«5 t. ham”... 4 2“. Jana. !. :x if 1. ‘ w, 2' f. .0 .V ». u.u.-l. 5.23.5 .. «.I; 3%. THESK This is to certify that the dissertation entitled Multifragment Emission in Central Collisions of 36Ar+197Au at E/A=50, 80 and 110 MeV presented by Larry William Phair has been accepted towards fulfillment of the requirements for Ph .0. degree in Physics M )0! professor Date 9/9/93 MSU is an Affirmative Action/Equal Opportunity Institution 0- 12771 uilirlliiinniwi 1 3129301043 7 LSBRARY Ml°hlgan State University E IN RETURN BOXto remove this modicum your record. FLAG TO AVOtD FINES return on or before date duo. DATE DUE ' DATE DUE DATE DUE l it"? ” usu loAnNflrmnttvo mum/5w oppmmmmon Wt MULTIFRAGMENT EMISSION IN CENTRAL COLLISIONS OF 36Ar+197Au AT E/ A = 50, 80 AND 110 MeV By Larry William Phair 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 1993 ABSTRACT MULTIFRAGMENT EMISSION IN CENTRAL COLLISIONS OF 35Ar+197Au AT E/A = 50, 80 AND 110 MeV By Larry William Phair Multifragment disintegrations of highly excited nuclear systems may carry information about the equation of state and the liquid-gas phase transition of low density nuclear matter. The mechanism causing multifragment decays is, however, not yet understood. To study this phenomenon we constructed a low threshold 4n: charged-particle detector, the Miniball, and studied the reaction of 36Ar+197Au at E / A=50, 80 and 110 MeV. For central collisions, the chance for equilibration of the system is highest. We therefore determined the most efficient method of selecting central collisions by comparing the ability of several global observables to select events with suppressed projectile-like fragment emission and small anisotr0pic azimuthal emission patterns. Similar event selection was obtained by using the following global observables: the charged particle multiplicity NC, the total transverse energy E,, and summed charge emitted at midrapidity 2,. An average multiplicity of 4 intermediate mass fragments (IMF) was observed for the most central collisions at E/A=110 MeV. These IMF multiplicities are consistent with predictions from a statistical model for evaporation from an expanding compound nucleus. The statistical decay model predictions are sensitive to the low-density nuclear equation of state. The IMF yields are also consistent with predictions from a standard percolation model for the 36Ar+197Au system. Assuming a spherical decay geometry, percolation fails, however, to predict the measured fragment multiplicities for the heavier system 129Xe+197Au at E/A=50 MeV investigated by our group. For noncompact decay configurations (bubbles and toroids) the percolation model can reproduce the large fragment multiplicities observed for this reaction. to Kendra, Nathan and Jordan iv ACKNOWLEDGMENTS I would first like to thank my advisor Konrad Gelbke for his time, patience, teachings and leadership. His ability to cut to the core of a problem has made my work at the laboratory fruitful and enlightening. Bill Lynch, Betty Tsang, Scott Pratt and Hartmut Schulz have given their friendship as well as invaluable instruction during the time of this analysis. Wolfgang Bauer suggested and oversaw several theoretical projects, as well as sitting on my guidance committee. Maris Abolins, Aaron Galonsky and SD. Mahanti gave a careful reading of this work and also served on my guidance committee. Three postdocs gave me my first exposure to experimental work here at the lab. Dave Bowman, Nelson Carlin and Romualdo de Souza taught by example the virtue of hard work and took the time to share their knowledge of experimental nuclear physics. As they left, other capable scientists took their places. Nicola Colonna, Graham Peaslee, Carsten Schwarz, Thomas Glasmacher, Bill Llope and Carlos Montoya have added their friendship and guidance to my research. My fellow graduate students have offered much support. Mike Lisa never minded being a sounding board for my ideas (some of which were not so bright, but he was always polite telling me so). He also had many helpful suggestions for my analysis. Erik Hendrickson, Paul (Rooster) McConville, Raman Pfaff, Don Sackett, George Jeffers, Mike Wilson and others helped me survive my course work during the first two years of graduate school. I am V especially indebted to two graduate students, Cathy Mader and Paul McConnville, without whom I am pretty sure I would not have passed the comprehensive exam the first time. Observing their approach to problem solving was very instructive. The graduate students in our group: Yeongduk Kim, Hongming Xu, Fan Zhu, Tapan N ayak, Wen Guang Gong, Damian Handzy, Wen Chien Hsi, Min-Jui Huang, James Dinius and Cornelius Williams have all spent long hours building and maintaining the Miniball and/ or running my thesis experiment. Easwar Ramakrishnan, Shigeru Yokoyama, Afshin Azhari, Mathias Steiner, Phil Zecher, Tong Li, Dietrich Klakow, Qiubao Pan, Brian Young, Jim Bailey, Michael Fauerbach, Stephan Hannuschke, John Kelley, Eugene Gualtieri and others have also helped to make my experience enjoyable and instructive. All of the staff deserve my thanks, but I would like to single out Jack Ottarson and Jim Vincent. I interrupted both of them countless times with different projects, requests, problems, etc., but every time they were always willing to help (quickly). This helped to smooth out what could have been rocky times during the construction of the Miniball. None of my education would have been possible without my parents and the selfless hours they gave in teaching me everything (from music to Scouting). They nurtured in me a desire to learn for which I am eternally indebted. And finally I am indebted to my wife, Kendra, for her love and patience and to our two boys, Nathan and Jordan. When I came home late at night there was something about a sleepy two-year-old saying "I love you to come home, Dad" that just helped keep things in perspective. vi Table of Contents List of Tables .................................................................................................................. ix ‘ List of Figures ................................................................................................................ x Chapter 1 Introduction ............................................................................................... 1 Chapter 2 MSU Miniball ............................................................................................ 4 2.1 Mechanical construction .......................................................................... 4 2.2 Detector design ........................................................................................... 11 Phoswich construction ....................................................................... 11 Uniform scintillation response of Csl(Tl) ...................................... 15 Scintillator foils .................................................................................... 21 Light pulser system .............................................................................. 25 Chapter 3 Experimental Setup and Data Reduction ............................................. 29 3.1 Experimental setup .................................................................................... 29 3.2 Data acquisition electronics ..................................................................... 31 3.3 Particle identification ................................................................................ 34 Bicron crystals ....................................................................................... 43 3.4 Energy calibration ...................................................................................... 47 Chapter 4 Impact Parameter Selection.............., ..................................................... 53 4.1 Motivation .................................................................................................. 53 4.2 Definition of filters .................................................................................... 55 4.3 Comparison of relative scales ................................................................. 62 vii viii 4.4 Suppression of projectile-like fragments .............................................. 71 4.5 Suppression of azimuthal correlations ................................................. 84 Azimuthal correlation functions ..................................................... 85 Comparison of relative scales ........................................................... 89 4.6 Directivity .................................................................................................... 103 4.7. Summary .................................................................................................... 107 Chapter 5 General Reaction Characteristics Selected by Impact Parameter ....................................................................................................................... 110 5.1 Multiplicity distributions ......................................................................... 110 5.2 Element distributions ............................................................................... 121 5.3 Angular distributions ............................................................................... 127 Chapter 6 Model comparisons .................................................................................. 132 6.1 Expanding emitting source model ......................................................... 133 6.2 Percolation ................................................................................................... 137 Toroids and bubbles ............................................................................. 147 Chapter 7 Fluctuations in multifragment emission ............................................ 157 Chapter 8 Summary .................................................................................................... 167 Appendix A Physics tape format .............................................................................. 170 Appendix B Temperature estimate from percolation theory ............................ 172 List of References ......................................................................................................... 173 List of Tables Table 2.1 Coverage in solid angle, polar angles, azimuthal angles and distance to the target (d) for individual detectors of the Miniball. ................... 9 Table 3.1 Fit parameters of Equation 3.8 for Figure 3.14 ...................................... 50 Table 3.2. Calculated energy for which the listed particle punch through 2.0 cm of CsI ................................................................................................................... 50 Table 4.1 Moving source parameters used to fit the energy spectra in Figures 4.10-12 ............................................................................................................... 76 Table A.1 Physics tape format for particle identification ..................................... 171 ix List of Figures Figure 2.1 Artist’ 5 perspective of the assembly structure of the Miniball 41: fragment detection array. For clarity, electrical connections, the light pulsing system, and the cooling system have been omitted .............................. 5 Figure 2.2 Half-plane section of the Miniball array. Individual detector rings are labeled 1 through 11. Numbers of detectors per ring are given in parentheses. The polar angles for the centers of the rings are indicated. The dashed horizontal line indicates the beam axis. ........................ 6 Figure 2.3 Front views of different detector shapes. The detectors are labeled by their ring number. Numbers of detectors per ring are given in parentheses ............................................................................................................... 8 Figure 2.4 Isometric view of the target insertion mechanism. .......................... 10 Figure 2.5 Schematic of phoswich assembly of individual detector elements. The u-metal shield covering the photomultiplier is not included. ........................................................................................................................ 12 Figure 2.6 Photograph of photomultiplier assembly. The scintillator and the first matching light guide are removed. The ring glued to the u-metal shield defines the alignment of the can housing the voltage divider. The can has been removed to expose the voltage divider. ................. 14 Figure 2.7 Schematic of the active voltage divider used for the Miniball detectors. ........................................................................................................................ 16 Figure 2.8 Relative variation of scintillation efficiency measured for two parallel surfaces of a CsI(Tl) crystal by using a collimated on-source. The axes of the coordinate system are parallel to the sides of the scintillator. The coordinates are fixed with respect to the scintillator. ............ 19 Figure 2.9 Variations of scintillation efficiency measured with a charge integrating ADC for two different time gates selecting the fast and slow components of scintillation for CsI(Tl). .................................................................. 20 xi Figure 2.10 Variations of scintillation efficiency detected with collimated a-particles of 8.785 MeV energy (solid points) and collimated y—rays of 662 keV energy (open points). The left-hand panel shows the measurement for a detector which was rejected. The right- hand panel shows the measurement for a detector which was incorporated in the Miniball. ................................................................................... 22 Figure 2.11 Relation between scintillator foil thickness and rotational frequency of spinning measured for solutions of Betapaint of different viscosity. The lines show fits with the power law to: v'“. ................................. 24 Figure 2.12 Schematics of light pulser assembly and LED trigger circuit ......... 26 Figure 2.13 Open points - gain variations of a CsI(Tl) photomultiplier assembly determined by measuring the detector response to 8.785 MeV (1 particles. Solid points - same data corrected for gain shifts in the off- line analysis by using information from the light pulser system. .................... 28 Figure 3.1 Schematic diagram of the data acquisition of the electronics of the Miniball. ............................................................................................................. 32 Figure 3.2 Timing and widths of the fast, slow and tail gates ............................ 33 Figure 3.3 Fast versus slow spectrum for detector 3—6 (ring 3, position 6). ..................................................................................................................................... 35 Figure 3.4 Tail versus slow spectrum for detector 3-6 (ring 3, position 6). ..................................................................................................................................... 36 Figure 3.5 Slow versus fast’ for detector 3-6. The solid line is used to separate light charged particles from intermediate mass fragments ................ 37 Figure 3.6 Schematic tail versus slow describing variables used in the construction of the PID function. ............................................................................. 39 Figure 3.7 PID versus slow for detector 3-6 ............................................................. 40 Figure 3.8 Z resolution of detector 3-6. Yield of Z (counts) as a function of Z. ................................................................................................................................. 41 Figure 3.9 Fast versus time for particles that stop in the fast plastic of detector 3-6. Time t=0 is arbitrary. ............................................................................ 42 Figure 3.10 Fast versus slow for a detector with a CsI(Tl) crystal from Bicron, detector 3-16. ................................................................................................... 44 xii Figure 3.11 Fast versus tail for a detector with a CsI(Tl) crystal from Bicron, detector 3-16 .................................................................................................... 45 Figure 3.12 Tail versus fast’ for a detector with a CsI(Tl) crystal from Bicron, detector 3-16. ................................................................................................... 46 Figure 3.13 CsI calibration of detector 3-6. The light output response (slow) as a function of energy deposited in the C51 (Egg). The circles represent all previous calibration data for this detector while the squares represent calibration taken during the present experiment. The solid lines are fits using Equation (3.8) for the elements listed. ........................ 48 Figure 3.14 CsI calibration of Miniball rings 2-4. The average light output response (slow) as a function of energy deposited in the CsI (ECsl)- The solid lines are fits using Equation (3.8) for the elements listed. The fit parameters are given in Table 3.1. .................................................. 49 Figure 3.15 The circles represent all previous calibration He calibration data for detector 3-6 while the squares represent calibration taken during the present experiment. Solid line - from Equation 3.8 using the fit parameters for detector 3-6. Dashed line — spline fit of a punch through point to low energy calibration ................................................................. 52 Figure 4.1 Correlations between charged-particle multiplicity NC, transverse energy 15,. intermediate rapidity charge Z, and identified hydrogen multiplicity N, observed for 36Ar+197Au collisions at E/A=50 MeV. Adjacent contours of different color differ by factors of 5 ........................ 57 Figure 4.2 Correlations between charged-particle multiplicity NC, transverse energy E“ intermediate rapidity charge Z,, and identified hydrogen multiplicity N, observed for 36Ar+197Au collisions at E/A=80 MeV. Adjacent contours of different color differ by factors of 5 ........................ 58 Figure 4.3 Correlations between charged-particle multiplicity NC, transverse energy E}, intermediate rapidity charge Z,, and identified hydrogen multiplicity N, observed for 36Ar+197Au collisions at E/A=110 MeV. Adjacent contours of different color differ by factors of 5 ........................................................................................................................................ 59 Figure 4.4 Reduced impact parameter scales 5(X) extracted from the measured quantities X = NC, E" Zyand N, . Each panel shows the relation extracted for the indicated observable ...................................................... 61 Figure 4.5 Conditional impact parameter distributions extracted for Y = 2, (top panels), Y = N, (center panels), and Y = E, (bottom panels) xiii for the 35Ar+197Au reaction at E/A=110 MeV. Left and right hand panels show distributions selected by impact parameter cuts 13(X) =0.05- 0.1 and 5(X)=0.35-0.45 on the indicated observables (X = N,, E,, and Z, ). All impact parameter scales were constructed according to Equation (4.4). ................................................................................................................................. 63 Figure 4.6 Conditional impact parameter distributions extracted for Y = Z, (top panels), Y=Nc (center panels), and Y = E, (bottom panels) for the 36Ar+197Au reaction at E/A=50 MeV. Left- and right-hand panels show distributions selected by impact parameter cuts 5(X)=0.05- 0.1 and 5(X)=0.35-0.45 on the indicated observables (X =NC, E, and Z,). All impact parameter scales were constructed according to equation (4.4). ................................................................................................................................. 64 Figure 4.7 Conditional impact parameter distributions extracted for Y = Z, (top panels), Y = NC (center panels), and Y = E, (bottom panels) for the 36Ar+197Au reaction at E/A=80 MeV. Left- and right-hand panels show distributions selected by impact parameter cuts 5(X)=0.05- 0.1 and 5(X)=0.35-0.45 on the indicated observables (X =Nc, E,, and Z, ). All impact parameter scales were constructed according to equation (4.4). ................................................................................................................................. 65 Figure 4.8 Conditional impact parameter distributions extracted for Y = Z, (top panels), Y = NC (center panels), and Y = E, (bottom panels) for the 35Ar+197Au reaction at E/A=110 MeV. Left- and right-hand panels show distributions selected by impact parameter cuts 5(X)=0.05- 0.1 and 5(X)=0.35-0.45 on the indicated observables (X =NC, E, and Z,). All impact parameter scales were constructed according to equation (4.4). ................................................................................................................................. 66 Figure 4.9 Centroids and widths at half maximum of conditional impact parameter distributions selected by cuts on impact parameters constructed from other observables X indicated in the individual panels. For ease of presentation, the open points have been displaced from the centers of the gates on 13(X). Top, center and bottom panels show the values for impact parameter scales constructed from NC, E, and 2,, respectively. Left, center and right hand columns show data for the 35Ar+197Au reaction at E/A=50, 80, and 110 MeV. ........................................ 69 Figure 4.10 Energy spectra of beryllium (left-hand panels) and carbon (right-hand panels) nuclei emitted in peripheral (5>0.6, top panels) and xiv central (5<0.3, bottom panels) 35Ar+197Au collisions at E / A=50 MeV. The exact cuts on the measured charged-particle multiplicity are given in the figure. The solid curves are fits with equation (4.5). The parameters are listed in Table 4.1. ............................................................................ 72 Figure 4.11 As Figure 4.10, for E/A=80 MeV. ......................................................... 73 Figure 4.12 As Figure 4.10, for E/A=110 MeV. ....................................................... 74 Figure 4.13 Relative contribution of projectile-like source extracted from the energy spectra of He, Li, Be and C nuclei selected by different cuts on charged-particle multiplicity for 36Ar+197Au collisions at E/A=110 MeV. A scale of the reduced impact parameter 5(Nc) and a pictorial illustration of the geometric overlap between projectile and target nuclei are included ........................................................................................... 78 Figure 4.14 Fast-particle fractions (fractions of particles detected at 0=9°- 23° with velocities greater than half the beam velocity) for 35Ar+197Au collisions at E/A=50 MeV. Solid points, open squares and open circles depict values determined for narrow cuts on the reduced impact parameters determined from NC, E, and Z,. Star-shaped points represent simultaneous cuts on 13(Nc), 5(E,) and 5(Z,). .................................... 81 Figure 4.15 As Figure 4.14, for E/A=80 MeV. ......................................................... 82 Figure 4.16 As Figure 4.14, for E/A=110 MeV83 Figure 4.17 Azimuthal correlation functions constructed from particle pairs of protons, deuterons, tritons and He nuclei emitted in peripheral 35Ar+197Au collisions (l;>0.75, NC =2-9) at E/A=50 MeV. The correlation functions were constructed for particles emitted at polar angles of 913b=31°-50° using an energy threshold Eth/A=12 MeV .................... 88 Figure 4.18 Azimuthal correlation functions for He nuclei emitted in 35Ar+197Au collisions at E/ A = 35 MeV. Panels from left to right show data selected by cuts on reduced impact parameters 5 = 0.8, 0.6, 0.4, and <02, respectively .......................................................................................................... 90 Figure 4.19 As Figure 4.18, for E/ A = 50 MeV. ....................................................... 91 Figure 4.20 As Figure 4.18, for E/ A = 80 MeV. ....................................................... 92 Figure 4.21 As Figure 4.18, for E/ A = 110 MeV. ..................................................... 93 XV Figure 4.22 Azimuthal correlation functions for He nuclei emitted in 125’Xe4-197Au collisions at E/ A = 50 MeV. Panels from left to right show data selected by cuts on reduced impact parameters 5 = 0.8, 0.6, 0.4, and <02, respectively .......................................................................................................... 94 Figure 4.23 Reduced-impact-parameter dependence of the coefficients 2.1 and 2.2 used to fit the measured azimuthal correlation functions for emitted He nuclei emitted in 35Ar+197Au collisions at E/ A = 35, 50, 80 and 110 MeV. ................................................................................................................ 96 Figure 4.24 Reduced-impact-parameter dependence of the coefficients l1 and 12 used to fit the measured azimuthal correlation functions of protons, deuterons, tritons or He-nuclei emitted in 36Ar+197Au collisions at E/ A = 50 MeV. ....................................................................................... 98- Figure 4.25 Reduced-impact—parameter dependence of the coefficients l1 and 2.2 used to fit the measured azimuthal correlation functions of protons, deuterons, tritons or He—nuclei emitted in 129Xe+197Au collisions at E/ A = 50 MeV. ....................................................................................... 99 Figure 4.26 Reduced-impact-parameter dependence of the coefficients 1.1 and 1.2 used to fit the measured azimuthal correlation functions of He nuclei emitted in 35Ar+197Au collisions at E/ A = 50 MeV. Results from different impact parameters filters are shown by the different symbols indicated in the figure ................................................................................. 101 Figure 4.27 Reduced-impact-parameter dependence of the coefficients l1 and It: used to fit the measured azimuthal correlation functions of He nuclei emitted in 129Xe-t-197Au collisions at E/ A = 50 MeV. Results from different impact parameters filters are shown by the different symbols indicated in the figure ................................................................................. 102 Figure 4.28 Conditional impact parameter distributions, dP[5(E,)]/ d5(E,), for 35Ar+197Au collisions at E/ A = 110 MeV selected by cuts on 5(NC) = 0.05 - 0.1 (dashed curve), directivity D S 0.2 (dotted- dashed curve), and for the simultaneous cuts 5(Nc) = 0.05 - 0.1 and D s 0.2 (solid curve). ........................................................................................................... 104 Figure 4.29 Distribution of transverse momentum directivity D for. 35Ar+197Au collisions at E/ A = 110 MeV selected by cuts on 5(Nc) <0.2 (dashed curve). The solid curve shows the distribution of the transverse momentum directivity obtained after randomizing the xvi azimuthal distribution of the emitted particles according to an isotropic distribution (solid curve). .......................................................................................... 