flR1=12x~1 and 15.0 fm) consistent. with breakup at densities of l/Gpo and 1/3/)(),
respectively, if half of the 200 nucleon source can be ”seen” by the correlation due to
collective effects (see sect. 5.3).
While the general trend of the predicted Coulomb correlation functions are simi-
lar to that (solid line) experimentally observed, the calculated correlation functions
underpredict. the measured one at very low relative energies and there are some differ-
ences in the overall shape. The extra measured yield in the data at very low relative
energies i'nay reflect long lived secondary decays that are not modelled by the equi-
librium Coulomb correlation function. Other (.lifferences, however, may reflect the
fact that the long ranged behavior of the Coulomb interaction makes it difficult to
('listinguisl‘i two body from multi-body Coulomb effects. For example, a third frag-
ment in between the two measured fragments would repel them in opposite directions,
widening the correlation function minimum. Thus we have decided not to insist upon
fitting the measured correlation function with Eq. 5.21, but simply fit an empirical
l‘)ackground function (solid line) to the data, instead.
We therefore extract information about the source volume by considering what is
required to fit the resonance peaks with the nuclear correlation function in Eq. 5.19.
This fit, however, requires the application of the empirical correction for collective
motion described in Section 5.3. When this correction is applied and the derivative
of the nuclear phase shift is given in a Breit-W’igner form [96]
d6}, ~ r,/2
dq "‘ (Em, — E:)2 + r3/4
(3.13.), (5.22)
90
the nuclear correlation function in Eq. 5.19 l‘)ecomes,
3
1?,.1,C(E,..-z) z . 1 . ’1 (ad/'1‘. n (,
(251+ I)(252 + I) 477 /f[l\/2,11.E,-€]
en
[\3
00
v
r,/2
>< :lr—Zczj. + 1) (BR)
For resonance states, the 1.)opulation will not be affected by collective motion
and if equilibrium is achieved, it should populate according to the breakup thermal
temperature Tmflml. If there is no collective motion, the relative energy spectrum
for the mixed event background in the denon'iinator of correlation function will also
be described by the same temperature Tmmnaz. On the other hand, due. to collective
motion of the hot source, the relative energy of the two particles in the event-mixing
case tends to be higher because the particles originate from different. places with
different collective velocities and because the reaction plane may also be different
from one event to another. Therefore, the effective temperature of the two particles
from different events is characterized by a. much higher value Tm”. This results in an
extra. factor in the resonance correlation Rmm where
1 1 1
chf 71thermal T001]
(5.24)
we find that d-a correlations are well described by assuming, T0” = 7 MeV (see
sect. 5.3) which could arise if the breakup temperature T,;,,_.,.,m,1 = 4 MeV and col-
lective ” temperature” is Ta,” = 9.3 MeV. Such values are typical of the temperatures
observed for isotopic and excited state temperatures (Tina-ml : 4 MeV) and for the
slopes of energy spectra (T6011 2 9.3 MeV) [30, 97]. If one is interested mainly in
resonance states near the threshold, i.e., Ere, _<_ 2 MeV, then the uncertainty of the
correlation Rune caused by the uncertainty of Te]; (i.e., varying between 5-10 MeV)
is less than 10%.
91
d-alpha correlation d-alpha correlation
2.186 MeV 3* 4 f 2.186 MeV 3"
4 .
3 - v,/v0=2.5
3
CC 5
3 CE 2 +-
4.31 MeV 2+
2 r
M»m..---_r -1 - 1 r
awn.— ._...._
1 I i o l l I l I A [H
4 6 8 1O 0 0.2 0.4 0.6 0.8 1 1.2 1.4
Erel (MeV) EmI (MeV)
Figure 5.16: The left panel shows the experimental d-a correlation function. The
solid, dashed and dotted lines are the different backgrounds used for the estimation
of uncertainty. The right panel exhibits the fit of the first resonance peak after
subtracting the background (solid line in the left panel).
Another imcertainty factor is the Coulomb correlation shape. We have used
particle-correlations in other channels that display fewer resonances to assess the
shape of the Coulomb correlation for interesting cases. For example, the p-d, and
t—t correlations have no resonances. Either may be used to estimate the background
of the p-t correlation. When we pay concentration on strong peaks, we also signifi-
cantly reduce the sensitivity to the Coulomb correlation uncertainties. Nevertheless,
this remains a significant potential source of uncertainty, which we must assess by
considering other shapes for the Coulomb correlation.
In the left panel of Figure 5.16 for the d-a correlation, we show three different
backgrounds (solid, dashed, dotted lines) that we have considered in order to assess
the uncertainty in the peak. In the right panel, we Show the nuclear correlation of
the first excited state of 6Li after subtracting the background correlation denoted
by the solid line in the left panel. The remaining correlation function can be di-
92
Table 5.3-
adopted frt ,
tion.
rectly comp
this spectru
Table 5.3. A
parameter. t
tion. We. fin
lfflb : 29]
nuclear sat in
certainty in ‘
this places 0
Z < 3 to see
The trito
States of 7Li I
7.")
3 at. 9.67
(Solid, (1
étSher
in the right y
extract the l;
J7r E” (MeV) Fm“, (MeV) Channel BR. (%)
6L1 3+ 2.186 0.024 d—o 100
2+ 4.31 1.7 d-o' 97
7Li 7/2- 4.63 0.003 t-a 100
.5/2— 6.68 0.88 t-a 100
5/2- 7.46 0.089 t-a 18
7/2- 9.67 0.4 t-a 40(fit)
3/2- 0.85 1.2 t—o 40(fit)
7Be 7/2’ 4.57 0.175 3He-a 100
5/2- 6.73 1.2 3He—a 100
8Be 0+ gs. 6.8eV 01-04 100
2+ 3.04 1.5 a-a 100
Table 5.3: Relevant spectroscopic information of 6Li, 7Li, 7Be and 8B0 which is
adopted from ref. [98] is listed for the correlation functions discussed in this sec-
tion.
rectly compared to the nuclear correlation Rnu.c(Erel). We have done so by fitting
this spectrum with Eq. 5.23 and using the known structural information, given in
Table 5.3. Applying Eq. 5.23 without any further considerations, we have only one
parameter, the free volume Vf, that can be varied to reproduce the correlation func-
tion. We find that the correlation is well described with free volume of the breakup
Vf/VO = 2.50 d: 0.27, where V0 is the volume of the total system (326 nucleons) at
nuclear saturation density. Uncertainties in the background constitute the main un-
certainty in the extracted volume. Before discussing the constraints in density that
this places on the breakup, we examine other correlations between fragments with
Z < 3 to see whether any consistent trends emerge.
