(n,n'y,) Reactions in 63 ,65 Cu and Background in Ov,/ Experiments by Dennis V. Perepelitsa Submitted to the Department of Physics in partial fulfillment of the Requirements for the Degree of BACHELOR OF SCIENCE at the MASSACHUSETTS INSTITUTE OF TECHNOLOGY June 2008 @2008 DENNIS V. PEREPELITSA All Rights Reserved The author hereby grants to MIT permission to reproduce and to distribute publicly paper and electronic copies of this thesis document in whole or in part. Signature of Author Department of Physics 15 May 2008 Certified by Joseph F. Formaggio Thesis Supervisor, Department of Physics Accepted by Professor David E. Pritchard Senior Thesis Coordinator, Department of Physics MASSACHUSETTS INSTMTUTE OF TEC-lINOLOGY JUN 13 2008 LIBRARIES Acknowledgments Compiling a full list of everybody who has helped me along the last four years at MIT is daunting if not impossible. First and foremost, I would like to thank my thesis advisor Joseph Formaggio for his guidance and support during the writing of this manuscript. I am grateful to the Weak Interactions Team at the Los Alamos National Laboratory for introducing me to experimental science at a national laboratory. Their instruction, mentorship and friendship was pivotal in my education as a scientist. I am indebted to my mentor Steve Elliott, to my friend Vince Guiseppe, to Team Leader Andrew Hime and to Vic Gehman. This one's for you. I would like to thank Matt Devlin, Ron Nelson and other members of LANSCENS for their help and technical instruction during the course of this experiment at LANSCE. This experiment was done in collaboration with Dong-Ming Mei and others at the University of South Dakota, and members of the MAJORANA collaboration. I would like to acknowledge my academic advisor, Peter Shor, and the competent and warm staff in MIT's Physics and Mathematics Departments, for their support during my time at MIT. Along the way, certain professors and individuals have inspired me to add physics as a major and, later, choose experimental physics as an academic career goal. My junior year was crucial to this development. I want to single out Professors Krishna Rajagopal for teaching me 8.05, Isaac Chuang for 8.13, Gerald Sussman for 6.946, Edmund Bertschinger for 8.224, Ulrich Becker for 8.14 and Eric Jonas for mentoring me in an undergraduate research position. I would like to thank the many friends who stuck by me through MIT. I cannot do justice to all of them here. I will always remember the kind words and sometimes stern encouragements of BJP, JTM, CTS and NLH. I am indebted to you beyond my ability to express. You showed me that no man is an island, but that no man need be one, either. Last but most important, I would like to thank my parents MLC and GCC, my brother CVP and my sisters CNC and PAC. You taught me that the only benchmark we must hold ourselves to are our own high standards. Your unwavering belief was a stronger encouragement than grades or personal achievement. With my family, I could climb any mountain. I would like to dedicate this manuscript to them. (n,n'y) Reactions in 63 '65 Cu and Background in OvPP Experiments by Dennis V. Perepelitsa Submitted to the Department of Physics on May 16, 2008, in partial fulfillment of the requirements for the Degree of Bachelor of Science in Physics Abstract Measurements of (n, xn'y) reactions in Cu are important for understanding neutroninduced background for certain underground double beta decay experiments. Neutroninduced gammas are a contribution to background for the next generation of double beta decay experiments, which are designed to reach the sensitivity of the atmospheric neutrino mass scale (45 meV). Measuring and understanding the high-energy neutron excitations of shielding materials such as natCu are crucial for establishing shielding requirements and understanding background. In particular, the regions around the Q-values of candidate OvP// decay isotopes must be investigated. Partial 'y-ray cross sections for a natural copper target were measured using the GEANIE spectrometer in a broad-spectrum neutron beam at LANSCE. The experimental apparatus, and sources of systematic and statistical error are discussed. The results provide useful data for benchmarking Monte Carlo simulation of background events in future experiments. Furthermore, measuring specific (n, n') excited state transitions in this material represents a nuclear structure contribution. Thesis Supervisor: Joseph Formaggio Title: Thesis Supervisor, Department of Physics Contents 1 Introduction 2 Theory of the Experiment 2.1 Background in Neutrinoless Double-beta Decay 2.2 Neutron-Induced Excitations . 