106 Figure 5.1 Upper panel — NC distributions. Lower panel - NM, distributions. Both from the reaction 36Ar+197Au at E/A=35, 50, 80 and 110 MeV. ................................................... , ..................................................................... 1 11 Figure 5.2 Measured relation between charged particle multiplicity, N,,, and total charge of identified particles (2525) for the reaction 35Ar+197Au at the indicated energies. Different colors represent contours that change by factors of 3. ........................................................................ 112 Figure 5.3 Percentage of detected charge Z", appearing in clusters (A>1) and intermediate mass fragments (IMF) as a function of reduced impact parameter for the reaction 35Ar+197Au at E/A=50, 80 and 110 MeV ................ 114 Figure 5.4 Measured IMF multiplicity distributions for the indicated gates on charged particle multiplicity N,,. Panels are labeled by incident energy. ........................... 116 Figure 5.5 First and second moments of IMF multiplicity distributions as a function of charged particle multiplicity, NC. Different symbols represent results for the indicated beam energies. ............................................... 117 Figure 5.6 First and second moments of IMF multiplicity distributions as a function of 5(Nc). The two circles show the approximate overlap between target and projectile for 13:02, 0.4, 0.6 and 0.8. ...................................... 119 Figure 5.7 First and second moments of IMF multiplicity distributions as a function of 13(E,). The two circles show the approximate overlap between target and projectile for 5:02, 0.4, 0.6 and 0.8. ...................................... 120 Figure 5.8 Elemental distributions at E/A=50 (circles), 80 (stars) and 110 MeV (diamonds) for I3=0.75 (top), 0.5 (middle) and 0.25 (bottom). Solid lines are exponential fits using Equation (5.1) ....................................................... 122 Figure 5.9 a as a function of 5(NC) for E/A=50, 80 and 100 MeV. ..................... 123 Figure 5.10 a as a function of 5(E,) for E/A=50, 80 and 100 MeV ...................... 124 Figure 5.11 Mass yields for a percolation simulation at fixed excitation energy. The open (solid) symbols correspond to distributions constructed for multiplicity N=67(77). The curves are power law fits .............. 125 xvii Figure 5.12 Evolution of the fit parameter I. as a function of multiplicity N for a percolation simulation at fixed excitation energy ............ 126 Figure 5.13 Angular distribution of elements Z=1-6 from the reaction 35Ar+197Au at E/A=50 MeV. Solid symbols - central cut 5(NC)<0.25. Open symbols - peripheral cut 5(NC)>0.75. ............................................................ 129 Figure 5.14 Same as Figure 5.13 for E/A=80 MeV. ................................................ 130 Figure 5.15 Same as Figure 5.13 for E/A=110 MeV. .............................................. 131 Figure 6.1 First and second moments of IMF multiplicity distributions as a function of charged-particle multiplicity, NC. Different symbols represent results for indicated beam energies. The solid, dashed, dotted- dashed and dashed-dotted-dotted curves show results calculated for the statistical decay of expanding compound nuclei of finite-nucleus compressibility K=144, 200, 288 and co, respectively. The dotted curves represent the calculations for K=200, filtered by the detector response. .......... 135 Figure 6.2 Elemental multiplicity distributions detected in 36Ari-197Au collisions at E/A=50, 80, 110 MeV (top panel) and in 129Xe+197Au collisions at E/A=50 MeV (bottom panel). The curves represent calculations with the bond percolation model (described in the text) for the indicated bond-breaking probabilities p. All calculations are filtered by the response of the experimental apparatus except for the dot-dashed curve. .............................................................................................................................. 139 Figure 6.3 Angular multiplicity distributions of light particles, Z=1,2 (circles), and intermediate mass fragments of Z=3-5 (squares) and Z=6- 12 (diamonds) detected in 35Ar+197Au collisions at E/A=50, 80, 110 MeV (top panel) and in 129Xe+197Au collisions at E/A=50 MeV (bottom panel). The curves represent calculations with the bond percolation model for a bond-breaking probability of p=0.7. The calculations have been normalized to the data at 0=45°. Dashed and solid curves show raw calculations and calculations filtered by the response of the experimental apparatus. ............................................................................................. 142 Figure 6.4 Relation between average IMF and charged-particle multiplicities detected in 35Ar+197Au collisions at E/A=50, 80, 110 MeV (open diamonds, open squares, and open circles, respectively) and in 129Xe+197Au collisions at E/A=50 MeV (solid circles). Thick and thin curves show the results of filtered and unfiltered percolation calculations, respectively. Details are given in text. ............................................. 144 Figure 6.5 IMF admixture as a function of 1, according to Equation (6.2). ....... 145 xvni Figure 6.6 Relation between average IMF and charged-particle multiplicities. Solid points represent values measured for 12‘e’Xe-i-197Au at E/A=50 MeV. Thin (thick) solid line shows the raw (efficiency corrected) percolation calculation for a solid sphere. The hatched area shows the range of average IMF and average charged-particle multiplicities predicted by percolation calculations for toroidal breakup configurations ............................................................................................................... 149 Figure 6.7 Relation between average IMF and charged-particle multiplicities. Solid points represent values measured for 129Xe+197Au at E/A=50 MeV. Thin (thick) solid line shows the raw (efficiency corrected) percolation calculation for a solid sphere. The hatched area shows the range of average IMF and average charged particle multiplicities predicted by percolation calculations for bubble-shaped breakup configurations ............................................................................................... 150 Figure 6.8 Extracted power-law exponents 1 fit to mass distributions predicted by the bond percolation model for the break up of toroidal systems as a function of Rt and p. ............................................................................ 153 Figure 6.9 Extracted power-law exponents A fit to mass distributions predicted by the bond percolation model for the break up of bubble shaped systems as a function of Rb and p. .............................................................. 154 Figure 6.10 It as function of p for a solid sphere (circles), a toroid of radius Rt = 3.0xa (diamonds), and bubble with inner radius Rb = 3.0xa (squares). ........................................................................................................................ 155 Figure 7.1 Upper part - Measured relation between transverse energy E, and total charged particle multiplicity for 36Ar+197Au reactions at E/A=110 MeV. Lower part - Charged particle multiplicity distributions for the cuts on E, indicated in the top panel. ........................................................ 160 Figure 7.2 Bottom, center and top panels show the mean values (NC), variances a: , and ratios 0'2 / (NC) of the charged particle multiplicity distributions for 36Ar+197Au reactions at E/A=110 MeV. These quantities were selected by various cuts on the transverse energy ................... 161 Figure 7.3 Relation between mean charged particle multiplicity (NC) and the ratios 0% / (NC). Solid circular points - experimental values extracted from near-central 35Ar+197Au reactions at E/A=35, 50, 80 and 110 MeV. Open symbols are explained in the text. ............................................... 163 xix Figure 7.4 Scaled factorial moments as a function of binning resolution. Solid points show experimental results for central 36Ar+197Au collisions at E/A=110 MeV. Open points show results from percolation calculations using p=0.7. Open squares depict calculations in which the total number of broken bonds is allowed to fluctuate ......................................... 165 Si 11‘. Chapter 1 Introduction Weakly-excited nuclei decay primarily by fission and light particle evaporation. For excitation energies much higher than the binding energy, explosive disintegration into light particles (232) takes place. Between these two extremes, there is a region in which copious production of intermediate mass fragments (IMF: 3.<_Z$20) is observed [Ogil 91, Bowm 91, Wadd 85]. The mechanism causing multifragment decays is not yet understood and is a subject of current debate. Many mechanisms have been proposed [Bert 83, Cser 86, Lync 87, Schl 87, Gros 90, Frie 90, More 75, Frie 83a, Frie 83b, Baue 87, Peil 89, Boal 88, Boal 90, Baue 85, Baue 86, Biro 86, Cerr 88] for the production of the fragments. For example, the action of thermal pressure may force hot nuclear systems to expand to low density where the exponential growth of density fluctuations may lead to a complete disintegration of the nuclear system [Bert 83, Schl 87], in analogy to a liquid-gas phase transition in infinite nuclear matter. Unfortunately, detailed comparisons of various fragment production models to experimental data have been lacking since few experiments [Doss 87, Boug 88, Troc 89, Boug 89, Kim 89, Blum 91, Ogil 91, Bowm 91] performed to date provided sufficient phase space coverage to allow the extraction of exclusive quantities. 2 Exclusive experiments have become necessary as different models, based on very different approximations, often give very similar results for inclusive data. Inclusive measurements suffer from the implicit averaging over impact parameter which makes it difficult to separate and understand statistical and dynamical effects. For studies addressing the thermodynamic properties of nuclear matter, the selection of central collisions is of particular interest since reaction zones formed in central collisions promise to reach the largest degree of equilibration. To address the question of multifragment decay of highly-excited nuclear systems and to provide exclusive (i.e. impact parameter selected) measurements for constraints on various theoretical models, we have measured fragmentation in the reaction 36Ar-i-197Au at E/A=50, 80 and 110 MeV. After addressing the technique of impact parameter selection we select central collisions and study the IMF yields making comparisons with two statistical models. The measured IMF multiplicities from central collisions are compared with predictions from the expanding emitting source model, a model in which an equilibrated source loses mass by evaporation as it expands. The fragment measurements are also compared to predictions from percolation theory. Simple percolation theory is of interest since it allows the study of finite system effects in a well defined model which exhibits a phase transition in the limit of infinite systems. Statistical models are of interest since microscopic transport calculations [Boal 88, Boal 90] capable of treating nonequilibrium-fragment emission predict fewer fragments than observed experimentally [Bowm 91]. Dynamics, as treated by these models, may not play the decisive role in fragment production. In this context, it is interesting to neglect dynamical effects and explore to what extent fragment formation 3 could be dominated by the geometric considerations contained in the percolation ansatz. The thesis is organized as follows: the experimental details are described in Chapters 2 and 3; Chapter 4 provides a detailed study of impact parameter selection for the 3‘5Ar+197Au reaction at intermediate energies; in Chapter 5 several reaction characteristics are evaluated as a function of bombarding energy and impact parameter; in Chapter 6 comparisons are made to predictions from the emitting expanding source model and percolation theory, and in Chapter 7 we look for intermittency signals in central collisions; a summary and conclusions are given in Chapter 8. Chapter 2 MSU Miniball 2.1 Mechanical construction The Miniball phoswich detector array is designed to operate in a vacuum vessel. An artist’s perspective of the three-dimensional geometrical assembly is shown in Figure 2.1. The array consists of 11 independent rings coaxial about the beam axis. For ease of assembly, as well as servicing, the individual rings are mounted on separate base plates which slide on two precision rails. The rings and detector mounts are made of aluminum. Good thermal conductivity between detectors and the mounting structure allows the conduction of heat generated by the photomultiplier voltage divider network into the array superstructure. This heat is removed from the Miniball by cooling the base plates to 15°C. By this means, constant operating temperature in vacuum is achieved after a brief equilibration time. The individual detector mounts are designed to allow the removal of any detector without interfering with the alignment of neighboring detectors. The entire assembly is placed on an adjustable mounting structure which allows for the alignment of the apparatus with respect to the beam axis. Figure 2.2 shows a half-plane section of the array in the vertical plane which contains the beam axis. Individual rings are labeled by the ring VSU-QO-O“ Figure 2.1 Artist’ 5 perspective of the assembly structure of the Miniball 41: fragment detection array. For clarity, electrical connections, the light pulsing system, and the cooling system have been omitted. MSU-90-046 Figure 2.2 Half-plane section of the Miniball array. Individual detector rings are labeled 1 through 11. Numbers of detectors per ring are given in parentheses. The polar angles for the centers of the rings are indicated. The dashed horizontal line indicates the beam axis. 7 numbers 1-11 which increase from forward to backward angles. For each ring, the number of detectors is given in parentheses. For a given ring, the detectors are identical in shape and have the same polar angle coordinates with respect to the beam axis. These angles are indicated in Figure 2.2. Since the angular distributions of the emitted particles are strongly forward peaked, the solid angle subtended by forward detectors is smaller than for backward detectors. Variations in solid angle were achieved largely by placing detectors at different distances from the target while keeping their size approximately constant. The front face geometries of the individual CsI(Tl) crystals are shown in Figure 2.3. Different detector shapes are labeled by the respective ring numbers with the number of detectors per ring given in parentheses (see Figure 2.2 for the definition of the ring numbers). The crystals are tapered such that the front and back surfaces subtend the same solid angle with respect to the target location. In order to reduce cost of fabrication, the curved surfaces were approximated by planar surfaces. The resulting loss in solid angle coverage is on the order of 2%, comparable in magnitude to the loss in solid angle coverage resulting from gaps between individual detectors (which must be provided to allow for mechanical tolerances and optical isolation between neighboring crystals). A listing of the detector solid angles is given in Table 2.1. An isometric drawing of the target insertion mechanism is shown in Figure 2.4. The targets are mounted on frames made of flat shim stock 0.2 mm thick. Each target frame is attached to an insertion rod. The insertion rods are mounted on a tray which can be moved parallel to the beam axis. An electromagnetic clutch provides the coupling to the insertion and retraction drive once a target rod is in the appropriate position. A third drive allows rotation of an inserted target about the axis of the insertion rod. This MSU-90-045 l 2 3 4 (:2) (IS) (20) (24) I 5 6 7 8 (24) (20) (20) (18) 9 IO ll (l4) (:2) (8) IL. 1 i i 1' 5 cm Figure 2.3 Front views of different detector shapes. The detectors are labeled by their ring number. Numbers of detectors per ring are given in parentheses. Table 2.1 Coverage in solid angle, polar angle, azimuthal angle and distance to the target (d) for individual detectors of the Miniball. Ring Detectors AQ (msr) 0 (°) A0 (°) A4) (°) d (m) 1 12 12.3 12.5 7 30.0 260 2 16 14.7 19.5 7 22.5 220 3 20 18.5 27.0 8 18.0 180 4 24 22.9 35.5 9 15.0 160 5 24 30.8 45.0 10 15.0 140 6 20 64.8 57.5 15 18.0 90 7 20 74.0 72.5 15 18.0 90 8 18 (-1) 113.3 90.0 20 20.0 70 9 14 135.1 110.0 20 25.7 70 10 12 128.3 130.0 20 30.0 70 11 8 125.7 150.0 20 45.0 70 10 Figure 2.4 Isometric view of the target insertion mechanism. Pht 5C5: l l rotation of the target is useful for the determination of the shadowing a detector experiences when it is located in the plane of the target frame. Also, sources can be mounted in the target position and rotated to point to different areas of the Miniball for debugging and calibration purposes. In its present configuration, the detector array covers a solid angle corresponding to about 89% of 41:. The loss in solid angle can be decomposed into the following contributions: (i) beam entrance and exit holes (4% of 41:); (ii) approximation of the curved surfaces corresponding to constant polar angle by planar surfaces (2% of 41:); (iii) optical isolation of detectors and allowance for mechanical tolerances (4% of 41:); (iv) removal of one detector (ring 8 position 6) at 0=90° to provide space for target insertion mechanism (1% of 41:). 2.2 Detector design Phoswich construction All phoswich detectors of the array are composed of a thin plastic scintillator foil, spun from Bicron BC-498X scintillator solution, and a 2 cm CsI(Tl) scintillator crystal. The foils in Rings 2-11 have an average thickness of 4 mg/cm2 or 40 um while Ring 1 foils have an average thickness of 5 mg/cm2 or 50 um. A schematic of the detector design is given in Figure 2.5. In order to retain flexibility in the choice of scintillator foil thickness, the scintillator foil is placed on the front face of the CsI(Tl) crystal without bonding material. The back face of the CsI(Tl) scintillator is glued with optical cement (Bicron BC600) to a flat light guide made of UVT Plexiglas. This light guide is glued to a 12 MSU-90-03l Fast scintillator L‘ght gunde PMT WA W \” Al-mylar foil \ Optical cement Figure 2.5 Schematic of phoswich assembly of individual detector elements. The u-metal shield covering the photomultiplier is not included. 13 second cylindrical piece of UVT Plexiglas (9.5 mm thick and 25 mm in diameter) which, in turn, is glued to the front window of the photomultiplier tube (Burle Industries model C83062). The photomultiplier tube and the cylindrical light guide are surrounded by a cylindrical u-metal shield (not shown in the figure). Front and back faces of the CsI(Tl) crystals are polished; the tapered sides are sanded and wrapped with white Teflon tape. The front face of the phoswich assembly is covered by an alurninized mylar foil (0.15 mg/cm2 mylar and 0.02 mg/cm2 aluminum). The primary scintillation of the plastic scintillator used has its maximum intensity at 370 nm. In bulk material of the scintillator, the intensity maximum is shifted to 420 nm by the addition of a wavelength shifter. Our scintillator foils are, however, too thin for an effective wavelength shift and maximum emission remains in the far blue region of the spectrum. The absorption of this light in CsI(Tl) places a constraint on the maximum useful thickness of the CsI(Tl) crystals. Additional absorption in the light guides can be minimized by using UVT Plexiglas rather than standard Plexiglas light guides. Such considerations become particularly important for phoswich detectors utilizing thin scintillator foils in efforts to reduce particle detection thresholds. Figure 2.6 shows a photograph of the basic photomultiplier assembly used for all detectors. The phoswich and matching first light guide have not yet been attached. A precision-machined aluminum ring is glued to the [.1- metal shield surrounding the photomultiplier and the cylindrical light guide. This ring provides the alignment for a precision-machined aluminum can which houses the voltage divider and which defines the detector alignment when bolted to the rings of the array support structure. In order to expose the voltage divider chain, this aluminum can has been removed and placed next 5'ng 5"“ In. unifies 3991‘, f6] l4 Figure 2.6 Photograph of photomultiplier assembly. The scintillator and the first matching light guide are removed. The ring glued to the u-metal shield defines the alignment of the can housing the voltage divider. The can has been removed to expose the voltage divider. 15 to the photomultiplier. The active voltage divider chain is soldered to the flying leads of the phototube. To prevent destruction of the FETs by sparking during operation in poor vacuum, the entire divider chain, including the leads to the photomultiplier tube, is encapsulated in silicone rubber (Dow Chemical Sylgard 184). Vacuum accidents occurring with fully biased detectors do not lead to divider chain failures. In fact, the detectors can survive a full pumping cycle from atmospheric pressure to vacuum with bias applied to them. The 10-stage Burle Industries model C83062E photomultiplier tube was chosen because of its good timing characteristics (tR=2.3 ns), its large nominal gain (=107) and its good linearity for fast signals. Since the apparatus is designed to operate in vacuum, active divider chains were chosen to minimize the generation of heat. A schematic of the active divider chain is given in Figure 2.7. The final stages of the divider chain are of the “booster" type which provide improved linearity for high peak currents generated by large signals of the fast scintillator. Uniform scintillation response of CsI(Tl) Previous experience with CsI(Tl) crystals used for the detection of energetic particles had revealed difficulties with the production of scintillators with uniform scintillation response [Gong 88, Gong 90]. Therefore, considerable attention was paid to select CsI(Tl) crystals of uniform scintillation efficiency. In previous tests of large cylindrical crystals [Gong 88, Gong 90], nonuniformities of the scintillation efficiency were detected by measuring the response to collimated y-rays. Such measurements are relatively easy to perform since they can be done in air. However, they are less suitable for small volume crystals, since collimated y-rays sample a relatively 16 MSU-SO-OSZ AhfiDE 50a COAXIAL CABLE ”(0%: l 7,- D 10% a, ‘n'uE-T 411,: Km“ 4722i ”II-HE F1EIOMQ 09F NHvAr I" T : W. t— 473+:31582M9. 08% \rw , , 47a VN2406L on '“F *— 4“ moist-'57:]: | E33 mo firvv . mess W. l— 479 i, 820 110 :: 3.9m: 05% MP F- 5; 680110. I— eeoko [Mr '"F 03% ”F 1—1 680m 02% W: I— um 01% “1,, 1— 680m PHoro- lnF *— "8 ”a SHv CABLE -HV E \I v 3‘— CATHODE , tom .1. lnF ' 7 J.- Figure 2.7 Schematic of the active voltage divider used for the Miniball detectors. 17 large volume of the crystal. Small scale fluctuations of the scintillation efficiency may remain undetected. In addition, measurements for small noncylindrical crystals are less precise, since the shape of the Compton background depends on the position of the collimated y-ray source. Such dependencies lead to additional uncertainties in the extraction of the photopeak position. It was determined, however, that nonuniformities of the scintillation efficiency can be detected very sensitively by scanning the CsI(Tl) crystals with a collimated a-source in vacuum. All crystals ordered from various manufacturers were rectangular in shape with dimensions of 2 in.x1.5 in.x 1 in. They were polished at the front and back faces (with dimensions of 2 in. x1.5 in.) and sanded at the sides. The back face was optically coupled to a clear acrylic light guide with the same dimensions as the crystal. This light guide, in turn, was optically connected to a photomultiplier tube of 1 in. diameter. The sides of the CsI(Tl) crystal and of the light guide were wrapped with white Teflon tape. By covering the front face of the CsI(Tl) scintillator with an aluminized mylar foil, a uniform light collection efficiency was achieved. (Without a reflective entrance foil, the light collection efficiency decreased by about 5% from the center of the front face to its sides.) The front face of the crystal was scanned in vacuum and the peak location of the 8.785 MeV a-line from a collimated 228Th a-source was monitored. In order to avoid edge effects, regions within about 2mm of the side boundaries of the crystal were not scanned. Most tests were performed with a simple multichannel analyzer equipped with a peak sensing ADC; in those instances the anode signal of the photomultiplier was shaped and amplified with standard electronics, using integration and differentiation times of 1 us. 18 Figure 2.8 shows the results of a scan for a crystal exhibiting a large gradient of the scintillation efficiency. The horizontal axis of the plot shows the location of the collimated a-source with respect to the center, along the short symmetry axis of the front face. Different surface treatments of the front face of the scintillator did not affect the measured variation of the scintillation efficiency. In order to demonstrate that such variations were related to the bulk material of the scintillator, we exchanged the role of front and back faces of this scintillator and performed an equivalent scan of the parallel surface (i.e. the previous back face). The results of the two scans are compared by the solid and open points in Figure 2.8 (The coordinate system was kept fixed with respect to the CsI(Tl) crystal.) Nearly identical variations of the scintillation efficiency are observed across the two parallel scintillator surfaces, indicating that the measured large gradient of the scintillation efficiency persists through the bulk material of the sample. The measurements shown in Figure 2.9 were performed by integrating the anode current of the photomultiplier with a charge integrating ADC using time gates of At=0.1-0.5 us and At=1.1-4.1 us, which select the fast and slow scintillation components of CsI(Tl). The fast component exhibits a larger variation of the scintillation efficiency than the slow component. Since the relative intensity of fast and slow scintillation components depends strongly on the T1 concentration [Birk 64, Mana 62], the observed variations of scintillation efficiency are most likely due to gradients in the T1 concentration. Crystals incorporated into the Miniball were preselected by scanning the 1.5 in. x 2.0 in. rectangular surface of original crystals along two perpendicular axes and requiring a uniformity of scintillation response better than 3%. The preselected crystals were then milled into their final shapes and 19 10 MSU-90-036 IIrIrrTTIITIIIIIIIrIIIIIIITITII — front side ' é --- back side i” lllllllllllllllllllllllllllllll Relative Light Output (70) N rlilijl'l'IIUIUIIIIIWUIIIUIIIII 6LllLLiLLlllJllllLLillLLlL4lll11 -15 -10 -5 O 5 - 10 15 X(mm) Figure 2.8 Relative variation of scintillation efficiency measured for two parallel surfaces of a CsI(Tl) crystal by using a collimated a-source. The axes of the coordinate system are parallel to the sides of the scintillator. The coordinates are fixed with respect to the scintillator. 