MEASURING NEUTRONS IN HEAVY ION COLLISIONS By Kuan Zhu A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of Physics — Doctor of Philosophy 2020 ABSTRACT MEASURING NEUTRONS IN HEAVY ION COLLISIONS By Kuan Zhu The behavior of the symmetry energy above nuclear saturation density plays a significant role in the properties of neutron stars, the structure of heavy nuclei, and the dynamics of nuclear reactions. To improve constraints on the symmetry energy term in the Nuclear Equa- tion of State (EOS), neutrons and charged-particles were measured with beams of 40,48Ca at 56,140 MeV/u on targets of 58,64Ni and 112,124Sn. The updated High Resolution Array (HiRA10) was used for charged-particle detection. A charged-particle veto wall was designed and constructed to eliminate charged-particle contamination in neutron spectrum measured with the Large Area Neutron Arrays (LANA). Several innovative methods were developed to improve LANA’s calibration and neutron/gamma discrimination procedures. Neutron light output are compared to the simulation results from SCINFUL-QMD to test the validity of neutron measurements and calculate the neutron detection efficiency. Various analysis pro- cedures needed for neutron measurements are demonstrated and a preview of the data to come is provided. ACKNOWLEDGMENTS First and foremost I could not have accomplished any of this without the guidance of my advisor, Professor ManYee Betty Tsang. She was always available for idea and discussion. I would like to further acknowledge my graduate guidance committee members: William Lynch, Pawel Danielewicz, Mark Dykman and Morten Hjorth-Jensen. Thank you for your time and thoughtful advice. Thanks to all previous and current HiRA lab members: Kyle Brown, Giordano Cerizza, Daniele Dell’Aquila, Genie Jhang, Pierre Morfouace, Chenyang Niu, Cl´ementine Santamaria, Rensheng Wang, Adam Anthony, Jon Barney, Justin Estee, Sean Sweany, Tommy Tsang, Chi-En Fanurs Teh and Joseph Wieske. It has always been joyful to keep learning and work with you. I would also like to thank the many talented undergraduates I worked with: Corinne Anderson, Hananiel Setiawan, Suhas Kodali, Mira Ghazali, Shuqi Han and Jiashen Ron Tang. Special thanks to Zbigniew Chajecki group in Western Michigan University. The ex- periments and the construction of the Charged-Particle Veto Wall can not be accomplished without the strong collaboration with his team. Last and certainly not least, thank you to all my fellow graduate students for your help and support. Maxwell Cao, Mengzhi Chen, Brandon Elman, Rongzheng He, Kuan-Yu Lin, Hao Lin, Dan Liu, Han Liu, Tong Li, Didi Luo, Xingze Mao, Samuel Marinelli, Pierre Nzabahimana, Omokuyani Udiani, Qi Wang, ChunYan Jonathan Wong, Huyan Xueying and Faran Zhou, thank you for your encouragement and friendship. This work was in part funded by National Science Foundation (NSF) and Michigan State University (MSU). iii TABLE OF CONTENTS LIST OF TABLES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . vi LIST OF FIGURES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . vii Chapter 1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1.1 The Atomic Nucleus . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1.2 Nuclear Equation of State . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1.3 Probing Effective Mass Splitting in Heavy Ion Collision . . . . . . . . . . . . 1.4 Neutron Detection . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1.5 Improvements from Previous Experiment . . . . . . . . . . . . . . . . . . . . 1.6 Organization of Dissertation . . . . . . . . . . . . . . . . . . . . . . . . . . . Chapter 2 Experimental Details . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.1 Microball 2.2 HiRA10 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.3 The Large Area Neutron Array (LANA) . . . . . . . . . . . . . . . . . . . . 2.4 Forward Array . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.5 Charged-Particle Veto Wall . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.5.1 Design and Prototype . . . . . . . . . . . . . . . . . . . . . . . . . . 2.5.2 Construction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Chapter 3 Data Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.1 Microball Impact Parameter Determination . . . . . . . . . . . . . . . . . . 3.2 LANA Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.2.1 LANA Calibration using cosmic rays . . . . . . . . . . . . . . . . . . 3.2.1.1 Impact position calibration . . . . . . . . . . . . . . . . . . 3.2.1.2 Time resolution and time offset calibration . . . . . . . . . . 3.2.1.3 Light attenuation length and photomultiplier gain-matching . . . . . . Light output calibration and position corrections 3.2.1.4 3.2.2 Time calibration with the Forward Array . . . . . . . . . . . . . . . . 3.2.3 Geometry Efficiency . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.2.4 Background Scattering Efficiency . . . . . . . . . . . . . . . . . . . . 3.3 Charged-Particle Veto Wall Pulse Height Calibration . . . . . . . . . . . . . 1 1 3 10 12 16 21 22 27 29 33 43 48 49 50 60 60 62 62 63 70 71 75 80 86 90 96 Chapter 4 Performance of LANA . . . . . . . . . . . . . . . . . . . . . . . . . . 104 4.1 Neutron/γ Pulse Shape Discrimination . . . . . . . . . . . . . . . . . . . . . 104 4.1.1 Traditional PSD & VPSD comparison . . . . . . . . . . . . . . . . . 106 4.1.2 PSD efficiency . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 116 4.2 Neutron Light Output Comparison with SCINFUL-QMD . . . . . . . . . . . 120 4.3 Preliminary Neutron Spectrum . . . . . . . . . . . . . . . . . . . . . . . . . 129 iv Chapter 5 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 133 BIBLIOGRAPHY . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 136 v LIST OF TABLES Table 1.1: Neutron designation table for different energy ranges. This table is adopted . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . from Ref. [1] Table 3.1: bmax table for all 16 beam-target-energy combinations in our experiment. The unit is barn, 100 fm2. . . . . . . . . . . . . . . . . . . . . . . . . . . . Table 3.2: The ∆X and v of each back LANA (NWA) bar. . . . . . . . . . . . . . . Table 3.3: The ∆X and v of each front LANA (NWB) bar. . . . . . . . . . . . . . . Table 3.4: The light attenuation length λ of each back LANA (NWA) bar. . . . . . . Table 3.5: The light attenuation length λ of each front LANA (NWB) bar. . . . . . . 14 62 67 67 73 73 vi LIST OF FIGURES Figure 1.1: The schematic showing the“plum pudding model”proposed by J.J. Thoma- son (left) and the model proposed by Rutherford (right) after his discovery experiment that showed back scattering of alpha particles. Alpha particles penetrating atoms are represented by horizontal black arrow. Blue circles represents electrons and red area shows the region with positive charge. The sizes of electrons and nuclei are not drawn in scale. . . . . . . . . . . Figure 1.2: Density dependence of the symmetry energy from the Skyrme interactions used in Ref. [2].The shaded region is obtained from HIC experiments. Picture is adopted from Ref. [3]. . . . . . . . . . . . . . . . . . . . . . . . Figure 1.3: (top panels) Y(n)/Y(p) ratios as a function of kinetic energy for 124Sn+124Sn at b=2fm. (bottom panels) DR(n/p) ratios as a function of kinetic energy. From left to right, the beam energies are 100 MeV, 150 MeV, 200 MeV, . . . . . . . . . . 300 MeV per nucleon. Figure is adopted from Ref. [4]. Figure 1.4: Neutron reaction cross sections adopted in SCINFUL-QMD model. Pic- . . . . . . . . . . . . . . . . . . . . . . . . ture is adopted from Ref. [5]. Figure 1.5: neutron/proton double ratio for 124Sn +124 Sn/112Sn +112 Sn reaction systems at 120 MeV/u for central collisions. The symbols are experimental data. Red line shows calculation using SLy4 which has m∗ p while blue line comes from calculation using SkM∗ which has m∗ p. The picture is adopted from Ref. [6]. Open circles are the original version and the close circles are data of the most updated version. The original data points did not take into account the efficiency corrections for charged- particles [7, 8]. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . n < m∗ n > m∗ Figure 1.6: Neutron Wall light vs TOF, where TOF is the time of flight for the ob- served particle to travel to the neutron walls. The prompt γ ray peak, charged-particle punch-through, and charged particle stopping PID lines are marked. Picture is adopted from Ref. [9]. . . . . . . . . . . . . . . . Figure 2.1: Experimental setup overview. Microball, Forward Array and HiRA10 are placed inside a vacuum chamber. The target is inside the Microball at the center of the chamber. The lid of the vacuum chamber is removed for this picture. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Figure 2.2: A schematic showing the experiment setup. It is a not-to-scale version of . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Figure 2.25. 2 7 13 15 17 20 23 24 vii Figure 2.3: Schematic diagram showing a vertical section of the Microball. The CsI(T1) crystals are shown in grey. The blue trapezoids behind crystals are light guides. The beam enters from the left. The numbers 1,2,3,to 9 are the ring numbers. The polar angular coverage for each ring are also shown together with the number of crystals contained in each ring. This picture is adopted from Ref. [10]. . . . . . . . . . . . . . . . . . . . . . . . . . . Figure 2.4: Ten microball detectors removed to allow charged particles emitted in reac- tions to be detected by HiRA10 without materials without going through blocking materials. The beam enters from left. . . . . . . . . . . . . . . . Figure 2.5: The target ladder used in the experiment. . . . . . . . . . . . . . . . . . Figure 2.6: Angular coverage for Forward Array, Microball, HiRA10 and LANA in lab Φ-Θ coordinates. The dashed blue rectangle represents the coverage of the Forward Array, green regions represent Microball, red regions represent twelve HiRA10 telescopes arranged in 4 towers, and the purple region represents LANA coverage. Ten crystals in total are removed on Microball ring #3, #4 and #5 leaving a direct path for the fragments to be detected by HiRA10. There is no such accommodation for LANA. Microball ring #6 is removed for target structure. The Microball covers part of the forward array from Θ = 13° to Θ = 28°. In the adopted co-ordinate from 0 and 360 deg, the coverage of LANA is split at 0 and 360 degree (top panel). The bottom panel shows LANA in continuous angular coverage by plotting the φ range from -60° to 300°. . . . . . . . . . . . . . . . . . Figure 2.7: (Left panel) A mechanical drawing of a HiRA10 telescope with its alu- minum container and associated electronics. The DSSSD (not shown) is placed in front of CsI crystals. (Right panel) A schematic drawing of a HiRA10 telescope showing the front and back sides of the DSSSD placed in front of the 4 CsI crystals. . . . . . . . . . . . . . . . . . . . . . . . . Figure 2.8: (Left panel) 12 HiRA10 telescopes arranged in 3 towers used in the present experiments. The mounts are designed in approximately in an arc so that the center of each telescope is located at 35 cm from the target. (Right panel) Angular coverage of 12 telescopes in HiRA10 on a (θ, φ) plane measured by Romer arm technology [11]. Each point represents a DSSSD (1.95mm × 1.95mm) pixel. . . . . . . . . . . . . . . . . . . . . . . . . . . Figure 2.9: Particle identification of a HiRA10 telescope. The Y axis is the calibrated DSSSD energy loss (∆E) versus the ADC channels of the residual energy (E) in the CsI 2D plot. The top right inset shows an extended plot up to 120 MeV of energy loss in DSSSD. Red lines are PID lines from simulation for various isotopes. This figure is adopted from Ref. [12]. . . . . . . . . 25 26 28 30 32 34 35 viii Figure 2.10: Mechanical cutaway drawing of one wall of the Large Area Neutron Array. [13] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Figure 2.11: One of LANA walls is illuminated with ultra-violet light after the front . . . . . . . . . . . . . . . . . . . . . . . . . aluminum cover is removed. Figure 2.12: 3D mechanical overview drawing showing the position of LANA being limited by the vault wall. During the move, some floor panels had to be removed. These panels were installed back again after the moving of LANA was done. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Figure 2.13: Angular coverage of the front LANA (NWB) and back LANA (NWA) in . . . . . . . . . . . . the lab frame based on laser position measurement. Figure 2.14: Schematic of the electronics setup (for one PMT) used to process signals from the LANA when the walls are used in standalone mode. An ex- act reproduction of these circuits is used to process signals obtained with each of the 100 individual PMTs. “Fast” and “total” are two copies of the original anode signal but integrated respectively by using a fast gate (con- taining only the fast component of the signal) and total gate (containing the whole signal) for PSD analysis. PMT dynode signals are used for the timing. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Figure 2.15: The front and back views of the original Forward Array without PMT bases. The front side of Forward Array is covered by a thin layer of Sn/Pb foil to protect itself from δ electrons emitted in reactions. . . . . . Figure 2.16: The front and back views of the original Forward Array assembled with . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . its backing. Figure 2.17: The side and back views of the upgrade Forward Array placed in position . . . . . . . . . . . . . . . . . . . . . . . . . . . . . during experiments. Figure 2.18: 3D sketch showing how 3 Charged-Particle Veto Wall scintillator bars are . . . . . . assembled at the bottom end with T-Slotted aluminum frame. Figure 2.19: Professor Chajecki is studying the signal characteristic of Charged-Particle Veto Wall bar prototype (at the right of the figure) with his students at Western Michigan University. . . . . . . . . . . . . . . . . . . . . . . . . Figure 2.20: Light guides are being glued on PMTs at Western Michigan Universty. . 36 38 40 41 42 44 46 47 51 52 53 ix Figure 2.21: (top left panel) Professor Chajecki is testing the Charged-Particle Veto Wall frame. (top right panel) Undergraduate students from Western Michigan University and Michigan State University are moving Veto Wall bar. (bottom left panel) Veto Wall bars are being taken out from shipment packages. (bottom middle and right panel) NSCL and WMU researchers are installing the Charged-Particle Veto Wall in front of LANA. . . . . . Figure 2.22: Charged-particle Veto Wall construction in progress. 22 bars have been installed in place. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Figure 2.23: The Charged-Particle Veto Wall installed in front of the LANA. In the top panel, the red and yellow tape marks on the ground indicate the polar angle with respect to the beam direction in lab frame. . . . . . . . . . . 54 55 57 Figure 2.24: (Top panel) 3D sketch of the Charged-Particle Veto Wall. (Bottom panel) 3D sketch of the Charged-Particle Veto Wall and the LANAs nested together. 58 Figure 2.25: Top view of experimental setup schematic. The schematic reflects the sizes and distances to the target of the Charged-Particle Veto Wall (VW) and LANA (NWA and NWB) in scale. The VW+LANA assemble is moved to 39.37 ° . All distances are drawn to scale and determined from laser measurements. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Figure 3.1: Correlations between the induced impact parameter ˆb and Microball mul- tiplicity Nc for 16 reaction systems. On the upper right corners of two panels, the reaction systems are labeled as target+beam. The top panels are for Ni target reaction systems and the bottom panels are for Sn target reaction systems. The left panels have lower beam energy of 56 AMeV and the right panels have higher beam energy of 140 AMeV. . . . . . . . Figure 3.2: (top panel) Time difference, tleft-tright distribution for cosmic muons ob- tained with bar #8 of the LANA. A numerical differentiation of this spec- trum (bottom panel) is used to determine the position of right and left edges of the tleft-tright distribution consistently as indicated by the dashed lines, which correspond to − L 2 respectively. . . . . . . . . . . . . . 2 and L Figure 3.3: A schematic drawing of a cosmic muon track penetrating 10 consecutive bars in the wall. Red stars represent the hit positions. The reconstructed muon track (blue slanted line) is obtained by fitting the measured hit positions on the bars with a straight line. ∆X indicates the position deviation between the expected and the actual hit position. ∆di,ref is the theoretical distance between two hits on a certain bar and reference bar respectively. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59 61 65 66 x Figure 3.4: Position deviation distribution for bar #8. . . . . . . . . . . . . . . . . . 68 Figure 3.5: δi,ref distributions for each of the bar in the array before (top panel) and after (bottom panel) the peaks are aligned. Y-axis is the bar number while i,ref − ∆T th X-axis shows the corresponding ∆T exp i,ref distributions. Bar #12 is used as the reference bar; therefore δ12,ref =0. . . . . . . . . . . . . . . Figure 3.6: Time deviation distribution for bar #8. . . . . . . . . . . . . . . . . . . . Figure 3.7: Equation (3.8) as a function of the position X along the bar #8 for cosmic ray data. The red dashed line is a linear fit of the distribution. The intercept of the best fit line is compatible with (0,0), indicating a good left-right matching of the PMT gains. The slope allows to extract the attenuation length of the bar. . . . . . . . . . . . . . . . . . . . . . . . . 69 72 74 Figure 3.8: (Top panel) Uncalibrated left-right GM as a function of the position along bar #8. Vertical cosmic rays are selected by restricting data to impinging angles −10° ≤ θ ≤ 10° with respect to the axis perpendicular to the bar length. Black points are the position of the cosmic ray MPV deduced with a Landau fit of data for each position bin. The blue line is the result of a quadratic fit of the cosmic MPV position dependence. (Bottom panel) Same as top panel after light calibration and position dependency correction. 76 Figure 3.9: An example of a Landau fit of the GM distribution for the position bin . . . . . . . . . . . . . . . . . . . . . . . . . . . . . -20 cm ≤ X ≤ 0 cm. 78 Figure 3.10: Position corrected light output spectrum obtained with an AmBe source for bar #8. The Compton edge is fitted with a Fermi function and the deduced position associated to the 4.44 MeV transition in 12C is indicated by an arrow and a dashed line. The theoretical value of the Compton edge is 4.2 MeV. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Figure 3.11: The cosmic-ray energy-light calibration curve (red line) obtained by using the 11.96 MeVee constraint of the cosmic muons and the zero-offset (blue solid circle and blue square, respectively). The open star corresponds to the Compton edge discussed in Figure 3.10. The open blue circles are two additional light-energy calibration points obtained by select cosmic muons that punch through the detector at 44.4 ± 5 and 56.3 ± 5 degrees with . . . . . . . . . . . . respect to the axis perpendicular to the bar length. 79 81 xi Figure 3.12: A time of flight spectrum from NWB with log (top panel) and linear (bottom panel) scale. The narrow peak at the left of the spectrum comes from prompt gamma rays, which arrives the LANA all together and earlier than any other particles. This plot is from 48Ca+124Sn at E/A = 56M eV reaction. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Figure 3.13: 2D plot of FA segment QDC raw channel versus the timing subtraction between NWB and a FA segment. Top panel is before walk in effect correction and bottom panel is after correction. The red curve in top panel is a fitted function in form of Eq. 3.11. . . . . . . . . . . . . . . . . Figure 3.14: Normalized ToF spectrum from NWB with FA Time Min (top panel) and . . . . . . . . . . . . . . . . . . . . . . . FA Time Mean (bottom panel). Figure 3.15: Hit distribution pattern in Φ-Θ plane for NWB using Geant4 simulation. In simulation, the LANA is placed in the same position as measured by laser measurements [14] in experiments. Then particles are simulated to emit from target uniformly in Φ-Θ plane. We can see that for different Θ, LANA’s coverage in azimuthal direction is different resulting in a geomet- ric efficiency difference. The red window is 0.2 ° wide binned at Θ = 30°. Figure 3.16: LANA fractional azimuthal coverage fΦ for front (NWB, top panel) and . . . . . . . . . . . . . . . . . . . . . . back (NWA, bottom panel) wall. Figure 3.17: A shadow bar used in experiments for background neutron scattering ef- . . . . . . . . . . . ficiency correction. Picture is adopted from Ref. [15]. Figure 3.18: NSCL staff is removing part of aluminum floor in S2 vault for leaving space in order to nest the LANA. Above S2 vault’s concrete floor, an aluminum . . . . . . . . . . layer of floor was built on which vacuum chamber sits. Figure 3.19: The front LANA’s (NWB) neutron hit pattern with 4 shadow bar in- stalled in Φlab-Θlab coordinates. 4 square areas with less counts of hits corresponds to 4 shadow bars labelled with A, B, C, D as shown. NWB bars are labelled from #1 to #25 corresponding to from the bottom bar to the top bar.Shadow bar A and B fully cover NWB bar #16 and shadow bar C and D fully cover NWB bar #8. The reaction system for this plot is 48Ca +124 Sn at E/A = 56 M eV . . . . . . . . . . . . . . . . . . . . . Figure 3.20: (left panels) Neutron hit position spectrum corresponding to shadow bar B (top) and D (bottom). Red lines indicates the fitted functions. (right panels) Scaled fitted functions indicating the background scattering neu- tron fraction. The lowest points’ Y values are both 0.26. The light output threshold is set to 5 MeVee for all 4 plots here. . . . . . . . . . . . . . . 83 85 87 88 89 90 91 93 94 xii Figure 3.21: (left panels) Neutron hit position spectrum corresponding to shadow bar A (top) and C (bottom). Red lines indicates the fitted functions. (right panels) Scaled fitted functions indicating the background scattering neu- tron fraction. The lowest points’ Y values are 0.27 and 0.29 for A and C respectively. The light output threshold is set to 5 MeVee for all 4 plots here. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Figure 3.22: Light output versus ToF 2D spectrum from front LANA (NWB). The top panel is the spectrum without using any VW information. The bottom panel is the same spectrum but with the condition that VW should not be triggered. We can clearly see that by requiring no hit in VW, the charged-particle events (those slanted lines in the top panel) are removed completely while keeping the rest of the spectrum undisturbed. . . . . . Figure 3.23: Light output versus ToF 2D spectrum from front LANA (NWB) gated on the events that must have VW triggered. The blue window (around 20 MeVee on Proton line) gating on proton line selects the events where a proton’s normalized ToF is in range [41.5, 43.5] ns and its light output is . . . . . . . . . . . . . . . . . . . . . . . . . in range [19.5, 20.0] MeVee. Figure 3.24: VW Geometric Mean(cid:112)Qtop ∗ Qbottom versus VW bar number 2D plot. 95 97 99 The top panel is before geometric mean calibration and the bottom panel is after geometric mean calibration using GMcalibrated = α ∗ GMraw. . . 100 Figure 3.25: VW bar #18’s geometric mean versus Y position 2D plot for events sitting in the blue window of Fig. 3.23. We can see that it is almost flat comparing to the top panel of Fig. 3.8. . . . . . . . . . . . . . . . . . . . . . . . . . 101 Figure 3.26: VW calibrated geometric mean versus NWB light output 2D plot. . . . . 103 Figure 4.1: A schematic of different pulse shapes for neutron and gamma rays. The peak heights are normalized to be the same. This picture is adopted from Ref. [9]. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 105 Figure 4.2: LANA geometric mean versus fast geometric mean 2D plots from a typical bar. The data was taken with 48Ca+124Sn at E/A = 56 MeV reaction. The top panel focuses on the low value end, which is a zoomed-in version of the bottom panel. We can barely see two lines in each panel, which are gamma rays (top) and neutrons (bottom). . . . . . . . . . . . . . . . . . 107 Figure 4.3: LANA flattened geometric mean versus fast geometric mean 2D plot. The data is the same as shown in figure 4.2 but the clustering of gamma rays (top) and neutrons (bottom) is much clearer. . . . . . . . . . . . . . . . 108 xiii Figure 4.4: A two-dimensional histogram of (Qf ast, Qtotal) from the left end of a LANA bar. The two solid curves are quadratic fits on the gamma ray cluster (top) and neutron cluster (bottom) respectively. Qf ast and Qtotal correspond to SHORT Q and LONG Q in this plot. This picture is taken from Ref. [16]. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 109 Figure 4.5: 2D vL,R plots for all the nine segments of one LANA bar, with 1/9 repre- senting the leftmost segment and 9/9 representing the rightmost segment. The solid line represents the n/γ separation line, while the dashed line is the line perpendicular to the solid line that pass through (-0.5, -0.5). This picture is taken from Ref. [16]. . . . . . . . . . . . . . . . . . . . . . . . . 110 Figure 4.6: PSD results comparisons between traditional flattened PSD method (left column) and VPSD method (right column) for both best (top row) and worst (bottom row) performance bars. . . . . . . . . . . . . . . . . . . . 113 Figure 4.7: (top panel) Neutron/γ pulse shape discrimination obtained by PPSD method. The X and Y axes correspond to vL and vR of Figure 7 in Ref. [16], after rotation to match for different segments on LANA bar. The light output threshold is set to 2 MeVee. Gamma rays aggregate in right bubble and neutrons are clustered in left bubble. (bottom panel) 1D projection plot from the plot in top panel. Two Gaussian functions are fitted to γ events peak (red dash line) and neutron events peak (green dash line). The blue solid line represents the sum of two Gaussian functions. Taking events with gate of P P SD-X < −0.6 ensures getting neutrons with nearly no gamma contamination. . . . . . . . . . . . . . . . . . . . 114 Figure 4.8: Gamma contamination as a function of particle incident energies, if the scintillation material used in this experiment to detect neutron has no PSD capability. The data was taken with 48Ca+124Sn at E/A = 56 MeV reaction. The light output threshold is set to 5 MeVee. . . . . . . . . . . 115 Figure 4.9: Pure neutron TOF spectra obtained by two correction methods. Black dash line is calculated based on the Gaussian fits in figure 4.7: Take neutron events with ”PPSD X>0.6” and then correct for 94.2% efficiency. Green solid line is the same green line shown in figure 4.10 using gamma peak matching method. . . . . . . . . . . . . . . . . . . . . . . . . . . . 117 xiv Figure 4.10: (top panel) Same as the bottom panel in figure 4.7 but set a gate at “P P SD-X = −0.1” to get pure gamma ray events and neutron with gamma contamination. (bottom panel) ToF spectra with different gate conditions. Spectrum of neutron and gamma where charged-particles are vetoed is drawn in black. Red line shows pure gamma ray events with condition “P P SD-X > −0.1” while blue line shows neutron with gamma contamination with condition “P P SD-X < −0.1”. Green line displays the ToF spectra after correction using gamma peak matching method, which is the same line drawn also in green in figure 4.9. . . . . . . . . . . . . . 119 Figure 4.11: Comparison of neutron light output spectra from experiments and SCINFUL- QMD with different incident neutron energies at 20 MeV (top panel) and 40 MeV (bottom panel). . . . . . . . . . . . . . . . . . . . . . . . . . . . 124 Figure 4.12: Comparison of neutron light output spectra from experiments and SCINFUL- QMD with different incident neutron energies at 60 MeV (top panel) and 80 MeV (bottom panel). . . . . . . . . . . . . . . . . . . . . . . . . . . . 125 Figure 4.13: Comparison of neutron light output spectra from experiments and SCINFUL- QMD with different incident neutron energies at 100 MeV (top panel) and 120 MeV (bottom panel). . . . . . . . . . . . . . . . . . . . . . . . . . . 126 Figure 4.14: Comparison of neutron light output versus incident neutron energy 2D plots from experiment data (left) and SCINFUL-QMD simulation (right). 127 Figure 4.15: The neutron detection efficiency of LANA as a function of neutron incident energy. Detection efficiencies for neutrons with incident energies more than 25 MeV are used. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 128 Figure 4.16: Neutron energy spectrum measured by the front LANA from 48Ca+64Ni at E/A = 140 M eV reaction. The data come from a batch of continuous runs of around 18 hours. The Y-axis is counts per bin and the X-axis is neutron energy in MeV in lab frame. The black line represents the raw spectra and the red line shows the spectra after detection efficiency correction using the efficiency data from Fig. 4.15. All data shown here have a deduced impact parameter gate from [0,0.4). . . . . . . . . . . . . 130 xv Figure 4.17: Neutron energy spectrum from 48Ca+64Ni at E/A = 140 M eV reaction with angular cuts. The Y-axis is count per bin from a batch of continuous runs of 48Ca+64Ni at E/A = 140 M eV reaction and the X-axis is neu- tron energy in MeV in lab frame. The error bars only reflect statistical uncertainties. From red, green and blue lines, different polar angular cuts from around 32°, 38° and 46° are applied respectively. Solid lines show the result after fitting with moving source model. The surface temperature of the moving source model is extracted to be 26±1 MeV. Similar moving source fits of the proton spectra in the same angular range yield similar results. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 132 xvi Chapter 1 Introduction 1.1 The Atomic Nucleus All visible things on earth are made of atoms. Each atom consists of a nucleus at the center and a cloud of electrons. Although this is common knowledge nowadays, atoms were regarded as the most fundamental particles until early 19th century. In 1897, Joseph John Thomson measured the charge-to-mass-ratio (e/m) of cathode ray particle using deflection in both electric and magnetic field. He discovered that the cathode ray particle turned out to be 2000 times lighter than hydrogen (the lightest atom). This discovery, for the first time, showed that atoms are not the indivisible particles of matter. With the discovery of electron, J.J. Thomson incorrectly postulated the “plum pudding model” assuming the negative-charged electrons were distributed throughout the atom in a uniform sea of positive charge. In 1919, with the help of Hans Geiger and Ernest Marsden, Ernest Rutherford devised an experiment that observed the deflection of alpha particles bombarded at a thin sheet of metal foil [17]. He expected that the positively charged alpha particles would pass straight through the foil with little deflection if “plum pudding model” was correct. However, Geiger and Marsden spotted that some alpha particles deflected at very large angles as shown in the left panel of figure 1.1. Considering the mass of an alpha particle is about 8000 times that of an electron, Rutherford realized that the mass of the atom must be concentrated 1 Figure 1.1: The schematic showing the “plum pudding model” proposed by J.J. Thomason (left) and the model proposed by Rutherford (right) after his discovery experiment that showed back scattering of alpha particles. Alpha particles penetrating atoms are represented by horizontal black arrow. Blue circles represents electrons and red area shows the region with positive charge. The sizes of electrons and nuclei are not drawn in scale. at a small point with positive charge. Almost all of the mass of an atom is located in the nucleus, with a very small contribution from the electron cloud. The radius of the atom can be from about 20,000 (uranium) to 60,000 (hydrogen) times bigger than the nuclear radius, which means the density of nucleus is extraordinarily high, in the order of 1017 kg/m3. The atomic nucleus composes of even smaller constituents. Every nucleus consists of protons and neutrons. In experiments that impinge alpha particles into pure nitrogen gas, scintillation detectors showed the signatures of typical hydrogen nuclei as a product. In 1919, Rutherford interpreted this observation as the process of alpha particle knocking a hydrogen nuclei out of nitrogen. After observing Blackett’s cloud chamber images in 1925, Rutherford realized that the actual process was the opposite case: the alpha particle is first 2 Thomson ModelRutherford Model captured by nitrogen, then a proton is emitted and an oxygen nucleus is left. This was the first reported nuclear reaction: 14N + α = 17O + p. Several years after the discovery of proton, in the year of 1930, Walther Bothe and Herbert Becker found that an unusually penetrating radiation was produced when energetic alpha particles were ejected on certain light elements such as beryllium, boron or lithium. In 1932, James Chadwick determined that this new type of radiation was not gamma rays but uncharged particles with about the same mass as the proton [18]. These particles are neutrons. Nuclear strong force binds protons and neutrons together against the repulsive electrical force between the positively charged protons. Studying the inner structure and interaction of such matter constituting of neutrons and protons is the topic of nuclear physics. 1.2 Nuclear Equation of State People have been attempting to understand nuclei through a variety of models such as liquid drop model (LDM) [19], droplet model (DM) [20], Thomas-Fermi model and density functional methods [21]. Here I will discuss the LDM, the simplest one among them. It was first proposed by George Gamow and further developed by Niels Bohr and John Archibald Wheeler [19]. This model was first formulated in 1935 by German physicist Carl Friedrich von Weizsacker. Based on the short range nature of the nuclear interaction and the saturation property of nuclear matter, LDM treats the nuclei as macroscopic drops of incompressible fluid of nuclear matter [22]. In LDM, the nuclear binding energy (the minimum energy required to completely disassemble a nucleus into separate protons and neutrons) of a nucleus 3 of protons Z, neutrons N and mass A=Z+N is written as: B(A, Z) = avA + asA 2 3 + ac Z2 1 3 A (N − Z)2 A + aa + O(A, Z) (1.1) The first term with av is the volume term: the binding energy is proportional to the volume of the nuclei, so it is also proportional to the mass A, the total number of nucleons. The second term with as is a correction to the volume term: the strong force that binds the nucleons together has a very limited range and a nucleon only interact strongly with its closest neighbors. Therefore, the nucleons on the surface of the drop has less neighbors to interact with, thus reducing the binding energy. The third term starting with ac takes the Coulomb repulsion between protons into account. The next term with aa is known as the asymmetry term. Based on the Pauli exclusion principle, no two identical fermions can occupy the same quantum state within a quantum system simultaneously. Consider a nucleus with imbalanced number of protons and neutrons, e.g., with more neutrons (N > Z), neutrons have to occupy higher single particle energy levels than protons, reducing its binding energy, when compared to a nucleus with equal number of protons and neutrons (N = Z). We call the excess in the proton or neutron number as the isospin asymmetry δ = N−Z Z . The last term O(A, Z) is called pairing term to refine the model. Due to Pauli principle, a nucleus would have lower energy if the number of protons with spin up equals to the number of protons with spin down. To describe the collective state of a macroscopic system such as neutron star, we must move beyond the liquid drop model. Nuclear equation of state (EOS) is used to describe the relationship between pressure, temperature, energy, and density of the system. Nuclear EOS can be approximately expressed (at zero temperature) in terms of the sum of energy 4 from a term that describes the symmetric nuclear matter with equal densities of neutrons and protons and energy from an asymmetric term consisting of nuclear matter with different densities of proton and neutron: E(ρ, δ) = E(ρ, δ = 0) + S(ρ)δ2 (1.2) Isospin asymmetry δ here is more generally written as: ρn − ρp ρn + ρp δ = (1.3) where ρn and ρp is the density of neutron and proton. We call the later term S(ρ)δ2 the symmetry energy. It is approximately the energy cost of converting symmetric nuclear matter into pure neutron matter. Nuclear EOS governs many fundamental properties of the nuclear matter and plays a crucial role in understanding not only terrestrial nuclei and nuclear reaction process, but also the evolution of nuclear structure and many astrophysical objects. Over the last four decades, much knowledge about the E(ρ, 0) term in Eq. 1.2, which is the EOS of symmetric nuclear matter, has been obtained through the hard work of scientists in both nuclear physics and astrophysics [23]. In more recent years, significant efforts have been taken to explore the relatively poorly known S(ρ). The symmetry energy plays a