2.3 Quantum Scattering ..... 2.4 Nuclear Database Crosschecks .. . . . . 3 Experimental Facility 3.1 GEANIE ............ 3.2 WNR Beamline ........ 3.3 4 Target Runs .......... Calibration and Data Analysis 4.1 Analysis Overview . . . . . . 4.2 Energy and Time Alignment . 4.3 Neutron Flux . . . . . . ... 4.4 Efficiency Calibration . . . . . 4.5 Gamma-ray Attenuation . . . 4.6 4.7 Neutron-induced Copper Spectrum . . . Gamma-ray Yields and Sensitivity Limits 25 27 30 35 38 42 4.8 Constructing the Integral Cross Section . 43 5 Error Analysis 5.1 Approximations and Systematic Error. 5.2 Statistical Error . . . . . . ... 5.3 Error Propagation . . . . . .. 5.4 Future Error Reduction . . . . . 6 24 24 Experimental Results 6.1 6.2 natCu 6.1.1 6.1.2 0v3/ 6.2.1 (n,xn'y) Cross Sections 63 ,65 Cu (n,xn'7) 65 Cu . 65 Cu (n,n'Hy) 65 CU . Regions of Interest . . 76 Ge ........... 44 44 47 48 49 6.2.2 6.2.3 6.2.4 6.2.5 6.2.6 6.2.7 82 Se 1 0Mo 0 116 C d 130 Te 136 Xe 150Nd ..... .......... . . . . . . ................. ................. . .... .................... . . . . . . . . . . . . . .. . . . . . . . . .............. ............. . . . . ... . . . . . 57 58 58 59 60 61 7 Conclusion 63 8 Appendix I: geanie.py Source Code 67 List of Tables 1 2 0O3/3 isotopes and Q-values ............. GEANIE detectors ................... 3 4 5 6 7 8 152Eu 9 ......... ....... and calibration source lines ................... 226 Ra and calibration source lines ................... Coaxial detector attenuation coefficients . ............... Neutron-induced 6 3 '65 Cu Spectrum Identification, 0-1MeV Neutron-induced 63,65 Cu Spectrum Identification, 1-2MeV Neutron-induced 63,65 Cu Spectrum Identification, 2-3MeV Experimental Ov/3 Region of Interest Results . ............ 16 22 .. . . .... ...... ...... .. 32 33 36 39 40 41 55 List of Figures LANSCE Experimental Facility Diagram ...... GEANIE Detector Position Diagram ......... Typical TDC Spectrum . ................ FC TOF: Time Axis .. ................. FC PH: Energy Axis . Neutron Flux Source . ................. ................. . . Radium-226 Calibration Source Spectrum . . . . . . . Europium-152 Calibration Source Spectrum . . . . . . GEANIE Array Absolute Energy Efficiency Fit . . . . Detector Attenuation Correction ............ . Neutron-Induced Natural Copper Spectrum, 3 < E, < 30 MeV Contributions to Error in the 962-keV Cross-Section . . . 63 Calculated and Simulated Cu 670-keV cross-section . . Calculated and Simulated 63 Cu 962-keV cross-section . . 63 Calculated and Simulated Cu 1327-keV cross-section . . Calculated and Simulated 63 Cu 1412-keV cross-section . . 65 Calculated and Simulated Cu 1115-keV cross-section . . Calculated and Simulated 65 Cu 1481-keV cross-section . . 76 Ge 0,33 Q-value ROI and sensitivity limit..... . 76 Ge Ovo,3 Q-value DEP ROI and sensitivity limit. . . 82Se OvP0P Q-value ROI and cross-section . . . . . . . 82Se Ov,/3 Q-value DEP ROI and sensitivity limits. . . 100Mo OvI/3 Q-value ROI and cross-section . . . . . 116 Cd Ov3/3 Q-value ROI and cross-section . ..... 6 11 Cd Ov0/ Q-value DEP ROI and sensitivity limit. . . 13 0 Te 0•O,3 Q-value ROI and cross-section . . . . . 30 ' Te Ov03 Q-value DEP ROI and sensitivity limit. . . 136 Xe OvI3f3 Q-value ROI and sensitivity limit. 1 36 Xe Ov/3/3 Q-value DEP ROI and sensitivity limit. . . 150 Nd Ov/3/3 Q-value ROI and sensitivity limit . . . . . . . 20 21 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26 28 28 29 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34 35 35 37 38 49 52 52 53 53 54 54 56 56 57 57 58 59 59 60 60 61 61 62 1 Introduction Several crucial questions in high-energy physics surround properties of neutrinos. The electron neutrino, postulated by Pauli in 1930 to carry away missing linear momentum and conserve lepton number in beta decay [26], was later confirmed in 1956 [12]. Thought to be massless for decades, recent experiments such as ones at the Sudbury Neutrino Observatory (SNO) [6], Kamioka Liquid Scintillator Anti-Neutrino Detector (KamLAND) [5] and Super-Kamiokande (Super-K) [7] have confirmed the phenomenon of neutrino oscillation, implying that at least two of the neutrino flavors have mass. It is a fascinating and elusive particle: it has