20 MSU-90-035 I I T I T I I I I I I T I I I I I I I r I j —— ADC gate : 0.1-0.5 ps u" ADC gate: 1.1-4.1 [LS NOOIFU‘O) Relative Light Output (%) Figure 2.9 Variations of scintillation efficiency measured with a charge integrating ADC for two different time gates selecting the fast and slow components of scintillation for CsI(Tl). Here, 21 scanned a second time, requiring uniformity of response within 2.5%. The preselection process avoided expensive machining of poor quality crystals; it was about 90% efficient for the selection of crystals of the desired quality. Figure 2.10 compares variations of scintillation efficiency detected with collimated a-particles of 8.785 MeV energy (source: 228Th) and y-rays of 662 keV energy (source: 137Cs). The left- and right-hand panels give examples for a rejected and an accepted crystal, respectively. The enhanced sensitivity of the a-particle scan is obvious. It is probably caused by the fact that a-particles sample a much smaller volume than y—rays and that the two kinds of radiation exhibit different sensitivities to the T1 concentration [Birk 64, Mana 62]. Scintillator foils Scintillator foils were spun [Meye 78, Norb 87] from Betapaint, Bicron BC-498X plastic scintillator dissolved in xylene. The original solution was ordered with a 40% weight ratio of solute to solvent. It was then diluted by adding xylene until the solution had the desired viscosity of 20-30 P. The viscosity was determined by measuring the terminal speed v of a steel ball sinking in a glass tube filled with a sample of Betapaint. Correcting Stoke's Law for the finite diameter of the glass tube gives the following expression for the viscosity [Dins 62]: 2 3 5 n= 23’ (”0 ’p)[1-2.104(L)+2.09(-'—) -o.95(-’-) ]. (2.1) 9v R R R Here, n denotes the viscosity, r and R are the diameters of the steel ball and the glass tube, g is the gravitational acceleration and p0 and p are the densities of the steel ball and the Betapaint, respectively. For the fabrication of scintillator foils, a glass plate of 23 cm diameter was mounted horizontally on a small platform connected to the drive of an 22 MSU-90° 034 4 IITIIIIIrTTIIIrII’TIIIIIIr IIIIIIITIIIIIIfll’TIIII’IrT 3— Rejected -— Accepted - Relative Light Output (7.?) O I -4 LllllllLlllllllLlllllllllzL IIllLllLllllllllllllllllll -10 -5 O 5 10 -10 -5 O 5 10 15 X(mm) Figure 2.10 Variations of scintillation efficiency detected with collimated 0:- particles of 8.785 MeV energy (solid points) and collimated y-rays of 662 keV energy (open points). The left-hand panel shows the measurement for a detector which was rejected. The right-hand panel shows the measurement for a detector which was incorporated in the Miniball. ho am PEI ii‘jc 63c] 23 electrical motor which allowed the plate to spin about its center at a preselected speed. To facilitate the removal of spun foils, the glass plate was covered successively with metasilicate solution and Teepol 610 and then wiped to leave only a thin film of the releasing agents on the glass substrate. An appropriate amount of Betapaint was poured on the center of a glass plate. In order to provide rapid spreading of the initial solution, the plate was spun at an enhanced speed for the first few seconds until the entire plate was covered with Betapaint. Following this rapid startup, the glass plate was spun at the preset rotational frequency for approximately 4 minutes until a solid foil had formed. After spinning, the glass plate was stored in a flow of dry nitrogen for about eight hours. The foil was then peeled from the glass plate, mounted on a frame, and placed in a dry nitrogen atmosphere for another 24 hours to allow further evaporation of residual xylene. We obtained good and reproducible results by using dilute solutions and spinning at low rotational frequencies. A number of measurements were performed to determine the relation between rotational frequency and foil thickness. The results of these measurements are shown in Figure 2.11. For each foil, thickness and homogeneity were determined by scanning the foil in vacuum with a collimated 228Th a-source and measuring the energy of the transmitted (it-particles in a calibrated silicon detector. The energy loss in the foil was then converted to an areal density according to reference [Lift 80]. The spun foils were uniform in thickness to within typically 1-2% over an area of 7x7 cmz. Scintillator foils used for instrumenting the Miniball in its present configuration were selected to have a thickness of 4.0 1:012 mg/cmz. 24 MSU-90-037 B I I I I I I r I’ T I I I I I I I I I I I I 7 — 1) = 41 poise _ 5 '" 1" x t cc V" — 1‘ I 618‘ 5 -— \ \ \a=0.67 .— o ‘ ‘ ~ ‘5 >0 4 — s ' —— E 17 = 25 poise v a.) 3 - _ 2 p l L l L 1 L I l l l l l l l L I. l l L L 200 300 400 500 600 v (rpm) Figure 2.11 Relation between scintillator foil thickness and rotational frequency of spinning measured for solutions of Betapaint of different viscosity. The lines show fits with the power law toe v'“. 25 Light pulser system Gain drifts of the photomultiplier tubes are monitored by a simple and compact light pulser system which operates in vacuum. In order to preserve the modularity of the device and avoid unnecessary removal of optical fibers during transport, each detector ring is provided with its own light pulser system. Figure 2.12 shows schematics of the mechanical assembly of the light pulser system and of the driving circuit for the light emitting diodes (LEDs) which is triggered by an external NIM logic signal. During experiments, the light pulser is triggered at a rate of about 1 Hz. Light is generated by simultaneously pulsing an array of eight LEDs (Hewlett Packard HLMP-3950) which generate light at wavelengths around 565 nm. The emitted light is diffused by reflection from an inclined Teflon surface. Light fibers which view only the scattered light transport the light to the individual photomultiplier tubes. Because of temperature fluctuations and aging effects, operation of light emitting diodes is not stable over long periods of time. Therefore, the intensity of each light pulse is monitored by two PIN diodes (Hamamatsu 5123) read out by standard solid state detector electronics. The ratio of the signals of the two PIN diodes can be used to monitor their stability. The ratio of PIN diode and photomultiplier signals can then be used to monitor the gain of the individual photomultiplier tubes according to the relation Ch’:Chx(_1:_Afl)_). 1+APMT (2.2) Here, Ch denotes the ADC conversion measured for a given event, Ch’ is the conversion corrected for gain shifts and APD and APMT are percentage changes (measured with respect to some arbitrary time t=0) of the average of the two PIN diodes and the individual photomultiplier signals for LED generated light pulser events. Better than 1% gain stabilization is achieved if 26 MSU-90-049 Fiber optics bundle LED drivercircuit +24V 560 353335333233333332323333:3}» HLMP rift-3:3: ----- 395° ........ 2N 2N2222 2700 2700 e .mv 1.5us Figure 2.12 Schematics of light pulser assembly and LED trigger circuit. 27 the temperature of the CsI(Tl) crystals is kept constant. (Variations of the scintillation efficiency of CsI(Tl) caused by temperature fluctuations cannot be detected with the light pulser.) It was verified, however, that active cooling of the base plate ensures rapid achievement of a stable operating temperature for the Miniball. Figure 2.13 illustrates the gain stabilization achieved with the light pulser system. The gain variations of a photomultiplier (enhanced by variations of the supply voltage) were directly measured by irradiating a CsI(Tl) crystal with a-particles emitted from a collimated 228Th source and monitoring the peak location of the 8.785 MeV a-line; they are shown by the open points in the figure. The solid points in the figure show the peak positions obtained in the off-line analysis after correcting the gain variations according to information obtained by the light pulser system. Gain stability to better than 1% was achieved. r- .- flu. C C>.~.~ ~w-mvv~ 28 MSU-90-048 14: I I I F I l I I | l - - 1 . o o . 1.3 _- O uncorrected '1 O corrected 3 1.2 —- —‘ .5 : 0 u : (U : . U 1.1 _— ; Q) h : .2. : ‘ *3 1.0_— --°- ~0- 0- ..-._ --4---o--‘°'~-O —: 5:6 : : 0.9 C.— ": - 0 0 : 0.8 E— —' .. u 9 : 0.7 - I I l l l l I l l L d O 1 2 3 4 5 6 '7 8 9 10 11 Run # Figure 2.13 Open points - gain variations of a CsI(Tl) photomultiplier assembly determined by measuring the detector response to 8.785 MeV 0: particles. Solid points — same data corrected for gain shifts in the off-line analysis by using information from the light pulser system. Chapter 3 Experimental Setup and Data Reduction Details of the 36Ar+197Au and 129Xe+197Au experiments along with techniques used to extract particle identification and energies are described in the following sections. As the thrust of this thesis work is the study of the reaction 36Ar+197Au at E/A=50, 80 and 110 MeV, the focus of this chapter will be on details of this experiment. Only sparse details will be given with respect to the 129Xe+197Au at 50 MeV/ nucleon and 3'6Ar+197Au at 35 MeV/ nucleon experiments. 3.1 Experimental setup The experiments were performed with 36Ar and 129Xe beams extracted from the K500 and K1200 cyclotrons at the National Superconducting Cyclotron Laboratory of Michigan State University. The argon beam energies were E/A=35, 50, 80 and 110 MeV, and extracted intensities were typically 108 particles per second. The xenon beam energy was E/A=50 MeV with intensities of about 107 particles per second. The areal density of the gold targets was approximately 1 mg/cmz. Light particles and complex fragments were detected with the MSU Miniball phoswich detector array. For the argon beam energy of E/A=35 MeV 29 30 the array covered scattering angles of Olab=16°-160° (rings 2-11) and a solid angle corresponding to 87% of 41:. At beam energies of E/A=50, 80 and 110 MeV, the array covered scattering angles of Olab=9°-160° (rings 1-11) and a solid angle corresponding to 89% of 41:. Details about the detector geometry are given in Chapter 2. For the 129Xe+197Au experiment the Miniball consisted of 171 detectors (rings 2-11 with 4 detectors missing from ring 2) with a solid angle coverage of approximately 87% of 41:. At very forward angles 2°-16°, fragments of charge Z=1-54 were detected with high resolution using a 16-element Si (300 um)- Si(Li) (5 mm) - plastic (7.6 cm) array [Keho 92] with a geometrical efficiency of 64%. Where counting statistics allowed, individual atomic numbers were resolved for Z=1-54. Representative detection thresholds for fragments of 2:2, 8, 20 and 54 fragments were approximately 6, 13, 21 and 27 MeV/ nucleon. Energy calibrations were obtained by directing 18 different beams ranging from 2:1 to 54 into each of the 16 detector elements [McMa 86]. The energy calibration of each of these detectors is accurate to better than 1%, and position resolutions of :15 mm are obtained. The complete detector system subtended angles from 2°-160° with respect to the beam axis and had a geometric acceptance of 88% of 41:. The detector array was actively cooled and temperature stabilized. Gain drifts of the photomultiplier tubes were monitored by a light pulser system (see Chapter 2). All events in which at least two detectors fired were recorded on magnetic tape. Random coincidences were negligible due to the low beam intensity. Each Miniball phoswich detector consisted of a 40 um (4 mg/cmz) thick plastic scintillator foil backed by a 2 cm thick CsI(Tl) crystal. All detectors had aluminized mylar foils (0.15 mg/cm2 mylar and 0.02 mg/cm2 aluminum) 31 placed in front of the plastic scintillator foils. As a precaution against secondary electrons, the detectors of ring 11 (Olab=140°-160°) were covered by Pb-Sn foils of 5.05 mg/cm2 areal density for the higher energy argon experiments (E/A=50, 80 and 110 MeV) and rings 2 and 3 were covered by aluminum foils of 0.81 mg/cm2 for the lowest bombarding argon energy. In the xenon induced reaction, rings 9-11 of the Miniball were covered with the Pb-Sn foils. Particles punching through the 4 mg/crn2 plastic scintillator foils were identified by atomic number up to Z=18. Hydrogen and helium were identified by isotope as well. Approximate energy thresholds are Eth/Asz MeV for Z=3, Eth/AE3 MeV for Z=10 and Eth/AE4 MeV for Z=18 fragments. Low energy particles stopped in the scintillator foils were recorded but could not be identified by atomic number. 3.2 Data acquisition electronics Figure 3.1 shows a block diagram of the data acquisition electronics for the 36Ar+197Au experiments at E/A=50, 80 and 100 MeV where the Miniball ran as an independent detector. A fast clear circuit (not shown) was added to the acquisition electronics for the 129Xe+197Au experiment where the Miniball ran as a slave to the forward array. In events where the Miniball detected particles but nothing was detected in the forward array, the fast clear vetoed the gate to the tail FERAs, cleared the fast, slow, tail and time FERA, and cleared the bit register. Figure 3.2 shows the gate widths for the fast, slow and tail and their relative timing along with a typical signal from a Miniball detector. During the CAMAC readout of the FERAs, integer‘2 words for fast, slow, tail and time are written to tape for detectors that have signals above threshold. The dynamic range of the FERA is 2048 channels. 32 x12 ECL x12 COMO“ SW! 150 as Figure 3.1 Schematic diagram of the data acquisition of the electronics of the Miniball. 33 F 0 1500: 15m |_J l l | "fast" "slow" "tail" 33m 390m 15'“ Figure 3.2 Timing and widths of the fast, slow and tail gates 3.3 Particle identification Typical spectra used for on-line analysis are shown in Figures 3.3 and 3.4. These spectra were generated for detector 3-6 using one run of the 36Ar+197Au reaction at 110 MeV/ nucleon (about 4 hours of beam time, =107 events in the Miniball). By plotting the fast versus slow components of the signal (here slow=slow FERA word/ 4 and fast=fast FERA word/ 4), clear element identification is obtained. Using the tail (tail=tail FERA word/4) versus slow, isotope resolution for H and He is obtained. The lines in Figure 3.4 are used later to make isotope identification easier. To facilitate the gate-setting procedure, we have constructed some simple particle identification functions. For element resolution we define fast - e(slow) 2 fast’ = (3.1) where fast is the fast FERA word, slow is the slow FERA word and e is a detector dependent constant which ranges from O to 1. A spectrum of fast’ versus slow (slow word/ 4 + 15 channels) for all runs (36Ar+197Au, E/A=50, 80 and 110 MeV) is shown in Figure 3.5. Of particular note is the separation of the double 0: line from Li. This separation allows a rather clean measurement of the number of intermediate mass fragments (particles which fall to the right of the red line with 223) from the light charged particles on an event by event basis. In order to facilitate setting the gates for isotope identification we constructed the PID function shown schematically in Figure 3.6. Using the upper and lower lines shown in Figures 3.4 and 3.6 we construct the following variables: tailr = tail + 6 (3.2) 08f 35 500 —I I I I I I I I I I I I I I I l I T I I I I l ‘I— 400 o” _. 300 .3 .1.) - (I) - c0 _ ‘H . 200 % n‘;v -4 100‘ -— l 1 l . l - O l. J l l l L I l 1 J. J l l L l l l l l l l l 1— 1 l 0 100 200 300 400 500 slovv Figure 3.3 Fast versus slow spectrum for detector 3-6 (ring 3, position 6). 36 500 400 r 300 tail 200 j- 100 Figure 3.4 Tail versus slow spectrum for detector 36 (ring 3, position 6). slow 37 500 —I I l l T I I I I I I I l I l I I I I I l I I ‘1— 400 300 slow 200 100 Figure 3.5 Slow versus fast’ for detector 3-6. The solid line is used to separate light Chal‘ged particles from intermediate mass fragments. Wher from whet Figui and lsoto slow gates Itjecl the V idem 38 where 6 is a random number between -0.5 and 0.5 (to remove digitization from the displayed spectra); tail, = line, (slow) (3.3) tail2 = line2(slow) (3.4) where line, and line2 are shown by the upper and lower lines respectively in Figure 3.4 and Figure 3.6 and are parameterized as a function of slow; Arm = tail, -tai12; (3.5) A = tail, - tailr (3.6) and A PID=512XA . (3.7) [DIX Isotope resolution when using this PID function is shown in Figure 3.7 (note: slow = slow word/ 4 +15, the extra 15 channels just to make it easier to set the gates). In this figure are plotted only the particles that fall to the left of the rejection line in Figure 3.5 (i.e. just light charged particles). Gates were set in the valleys between different isotopes of Figure 3.7 to determine particle identification. On the other hand, for element identification, rather than set gates in the valleys, the gates were set along the ridges in Figure 3.5. A linear interpolation between ridges was then used to determine a ”real” (as opposed to ”integer”) Z value for the point (fast’,slow) in question. The resulting identification of elements for detector 3-6 is histogrammed in Figure 3.8 for the 110 MeV per nucleon runs. For particles which stop in the fast plastic (the horizontal line at the bottom of Figure 3.5) we cannot extract the Z value. However, it is possible to make some classification of the particles using time of flight. Figure 3.9 shows the fast signal versus time (time=tdetector-tRF, the time difference between the 39 500”"I""I l“"lr‘r'1 400:.— (slow, tail,) i 300— A _‘ 08 : 4.) 200—- I I I I 100 Figure 3.6 Schematic tail versus slow describing variables used in the construction of the PID function. 4O 500 '— 400 300 PID 200 m 100 — l O 1 l l l l l l l l L J l l L l O 100 200 300 slow Figure 3.7 PID versus slow for detector 3-6. 39 I I I I I I I I I 500 I I I I I 400 (slow, tail,) 300 _. tail 200 100 T I I r I.... l . . . .l O 100 200 300 400 500 Figure 3.6 Schematic tail versus slow describing variables used in the construction of the PID function. 40 500 :I I I I I I I I I I I I I I I I I I I I I L I .4 400 — 300 — Q . I—t Q" i 200 ~ 100 — — O F- l L l L l l I 1 l J l l l l l I I l l I L [d 0 100 200 300 400 500 slow Figure 3.7 PID versus slow for detector 3-6. 41 2000 I I I I I I I I I I I I I I : Det 3-6 : 1500 —' '— c _ \N, 1000 w >.. _ j H 500 —- 0 _L1 1 Figure 3.8 Z resolution of detector 3-6. Yield of Z (counts) as a function of Z. 42 500 *- 400 — 300 *— fast 200 100 tjnle (us) I J l L l 1 L1 1 14 l or ill 1 Figure 3.9 Fast versus time for particles that stop in the fast plastic of detector 3-6. Time t=0 is arbitrary. 43 rf time from the cyclotron and the time that a detector fires) for particles that stop in the plastic. Two branches are clearly seen — a prompt branch (on the left) and a slower branch (on the right), presumably due to fission of the heavy target. In this figure the time t=0 is arbitrary - only the relative timing between the two branches can be extracted. Bicron crystals The identification techniques described above work well for more than two-thirds of the detectors in the Miniball. The remaining one-third have been classified as ”Bicron-like” detectors and require special attention. Figure 3.10 shows a typical fast versus slow spectrum for detector 3-16 which has a CsI(Tl) crystal from Bicron. The characteristics of a detector which we call ”Bicron-like” are two-fold: a large Z separation in fast for small slow (which quickly disappears for large slow) and poor separation of Li, He, and H particles. Improved resolution (shown in Figure 3.11) can be obtained by plotting the fast versus tail components of the PMT signal. It is now easier to separate He from Li fragments. Figure 3.12 shows fast’ (replacing slow with tail in Equation 3.1) versus tail for several runs. Again a rejection line can be set to separate the Li fragments from the light charged particles. Light charged particles fall below the rejection line and are further identified by isotope using PID versus slow. 300 400 500 slovv 200 100 Figure 3.10 Fast versus slow for a detector with a CsI(Tl) crystal from Bicron, detector 3-16. 45 500 t__'_‘ I I' I I 1T T I I I I I I I I I I I I I I I I I L . 1 400 —' °. . - "J 300 fast 200 100 Figure 3.11 Fast versus tail for a detector with a CsI(Tl) crystal from Bicron, detector 3-16 46 500 I I I IfTI T—III I I I 400 300 tail 200 100 7 1 J_L L JL I; ;L LJ 1 l I L l I 0 100 200 300 400 500 fast’ Figure 3.12 Tail versus fast’ for a detector with a CsI(Tl) crystal from Bicron, detector 3-16. 47 3.4 Energy calibration Energy calibrations of forward Miniball detectors were obtained by measuring the elastic scattering of 4He, 6Li, 10B, 12C, 160, 20Ne and 35Cl beams from a 197Au target at incident energies of E(4He)/A=4.5, 9.4, 12.9, 16 and 20 MeV; E(6Li)/A=8.9 MeV; E(10B)/A=15 MeV; E(12C)/A=6, 8, 13 and 20 MeV; E(16O)/A=16 and 20 MeV; E(20Ne)/A=1O.6, 11.3, 13.3, 15.0 and 19.8 MeV; and E(35Cl)/A=8.8, 12.3 and 15 MeV. Figure 3.13 shows the response of the light output as a function of energy deposited in the C51 crystal of detector 3-6 for the listed elements. For calibrations we assumed a functional form for the light output (slow channel) as slow = yEC, + mew-E“ -1). (3.8) yis the slope of the light output in the region linear with energy. The values of the parameter [3 come from the extrapolation of the asymptote (yEcsl-fi) to the slow axis. The values of a come from the data points in the low energy curved region of the response where quenching effects make the response nonlinear. For detectors in rings 1-4, these calibrations are estimated to be accurate within 5%. Calibrations of more backward detectors were obtained by using the energies of light particles punching through the CsI(Tl) crystals to normalize to extrapolations of the average response of detectors at more forward angles. The average response of rings 2-4 is shown in Figure 3.14. For elements where the light output as function of ECsI has not been measured, a linear interpolation between measured curves is used. Table 3.1 contains the fit parameters for the curves in Figure 3.14. The resulting uncertainties in energy calibration at more backward angles are considerably larger, typically of the order of 10% and for some detectors as large as 20%. In the 35Ar+197Au at E/A=110 MeV we had for the first time He and Li punching through the 2.0 cm CsI crystals in the forward detectors. The ~m.v:~.~.$ -nv \SOTfl Detector 3—6 600 I I I I I I I T I I I I l I I lllJl 500 z=2 3 5 6 a 10 17 400 300 200 slow Channel 100 IIIIIIIIIIjIIIIIIIIIWIIIIIII llllllJlllllllJJlllLLll O H O O N O O 03 O O ECsI (MeV) Figure 3.13 CsI calibration of detector 3-6. The light output response (slow) as a function of energy deposited in the CsI (ECsI). The circles represent all previous calibration data for this detector while the squares represent calibration taken during the present experiment. The solid lines are fits using Equation (3.8) for the elements listed. 49 Energy Calibration 1200 .- I T I I I I I r I r I I I II, r I I/ ,«r 1 ‘ E I l/ : L Ring 2.3.4 I ,’ 1000 — Z=2 3 I 5 6 / 8 -- I- I I . : I, / I 10 : 800 :- 2 I I, ' .. 0/ . .. E / II o .4 =3: - /I . I, : I g 600 — / :/< 1 “j 0 '_'_ ’ y . J . D I I L . ,’ ‘ J 400 :- I’ . 200 —- '1 O - 1 I l L l 1 l l l ' I [A l L d 0 200 300 400 500 EN(MeV) Figure 3.14 CsI calibration of Miniball rings 2-4. The average light output response (slow) as a function of energy deposited in the CsI (ECsI)- The solid lines are fits using Equation (3.8) for the elements listed. The fit parameters are given in Table 3.1. 50 Table 3.1 Fit parameters of Equation 3.8 for Figure 3.14 Z y 6 a 2 5.00842 84.1273 0.0293086 3 4.32194 86.1841 0.023502 5 3.9500 199.942 0.0138215 6 3.46959 189.468 0.0164807 8 2.71884 126.610 0.0194922 10 2.4808 178.975 0.016974 17 2.01865 261.832 0.00867138 Table 3.2. Calculated energy for which the listed particle punch through 2.0 cm of C31. Z A E (MeV) 1 1 75.2 1 2 100.6 1 3 119.4 2 3 266.7 2 4 300.8 3 6 576.0 3 7 615.2 51 calculated punch through energies are listed in Table 3.2. These punch through points can be used to extend the calibration to higher energies. In doing so, it was discovered that the punch through points lie below the extrapolated He and Li curves of Figure 3.14 by typically 20%. Clearly there is a problem region (above 80 MeV for He) for which we have only the calculated punch through energy with which to calibrate. Our crude solution to this problem consisted of using the measured calibration curves out to the highest calibration point and then performing a spline fit out to the punch through point that connects the two regions (see Figure 3.15). 52 1500 I I I I I I I I I I I I I I 3 l l - - low energy fit p : 1250 _—— - including punch through / —: _ / -—-I " / : cu - ’ - Ci 1000 :— / / —.. _ / _ g .. , _ 4: 750 I / _ _. / _ U — / _ — / — 3 E / " ,9. 500 — / / —~ U) _'_ / - _ / Z 250 — —: p _ O _ l L l l I L l l L I l I J L I I I-4 0 100 200 300 ECsI (MeV) Figure 3.15 The circles represent all previous calibration He calibration data for detector 36 while the squares represent calibration taken during the present experiment. Solid line - from Equation 3.8 using the fit parameters for detector 3-6. Dashed line — spline fit of on punch through point to low energy calibration. Chapter 4 Impact Parameter Selection 4.1 Motivation Still today, it is not yet clear how experimental observables from nuclear collision experiments can provide quantitative information about phase transitions in nuclear matter [Saye 76, Jaqa 83, Jaqa 84, Curt 83, Rose 84, C011 75, Morl 79, Shur 80, McLe 81, Kuti 81, Snep 88, Schl 87, Cser 86, Lync 87] - whether it is a liquid-gas phase transition at moderate temperatures and low densities, or a transition between a nucleon gas and a quark-gluon plasma at high densities and temperatures. Intermediate energy nucleus-nucleus collisions may produce finite nuclear systems at temperatures and densities commensurate with a liquid-gas phase transition in infinite nuclear matter. However, interpretations of inclusive measurements are complicated by the implicit average over impact parameter which makes it difficult to unravel the complex interplay between statistical and dynamical effects. Comparisons between experiment and theory are expected to become more tractable and more sensitive to unknown model parameters as research becomes more focused upon exclusive experiments in which specific reaction filters are employed to select narrow ranges of impact parameter. 