The triton—alpha correlation function is shown in Figure 5.17 with the resonance
states of 7Li (see Table 5.3): 7/2‘ at 4.63 MeV , 5/2‘ at 6.68 MeV, 5/2‘ at 7.46 MeV,
7/2“ at 9.67 MeV and 3/2‘ at 9.85 MeV. In the left panel, different backgrounds
(solid, dashed and dotted lines) are plotted for estimating the uncertainties. As shown
in the right panel, the very pronounced first resonance above the threshold is used to
extract the breakup volume and this gives a value Vf/VO = 2.55 :i: 0.28 that is very
93
t-alpha correlation t-alpha correlation
2.5
1
2 c 4 63 MeV 7/2‘ V,/V0 = 2.55
6 68 MeV 5/2' 2
l | 7.46 MeV 5/2' 0'8 4.63 MeV 72‘
1 5 - a, I 9.67 MeV 7/2‘
[I ' 3' 1 [9.85 MeV 312' 8 0.6 . ,
+ ; ‘ ~
‘— 7’, "\ now-v . l m
w foMm ‘ ‘-‘M:n-'/.W--v L ‘- .
0.4 2
0.2 .
f
' lllllflmttftw “WWW
O I I l J— l l l L l 1
O 2 4 6 8 1O 0 0.5 1 1.5 2 .5 3 3.5 4
Erel (MeV) Erel (MeV)
Figure 5.17: The experimental t-o correlation function is shown in the left panel. The
solid, dashed and dotted lines are the different backgrounds used for the estimation
of uncertainty. The right panel exhibits the fit of the first resonance peak after
subtracting the background (solid line in the left panel).
similar to the one obtained from the d-a correlation function.
Similarly the resonance states of 4.57 MeV 7/2‘ and 6.73 MeV 5/2‘ (see Table
5.3) of the mirror nucleus 7Be are depicted in the 3He—alpha correlation in Figure 5.18.
The 5 / 2‘ state at 7.21 MeV is not shown due to a large decay branching ratio of about
97% through the p—GLi channel. The rise close to the threshold most likely comes from
the contaminant in the 3He PID from alpha particles and what one sees is the ground
state of 8Be from the alpha—alpha correlation (see below). Nevertheless, it does not
detract from our efforts to extract the free volume, which yielded Vf/VO = 2.22 :t 0.36
from the fit to the resonant state at 4.57 MeV.
In the case of the alpl'ia-alpl'ia correlation as shown in Figure 5.19, the ground
state 0+ of 8Be and its first excited state 2+ at 3.04 MeV are illustrated (see Table
5.3). In addition, the peak at about Ere, : 0.6 MeV is mainly from the decay of the
2.43 MeV state of 9Be [81]. In this example the identical particle effect is observed.
94
1+R
Figure 5.18
Solid. daslu
0f uncertai
SlIl')ira(‘llll“
’l
1+R
3He-alpha correlation
0.6 *
4.57 MeV 7/2’
1.5 ~ 6.73 MeV 5/2'
x it I 0.4 ‘
n:
0.2
0 l l l 0
O 2 4 6 10
Erel (MeV)
3He-alpha correlation
4.57 MeV 7/2'
v,/vo = 2.22
l + H +
[Lilli ll 1"] i
Hffil‘l
0
1 2 3
E... (MeV)
Figure 5.18: The left panel shows the experimental 3He-a correlation function. The
solid, dashed and dotted lines are the different backgrounds used for the estimation
of uncertainty.
subtracting the background (solid line in the left panel)
1+R
alpha-alpha correlation
, 25
9.5. 0
2.5 --
20
2 r
l 98e'(2,43MeV) w 15
1 .5 fl 3.04 MeV 2’ (ES
10
5
0 1 I l
0 2 4 6 1O
Ere, (MeV)
alpha-alpha correlation
The right panel exhibits the fit of the first resonance peak after
9.5. 0‘
+
V'IV0 = 2.61
+ J l J
Ere, (MeV)
0
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4
Figure 5.19: The experimental o-a correlation function is shown in the left panel. The
solid, dashed and dotted lines are the different backgrounds used for the estimation
of uncertainty.
subtracting the background (solid line in the left panel).
95
The right panel exhibits the fit of the first resonance peak after
That is. th.
consequent .
By taking .
hunt the 3n
In all t.
this study.
temperatui
test t0 see i
I0 make tl
by the pro]
rcsenance ]
given belov
Performing
”I Fig- 5.21
excitation
essent ial a:
with T6” :
over this ‘
0f the ble-
lif I/l‘e : D
-
That is, the phase space of the two identical particles is reduced by a factor of 2 and
consequently the resonance correlation Rm“, becomes twice as large as in Eq. 5.23.
By taking into account this effect, one obtains a free volume of Vf/Vo = 2.61 i 0.03
from the ground state of unstable {Be as shown in the right panel of Figure 5.19.
In all the above correlations, only the first most pronounced peak is fitted for
this study, which is sensible in reducing the sensitivity to the uncertainty of the
temperature factor and the background. Nevertheless, it is an interesting consistency
test to see if we can fit all the peaks we observed. In figure 5.20, we explore this issue.
To make these higher peaks more visible, we divide the nuclear correlation function
by the prefactors in Eq. 5.23. This leaves only the free volume and the density of
resonance states coming from the derivative of the strong interaction phase shifts
given below
3 ‘1
1 . h (Era/W (5.25)
(2J1 + 1)(2]2 + I) 47TVO/LV (ZIMETCJ
1 1 r./2
= —— 2.1. 1 . .
vf/v0 7r 2,] + )(E,.,_., — E:)2 + r3/4
Rnuc(Erel) '
(13.11.) E t;,/V,,),.,.,(E,.e,).
Performing this operation on the correlation function data provides those data points
in Fig. 5.20. The solid lines are the sum of the resonance line shapes calculated using
the parameters in Table 5.3. To make a good representation of these peaks at higher
excitation energies, a careful normalization of the Coulomb correlation background is
essential as is the correction due to collective motion corresponding to the exponential
with T6,, = 7 MeV. In other words, fitting over the domain of these higher resonance
states is useful to constrain the background and the parameter Teff. After fitting
over this extended excitation energy range, the numbers obtained for free volume
of the breakup are changed very slightly: ’f/VO = 2.63 for the d—alpha correlation;
Vf/VO = 2.79 for the t-alpha correlation; V f / V0 2 2.44 for the 3He-alpha correlation;
96
d-alpha correlation
VO/Vi pres(Erel)
2.186 MeV 3+
30 r
q V/V0=2.63
we
v3 20 -~
Cf
3°
10 r 4.31 MeV 2”
O ...-"' 1 1 1 L 1H
0 0.5115 2 2.5 3 3.5 4
Ere, (MeV)
3He-alpha correlation
125 P 4.57MBIV7/2‘
10 ‘
u} V/Vo=2.44
g 7.5
Q.
:0 5 L. 6.73MeV5/2'
2.5 1. i” .