fundamental role especially in the stability of neutron stars. The pressure that supports the neutron star against the collapse from gravitational force, mainly comes from the symmetry energy [24]. It also strongly influences neutron star structure, radius, moment of inertia, cooling process and prevents the collapse of a neutron star into a black hole [25, 26, 27, 28, 29, 30]. The first observation of a neutron- 5 star merger event in 2017, GW170817 (GW), has provided new insights to the equation of state of neutron star [31, 32, 33, 34, 35, 36]. Much progress has been made in constraining the two coefficients S0 and L in the Taylor expansion of S(ρ) around the saturation density ρ0 (0.16 nucleon/fm3): S(ρ) = S0 + (ρ − ρ0) + L 3ρ0 Ksym 18ρ0 (ρ − ρ0)2 + O((ρ − ρ0)3) (1.4) Constraints of S0 and L at sub-saturation density was obtained through various types of laboratory experiments including excitation energies of Isobaric Analog States (IAS) [37], Pygmy Dipole Resonances (PDR) [38, 39, 40], atomic mass analyses [41, 42], electric dipole polarizability [43, 44], neutron skin thickness of heavy nuclei [45] and isospin diffusion in heavy ion collision [46, 47, 48, 49, 50]. While much knowledge is obtained about S(ρ) at sub-saturation density, the high density behavior of S(ρ) remains rather uncertain. A set of Skyrme parameterizations, which is selected based on the fit of binding energy difference between 100Sn and 132Sn nuclei, gives very different predictions about the energy per nucleon in pure neutron matter at densities above saturation density [3]. Figure 1.2 shows that the uncertainties remain large when S(ρ) with ρ > ρ0 with selected Skyrme interactions from Ref. [2]. For simplicity, it is common to write the density dependence of S(ρ) in the following form: S(ρ) = Skin( ρ ρ0 2 3 + Sint( ) )γ ρ ρ0 (1.5) The first term with Skin is the Thomas-Fermi kinetic energy. In a uniform system of Fermions of spin S = 1 2 in 3 dimensions, the relationship between Fermi momentum kF and the 6 Figure 1.2: Density dependence of the symmetry energy from the Skyrme interactions used in Ref. [2].The shaded region is obtained from HIC experiments. Picture is adopted from Ref. [3]. density ρ is: kF = 3πρ3. Then the kinetic energy is proportional to the square of the Fermi momentum kF 2. Overall, the kinetic energy part in S(ρ) depends on ρ 2 3 . The second term with Sint is the interaction energy term. It is modeled to have power law of density dependence with factor γ. The value of γ is usually chosen between 0.5 to 2. The symmetry energy is “soft” when γ < 1 and “stiff” when γ > 1. It is worth mentioning that the symmetry energy S(ρ) is closely related to the neutron- proton effective mass splitting. From the Fock exchange term, finite range and correlation effects, the nuclear mean-field potential has momentum dependencies [51, 52, 53, 54]. The concept of effective mass was introduced to simplify the description of nucleons that are moving in a momentum-dependent mean-field potential. Consider the following Hamilton’s 7 Equation for a nucleon: ˙x = = ∂H ∂p p m + ∂V ∂p Define the effective mass m∗ such that: ˙x = p m∗ Combine Eq. 1.7 and Eq. 1.8, we have: m∗ m = 1 1 + m p ∂V ∂p (1.6) (1.7) (1.8) (1.9) Nucleons can be described as moving with their effective masses, instead of with their free masses in a momentum-dependent potential. An isoscalar effective mass m∗ the mean-field potential with m∗ s is derived from ≈ 0.65 − 0.75 where mN is the nucleon mass [7, 55]. Due s mN to the momentum dependencies in the isovector mean-field potential, the effective masses for neutron m∗ p are different from each other [52], resulting in the effective mass n and proton m∗ splitting: ∆m∗ np = m∗ n − m∗ mN p (1.10) Based on the well-known Lane potential [56], the single nucleon potential Vτ (ρ, k, δ) in 8 Equ. 1.9 can be well approximated by: Vτ (ρ, k, δ) = V0(ρ, k) ± Vsym(ρ, k) · δ (1.11) where the V0(ρ, k) is the nucleon isoscalar potential and Vsym(ρ, k) is the isovector (sym- metry) potential for nucleons with momentum k and isospin asymmetry δ at density ρ. τ = n for neutrons and τ = p for protons. With the isoscalar asymmetry δ considered, equations 1.10 can be written as: ∆m∗ np = (1 + mN ¯h2kF mN ¯h2kF (dVp/dk − dVn/dk) dVp/dk)(1 + mN ¯h2kF dVn/dk) (cid:12)(cid:12)(cid:12)(cid:12)kF (1.12) For practical systems, δ is small and can be neglected in the denominator, but the leading order term in the numerator is proportional to δ. So to a good approximation: ∆m∗ np ≈ −2δ · = −2δ · ( dVsym mN ¯h2kF dVp/dk)(1 + dk mN ¯h2kF dVsym dk ) mN ¯h2kF (cid:12)(cid:12)(cid:12)(cid:12)kF dV0 dk )2 mN ¯h2kF )2 · ( (1 + m∗ s mN (cid:12)(cid:12)(cid:12)(cid:12)kF (1.13) (1.14) Here, m∗ s is the nucleonic effective mass in symmetric matter, where Vsym vanishes. On the other side, the symmetry energy Esym(ρ) and its slope L(ρ) can be expressed as: Esym(ρ) = L(ρ) = 1 3 2 3 ¯h2K2 F 2mN ¯h2K2 F 2mN + + 1 2 3 2 Vsym(ρ, kF ) Vsym(ρ, kF ) + ∂Vsym ∂k (cid:12)(cid:12)(cid:12)(cid:12)kF kF (1.15) (1.16) Combining equations 1.12 to 1.16, the effective mass splitting depends on asymmetry δ 9 in such form [57]: ∆m∗ np ≈ δ · 3Esym(ρ0) − L(ρ0) − 1 mN m∗ 3 0 mN m∗ )2 0 EF (ρ) · ( EF (ρ0) (1.17) where m∗0 = 0.7mN at normal density. This effect influences the magnitude of shell effects in nuclei far from stability and some properties of neutrons star especially the cooling process via neutrino emission [58, 59]. However, both the sign and magnitude of the effective mass splitting ∆m∗ np are poorly determined. Different Skyrme parameterizations contained different effective mass splitting. Moreover, some microscopic calculations performed in Landau-Fermi-liquid theory [60] and the nonrelativistic Brueckner-Hartree-Fock approach [61] have predicted that m∗ n > m∗ p while other calculations using relativistic mean-field theory and relativistic Dirac-Brueckner theory [52, 62] predict that M∗ p . Additional experimental data and theoretical studies n < M∗ are urgently needed to resolve this issue. 1.3 Probing Effective Mass Splitting in Heavy Ion Collision Heavy Ion Collision (HIC) is one of the main tools used to explore the density dependence of the symmetry energy and the momentum dependence of the symmetry potential. A large range of densities can be obtained in HIC by choosing different projectile-target systems [63, 64, 65] and bombarding energies. In the energy above the Fermi energy but below around 150 MeV per nucleon, nucleons that participate in the collision first form a compressed, high-density region, then expand with the emission of light clusters. The emitted particles carry the information of nuclear symmetry energy because the motions of ejected particles 10 are determined by the nuclear interaction during such compression-expansion process. The measurements of particles emitted from HIC include neutron/proton yield ratio spectra and isospin diffusion. This strategy has been successfully applied to constrain the symmetry energy at sub-saturation density using Sn+Sn collisions [64, 66, 47, 67, 68, 69]. Transport models that take EOS as an essential input is used to interpret the results measured in HIC and to constrain EOS quantitatively. Calculations performed in trans- port models have been successful in describing observables in HIC such as isoscaling ratios and elliptic flow [50, 70]. Calculations typically employs a set of parameterizations such as Skyrme interactions that characterize the density dependence of the nuclear matter energy including the symmetry energy. Aside from the density dependence of the symmetry en- ergy, other inputs such as neutron-proton effective mass splitting ∆m∗ np that results from a momentum-dependent mean-field potential and in-medium nucleon-nucleon cross sections δN N also impact the propagation of nucleons in collisions. Results from transport model calculation predict that neutrons with high momentum emitted from the compressed par- n < m∗ n < m∗ p and calculations using p, compared to protons with same momentum [4, 22]. This effect will enhance ticipant region will experience a more repulsive potential thus a higher acceleration process if m∗ ratio of neutron over proton (n/p) at high momentum for m∗ m∗ n > m∗ p give predictions that the acceleration of protons are enhanced, resulting in a lower n/p spectral ratio. If we use M to represent the multiplicities for a certain particle, the differential yield defined as Y (n) = dMn dE and Y (p) = dMp dE and the spectral yield ratio of n/p is written as: Rn/p = Y (n) Y (p) = dMn/dE dMp/dE (1.18) 11 Eq. 1.18 is called “single ratio”. Another observable is defined as “double ratio”: DR(n/p) = (Rn/p)system1 (Rn/p)system2 (1.19) which takes the ratio of single ratios from two reaction systems. Double ratio is commonly used to take the advantage of cancellation of systematic errors which are difficult to determine experimentally such as detector efficiencies and to a certain extent, inadequacies in models. such as the Coulomb effects and clustering mechanisms. Figure 1.3 shows simulation results of Rn/p and DRn/p using Improved Quantum Molec- ular Dynamics (ImQMD) model [71, 72, 73] with two Skyrme interaction parameter sets, SLy4 [74] and SkM∗ [75]. SLy4 and SkM∗ have similar symmetry energy coefficient S0, slopes of symmetry energy L and isoscalar effective mass m∗, i.e., S0 = 32 ± 2MeV, L = 46MeV and m∗/mN = 0.7 ± 0.1. However, the effective mass splitting is different with m∗ n < m∗ for SLy4 and m∗ p for SkM∗. In figure 1.3, DR(n/p) is roughly flat or decrease slightly with nucleon kinetic energy for SkM∗ while they increase for SLy4. The sensitivity of the n > m∗ p DR(n/p) to effective mass splitting decrease with greater incident beam energies due to in- crease in nucleon-nucleon scattering [4]. These calculations suggest incident energy less than 200 MeV is best for the effective mass splitting study. 1.4 Neutron Detection Neutron detection is very difficult and this is why the discovery of neutron was relatively late in 1932 by Chadwick. Unlike protons and other charged-particles, neutron has no electric charge and does not interact with other particles through electromagnetic force. 12 Figure 1.3: (top panels) Y(n)/Y(p) ratios as a function of kinetic energy for 124Sn+124Sn at b=2fm. (bottom panels) DR(n/p) ratios as a function of kinetic energy. From left to right, the beam energies are 100 MeV, 150 MeV, 200 MeV, 300 MeV per nucleon. Figure is adopted from Ref. [4]. Neutrons are stable when bounded in nuclei. However, as a free particle, neutron beta decays with a half-life of about 880 s. It primarily interacts with other particles by strong force (hadronic interaction), which is much stronger than electromagnetic interaction but with very limited range of a few fermi. As neutral particles, neutrons do not ionize directly and their paths are weakly affected by electric and magnetic fields. They are usually detected by nuclear interactions that produce secondary charged-particles with relatively low interaction probability. Because the nuclear force has extremely short range, neutrons need to be very close within around 10−15 m of another nucleus in order for scattering to take place. Given 13 the radius of a nucleus is at least 10,000 times smaller than the radius of an atom, the probability of neutrons interact with normal material is relatively low. Thus, the detection of neutrons represents a considerable technical challenge. The energy of neutrons to be detected can vary from below meV (milli-electron volt) called ultracold neutrons up to more than TeV (Tera-electron volt) which are mainly created in large accelerators or in certain astrophysical activities. Neutrons are commonly classified according to their energies as listed in the following table [1]: Designation Ultracold neutrons Cold neutrons Thermal neutrons Epithermal neutrons Intermediate neutrons Fast neutrons High-energy neutrons Neutron Energy < 0.2 meV 0.2 meV - 2 meV 2 meV - 100 meV 100 meV - 1 eV 1 eV - 10 keV 10 keV - 20 MeV > 20 MeV Table 1.1: Neutron designation table for different energy ranges. This table is adopted from Ref. [1] Neutrons at different energy range require different detection techniques. The energy range of neutrons we measured in our experiment is from tens of MeV to a couple hundred MeV; the fast and high energy neutrons. The principles of detection of fast neutrons are based on ionizing radiation. Because neutrons do not directly ionize in material, charged- particles need to be induced by neutrons first, and then the energy of charged-particles are converted into different types of signal through ionization processes. The main interactions of neutrons with nuclei are the following three processes: elastic scattering, inelastic scattering and radiative neutron capture [1]. The most important process for such neutrons is elastic scattering on light target nuclei, especially elastic n-p scattering due to the higher cross- sections compared to other reaction channels, as shown in figure 1.4. 