53 54 In most experiments, information about the impact parameter is extracted from quantities which relate to the collision geometry via simple intuitive pictures. Many impact parameter filters represent some measure of the "violence" of the reaction which, in turn, is assumed to be related to the collision geometry. Common impact parameter filters are based upon the measured multiplicity of charged particles [Tsan 89a, Stoc 86, Stéc 86, Cava 90], the transverse energy [Ritt 88], or the summed charge of particles emitted at intermediate rapidity [Ogil 89]. For collisions with incident energies of a few hundred MeV per nucleon, the summed charge, Zbow, of particles with atomic number Z22 [Hube 91] has also been used. This quantity is the complement of the combined p, d, and t multiplicity. At lower energies, E/A=20-50 MeV, comparable information on impact parameter has been extracted from measurements of the velocities of fusion-like residues [Cali 85, Awes 81, Chen 87], charged-particle multiplicities [Chen 87, Tsan 89b], or neutron multiplicities [Gali 85, Morj 88]. A recent analysis of data with solid angle coverage restricted to forward angles (elabgm) suggests that improved selectivity for central collisions could be achieved by introducing a new observable, the transverse momentum directivity (defined in Section 4.6) and by simultaneous cuts on large charged-particle multiplicities and small transverse-momentum directivities [Alar 92]. A priori it is unclear to what extent the various techniques select similar or equivalent impact parameters, and whether one technique provides superior resolution to another. At low energies, cross calibrations have been performed between the linear momentum transfer techniques and the emitted charged [Tsan 89b] or neutral particle multiplicities [Gali 85]. In this chapter, we investigate 36Ar+197Au collisions at incident energies of E/A=50, 80, and 110 MeV and 129Xe+197Au collisions at E/A=50 MeV and IIkIet pafidex waged] hydrogex uefiflna Th wiousi duhtd] mrdadc fitsls: ganth be ‘mpaa I?pts Pflflnme 0ndu§o .l. I .ml. “-lPie 1 RIan 3.7;: by at ‘1‘! 033 EctDISH 55 explore the relation between impact parameter filters based upon the charged- particle multiplicity, NC, the total transverse kinetic energy of detected charged particles, E,, the mid-rapidity charge, Z), and the multiplicity of hydrogen nuclei, N1 (the complement of med). We also examine the usefulness of a directivity cut in selecting central collisions for these systems. The chapter is organized as follows: in section 4.2, we define the various impact parameter filters; in section 4.3, we compare the relative scales derived from the different observables and investigate their cross- correlations; in section 4.4, the efficiency of the individual impact parameter filters is quantified in terms of their ability to suppress contributions from projectile fragments and fast particles emitted at forward angles; in section 4.5, the impact parameter filters are studied again, in terms of their ability to suppress azimuthal correlations; in section 4.6, we study the effects of impact parameter selection using directivity cuts; in section 4.7, a summary and conclusions are given. 4.2 Definition of filters Throughout this work, we will use the following quantities to extract information on the magnitude of the impact parameter: (1) The charged-particle multiplicity, NC. This quantity includes all charged particles detected by the Miniball (and the forward array for the xenon induced experiment), even if they are not identified. For example, heavy fragments stopped in the scintillator foils are included in the definition of NC. Multiple hits in a single detector module are counted as single hits, even if they can be clearly identified as double hits (as is generally the case for double hits by (at-particles). The number NC is therefore equal to the number of detectors in which at least one charged particle is detected in a given event. [1560' Here, angle the M charge OI, eq'. Here, - iOIAI S: referer Sj‘sterr ddlneg Where 56 (2) The total transverse kinetic energy of identified particles, 15,, defined [Tsan91]as E: ZEsin25=9 2— “52mm (4.1) Here, 5,, pi, and 9,. denote the kinetic energy, momentum and emission angle of particle i with respect to the beam axis. E, was calculated using only the Miniball detectors. (3) The mid-rapidity charge, Z,, defined [Ogil 89a] as the summed charge of all identified particles of rapidity y with 0.25),“ S y S 075ij + 0.25))” (4.2a) or, equivalently, 0.75%, S y’ S 0.75y;mj (4.2b) Here, primed quantities are defined in the center-of—mass rest frame of the total system and unprimed quantities are defined in the laboratory frame of reference; yam, y,” and ym. denote the rapidities of the total center-of-mass system and of target and projectile, respectively. The rapidity, y, of a particle is defined [Gold 78] as 2 2 y=%m[\’m +P +pcos9]=mnh—1(ficosg), (4.3) \[mz +p2 —pcosG where m, ,8, and p denote the particle's mass, velocity and momentum, respectively. (4) The identified hydrogen multiplicity, N,. This quantity is defined as the number of detectors in which a Z=1 particle is identified. The definition includes hydrogen nuclei which punch through the CsI(Tl) crystals and double hits by p, d, or t Nl is the complement of Zbow, [Hube 91], the summed charge of particles with atomic number of 222. Figures 4.1-4.3 show the measured correlations between the quantities NC, 5,, Z, and N,. At all incident energies, the four quantities are strongly 57 E/A=50MeV Ar + Au I 0 200 400 500 800 c Et(MeV) Figure 4.1 Correlations between charged-particle multiplicity NC, transverse energy E,. intermediate rapidity charge Z, and identified hydrogen multiplicity Nl observed for 36Ar+197Au collisions at E/A=50 MeV. Adjacent contours of different color differ by factors of 5. 30 20 60 N 40 20 F ‘Blile 42 9.] . , 9'3" 5,. muifiplicn 58 E/A=80MeV Ar + Au 30 I l I I I I 40 130 -20 oz ‘10 0 '60 —40 4“ '20 0 0 10 20 30 O 300 600 9001200 NC Et(MeV) Figure 4.2 Correlations between charged-particle multiplicity NC, transverse energy E,, intermediate rapidity charge 2,, and identified hydrogen multiplicity Nl observed for 36Ar+197Au collisions at E/A=8O MeV. Adjacent Contours of different color differ by factors of 5. 3C 2C 1C BC N4C 2C 59 E/A=110MeV Ar + Au 30 I I I 45 20 ‘40 ‘20 O 0 15 3O 0 500 1000 1500 E*(MeV) Figure 4.3 Correlations between charged-particle multiplicity NC, transverse energy 8,, intermediate rapidity charge Z}, and identified hydrogen multiplicity NI observed for 36Ar+197Au collisions at E/A=110 MeV. Adjacent Contours of different color differ by factors of 5. 60 correlated. In general, an increase in the value of one observable is accompanied by increases in the values of the other three observables. From this observation one may already conclude that all four quantities (NC, E,, Z, and N,) are suitable for impact parameter selection — or none of them are. However, the correlation between E, and NC gives evidence for a slight saturation of NC at high E,, indicating that the transverse energy might provide a better central collision trigger than the charged-particle multiplicity. In order to construct an approximate scale for the impact parameter, we adopt the geometrical prescription proposed by Cavata [Cava 90]. For each of the quantities NC, E,, Zy and N,, we assume a monotonic relationship to the impact parameter and define the reduced impact parameter scale via :00 =b(X)= {jde—L-dx’f ) —dX’}/ (4.4) where X =Nc, E,, Z, and N,. dP(X) / dX is the normalized probability distribution for the measured quantity X, and bu,“ is the maximum impact parameter for which particles were detected in the Miniball (NC 22). In the following, we will use the reduced impact parameter scale 5 which ranges from [3:1 for glancing collisions to 5:0 for head-on collisions. The quantitative relation between the reduced impact parameters 5(X) and the measured observables X (X =NC, E,, Z, and N,) is shown in Figure 4.4. Individual panels present the relationships between the reduced impact parameters 5(X) and the observables X =NC, E,, Z, and Nl shown on the abscissae, and the different curves show the relationships extracted for the three different incident energies. While geometric prescriptions implicit in equation (4.4) may provide reasonable scales for the average relationship between charged-particle multiplicity and impact parameter, it is not clear, a priori, whether the scales 61 NC E,(MeV) 1 O0 10 20 30 40 0 250 500 750 1000 1250 . IIIIIIIIIIIII IIIIIVIII:~L: FIIIIIIIIIITfIIIITIIIIIQg °°Ar+‘"Au. E/Aa :: E 0.8 110MeV TT 9’ aouev 3: g ......... IIII 1111 1111 IIII IIII IIrI .4 .__4 1 ... .4 11111111 111111111 LLLII1111I111111111I1 1111'1111'1111I1L11I IIIIleleIIIIIIIIII \ ._. \ 1 1111IllllIlLllIlLlJIl.PLh 5 10 15 20 25 30 ZY' DI1 111111111111111 O N O 4:. O O} O 0 Figure 4.4 Reduced impact parameter scales 13(X) extracted from the measured quantities X =NC, E,, Z, and N,. Each panel shows the relation extracted for the indicated observable; different curves represent the relations extracted at the three incident energies. 62 extracted from the various quantities are commensurate. Furthermore, for collisions at fixed impact parameter, these quantities exhibit fluctuations of unknown magnitude. Therefore, reaction filters constructed from the various observables could have different resolutions. This question will be addressed in the next sections. 4.3 Comparison of relative scales In order to investigate the relationship between impact parameter scales extracted via equation (4.4) from the various measured observables, we have set narrow gates on impact parameters 5(X), defined by means of an observable X, and determined the conditional distributions of impact parameters 5(1’), constructed from different observables Y (Y at X and X ,Y = NC, E,, 2,, N,). Conditional impact parameter distributions are presented in Figures 4.5-4.8. Individual panels of these figures show conditional impact parameter dP(13(Y)) distributions T determined from the indicated observables Y. Left and right hand panels show distributions extracted for the cuts 13(X)=0.05-0.1 and 5(X)=0.35-0.45, respectively. Dashed and dotted curves show results obtained by cuts placed on one observable, and solid curves show results obtained by simultaneous cuts placed on two observables. The observables X used for these cuts are indicated in the individual left-hand panels; the conventions for left and right panels are identical. For better comparison, all conditional impact parameter distributions are normalized to unit area. For the present reaction, impact parameter filters based upon N1 (the complement of 2,0“) are considerably less selective than reaction filters based upon NC, E, and 2b,“. This effect is clear from Figure 4.5, which compares 63 36Ar + 1971.11, E/A=110MeV 6 : I I I I I I I I I I I I I I I I I I I I I I I::I I I I I I I I II I I I I I I I II I I I I I g .. .1.- m .7.— “ A ._‘.‘Z_ A A .—'1 C: 5 E b(z,) ; b(X) = 0.05-0.1 3E b(z,) : 1100 = 0.35-0.45 3 3° 4 E- """ X a N‘ iii—- _-:1 ('0 3:—// \.... “""E‘ -—::—- -—i§ E I .'\ x = E" N1 55 E 2 r- -' \ '- 1:— '1 El . \ 3E -, : 1 : \ ‘71:?— _5 1'1111111111 11-111111111::'11111111111 "2111111: 0 :I I I I I I I I I I I I I I I I I I I I I I I I::I I I I I I I I I I I I I I | I I I I I II I I: 5 g— S(N,) : Soc) - 0.05—0.1 "51-: $01,) :B(X) - 0.35-0.45“; 4 :_- """ X 8 Zr :: —§ E ‘- ‘- x a E :1: : 3 :— ‘ -::- -— 25 — X =- 31.2, 55 j 1 E— ‘Z’é‘ “S ' 1" I 1 1 L121: 1 I 1 1 1 = 0 I I I I I I I I I I::I I l I I I I I l I I I I: 5 Bus.) : 1300 = 0.05-0.1 1:? 1:03,) 801) = 0.35-0.45 "5 4 ----- x - zY -E':'- -E 3 — - X " N! ._:.E_ _: 2 — x " "v z, 55 3 EE ".\ E 1 .' \ -—ZE— , / ‘- —5 O{ .'-11~11111111112E 11111111111'31L111L: 0.0 0.2 0.4 0.6 0.8 0.0 0.2 0.4 0.6 0.8 1.0 Figure 4.5 Conditional impact parameter distributions extracted for Y = Z, (top panels), Y = N, (center panels), and Y = E, (bottom panels) for the 36Ar+197Au reaction at E/A=110 MeV. Left and right hand panels show distributions selected by impact parameter cuts 5(X)=0.05-0.1 and 5(X)=0.35-0.45 on the indicated observables (X =N,, E,, and Z). All impact parameter scales were Constructed according to Equation (4.4). dP(b)/db 64 36Ar + 19713.11, E/A=5OMeV GE IIIIWIIIII IIIIIII I I I IIIEErII IIIFIIIFITI I I I I II—FTII: 5 :_ - - _::_ - - _: c: 5 : b(z,) : b(X) = 0.05-0.1 :: b(z,) : b(X) = 0.35-0.45 : 4, IL— ..... ._":_ _: N 4 : x " N0 :: :1 I 3 E. " - X " Br .55. .3 2 E ,—."' _X=EvN EE 5° 2 - ' \ c l-‘_ d - -' . —::_' _: 1 :— <-- 55—- —; ”.1"111|1 1|1111"'1_IL111:: 1 llllllllll. 41: 0 : I II I I I I I I I I I I I I I I I lira-1:1 I I I I I I I I I I I I I I I I I I I rrI: 5 E’ b(Nc) : b(X) - 0.05-0.1 ‘3: b(Nc) : b(X) - 0.35-0.45 3 4 E— ..... x - z, -—:E— -2 d— -1 E - — x = E :: : 3 _ £ —_ —— 25 -.._—X=Evzr ii i E ' ' ii i 1 : .- 75— ": "1.111|1111|1111 ‘~11|1111:: LL: 0 :I I I I I I I I I I I II I I I I I I I I I I I::I I I: L - 3: .3 5 ~_- 5(3.) : b(X) = 0.05-0.1 55 : 4 E- ''''' x = z, —§§— —: 3 E— "' " X 3 Ne __::_ __'-‘ E (~--\ ——x-Nc.z HE E 2 - A“ \. I _:.-__ .4 E 1" V. 35 E 1 ' ‘ \". ._‘:_ . '- ._"‘ . V._ :I- ' \I. _ :1 0 I l l l I l l l l I l l l "II-44 l l l LLL: LI 1 l l l I l l I l I 1 L ' 4.1 L l: 0.0 0.2 0.4 0.6 0.8 0.0 0.2 0.4 0.6 0.8 1.0 A Figure 4.6 Conditional impact parameter distributions extracted for Y = Z, (top panels), Y = NC (center panels), and Y = E, (bottom panels) for the 36Ar+197Au reaction at E/A=50 MeV. Left- and right-hand panels show distributions selected by impact parameter cuts 5(X)=0.05-0.1 and 13(X)=O.35—0.45 on the indicated observables (X =NC, E, and Z). All impact parameter scales were constructed according to equation (4.4). 65 36Ar + 19713.11, E/A=80MeV 0") 6 I—I I I I I I I I I I I I I II I I I I Can-I I If I I I I IT I I I I I I I I I 5 I A I A I I a; I A I A I I 5 5 E“ b(z.,) : b(X) = 0.05-0.1 3E— b(z,) : b(X) = 0.35-0.45“: 4 E— """ X 3 NC i}. -_: 3 a. , .- - - x = E. _fia. __= I / -' '. __ - "b :1 2 = / x - 3.. N. 3%.. _a I: .- \ .. :: d 1 = _- \a, .55. _3 \2 35 ‘ a 'D IILIIIIILII .' _LIILIIIII‘” ' LIILILIIIILI' IJII" 0 —I I I I I I I I I II I I I I I I I I I—IEI I I I I I I I I I I I I I I I I I I I-1 a 5 l A I A l I a. I I A I I 5 > 5 E— b(Nc) : b(X) - 0.05-0.1 _EE— awe) : b(X) - 0.35-0 45"; .0 4 E._ ..... x - zY _g_ _i IE: 3 3.. / -'-’XI= Et ._3E_. : p- / ..' -.. a _— a _ _. me; E .- ' db— .4 : .' :t: z 1 I .' -. —‘E— .5 0 ‘i111 IILII 11 - 111 LIL41.§E ILI : :Fl I I I I I I I I I III I I—II II I II I3:I I I II I: 5 g" 5(3‘) : Soc) = 0.05-0.1 —§E— S(E,) : b(X) = 0.35-0.45 "5 4' :'—- """ X 3 z? '—::—- —: 3 E " ’ x ‘ "c €5— ——f 2 E _ x ‘ Nc- 2? SE 5 a '\. EE :3 1 ’ ‘- ‘73;— *5 ~$L4I I I I L I I I I F: I l I I I I I I II " I I LL I I: 0 0.0 0.2 0.4 0.6 0.8 0.0 0.2 0.4 0.6 0.8 1. IIIO-ZB-GSW Figure 4.7 Conditional impact parameter distributions extracted for Y = 2, (top panels), Y = NC (center panels), and Y = E, (bottom panels) for the 35Ar+197Au reaction at E/A=80 MeV. Left- and right-hand panels show distributions selected by impact parameter cuts 5(X)=0.05-0.1 and 13(X)=0.35-0.45 on the indicated observables (X =NC, E,, and 2,). All impact parameter scales were constructed according to equation (4.4). 66 36Ar + 197Au, E/A=110MeV ZPO-ZG-OSK 6 -I I II I I II III I I I II I I I I—II—I I II II I f I I II I I I I II fl 5 I AI A I I 55 I AI A I I i 5 :_ b(z,) : 1300 = 0.05-0.1 "E: b(Z.,) : bot) = 035-045“; 4 ;— , ----- x - a. -:;—- —-3 3%.—Is" ——X'E‘ -—f§- -—f E /-' \'2 — x 3 Ev Ne 55 E 2 : \°. —::— '1‘ : _' \'- EE '3 1 .' \'-. “:5— _z oIIIIIILII \' 1111'1111:: ~' 1 1'111 :1 : I I I II I I I I I I II I I I I I I I IE:I I I I II I I I I I I II :1 5 E" Sm) : Sm - 0.05-0.1 ‘gg‘ Ewe) : Bot) - 0.35-0.45? 4;.— ..... x-zr __;§_ —; 3 FE.— /4 - _ - x " E‘ -—fE—- —5.‘ : I". \'-A — x - E. z, :: : I .' \'. ‘I—- -: 2 5 ’-' \'-.' I; . - a 1' .- \n. ‘55-— - "E OIILII111L1'1IL111I111F: ”1'1111 IIII..'IIIIIL: —II II I I I I I I II I I I I I I I I—u—I I II I I I I I I I I I I I I III I- E. I A I A I I 35 I A I A I l g 5 : baa.) : b(X) = 0.05-0.1 “5.?" has.) : b(X) = 0.35-0.45 ‘5 : ' I: : 4E- '\ “““ x-z, 15— -= 3 3 ' " "‘ X ' N0 .35.. .5 2% _ x ‘ NI» 2, ii i 1“ -—;;— -: 0L '141I1L11L111F” 11Lfi 0.0 0.2 0.4 0.6 0.8 0. 0.8 1.0 A b Figure 4.8 Conditional impact parameter distributions extracted for Y = Z, (top panels), Y = NC (center panels), and Y = E, (bottom panels) for the 36Ar+197Au reaction at E/A=110 MeV. Left- and right-hand panels show distributions selected by impact parameter cuts 5(X)=0.05-O.1 and 13(X )=0.35—0.45 on the indicated observables (X =NC, E, and 2,). All impact parameter scales were constructed according to equation (4.4). 67 conditional distributions based upon N,, E, and Z, for collisions at E/A=110 MeV. The reduced resolution of impact parameter filters based upon N1 is most likely of statistical origin. For collisions at fixed impact parameter, the relative magnitude of statistical fluctuations is enhanced for the observable Nl because it contains, by definition, only a subset of the emitted particles. Figures 4.6, 4.7 and 4.8 present conditional impact parameter distributions at different energies. In general, the conditional impact parameter distributions become narrower with increasing particle energy. Qualitatively, this can be understood in terms of statistical fluctuations: at fixed impact parameter, statistical fluctuations lead to distributions in X =Nc, E,, and Z, of finite widths AX. For collisions at fixed impact parameter, the mean values of all quantities (X =NC, E, and 2,) increase with increasing projectile energy and the relative fluctuations, AX / X, decrease. Hence, impact parameter determinations should become more accurate at higher energies. At a given incident energy, rather similar conditional impact parameter distributions are extracted from the observables NC, E, and 2,, With little sensitivity to the applied cut on other observables. While some differences exist, they are generally small. Cuts on small impact parameters, 5(X)=0.05—0.1, generally produce conditional distributions that peak at larger impact parameters, 13(Y) =0.2. Conditional impact parameter distributions extracted for simultaneous cuts on Small impact parameters, 5(X,)=5(X2)=O.OS-O.1, are slightly narrower and the)’ peak at lower impact parameters, 13(Y)=O.12-0.16, than those obtained fr om cuts on a single observable. Hence, somewhat improved selection of cen trail collisions can be obtained from multi-dimensional cuts. One-dimensional cuts on 13(E,)=0.05-O.1 produce narrower distributions m bCNC) and 5(2,) than the alternative one-dimensional cuts on 5(2,) and 68 5(NC), respectively. Furthermore, two-dimensional cuts on small impact parameters produce narrower distributions for 13(E,) than for 5(NC) and 5(2)). These observations suggest that filters based on E, may be more effective in selecting central collisions than filters based on the other observables investigated in this work. For cuts on intermediate impact parameters, 13(X)=0.35-O.45, the conditional distributions are peaked at values close to 5(Y) =O.4, i.e. close to the cut on 5(X). This correspondence between the impact parameter scales derived from different observables improves at higher beam energies, indicating that all impact parameter filters have improved resolution at higher energies. Again, small improvements in resolution are obtained by the application of two-dimensional cuts. However, these improvements are less pronounced than those seen for very small impact parameters. The similarities of the various conditional impact parameter distributions shown in Figures 4.6-8 indicate that the three observables (NC, 5, and 2,) have similar selectivity on impact parameter. Figure 4.9 summarizes and corroborates these findings in a compact form. The individual panels of the figure show the centroids (points) and widths at half maXimum (vertical bars) of conditional impact parameter distributions selected by narrow cuts on impact parameters, centered at 5(X)=0.1...0.9 and extracted from other observables X . Left, center and right columns show data at E/ A=50, 80, and 110 MeV, respectively. Overall, the relation between the individual impact parameter scales is fairly linear with significant deviations occurring only for small and large l.mIDact parameters. Such deviations must be expected at the edges of the l.mlDact parameter scales whenever the two-dimensional correlations shown i - . . . n Izlgtues 4.1-3 have finite Widths and some curvature. f P‘- \ a . HIKE» 1.0 0.8 A ’chei 2 <5 0.4 0.2 0.0 0.8 A ’2 0.6 51 (.3 0.4- 0.2 0.0 0.8 A ’2 0.6 S ('3 0.4 0.2 0.0 0.0 0.2 E/A=50MeV 69 “Ar-t 197Au E/A=80MeV E/A=110MeV __ I III - III {III EIIIIIW ,EIIIIIIIII IV'VTI'ITI Iv VIITIITT' 'UT'IITII TV :II P 1 I l I a,l l I l l EIIII L I, r l 1 I . l Bot) Figure 4.9 Centroids and widths at half maximum of conditional impact p‘h‘alneter distributions selected by cuts on impact parameters constructed from other observables X indicated in the individual panels. For ease of pr eSentation, the open points have been displaced from the centers of the gates on 13(X). Top, center and bottom panels show the values for impact Parameter scales constructed from NC, E, and 2,, respectively. Left, center and right hand columns show data for the 36Ar+197Au reaction at E/A=50, 30, and 110 MeV. CfO‘ZB‘flSN 0.4 0.6 0.8 0.0 0.2 0.4 0.6 0.8 0.0 0.2 0.4 0.6 0.8 1.0 (Jfl‘ ._ Cr (C) ‘ré Pm. 51v Parr 70 Furthermore, the widths of the various conditional distributions are rather similar. Some modest improvement in impact parameter selection can be obtained by using multi-dimensional cuts (solid circular points). These improvements are difficult to quantify since none of the conditional distributions can be narrower than their natural widths at a given sharp impact parameter. Since the widths from one— and two-dimensional cuts are rather similar, one may surmise that the widths extracted from the two- dimensional cuts are close to the intrinsic resolution of the respective impact parameter filters. Improved impact parameter selection could be expected from observables (or combinations of observables) which are less subject to statistical fluctuations. For example, one may argue that central collisions could be reconstructed by an accurate determination of the energy deposited into internal degrees of freedom. For fixed excitation energy, the number of emitted charged particles will exhibit considerable fluctuations due to statistical partitions of the de—excitation energy among neutrons, protons and complex particles, all of which are emitted over a broad energy spectrum. Improved reconstructions of the initial excitation energy could be expected from simultaneous measurements of neutral and charged-particle multiplicities, but some fluctuations would remain due to the finite energy distributions of the emitted particles (and varying separation energies). One should expect statistical fluctuations to be smaller for the total transverse energy of the emitted charged particles since this quantity suffers only from the random partition of the excitation energy into charged and neutral particles and into longitudinal and transverse velocities. Close inspection of Figures 4.5-8 reveals that the double-gated 5(E,) distributions are slightly narrower than the other impact parameter distributions, possibly indicating 71 that the transverse energy exhibits a particularly good selectivity for central collisions. At higher energies and for noncentral collisions, geometric considerations may lead to the expectation that the intermediate rapidity charge might become competitive or even more appropriate since it could provide a better measure of the participant zone. Unfortunately, microscopic reaction models do not yet describe realistic descriptions of complex particle emission processes, and fluctuations in Z, due to fragment formation at the interfaces of the participant and spectator zones are difficult to assess. Since a substantial portion of the emitted nuclear matter emerges in the form of bound clusters, it is not clear which observables (or combinations of observables) provide optimal impact parameter selectivity. In the next section, we explore this question further by investigating an alternative measure for the selectivity of various impact parameter filters which is based upon their ability to suppress the emission of fast fragments at forward angles. 4.4 Suppression of projectile-like fragments In this section, we explore the efficiency of various impact parameter filters in terms of their ability to suppress projectile-like fragments emitted with near-beam velocity at forward angles. For illustration, Figures 4.10-12 show the energy spectra of beryllium (left-hand panels) and carbon (right-hand panels) nuclei detected in Rings 1- 10 in the Miniball for incident energies of E/A=50, 80, and 110 MeV, respectively. Top and bottom panels show the energy spectra gated by cuts on the charged-particle multiplicity corresponding to 5(NC)>O.6 and 5(NC)<0.3, respectively. At all incident energies, the energy spectra gated by large impact 72 36Ar+197Au, E/A=5OMeV IIII FlrrrIIIIIIIIT IIIIIIIrjIIIIIIIrlrg 10-3 80. zsncsia 7 c. 2511,3513 I" _ .0. .. § _ 71:}. '00-.‘ ‘. .‘.".,l... O 10 4 J _.-... ..... , ...... . g . 0.. o. . i.‘ .. 0.... ... A ‘. “ .l . ' .~ ‘. . > ‘ 0 g \! , .' cg °o 7 ' a l l I l l l l l l l l L L1 1 1+1 l L 1 J_L ; I l I I I T I Al I I I I T r FYI I I I I I 12.9 ' 4w, mucus a 19.5' 5’3. ~. 3 27° - ' ' 35.9 ‘ , . No. . ‘5, _ , . .' a 1!! ‘6 57.5. \s , '. on 7:29 p - . ‘ ’. 90° .. V ,_ 110° - ! . 130° 10—7_ILLJ[LLIIILI;LL111 LllLlLL.Ll|llllLlLll O 200 400 600 800 200 400 600 800 E (MeV) Figure 4.10 Energy spectra of beryllium (left-hand panels) and carbon (right- hand panels) nuclei emitted in peripheral (1350.6, top panels) and central (5<0.3, bottom panels) 36Ar+197Au collisions at E/A=50 MeV. The exact cuts on the measured charged-particle multiplicity are given in the figure. The solid curves are fits with equation (4.5). The parameters are listed in Table 4.1. dzP/dfldE (sr"1 MeV“) 73 36Ar+197Au, E/A=80MeV I I I I I T I I I I I I I I I—T I I I r1 I I I Ii I I I I fiT I I I I I 38.25Nc516 - C.2$N§SIB ’1‘? H .‘.... -- - ...... 4:31... .- . --.. 3" .~ ito,‘ a... . ... . -\ o... ...“.°°°°...".--. E; .0 . ..I .o 3‘ .’ . .0 _, o 3‘ . ‘3‘- 00 I. a \‘ , II.... DO \ o I ‘9 9!! -. 1 . a if i J 1 [J l L1 1 1 [I l l l I l I l 1 L l l l L L L l l l l l l l l l l 1 Fl I I I I I r I I I II I I I I I I I TI I I I I Fl I I I I I I I I I I ix - jab-k3 Be. N :2: 125° 1-13.1. c was ,, . .. 6 J4 . '. ‘. 19.5- . '. «x1 ‘ m. °. 35.9 ‘ " . '. ‘ 'l‘. , - 45' - 0 lg 'l‘ ' 57.5- - _' 'o- ’ Q : ~. '. i m- ,- fig ' f 90. ° if _- ' «. ‘ 110' i 130‘ § LllLngLllllLlLLLll lllllllLllLlllllllL O 200 400 600 800 200 400 600 800 E (MeV) Figure 4.11 As Figure 4.10, for E/A=80 MeV. QVO-ZB-HSN LIV .32: :u\Ln 3 (IN 1 f“Ed-re 4. 74 36Ar+197Au, E/A=110MeV 10-3 IIIIIIIIIIIIIIIrrIfIIIIIIjTIIII—IIIIIIT _4 Be. 25N¢$18 «"00... 12.5- c. 2311:5111 10 , a; ' 19.5' - ,3“ .0. i. - - . 27° ' "“ ‘1‘ .0‘ - . ........ .V\{‘ o i . ' . . Do 35.5. _ .‘ . ‘ 3» o. . ..-... I A 3‘ . O ... . ‘5. ‘ . . o .. T . ' . 57.5' \i .. '- ° 3’ O O 725' I! o '0 ' § °°° 90' 21-"- ”9 ! 