. ,, if
0 ++x+fi++§+ H’ 1 1
0 1 2 3 4 5
Erel (MeV)
VO/Vf pres(Erel)
t-alpha correlation
15 -
4.63 MeV 7/2'
V/vo=2.79
10 l
6.68 MeV 5/2‘
l7.46MeV5/2'
9.67 MeV7/2'
9.85 MeV3/2’
5 ~ I
l
I! l J i
O 1 2 3 4 5 6 7 8
E (MeV)
alpha-alpha correlation
20 g.s.0‘
15 , V/Vo=2.54
10 ~
5 h °Be'(2.43MeV) 3'04 MW 2‘
‘ 1
0 V i 1 '
0 1 2 3 4 5
Ere, (MeV)
Figure 5.20: The densities of the resonance states from d-a, t-a, 3He—cr and CH}
correlations are fitted after subtracting the backgrounds carefully selected by the
solid lines in Figures 5.16-5.19.
Vf/VO = 2.54 for the alpha-alpha correlation. The free volume measurements are
given in Table 5.4.
The cm‘isistency between the 1')reli1'ninary numbers given for the free volume in
the ”VI/V0” column in Table 5.4 is surprising. However, further estimates on the
breakup density need the consideration of corrections from collective motion and
from the secondary decay of heavier particle unstable isotopes that can produce the
particles in the correlation as well as nuclei in the resonance states we observe. Here we
attempt a correction for such effects. To consider the effects of collective motion, we
need to revisit the discussion in Section 5.3 on collective motion where the reduction
of the source size due to the interplay of collective and thermal motion is discussed.
The problem we need to solve is more complex than the problem we needed to solve
in order to determine the density from the p—p correlation function and from the
second peak in the d-a correlation function. In those latter cases, the shape of the
resonance was sufficient to determine the volume of the source while one needed the
simulation to determine the mass in the source. Here we need to consider how the
collective motion influences the magnitudes of both the resonant, non-resonant and
mixed event yields. (The previous analysis showed that the latter two are influenced
similarly. )
Generally speaking, rearranging Eq. (5.2) yields the correlation function 1 + R(q)
expressed as a ratio of 5 dimensional integral over the coincidence yield divided by a
corresponding 5 dimensional integral over the mixed event yield product. If we try a
Monte Carlo simulation as in Section 5.3 of these yields, we break down the spatial
origins of the contributions to the momentum distributions. This means that starting
from these phase space distributions at breakup, we need to do an 11 dimension inte-
gral: 3 spatial and 3 momentum integrations for each particle minus one integration
corresponding to the non-integrated dependence on E,.._.,. Collective motion builds
98
correlations between the emission points and momenta of the emitted particles that
effectively reduce the fraction of the source contributing to the correlation function.
This reduction factor, however, is depem'lent on EN, as the exponential dependence
in Fig. 5.11 demonstrates.
Fig. 5.11, however, suggest. a simple way to deal with collective motion. Let us
assume that the data are consistent. with T.” = 7 MeV. Thus, for the source volume
of this simulation and this specific collective velocity field, the resonant correlation
function should be proportional to the middle panel. Without collective motion, the
same source volume leads to the lower panel. The ratio of the middle panel divided
by the lower panel is what one would expect for the ratio of correlation function from
a source of the same dimension with collective motion divided by the correlation
function without it. This ratio is independent of the actual volume of the source.
Like the volume reduction factors in Fig. 5.9, it depends 011 vhmax, otmax, and the
random velocity 1’11; = y/3T/m, but not on the source radius R. Thus the inverse of
this ratio frat,” provides a way to correct for the collective motion effects and obtains
what the correlation function should be without collective motion. Fig. 5.11 shows
what the correction factor frat“, should be for the d-a correlation function in figure
5.15.
In applying this c<:>rrection factor, we simply need to extract the value for the cor—
rection factor flan-(,(Eml) at the energy of the resonance in question. This dictates the
factor by which the correlation functions would have been smaller without collective
motion. This smaller correlation function would have resulted in the extraction of a
new free volume that would be larger by a factor of fm,,-O(E,.el)’1. The production
of heavier particle unstable nuclei and their decay into the particles included in the
correlation function or into the observed resonance peak also alter the correlation
function by increasing the yields of the involved nuclei. It is the correlation function
99
of primary fragments before decay which has the simple dependence on the emission
volume predicted by Eq. 5.19. Preliminary calculations of secondary decay suggest
that the shape of the spectrum after decay is not very different from the shape before
decay. Under this assumption, the primary correlation functions before decay differ
from the corresponding ones after decay by a multiplicative factor Elem where in the
case of the d-a correlation
(5.26)
where the factor DRd reflects the multiplicative factor by which the yield for deuterons
is enhanced by secondary decay, and the other factors in Fdec reflect the application
of the same considerations to the yields of 0’s and to the resonance in 6Li as well.
Secondary decay correction factors were obtained by calculating the equilibrated mul—
tifragment decay of a system consisting of 260 nucleons and 106 protons at an exci-
tation energy of E*/A =4 MeV using the SMM model described in the next chapter
of this dissertation. The secondary (‘lecay factors for the correlation functions shown
in Figs. 5.16—5.19 are given in the third column of Table 5.4.
we note that these secondary decay corrections are rather large. In particular, a
value for Fdec of 6.09 is obtained for the d-a correlation, which should be compared
to the reciprocal of the value for A = 0.11 that was obtained by fitting the shape for
the second peak of the inclusive d—a correlation in Section 5.3. This suggests that
the SMM predicts about the same (differ by 49%) secondary decay contributions as
the Koonin-Pratt approach. To get a feeling for what a 49% correction implies, we
increase all the Fdec factors in the third column of Table 5.4 by 49%. These modified
values of Fdec(KP) are also shown in Table 5.4. These modified values should be
viewed with caution beyond some indication of how sensitive the densities are to the
secondary decay corrections.
100
Based 011 the above discussions. a correlation function Rnucpri without collective
motion before secondary decay can be obtained by an‘pr, ~ Rum - E1“ . fmm, (EM).
Since the factor exp(—E,.(,1 [11,”) consistent with Fig. 5.11 has already been taken into
account in Eq. 5.23 in extracting the free volumes for the correlation functions shown
in Figs 5.16-5.19, the normalization of the exponential factor frauofl :—: f,.(,,,-0(En.1 = 0)
should be directed out to correct the free volumes for the remaining effects of collective
motion. Therefore an estimate of the expanded free volume after corrections for
secondary decay and collective motion might. be given by
Vf
V .source 1: _~ 527
f chc ' fratio.0 ( )
and the freezeout density p/po by
p
P/Pn = +—’—— 15.28)
Pf + 100
where
A we runner) 0-8fra io.f ' Flier:
mm, = —,’— = ’ ’ (5.29)
fgsourcept) L'f/f’f)
where the correction factor fmhoy) = 0.15 is obtained from Fig. 5.11 for the d—a
correlation and for simplicity it is used for calculating the densities for the correlations
shown in Figs. 5.16—5.19 assuming they all have the similar collective correction
factors. We note that they all have similar effective temperature of T8” = 7 MeV.