14 Figure 1.4: Neutron reaction cross sections adopted in SCINFUL-QMD model. Picture is adopted from Ref. [5]. Popular types of neutron detectors include gas-filled detectors, semiconductor and scintil- lators [76]. Organic plastic and liquid scintillators composed of large percentage of hydrogen are widely used in neutron measurement. Not only are they fast detectors, a requirement for Time of Flight (ToF) measurement to determine the energy of the detected neutrons, but they are robust and relatively cost effective to build a large detector like LANA. Since recoiling protons is the dominant neutron interaction in scintillators, neutrons and charged- particles emitted from nuclear reaction cannot be distinguished from the light they produced in the scintillator. This is why we designed and built a thin plastic scintillator wall which has 1% neutron detection efficiency and 100% charged-particle detection efficiency. It will be placed in front of existing Large Neutron Area Array (LANA) at NSCL to provide extra in- formation in order to distinguish and veto charged-particle events from neutrons detected in LANA. This wall is called Charged Particle Veto Wall (VW) and its detail will be presented 15 in Chapter 2. 1.5 Improvements from Previous Experiment Due to difficulties in measuring neutrons, few experimental data on neutron and protons yield or spectra ratios exist. In 2009, the HiRA group measured the neutron and protons spectra emitted from Sn+Sn reactions [6, 7]. Conclusions from that study regarding the effective mass are ambiguous both due to quality of the data. In the current experiment, We studied smaller, near symmetric reaction systems such as Ca+Ni and Ca+Sn to allow extensive calculations with CPU intensive transport models such as the AMD models which have sophisticated treatment of cluster production. We also study very asymmetric systems such as Ca+Sn to study the in-medium NN cross sections. Based on experience with previous experiments, we have also made major improvement in the experimental setup regarding detection of charged particles and the neutrons. The work of this thesis mainly focuses on the improvement made in neutron detection while the thesis of Sean Sweany focuses on the improvement in detecting charged particles in this experiment. Fig. 1.5 shows the DR(n/p) spectral double ratios measured in previous experiment on central collisions of 124Sn+124Sn and 112Sn+112Sn systems at E/A=120 MeV. The red and blue shaded bands correspond to the predictions from ImQMD using SLy4 and SkM* interactions. As expected the results split starting at center of mass energy of the nucleons around 50 MeV/u. Unfortunately due to limitation of the charged particle detector, the data does not extend beyond 80 MeV/u in the center of mass energy. The large statistical uncertainties mainly arise from contamination of charged particles in the neutron spectra due to deficiencies of the old veto arrays. Because of the difficulties for most transport models 16 Figure 1.5: neutron/proton double ratio for 124Sn +124 Sn/112Sn +112 Sn reaction systems at 120 MeV/u for central collisions. The symbols are experimental data. Red line shows calculation using SLy4 which has m∗ p while blue line comes from calculation using SkM∗ which has m∗ p. The picture is adopted from Ref. [6]. Open circles are the original version and the close circles are data of the most updated version. The original data points did not take into account the efficiency corrections for charged-particles [7, 8]. n > m∗ n < m∗ 17 in producing the relative abundances of light isotopes in HIC, instead of comparing free neutrons and protons spectra, we use the observable the coalescence invariant (CI) neutron and proton spectra by combining the free nucleons with those bound in light isotopes with 1 −0.1” as pure gamma ray events and the rest with “P P SD-X < −0.1” as neutron with gamma contamination. The time of flight spectrum for pure gamma events with “P P SD-X > −0.1” are shown in the bottom panel of figure 4.10. ToF spectrum for “P P SD-X < −0.1” corresponding to neutron with gamma 116 Figure 4.9: Pure neutron TOF spectra obtained by two correction methods. Black dash line is calculated based on the Gaussian fits in figure 4.7: Take neutron events with ”PPSD X>0.6” and then correct for 94.2% efficiency. Green solid line is the same green line shown in figure 4.10 using gamma peak matching method. 117 Corrected using Gaussian Fit MethodCorrected using Gamma Peak Matching contamination is drawn as the green line in figure 4.10. Red line representing pure gamma ray events has a strong prompt gamma peak at around 16 ns. It also has a long tail from reactions on surrounding material that looks like neutrons. While the majority of blue line includes ALL of the neutron events, there are also gamma ray events from the tail of the gamma ray Gaussian, and the (n,gamma) background reactions mixed in. To get rid of the gamma contamination in blue line, we first assume that the gamma events mixed in blue line have the same ToF spectrum shape as pure gamma ray ToF spectrum in red with some scaling constant. Then we use the prompt gamma peaks in red and blue lines as an indicator of this scaling constant. We can calculate this scaling constant by doing division between the prompt gamma peak count heights of red and blue lines. In the case of figure 4.10, the scaling constant is 0.449. To get pure neutron spectrum after correction, we multiply the pure Gamma Tof spectrum by 0.449 and subtract it from neutron ToF with gamma contamination. The neutron ToF after correction is shown in green line both in figure 4.10 and figure 4.9. Figure 4.9 compares the pure neutron ToF spectra after correction using two methods described above. The black dots are the pure neutron spectrum rescaled by 1/0.94 as dis- cussed in the first procedure. The green line is what we did second by taking everything in the neutron PSD peak and subtracting the gamma contamination. These two spectra agree with each other within 2 percent. The gamma contamination rate in these spectra is within 1 percent based on previous analysis and indeed we can see a small prompt gamma peak at around 16 ns. Because the speed of neutrons emitted from our experiments can not be close to the speed of light, particles with TOF below 20 ns will be eliminated from neutron events in further neutron energy and momentum spectra analysis. This will make the gamma contamination in neutron spectra much smaller than 1 percent. 118 Figure 4.10: (top panel) Same as the bottom panel in figure 4.7 but set a gate at “P P SD-X = −0.1” to get pure gamma ray events and neutron with gamma contamination. (bottom panel) ToF spectra with different gate conditions. Spectrum of neutron and gamma where charged- particles are vetoed is drawn in black. Red line shows pure gamma ray events with condition “P P SD-X > −0.1” while blue line shows neutron with gamma contamination with condition “P P SD-X < −0.1”. Green line displays the ToF spectra after correction using gamma peak matching method, which is the same line drawn also in green in figure 4.9. 119 Neutron + GammaPure Gamma RayNeutron with Gamma ContaminationPure Neutron after correction 4.2 Neutron Light Output Comparison with SCINFUL-QMD To successfully detect a neutron in LANA, the neutron needs to interact with the scintillating material and deposit enough that it is above a threshold value for which electronics can generate a trigger signal reliably. Because neutrons are neutral, a neutron interacts with the detector via the strong interaction making the detection probability significantly less than unity (<10 percent) for a thickness of 6.35 cm of the LANA scintillator bars. There are two different ways that one can define the efficiency; therefore it is easy to become confused about what one is discussing. This is because the efficiency can be defined in a way that it includes the solid angle and it can also be defined in a way that it includes only the probability that a neutron interacts deposits energy above the threshold value if it goes through the solid angle subtended by the detector. For the following discussion, we take the second definition and define the neutron de- tection efficiency as the probability of detecting a neutron that traverses a detector with a particular incoming kinetic energy. Determining this efficiency is indispensable in extracting the absolute magnitude and shape of neutron energy spectrum. It depends on the thickness of the scintillator and of the electronic threshold applied to the scintillator signal that must be overcome by the induced signal. The NE213 scintillator in LANA is a hydrocarbon fluid containing both protons (hydro- gen nuclei) and carbon atoms in it. Neutrons below 10 MeV are typically scattered elastically by these protons and subsequent interactions of the recoiling protons with the scintillators create the photons to be detected. Neutrons can be detected also through inelastic, fusion, or breakup nuclear reactions on the carbon nuclei in the scintillator fluid. This becomes increasingly important for more energetic neutrons. When a neutron reacts with a carbon 120 nucleus, one or more charged particles can thereby be generated. Besides considering various reaction channels between incident neutrons and scintillation material, we need to take how the scintillator geometry influences the induced signal into account because events can occur that involve recoiled protons escaping the detector without depositing all of their energies. The resulting signal may be above the detection threshold and be counted or below the detec- tion threshold and be lost. Developing a simulation codes that can handle all these processes is a complex task because some of the reaction channels are not well described in current detector simulations and it is an ongoing project for one of our Korean collaborators. To obtain preliminary neutron spectra in this thesis, we used a code called SCINFUL-QMD [5] developed by Satoh et al. at the Japan Atomic Energy Agency. This code is one of the more complete simulation in modeling neutron reaction processes in the detector and reproducing experimental data. Using SCINFUL-QMD, we are able to determine the neutron detection efficiency in LANA. To validate the code, we compare the simulated neutron light response function to our experimental data. SCINFUL-QMD code simulates the NE-213 scintillator light response of neutrons from 0.1 MeV to 3 GeV using the Monte Carlo technique. Above 150 MeV, multiparticle breakup is an important decay mode and to model this, SCINFUL-QMD incorporates a quantum molecular dynamics model (QMD) and the statistical decay model (SDM). The QMD semi- classically describes the behavior of nucleons and mesons during nucleus collision and thus simulates the process for these more complicated nucleon reactions. The current version of SCINFUL-QMD code restricts the configuration of detector to be in the shape of a cylinder with radius and height as input variables. We adjust the input variables including cylin- der’s height, radius similar to the geometry of our unit cell of 6.35cm and 7.62cm as depth and height, and light attenuation factor is then adjusted by comparing the simulated light 121 response to the data. Figures 4.11 , 4.12 and 4.13 show a more comprehensive neutron light response compar- ison between experimental data and SCINFUL-QMD simulation using neutron light output versus incident neutron energy. Figure 4.14 shows a two dimensional representation of the light output response as a continuous function of the neutron energy En. In general, we find that SCINFUL-QMD reproduced the shapes of light output spectra pretty well for different incident neutron energies. This is consistent with the study by Coupland et al. [9]. For that work, accuracy in the shape of the spectrum was extremely important as the veto wall was not available then to eliminate charged particle contamination in the neutron walls. So it was necessary to know details of the light output