1! 9 110° " g 7 130° ‘ ' "‘ | l l U! 1 1 1_L 1 1 1 1 1 1 1 1 1 1 1 1 L V I I I FT I I I I I I I I TfIf I I I C23 - .-_. ,O ’14:“ . ‘ Bo. mac 9.? ° "“11 \ 0 ‘ ‘ . 0. ° ' ' N o 11 'o '4" ! .0 f 10“? lllllllllllLllllllJ—lllllllllllllllllll O 200 400 600 800 200 400 600 800 E (MeV) Figure 4.12 As Figure 4.10, for E/A=110 MeV. BIO-ZO-OSN 75 parameters exhibit pronounced maxima at forward angles which correspond to fragment velocities close to the beam velocity. Such projectile-like contributions are strongly suppressed in the energy spectra selected by high- multiplicity cuts corresponding to 13(NC)<0.3, as qualitatively expected from simple geometrical considerations within the framework of a participant- spectator approximation. These qualitative findings are consistent with observations at lower energy, E/A=3S MeV [Kim 92]. They indicate that impact parameter filters based on charged-particle multiplicity are rather effective in selecting collisions with large geometric overlap between target and projectile, even in situations where simple participant-spectator models are not expected to be accurate in detail. In order to provide a more quantitative basis for discussing the effects of impact parameter filtering on the measured energy spectra, we fit the energy spectra with a simple three-source parameterization corresponding to the superposition of three Maxwellian distributions centered at velocities v, and characterized by temperature parameters Ti: 2 3 3 —E—V +Ei—21/E,(E-V )cosB fig: 20‘.(E,9) = XNiJE-vc exp ( C T C ) ,(4.5) i=1 i=1 1' where E. = §mvi2 (4-6) is the energy of a particle at rest in source 2'. The parameter VC is introduced to roughly account for Coulomb repulsion from a heavy charge assumed, for simplicity, at rest in the laboratory system [Awes 81, Chit 86]. Fits obtained with this parameterization are shown as solid curves in Figures 4.10-12. The corresponding parameters are listed in Table 4.1. A word of caution is necessary. Since our energy calibrations at backward angles have considerable uncertainties, the source parameters listed 76 Table 4.1 Moving source parameters used to fit the energy spectra in Figures 4.10—121 IP/A 2 NC N1 Bl T1 N2 [32 T2 N3 B3 T3 0 4 7—13 106 0.273 9.2 40.2 0.180 17.0 55.2 0.079 12.6 1 0 6 2-13 179 0.275 9.7 24.1 0.177 17.4 39.6 0.071 14.1 0 4 219 51 0.196 14.5 231 0.098 14.3 171 0.036 8.6 0 6 219 2.07 0.221 16.3 69.3 0.141 14.9 211 0067 14.6 0 4 2-16 107 0.303 11.1 16.3 0.190 20.2 44.1 0.073 12.9 0 6 2-16 54.7 0.313 12.5 5.46 0.169 20.0 34.6 0.66 12.4 0 4 225 28.1 0.218 21.8 152 0.106 21.9 212 0.046 12.0 80 6 225 1.08 0.257 24.5 31.5 0.142 24.7 238 0.071 16.9 110 4 2-18 26.1 0.326 17.5 6.88 0.165 24.3 34.6 0.060 12.6 110 6 2-18 4.37 0.333 19.4 1.67 0.155 22.3 25.4 0.057 12.4 110 4 230 9.48 0.248 31.9 91.8 0.123 29.9 251 0.055 15.8 110 6 230 12.1 0.406 24.3 17.4 0.142 34.8 218 0.068 19.4 in Table 4.1 may have large systematic uncertainties. In particular, the parameters of the slow target-like sources (i=3) must be viewed with caution, and the temperature parameters T3 should not be misconstrued as accurate temperature measurements for target—like residues. Nevertheless, the fits allow estimates of the relative contributions from fast projectile-like sources and intermediate-velocity ”nonequilibrium” sources. At all energies, peripheral collisions (5>0.6) are characterized by 'strong contributions from projectile-like sources. Such contributions are strongly suppressed for more ”central” collisions (5<0.3). However, central collisions have significant contributions from intermediate-velocity ”nonequilibrium” sources representing emission during the early, nonequilibrated stages of the reaction. Our present findings are consistent with previous observations at a lower energy [Kim 92]. Clearly, accurate descriptions of energy spectra and angular distributions will require lFor beryllium and carbon nuclei, the Coulomb parameters were Vc=32.9 and 47.2 MeV, respectively. The normalization constants Ni are given in units of 10’6/(8nMeV3’2); units for EIA and Ti are in MeV. At the energies E/A=50, 80. 110 MeV. the velocities of the projectile are: Bc.m.=0.051, 0.065 and 0.07, respectively. 77 theoretical treatments which allow the incorporation of emission from the earlier non-equilibrated phases of the reaction, as well as from the later more equilibrated stages. In order to provide a more quantitative measure of the selectivity of the impact parameter filter based upon the charged-particle multiplicity, we have fitted the multiplicity selected energy spectra of representative complex particles (He, Li, Be, C) with equations (4.5-6) and determined the relative contribution, om} / 0,0,, of the projectile-like source to the total particle yield. This contribution was evaluated by integrating the respective sources over all angles and energies: 0W. 2 H (II (E, 9)dEdQ , (4.7) 0,0, = 23“” 0,05, mama , (4.8) We did not analyze the energy spectra of hydrogen nuclei, because of the restricted dynamic range of the Miniball (75 MeV protons punch the C51 crystals). Moreover, light particles (especially nucleons) are less suitable for such an analysis, because of their comparatively large mean free path and because ”thermal” smearing of the energy spectra is more serious for light particles than for intermediate mass fragments. As a consequence, collective source velocity components are more difficult to unravel from the energy spectra of light particles. Figure 4.13 shows ratios 0,,0, / 0,0, extracted from the energy spectra of He, Li, Be and C nuclei selected by different cuts on charged-particle multiplicity for the 36Ar+197Au reactions at E/A=110 MeV. For illustration, a scale of the reduced impact parameter 5(NC) and a pictorial illustration of the geometric overlap between projectile and target nuclei are included in the figure. While such a simplistic graphic visualization must not be taken too 78 36Ar +197Au, E/A=110MeV 1 I I Tfir I I I I fl I I I I I I I I I I I I I I I I fl T I I l 1 l I z - A .. - b (3' 9 -§ - .9 .8 .7 .6 .5 .4 .3 .2 r g 0.8 — l r l l l l 1 l — g — . _. - CJ He ~ § 06 "" I O C U '7 Q : . 0 Be : '§ _ ' ’ I C . b“ 0.4 -— . _.. - C] O - _ C] C] S C] - . [:1 a 0.2 —— ' 0 — _ D , _ I ‘ 7 O P L 1 l 1 I. l l L l I l l 1 1 l l l l l L #1 ! JLI_L L! l l l 1 9 l q 0 5 10 15 20 25 30 35 NC Figure 4.13 Relative contribution of projectile-like source extracted from the energy spectra of He, Li, Be and C nuclei selected by different cuts on charged- particle multiplicity for 36Ar+197Au collisions at E/A=110 MeV. A scale of the reduced impact parameter 13(NC) and a pictorial illustration of the geometric overlap between projectile and target nuclei are included. 79 seriously, it nevertheless illustrates that complete overlap between projectile and target nuclei is only achieved for relatively small impact parameters, representing less than 10% of the total reaction cross section. Even in such a naive geometric picture, some emission from projectile-like sources must be expected down to impact parameters of 5:0.4. These simple expectations are fulfilled rather nicely for the emission of intermediate mass fragments for which projectile-like contributions are strongly suppressed at large multiplicities (small impact parameters). The suppression of projectile-like contributions is more effective for heavier (e.g. carbon nuclei) than for lighter particles (e.g. a-particles). In fact, the emission of fast a-particles does not follow the simple trends expected from simple geometric arguments. Possibly a-particle emission already sets in at the early contact phase of the reaction and can, therefore, not be described by a simple participant-spectator picture. In order to compare of the effects of different impact parameter filters on the shapes of the energy spectra at forward angles, we have analyzed the energy spectra of particles detected in rings 1 and 2 (&=9°-23°) and determined the ”fast-particle fraction” 0(v > vap)/ 0,0,, defined as the fraction of particles detected in rings 1 and 2 with velocities larger than half the projectile velocity. This simple quantity provides qualitatively similar insight as the quantity om]. / 0,0, obtained from the moving source decomposition, without necessitating cumbersome multi-parameter fits with equation (4.5). The fast- particle fraction is well defined and it can be established with good statistical accuracy even for narrow cuts on impact parameter. Furthermore, this quantity is insensitive to source-parameter ambiguities associated with fits to energy spectra which have poor statistical accuracy due to narrow cuts on impact parameter. 80 Fast-particle fractions extracted for various cuts on impact parameter are shown Figures 4.14-16. Different figures show results for the three incident energies. Individual panels show the fast-particle fractions for He, Li, Be and C nuclei, and different symbols depict results obtained by different impact parameter filters. At all energies, the fast-particle fractions are monotonic functions of the reduced impact parameter. The suppression of fast particles for cuts on small impact parameters is particularly effective for beryllium and carbon nuclei. The suppression of fast particles is less effective for 131-particles. These qualitative observations are consistent with the results obtained with the moving source decomposition shown in Figure 4.14. For impact parameters 5<0.6, all three impact parameter filters produce rather consistent fast-particle fractions. For intermediate and small impact parameters (5<0.6), filters constructed from NC, E, and Zy appear to provide comparable resolution. Slightly better suppressions of the fast-particle fractions can be obtained by employing triple cuts on values of the impact parameters reconstructed from NC, E, and 2, (see star-shaped points). Consistent with our previous findings, these improvements are relatively inconspicuous. For larger impact parameters, however, the differences between the various techniques become more significant, particularly at the higher energies. Such differences may be caused by statistical fluctuations of the quantities NC, E, and 2,. The relative magnitude of these fluctuations should be largest for peripheral collisions which are characterized by small mean values , < E,>, and . As a consequence, selections of large impact parameters may be associated with larger uncertainties than selections of smaller impact parameters. U(v>vP/2)/atot 81 36Ar+ 1971111, E/A=50MeV 1.0 .- I I I I I I I I I I r I T I Id-I I I I I I t I I I I I I I I I I . .q : : He I I ; U I U 1: Li a H 3 I: 3 (I? 0.8 :- g fl . _‘__ O i i n ‘1‘ ,3 _ 9 g _1. O t .1 g 1- x .0. O __l a 067 77“ . 0.4 :- ’ 1‘”"1: __‘I" '—_: : D X=Zy .tt -4 "' O xag, __:_ _“ 0-2 Z— x X=N¢. 2,. E. j- ~ :I 1 LJ 1 L L I l 1 1 I 1 I I I I l 1"! L I I l I I I 1 I I I ' 1 1‘ 0.0 '- I I I I I I I I I I I I I I I I I:: I I I I I I I I 1 I I I I I I 4 1.0 :- __ E n _1 : Ha I I: C :8! 5 5 : 0.8 :— 5 3 8 01:- g -_j : I 2: 5 § 2 0.6 _— 9 O i ‘17 a 1 ‘1 fl ._ 8 : a n «l- 9 n 1 0.4 :- . “__T . X -—-1 1— x .1... x d 0.2 E- __ —: O O :l l L I L l I L l l l L l I L I l l I]- l I I I l l I I l_ l l I l L L I 1 1 1d 0.0 0.2 0.4 0.6 0.8 0.0 0.2 0.4 0.6 0.8 1.0 1301) Figure 4.14 Fast-particle fractions (fractions of particles detected at 9=9°-23° with velocities greater than half the beam velocity) for 3‘5Ar+197Au collisions at E/A=50 MeV. Solid points, open squares and open circles depict values determined for narrow cuts on the reduced impact parameters determined from NC, E, and 2,. Star-shaped points represent simultaneous cuts on 13(Nc), 1305,) and 5(z,). 82 36Ar+197Au, E/A=80MeV 1.0 _ VT I I I I I r I I T I I 1 I4 I I I r r I 1 1 1 I I I T I 1 r _ g : He 1 l I l a It L, l I V a 2 5% 0.8 — x 1‘ —— g 3 9 e -—_ e I . 3 1: a .1 é o 6 ’ I " ' —3 °° ' : 9 g 2: o 9 : "' x '1- O I d 0.4 :— ' X'Nc ‘1" Q —. .8 _ DX=ZY __ _, e ' ”=8. .2; _: g 0.2 : )1 xauc, 2,, 3,1: 2 a '- l I l I 1 L1 I l L I I I l I I l l Id-l I I I I L1 I I I I IT I I L d \g 0.0 - r I I I I I I I I I I I I I I I I r___‘1 I I I I I I i 9 1 > " Be _.,.. C a 1 A - ' ’ )1 ll ‘1' _ 3, 0~8 .— 3 a 6'2:- x , . _ b : u I :t a 8 ° C : 0'6 '— a i: 9 i .. .1.- 8 .1 0.4 :- 8 E f:- ° -: .. 5 x -- a 3 J 02 :— )1 ‘1- a 2 1 .. - i .1 O 0 I. l l I I l I I I I l I L l l l I I l 4‘; L L I I I 1 LI I l 1 I 1 L1 I I I i- '0.0 0.2 0.4 0.6 0.8 0.0 0.2 0.4 0.5 0.8 1.0 b(X) Figure 4.15 As Figure 4.14, for E/A=80 MeV. U(v>vp/2)/Utot 83 36Ar+ ‘97Au, E/A=110MeV 1.0 I I I rI I I I I I I I I I I I I I fiI I I I T I I r I I r - I I I -- I I d = : He I g E‘ :: Li I I x E : .9 L. E .. 3 ' I c': - I 4- d g 0.6 _— 1 f;— 3 -: : q 5 :: fl : 0.4 — X ' X=Nc ‘Tf' g ‘— I o x=z, _ 9 I 0 2 :_ 0 Hz 3:. x _: . - XX=N0 Zr.E‘ :: -4 ”I l I I L L L I L LJLI l I L I I 1 1.0-1 1 I I l I I LLL ILLI l I L! I 1 -‘ 0.0 .- I I fl, I I I I I I I I T I I I I I I—-- r I I I I r I I I I I I I I I I I i d : Be x 2: C : 0.8 :- xz __.— —_ .. _. X .. o 6 ' g g a g " x ‘ __. __.... x _. . __ g .- g . _ .. -- B o 9 ‘ I— 8 ‘P ' 8 _ 0.4 :— 0 -::- l O _: - g R d- . c: — -- -I 0-2 3'3 ‘5' a 9 ‘: I.1 I I I l L 1 I 1 1 L I L l l I L l I‘I-J L l g L 1 I Ii 1 I I L L L I L L L‘ 0.0 0.0 0.2 0.4 0.6 0.8 0.0 0.2 0.4 0.6 Soc) Figure 4.16 As Figure 4.14, for E/A=110 MeV. 0.8 1.0 84 4.5 Suppression of azimuthal correlations In the previous section we have shown that impact parameter selection techniques based on NC, E, and Z, provide rather similar event selection and that, indeed, cuts on small, reduced impact parameters strongly suppress contributions from projectile-like fragments as expected from qualitative arguments. A recent analysis of data with solid angle coverage restricted to forward angles (9lab S 30°) suggests that improved selectivity for central collisions could be achieved by introducing a new observable, the transverse momentum directivity (defined in equation (4.11), below) and by simultaneous cuts on large charged-particle multiplicities and small transverse-momentum directivities [Alar 92]. In the next two sections, we address two questions: (i) can previously used impact parameters filters (NC, E, and Zy ) be used to select central collisions, i.e. collisions with small angular momenta?, and (ii) do cuts on the directivity significantly improve central event selection when used with detectors providing 41: coverage? In order to quantify the selection of collisions with small impact parameters, we explore the azimuthal correlations between emitted light particles [Tsan 84a, Chit 86, Fiel 86, Tsan 90, Ardo 90, Elma 91, Wang 91]. For truly central collisions, a reaction plane is undefined and the azimuthal distribution of emitted particles must be symmetric about the beam axis. If the azimuthal correlations between two emitted particles reflect their single-particle emission patterns, the azimuthal correlation function must become flat for central collisions. (Deviations from strict azimuthal isotropy may arise from final state interactions such as the sequential decay of primary reaction products produced in particle unbound states [Poch 87] or, for small systems, from momentum conservation effects [Chit 86, Lync 87].) For peripheral collisions on the other hand, transverse 85 flow effects or other ordered motion in the reaction plane [Tsan 84a, Chit 86, Fiel 86, Tsan 90, Ardo 90, Elma 91, Wang 91, Tsan 84b, Tsan 86, Tsan 88, Wils 90, Tsan 91] can cause large anisotropies in the azimuthal correlations. We apply this analysis to our previously measured [Phai 92a, Tsan 88, Bowm 91, deSo 91, Kim 92, Bowm 92] data of 36Ar+197Au collisions at E/A=3S, 50, 80 and 110 MeV, and 129Xe+197Au collisions at E/A=50 MeV. In this section we will define the azimuthal correlations and use them to test the effectiveness of the different impact parameter filters. In section 4.6 we explore the usefulness of a directivity cut in selecting central collisions. A summary and conclusions are given in section 4.7. In the previous section it was demonstrated that impact parameter filters based upon NC, E, and Z, were similarly effective in suppressing particles emitted with near-projectile velocities when cuts on small reduced impact parameters were applied. While the suppression of beam velocity particles is qualitatively expected for collisions between a relatively small projectile and a relatively large target nucleus, a suppression of beam velocity fragments may not necessarily be the best indicator for event centrality. Therefore, in the following sections we evaluate an alternative observable which depends less on a participant-spectator picture and which is more closely related to the angular momentum effects as evidenced by an ordered motion of the emitted particles in the entrance channel reaction plane. Azimuthal correlation functions Ordered motion of the emitted particles can be detected by measurements of azimuthal correlation functions [Tsan 84a, Chit 86, Fiel 86, Tsan 90, Wang 91] defined by the ratio 86 Y(A¢) Y'(A¢) 9,; Here, Y(A¢) is the coincidence yield of two (identical) particles emitted with = c[1+ R(A¢)]|9j. (4.9) relative azimuthal angle A4) at a polar laboratory angle 6 and in collisions selected by a specified cut on reduced impact parameter 5; Y ’(A¢) is the background yield constructed by mixing particle yields from different coincidence events, but selected by identical cuts on the reduced impact parameter; C is a normalization constant such that the average value of the correlation I+R(A¢) is one. All azimuthal correlation functions presented in this paper were constructed from particles detected at 9 = 31° - 50°. Azimuthal correlation functions may provide an additional diagnostic tool with regard to the selection of central collisions which is complementary to the suppression of beam velocity particles investigated in the previous section. For example, if projectile and target nuclei fuse at finite impact parameters, the emission of beam velocity particles will be suppressed for all impact parameters below a certain value. However, off-center fusion reactions will induce a collective rotation of the residue. The angular velocity of this collective rotation depends on impact parameter. Particle emission from a rotating compound nucleus is focused in a plane perpendicular to the angular momentum vector, i.e. the emission will be enhanced in the reaction plane [Chit 86]. The degree of this enhancement increases with the angular velocity of rotation, and it decreases as a function of temperature. The effect becomes more pronounced for heavier emitted particles [Chit 86]. For the case of equilibrium emission from a long-lived rotating system, the emission becomes left-right symmetric and the azimuthal correlation functions exhibit a characteristic V-shape [Chit 86]. In the limit of 5-9 0, the collective rotation ceases and the azimuthal correlation function becomes flat, i.e. R(A¢) -> 0. 87 Nonvanishing azimuthal correlation functions have been observed in a large number of intermediate-energy heavy ion collision experiments. In many instances, slightly distorted ”V”-shapes were observed [Tsan 84a, Chit 86, Fiel 86, Tsan 90, Elma 91] which could be understood in terms of a collective rotational motion in the reaction plane caused by the attractive mean nuclear field [Tsan 84a, Tsan 86, Tsan 88, Wils 90]. A number of other physical effects can influence the shape of the azimuthal correlation functions and lead to deviations from symmetric V-shapes. An important example is the directed transverse flow caused by the interplay of mean field deflection and pressure due to nuclear compression [Tsan 89a, Moli 85, Ogil 90, Wils 90, Sull 90]. Additional distortions may arise from phase space constraints imposed on finite systems by momentum conservation [Lync 82, Ogil 89b] or final state interactions [Elma 91]. Figure 4.17 shows azimuthal correlation functions of protons, deuterons, tritons and He nuclei (3He and 4He combined) detected in peripheral (13> 0.75) 35Ar + 197Au collisions at E / A=50 MeV. (For brevity of notation, all emitted He nuclei are denoted by the symbol a in the figures; contributions from 3He are smaller than those from 4He by a factor of about 5). A uniform software energy threshold of E/A=12 MeV was applied. In addition, the energies of protons, deuterons and tritons were required to be smaller than 75, 100 and 119 MeV respectively (the punch-through energies for the 2 cm thick CsI crystals). For He nuclei, no upper energy threshold was imposed since punch-through He nuclei could still be cleanly identified since no distinction is being made between 3He and 4He. Consistent with previous observations for slightly different systems [Tsan 84a, Chit 86,]ie1 86, Tsan 90, Ardo 90, Elma 91], the azimuthal correlation functions exhibit (slightly distorted) V-shaped patterns with a clear minimum at A¢ a 90°, reflecting the 1 +12(A¢xx) 88 °°Ar + 197Au, E/A=50 MeV 2.0 l I I I I I I I I I I r I T7 I I r f 5 ” ‘ ‘F T 8 x: “E ” 0 p . i ‘ 55 _ 0 d 91.b=31°-50° - 1.5 —- o t Nc=2-9 i — _ o a 8 _ 1+R(A¢XX) ’ i Q o o _ § § _ 0.5_ 0 o ‘— l—- o —1 L l l i I l l L I l l l l l l l l l L 0 50 100 150 A¢xx Figure 4.17 Azimuthal correlation functions constructed from particle pairs of protons, deuterons, tritons and He nuclei emitted in peripheral 36Ar+197Au collisions (5>0.75, NC =2—9) at E/A=50 MeV. The correlation functions were constructed for particles emitted at polar angles of 91ab=31°'50° using an energy threshold Eth/A=12 MeV. para" 1. Com 89 known preferential emission of nonequilibrium particles in the entrance channel reaction plane [Tsan 84a, Chit 86, Fiel 86, Tsan 90, Ardo 90, Elma 91, Tsan 84b, Tsan 86, Tsan 88, Wils 90, Tsan 91]. Again consistent with previous observations [Tsan 84a, Chit 86, Fiel 86, Tsan 90, Ardo 90, Elma 91, Tsan 84b, Tsan 86, Tsan 88, Wils 90, Tsan 91], the azimuthal anisotropies become stronger with increasing mass of the particle pair. Since the effect is particularly pronounced for He nuclei and since He nuclei are emitted in great abundance, we utilize the He azimuthal correlation function as a diagnostic tool for assessing whether cuts on small reduced impact parameters do, indeed, select central collisions. Comparison of relative scales Figures 4.18-22 show azimuthal correlation functions for He nuclei emitted in 36Ar + 197Au collisions at E/ A = 35, 50, 80 and 110 MeV and for 129Xe +197Au collisions at E/ A = 50 MeV. A software threshold of E/A=8 MeV was used in selecting the He nuclei. Different panels of the figures show results for different cuts on the charged-particle multiplicity NC. For each panel, the overlapping circles present a simple geometric picture of the collision geometry deduced by means of equation (4.4). At large impact parameters, i.e. low values of NC, the correlation functions show a strong preference of emission at relative azimuthal angles of A4) = 0° and 180°, characteristic of preferential emission in the reaction plane. At larger values of NC, the azimuthal correlation functions become more isotropic. For a given cut on reduced impact parameter, the azimuthal correlation functions become increasingly damped as the beam energy is increased. The effect may be related to the disappearance of flow predicted and observed [Moli 86, Ogil 90, Wils 90, Sull 90] in symmetric projectile-target collisions at comparable I ~f“ I?(A¢aa) 2.0 0.0 O 50 100150 0 50 100150 0 50 100150 0 50100150 90 36Ar + 19725.11, E/A=35 MeV MSU—QG-OIG IIII IIII IIIIIIIIII IIIILIJIIJI IIII .. .— i— .- _ "U .— .- .— 1111 1111 IIIIIIIIIIIIIIIIIII ® Nc=11 lLlIllIIIllIII IIII I I I I I I I I 0 | I I I I I I I I - - d d 1 1111 C9 IIIIII'IIIIIIIII q II—IIIIII I I I 1 d d - IIIIIIWIIIIIIIIIII <9 I I I I 1 I I I 1 I 1 J I IIIIIIIII Nc 14 l1111|llll|111 < I I I 1 chie J 1 I I I "lLii 1111l111LlLLlLlllll A¢aa(degrees) Figure 4.18 Azimuthal correlation functions for He nuclei emitted in 36Ar+197Au collisions at E/ A = 35 MeV. Panels from left to right show data selected by cuts on reduced impact parameters 5: 0.8, 0.6, 0.4, and <02, respectively; the actual cuts on charged-particle multiplicity NC are indicated in the figure. The circles show the approximate geometric overlap between target and projectile for the different cuts in NC. 2.0 91 °°Ar + 197Au, E/A=50 MeV IIIIIIIIIIII IIIIIIIIIIIIIIIIIIT IIIITIIIMISIIIII??IIOII? -- —u— A)— -n- —: h -0- db db "‘ - -- CED -~ -- © - I- q- -4- «I- -: _9. __ __ __ _ 1_ -_ -- _ u- «u— -I— .0. .1 .. __.. .0. "F' .. - d- ‘ I- d _ __. _w .. L - .. - _ . .- I .- _ - «r- -n- -1— -¢ :— cu— ~1- qp— - _. __ __.. __. y- -1P -I— dl— _ L - =14 -- a. Ncgm - i- l- -I- —II— -I u— I- -- c-h— ct [111111111l1111 [11111111111111 IIIIlU11IIIIIl1111 1I11l11111111111111 0.0 0 50 100150 0 50 100150 0 50 100150 0 50100150 A¢aa(degrees) Figure 4.19 As Figure 4.18, for E/ A = 50 MeV. 92 36Ar + 197.111, E/A=80 MeV 20 MSU-93‘0”! o _IIIIIIIIIIIIIIIIIII-O—IIIIIIIIIIIIIIIIIIL-IIIIIIIIIIIIII'IIIIq-IIIIIIIIIIIIIIIIIIII _ CD I CD I C9 -- © 4 1.5— —— —— —— —— A * "I n -- I B b all” I- d. - B b db d— -- .1 2 ~ -. -- -- - V 10_—- —:_ —V’——W ‘13 ‘_ -- -- -- I + - -- -- _- - H b -b -b --b -4 0.5— —— —— —— - Nc=9 ~— Nc=17 «- Nc=23 -- Ncgz'z - I”II[I'llIlllllllllllull-IILIIIIII'IllllllllflIIJllllIllllllll1ll-H-111111111l1111[1111" 0.0 . 0 50 100150 0 50 100150 0 50 100150 0 50 100150 A¢aa(degrees) Figure 4.20 As Figure 4.18, for E/ A = 80 MeV. 93 °°Ar + 1971111, E/A;110 MeV 1180-93-016 2.0 IIIIIIIIIIIIW[IIII IIIIFIIIIIIIIIIIII IIIIIIIIIIIIIIIIIII On '0» <0 fiIIlIIIIlIIIIIIIII cub p- - CC) 1- I 1 I 1.5 IIfilIIII ILIIIIIII IIIIIIIIII I I I I I I 1 I I I I I 1 I I 1 J J I < I I I I I I I I [<1 1 I I I I I 1 IIII [7 I II ILII I I fiI I I II Ij 1+R(A¢aa) 3 I I 1 I .- .- b — — I— I- L. b .— I- n I- h- P— 0.5 Nc=9 Nc=19 Nc=z7 NCZSZ IIIIIIIII JIIIIIIII .- .- .