Values for the density, in units of saturation density p0 are calculated from Eqs.
5.28-5.29 and given in Table 5.4, indicating that the freezeout density of the system is
about 1/5-1/3p0. For comparisons, the density values of p/ p0(K P) are also shown by
applying the source reduction factor f2,” = 0.2 obtained for the Koonin-Pratt analyses
instead of the factor flan-0,0 discussed here, and using the modified secondary decay
factor EMU/(P). If we omit the contributions of the excluded volume as in the case of
101
the Koonin—Pratt approach and apply the same secondary decay and collective motion
corrections as the results from the Koonin-Pratt approach. the pf/pn(KP) should be
comparable to the density obtained from the Koonin-Pratt approach in Sect. 5.3.
For the do correlation, indeed, the value of the density (pf/pUUX'P) = 0.57 in Table
5.4) extracted in the equilibrium limit are similar to the value (p/po = 0.54 in Table
5.2) extracted from the Koonin-Pratt analysis. Such a conclusion is in agreement
with the work by Jennings et al. [95], which equate the thermal and Koonin—Pratt
approaches for Gaussian sources under the assumptions of instantaneous emission
and no corrections from secondary decay and excluded volume.
The density extracted from the equilibrium approach has similar large uncertain-
ties as in the Koonin-Pratt analysis, which have been discussed in the end of Sect.
5.3. Especially the estimates of the correlation enhancement factor f,.at,,,(E,.€,) due to
collective motion are in need of better accuracy. On the other hand, the secondary
decay process significantly modifies the extracted free volume in the opposite direc-
tion. The resulting correction factor Fder: also need to be studied more carefully by
accurately checking the modelling of secondary decay by the statistical model.
Correlation Vf/VO Em, E16,.(KP) pf/po p/po pf/pn (KP) p/p0(KP)
d-o' 2.50 :t 0.27 6.09 9.09 0.29 0.23 0.57 0.37
t—o' 2.55 :1: 0.28 7.09 10.6 0.33 0.25 0.66 0.40
3He-(r 2.22 d: 0.36 8.45 12.6 0.46 0.31 0.90 0.47
0-0' 2.61 :l: 0.03 4.76 7.11 0.22 0.18 0.43 0.30
Table 5.4: List of quantities are shown for the d-a, t-a, 3He—a and 01-0 correlations
(Figs. 5.16—5.19). The values of Vf/VO are obtained from the corresponding experi-
mental correlation functions. chc is the correction factor in Eq. 5.26 calculated from
secondary decay contributions in the SI\«"Il\-~'I. The density p/po is estimated in Eq. 5.28
using the secondary decay correction factor Fdec from the SMM while p/p0(KP) is
obtained using Fdec(K P) (see text). The corresponding pf/po and pf / p0(K P) are
calculated in Eq. 5.29 without considering excluded volumes.
Besides the density extraction, it is interesting to see via the above multi-peak
102
fittings that. unknmvn spin of one resonziu‘ice can possibly be obtained if the structural
information is well known for the other states. In the following section we will discuss
how to use this technique to determine spins of particle unstable nuclei.
5.6 Spin Determination of Particle Unstable States
One of the striking observations of this work is the degree to which the equilibrium
assumption can describe the relative magnitudes of the various peaks in the corre-
lation function. This is illustrated by the good agreement between the calculated
and measured values of the nuclear correlation function in Fig. 5.20. Noting the
dependence of the correlation function on the spin of the state (see Eq. 5.23), we
now explore whether this can be used to determine the spin of a nuclear state, if that
spin is unknown. By armlyzing density spectra. of resonance states like those shown
in Fig. 5.20, we try to determine the unknown values of spins and even unveil new
structures of nuclei.
To ensure the validity of applying this technique on studying the structures of
unbound nuclei, one simple case is examined in the following. In Figure 5.21, the
proton—7L1 correlation function is shown with the resonance states of excited 8Be that
are close to the tl'iresl‘iold: 17.64 MeV 1+, 18.15 MeV 1+, 18.91 MeV 2", 19.07 MeV
3+ and 19.24 MeV 3+ [98]. The branching ratios of these states through the proton
decay channel are as follows [98], the 17.64 MeV state is 100%, the 18.15 MeV state
is 100% where 96% goes to the ground state of 7Li and 4% to the first 1/2' excited
state of 7Li at 0.478 MeV, the 18.91 MeV state is 86%, the 19.07 MeV state is 100%
and the 19.24 MeV state is 50%.
The reason why we choose the p—7Li correlation as an example is the structural
information of 8Be is quite complete and these pronounced resonance states are very
103
p-7Li correlation
17.64MeV 1+
f
1.5
18.15MeV cm
*1
1
18.91MeV 2'
19.07MeV 3+
19.24MeV 3+
+9311,
1’
+H’
1 + + H ““3
+ [#1 i “:1?“ 1’ +++++ff++f+m + + ” + + + + + + + +
1 — {1+1 ____________ {ft Effimt’t +1.”; #1- #41511; ‘bfitfl- *- Mew
Y'- f I I
0.5
O l l l l
2 3
Erel (MeV)
Figure 5.21: The proton-7Li correlation function is plotted. The dashed line is the
selected background. See text for details.
104
close to the threshold, which can help reduce the sensitivity to the background se-
lection. For simplicity, we will parameterize the background correlation function as
follows,
1 + R59 2 1 — exp[—(ET€1/EC)”] (5.30)
which vanishes at zero relative energy and reaches unity at large relative energy. By
comparing with the proton-“Li and proton-8M correlations, the background of the
proton-7Li correlation function (the solid line in Figure 5.21) is selected with parame-
ters EC 2 0.16 MeV and a = 0.5 since these correlations should have similar Coulomb
interaction. After subtracting this background, the density of resonances is plotted
in Figure 5.22. The dotted lines represent the individual resonance states and the
solid line is the corwoluted fit by applying the known structural information of these
resonances. The fit is good everywhere except at the high energy end where the con-
tributions from higher lying states are not included due to incomplete spectroscopic
information.