a miniscule mass, different mass and flavor eigenstates, interacts only weakly (through the only CP-violating interaction) and presents serious technical challenges in its detection. Importantly for this experiment, it may also be a massive Majorana fermion. Besides trying to measure the neutrino masses and their mixing angles, scientists are interested in questions relating to the chirality of observed neutrinos. It seems that the only observed neutrinos are left-handed and the only anti-neutrinos are righthanded. Because the neutrino is not massless, there are reference frames in which the chirality appears flipped. But since right-handed neutrinos and left-handed antineutrinos are not observed, could this mean that left-handed neutrinos are just righthanded neutrinos in another frame? In other words, could the neutrino be its own antiparticle? In other words, is it a Majorana, instead of a Dirac, fermion? If this is the case, then observing (and later measuring) processes which rely on the neutrino to act as its own antiparticle would confirm the Majorana nature of the neutrino (and help constrain other parameters, like neutrino mass). Some experiments to measure one of these theoretical processes, double-beta decay, are taking data and others are in development. In fact, a recent experimental result from the Heidelberg-Moscow Experiment claims to have observed 0V/3 [13]. Before this (rare) process can be confirmed and measured in future experiments, scientists must undertake an investigation of auxiliary issues such as background. In the present work, we investigate an important background measurement in preparation for these experiments, and construct several nuclear physics database cross-checks. The motivation for this is to supplement the MAJORANA collaboration and other neutrino researchers with necessary data. Data for this undergraduate thesis was taken during a summer internship at the Los Alamos National Laboratory. In effect, this is a nuclear physics measurement taken in the context of neutrino science. Though inspired by members of the MAJORANA collaboration, its results are applicable and important to other neutrinoless double-beta decay experiments as well. 2 2.1 Theory of the Experiment Background in Neutrinoless Double-beta Decay Observation of neutrinoless double-beta decay would show that the neutrino is Majorana in nature. Since the rate of this hypothetical process is related to the effective Majorana mass, measuring it will give an experimental constraint on neutrino parameters. Unfortunately, current experimental limits on the neutrino masses indicate that the next generation of experiments will have to be sensitive to the atmospheric mass scale (- 45 meV) and suppress background by an order of magnitude more than is currently possible. Only a handful of nuclear isotopes have the energetics to be Ov/3 candidates. Scientists will be looking for a specific single-site, two-electron signal with an energy equal to the Ov/ Q-value (e.g. for 76 Ge, this is 2039-keV). Again, as the process is rare, every source of background radiation that falls within this region of interest must be understood. Without a handle on the contribution from background, experimenters cannot claim one way or the other that they have measured this process or put limits on their apparatus sensitivity. There are other powerful backgroundrejection technologies to signal from background, each appropriate for a different set of experiments. For example, multi-site rejection and pulse shape analysis to distinguish gammas from electrons in MAJORANA and the GERDA experiments [9]. However, in solid-state detectors, a double escape peak from a higher-energy gamma can become an electron overlapping the Q-value. Thus, the area of the spectrum 1022-keV above the Q-value is also a region of interest (ROI). To minimize background radiation, experimental sites deep underground are being investigated for their sources of radiation. [21] Although many forms of background radiation are suppressed or lessened under the earth's surface, it is important to quantitatively understand the ones that do exist. Two of these sources of background are thermal and fast neutrons. These neutrons are created by spallation processes caused by high-energy cosmic rays, and can contaminate an experimental site. The free neutrons are likely to interact with material likely to be around the experimental site: enriched germanium