— L. IIIIIIIIII11IIII111 L111ILILIIIIIII11II IIIIIIIIII11111111 L1111L111I11L1I1111 0.0 0 50 100150 0 50 100150 0 50 100150 0 50100150 A¢aa(degrees) Figure 4.21 As Figure 4.18, for E/ A = 110 MeV. 94 12°Xe + 1""1111, E/A=50 MeV MSU-93-017 2.0 IIII[IIIIIIIIIIIIII IIIIIIIIIIIIIIIIIII IIIIIIIIIIIIIIIIVII IIIIIIIIII’IIIIIIIII OD: 01> c030 I IIITIII H '01 bIIIf I I I I I I I I I I I I I I I I I I 1 I I I I I I I I I I T I IILIIIILI I < i I 1 1 I IIIIIIIII IIIII IIIII JIIIIIIII IIIIIIIIIIIIII 0.5 Nc=12 -‘ Nc=22 "' Nc=3o “'- Nc235 0.0 1111IIIIIIIIIIIIIII.LIIIIIIIIIIJLLILLll-bLlLL llllIIIIIIIIII-V-IIIIIIIIIIIIIIIIIII 0 50 100150 0 50 100150 0 50 100150 0 50100150 A¢aa(degrees) Figure 4.22 Azimuthal correlation functions for He nuclei emitted in 129Xe+197Au collisions at E/ A = 50 MeV. Panels from left to right show data selected by cuts on reduced impact parameters 13: 0.8, 0.6, 0.4, and <02, respectively; the actual cuts on charged-particle multiplicity NC are indicated in the figure. The circles show the approximate geometric overlap between target and projectile for the different cuts in NC. 95 energies, possibly reflecting an increased balancing of attractive and repulsive forces from the mean nuclear field and from pressure, respectively. In order to allow a more compact presentation of the main features of the observed azimuthal correlation functions, we have fitted them by functions of the form 1+ 1.1 cos(A¢) + 2.2 cos(2A¢) , (4.10) where 1.1 and 12 are treated as free parameters. Large values of 22 may be associated with collective motion resembling a rotation [Lace 93]. Positive (negative) values of M indicate preferential emission of the particle pair to the same side (opposite sides) of the beam. Positive values of M can come from large final state interactions (e.g. the decay 8Be —-) 20:) or, alternatively, from directed sideward flow. Negative values of M (preferred emission on opposite sides of the beam) may reflect phase space constraints for small systems due to momentum conservation [Lync 82, Ogil 89b]. For isotropic distributions l1=l2=0. Figure 4.23 presents the values of the parameters 11 and 12, extracted as a function of reduced impact parameter 5(NC) for 36Ar + 197Au collisions at the different bombarding energies. At all bombarding energies, 12 decreases as a function of decreasing reduced impact parameter. Indeed, for E/ A 2 50 MeV, 12 -) 0 as 5(Nc) —) 0 with rather good accuracy, indicating that small reduced impact parameters do, indeed, select near-central collisions for which R(A¢) z 0 by necessity. At E/A=35 MeV, the selection of central collisions appears to be of slightly reduced quality since 12 2 0.1 even for very small values of 5(NC). To some extent, reduced selectivity for very central collisions at the lower energy may result from a loss of statistical resolution as ANC [NC becomes larger at lower incident energy. It is more likely, however, that this loss of resolution is related to the breakdown of the participant-spectator picture 96 1150—93-01:; :I I II I I I I III I I I7 I I I l I I I I: 0.6 — 36Ar+197Au 0‘: Q": C E/A(MeV) = gt: §§§ j “ E] 35 Q Q - 0.4:... 50 D .I Qi 02 Z 0 30 D D .' 099 ~ ’< P I 110 D ' go i Z I D ° 000 _ 02 :— C] D . j i _ _ 0 CI D . . . 00 I.- -4 _ . I ‘ 00 _. iIO 909099?" __‘ 0.2:HHHHHHiHHiHH: -’- 3 0.1 — .. ; o D D D I: D I: 99 g I r . (4.11) , y.2y.. In equation (4.11), the sum includes all identified particles with rapidities larger than the center-of—mass rapidity, and pii denotes the transverse momentum of detected particle i. The momenta are calculated assuming A=ZZ for 332318. Following the procedure outlined in Section 4.3, we explored the conditional distribution of impact parameters 505,) selected by the cuts 5(NC) = 0.05 —O.1 and D s 0.2. The results are shown in Figure 4.28. The 104 36Ar+197Au, E/A=110 Me 5 . MSU—93—023 I I I I — — — - b(Nc) = 0.05—0.1 / -‘-‘-' D§0.2 Simultaneous cut A dP(B)/db a . .-.-C-O‘O-o_0- ‘ . Figure 4.28 Conditional impact parameter distributions, dP[5(E,)]/ d5(E,), for 36Ar+197Au collisions at E/ A = 110 MeV selected by cuts on 5(NC) = 0.05 - 0.1 (dashed curve), directivity D S 0.2 (dotted-dashed curve), and for the Simultaneous cuts 5(NC) = 0.05 - 0.1 and D S 0.2 (solid curve). 105 dashed and dotted-dashed lines show the results for cuts on 13(NC) and D only, and the solid line shows the results for a simultaneous cut on 5(NC) and D. A single cut on D alone provides little selectivity of central collisions, and no improvement in the 5(E,)-distribution is observed if an additional cut on D1) and bound in intermediate mass fragments to the detected total charge, 2”. For this measurement all hydrogen nuclei punching through the C51 crystals were treated as mass number A=1. The percentage of the detected charge Zm observed in clusters (open symbols) and in intermediate mass fragments (solid symbols) is plotted as a function of 5(NC) for the three bombarding energies E/A=50, 80 and 110 MeV. At all bombarding energies nearly 30% of the detected charge appears in the form of intermediate mass fragments for central collisions. The charge bound in clusters accounts for about 80% of the detected charge, independent of beam energy. 114 36 197 Ar+ Au 100 I I I I I l I I I I I I I I I I I I I I I so M83886? €510<>CI<>OGO<> 00001:] [100000 _— 0 000 0000 D D 000 — 0 Do _ /\ ‘ <><><> U - E 60 — IMF A>1 E/A(MeV) 00 o _— 5’, ' I D 50 00 ° ‘1 N : . o 80 O o : qES - 0 O 110 0 o; N 40 — O — v - O- y'rfrfin-u {,3 I n ._ 20!. B Figure 5. 3 Percentage of detected chargeZ M appearing in clusters (A>1) and intermediate mass fragments (IMF) as a function of reduced impact parameter for the reaction 36Ar+197Au at E/A=50, 80 and 110 MeV. 115 To better characterize the various classes of multifragment events, we have extracted the probability distributions P(N,MF) of detecting NM, intermediate mass fragments in a single collision for different gates on the charged particle multiplicity NC. The four panels in Figure 5.4 depict these distributions. At each energy, the IMF multiplicity distributions become wider and shift toward higher average multiplicities as the charged-particle multiplicity increases. Peripheral reactions, selected by NC <7, exhibit narrow IMF multiplicity distributions peaked at NW =0. For these collisions, IMF emission is an unlikely process. On the other hand, for central collisions selected by large values of NC, IMF emission is a common process for which the average IMF multiplicity increases from (N,MF)=1 at E/A=35 MeV to (N,MF)=4 at E/A=110 MeV. At the highest incident energy, events are observed in which as many as 10 intermediate mass fragments are detected in the exit channeL To allow quantitative comparisons of IMF multiplicity distributions, we have determined their first and second moments, (NW) and of”. The dependence of these moments on the total charged particle multiplicity NC is shown in Figure 5.5. Values extracted at different incident energies are shown by different symbols as indicated by the key in the figure. At all energies, (MW) and 0er exhibit an initial, approximately linear increase as a function of NC. The slope of this increase is rather similar for the four energies investigated. For large values of NC, corresponding to the extreme tails of the respective NC distributions, both (NW) and 0er increase only marginally as a function of NC. This saturation arises from working in the high multiplicity tail of the NC distribution where the excitation energy is roughly constant. For lower NC , the average violence of the collision is smaller when changing the gate on NC. But making a higher NC gate does not increase the average 116 100 I I I I E/A = 50MeV _1 Ne 10 C] 2—9 I 10-16 0 17-20 10"2 O 221 10-3 A (a. E§ Z I I I : I \_/ l 0.. E/A = 110MeV 10‘1 "° I3 2-9 I 10-23 0 24-31 10'”2 O 232 10-3 10—4 . . Figure 5.4 Measured IMF multiplicity distributions for the indicated gates on charged particle multiplicity NC. Panels are labeled by incident energy. 8931-88-08" 117 III .0 O 36 Ar+197Au 4~fiTI I I I I I I I l l I I I rTI III I- . .. ‘ I .0. Z 3“ 00 — - 0'0 - /\ ~ 0. - Ln _ o - E 2:. O . - z " IOI'I ~ v : I'O ° ~ ” .580. : 1_ CF? 0 —_ O I - .1— ——P ~0- ~1— —r— —1-— _— Ill- ~— qr— ~0- _— _- .1}- _— -II- _— *I‘ .— 3}- —‘ _ o ' " m h '0';0 . : e2- 00"” —' ._. I 00 E/A(MeV) j b ” 1"..- 035 "' 1:. lb”. .50 _: 60.0.. 251320 Ofll....l....l....l.f O 10 20 3O 40 Figure 5.5 First and second moments of IMF multiplicity distributions as a function of charged particle multiplicity, NC. Different symbols represent results for the indicated beam energies. 118 excitation energy of your event selection, thus giving a saturated value of (NW)- This same information can be plotted as a function of 5 instead of NC and this is done in Figure 5.6. This allows an easier comparison of the different reactions as a function of impact parameter. Most striking is the saturation of (NW) at E/A=50 MeV. This could come from a breakdown in the assumption of the participant spectator picture upon which 5 is based. Another possibility is that NC is not the best measure of centrality at this bombarding energy. As an alternative scale we plot the (MM) and of“, as a function of 5(E,) (Figure 5.7). For this impact parameter scale the saturation in (MW) and 0qu is much less pronounced. In addition, a central cut (5<0.3) using the impact parameter scale 13(E,) selects events with larger numbers of intermediate mass fragments than that from the corresponding cut using 5(NC). The (NW) measured for the two impact parameter scales (from NC and 5,) are similar down to 52.03 for the highest bombarding energies (80 and 110 MeV/ nucleon). For the 50 MeV/ nucleon reaction the measurements diverge for 5=0.4. A saturation in (NW) as a function of 13(E,) is not observed because E, is perhaps a better measure of the violence of the collision than NC. E, is also “Ct a bounded quantity. At very high bombarding energies, NC is limited by For very large NC an auto-correlation between NC and the faCt that NC S Zsystem ° where by (NIMF) is introduced. The obvious extreme example is NC=Z,y,,,,,,, CorlStraint N W =0. 119 36 197 Ar+ Au 4 I r. I I l I I r I I I I I I T I I r I I I r - .~ .. 2 ° 0 . I 3 _QDO o o T . O .. Ah. - o. - E. 2 L O. __i Z 2" ' l I .I.) I V : e : 1 __.. 0c, _ _ Col _ O _— .— nu— HF qr— ‘— up— .- __. _— ‘- .‘F- -n-- .0— —I— .m— r- l 1 C9 C9 GD OD 3_ _‘ me. 3:5. 0 . E/A(MeV) I . O . - tag 27 O O 0' :23 T ;_ "a. __' 1_ (’- _ I 0; : O'....l...,l....l....QQILQ_] 0.0 0.2 0.4 0.6 0.8 1.0 b(Nc) Figure 5.6 First and second moments of IMF multiplicity distributions as a f“fiction of 13(NC). The two circles show the approximate overlap between t"=ll'get and projectile for 5:0.2, 0.4, 0.6 and 0.8. 120 36 197 Ar+ Au 5 it I 1 I I I r I I I I I I I I r I T I I I I I : 4 '0. ‘: I O .. : - O O - Ann. 3 _ Oo .0 —_ 2 o 00. q "' - I oo 3 z 2 :_ I I I I C’s __ V I l I a I : "I : 1 :— ‘3'9' . —.‘. I 0.0. I O .- 1 1 1 1 l 1 I I Fl I I I I l 1 1 1.L 1m: _ I I I I I I I I I r I I I I - I © (9 CD 03 I 3:‘ i l N _Cbo~.. E/A (MeV) I an. _ '0 - I : I I I I I E. I 110 : 1 7 “W. ‘i .. 0:). _ - . - O b 1 1 L l 1 1 1 1 l I 1 14 l 1 1 1 1¢ m (10 (I2 04. (16 013 ‘L0 b(Et) Figure 5.7 First and second moments of IMF multiplicity distributions as a fuIlction of 5(5). The two circles show the approximate overlap between tal‘get and projectile for 13:02, 0.4, 0.6 and 0.8. 121 5.2 Element distributions Element distributions for three different impact parameter cuts are shown in Figure 5.8 for the three bombarding energies. The solid lines are fits for Z=3-20 assuming a functional form for the average yield of N(Z) o. e‘” (5.1) The different distributions are normalized such that N(Z) is the average multiplicity of Z for a given cut on 5. The steeper distributions come from the highest bombarding energy. This can be seen more clearly in Figure 5.9. The fit parameter a is plotted as a function of 5 (constructed from NC) for the three bombarding energies. With this impact parameter scale the flatness of the element distributions (as measured by (1) seems to saturate at about 5(NC)=O.5. For 5(NC)" I : 102 :— o N=67 —: : o N=77 1 - s> - :; Y(A)«A"‘ i 4 ' 1. 101 l l l l I l l l l I l l l l I l l l l o 5 1o 15 20 Figure 5.1] Mass yields for a percolation simulation at fixed excitation energy. The open (solid) symbols correspond distributions constructed for multiplicity N=67(77). The curves are power law fits. 126 ”I I I I I I I I I I I I I I I q t. d r1 2.4 — _ 2.3 _ g i _ ,< - D . . o _ - o . 2.2 — o ._ L. 0 q . O a _ O a I- Q . I”I l 1 l l l l l l 4 I l 1 l l IJ 65 7O 75 80 Figure 5.12 Evolution of the fit parameter A as a function of multiplicity N for a percolation simulation at fixed excitation energy. 127 assigning a fixed number of bonds to be broken (representing a fixed excitation energy) and then randomly breaking the bonds between nucleons in the compound system. Clusters with their bonds intact are identified by mass in each event. Each event is also identified by the multiplicity N , the total number of clusters emitted. In Figure 5.11 is plotted the mass distributions for breakup of an A=233 system at a fixed excitation energy (i.e., fixed number of bonds broken). The two distributions plotted are selected for multiplicities N of 67 (open circles) and 76 (solid circles). The solid lines are a fit with a power law function Y(A) cc A"? The higher multiplicity cut gives a steeper mass distribution even though the two distributions correspond to events with the same excitation energy. Figure 5.12 shows the auto correlation more clearly. The fit parameter A is plotted as a function of multiplicity. Again we observe larger values of 2. (steeper mass distributions) for increasing N. We therefore conclude that the rise of a for very small 13(NC) likely an artifact produced by an autocorrelation of the particular impact parameter filter. This autocorrelation is not apparent for the impact parameter filter based upon the observable E,. 5.3 Angular distributions Angular distributions as a function of bombarding energy and event Cer'ttrality for elements with Z=1-6 are shown in Figures 5.13-15. The distributions are normalized such that the integrated yield (over solid angle aCceptance) is equal to the average multiplicity of a given element for the giVen impact parameter cut. The open circles are the angular distributions for a peripheral cut, 5>0.75, and the solid circles are for a central cut, l3<0.25. The yield for lithium is not plotted for the last two rings. Poor separation of 1ithium and helium in these detectors prevented an accurate measurement of the lithium yield at these backward angles (see discussion in Section 3.3). The 128 fragment yield is suppressed at 9=90° because of target shadowing at this angle. The fragment yield is also suppressed at the most backward angle (9=150°) because of the Pb-Sn foils used to suppress electron detection (see Section 3.1). The slopes of the angular distributions are steeper for peripheral collisions than for central collisions which indicates more equilibration and smaller contributions from projectile-like sources for events with large multiplicities (see Section 4.4). In addition, for central collisions the distributions become less steep as one goes up in bombarding energy. 129 E/A=50 MeV .- I I I I rfle I I I I I I I I I I I I I ITI I I I I I III -1 10’2 r "a 330. E . o O . 10’3 g- 0 .' 0 . ‘2 : ° . O o 3 . o o . u _4 I- o o O . I. .. 10 r O o O O . . '1: E o 0 o o o 5 _5 _ A 00 O o q 10 5‘ . 33(Nc)O.75 : 'L. 10—6 5‘ ° ° ‘21 —7 I . L L I . . 1 . l 1 I . . l . I \ ’ 10-2 _I I I rI IfTII I I I I I I II I I II I I I I I I I I r II I I I I f -1 c: 1° :— z=4 z=6 I "O E. 5 10_4 g9... .- E E 0 ' .0 5 _5 I- O . . 0 I -1 10 E— 0 . . . O 1 E O I I o . :1 : ° 0 . ° . I 10_6 E" O o 0 -§ 5 ° ° 0 o o ' 3 10-7 .- 14LIIILIJLIAITII 1 i llljijbll 1I141?1?1?lIl1-‘ O 50 100 150 0 50 100 150 0 50 100 150 9 (degrees) 1‘I‘igure 5.13 Angular distribution of elements Z=1-6 from the reaction 36Ar+197Au at E/A=50 MeV. Solid symbols - central cut 13(NC)<0.25. Open Symbols - peripheral cut 5(NC)>O.75. dN/dQ OnsnI) 10-5 10-2 36 197 Alri- 130 Au, E/AzBO MeV 9 (degrees) Figure 5.14 Same as Figure 5.13 for E/A=80 MeV. I I III fI I I I I III I I I I I er1 I II III I I I I I I I II I '4 : O .I 8.. i o . I .I E— o . 0 ° . o _5 5 o . ° . o o 3 ,_ o O O I . O. 4 5' ° ° 0 0° 0 ° “3 E I O 0 O O o . o E I' o " E‘ . b(N¢)<0.25 0° 1: : ° b(Nc)>O.75 o 3 F ° 1 5 z=1 2:2 2:3 .3 stale. :1: fits Mtltt .llh.‘ rtlH :1: 'I':‘ h I I I I I I ' I I I ' I ' ~ f 2:4 2:5 2:6 E 5' ‘I :I. .. . : o o 0 EU I.. ... . o I.. 13 I . . ' I .. o . o ' o ° ’ o o o o 1 ° 0 o o ' 2: ° ' 0 ° 3 o o . 0 ° 0 ° o ‘3 ° ° 0 o : o . l l I l l LLLHLATI 1 7 LII l l l 1 L11 ll Lié L1 L111 1?]?lelll l O 50 100 150 0 50 100 150 0 50 100 150 10“2 10-3 1o-4 10-5 A I ._ h 10 6 .U) 33, 10’7 c: 10-2 3 3 0- 2: 1 "o 10“4 10“5 1o—6 10-7 131 36Ar+197Au, E/A= 110 MeV r fl I rI I I I I I I I I I I I I I I I I I I I I IT I I I rI' r7 I I I T I d . ' ‘5 o. : 0 . 0 g... : o O o . I . 1 oo ' 0 ° 0 . '0. i o oo o . o '. . . ° 0 0 ° 0 o 0 -—§ 0 O O O O o I . i A o I o b(Nc)O.75 o o o : 0 2:1 z=2 2:3 I Z=4 2:5 2:6 I ‘5 . o 1 I... . I... . Co... 1: o . o . 5 o . . o . . . a O . O . . .1: O o o o I 2 o o °° o °° ‘ o O o 0 1 O O o O O O O O i l 11 l l l I l l l 1 IT]. 1 l l l l l l 1 l l I l l L 1 ¢ L l l I l l 191? l 9 I; l L‘ 50 100 150 O 50 100 150 O 50 100 50 9 (degrees) Figure 5.15 Same as Figure 5.13 for E/A=110 MeV. Chapter 6 Model comparisons Over the broad range of incident energies in this study of 36Ar+197Au, the mean values and variances of the intermediate mass fragment multiplicity distributions exhibit an approximate scaling with the total charged-particle multiplicity (see Figure 5.5). The relationship between the IMF and total charged-particle multiplicities can be qualitatively understood by assuming that the charged-particle multiplicity is strongly correlated with energy deposition and that the production of intermediate mass fragments depends primarily upon this energy deposition. The nearly universal increase of (NW) as a function of NC can thus be viewed as due to the selection of interactions involving a progressively increasing amount of internal energy deposition. Deviations from a universal relationship between (NW) and NC can arise from different amounts of preequilibrium emission and/ or angular momentum transfers to the equilibrated nuclear systems at the different incident energies. In addition, different impact parameters can be expected to involve decaying systems of different mass. The measured multiplicities of light particles and intermediate mass fragments are compared in this chapter with both a model involving statistical decay of an expanding compound nucleus and with a model involving percolation theory. 132 133 6.1 Expanding emitting source model We have compared the experimentally measured IMF multiplicity distributions with the predictions of the expanding emitting source model of Friedman [Frie 90]. This schematic model couples the phase-space features of statistical decay with the dynamical features of expansion driven by thermal pressure. It has the capacity for predicting multiplicity distributions for different species of particles (IMFs in particular). The model has been successfully employed [Frie 88] to interpret the low emission temperatures deduced from the relative population of states [Chen 87], as well as trends in IMF multiplicity with excitation energy [Troc 89] where the IMF multiplicity was found and predicted to be less than one [Frie 90] (lower than predicted by the instantaneous breakup model of [Bond 85b]). The model also predicted [Frie 90] a sharp rise in multiplicity for excitation energies on the order of 8 MeV/ nucleon. The expanding emitting source (EES) model characterizes an ensemble of emitting sources by a time-varying average density p(t) to which the level densities of the source are calculated via the Fermi-gas approximation. For the purpose of calculating the mean collective expansion energy, the model assumes that the density of the source is uniform. This assumption requires that the collective radial velocity increase linearly with radius. In the EES model, the dynamical response of the source is governed by the interplay of thermal pressure and the nuclear binding forces, which tend to return the density to its equilibrium value p0. The binding effects are parameterized in terms of the finite-nucleus compressibility. For simplicity, the binding energy per nucleon at densities different from p0 is assumed to deviate from liquid-drop values by a quadratic function of density, 134 A A 18 p0 ' where the first term on the right represents the liquid drop binding energy values (~-8 MeV) and K is the finite-nucleus compressibility coefficient. This compressibility includes surface and Coulomb effects. The calculations require an assumption of initial source mass, charge and thermal excitation energy when the system is at normal density. For simplicity we assumed the decaying source to be the full composite system (A=233, 2:97), and we further assumed that the system expanded with no initial expansion velocity from normal density. For different compressibilities we calculated the correlation between the predicted total charged-particle and IMF multiplicity distributions. The results are shown by the curves in Figure 6.1 (the data were described previously in Chapter 5, Figure 5.5). For orientation, the upper panel of Figure 6.1 includes an approximate scale of the relation between the excitation energy of the emitting system and the mean charged-particle multiplicity calculated from the BBB model. In the extreme tails of the NC distributions, the correlation between internal energy and NC becomes dominated by fluctuations of the charged-particle multiplicity. Hence, very large values of NC become ineffective in selecting nuclei of increasing internal energy thus causing the observed saturation of (MW) and of”, at large values of NC (see discussion in Section 5.1). The calculations are sensitive to the nuclear compressibility at low density. The solid, dashed and dashed-dotted curves in Figure 6.1 show predictions for the relationship between the (Nm)and (NC) for finite-nucleus compressibilities of K=144, 200 and 288 MeV, respectively [Blai 80]. IMF 6 .- I I I I I I I I I I I I I r I I I 7 l’ I :‘ : 5 1. 1.5 2 2.5 d g 5 :— l l l l . f j In : E. (08V) ..-'// j ('0 .. / ., A : 3“... : In . ., . a 3 :— 0‘? ’9" 1 6 2: .O' .............. 1 1 .— fl m 1 14 I 1 1 1 1 l l 1 1 L l i L: O l. I I I I I I I I I I I l r I I I lfi I‘ - 1 3 ‘_ _. NEH " 'l 2 2 r- 1 6" . -< 1 _— T O ‘ L 0 Figure 6.1 First and second moments of IMF multiplicity distributions as a function of charged-particle multiplicity, NC. Different symbols represent results for indicated beam energies. The solid, dashed, dotted-dashed and dashed-dotted-dotted curves show results calculated for the statistical decay of expanding compound nuclei of finite-nucleus compressibility K=144, 200, 288 and co, respectively. The dotted curves represent the calculations for K=200, filtered by the detector response. 136 multiplicities for a non-expanding compound nucleus, corresponding to the limit K—->oo, are shown by the dashed-dotted-dotted curves. To illustrate instrumental distortions, the dotted curves show the calculations for K=200 MeV, filtered by the response of the experimental apparatus. At low multiplicities, corresponding to low excitation in this model, the nucleus does not expand and all calculations predict IMF multiplicities consistent with the measured values. For multiplicities larger than NC =20, however, expansion strongly influences the predicted number of clusters in the final state. Large observed IMF multiplicities, comparable to the measured values, are only predicted for an equation of state that is sufficiently soft to allow the nucleus to expand in response to thermal pressure. Very stiff low-density equations of state hinder the expansion of the system leading to a suppression of the production of multifragment final states and a consequent underprediction of the observed mean IMF multiplicities by nearly a factor of two. The model ignores angular momentum, fluctuations in source size and excitation energy, and the lack of thermal equilibrium. While the inclusion of such effects might affect the final observations, we nonetheless find that the essential features of the observed data are included in the predictions of the schematic model shown by the curves in Figure 6.1. The sensitivity to the source size was explored by changing the initial mass and charge by 20%. The qualitative conclusions remain valid despite this change. The calculations with the schematic model suggest that multifragment decays of highly-excited nuclear systems may exhibit considerable sensitivity to the low-density nuclear equation of state. They indicate that the expansion dynamics, which is governed by the compressibility, may be intimately connected to the production of IMFs. A more quantitative exploration of 137 these properties, however, will require a more complete model which also incorporates non-equilibrium effects. 6.2 Percolation In the search for a signature of the liquid-gas phase transition in low density nuclear matter, percolation models are attractive since they exhibit a well-defined phase transition for infinite systems and since they allow straightforward generalizations to finite systems and the incorporation of important geometrical ingredients [Biro 86] for multifragmenting systems. In rather general terms [Siem 83, Hirs 84], the fluctuations at the critical point of the nuclear matter phase diagram are expected to lead to mass distributions which follow a power law, o(A)ocA"‘, with a critical exponent of the order of 1:22-23. A number of theoretical investigations of phase transitions in finite nuclear systems have been based on percolation models [Baue 85, Baue 86, Baue 88, Camp 86, Camp 88, N g6 90, Biro 86, Jaqa 90]. Most percolation models are governed by a single bond-breaking or site-vacancy parameter and cannot be expected to reproduce the two- dimensional phase diagram of nuclear matter in the temperature vs. density plane. Despite this limitation, they have been rather successful [Baue 85, Baue 86, Baue 88] in describing the observed [Hirs 84] power-law behavior of measured fragment mass distributions and in developing techniques to extract critical exponents from exclusive fragmentation data [Camp 88, N g6 90]. In this section, we provide a test of the bond percolation model of references [Baue 85, Baue 86, Baue 88] and compare its predictions to fragment yields measured for the reactions 129Xe+197Au at E/A=50 MeV and 36Ar+197Au at E/A=50, 80 and 110 MeV. 138 Calculations were performed with the bond percolation model of ref. [Baue 85, Baue 86, Baue 88]. In this model, the nucleus is considered to be a cubic lattice, the sites of which are randomly occupied by protons and neutrons within a spherical volume of radius R ——= (:3; A)M a, where a is the lattice parameter (a = p?” z 1.8 fm). Initially, all nucleons are connected in one cluster. The bonds between the sites are randomly broken with a probability p. Each nucleon is assigned a random momentum consistent with the momentum distribution of a Fermi-gas of temperature T estimated [Baue 88] from the bond-breaking probability as T=11.7JE MeV (see Appendix B). Emitted fragments are defined in terms of connected clusters. Initial fragment energies are calculated from the total momenta of the clusters. Final kinetic energies are calculated by incorporating the final state Coulomb repulsion between the fragments. For this purpose a given fragment partition is translated into a spatial distribution of clusters characterized by an average freeze-out density p=0.2p0 , where p0 =0.17 fm'3 is the density of normal nuclear matter. (Different choices of the freeze-out density lead to slightly different shapes of the low-energy portion of the energy spectrum due to changes in the Coulomb repulsion between the fragments. These Coulomb barrier fluctuations are of minor importance in the present context.) The final momenta of the emitted fragments are boosted by the velocity of the center- of-mass of projectile and target. In order to allow meaningful comparisons with our data, we have filtered the theoretical distributions with the response of the experimental apparatus. Elemental multiplicity distributions measured for the reactions 36Ar+197Au and 129Xe+197Au are shown (as points) in the top and bottom panels of Figure 6.2. These distributions were obtained by integrating all identified fragments over all detectors. In order to select central collisions, 139 MSU-92-019 I I I l I I I I I I I I I I I I I I I ”A: + "'Au. E/A = o 50MeV, Mezzo A 100 \ “I.” o aouev, Ncgze g 1 \\ o 110MeV, chao _. \ ‘ 10 \ ‘ - ~.