To further test. the reliability of the background selection and the sensitivity to
the spin factor, a fit to the overall correlation function is carried out by varying
the parameters in the background expression (Eq. 5.30) and the spin value of the
17.64 MeV state. In Figure 5.23, the best fit (solid line) yields a background shown
as the dashed line with parameters of EC 2 0.152 MeV and a = 0.547 which are
very similar to what we have assumed above and a spin value of 1.06 i 0.1 which is
consistent with the experimental value. To illustrate the sensitivity of this technique
to the spin determination, calculated correlations (dotted lines) are also shown in
Figure 5.23 assuming that the spin of the 17.64 l\x‘leV state is 0 and 2, respectively
and meanwhile keeping the other parameters the same. The significant separation of
J = 0, 2 calculations from the J = 1 fit shows a good sensitivity to spin determination.
105
p-7Li correlation
18.91MeV 2'
1 * 19.07MeV 3+
19.24MeV 3+
17.64MeV 1+ [ [ [
”7:, 0.75 ” [ [
18.15MeV1+
0.5 T [1H 1‘]
const preS(E
D
O
.l
Figure 5.22: The density profile of the resonances of 8Be is shown in the p—7Li corre-
lation after subtracting the background selected in Figure 5.21.
106
p-7Li correlation
17.64MeV 1+
18.91Mevz;
.; 19.07MeV 3+
15 _ 53 19.24MeV3
- 18.15MeV? [ I [
O
- +
. o -------------------------- Jbtt nt‘]
0 l l 1 l
0 0.5 1 1.5 2
Erel (MeV)
Figure 5.23: The best fit (solid line) is performed for the p-7Li correlation by varying
the background and the spin value of the 17.64 MeV state. The dashed line is the
fitted background. Two calculations are shown as the dotted lines assuming that the
spin of the 17.64 MeV state is 0 and 2, respectively and keeping the other parameters
the same.
107
As (1111‘
smnhdma
dwwnfor
obvious [1
state at 2
stzitc 1111s
.\le\" stat
5Li. a 1+
iMXC}Pt[
tlu?inter(
functhn1,
a 8D111 \"dl
Houe‘
humthuf
buiknov
Infigun
sphivahi
lhuflissl
neonanc
Syfienir
finnihul
Welnay
SDinoft
C‘dSeS’ th
of other
experhiu
As one can see from the above example. this technique can provide access to the
spin information of resonances. 111 Figure 5.24. the proton-TBe correlation function is
shown for an example with an unknown spin to be determined. In this (’tase the two
obvious peaks correspond to the first excited state of 8B at 0.774 MeV and the 3+
state at. 2.32 MeV, respectively [98]. Interestingly the spin value for the 0.774 MeV
state has not been measured although it is 1,)elieved to be the 1+ analog of the 17.64
MeV state in 8Be [98, 99]. Also from the corresponding state in the mirror nucleus
8Li, a 1+ assignment would be expected for this state. However, experimental efforts
have yet. to justify that statement. If we assume that only these two states exist in
the interesting region, the solid line in Figure 5.24 is the best fit to the correlation
function, which fits a background (dashed line) along with the resonances and yields
a spin value of J 1 = 0.98i 0.29. The result confirms the hint from the mirror nucleus.
However, the fit does not present all the features of the experimental correlation
function. Especially a small bump at about 1.4 MeV seems to exist in between the
two known states. The calculations in ref. [100] predict a 1+ state of 8B at 1.4 MeV.
In Figure 5.25, the three resonances including the one at 1.4 MeV are fitted. A similar
spin value J1 = 0.95 i033 of the first state is extracted. But the background (dashed
line) is sluiillower than that of the previous fit to accommodate the additional 1.4 MeV
resonance. If we can reduce the background by 111easuring a reaction with a small
system or constrain more strongly the background contribution to the correlation
function in some way and have better statistics and resolution in the correlation,
we may determine further if there indeed is a state at 1.4 MeV. Nevertheless the
Spin of the first state at 0.774 MeV is confirmed to be one in either case. In both
cases, the underestimated tail of the 2.32 MeV state seems to indicate the existence
of other nearby states at slightly higher energy which have not been identified yet in
experiments. However, we don’t have the resolution to distinguish these states.
108
p-7Be correlation
spin unkown
1.5 —- 0.774MeV
2.32MeV3”
+
- ++
CE 1 F ‘ + _______ .Jf .............. Jr—Hr
i V ,,,,, +~ ‘
0.5 c ff
0 ' L J I 1
0 0.5 1 1.5 2 2.5 3
Erel (MeV)
Figure 5.24: The p—7Be correlation function is fitted by the solid line assuming only
two states at 0.774 MeV and 2.32 MeV. The dashed line is the fitted background.
109
P-7Be correlation
Spin UnkOWn
1.4Mev 1+
1 ‘ ++++.
1 f— k
a: + " _____________ w
+ I .................
‘— + j ,,,,,,,
0.5 7- "I,
0 . 1 1 I I
0 0.5 1 1.5 2 2.5 3
Ere! (MGV)
Figure 5.25: The p—7Be correlation function is fitted by the solid line assuming the
existence of an additional state at 1.4 MeV. The dashed line is the fitted background.
110
01-6Li correlation
B.R.=100% (B.R.=55°/o)
[ 3 ...--..,/J =4
'.'-. ' .............. x/J =3
. ------- . J=2(J=4)
J=1
P O ..
‘ O
. .§ ‘I. o
I
~ ' . ‘.. \
[_. .4 Q '
~~ ..C ‘-
d
b ~~ ..
~
~ . 0..
~~ o
1.4 T F‘s .'
s '0'
C
1- “
.
2.2 2.4
1+R
(.0
O 0.5 1 1.5 2 2.5 3
Ere,(MeV)
Figure 5.26: The a-SLi correlation function is fitted by taking into account all the
spectroscopic information shown in Table 5.5 except for the 6.56 MeV state. By
varying the spin of the 6.56 MeV state from 1 to 4 and assuming the decay branching
ratio is 100%, one obtains the dashed, solid, dotted and dot—dashed fitting lines,
respectively. If a branching ratio of 55% is assumed, the solid line represents a fit of
J=4.
111
In the following case. the c1-6Li correlation is shown with ample resonance states
forming peaks near the decay threshold. The structural information of 10B resonances
is listed in Table 5.5 where we notice that the spin of 6.56 MeV state is not well
measured (J24 is tentatively suggested [98]). Figure 5.26 shows the fit with the
known structural information except for the state at 6.56 MeV. The spin J =1 of the
7.002 MeV state and its B.R. of 30% (see the d-a-a correlation in sect. 5.7) are
used in the fitting, which have little effects on determining the spin and B.R. of the
6.56 MeV state. In this fit we vary the spin value of this state from 1 up to 4 while
assuming that 100% of this state decays to the ground state of 6Li. As we can see,
it seems J =2 gives the best fit while the cases of J21,3 are still possible due to poor
resolution. On the other hand, if J = 4 is assumed for the 6.56 MeV state, then a
branching ratio of this state decaying to alpha—6U channel can be obtained as about
50%.