in the detectors, natural lead and copper shielding, and other detector and shielding materials. Thus, inelastic neutron scattering measurements are important for the next generation of neutrinoless double-beta decay experiments. [20] By measuring the gammaray production cross-sections of overlapping lines in shielding and detector materials, scientists can provide benchmarks for background simulations and analyses. Though the current effort is related to the MAJORANA experiment, which uses enriched 76 Ge as the double-beta decay candidate, examining the regions of interest for any Q-value is important. Experimental attempts to measure Ov03 in a variety of isotopes are underway in underground laboratories around the world [4]. Some of these isotopes (and their Q-values) taken from [35] are given below in Table 1, updated with recent results from [29], [1] and [30]. Table 1: 0/33 isotopes and Q-values Transition Q-value (keV) 7Ge 82 Se _ 76Se - 1oMo 16-Cd 2995.5(1.9) 86Kr -- 10 2039.00(0.05) Ru 3034.40(.17) 116 SSn 2809(4) 13OTe 1 0 Xe 2530.3(2.0) 136 Xe 1366 Ba 2457.83(.37) 3367.7(2.2) 1oNd -•15 Sm This table lists eight isotopes that could undergo neutrinoless double-beta decay, along with a recent measurement of the Q-value. 2.2 Neutron-Induced Excitations To investigate the potential contaminating background radiation in double-beta decay experiments, data for this experiment was taken using the broad-spectrum neutron beam of a linear accelerator. Therefore, it is important to understand how neutrons over a wide energy range hitting a target cause gamma radiation. In this experiment, we chose to study (n, xn'T) reactions. Specifically, reactions activated by an incident neutron which produced one more (x) neutrons and one or more photons. In x = 1 reactions, the neutron excites the nucleus and leaves the target area while the nucleus decays. In x > 1 reactions, the neutron is so energetic that it knocks out additional neutrons, leaving the nucleus an isotope with lower A number in an excited state. Of course, other interactions take place (for example, an 63 (n, py) reaction on 63 Cu would convert this isotope to an excited state of Zn, and a (n, n'c-y) reaction on 65 Cu would convert this isotope to an excited state of 60 Co) but (n, xn'l) interactions are very likely to happen and do not change the identity of the decaying element. Instead of decaying to a lower-state by gamma emission, an excited nucleus can decay by releasing a conversion electron and a characteristic X-ray. When measuring gamma ray production cross sections, the possibility of missed events because of internal conversion should be considered. The TOI 2003 [17] described how likely this was for a given transition. Additionally, a high-energy incident neutron is likely to impart much of its kinetic energy to the nucleus and excite it to a high level. Rather than decay immediately to the ground state (which is sometimes not feasible due to angular momentum conservation or other reasons), the nucleus will often give off a gamma cascade and transition in steps down to the ground state. For example, an observed (les -4 gs) event could have come from the nucleus excited to the first excited state decaying, or part of a cascade from a much higher level. The relationship between a high-energy excited state and probable gamma cascades can be reconstructed if the branching ratios from each level are known. In principle, we could choose to measure the level cross sections for a given excited state (that is, how likely is a neutron to put the nucleus into a certain state). In practice, however, we are interested in the gamma ray production cross sections for a given transition. One of the reasons is that we are interested in the end result of what gamma radiation is produced by a neutron and not the exact decay schemes by which it is produced. Finally, to correlate which decay events are activated by which incident neutrons, the gamma ray production must be prompt. That is, the decay should occur within the timing resolution of the experimental apparatus. In practice, this meant that the half-life of any excited state reached should be under 15ns to give meaningful results about how this depends on neutron energy. 