response in the regions of neutron energy and light output where the charged particle contamination was absent. The efficiency for detecting a neutron of a given energy in a given scintillator bar corre- sponds to the probability of getting a signal in the neutron wall that is over the experimental threshold when a neutron of that energy hits that bar in the neutron wall. To calculate the efficiency for a neutron at certain energy,we need to choose the threshold in the geometric mean signal in the bar the neutron signal must exceed. During the experiment, we set a hardware threshold on the bars so that a signal of 3 MeVee would be detected everywhere on the bar. Here, we raise that threshold to 5 MeVee. To calculate the efficiency for detec- tion of an energy deposition above this threshold, we integrate the properly normalized light output spectra e.g. figures 4.11 , 4.12 and 4.13, over all light output values higher than that threshold value. Each of these light output curves gives the efficiency at one energy. Doing this at one MeV intervals in the incident neutron energy provides the neutron efficiency as a function of energy. There are discrepancies between the observed light output in Figures 4.11, 4.12 and 122 4.13 and the simulation. These discrepancies between data and simulation may be caused by the significant limitations of changing the detector geometries and moving the position of neutron source in SCINFUL-QMD code. It is also hard to reproduce the light attenuation, scattering and collection processes in simulation. Ongoing efforts are being taken to embed SCINFUL-QMD into GEANT4 platform to make the code more versatile and more accurate for our purposes. The final neutron spectra will be obtained after that effort is complete. Nevertheless, we now examine the neutron spectra using the present efficiency calculations. Based on SCINFUL-QMD simulation with a threshold of 5 MeVee, the neutron detection efficiency of LANA for neutrons with energies greater than 25 MeV is calculated as shown in figure 4.15. Low energy neutrons have larger contamination due to scattering from the beam dump and the floor and ceiling of the experimental vault and we have not studied that contamination extensively. 25 MeV also corresponds to the energy threshold of the charged particle detected in HiRA. 123 Figure 4.11: Comparison of neutron light output spectra from experiments and SCINFUL- QMD with different incident neutron energies at 20 MeV (top panel) and 40 MeV (bottom panel). 124 Figure 4.12: Comparison of neutron light output spectra from experiments and SCINFUL- QMD with different incident neutron energies at 60 MeV (top panel) and 80 MeV (bottom panel). 125 Figure 4.13: Comparison of neutron light output spectra from experiments and SCINFUL- QMD with different incident neutron energies at 100 MeV (top panel) and 120 MeV (bottom panel). 126 Figure 4.14: Comparison of neutron light output versus incident neutron energy 2D plots from experiment data (left) and SCINFUL-QMD simulation (right). 127 ExperimentSCINFUL-QMD100 Figure 4.15: The neutron detection efficiency of LANA as a function of neutron incident energy. Detection efficiencies for neutrons with incident energies more than 25 MeV are used. 128 4.3 Preliminary Neutron Spectrum We extracted preliminary neutron spectrum from 48Ca+64Ni at E/A = 140 M eV reaction as shown in figure 4.16. Firstly, when calculating the raw energy spectra of neutron drawn in black, we adopt a light output threshold of 5 MeVee to ensure the whole LANA can detect particle hits uniformly. The majority of gamma contamination comes as prompt gamma so they are filtered out naturally by the time-of-flight gate cut. For gammas that hit the LANA within the time window of neutrons, most of them come from the experiment vault background. Most of them are eliminated by utilizing the neutron/γ discrimination method described in Chapter 4.1. To measure the neutrons from central collisions, a deduced impact parameter gate from [0,0.4) is applied. The red line above the black line in figure 4.16 takes the detection efficiency correction into account by using the SCINFUL simulation results from figure 4.15. In figure 4.17, neutron spectra with several angular cuts are shown for the same reaction system. The Y-axis is count per bin and the X-axis is neutron energy in MeV in lab frame. A moving source model is applied to extract the following parametrizations of this moving source: apparent temperatures T , velocities Vs, multiplicities N and Coulomb energy Ec. Here, the Coulomb energy is the energy gained by Coulomb repulsion of the exciting proton by the rest of the system. In the context of compound nucleus, that would be called the Coulomb barrier. In this model, the particles are assumed to be emitted isotropically from one moving sources, following Maxwell-Boltzman distribution in source rest frame [107]: d2σ dΩdE = N 2(πT )3/2 (E − Ec)1/2exp[−(E − Ec)/T ] (4.5) To apply this model to neutron energy spectrum measured in lab frame, we need to do 129 Figure 4.16: Neutron energy spectrum measured by the front LANA from 48Ca+64Ni at E/A = 140 M eV reaction. The data come from a batch of continuous runs of around 18 hours. The Y-axis is counts per bin and the X-axis is neutron energy in MeV in lab frame. The black line represents the raw spectra and the red line shows the spectra after detection efficiency correction using the efficiency data from Fig. 4.15. All data shown here have a deduced impact parameter gate from [0,0.4). 130 the following transformation [108]: d2σ dΩlabdElab (cid:48) E (cid:48)(cid:48) E N = 2(πT )3/2 (cid:114) = Elab − Ec (cid:48) − 2 = E (cid:48) )1/2exp[−(E (cid:48)(cid:48) )/T ] (E (cid:48) ( 1 2 mV 2 s )cos(θlab) + 1 2 mV 2 s E (4.6) (4.7) (4.8) where θlab is the emitting angle measured in lab frame, m is the mass of the emitted particle and Elab is the energy of the emitted particle measured in lab frame. To do this moving source fit for figure 4.17, because neutron has no charge, we set Ec to be 0. T , N and Vs are the global free parameters for 3 spectra with different angular cuts. The surface temperature is determined to be 26±1 MeV from the fitting using moving source model. The best fit value for the source velocity Vs is found to be 0.48 times of the beam velocity. The extracted temperature is consistent with a mid rapidity source created in heavy ion collisions at 140 MeV/u incident energy1. Consider the case of a moving black body: its black body spectrum shows a frequency shift due to the relativistic Doppler effect which depends on the angle α between the observer and the source: f(cid:48) = f . This leads 1− v (cid:113) c cos(θ) 1−( v c )2 to an angle dependent temperature and the observer will not detect an isotropic black body spectrum [109, 110]. 1Transformation assumes same temperature in the center-of-mass frame. However, recent studies suggest that the temperature could be different in different frame. 131 Figure 4.17: Neutron energy spectrum from 48Ca+64Ni at E/A = 140 M eV reaction with angular cuts. The Y-axis is count per bin from a batch of continuous runs of 48Ca+64Ni at E/A = 140 M eV reaction and the X-axis is neutron energy in MeV in lab frame. The error bars only reflect statistical uncertainties. From red, green and blue lines, different polar angular cuts from around 32°, 38° and 46° are applied respectively. Solid lines show the result after fitting with moving source model. The surface temperature of the moving source model is extracted to be 26±1 MeV. Similar moving source fits of the proton spectra in the same angular range yield similar results. 132 Chapter 5 Summary The goal of detecting neutrons in experiment 14030 and 15190 is to construct neutron/proton ratios observable to study the symmetry energy. Detection of the charged-particles (including protons) using the upgraded HiRA10 array has been analyzed. This dissertation describes the initial effort to analyze the neutrons. Following is a list of major improvements of this thesis experiment compared to the earlier version of experiment done in our group: 1. Full removal of charged-particle contamination from neutron spectra using the newly- built Charged-Particle Veto Wall. 2. Development of an innovative method that can calibrate the time, position and light output of neutron scintillation detectors using only cosmic rays. 3. Development of a new algorithm for neutron/γ pulse shape discrimination using VPSD method. 4. Measurement of charged-particles with greater extended kinetic energy ranges. For example, HiRA10 can measure protons with kinetic energy up to 200 MeV, compared to 116 MeV for the original HiRA. The ability to fully remove charged-particle contaminations from neutron spectra is one of the essential requirements to successfully compare experimental n/p yield ratios with 133 simulation results, especially for the high kinetic energy regions where the observable of yield is much more sensitive to the effective mass splitting as shown in Fig. 1.3. By successfully vetoing the charged-particles in the LANA walls, we can successfully measure the neutron energy spectra beyond 200 MeV which was not achieved in previous study. This thesis describes the measurement of neutrons emitted in heavy ion collisions. In experiment E14030 and 15190, we studied 16 reaction systems including 2 beam isotopes of 40Ca and 48Ca, 2 beam energies of E/A = 56 MeV and 140 MeV, and 4 targets of 58Ni, 64Ni, 112Sn and 124Sn. Neutrons and charged-particles were measured in these reaction systems in order to ultimately identify and construct observables that are sensitive to symmetry energy, and extract EoS constraints using transport calculations. The major part of this dissertation focuses on the measurements of neutrons with im- proved detection equipment and analysis methods. A Forward Array used as the start timing detector for the neutron wall, was upgraded to increase its forward angle coverage. A new Charged-Particle Veto Wall was designed, constructed and placed in front of LANA to pro- duce clean neutron spectrum without charged-particle contamination. An accurate, fast and convenient calibration method of LANA using cosmic muons was tested [111] and used to de- termine the intrinsic timing and position resolution of the LANA. Amazingly, the resolution is nearly the same as 20 years ago. We also developed a new neutron/γ discrimination pro- cedure based on LANA’s PSD capability which can better separate neutrons from gammas when compared to the traditional method. The discrimination provides accurate quantita- tive gamma contamination and neutron loss estimation [16] suggesting that PSD is needed in heavy ion collisions as contamination from gammas reaches 50 percent for high energy neutrons. We found that neutron light output spectrum acquired from LANA is comparable to the simulation results of SCINFUL-QMD code [5] by adjusting the attenuation length 134 parameter in the code. This allows us to use SCINFUL to simulate the neutron efficiencies of LANA from 10 to 100 MeV neutrons. To validate the analysis procedure, we generate the efficiency corrected neutron spectra for 48Ca+64Ni reactions at E/A = 140 M eV . Us- ing a moving source model, the source’s surface temperature (20±1 MeV) and velocity (0.5 vbeam) are extracted. These parameters are reasonable and similar to those extracted from the proton spectra. Due to the complexity of the experiment and the large amount of data acquired (27 TB), only the neutrons emitted in two reactions and WallB of the two LANA walls have been studied in details. The data presented in this dissertation mainly comes from 48Ca+124Sn and 48Ca+64Ni reaction system to demonstrate various analysis procedures needed for neu- tron measurements. Care is needed to extend the analysis to the remaining 14 systems. This thesis provides more of a preview of the data to come than a conclusion of the project. More careful investigation in the future is needed to validate every analysis step and determine the corresponding uncertainties. On the transport model side, efforts are being taken to identify better observables and address uncertainties in the simulations of heavy ion collisions. 