.17r 13* (MeV) Fem, (keV) B.R. (%)
3+ 4.774 8.4x10-3 100
2‘ 5.1103 0.98 100
2+ 5.1639 1.76x10‘3 13
1+ 5.180 110 100
2+ 5.9195 6 100
4+ 6.025 0.05 100
3‘ 6.1272 2.36 97
(4)"? 6.560 25.1 100‘?
1“ 6.873 120 38
(1,2) + 7.002 100 small
2" 7.43 100 <30
2+ 7.478 74 <35
Table 5.5: Spectroscopic information of 108* ——> a+6Li adopted from ref. [98]. The
state at the 6.56 MeV is assigned tentatively with J=4 and only alpha decay is
confirmed. The last three states contribute very little to the fitting. For details see
text.
An extra 3+ state is predicted at about 6 MeV in ref. [101]. If this additional
3+ state exists and all states are populated with their full statistical weights, we
112
must. have overestimated that background in the fit above. If instead, one lowers
the background to accon‘miodate the 3+ state, the fitted spin value for the 6.56 MeV
statecould be a little higher. But it still can not account for a spin of J=4 if we
assume that the branching ratio to the a-GLi channel is 100%. In the next section,
we see some weak evidence of a resonance at Em, : 0.6 MeV in the deuteron—cr-a
correlation indicates that some of the 6.56 MeV state decays via either the d-cr-oz
three—particle channel or the d-8Be two-particle channel. If it accounts for half of the
decay of the 6.56 MeV state, then the best fit (solid line) in Figure 5.26 favors a spin
of J = 4. We don’t have enough statistics, however, to determine the branching ratio
for the decay of the 6.56 MeV state to this channel, leaving the determination of spin
for the 6.56 MeV state open at the present time. Further studies 011 the branching
ratios are needed to clarify this issue.
To further apply this technique on probing nuclear structures, better statistics
and resolution are important to distinguish overla.11)ping resonances. Better statistical
accuracy and lower backgrounds can help reduce the uncertainty in the fit. Lower
backgrounds can be attained by using a smaller system. However, the agreement with
equilibrium correlation functions could be worse because there may be stronger non-
equilibrium contributions. Resonance peaks far from the threshold are suppressed
significantly due to the energy dependent suppressing factor in Rnuc which comes
from two particle phase space and collective effects. Therefore the states close to the
decay threshold are easier to study by this technique. As far as the secondary decay
effects concerned, proton—rich nuclei are preferred in this technique. In addition, all
the decay branching ratios, if not known, for the low-lying resonances of proton—rich
nuclei can in principle be determined by studying all relevant particle correlations.
In this case we don’t need to deal with neutron decay which can be omitted for
proton-rich nuclei. It will be an advantage when only charged particles are detected
113
in experiments.
5.7 Multiple Particle Correlations
Similar to two particle correlation functions, correlation of three or more particles can
be constructed in heavy ion reactions. W hile we have not analyzed them in detail, it
is nevertheless interesting for future studies to examine them. Choosing some of the
stronger correlation functions, we discuss some of their properties in this section.
In Figure 5.27 the three-alpha correlation function is shown. The resonance peaks
of 12C [102] are labelled in the plot. In two particle correlations one sees a weak energy
dependent factor 1/\/E,.el in Eq. 5.23 which stems from the two-body phase space.
In contrast to two particle correlations, three body phase Space yields a more strong
and singular energy dependent factor in three-particle correlation functions. For a
three body decay, this factor can be as strong as l/Efd and consequently increases
the height of resonance states well below 1 MeV enormously and suppresses the peaks
at higher energies. Due to this singular factor the first peak is extremely pronounced
corresponding to the 0+ state of 12C at 7.654 MeV. The second peak showing the
3‘ state at 9.64 MeV is much reduced compared to the first peak even though it
has a larger spin degeneracy factor. The states of 0+ at. 10.3 MeV and 1“ at 10.844
MeV are less evident because they are much broader and have low spins. A broad
pronounced group is observed consisting of five states of 2+ at 11.16 MeV, 2‘ at
11.828 MeV, 1+ at 12.71 MeV, 2" at 13.352 MeV and 4+ at 14.083 MeV, which can
not be distinguished from each other due to poor resolution. Beyond the broad 2+
state at 15.44 MeV the proton decay channel becomes dominant.
111 Figure 5.28 the proton-proton-alpha correlation is plotted depicting the reso-
nances of the 0+ ground state and the 2+ state at 1.67 MeV from 6Be [98]. Figure
114
5.29 shows the proton—a11.111a—al})lia correlation function from the decay of 9B. The
first resonance of the 3/2" ground state of 9B is so pronounced because of the singular
three-body phase space factor (liscusst—ét’l above. And the 5/2‘ state at 2.361 MeV is
also pronounced while there is a weak evidence showing the existence of a broad state
at about 1.6 MeV [98].
In Figure 5.30 the deuteron-alpha—alpha correlation function shows the resonance
peaks from the decay of 1”B in the range of 6-8 MeV [98]. Although poor statistics,
one can see a suggestion of a resonance from the 6.56 MeV state to this decay channel,
which is just. above the detection threshold. This might indicate that the branching
ratio for the o-GLi channel from the 6.56 MeV state is less than 100%. Thus there
is a possibility that the spin of this state could be larger than the magnitude of the
peak in the (if-6L1 channel might indicate. The second peak in Fig. 5.30 could be the
decay of the 1‘ state at 6.873 MeV and the (1, 2)+ state at 7.002 MeV. A shoulder
on the spectrum corresponds to the decay from the 2~ state at 7.43 MeV and the 1‘L
state at 7.67 MeV and there is a pronounced peak showing the decay of the 2+ state
at. 8.07 MeV.
As well as identifying resonance states of particle unstable nuclei, we can gain
access to other valuable information in correlatitm functions. which is difficult to
be obtained elsewhere. The three-body Coulomb correlation can be examined in
these three particle correlation functions. Although the resonance states far from
the threshold are difficult to access due to the large phase space suppressing factor
1/E3d, this factor also makes the peaks very close to the threshold astoundingly
pronounced, which may be useful for further studies on the properties of sources and
so on. However, more needs to be done 011 the formalism for three body correlations
before this can be a quantitative tool.
alpha-alpha-alpha correlation
12
C 100 ~
. l 7.654lvleV
1116Mev
2.5 7’ ‘1828MCV
12.71MeV -
13.352Mev 50 _
14.083MeV
—L I l-
” [ [1+
1 +111 .
0 ......... tL +4)
m 2. [ [l] +++MW++++*+*+ m 0 0.5
+ 1.5 [ [ [#11 f +23 ”,4;
‘— 1 1 f 1 m, 1 Wffwowm 4
[ H [1 1'1 ++ fwflW’”’*’“’°-M’Jw
1 _ [[1] [_J 15.44MeV
.. [ 10.3Mev
10.844lv1eV
0.5 r
T7654Mev
O l l l l l l l
O 2 4 6 8 101214
Erel (MeV)
Figure 5.27: The a-oz-a correlation function is shown. Resonances from the excited
states of 12C are labelled with the first peak seen more clearly in the inner upright
panel.