2.3 Quantum Scattering In a decay event, photons carry away units of angular momentum. Since this is a conserved quantity, given the angular momentum quantum numbers of the initial and final state, and current spin of the nucleus, the gamma will tend to head in some directions over others. In the normal decay of a radioisotope with nuclear spin, the nuclear spins of the sample point in all directions with equal probability, so any angular dependence in the decay event is averaged out. Thus, the cross-section of a transition could display some angular distribution u(0, 4). In our setup, the azimuthal 0-symmetry is not broken, but the presence of the neutron beam breaks the 0-symmetry in (n,n'y) reactions. It is known that when a sample is placed in the flight path of an active neutron beam, the nuclear spins tend to point in the plane anti-parallel to the direction of the beam. Any azimuthal dependence in the cross-section will be averaged out, but there is some theoretical dependence on 0, the angle of incidence with the beam. Quantum scattering theory says that the differential cross-section can be expressed as the weighted sum of even Legendre polynomials in sin 0 [10]: do(0) dO 4r 4w aP,(sinO) (1) n=0,2,4 Above, o0 is the integral cross section, which can be obtained by integrating the differential cross-section over the range of sin 0. Roughly, the weights of the different polynomials measure the "anisotropy" of the event. ao measures how isotropic the decay is, a2 measures any overall dipole moment, a4 measures any overall quadrupole moment, etc. Since the sense of absolute scale is set by o0 ,ao - 1 by definition, and the weights of the other polynomials are expressed as unitless ratios to the weight of the isotropic term. A key mathematical property of even Legendre polynomials is that for every n > 2, they integrate out to zero. This means that even though there may be a contribution from the quadrupole at a given angle 0, the net addition to the integral cross-section over all angles will be zero. In practice, since no apparatus can sample continuously at all angles, approximations are used to connect the differential cross section sampled at certain angles to the integral cross section. Some researchers reason that the contribution from a 4 and higher terms is negligible and sample at an angle which is a zero of P2 to eliminate the a2 term and measure ao directly [32]. Another possibility is to find optimal sampling angles 0 and then use numerical quadrature to estimate ao from measurements of d [22]. Still, other researchers measure the cross-section as if their array had complete coverage, calculate the theoretical paramters an and then correct their measurement for this. Although GEANIE allows the experimenter to sample the differential cross-section at multiple values of sin 0, in this paper, we circumvent a full treatment of the angular problem by treating our detector as having close to complete coverage. 2.4 Nuclear Database Crosschecks Finally, data obtained from the experiment is useful for measuring other quantities directly comparable with the literature. Of the published (n, xn'y) data, cross-sections against neutron energy tend to be well-measured only for specific transitions, or in common isotopes, or at single energy values, or with low statistics or with significantly small amounts of angular coverage. Constructing cross-section measurements at this apparatus serves two purposes. The comparison with established results gives us confidence in results using the same analysis process or provides an overall normalization for the same data. The other purpose is to contribute not well-measured results to the nuclear physics community. In addition, nuclear reaction simulation packages such as TALYS [15] or GNASH [34] are usually only accurate in modeling prominent or low-level transitions. Nuclear (n, xn') data is useful for providing benchmarks for further development. In addition to cross-sections, other calculated values such as branching ratios, level-production cross-sections and other measurements can be. 3 Experimental Facility Data for this experiment was taken at the linear accelerator at the Los Alamos Neutron Science CEnter (LANSCE), a major experimental science facility at Los Alamos National Laboratory. Some references which describe these facilities in more detail are given below. Similar experiments have been taking place at this facility for over a decade, and the specific apparatus used in this experiment is well-documented. The LASNCE experimental facility is shown in Figure 1. The LANSCE linac is an 800 MeV proton beam, which is converted using a