135 BIBLIOGRAPHY 136 BIBLIOGRAPHY [1] A. Klett, Neutron Detection, pp. 759–790. Berlin, Heidelberg: Springer Berlin Heidel- berg, 2012. [2] B. Alex Brown, “Neutron radii in nuclei and the neutron equation of state,” Phys. Rev. Lett., vol. 85, pp. 5296–5299, Dec 2000. [3] M. B. Tsang, J. R. Stone, F. Camera, P. Danielewicz, S. Gandolfi, K. Hebeler, C. J. Horowitz, J. Lee, W. G. Lynch, Z. Kohley, R. Lemmon, P. M¨oller, T. Murakami, S. Ri- ordan, X. Roca-Maza, F. Sammarruca, A. W. Steiner, I. Vida˜na, and S. J. Yennello, “Constraints on the symmetry energy and neutron skins from experiments and theory,” Phys. Rev. C, vol. 86, p. 015803, Jul 2012. [4] Y. Zhang, M. Tsang, Z. Li, and H. Liu, “Constraints on nucleon effective mass splitting with heavy ion collisions,” Physics Letters B, vol. 732, pp. 186 – 190, 2014. [5] D. Satoh, T. Sato, N. Shigyo, and K. Ishibashi, “Scinful-qmd: Monte carlo based computer code to calculate response function and detection efficiency of a liquid organic scintillator for neutron energies up to 3 gev.” [6] D. D. S. Coupland, M. Youngs, Z. Chajecki, W. G. Lynch, M. B. Tsang, Y. X. Zhang, M. A. Famiano, T. K. Ghosh, B. Giacherio, M. A. Kilburn, J. Lee, H. Liu, F. Lu, P. Morfouace, P. Russotto, A. Sanetullaev, R. H. Showalter, G. Verde, and J. Winkelbauer, “Probing effective nucleon masses with heavy-ion collisions,” Phys. Rev. C, vol. 94, p. 011601, Jul 2016. [7] P. Morfouace, C. Tsang, Y. Zhang, W. Lynch, M. Tsang, D. Coupland, M. Youngs, Z. Chajecki, M. Famiano, T. Ghosh, G. Jhang, J. Lee, H. Liu, A. Sanetullaev, R. Showalter, and J. Winkelbauer, “Constraining the symmetry energy with heavy- ion collisions and bayesian analyses,” Physics Letters B, vol. 799, p. 135045, 2019. [8] P. Morfouace, W. Lynch, and M. Tsang, “Charged-particle detection efficiencies of close-packed csi arrays,” Nuclear Instruments and Methods in Physics Research Section A: Accelerators, Spectrometers, Detectors and Associated Equipment, vol. 848, pp. 45 – 53, 2017. [9] D. Coupland, Probing the nuclear symmetry energy with heavy ion collisions. PhD thesis, Michigan State University, 2013. [10] D. Sarantites, P.-F. Hua, M. Devlin, L. Sobotka, J. Elson, J. Hood, D. LaFosse, J. Sarantites, and M. Maier, “The microball design, instrumentation and response 137 characteristics of a 4¨I-multidetector exit channel-selection device for spectroscopic and reaction mechanism studies with gammasphere,” Nuclear Instruments and Methods in Physics Research Section A: Accelerators, Spectrometers, Detectors and Associated Equipment, vol. 381, no. 2, pp. 418 – 432, 1996. [11] “Faroarm.” [12] D. DellˆaAquila, S. Sweany, K. Brown, Z. Chajecki, W. Lynch, F. Teh, C.-Y. Tsang, M. Tsang, K. Zhu, C. Anderson, A. Anthony, S. Barlini, J. Barney, A. Camaiani, G. Jhang, J. Crosby, J. Estee, M. Ghazali, F. Guan, O. Khanal, S. Kodali, I. Lom- bardo, J. Manfredi, L. Morelli, P. Morfouace, C. Niu, and G. Verde, “Non-linearity effects on the light-output calibration of light charged particles in csi(tl) scintillator crystals,” Nuclear Instruments and Methods in Physics Research Section A: Acceler- ators, Spectrometers, Detectors and Associated Equipment, vol. 929, pp. 162 – 172, 2019. [13] P. Zecher, A. Galonsky, J. Kruse, S. Gaff, J. Ottarson, J. Wang, F. De´ak, A. Horv´ath, A. Kiss, Z. Seres, K. Ieki, Y. Iwata, and H. Schelin, “A large-area, position-sensitive neutron detector with neutron/γ-ray discrimination capabilities,” Nuclear Instruments and Methods in Physics Research Section A: Accelerators, Spectrometers, Detectors and Associated Equipment, vol. 401, no. 2, pp. 329 – 344, 1997. [14] “Faro vantage laser trackers.” [15] R. H. Showalter, Determination of density and momentum dependence of nuclear sym- metry potentials with asymmetric heavy ion reactions. PhD thesis, Michigan State University, 2015. [16] F.C.E. Teh, “Value-assigned pulse shape discrimination for neutron detectors,” 2020. [17] E. Rutherford, “Retardation of the alpha particle from radium in passing through matter,” Philosophical Magazine, vol. 12, pp. 134–146, 1906. [18] C. J., “The existence of a neutron,” Proc. Roy. Soc. A, vol. 136, pages = 692-708, year = 1932,. [19] W. D. Myers and W. J. Swiatecki, “Nuclear masses and deformations,” Nuclear Physics, vol. 81, no. 1, pp. 1 – 60, 1966. [20] W. D. Myers, “Droplet model nuclear density distributions and single-particle potential wells,” Nuclear Physics A, vol. 145, no. 2, pp. 387 – 400, 1970. [21] M. Bender, P.-H. Heenen, and P.-G. Reinhard, “Self-consistent mean-field models for nuclear structure,” Rev. Mod. Phys., vol. 75, pp. 121–180, Jan 2003. 138 [22] M. Baldo and G. Burgio, “The nuclear symmetry energy,” Progress in Particle and Nuclear Physics, vol. 91, pp. 203 – 258, 2016. [23] B.-A. Li, “Nuclear symmetry energy extracted from laboratory experiments,” Nuclear Physics News, vol. 27, no. 4, pp. 7–11, 2017. [24] J. M. Lattimer and M. Prakash, “Neutron star structure and the equation of state,” The Astrophysical Journal, vol. 550, pp. 426–442, mar 2001. [25] H. Shen, “Complete relativistic equation of state for neutron stars,” Phys. Rev. C, vol. 65, p. 035802, Feb 2002. [26] D. Page, J. M. Lattimer, M. Prakash, and A. W. Steiner, “Minimal cooling of neu- tron stars: A new paradigm,” The Astrophysical Journal Supplement Series, vol. 155, pp. 623–650, dec 2004. [27] A. Steiner, M. Prakash, J. Lattimer, and P. Ellis, “Isospin asymmetry in nuclei and neutron stars,” Physics Reports, vol. 411, no. 6, pp. 325 – 375, 2005. [28] A. W. Steiner, “Neutron star inner crust: Nuclear physics input,” Phys. Rev. C, vol. 77, p. 035805, Mar 2008. [29] B. G. Todd-Rutel and J. Piekarewicz, “Neutron-rich nuclei and neutron stars: A new accurately calibrated interaction for the study of neutron-rich matter,” Phys. Rev. Lett., vol. 95, p. 122501, Sep 2005. [30] J. M. Lattimer and M. Prakash, “The physics of neutron stars,” Science, vol. 304, no. 5670, pp. 536–542, 2004. [31] B. P. Abbott et al, “Gw170817: Observation of gravitational waves from a binary neutron star inspiral,” Phys. Rev. Lett., vol. 119, p. 161101, Oct 2017. [32] B. P. Abbott et al, “Gw170817: Measurements of neutron star radii and equation of state,” Phys. Rev. Lett., vol. 121, p. 161101, Oct 2018. [33] B. P. Abbott et al, “Properties of the binary neutron star merger gw170817,” Phys. Rev. X, vol. 9, p. 011001, Jan 2019. [34] K. Chatziioannou, C.-J. Haster, and A. Zimmerman, “Measuring the neutron star tidal deformability with equation-of-state-independent relations and gravitational waves,” Phys. Rev. D, vol. 97, p. 104036, May 2018. [35] M. Tsang, W. Lynch, P. Danielewicz, and C. Tsang, “Symmetry energy constraints from gw170817 and laboratory experiments,” Physics Letters B, vol. 795, pp. 533 – 536, 2019. 139 [36] C. Tsang, M. Tsang, P. Danielewicz, F. Fattoyev, and W. Lynch, “Insights on skyrme parameters from gw170817,” Physics Letters B, vol. 796, pp. 1 – 5, 2019. [37] P. Danielewicz, P. Singh, and J. Lee, “Symmetry energy iii: Isovector skins,” Nuclear Physics A, vol. 958, pp. 147 – 186, 2017. [38] A. Klimkiewicz, N. Paar, P. Adrich, M. Fallot, K. Boretzky, T. Aumann, D. Cortina- Gil, U. D. Pramanik, T. W. Elze, H. Emling, H. Geissel, M. Hellstr¨om, K. L. Jones, J. V. Kratz, R. Kulessa, C. Nociforo, R. Palit, H. Simon, G. Sur´owka, K. S¨ummerer, D. Vretenar, and W. Walu´s, “Nuclear symmetry energy and neutron skins derived from pygmy dipole resonances,” Phys. Rev. C, vol. 76, p. 051603, Nov 2007. [39] O. Wieland and A. Bracco, “The pygmy dipole resonance in 68ni and the neutron skin,” Progress in Particle and Nuclear Physics, vol. 66, no. 2, pp. 374 – 378, 2011. Particle and Nuclear Astrophysics. [40] A. Carbone, G. Col`o, A. Bracco, L.-G. Cao, P. F. Bortignon, F. Camera, and O. Wieland, “Constraints on the symmetry energy and neutron skins from pygmy resonances in 68Ni and 132Sn,” Phys. Rev. C, vol. 81, p. 041301, Apr 2010. [41] M. Kortelainen, T. Lesinski, J. Mor´e, W. Nazarewicz, J. Sarich, N. Schunck, M. V. Stoitsov, and S. Wild, “Nuclear energy density optimization,” Phys. Rev. C, vol. 82, p. 024313, Aug 2010. [42] B. A. Brown, “Constraints on the skyrme equations of state from properties of doubly magic nuclei,” Phys. Rev. Lett., vol. 111, p. 232502, Dec 2013. [43] D. M. Rossi, P. Adrich, F. Aksouh, H. Alvarez-Pol, T. Aumann, J. Benlliure, M. B¨ohmer, K. Boretzky, E. Casarejos, M. Chartier, A. Chatillon, D. Cortina-Gil, U. Datta Pramanik, H. Emling, O. Ershova, B. Fernandez-Dominguez, H. Geis- sel, M. Gorska, M. Heil, H. T. Johansson, A. Junghans, A. Kelic-Heil, O. Kiselev, A. Klimkiewicz, J. V. Kratz, R. Kr¨ucken, N. Kurz, M. Labiche, T. Le Bleis, R. Lem- mon, Y. A. Litvinov, K. Mahata, P. Maierbeck, A. Movsesyan, T. Nilsson, C. Nociforo, R. Palit, S. Paschalis, R. Plag, R. Reifarth, D. Savran, H. Scheit, H. Simon, K. S¨um- merer, A. Wagner, W. Walu´s, H. Weick, and M. Winkler, “Measurement of the dipole polarizability of the unstable neutron-rich nucleus 68Ni,” Phys. Rev. Lett., vol. 111, p. 242503, Dec 2013. [44] A. Tamii, I. Poltoratska, P. von Neumann-Cosel, Y. Fujita, T. Adachi, C. A. Bertu- lani, J. Carter, M. Dozono, H. Fujita, K. Fujita, K. Hatanaka, D. Ishikawa, M. Itoh, T. Kawabata, Y. Kalmykov, A. M. Krumbholz, E. Litvinova, H. Matsubara, K. Nakan- ishi, R. Neveling, H. Okamura, H. J. Ong, B. ¨Ozel-Tashenov, V. Y. Ponomarev, A. Richter, B. Rubio, H. Sakaguchi, Y. Sakemi, Y. Sasamoto, Y. Shimbara, Y. Shimizu, F. D. Smit, T. Suzuki, Y. Tameshige, J. Wambach, R. Yamada, M. Yosoi, and J. Zeni- 140 hiro, “Complete electric dipole response and the neutron skin in 208Pb,” Phys. Rev. Lett., vol. 107, p. 062502, Aug 2011. [45] Z. Zhang and L.-W. Chen, “Constraining the symmetry energy at subsaturation densi- ties using isotope binding energy difference and neutron skin thickness,” Physics Letters B, vol. 726, no. 1, pp. 234 – 238, 2013. [46] H. S. Xu, M. B. Tsang, T. X. Liu, X. D. Liu, W. G. Lynch, W. P. Tan, A. Vander Molen, G. Verde, A. Wagner, H. F. Xi, C. K. Gelbke, L. Beaulieu, B. Davin, Y. Larochelle, T. Lefort, R. T. de Souza, R. Yanez, V. E. Viola, R. J. Charity, and L. G. Sobotka, “Isospin fractionation in nuclear multifragmentation,” Phys. Rev. Lett., vol. 85, pp. 716– 719, Jul 2000. [47] M. B. Tsang, T. X. Liu, L. Shi, P. Danielewicz, C. K. Gelbke, X. D. Liu, W. G. Lynch, W. P. Tan, G. Verde, A. Wagner, H. S. Xu, W. A. Friedman, L. Beaulieu, B. Davin, R. T. de Souza, Y. Larochelle, T. Lefort, R. Yanez, V. E. Viola, R. J. Charity, and L. G. Sobotka, “Isospin diffusion and the nuclear symmetry energy in heavy ion reactions,” Phys. Rev. Lett., vol. 92, p. 062701, Feb 2004. [48] M. B. Tsang, C. K. Gelbke, X. D. Liu, W. G. Lynch, W. P. Tan, G. Verde, H. S. Xu, W. A. Friedman, R. Donangelo, S. R. Souza, C. B. Das, S. Das Gupta, and D. Zhabinsky, “Isoscaling in statistical models,” Phys. Rev. C, vol. 64, p. 054615, Oct 2001. [49] H. L. Liu, F. R. Xu, S. W. Xu, R. Wyss, and P. M. Walker, “High-spin isomeric structures in exotic odd-odd nuclei: Exploration of the proton drip line and beyond,” Phys. Rev. C, vol. 76, p. 034313, Sep 2007. [50] Z. Y. Sun, M. B. Tsang, W. G. Lynch, G. Verde, F. Amorini, L. Andronenko, M. Andro- nenko, G. Cardella, M. Chatterje, P. Danielewicz, E. De Filippo, P. Dinh, E. Galichet, E. Geraci, H. Hua, E. La Guidara, G. Lanzalone, H. Liu, F. Lu, S. Lukyanov, C. Maiolino, A. Pagano, S. Piantelli, M. Papa, S. Pirrone, G. Politi, F. Porto, F. Rizzo, P. Russotto, D. Santonocito, and Y. X. Zhang, “Isospin diffusion and equilibration for Sn + Sn collisions at e/a = 35 mev,” Phys. Rev. C, vol. 82, p. 051603, Nov 2010. [51] S. F. Ban, J. Li, S. Q. Zhang, H. Y. Jia, J. P. Sang, and J. Meng, “Density dependencies of interaction strengths and their influences on nuclear matter and neutron stars in relativistic mean field theory,” Phys. Rev. C, vol. 69, p. 045805, Apr 2004. [52] B. Liu, V. Greco, V. Baran, M. Colonna, and M. Di Toro, “Asymmetric nuclear matter: The role of the isovector scalar channel,” Phys. Rev. C, vol. 65, p. 045201, Mar 2002. [53] K. A. Brueckner, “Two-body forces and nuclear saturation. iii. details of the structure of the nucleus,” Phys. Rev., vol. 97, pp. 1353–1366, Mar 1955. 141 [54] W. Zuo, A. Lejeune, U. Lombardo, and J.F. Mathiot, “Microscopic three-body force for asymmetric nuclear matter,” Eur. Phys. J. A, vol. 14, no. 4, pp. 469–475, 2002. [55] B.-A. Li, “Constraining the neutron-proton effective mass splitting in neutron-rich matter,” Phys. Rev. C, vol. 69, p. 064602, Jun 2004. [56] A. Lane, “Isobaric spin dependence of the optical potential and quasi-elastic (p, n) reactions,” Nuclear Physics, vol. 35, pp. 676 – 685, 1962. [57] B.-A. Li and X. Han, “Constraining the neutron–proton effective mass splitting using empirical constraints on the density dependence of nuclear symmetry energy around normal density,” Physics Letters B, vol. 727, no. 1, pp. 276 – 281, 2013. [58] J. Dobaczewski, “Structure of nuclei at extreme values of the isospin,” ACTA PHYSICA POLONICA B, vol. 30, p. 1647, 1999. [59] M. Baldo, G. F. Burgio, H.