116
1+R
proton-proton-alpha correlation
1.5
1 _
0.5 c
]
11*
l
0+
1“+
F 9.3.
1.67MeV 2+
+++,,++++++
H++++++ when: w.»
+
1
6Be
4* + «4+ n *
9 + 0".“ *N“”§ H, ”0.. I”O.+N«
L
Figure 5.28: The p-p—a- correlation function is shown.
2 4 6
Erel (MeV)
117
8
1O
12
p-4He-4He correlation
QB 9.8. (3/2)’
I I t
60
I
(.—
2-5 W 9.8. 40
Ur'TUrTI’I
(5/2)' 20
2 “’ 2.361MeV 03.11.}?
1+R
01
0 1111111
012345678
Erel (MeV)
Figure 5.29: The p—(I-a correlation function is plotted. The resonances from the decay
of 9B are shown.
118
d-oc-oc correlation
3
CD CD (41316.56 MeV
® 1'at6.873 MeV
2.5 *- l®® @ ® (1,2)+ at 7.002 MeV
I] @ @ 2‘" at 7.43 MeV
@ © (1*) at 7.67 MeV
2 ~ I <6) 2* at 8.07 MeV
++
qI" 1 5 T [+1 ++++++++W++ ++ +
‘— [ + + ++ +++W++N+++++fiw
1 _ ii
0.5 _
O l 1 1 l
O 2 4 6 8 10
Erel (MeV)
Figure 5.30: The d—a-a correlation function is shown. The resonance states of 10B
that contribute this decay are listed. See detailed discussions in text.
119
Chapter 6
Statistical Multifragmentation
Model with Empirical
Modifications
The statistical multifragmentation model (SMM)[36] has been successful in describing
the thermmlynanrical properties of multifragmeiitation processes, especially for corn-
parisons to measured fragment multiplicities, mass and charge distributions and so
forth[19, 21]. Further studies on the degree of thermalization of such reactions involve
accurately constructing thermometric observables such as isotopic thermometers[54,
31]. Isospin effects in multifragn1entation are also being explored for the study of
isospin fractionation[41], isotope scaling[39] and even the isospin dependence of the
nuclear equation of state[42]. Such studies, however, have been rendered less conclu-
sive because of the inaccurate modelling of the later stages of the breakup in current
models where detailed nuclear structural infornu-ition are critical to accurately model
the secondary decay process. In this chapter, an empirical statistical multifragmen-
tation model incorporated with an empirical secondary decay procedure is described
by taking into account experimental information such as binding energies and level
densities as much as possible. By means of this improved model, nuclear thermometry
and isospin effects are studied and compared with experimental data.
120
6.1 Microcanonical Statistical Multifragmentation
Model (SMM)
6.1.1 Underlying Formalism
In the SMM, it is assumed that an excited, expanded and equililn‘ated source is
formed after the most violent stages of a heavy ion reactitm and it then decays simul-
taneously and statistically. The Inicrocanonical version of the SMM uses the Monte
Carlo method and averages oliiservzitbles with the statistical weight over fragmentation
1nodes[3(5]. A multifragment decay mode is defined in the SMM approach as a specific
set of emitted fragments and light particles. For simplicity, each fragmentation mode
in the SMM approach is weighted according to the entropy of the mode. This entropy
is approximated by analytical expressions rather than by an event by event sampling
of the phase. space as in ref. [11]. These approxinlations rely upon that fact that the
domiimnt contribution to this entropy comes from internal phase space of fragments
which plays the role of a heat bath within the. SMM approach just as an excited
residue plays the role of a heat bath within compound inn-lear decay theory [43].
The characteristics of this source, such as its thermal excitation energy E *, density,
mass and atomic numlmrs, A0 and 20, respectively, are taken for granted in the model.
For a given fragmentation mode, mass and charge conservation is strictly imposed:
A0 = Z NAZ A and Z0 2 : NAZ Z (6.1)
{AZ} {AZ}
where NAZ is the multiplicity of a fragment with mass number A and Charge number
Z. By making a Wigner Seitz approximation to the Coulomb energy, energy conser-
vation within the SMM approach leads to the expression [36]
, , Z2 v ”3
Eg‘ + O 2 (1C—% (V?) + Z IVAZEAZ (6.2)
A0 {AZ}
121
where E85 is ground state energy of the fragmenting source. The first term on the
right hand side stands for the Coulomb energy of a homogeneous sphere of matter
containing the total charge ZO and mass A0 at a (‘lensity p 2 p0 (Hy/V) where if) is the
normal volui'ne at saturation density [)0 and V is the, breakup volume occupied by the
system. The second term is the sum of the energy of each individual fragment in this
decay mode. Enz is the kinetic plus internal energy for each of these fragments. It is
related to the temperature by assuming all fragments are at a common temperature
as follows,
EM = 3T + 13pm + E552 — Big (63)
where the internal excitation energy of the fragments, 32(T), may be approximated
by an extension of the. semi—ei'npirical mass formula to finite temperatures [36], and
the extra Coulomb energy of the fragment in the fragmentation volume, Effz, may
be calculated within the VVigner-Seitz approximation (see below). B AZ stands for
the ground state binding energy for the fragn‘ient. Eqs. 6.2 and 6.3 result from an
average of the niicrocanonical expression for energy conservation over the phase space
corresponding to the specific fragmentation mode.
By applying the energy conservation relationship in Eqs. (6.2-6.3) one obtains
a ten’iperature T that describes the internal excitation and translational energies of
fragments within a given fragn'ientation mode. Even though the overall system is
assumed to be in equilibrium at a fixed excitation energy E5, different decay modes
have different Coulomb, binding, and translational energies and, consequently, dif—
ferent excitation energies of the emitted fragments. Consistency with Eqs. (6.2-6.3)
therefore requires that the temperature T of the fragments varies from one decay mode
to another, reflecting the differences between the Coulomb, binding and translational
energies of the various fragmentation modes.
122
Labelling the fra mentation mode N43 with the index, m, a )hvsical observable
b ..