spallation target to a neutron beam in the Weapons Neutron Research (WNR) Facility, and the 200 beam flight path was used in this experiment. The flight path runs through GEANIE, a high- resolution gamma ray spectrometer used in previous (n, xn'^y) reaction measurements, fission studies and experiments concerning nuclear spectroscopy, nuclear reactions and nuclear structure. GEANIE is operated by LANSCE-NS, the "Neutron and Nuclear Science" group of Los Alamos National Laboratory. Established literature describes the operation and calibration of the experimental facility, including addressing the angular distribution problem and performing benchmarking measurements of the 5"Fe (n,xn'y) 846-keV 2+ --+ gs line [19]. Another excellent reference that discusses GEANIE, the WNR and calculates 2 38 u (n,xZ7Y) production cross sections is [8]. A more involved description of how the fission chamber functions is given in [33]. During the summer of 2007, I visited the WNR facility at least once a week to actively oversee the experiment while data was being taken. I I Figure 1: LANSCE Experimental Facility Diagram This is a diagram showing the proton beam, spallation target, neutron beam flight path, collimation, fission chamber, detector array and lead beam stop. Adapted from [8]. 3.1 GEANIE The detector array used in this experiment is the GErmanium Array for NeutronInduced Excitations (GEANIE). GEANIE normally consists of 26 high-purity, highresolution germanium detectors. Sixteen of these detectors are coaxial with an energy range up to roughly 4 MeV (8000 channels, 2 channels per keV), and ten are planar with an energy range up to roughly 1 MeV (8000 channels, 8 channels per keV). Since most of the examined gamma ray energies of interest were greater than 1 MeV, the planar detectors were seldom used. For the rest of this analysis, we focus on the coaxial detectors. 6, 12, 15, 24 ~EZIZIXIZZIZI-zz~ 14, 17. 22 7, 16, 19, 25 Figure 2: GEANIE Detector Position Diagram The figure on the left is a top-down view showing the values of 0 that correspond to active coaxial detectors. The figure on the right is a side-view showing the three halos (q = -29', 00, 290) which have active coaxial detectors. The detector array was arranged in a set of four "halos" (mounted rings). Each ring is defined by a value of q, which is the angle that the cone between the target and the ring makes with the plane parallel to the ground. Historically, the first three halos (at q = -29", 00, +290) were installed first and contain seven, six and seven detectors at equally spaced locations along the ring. The fourth halo at q = +550 consisting of six detectors was installed later. Since all of the detectors in this last halo were planar detectors, only detectors from the first three halos were primarily used in this analysis. Along each ring, a detector position is also defined by 0, the angle of incidence with the vertical plane of the beam, with 0 = 0 corresponding to directly behind the target. Looking down on the array from above, the positive values of theta go counter-clockwise. Det # 1 2 Type P1 P1 P1 P1 P1 Activre? Table 2: GEAN\JIE detectors 0 0 Distance cos Ob Comments 29.00 -152.8' 14.415 cm 0.778 Yes Yess -29.00 -154.00 14.442 cm 0.786 Yes s 29.00 3 157.00 14.379 cm 0.805 4 Yess -29.00 157.90 14.318 cm 0.810 Yess 56.50 27.00 5 16.535 cm -0.492 Yess 6 29.00 102.00 14.237 cm 0.182 Cx 7 Cx Yes -29.00 102.50 14.288 cm 0.189 P1 Yess 8 55.00 77.00 18 cm -0.129 0 P1 No 9 55.00 -129.0' 18 cm 0.361 No data P1 No 10 55.00 -77.00 0 18 cm -0.129 Poor statistics 0 11 No Cx 26.50 14.308 cm -0.895 Poor resolution 0.00 1.00 Yess 29.00 12 Cx 14.379 cm -0.874 Cx No -29.00 1.20 13 14.773 cm -0.874 Low statistics Yess 14 Cx 0.00 -25.20 14.392 cm -0.905 Yess 29.00 -51.10 14.392 cm -0.549 15 Cx Yess Cx -29.00 -51.00 13.846 cm -0.550 16 Yess 17 Cx 0.00 -76.90 14.442 cm -0.227 Yess P1 56.50 -25.00 16.378 cm -0.500 18 Yess -29.00 -102.0 0 14.308 cm 0.182 19 Cx P1 No 20 0.00 -128.0 14.161 cm 0.616 Unstable gain No 55.50 129.0c 17.089 cm 0.356 21 Cx No data Yess 22 14.917 cm -0.199 Cx 0.00 78.50 Cx Nc 23 0.00 129.5c 14.237 cm 0.636 Poor statistics 0 Yess 29.00 -101.7 24 14.176 cm 0.177 Cx Ye•s Cx -29.00 53.50 14.435 cm -0.520 25 Nc 29.00 53.00 Cx 26 14.455 cm -0.526 No data This table lists each of the 26 detectors and indicates if it is planar or coaxial, whether or not it was included in the final analysis, its position (0, phi), its