-J. Schulze, and G. Taranto, “Nucleon effective masses within the brueckner-hartree-fock theory: Impact on stellar neutrino emission,” Phys. Rev. C, vol. 89, p. 048801, Apr 2014. [60] O. Sjoberg, “On the Landau effective mass in asymmetric nuclear matter,” Nucl. Phys. A, vol. 265, pp. 511–516, July 1976. [61] W. Zuo, A. Lejeune, U. Lombardo, and J.F. Mathiot, “Microscopic three-body force for asymmetric nuclear matter,” Eur. Phys. J. A, vol. 14, no. 4, pp. 469–475, 2002. [62] F. Hofmann, C. M. Keil, and H. Lenske, “Density dependent hadron field theory for asymmetric nuclear matter and exotic nuclei,” Phys. Rev. C, vol. 64, p. 034314, Aug 2001. [63] P. Danielewicz, R. Lacey, and W. G. Lynch, “Determination of the equation of state of dense matter,” Science, vol. 298, no. 5598, pp. 1592–1596, 2002. [64] B.-A. Li, L.-W. Chen, and C. M. Ko, “Recent progress and new challenges in isospin physics with heavy-ion reactions,” Physics Reports, vol. 464, no. 4, pp. 113 – 281, 2008. [65] Y. Zhang, D. D. S. Coupland, P. Danielewicz, Z. Li, H. Liu, F. Lu, W. G. Lynch, and M. B. Tsang, “Influence of in-medium nn cross sections, symmetry potential, and impact parameter on isospin observables,” Phys. Rev. C, vol. 85, p. 024602, Feb 2012. [66] M. A. Famiano, T. Liu, W. G. Lynch, M. Mocko, A. M. Rogers, M. B. Tsang, M. S. Wallace, R. J. Charity, S. Komarov, D. G. Sarantites, L. G. Sobotka, and G. Verde, “Neutron and proton transverse emission ratio measurements and the density depen- dence of the asymmetry term of the nuclear equation of state,” Phys. Rev. Lett., vol. 97, p. 052701, Aug 2006. 142 [67] M. B. Tsang, Y. Zhang, P. Danielewicz, M. Famiano, Z. Li, W. G. Lynch, and A. W. Steiner, “Constraints on the density dependence of the symmetry energy,” Phys. Rev. Lett., vol. 102, p. 122701, Mar 2009. [68] T. X. Liu, W. G. Lynch, M. B. Tsang, X. D. Liu, R. Shomin, W. P. Tan, G. Verde, A. Wagner, H. F. Xi, H. S. Xu, B. Davin, Y. Larochelle, R. T. d. Souza, R. J. Charity, and L. G. Sobotka, “Isospin diffusion observables in heavy-ion reactions,” Phys. Rev. C, vol. 76, p. 034603, Sep 2007. [69] W. G. Lynch and M. B. Tsang, “The nuclear symmetry energy at sub-saturation den- sities,” 2018. [70] P. Danielewicz, “Determination of the mean-field momentum-dependence using elliptic flow,” Nuclear Physics A, vol. 673, no. 1, pp. 375 – 410, 2000. [71] N. Wang, Z. Li, and X. Wu, “Improved quantum molecular dynamics model and its applications to fusion reaction near barrier,” Phys. Rev. C, vol. 65, p. 064608, May 2002. [72] N. Wang, X. Wu, and Z. Li, “Dynamic study of fusion reactions for 40,48Ca+90,96Zr around the coulomb barrier,” Phys. Rev. C, vol. 67, p. 024604, Feb 2003. [73] N. Wang, Z. Li, X. Wu, J. Tian, Y. Zhang, and M. Liu, “Further development of the improved quantum molecular dynamics model and its application to fusion reactions near the barrier,” Phys. Rev. C, vol. 69, p. 034608, Mar 2004. [74] Douchin, F. and Haensel, P., “A unified equation of state of dense matter and neutron star structure,” Astron. Astrophys., vol. 380, no. 1, pp. 151–167, 2001. [75] J. Bartel, P. Quentin, M. Brack, C. Guet, and H.-B. H ˜A¥kansson, “Towards a better parametrisation of skyrme-like effective forces: A critical study of the skm force,” Nuclear Physics A, vol. 386, no. 1, pp. 79 – 100, 1982. [76] G. F. Knoll, Radiation detection and measurement; 4th ed. New York, NY: Wiley, 2010. [77] M. Wallace, M. Famiano, M.-J. van Goethem, A. Rogers, W. Lynch, J. Clifford, F. De- launay, J. Lee, S. Labostov, M. Mocko, L. Morris, A. Moroni, B. Nett, D. Oostdyk, R. Krishnasamy, M. Tsang, R. de Souza, S. Hudan, L. Sobotka, R. Charity, J. Elson, and G. Engel, “The high resolution array (hira) for rare isotope beam experiments,” Nuclear Instruments and Methods in Physics Research Section A: Accelerators, Spec- trometers, Detectors and Associated Equipment, vol. 583, no. 2, pp. 302 – 312, 2007. [78] B. Davin, R. de Souza, R. Yanez, Y. Larochelle, R. Alfaro, H. Xu, A. Alexander, K. Bastin, L. Beaulieu, J. Dorsett, G. Fleener, L. Gelovani, T. Lefort, J. Poehlman, 143 R. Charity, L. Sobotka, J. Elson, A. Wagner, T. Liu, X. Liu, W. Lynch, L. Morris, R. Shomin, W. Tan, M. Tsang, G. Verde, and J. Yurkon, “Lassa: a large area silicon strip array for isotopic identification of charged particles,” Nuclear Instruments and Methods in Physics Research Section A: Accelerators, Spectrometers, Detectors and Associated Equipment, vol. 473, no. 3, pp. 302 – 318, 2001. [79] E. Pollacco, D. Beaumel, P. Roussel-Chomaz, E. Atkin, P. Baron, J. P. Baronick, E. Becheva, Y. Blumenfeld, A. Boujrad, A. Drouart, F. Druillole, P. Edelbruck, M. Gelin, A. Gillibert, C. Houarner, V. Lapoux, L. Lavergne, G. Leberthe, L. Leter- rier, V. Le Ven, F. Lugiez, L. Nalpas, L. Olivier, B. Paul, B. Raine, A. Richard, M. Rouger, F. Saillant, F. Skaza, M. Tripon, M. Vilmay, E. Wanlin, and M. Wittwer, “Must2: A new generation array for direct reaction studies,” in The 4th International Conference on Exotic Nuclei and Atomic Masses (C. J. Gross, W. Nazarewicz, and K. P. Rykaczewski, eds.), (Berlin, Heidelberg), pp. 287–288, Springer Berlin Heidel- berg, 2005. [80] L. Acosta, E. Pagano, T. Minniti, G. Verde, F. Amorini, A. Anzalone, L. Audi- tore, M. Buscemi, G. Cardella, A. Chbihi, E. De Filippo, L. Francalanza, E. Geraci, S.Gian ˜A¬, C. Guazzoni, E. Guidara, G. Lanzalone, I. Lombardo, S. Nigro, and M. Vig- ilante, “Farcos, a new array for femtoscopy and correlation spectroscopy,” The European Physical Journal Conferences, vol. 31, pp. 1–8, 07 2012. [81] Eljen Technologies, “Ej-301,” 2016. [82] J. Scherzinger, R. A. Jebali, J. Annand, K. Fissum, R. Hall-Wilton, K. Kanaki, M. Lundin, B. Nilsson, H. Perrey, A. Rosborg, and H. Svensson, “The light-yield response of a ne-213 liquid-scintillator detector measured using 2ˆa6mev tagged neu- trons,” Nuclear Instruments and Methods in Physics Research Section A: Accelerators, Spectrometers, Detectors and Associated Equipment, vol. 840, pp. 121 – 127, 2016. [83] “Costruzioni apparecchiature elettroniche nucleari s.p.a..” [84] “Lecroy corporation.” [85] M. Youngs, Using light emitted clusters as a probe of the symmetry energy in the nuclear equation of state. PhD thesis, Michigan State University, 2013. [86] “MSU Electrical and Computer Engineering Department ECE Testing Facility.” [87] Eljen Technologies, “EJ-200,” 2016. [88] R. D. Souza, N. Carlin, Y. Kim, J. Ottarson, L. Phair, D. Bowman, C. Gelbke, W. Gong, W. Lynch, R. Pelak, T. Peterson, G. Poggi, M. Tsang, and H. Xu, “The msu miniball 4¨I fragment detection array,” Nuclear Instruments and Methods in Physics 144 Research Section A: Accelerators, Spectrometers, Detectors and Associated Equipment, vol. 295, no. 1, pp. 109 – 122, 1990. [89] SYLGARD tech, “Sylgard 182 silicone elastomer.” [90] C. Cavata, M. Demoulins, J. Gosset, M.-C. Lemaire, D. L’Hˆote, J. Poitou, and O. Valette, “Determination of the impact parameter in relativistic nucleus-nucleus col- lisions,” Phys. Rev. C, vol. 42, pp. 1760–1763, Oct 1990. [91] T. X. Liu, W. G. Lynch, R. H. Showalter, M. B. Tsang, X. D. Liu, W. P. Tan, M. J. van Goethem, G. Verde, A. Wagner, H. F. Xi, H. S. Xu, M. A. Famiano, R. T. de Souza, V. E. Viola, R. J. Charity, and L. G. Sobotka, “Isospin observables from fragment energy spectra,” Phys. Rev. C, vol. 86, p. 024605, Aug 2012. [92] J. R. Winkelbauer, PRECISION MEASUREMENT OF ISOSPIN DIFFUSION IN PERIPHERAL SN+SN COLLISIONS AT 70 MEV/U. PhD thesis, Michigan State University, 2015. [93] L. Phair, D. Bowman, C. Gelbke, W. Gong, Y. Kim, M. Lisa, W. Lynch, G. Peaslee, R. de Souza, M. Tsang, and F. Zhu, “Impact-parameter filters for 36ar+197au collisions at ea = 50, 80 and 110 mev,” Nuclear Physics A, vol. 548, no. 3, pp. 489 – 509, 1992. [94] 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, “Intermediate mass fragment emission in 36Ar+197au collisions at e/a=35 mev,” Phys. Rev. C, vol. 45, pp. 338–352, Jan 1992. [95] K. Drozdowicz, M. Hoek, and D. Aronsson, “Energy calibration of neutron detectors for the neutron spectrometer tansy,” Nuclear Instruments and Methods in Physics Research Section A: Accelerators, Spectrometers, Detectors and Associated Equipment, vol. 306, no. 1, pp. 315 – 330, 1991. [96] S. Meigo, “Measurements of the response function and the detection efficiency of an ne213 scintillator for neutrons between 20 and 65 mev,” Nuclear Instruments and Meth- ods in Physics Research Section A: Accelerators, Spectrometers, Detectors and Asso- ciated Equipment, vol. 401, no. 2, pp. 365 – 378, 1997. [97] N. Nakao, T. Kurosawa, T. Nakamura, and Y. Uwamino, “Absolute measurements of the response function of an ne213 organic liquid scintillator for the neutron energy range up to 206 mev,” Nuclear Instruments and Methods in Physics Research Section A: Accelerators, Spectrometers, Detectors and Associated Equipment, vol. 463, no. 1, pp. 275 – 287, 2001. [98] W. Betts, F. Bieser, R. Bossingham, M. Botlo, M. Cherney, J. Chrin, P. Colarco, H. Crawford, K. Dao, H. Diaz, D. E. Greiner, L. Greiner, E. L. Hjort, S. Jacobson, R. C. Jared, E. Judd, S. R. Klein, A. N. Lebedev, M. J. LeVine, V. Lindenstruth, M. A. 145 Lisa, K. Marks, C. McParland, T. S. McShane, J. Meier, M. T. Nguyen, D. L. Olson, I. Sakrejda, J. Schambach, R. A. Scheetz, N. T. B. Stone, C. E. Tull, G. Visser, C. Vu, H. Wieman, and E. Yee, “Results from the star tpc system test,” IEEE Transactions on Nuclear Science, vol. 44, pp. 592–597, June 1997. [99] B. C. Rastin, “An accurate measurement of the sea-level muon spectrum within the range 4 to 3000 GeV/c,” Journal of Physics G: Nuclear Physics, vol. 10, pp. 1609–1628, nov 1984. [100] J. Wiechula, “Commissioning and calibration of the alice tpc,” Nuclear Physics A, vol. 830, no. 1, pp. 531c – 534c, 2009. Quark Matter 2009. [101] C. Hagmann, D. Lange, and D. Wright, “Cosmic-ray shower generator (cry) for monte carlo transport codes,” in 2007 IEEE Nuclear Science Symposium Conference Record, vol. 2, pp. 1143–1146, Oct 2007. [102] “Physical constants,” Physics Letters B, vol. 204, pp. IN3 – 127, 1988. [103] E. Siciliano, J. Ely, R. Kouzes, J. Schweppe, D. Strachan, and S. Yokuda, “Energy calibration of gamma spectra in plastic scintillators using compton kinematics,” Nuclear Instruments and Methods in Physics Research Section A: Accelerators, Spectrometers, Detectors and Associated Equipment, vol. 594, no. 2, pp. 232 – 243, 2008. [104] NSCL, GANIL, GSI, and RIKEN, “LISE,” 2003. [105] F. Brooks, “A scintillation counter with neutron and gamma-ray discriminators,” Nu- clear Instruments and Methods, vol. 4, no. 3, pp. 151 – 163, 1959. [106] N. Zaitseva, B. L. Rupert, I. Paweaczak, A. Glenn, H. P. Martinez, L. Carman, M. Faust, N. Cherepy, and S. Payne, “Plastic scintillators with efficient neutron/gamma pulse shape discrimination,” Nuclear Instruments and Methods in Physics Research Section A: Accelerators, Spectrometers, Detectors and Associated Equipment, vol. 668, pp. 88 – 93, 2012. [107] R. Wada, D. Fabris, K. Hagel, G. Nebbia, Y. Lou, M. Gonin, J. B. Natowitz, R. Billerey, B. Cheynis, A. Demeyer, D. Drain, D. Guinet, C. Pastor, L. Vagneron, K. Zaid, J. Alarja, A. Giorni, D. Heuer, C. Morand, B. Viano, C. Mazur, C. Ng, S. Leray, R. Lucas, M. Ribrag, and E. Tomasi, “Temperatures and excitation energies of hot nu- clei in the reactions of 32S+ag and 16O+ag at 30 mev/nucleon,” Phys. Rev. C, vol. 39, pp. 497–515, Feb 1989. [108] R. Wang, Investigating the symmetry energy effect in fission reaction from the emission of the light charged-particle. PhD thesis, Tsinghua University, 2015. 146 [109] P. Landsberg and G. Matsas, “The impossibility of a universal relativistic temperature transformation,” Physica A: Statistical Mechanics and its Applications, vol. 340, no. 1, pp. 92 – 94, 2004. News and Expectations in Thermostatistics. [110] P. T. Landsberg and G. E. Matsas, “Laying the ghost of the relativistic temperature transformation,” Physics Letters A, vol. 223, no. 6, pp. 401 – 403, 1996. [111] K. Zhu, M. Tsang, D. Dell’Aquila, K. Brown, Z. Chajecki, W. Lynch, S. Sweany, F. Teh, C. Tsang, C. Anderson, A. Anthony, J. Barney, J. Crosby, J. Estee, I. Gasparic, G. Jhang, O. Khanal, S. Kodali, J. Manfredi, C. Niu, and R. Wang, “Calibration of large neutron detection arrays using cosmic rays,” Nuclear Instruments and Methods in Physics Research Section A: Accelerators, Spectrometers, Detectors and Associated Equipment, vol. 967, p. 163826, 2020. 147