__ \ ~~~~~~~~ - \ 10—2 \ \ \ \ \ ...... e a 10_3 : % = I if i : : I : l l [_— h l l l 1" I I I I I l I I I r 101 ‘9. mice + 1""111. E/A = 50MeV \ kn, o Nc233 A 100 \ \OKI, . v \ ‘°"\., . . . z 10‘1 \ "\"r.°° ‘,~ ~.-. . - -P=0.8 \\ 'O-g- - _2 P=O.8 \ \ . 1° —P=0.7 \ g " _ . -p=o.7 (unfiltered) \ 10—3 I I I I L I I L I I I I I LL I I I 1 l g 0 5 10 15 20 Z Figure 6.2 Elemental multiplicity distributions detected in 36Ar+197Au collisions at E / A=50, 80, 110 MeV (top panel) and in 129Xe+197Au collisions at E/A=50 MeV (bottom panel). The curves represent calculations with the bond percolation model (described in the text) for the indicated bond-breaking probabilities p. All calculations are filtered by the response-of the experimental apparatus except for the dot-dashed curve. 140 these Z-distributions were selected by the total charged-particle multiplicity cuts indicated in the figure. When one adopts a strictly geometric interpretation of the measured charged-particle multiplicity [Cava 90, chapter 4], these cuts were chosen to represent the range of impact parameters of 550.3 . The elemental distributions observed for the various reactions exhibit rather similar shapes. For the 36Ar+197Au system (top panel), the elemental multiplicity distributions become slightly steeper with increasing bombarding energy. The multiplicities for heavier IMFs are significantly larger for the 129Xe+197Au system (bottom panel) than for the 36Ar+197Au system (top panel). Part of this difference in the observed yields of heavier fragments may be due to an increased detection efficiency in 129Xe+197Au reactions resulting from the larger center-of—mass velocity. The curves in Figure 6.2 depict elemental multiplicities predicted by the standard bond percolation model and filtered by the detection efficiency of the experimental apparatus. Calculations are shown for representative bond- breaking probabilities above and below the near-critical bond-breaking parameter of p=0.7. (In the percolation model, the critical point marks a second order phase transition: for p>0.7, the percolation cluster disappears and the system breaks up completely.) Calculations performed for bond-breaking probabilities much larger or smaller than the critical value predict Z- distributions which are too steep. For the 36Ar+197Au reactions, the overall magnitudes and shapes of the experimental Z-distributions are in reasonable agreement with predictions of the percolation model when the bond-breaking parameter is taken close to the critical value. However, for the 129Xe+197Au reaction, the percolation model underpredicts the yield of heavier fragments (Z=6-20) for any choice of bond-breaking parameter. 141 Representative angular distributions of emitted particles are shown in Figure 6.3. The top and bottom panels show results for the 36Ar+197Au reaction at E/A=110 MeV and for the 129Xe+197Au reaction at E/A=50 MeV, respectively. The angular distributions are shown for three different ranges of element numbers, Z=1-2 (circles), Z=3-S (squares), and Z=6—12 (diamonds). For both systems, the angular distributions become more forward peaked with increasing fragment charge. Angular distributions for the 129Xe+197Au system are more forward peaked than those for the 36Ar+197Au system. These effects are largely due to kinematics. Differences between the angular distributions of the two reactions arise primarily from the larger velocity of the emitting source of the 129Xe‘I-197Au system. The curves in Figure 6.3 show results of percolation calculations. To facilitate a better comparison of shapes between observed and predicted angular distributions, the calculated angular distributions were normalized to the experimental yields at 6=45°. To display the effects of the detector response, filtered and unfiltered calculations are shown by solid and dashed curves, respectively. In view of the fact that the percolation model does not include dynamical preequilibrium effects, the shapes of the experimental angular distributions are reasonably well reproduced by the calculations. Discrepancies between theoretical and experimental angular distributions due to preequilibrium emission are most pronounced for lighter fragments emitted at forward angles. For such fragments the measured angular distributions are slightly more forward peaked in the laboratory than the calculated angular distributions. Similar discrepancies exist for other treatments in which statistical equilibrium is assumed. For this purpose, the calculated angular distributions may be sufficiently realistic to assess effects of 142 1 HSU-QZ-OO 1 10-. “ l I I FT] I I I I I T’I I I I I II 3'Ar + mAu 10"2 '0 10‘3 10’4 .. 12.}(6 + "7A1: E/A = souev. chss dN/do (msr’l) 10"4 . I 3-5 0 6-12 10"5 - - P-0.7 — P-0.7 (filtered) 6 lab Figure 6.3 Angular multiplicity distributions of light particles, Z=1,2 (circles), and intermediate mass fragments of Z=3-5 (squares) and Z=6—12 (diamonds) detected in 35Ar+197Au collisions at E / A=50, 80, 110 MeV (top panel) and in ‘ 129Xe+197Au collisions at E/A=50 MeV (bottom panel). The curves represent calculations with the bond percolation model for a bond-breaking probability of p=0.7. The calculations have been normalized to the data at 9=45°. Dashed and solid curves show raw calculations and calculations filtered by the response of the experimental apparatus. 143 instrumental distortions on the energy- and angle-integrated particle distributions presented in Figures 6.2 and 6.3. In order to display more clearly the fraction of IMFs among the emitted charged particles, Figure 6.4 presents the average IMF multiplicity, (Mm), as a function of charged-particle multiplicity, NC [Bowm 91] (see Figure 5.5). Points in the figure show the average IMF multiplicity (MW) as a function of charged-particle multiplicity NC. For a given charged-particle multiplicity, more intermediate mass fragments are observed for the 129Xe-I-197Au reaction than for the 36Ar+197Au reactions. At large charged-particle multiplicities (NC 230), IMF admixtures of (NW) /NC =0.18 and 0.1 are observed for the systems 129Xe+197Au and 36Ar+197Au, respectively. If the elemental distributions strictly followed a power law distribution, 0(2) = 0'02", the IMF-admixture would be determined by the exponent 1:: 20 002" ”mp = 2:3 N Zw C 2 0.02-1 2:1 For power law distributions, the (raw detected) ratios of (N,,,F) / NC =0.18 and (6.2) 0.10 correspond to exponents of 122.15 and 2.6 (see Figure 6.5). If we use the shapes of the energy and angular distributions predicted by the percolation model to correct for the detector efficiency, we obtain efficiency corrected values of (N,MF)/NC =0.19 and 0.116, corresponding to 1:--2.12 and 2.5. These values are slightly below and above the value, “rm-32.2, at the critical point. According to Figure 6.2, the percolation model underpredicts the yield of IMFs with Z>5 for the 129Xe+197Au reaction. This failure is displayed more clearly by the curves in Figure 6.4 which show maximum IMF admixtures predicted by the percolation model. Thick and thin curves represent results of filtered and unfiltered percolation calculations, respectively, using the near- ‘Y I Figu dElEI Open E/A. and l 144 MSU-92—055 I I I I I I I I I W I I I -‘ l. o ‘”Xe+""Au, E/A = sonev ,' . 8 __. o “A:+‘°"Au. E/A = 110MeV ,.’ fl _ o ”Ar+‘"Au. E/A = aouev l.’ - . o ”Ar+‘°"Au. E/A = souev q .r' . 6 — _. A r- _ In - .. a I z «- v 4 ~— — 2 — _ - 197+129 . . Cb . 0 l L l l l 1 1 1 1 1 1 J_ l 0 20 40 60 Figure 6.4 Relation between average IMF and charged-particle multiplicities detected in 35Ar+197Au collisions at E/A=50, 80, 110 MeV (open diamonds, open squares, and open circles, respectively) and in 129Xe+197Au collisions at E/A=50 MeV (solid circles). Thick and thin curves show the results of filtered and unfiltered percolation calculations, respectively. Details are given in text. 145 0.25 0.20 0.15 0.10 NIMF/NC |llllllJlLLLlllLllll 0.05 IlIfilTIIIIjIIIIIIIII llll 0.00 llllllllllllllLlLJ 2.00 2.25 2.50 2.75 3.00 T )— Figure 6.5 IMF admixture as a function of 1, according to Equation (6.2). 146 critical bond-breaking parameter, p=0.7. (For the percolation model, these calculations give upper bounds for the admixture of IMFs among the emitted charged particles. Smaller IMF admixtures can be obtained by using larger or smaller bond-breaking parameters.) The dashed curves in Figure 6.4 show percolation calculations for the combined 36Ar+197Au system. For this system, filtered and unfiltered IMF admixtures are very similar, and the relative abundance of Ms observed in central collisions can be reproduced by the model calculations. The solid curves represent percolation calculations for the combined 129Xe+197Ausystem. Here, the filtered calculations represent slightly higher IMF admixtures than the unfiltered calculations. This effect is largely due to an increased IMF detection efficiency resulting from the larger center-of-mass velocity of the 129Xe+197Au center-of—mass system. However, both filtered and unfiltered calculations predict IMF admixtures that are too small. This failure is most dramatic for central 129Xe-I-197Au collisions for which the comparison with an equilibrium model is most meaningful. As an alternative scenario, we have also performed percolation model calculations for the separate multifragment decay of excited projectile and target nuclei. The unfiltered calculations, shown by the thin dot-dashed curve, predict slightly higher IMF admixtures. The thick dot-dashed curve illustrates the effect of filtering for an extreme two-source scenario in which the relative velocity of projectile and target was reduced by only 50% from the initial value, taking total momentum conservation into account. Even in such an extreme scenario, the major discrepancy remains. High-resolution coincidence experiments indicate that a significant portion of primary fragments can be expected to be produced in highly excited, particle unbound states which decay by light particle emission [N aya 92]. Such sequential decay processes will result in secondary fragment yields which are 147 smaller and secondary light particle yields which are larger than the corresponding primary yields. The portion of primary fragments may therefore be even larger than the portion of particle-stable secondary fragments. This aggravates the failure of the percolation model to predict the large proportion of intermediate mass fragments among the particles emitted in the 129Xe+197Au reaction. The inability of the bond-percolation model to reproduce the large intermediate mass fragment multiplicities observed for the 129Xe+197Au system is unexpected and represents, to our knowledge, the first significant failure of the percolation model. This model is only one representation of a large number of phase transition models which all belong to the same universality class and should therefore show similar deficiencies. At present it is not clear whether one can rule out all such models or whether one may be forced to consider dynamical enhancements of fragment yields due to collective expansion or rotation. Toroids and bubbles In the bond-percolation calculations of the previous section, the disintegrating systems were assumed to have compact spherical configurations. Recent microscopic transport calculations [Baue 92, More 92] indicate, however, that multifragment decays may proceed via more complex toroidal or bubble-shaped decay configurations. In this section, we employ the bond-percolation model used previously to investigate how multifragment disintegrations might be affected by the occurrence of ring- and bubble-shaped decay configurations. All calculations presented in this section were performed for a system consisting of A=N+Z=250 nucleons, with Z=102. The decaying system is a 1 F0 ad me sol per nea ma; 3 CC mu] thic. line eXpE for t: Prob; tOl’OiI the 5 defim' 148 represented by those points on a simple three-dimensional cubic lattice which fall within the specified decay volume. Each site represents a nucleon that is "bonded" to its nearest neighbors as described in the previous section. In the following discussion of non-compact breakup geometries, instrumental distortions will be ignored since we wish to outline some general trends without trying to fit a specific set of data. Nevertheless, it is useful to provide a reference which allows the reader to gauge the magnitude of various effects. For this purpose, the solid points in Figures 6.6 and 6.7 show the fragment admixtures (i.e. the mean number of detected intermediate mass fragments, (N,,,F), as a function of the detected charged-particle multiplicity, NC) measured [Bowm 91] for the 129Xe + 197Au reaction at E/ A = 50 MeV. The solid lines in the figures show previous calculations with the bond percolation model for a compact spherical breakup configuration using a near-critical bond-breaking parameter p = 0.7. These calculations represent the maximum fragment admixtures predicted by the bond-percolation model for a compact spherical geometry; they underpredicted the measured fragment multiplicities (see previous section). The difference between the thin and thick lines illustrates the magnitude of instrumental distortions. The thin line represents the "raw" calculation (not corrected for the acceptance of the experimental apparatus), and the thick line represents the calculation filtered by the acceptance of the experimental apparatus. The hatched area in Figure 6.6 shows fragment admixtures predicted for toroidal breakup configurations. In these calculations, the bond-breaking probabilities were varied between p = 0.5 - 0.8, and the central radii of the toroids were varied between Rt = (2.0 - 4.5)xa, where a = p01 / 3 ~ 1.8 fm denotes the spacing between adjacent lattice sites (see insert in the figure for a definition of the geometry). Because of volume conservation, the thickness d 149 MSU-93-039 10 h I I I T‘r I I I T I I I I 77 I I I I I H : o 12"Xe+""’Au, E/A=50 MeV : " sphere, A=326, p=0.7 « 8 — fl : 1 /\ ‘_ _“ k. 6 E ' 1 Z _ toroid . V - A=250 — 4 _— P=0.5-0.8 —_ _ R,=2.o-4.5 . 2 ‘ CH9 ‘ F / a l l '.l l I l L l l l l l I l L L l W O I | 0 20 4O 60 80 Figure 6.6 Relation between average IMF and charged-particle multiplicities. Solid points represent values measured for 129Xe + 197Au at E/A=50 MeV. Thin (thick) solid line shows the raw (efficiency corrected) percolation calculation for a solid sphere. The hatched area shows the range of average IMF and average charged-particle multiplicities predicted by percolation calculations for toroidal breakup configurations. 150 MSU—93—040 10 '- l I I I l I I I I \ I I I I 4 I o 12")(e+“"’Au. E/A=50 MeV \ Z - sphere, A=326, \ I 8 _— p=m \ ‘1 _ %-\\u . - §§ \\ . A 6 — §~ § 3 — in P §§5’ E . 2 § §:: E . z : _5. $3: bubble 3‘ ‘ V - .5 §f A=250 ‘3 4 — Q; P-O 5—0 a I 5%; - ' ‘ - ' r§§ Rb=0.0-4.0 2 r—- § > — L ~§ . .. § - . § . O b L .1 ’ 1 I l g l 1 l 1 1 l L l l l 1 l ‘* 0 20 40 60 80 NC Figure 6.7 Relation between average IMF and charged-particle multiplicities. Solid points represent values measured for 129Xe + 197Au at E/A=50 MeV. Thin (thick) solid line shows the raw (efficiency corrected) percolation calculation for a solid sphere. The hatched area shows the range of average IMF and average charged particle multiplicities predicted by percolation calculations for bubble-shaped breakup configurations. 151 of the toroid is defined by Rt and the nucleon number: A a antdz/ 2. For A=250, the toroid has a hole in the center only for Rt > 2.5xa. The upper boundary of the hatched area is determined by the breakup of a toroid of radius Rt = 4.5xa, and the lower boundary represents the breakup of an oblate object of Rt = 2.0xa. Over the range of NC z 30 - 50, the percolation model can produce significantly larger fragment multiplicities for toroidal than for spherical breakup configurations. Qualitatively such an effect may be expected since the surface of a toroid is larger than that of a sphere. Enhanced fragment admixtures can also be obtained for bubble-shaped breakup configurations. In Figure 6.7 the hatched area shows the range of average IMF multiplicities predicted for bubble shaped density distributions using bond-breaking probabilities of p = 0.5 - 0.8 and inner bubble radii of Rb = (0.0 - 4.0)xa (see insert in the figure for a definition of the geometry). For A = 250 and Rt > 4.0xa, the bubble has a thickness less than 1.0xa, and the simulation models the breakup of a thin sheet. The upper rising boundary of the hatched area corresponds to calculations with a fixed bond-breaking probability of p = 0.5 and varying inner radii, Rb = (0.0 - 4.0)xa. The falling part of the upper boundary represents calculations for an inner radius of R], = 4.0xa and varying bond-breaking probabilities p = 0.5 - 0.8. For (NC) 5 70, the lower boundary is given by the breakup of a sphere with varying bond-breaking probabilities, p = 0.5 - 0.8. The results in Figures 6.6 and 6.7 demonstrate that objects with larger surfaces can produce more fragments than objects with smaller surfaces. Hence, geometrical considerations may play an important role for multifragment disintegrations. A relatively cool object (small bond-breaking probability) with a large surface may decay into more fragments than a hotter object (larger bond-breaking probability) with a smaller surface. For 152 noncompact breakup geometries, the large fragment multiplicities observed for the 129Xe + 197Au reaction can be reconciled with predictions of the bond percolation model. A number of investigations have aimed at obtaining information of near-critical behavior from the shape of fragment mass or charge distributions [Gros 90, Pana 84, Mahi 88, Ogil 91, Li 93]. Near the critical region, scaling theory of large systems predicts mass distributions of the form: N409) .. A"f(A"(p - 19.)). (63) Here, 1: and o are critical exponents and f (A°( p — 19.)) is a scaling function that modulates the power law behavior near the critical bond-breaking probability, p=pc, above which the "infinite" percolation cluster ceases to exist. In the Fisher droplet model f is an exponential so that NA (P) °¢ A.f exp(const- A"(p - pc )). (6.4) f has the appropriate limiting behavior of f =1 at p = Pa so that N A cc A" In practice, the mass or charge distributions are often fit [Pana 84, Mahi 88, Ogil 91, Li 93] by a simple power law, Y(A) .. A“, (6.5) where A is treated as a fit parameter, and the critical exponent 'c is identified with the extracted minimum value of A [Baue 85]. We will now show that this empirical approach can lead to misleading results if the geometry at breakup is not compact. For this purpose, we performed power law fits to the mass distributions predicted by the bond percolation model over a broad range of parameters and geometrical configurations. The results of these calculations are summarized in Figures 6.8 and 6.9 for toroidal and bubble-shaped geometries, respectively. In both cases, the best fit-parameter A. exhibits a clear valley as a function the bond- 153 toroid, A=250 Figure 6.8 Extracted power—law exponents 7L fit to mass distributions predicted by the bond percolation model for the break up of toroidal systems as a function of Rt and p. wo-co-nsw 154 bubble, A=250 2.75 2.5 2.25 Figure 6.9 Extracted power-law exponents 7» fit to mass distributions predicted by the bond percolation model for the break up of bubble shaped systems as a function of Rb and p. zvo-co-nsw 155 MSU-93-043 3 5 L: I l I I I I I I I I r I I I I I l r _J - A=250 - ' o sphere . ‘- 3 0 ’_ g o toroid R.=3.0 _: ° _ g g I bubble Rb=3.0 - _ I . < 5 . I « 2. — -— _ I 0 fl - § § ' . I O . I- § . s a —I " I 3 3 ° ‘ 2.0 — _ . — h- . . - - -1 1 5 ln] 1 l l l l l [A l l L L l l l J l 1 Id 0 5 0 6 0.7 0 8 P Figure 6.10 A as function of p for a solid sphere (circles), a toroid of radius Rt = 3.0xa (diamonds), and bubble with inner radius Rb = 3.0xa (squares). 156 breaking (p) and geometry (Rt and Rb) parameters. In this valley, smaller values of 1 occur for less compact breakup geometries. The strong geometry-dependence of the relation between the power- law exponent A. and the bond-breaking parameter p is depicted more clearly in Figure 6.10 for three representative breakup configurations, a solid sphere (solid circles), a toroid of central radius Rt = 3.0xa (open diamonds) and a bubble with inner radius Rb = 3.0xa (solid squares). Microscopic transport calculations [Baue 92, More 92] predict a strong dependence of the breakup geometry upon beam energy, impact parameter and projectile-target combination. Indeed, under favorable conditions, the formation of unstable bubbles and rings has been predicted [Baue 92, More 92]. The present model calculations for finite systems indicate a strong dependence of extracted "critical" parameters (1 and pc) on the geometrical configuration of the system at breakup. Compilations of power law exponents 1 determined from data for different entrance channels [Pana 84], for different impact parameters [Ogil 91] or for excitation functions covering broad ranges of energies [Pana 84, Mahi 88, Li 93] may contain samples representing sufficiently different geometrical configurations to render a minimum in A difficult to interpret. For infinite systems, critical exponents govern the scaling laws near critical points. We suspect that the application of scaling laws to finite systems of potentially complex breakup geometries is much less straight forward than originally surmised [Camp 92]. Chapter 7 Fluctuations in multifragment emission A recent analysis [P105 90] of fragment-size distributions observed in reactions induced by gold on emulsion at E/A=1 GeV [W add 85] saw evidence for intermittency, which might indicate that fragmentation processes are scale invariant. We take up the problem of intermittency by analyzing the factorial moments of charge distributions observed in 36Ar+197Au reactions at beam energies between 35 and 110 MeV per nucleon [deSo 91]. The occurrence of intermittency is deduced from the factorial moments [Bial 86, Bial 88] 2% 2(N,(N, -1)...(N,. -k + 1)) F.(A)= z.) 201,)" i=1 (7.1) where 20 is the total charge of the disintegrating nuclear system, A is a binning parameter, and N,- is the number of fragments with charges in the interval (i- 1)A <2 5 iA where i=1,...,Zo/ A. The ensemble average < > is performed over all fragmentation events considered. Intermittency is defined by a relation [Bial 86, Bial 88] F,(A’) = F,(aA) = a‘“’"Fk(A) (7.2) 157 158 between factorial moments F ,(A’) and F ,(A) obtained for two different binning parameters A and A’ = 0A. Generally, evidence for intermittency has been obtained by examining the double logarithmic plot of lnF, versus -lnA. Plots which show lines of positive slope are consistent with nonzero fractal dimension and considered to display intermittency. By construction, the factorial moments F k (A) are unity for Poisson distributions. Hence, Poisson distributions do not exhibit intermittency. The expression for the second factorial moment can be written as F2(A) = 1 +2410} — W» (7.3) 2.002 where (N,)and of denote the mean value and the variance of the multiplicity distribution in the ith bin. Distributions which are narrower (broader) than Poisson distributions possess moments which are less (greater) than unity. For A = Z0, the right hand side of Equation (7.3) reduces to 1+(ofi —(NC))/ (NC)2, where (NC) and of; are the mean value and the variance of the charged particle multiplicity distribution. Intermittency as an indicator of nontrivial physics is generally sought for in systems exhibiting larger than Poisson fluctuations. Evidence for such large fluctuations must be sought in events representing similar initial conditions. However, constraints from conservation laws may lead to reduced fluctuations because statistical independence of individual bin occupations is lost. As the number of bins becomes larger, the individual bin occupations may become more independent, and the moments may increase. The corresponding rise in the factorial moments as a function of decreasing bin size would be of little interest. In order to explore whether there is a basis for non-trivial intermittent behavior in a reaction in which multifragment emission has been observed, 159 we analyzed the first and second moments of the charged particle multiplicity distributions measured (see Section 5.2), with the Miniball [deSo 90] for 36Ar-I-197Au collisions over a broad range of energies, 355 E/ A5110 MeV. Event selection was performed by cuts on the total transverse energy, E, = E sin2 0, of the emitted charged particles. For orientation, we also provide an empirical impact parameter scale by using the geometrical prescription of Section 4.2 to construct a ”reduced” impact parameter IS. The reduced impact parameter assumes values of I; zl for peripheral collisions and 5 z 0 for the most violent collisions characterized by large values of E,. The top panel in Figure 7.1 shows the measured two-dimensional correlation between transverse energy E, and charged particle multiplicity NC for 36Ari-197Au collisions at E/A=110 MeV. The dashed and dot-dashed curves in the bottom panel of Figure 7.1 depict charged particle distributions selected by narrow cuts on E,, corresponding to reduced impact parameters 5:101 and 0.6. The dotted curves illustrate the effects of increasing the widths, A5,, of the cuts as indicated by dotted horizontal lines in the top panel. For central collisions, the NC distribution is rather insensitive to AE,, but for more peripheral collisions it suffers considerable broadening as AB, is increased. The broadening due to impact parameter averaging is illustrated more quantitatively in Figure 7.2. Top, center and bottom panels of the figure depict the quantities 02 /(NC), a: and (NC), respectively, as a function of transverse energy. Solid circular, open square-shaped and star-shaped points show values obtained for cuts of widths AE,=20, 180 and 340 MeV, respectively. For narrow cuts on AE,, the charged particle multiplicity distributions are inconsistent with F2(A = 20) > 1, since of. /(NC) < 1. For near-central collisions, 5<0.3, the extracted values of of. / (NC) exhibit little dependence on the 160 MSU-92-038 12507....1. "' "' H'flfl 1000' — = = = 0.1 I?) 750 ‘ 0.2 ‘2: _ «0.4 m 500 ............... I b -:::::::::= 06 250 """""" ' """ ‘ 0.8 O . l *t t q 0.10 _ € 0.05 " ‘ 000 .u-;./l.. . Li". r\n'-°1'uuL~L 1 .l'x-L‘ ‘ 0 10 20 30 40 50 Figure 7.1 Upper part - Measured relation between transverse energy E, and 0") total charged particle multiplicity for 36Ar+197Au reactions at E/A=110 MeV. Lower part - Charged particle multiplicity distributions for the cuts on E, indicated in the top panel. 