0, such as the yield of a fragment or the temperature, can be expressed by a weighted
average over decay motles as,
:77] L1If"? ()171
O = 6.4
< > Zn: ”km- ( )
where the statistical weight associated with the mode m,
l’Vm : exp 2 NAZSAZfT) (6.5)
{AZ}
may be found by expressing the entropy of the fragments, SAZ, through the thermo-
dynamical relation with free energy,
8F
8T ’
S = — (6-6)
The corresponding free energy associated with each fragment can be written as,
2‘2 / 1/3
FAZ(T) : —BAZ — €1ch (V0)
+ BMW MT), (67)
where the kinetic term corresponds to:
771,, AT] 3/2
2nh2
+ T111 (1V4Zl) /]V/lZ . (6.8)
Ffzm = —-Tln g,.,zvf[
where V, = V — V0 is the free volume, m" represents the nucleon mass, and 942 is
the spin degeneracy factor. Empirical values at ground state are used for A < 5 since
one assumes that these nuclei have no excited states, except 4He. In all the other
cases, gAZ = 1 because we assume that this effect is, to some extent, already taken
into account by the level density used in the model. The Coulomb term in Eq. (6.7)
123
is associated with the remaining corrections in the \Vigner-Seitz approximation. The
excitation of the intrinsic degrees of freedom is taken into account by F§Z(T) and
therefore it is zero for the light particles as just mentioned.
To calculate the properties of the multifragment (‘nnission from the excited source,
one should sum the contributions of all the fragmentation modes consistent with mass,
charge and energy conservation imposed by Eqs. (GU—(6.2). However, this would be
extremely time consuming owing to the huge number of possible fragmentation modes
[103]. Therefore, the present. approach samples the more probable modes via a Monte
Carlo calculation. This is carefully discussed in ref. [103]; we note in passing that the
l\lonte Carlo procedure introduces a bias since not all the mass and charge partitions
enter with the same weight. Therefore a)", must be modified to correct for this bias
[103].
In the standard SMM formalism, simple paramet.rizations [36] of the ground state
energy and internal free energy are adopted for convenience. In order to have accurate
calculations, however, more careful treatments related to the nuclear structures are
needed, especially for the secondary decay process. Before we introduce an empirically
modified SMM with an consistent empirical secondary decay procedure in section 6.2,
issues related to primary temperatures need to be addressed in the following.
Temperatures have been extracted from the isotopic yields of heavy ion collisions
using the isotope thermometry method proposed by Albergo et al. [54]. The idea of
the method is to obtain temperature from the double ratios of the yields of four suit-
ably chosen isotopes, (A1, 21), (.41 + 1, 21 ), (142,22), (A2 + 1, 2;) via the relation[54],
AB
T - = ————— 6.9
150 ln(aAY) ( )
124
where
Y(AlaZI)/Y('41+1’Zl)
A}, : ‘ (.10
l/(flg, ZQ)/}r(A2 +122)! () )
AB = B(.4,, 21) — B(A1 + 1.21)
_B(.42.Z2) + B(.42 +1.22). (6.11)
and
a 2 (2122-42 + 1) (2.121,..n+-1 +1) .42 (A, +1) 3/2 (612)
(2'121v/11+1)(2J22..42+1 +1) Al (42 +1) ‘ -
Here Y(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 JZA is the ground state spin of a nucleus with
charge Z and mass A. However, this isotope thermometry method is derived within
the context of the grand canonical ensemble which could be doubtful as compared
with experiments which are probably closer to microcanonical limit. In the following
sections, the validity of the Albergo formula (Eq. 6.9) will be tested within this
approximate microcanonical SMIV’I.
6.1.2 Temperature Distributions
The SMM procedure expressed in Eqs.(6.2-6.4) leads to a distrilmtion of the temper-
atures of the fragmenting system for a given excitation energy in the same sense that
the temperature of the daughter nucleus in compound nuclear decay theory varies as
a function of the Coulomb barrier and separation energy of each decay channel. The
points in Fig. (6.1) denote the temperature distributions for the fragn‘ientation of an
excited “2872. nucleus at three different excitation energies obtained with the SMM.
These distributions are well fitted by gaussian functions, shown by the curves in the
figure, with variances 0:}. that are fairly independent of the energy, O'T z 0.4 MeV,
in the range 3 MeV S Eg/A S 10 MeV. At each excitation energy, we average over
125
- —c+ EJA=4Mev
_.. EJA=6Mev
0.12 r
.0
o .0
m _s
W) (MeV‘)
.0
8
0.04
0.02 *
Figure 6.1: The points denote distributions of temperatures calculated with the SMM
approach for the decay of a 112571 nucleus at three different excitation energies. The
lines denote gaussian fits to the calculated distributions.
all of the fragmentation modes and define this average value as the ”approximate
microcanonical” temperature TM 1C.
Since each of the isotopes employed in the thermometer has a specific mass, charge
and binding energy, the application of conservation laws sets a constraint on the values
available to the remainder of the system. Because of this finite size effect, the tem-
perature distribution obtained when a. specific isotope is present is slightly different
from the one obtained when all fragmentation modes are considered. In particular,
a small difference (3 0.1 MeV) is observed between the average temperatures for the
various isotopes; this is illustrated in Fig. (6.2) for carbon isotopes from the fragmen-
tation of a ”2572. nucleus at E} /A = 6 MeV. Even though the average temperatures
126
0.12
.0
o
m
,
l l l l
o- -
3.5 4 4.5 5 5.5 6 6.5 7
T (MeV)
Figure 6.2: The points denote temperature distributions calculated with the SMM
approach for the different isotopes considered in the carbon thermometer for an ex-
citation energy of Ea/A = 6MeV. The lines denote gaussian fits to the calculated
distriljmtions.
are different reflecting the different binding energies of the three isotopes, all these
distributions are gaussians with nearly the same variances. we can extract another
temperature T, M p by averaging over fragmei‘itation modes which contain an Interme-
diate Mass Fragment (IMF) with 3 _<_ Z S 10. It’s interesting to note that TM”; can
exceed TIMF at low energies by as much as 0.2 MeV, in part because it takes more
energy to emit an IMF than to emit an equivalent mass in the form of alpha particles,
leaving less energy for thermal excitation.
The basic idea contained in Eq. (6.9) was derived under the assumption that the
primary yields are well represented by the grand canonical approximation at a single
breakup temperature; the double ratio was invoked to cancel out the contribution to
127
the yields coming from the neutron and proton chemical potentials. In the Sh‘lhx‘l,
however, the temperature varies from one fragmentation mode to another and the
chemical potentials, which appear within the grand canonical formalism as Lagrange
multipliers that conserve charge and mass, are not explicitly invoked. Thus, we can
not presume the validity of the Albergo’s formula ( Eq. 6.9) in the SMM and must
test its validity instead.
We begin with a test of the validity of Eq. (6.9) when one employs the primary
yields. The primary yield for the ground state can be related to the total yield by
N912 = 1 AZ -yf§}€X1)lF.§z('T)/’Tl (6-13)
for a given fragmentation mode. Following the procedure described in the previous
section, we will use this expression and Eq. (6.4) to obtain the average g.s. yield
distribution (N512). This, in turn, can be used in Eq. (6.9) to extract isotopic
temperatures as follows,
r93 1 ms