distance to the detector, which values of cos Ob it is sampling and any reasons for it not being included in the analysis. These values of (0, 0) are different from the angle of incidence with the beam 0 b, with the conversion given by cos Ob = - cos 0 cos 0. For various timing-, resolution- or gain-related reasons, some detectors were not included in the analysis. Additionally, a few offline detectors had their data acquisition channels used in other, unrelated experiments. Table 2 lists the detectors, their type (coaxial or planar), reference angles, distance to target and whether or not they were used in this analysis (along with the reason if the detector is challenged). 3.2 WNR Beamline To create a broad-spectrum neutron beam, the monoenergetic 800 MeV proton beam is run through an unmoderated natW spallation source, resulting in neutrons with energies ranging from about 0.1 MeV to 600 MeV. The beam delivers between 2 and 4 yA of current with a specific time structure. The beam is composed of 40 Hz macropulses 625 ps wide, each of which is composed of micropulses spaced 1.8 Ps apart. The beam area on target was trimmed with lead collimaters after the spallation source but before the fission chamber and detector array. For this experiment, the area of the beam on target was reduced to 1.9 cm radius with lead collimators. This ensures that the beam area would fall entirely within the copper target. The GEANIE flight path runs through a fission chamber 18.495 m after the spallation target, consisting of 235 U and 23 8 U foils, at which fission reactions ((n, f), with fission products f) are measured by the data acquisition system. Time of flight (TOF) information, along with a pulse height proportional to the number of fission events, was stored during each run. The center of the array, where the target sits, is 184.5cm after the fission chamber. Time-of-flight analysis was used to reconstruct neutron energies. A sharp gamma flash at the beginning of the TOF spectrum signified the arrival of gammas from the spallation chamber, followed by the fastest neutrons. Since the speed of light, the distance between chamber and spallation source, and the neutron mass is known, the velocity of any incident neutron is related to its energy by E, = Eo/V1 - v2/ 2 , where Eo = 939.6MeV is the neutron rest mass. The time resolution of the data acquisition system is 15ns. For a given neutron energy, the pulse height information in the fission chamber can later be combined with the known fission cross-sections of isotopic uranium to give a calculation of neutron flux. 3.3 Target Runs In the summer of 2007, runs were taken with a natural copper target between 6/18/08 and 7/1/08. The target was composed of three half-millimeter sheets measuring four square inches in area, which covered all of the beamspot. natCu is 69.15% 6 3 Cu (62.93 amu) and 30.85% 65 Cu (64.93 amu). At room temperature, natural copper has a density of 8.96 grams per cubic centimeter. To convert recorded events to a cross section, the thickness of a copper isotope seen by an incident neutron needs to be determined. The thickness in atoms per barn of the target is given by the following calculation: (8.96 g/cm3 ) x (1.5 mm thick target) x (10-24 cm 2 /barn) atoms (6.022 x 1023 g/amu) x (.6915 x 62.93 + .3085 x 64.93 amu/atom) barn Two efficiency calibration runs using known-activity were performed, during which the beam was turned off. An efficiency calibration run with 152 Eu was performed on 7/3/08 and one with 226 Ra from 7/3/08 to 7/5/08. Both copper isotopes are stable, so there should be no background caused by radioactive decay of the unactivated target. Still, to investigate sources of background in the experiment that are not from (n, n') reactions on natCu, data was taken with the neutron beam turned on and a blank capsule as the target. 4 Calibration and Data Analysis The data acquisition system at LANSCE has the interesting property that it records information from both the fission chamber and the array spectrometer. Interpreting the experimental results requires an understanding of the beam structure and format of the information output by the apparatus. While data is being taken, individual events detected by the system are written to disk for later off-line processing and analysis. 