161 2.0 I ' . g " 1 M I ‘3] I? A 1.57 \ X fl co 0 .. ' . N Z : 5’ ‘- c'; \/ 1.07 \ 1 21’ \ : fix. Nb!) _ \0\71\ d 0.5 I All IIUU'I'UI 2 C H 01 II All}.a 0 ZOMeV I 0 1801191! )1 340MeV I UIVUTUIUU IU'UIVIIUIYIU' I II 8 'l 0'.DLL...L...1..11.AJ 0 200 400 600 800 1000 Et(MeV) Figure 7.2 Bottom, center and top panels show the mean values (NC), variances oi , and ratios 0% / (NC) of the charged particle multiplicity distributions for 36Ar+197Au reactions at E/A=110 MeV. These quantities were selected by various cuts on the transverse energy; the mean values of these cuts are given by the abscissa and the widths are given in the figure. The upper scale gives the reduced impact parameter 5(E,) . 162 widths of the applied cuts. However, for larger impact parameters, 13>0.4, wide cuts on E, cause an artificial broadening of the multiplicity distributions resulting from the superposition of distributions with different centroids. Poorly defined ensembles of events may therefore exhibit larger than Poisson variances, of. /(NC)>1. However, these large variances are an artifact from impact parameter averaging, and they do not represent intrinsic fluctuations of the decaying system. Figure 7.3 depicts the relation between (NC) and 0% / (NC) extracted for central (13<0.3) 3'5Ar+197Au collisions at the incident energies of E/A=35, 50, 80 and 110 MeV. At all energies, the fluctuations of the charged particle multiplicity are considerably smaller than expected for Poisson distributions. In order to explore effects resulting from phase space constraints such as energy conservation, we performed calculations with the bond-percolation model of refs. [Baue 85, Baue 86, Baue 88]. For simplicity, we assumed the decay of the composite system (A=233, Z=97). Calculations with bond-breaking parameters close to the critical value of p=0.7 have already been shown (Section 6.2) to reproduce the element distributions measured for the present reaction. Standard percolation calculations (in which the number of broken bonds is allowed to fluctuate from event to event) predict fluctuations in NC which are much larger than observed experimentally. Open circular points in Figure 7.3 show representative results for p=0.6 and 0.7. The indicated shift from the open circle to the open triangular point illustrates the magnitude of instrumental distortions for the case p=0.7. These distortions are too small to affect our conclusions. If one introduces a constraint analogous to energy conservation by requiring a fixed number of broken bonds, much narrower charged-particle distributions are produced (see open diamonds). These illustrative calculations suggest that the widths of impact-parameter—selected 163 I IFIII I I I I II I I [I IIIII I I K .. a g 1.0— - | . O 0 .§ b / " C CD .- u O. 0.8"“ A - A . -I U z 0.6— - v r , \ . o . NCO . I 0.4. . . O . o o o . ' 'l 0.2- )1 31 .- . x0 . . [31:1 . 00 1 lklllLLl [1 L1 1 LLI 111.1 1 1 1 0 10 20 30 40 50 No Figure 7.3 Relation between mean charged particle multiplicity (NC) and the ratios of. / (NC). Solid circular points — experimental values extracted from near-central 36Ar+197Au reactions at E/A=35, 50, 80 and 110 MeV. Open symbols are explained in the text. 164 charged-particle-multiplicity distributions are strongly affected by phase space constraints due to energy conservation. This conclusion is corroborated by more realistic statistical model calculations which incorporate energy conservation on an event-by-event basis. The open square and star-shaped points show predictions of the sequential decay model GEMINI [Char 88] and of the Copenhagen fragmentation model [Bond 85a, Bond 85b, Barz 86], respectively. (The individual points represent results obtained for the decay of heavy compound nuclei at various excitation energies.) Both microcanonical and sequential decay models predict ratios of. / (NC) somewhat smaller than observed experimentally, and they do not show intermittency [Ella 92, Barz 92]. These smaller ratios from theoretical predictions may not indicate an inconsistency with experiment since we do not select a sharp value of impact parameter and/ or excitation energy for the data in our event selection. In other words, the data contain some residual impact parameter averaging and represent an upper limit on the ratio of. / (NC). In Figure 7.4, second factorial moments calculated from Equation (7.1) are presented as a function of binning resolution. The solid points represent experimental data selected by a narrow cut on central collisions at E/A=110 MeV. A slightly positive slope is observed, but the moments are smaller than unity. As argued above, this positive slope may be of trivial origin. In order to corroborate this point, we include the results of percolation calculations performed for p=0.7. Standard calculations for which the total number of broken bonds is allowed to fluctuate (open squares) predict large factorial moments, but no intermittency. When the number of broken bonds is constrained to be constant (open diamonds), the factorial moments are strongly reduced in magnitude, and they exhibit a small increase as a function 165 I ‘ I I ‘ I g 9: ’ P=0.7 do 0.000 - - I}: " %B:E a a . B . as E] °’ 1- \S \ fl 4 —0.002 ~ _ _ fl— ’1 . . . . 1 L . . . 1 A . . . 1 ‘ I I ’ II I P=0.7 q f?) —0.016 _ energy conserved _ Ln. - e . .. 5 L W090 3 . -0.018 2' -. 1 . . . . 1 . . . . 1 A . . . 1 O 022 I . r . T r . . . . I r . r . I - ' _ “Ar + ""Au. E/A=110MeV - ‘ central collisions . 1 . . 1 L . . 1 . 1 1 -3 -2 -1 O —ln(A) Figure 7 .4 Scaled factorial moments as a function of binning resolution. Solid points show experimental results for central 36Ar+ 197Au collisions at E/A=110 MeV. Open points show results from percolation calculations using p=0.7. Open squares depict calculations in which the total number of broken bonds is allowed to fluctuate; open diamonds represent calculations in which the total number of bonds is kept fixed. 166 of binning resolution, similar to that observed experimentally. This increase appears to be of little significance. It has recently been shown [Elat 92] that rate-equation models used to describe sequential decay processes are incapable of producing intermittency in fragment mass distributions. The few statistical models that could produce intermittency signals [DeAn 92, Barz 92] had to resort to mixing fragmentation events of very different initial excitation energies (effectively averaging over impact parameter). Our results are consistent with the small fluctuations predicted from the statistical multifragmentation approach for initial conditions with a narrow range of excitation energies. Chapter 8 Summary In this work we have studied multifragment emission in 35Ar+197Au reactions at incident energies of E/A=50, 80 and 110 MeV. Of particular interest are central collisions, where the chance for equilibration of the system is the highest. To select such collisions we have made a detailed comparison of impact parameter scales constructed from the following global observables: the charged particle multiplicity NC, the total transverse energy of an event E,, the sum of the charge emitted at midrapidity Zy and the hydrogen multiplicity N, (the complement of 2M, the total charge bound in clusters of ZZZ). Each of these observables was evaluated in its ability to select events with suppressed projectile-like fragment emission and azimuthally anisotropic emission patterns. We found that scales constructed from NC, E, and Z’ provide comparable measures of impact parameter. Impact parameter scales based on E, provide slightly better resolution those based on NC or 2,, while the scale constructed from N1 is the worst. N 0 additional selectivity for central collisions was found by imposing additional cuts on the transverse momentum directivity D, a variable which has been applied successfully in symmetric systems at higher bombarding energies where flow effects are important. The mean values and variances of the multiplicity distributions of intermediate mass fragments were found to increase as a function of the total 167 168 charged-particle multiplicity; this may be interpreted as a rough measure of the internal energy of the fragment emitting system. An average multiplicity of 4 intermediate mass fragments is observed for the most central collisions at E/A=110 MeV. These large IMF multiplicities are consistent with predictions of a statistical model for evaporation from an expanding compound nucleus. The statistical decay model predictions are sensitive to the low-density nuclear equation of state. In our investigation of fragment admixtures predicted by a standard bond-percolation model, we found rather good agreement with the values measured for 36Ar + 197Au collisions. On the other hand, larger fragment admixtures measured for 129Xe + 197Au collisions were found to be inconsistent with the model (for compact geometries). We have also explored the multifragment breakup of different geometrical configurations (toroids and bubbles) with a bond-percolation model. Calculations for these finite systems predict enhanced fragment production for less compact decay configurations. For noncompact breakup geometries, the large fragment multiplicities observed for the 129Xe + 197Au reaction can be reconciled with predictions of the bond percolation model. Power law fits to the predicted mass distributions reveal a strong sensitivity of the extracted critical exponents to the geometry of the decaying system. Of course, the existence of such non-compact breakup configurations is not yet established. This question still remains to be answered. In this fragmentation study we also include an intermittency analysis based on factorial moments that should permit the establishment of deviations from Poissonian fluctuations in the fragmentation process. We argue that an intermittency signal is meaningful only when it is observed in data selected by a narrow cut in impact parameter, and only when F,(A)>1. 169 These two conditions are never simultaneously met in our data. A surprisingly small value of about 0% /(NC) = 0.3 was measured in central collisions, independent of beam energy. The small widths of the charged particle distributions may be caused by the constraints imposed by energy conservation. 170 Appendix A Physics tape format The information is stored in buffers 4096 words long (integer‘2). The first 16 words are the header. The first word past the header is the multiplicity of the event. This number is negative. The second word is the ringmask (one bit is set in this word for every ring that is hit). The third word is the RF word. The fourth is the number of intermediate mass fragments in the event. Knowing the event multiplicity (number of detectors that fired), we can calculate the length of the event. length = (multiplicity‘S) + 4 (AJ) To get to the next event you need to increment the pointer in the buffer by this length, i.e. ‘ next event = last event + length. (A2) The information of the current event is stored and accessed as follows: pointer = pointer + 4 ! move past the multiplicity, ringmask, RF, ! and NM words do ii = pointer, next event, 5 detector number = buffer(ii) ! from 1-188 slow = buffer(ii+1) ! slow channel number 1-2048 "real Z": buffer(ii+2) ! 2’25 from the PID maps PID number = buffer(ii+3) ! two byte word: first byte=ISO, for ! isotope resolution ! second byte=IZ of the particle energy = buffer(ii+4) ! in 100 keV units end do 171 Particle identification is accomplished through the two words 150 and 12 as shown in the following table: Table A.1 Physics tape format for particle identification. firticle type Hydrogen punch through Proton Deuteron Triton 3He 4He 6He or Be decay to 20: He punch through Li punch through 2a punch through Unidentified particles Fragments that punch through fast plastic Fission fragment Fragments that stopped in the plastic, above the LCP reject line, in the prompt branch of fast- time spectrum Particles that stopped in the plastic, below the LCP 0 11 reject line, in the prompt branch of fast-time NgNO moan NNNHHHHE H N spectrum No isotope resolution 1,2 5 Hit detector (ring 1, position 5) 0 0 The following detectors did not work properly during the experiment: fietector fiProblem fifesult 5 PMT pulsed Hit detector 16 double image no isotope resolution 48 discriminator double fired no Z greater than 2 70 double image no Z greater than 2 169 very small signal no Z greater than 2 None of the above detectors have energies. 172 Appendix B Temperature estimate from percolation theory In the standard percolation model the bond breaking probability p is linearly related to the excitation energy per nucleon E‘ of the compound system such that E‘ E' p z = Ea Emir where E, is nuclear matter binding energy per nucleon (16 MeV), E,,“, is the (B.1) energy required to break one bond and z is the number of nearest neighbors on the lattice (2:6). (By placing the nucleons on a lattice, the volume and surface effects in the binding energy are well approximated [Baue 88].) The total excitation energy of the system is simply E5... = pEaAm- (3.2) EL, in the Fermi gas model is 5;, = ar2 (13.3) where a is the level density parameter (a= A,”/ 8.5 MeV) and T is the temperature. The expression for the temperature is then E A T = .”+"’ =117Mev43. (13.4) For the near-critical bond-breaking probability p=0.7, T=9.8 MeV. [Alar 92] [Ardo 90] [Awes 81] [Barz 86] [Barz 92] [Baue 85] [Baue 86] 173 List of References ].P. Alard, Z. Basrak, N. Bastid, I.M. Belayev, M. Bini, Th. Blaich, R. Bock, A. Buta, R. Caplar, C. Cerruti, N. Cindro, J.P. Coffin, M. Crouau, P. Dupieux, I. Bro, Z.G. Fan, P. Fintz, Z. Fodor, R. Freifelder, L. Fraysse, S. Frolov, A. Gobbi, Y. Grigorian, G. Guillaume, N. Gerrmann, K.D. Hildenbrand, S. Holbling, O. Houari, S.C. Jeong, M. Iorio, F. Jundt, J. Kecskemeti, P. Koncz, Y. Korchagin, R. Kotte, M. Kramer, C. Kuhn, I. Legrand, A. Lebedev, C. Maguire, V. Manko, T. Matulewicz, G. Mgebrishvili, I. Mfisner, D. Moisa, G. Montarou, P. Morel, W. Neubert, A. Olmi, G. Pasquali, D. Pelte, M. Petrovici, G. Poggi, F. Rami, W. Reisdorf, A. Sadchikov, D. Schiill, Z. Seres, B. Sikora, V. Simion, S. Smolyankin, U. Sodan, N. Taccetti, K. Teh, R. Tezkratt, M. Trzaska, M.A. Vasiliev, P. Wagner, J.P. Wessels, T. Wienold, Z. Wilhelmi, D. Wohlfarth, and A.V. Zhilin, Phys. Rev. Lett. 69, 889 (1992). D. Ardouin, Z. Basrak, P. Schuck, A. Péghaire, F. Saint-Laurent, H. Delagrange, H. Doubre, C. Grégoire, A. Kyanowski, W. Mittig, J. Péter, Y.P. Viyogi, J. Québert, C.K. Gelbke, W.G. Lynch, M. Maier, J. Pochodzalla, G. Bizard, F. Lefebvres, B. Tamain, B. Remaud, and F. Sébille, N ucl. Phys. A514, 564 (1990). TC. Awes, G. Poggi, C.K. Gelbke, B.B. Back, B.G. Glagola, and V.E. Viola, Jr., Phys. Rev. C24, 89 (1981). H.W. Barz et al., N ucl. Phys. A448, 753 (1986). H.W. Barz, ].P. Bondorf, A.S. Botwina, R. Donangelo, LN. Mishustin, H. Schulz and K. Sneppen, Nucl. Phys. A545, 213C (1992). W. Bauer, D.R. Dean, U. Mosel, and U. Post, Phys. Lett. B 150, 53 (1985). W. Bauer, U. Post, D.R. Dean and U. Mosel, N ucl. Phys. A452, 699 (1986). [Baue 87] [Baue 88] [Baue 92] [Bert 83] [Bial 86] [Bial 88] [Birk 64] [Biro 86] [Blai 80] [Blum 91] [Boal 88] [Boal 90] 174 W. Bauer et al., Phys. Rev. Lett. 58, 863 (1987). W. Bauer, Phys. Rev. C 38, 1927 (1988). W. Bauer, G.F. Bertsch and H. Schulz, Phys. Rev. Lett. 69, 1888 (1992). G. Bertsch and PJ. Siemens, Phys. Lett. 1263, 9 (1983). A. Bialas and R Peschanski, N ucl. Phys. B273, 703 (1986). A. Bialas and R. Peschanski, N ucl. Phys. 3308 (1988) 857. LB. Birks, The Theory and Practice of Scintillation Counting (Pergamon Press, New York, 1964) p.2 T.S. Biro, J. Knoll and R. Richert, Nucl. Phys. A459, 692 (1986). ].P. Blaizot, Phys. Reports 64, 171 (1980). Y. Blumenfeld et al., Phys. Rev. Lett. 66, 576 (1991). D.H. Boal and IN. Glosli, Phys. Rev. C37, 91 (1988). D.H. Boal and IN. Glosli, Phys. Rev. C42, R502 (1990). [Bond 85a] LP. Bondorf et al., Nucl. Phys. A443, 321 (1985). [Bond 85b] LP. Bondorf et al., Nucl. Phys. A444, 460 (1985). [Boug 88] [Boug 89] R. Bougault, et al., N ucl. Phys. A488, 255C (1988). R. Bougault et al., Phys. Lett. 8232, 291 (1989). [Bowm 91] DR. Bowman, G.F. Peaslee, R.T. de Souza, N. Carlin, C.K. Gelbke, W.G. Gong, Y.D. Kim, M.A. Lisa, W.G. Lynch, L. Phair, M.B. Tsang, C. Williams, N. Colonna, K. Hanold, M.A. McMahan, GJ. Wozniak, L.G. Moretto, and W.A. Friedman, Phys. Rev. Lett. 67, 1527 (1991). [Bowm 92] DR. Bowman, C. M. Mader, G. F. Peaslee, W. Bauer, N. Carlin, R.T. de Souza, CK. Gelbke,W..G Gong,Y.D. Kim, M.A. Lisa, W.G. Lynch, L. Phair, M. B. Tsang, C. Williams, N. Colonna, K. Hanold, M.A. McMahan, GJ. Wozniak, L.G. Moretto, and W.A. Friedman, Phys. Rev. C46, 1834 (1992). [Camp 86] X. Campi, J. Phys. A19, L917 (1986). [Camp 88] X. Campi, Phys. Lett. 208B, 351 (1988). [Camp 92] [Cava 90] [Cerr 88] [Char 88] [Chen 87] [Chit 86] [C011 75] [Cser 86] [Curt 83] [DeAn 92] [deSo 90] [deSo 91] [Dins 62] [Doss 87] [Elat 92] [Elma 91] 175 X. Campi and H. Krivine, Z. Phys. A344, 81 (1992). C. Cavata, M. Demoulins, I. Gosset, M.C. Lemaire, D. L'Héte, J. Poitou, and O. Valette, Phys. Rev. C42, 1760 (1990). C. Cerruti et al., N ucl. Phys. A476, 74 (1988). RJ. Charity et al., Nucl. Phys. A483, 371 (1988). Z Chen, C.K. Gelbke, I. Pochodzalla, C.B. Chitwood, DJ. Fields, W.G. Gong, W.G. Lynch, and MB. Tsang, Nucl. Phys. A473, 564 (1987). CB. Chitwood, DJ. Fields, C.K. Gelbke, D.R. Klesch, W.G. Lynch, M.B. Tsang, T.C. Awes, R.L. Ferguson, F.E. Obenshain, F. Plasil, R.L. Robinson and GR. Young, Phys. Rev. C34, 858 (1986). ].C. Collins and MJ. Perry, Phys. Rev. Lett. 34, 1353 (1975). LP. Csernai, and 1.1. Kapusta, Phys. Rep. 131, 233 (1986), and references therein. M.W. Curtin, H. Toki, and BK. Scott, Phys. Lett. 3123, 289 (1983). AR. DeAngelis, D.H.E. Gross and R. Heck, N ucl. Phys. A537, 606 (1992). RT. de Souza et al., Nucl. Instr. and Meth. A295, 109 (1990). RT. de Souza, L. Phair, D.R. Bowman, N. Carlin, C.K. Gelbke, W.G. Gong, Y.D. Kim, M.A. Lisa, W.G. Lynch, G.F. Peaslee, M.B. Tsang, H.M. Xu, F. Zhu, and W.A. Friedman, Phys. Lett. 8268, 6 (1991). A. Dinsdale and F. Moore, ”Viscosity and its Measurement”, Chapman-Hill, London, 1962. K.G.R. Doss et al., Phys. Rev. Lett. 59, 2720 (1987). B. E1atari,]. Richert and P. Wagner, Phys. Rev. Lett. 69, 45 (1992). A. Elmaani, N .N . Ajitanand, I.M. Alexander, R. Lacey, S. Kox, E. Liatard, F. Merchez, T. Motobayashi, B. N oren, C. Perrin, D. Rebreyend, T.U. Chan, G. Anges, and S. Groult, Phys. Rev. C43, R2474 (1991). 176 [Fiel 86] DJ. Fields, W.G. Lynch, T.K. N ayak, M.B. Tsang, C.B. Chitwood, C.K. Gelbke, R. Morse, J. Wilczynski, T.C. Awes, R.L. Ferguson, F. Plasil, F.E. Obenshain, and GR. Young, Phys. Rev. C34, 536 (1986). [Fox 93] D. Fox, R.T. de Souza, L. Phair, D.R. Bowman, N. Carlin, C.K. Gelbke, W.G. Gong, Y.D. Kim, M.A. Lisa, W.G. Lynch, G.F. Peaslee, M.B. Tsang and F. Zhu, Phys. Rev. C47, R421 (1993). [Frie 83a] W.A. Friedman and W.G. Lynch, Phys. Rev. C28, 16 (1983). [Frie 83b] W.A. Friedman and W.G. Lynch, Phys. Rev. C28, 950 (1983). [Frie 88] W.A. Friedman, Phys. Rev. Lett. 60, 2125 (1988). [Frie 89] W.A. Friedman, Phys. Rev. C40, 2055 (1989). [Frie 90] W.A. Friedman, Phys. Rev. C42, 667 (1990). [Gali 85] I. Galin, N ucl. Phys. A447, 519C (1985). [Gold 78] AS. Goldhaber and H.H. Heckman, Ann. Rev. Nucl. Part. Sci. 28, 161 (1978). [Gong 88] W.G. Gong, Y.D. Kim, G. Poggi, Z. Chen, C.K. Gelbke, W.G. Lynch, M.R. Maier, T. Murakami, M.B. Tsang, H.M. Xu and K. Kwiatkowski, N ucl. Instr. and Meth. A268, 190 (988). [Gong 90] W.G. Gong, N. Carlin, C.K. Gelbke and R. Dayton, Nucl. Instr. and Meth. A287, 639 (1990). [Gros 90] D.H.E. Gross, Rep. Prog. Phys. 53, 605 (1990), and refs. therein. [Hirs 84] AS. Hirsch, A. Bujak, LE. Finn, LJ. Gutay, R.W. Minich, N .T. Porile, R.P. Scharenberg, B.C. Stringfellow, and F. Turkot, Phys. Rev. C29, 508 (1984). [Hube 91] J. Hubele, P. Kreutz, ].C. Adloff, M. Begemann-Blaich, P. Bouissou, G. Imme, I. Iori, GJ. Kunde, S. Leray, V. Lindenstruth, Z. Liu, U. Lynen, RJ. Meijer, U. Milkau, A. Moroni, W.F.]. Miiller, C. N g0, C.A. Ogilvie, I. Pochodzalla, G. Raciti, G. Rudolf, H. Sann, A. Schiittauf, W. Seidel, L. Stuttge, W. Trautrnann, and A. Tucholski, Z. Phys. A340, 263 (1991). ana 83] HR. Jaqaman, A.Z. Mekjian, and L. Zamick, Phys. Rev. C27, 2782 (1983). ana 84] HR. Jaqaman, A.Z. Mekjian, and L. Zamick, Phys. Rev. C29, 2067 (1984). 177 I'Jaqa 90] HR. Jaqaman, G. Papp, and D.H.E. Gross, Nucl. Phys. A514, 327 (1990). [Keho 92] W.L. Kehoe, A.C. Mignerey, A. Moroni, I. Iori, G.F. Peaslee, N. Colonna, K. Hanold, DR Bowman, L.G. Moretto, M.A. McMahan, ].T. Walton, and GJ. Wozniak, Nucl. Instr. Meth. A 311, 258 (1992). [Kim 89] Y.D. Kim et al., Phys. Rev. Lett. 63, 494 (1989). [Kim 92] Y.D. Kim, R.T. de Souza, D.R. Bowman, N. Carlin, C.K. Gelbke, W.G. Gong, W.G. Lynch, L. Phair, M.B. Tsang, and F. Zhu, Phys. Rev. C45, 338 (1992). [Kuti 81] J. Kuti, I. Polbnyi and K. Szlachanyi, Phys. Lett. B98, 199 (1981). [Lace 93] R. Lacey, A. Elmaani, J. Lauret, T. Li, W. Bauer, D. Craig, M. Cronqvist, E. Gualtieri, S. Hannuschke, T. Reposeur, A. Vander Molen, G.D. Westfall, W.K. Wilson, ].S. Winfield, I. Yee, S. Yennello, A. Nadasen, R.S. Tickle, and E. Norbeck, Phys. Rev. Lett. 70, 1224 (1993). [Li 93] T. Li et al., Phys. Rev. Lett. 70, 1924 (1993). [Lift 80] U. Littrnark and IF. Ziegler, Handbook of Range Distributions for Energetic Ions in All Elements, vol. 6 of The Stopping and Ranges of Ions in Matter, ed. ].F. Ziegler (Pergamon, New York, 1980). [Lync 82] W.G. Lynch, L.W. Richardson, M.B. Tsang, R.E. Ellis, C.K. Gelbke and RE. Warner, Phys. Lett. 1088, 274 (1982). [Lync 87] W.G. Lynch, Ann. Rev. N ucl. Part. Sci. 37, 493 (1987), and references therein. [Mahi 88] M. Mahi et al., Phys. Rev. Lett. 60, 1936 (1988). [Mana 62] W.W. Managan, IRE Trans. N ucl. Sci. NS-9 (3) (1962) 8. [McLe 81] L.D. McLerran and B. Svetitsky, Phys. Lett. 898, 195 (1981). [McMa 86] M.A. McMahan, GJ. Wozniak, C.M. Lyneis, D.R. Bowman, RJ. Charity, Z.H. Liu, L.G. Moretto, A.C. Mignerey and MN. Namboodiri, Nucl. Instr. Meth. A 253, 1 (1986). [Meye 78] D. Meyerhofer, J. Appl. Phys. 49, 3993 (1978). [Moli 85] 1.]. Molitoris and H. Stocker, Phys. Lett. 1628 (1985) 47. [More 75] LG. Moretto, Nucl. Phys. A247, 211 (1975). [More 92] [Morj 88] [Morl 79] [N aya 92] [N30 90] [Norb 87] [Ogil 3931 [Ogil 89b] [Ogil 90] [Ogil 91] [Pana 84] [Peil 89] [Phai 92a] [Phai 92b] [Plos 90] [Poch 87] 178 L.G. Moretto et al., Phys. Rev. Lett. 69, 1884 (1992). M. Morjean, et al., Phys. Lett. 8203, 215 (1988). RD. Morley and MB. Kislinger, Phys. Rep. 51, 63 (1979) and references therein. T.K. N ayak, T. Murakami, W.G. Lynch, K. Swartz, DJ. Fields, C.K. Gelbke, Y.D. Kim, J. Pochodzalla, M.B. Tsang, F. Zhu, and K. Kwiatkowski, Phys. Rev. C65, 132 (1992), and refs. therein. H. N g0, C. N g6, F.Z. Ighezou, J. Debois, S. Leray, and Y.-M. Zheng, Z. Phys. A337, 81 (1990). E. N orbeck, T.P. Dubbs and L.G. Sobotka, Nucl. Instr. and Meth. A262, 564 (1987). CA. Ogilvie, D. Cebra, 1. Clayton, S. Howden, J. Karn, A. Vander Molen, G.D. Westfall, W.K. Wilson, and ].S. Winfield, Phys. Rev. C40, 654 (1989). CA. Ogilvie, D.A. Cebra, J. Clayton, P. Danielewicz, S. Howden, J. Karn, A. Nadasen, A. Vander Molen, G.D. Westfall, W.K. Wilson, and LS. Winfield, Phys. Rev. C40, 2592 (1989). CA. Ogilvie, W. Bauer, D.A. Cebra, J. Clayton, S. Howden, I. Karn, A. Nadasen, A. Vander Molen, G.D. Westfall, W.K. Wilson, and ].S. Winfield, Phys. Rev. C42 (1990) R10. C.A. Ogilvie et al., Phys. Rev. Lett. 67, 1214 (1991 ). A.D. Panagiotou et al., Phys. Rev. Lett. 52, 496 (1984). G. Peilert et al., Phys. Rev. C39, 1402 (1989). L. Phair, D.R. Bowman, C.K. Gelbke, W.G. Gong, Y.D. Kim, M.A. Lisa, W.G. Lynch G.F. Peaslee, R.T. de Souza, M.B. Tsang and F. Zhu, Nucl. Phys. A458, 489 (1992). L. Phair et al., Phys. Lett. B 285, 10 (1992). M. Ploszajczak and A. Tucholski, Phys. Rev. Lett. 65, 1539 (1990). J. Pochodzalla, C.K. Gelbke, W.G. Lynch, M. Maier, D. Ardouin, H. Delagrange, H. Doubre, C. Gregoire, A. Kyanowski, W. Mittig, A. Péghaire, J. Péter, F. Saint-Laurent, B. Zwieglinski, G. Bizard, F. Lefebvres, B. Tamain, and]. Québert, Y.P. Viyogi, W.A. Friedman, and DH. Boal, Phys. Rev. C35 (1987) 1695. [Rift 88] [Rose 84] [Saye 76] [Schl 87] [Shur 80] [Siem 83] [Snep 88] [Stoc 86] [$th 86] [Sull 90] [Troc 89] [Tsan 84a] [Tsan 84b] [Tsan 86] [Tsan 88] [Tsan 89a] 179 HO. Ritter, Nucl. Phys. A488, 651C (1988). ].H. Rose, J.P. Vary, and JR. Smith, Phys. Rev. Lett. 53, 344 (1984). G. Sayer, H., Chandra, and U. Mosel, U. Nucl. Phys. A24, 221 (1976). TJ. Schlagel, and V.R. Pandharipande, Phys. Rev. C36, 162 (1987). E.Shuryak, Phys. Rep. 61, 71 (1980) and references therein. PJ. Siemens, Nature 305, 410 (1983). K Sneppen and L. Vinet, N ucl. Phys. A480, 342 (1988). R. Stock, Phys. Reports. 135, (1986) 259. H. Stocker and W. Greiner, Phys. Reports 137, 277 (1986). JP. Sullivan, J. Péter, D. Cussol, G. Bizard, P. Brou, M. Louvel, J.P. Patry, R. Regimbart, ].C. Steckmeyer, B. Tamain, E. Crema, H. Doubre, K. Hagel, G.M. Jin, A. Péghaire, F. Saint-Laurent, Y. Cassagnou, R. Legrain, C. Lebrun, E. Rosato, R. McGrath, S.C. Jeong, S.M. Lee, Y. N agashima, T. N akagawa, M. Ogihara, J, Kasagi, and T. Motobayashi, Phys. Lett. 8249 (1990) 8. R. Trockel et al., Phys. Rev. C39, 729 (1989). MB. Tsang, W.G. Lynch, C.B. Chitwood, DJ. Fields, D.R. Klesch, C.K. Gelbke, G.R Young, T.C. Awes, R.L. Ferguson, F.E. Obenshain, F. Plasil and R.L. Robinson, Phys. Lett. B 148, 265 (1984). MB. Tsang, C.B. Chitwood, DJ. Fields, C.K. Gelbke, D.R. Klesch, W.G. Lynch, K. Kwiatkowski and V.E. Viola, Jr., Phys. Rev. Lett. 52, 1967 (1984). MB. Tsang, R.M. Ronningen, G. Bertsch, Z. Chen, C.B. Chitwood, DJ. Fields, C.K. Gelbke, W.G. Lynch, T. N ayak, J. Pochodzalla, T. Shea, and W. Trautrnann, Phys. Rev. Lett. 57, 559 (1986). MB. Tsang, W.G. Lynch, R.M. Ronningen, Z. Chen, C.K. Gelbke, T. Nayak, J. Pochodzalla, F. Zhu, M. Tohyama, W. Trautmann, and W. Diinnweber, Phys. Rev. Lett. 60, 1479 (1988). MB. Tsang, G.F. Bertsch, W.G. Lynch, and M. Tohyama, Phys. Rev. C40, 1685 (1989). [Tsan 89b] [Tsan 90] [Tsan 91] [Wadd 85] [Wang 91] [Wils 90] 180 MB. Tsang, Y.D. Kim, N. Carlin, Z. Chen, R. Fox, C.K. Gelbke, W.G. Gong, W.G. Lynch, T. Murakami, T.K. Nayak, R.M. Ronningen, H.M. Xu, F. Zhu, L. Sobotka, D. Stracener, D.G. Sarantites, Z. Majka, V. Abenante, and H. Griffin, Phys. Lett. 8220, 492 (1989). MB. Tsang, Y.D. Kim, N. Carlin, Z. Chen, C.K. Gelbke, W.G. Gong, W.G. Lynch, T. Murakami, T. Nayak, R.M. Ronningen, H.M. Xu, F. Zhu, L.G. Sobotka, D.W. Stracener, D.G. Sarantites, Z. Majka, and V. Abenante, Phys. Rev. C42, R15 (1990). MB. Tsang, R.T. de Souza, Y.D. Kim, D.R. Bowman, N. Carlin, C.K. Gelbke, W.G. Gong, W.G. Lynch, L. Phair, and F. Zhu, Phys. Rev. C44, 2065 (1991). CJ. Waddington and PS. Freier, Phys. Rev. C 31, 888 (1985). S. Wang, YZ Jiang, Y.M. Liu, D. Keane, D. Beavis, S.Y. Chu, S.Y. Fung, M. Vient, C. Hartnack and H. Stacker Phys. Rev. C44, 1091 (1991). W.K. Wilson, W. Benenson, D.A. Cebra, J. Clayton, S. Howden, I. Karn, T. Li, C.A. Ogilvie, A. Vander Molen, G.D. Westfall, LS. Winfield, and A. N adasen, Phys. Rev. C41, R1881 (1990). "llllllllllllllllf