4.1 Analysis Overview Data analysis for this experiment was done using a number of pieces of legacy software written by GEANIE collaborators at LANL, in addition to a software suite developed by the author for this experiment. To unpack the event files written to disk into a more accessible format, code from the internal tscan analysis package was used. This is the step at which time and energy gain-alignment occured and individual detectors and information channels were turned on or off. Then, channels from the resulting data could be extracted using the rgmt code from the tscan package. This includes neutron flux data, calibration spectra and livetime spectra. The excite program was used to obtain neutron-induced gamma spectra gated on netron energy bins. To analyze the yields for a given gamma peak yield seen by a given detector in a given neutron energy bin, the gf3 program from the RADWARE gamma-ray analysis software was used [28]. After the yield, livetime, neutron flux and calibration information was extracted, the author synthesized this data with a python module, called geanie.py, written for this purpose, heavily relying on the popular pylab [18] and scipy [31] Python packages. The commented source code is attached in the appendix at the end of this manuscript. geanie.py is a library that defines useful functions for analysis. It is imported at the start of any script which performs data analysis. At various stages of the analysis bash and FORTRAN scripting were also used. To construct the integral cross-section a for a given transition -y at a neutron energy E,,several measurements have been synthesized in the expression: or = N -I(y, En)/c(Y) t 4(En)" F(-) (3) Equation 3 is the focus of this experiment. I is the corrected yield for the transition, E is the calibrated efficiency, ( is the corrected neutron flux, F is an attenuation correction and N is an overall normalization. In the sections that follow, we describe each of these in detail, as well as discuss errors that arise from each of its terms and alternate formulations of it. We return to the equation in Section 4.8. 4.2 Energy and Time Alignment After the event file is processed, two spectra are compiled for each detector. The first of these is the 8192-channel ADC energy spectrum, in which events are sorted by energy with a gain depending on whether the detector is planar or coaxial. In addition to these ADC spectra, the data acquisition system records events measured while the beam is off, but supresses them by a factor of 8, which will be important later while performing an efficiency calibration. Similarly, the events are sorted by time in a 8192-channel "TDC" time spectrum with a gain of 4.0 ns / channel (eight times as compressed as other time of flight spectra), with each macropulse being written from left to right (since the fastest neutrons arrive first, we later reverse the spectrum so that the lower energy ones are on the left). Of course, the macropulse has several micropulses as a substructure. Thus, the TDC spectrum looks like many reversed time-of-flight spectra put together side by side. Both types of spectra need to be calibrated. The energy spectrum for each detector was aligned using a pair of known transitions as reference lines for a linear calibration. For the coaxial detectors, a gain of 0.5 keV/channel was used with channel zero coinciding with 0 keV by calibrating the 670-keV and 962-keV transitions in 63 Cu to channels 1339 and 1924, respectively. Planar detectors were aligned using the same lines with a gain of 0.125 keV/channel to channels 5357 and 7696.5, respectively. In general, once calibrated individual detectors drifted only a few channels at most during the time that data was being taken. To combine the TDC spectra, the width between the individual micropulses was measured to be 447 channels (1.788 ps using the TDC compression), and these individual spectra were added together to give a single time-of-flight spectrum compiled from all the micropulses. A TDC spectrum before this operation is shown in Figure 3. The sharp peak, caused by prompt gammas from reactions in the spallation target, must be aligned in each detector. The offset for each detector was calibrated so that the tip of the gamma peak falls at channel 100. ouu I II TDC Counts I I'-• TDC'Counts I 500 400 SIA -C o 200 U 100 9 LYJ' ~ 4 600 YF·*LYR 4500 PI1·UI· 5000 1'IYI1~ 5500 I ~ri·~rC1 6000 I ~liY 6500 iLl ýý I ~~ I~i~iT 7000 ~rL·IY- I~ 7500 ·n 8000 Time channel Figure 3: Typical TDC Spectrum This is a typical time spectrum seen by a detector. Each peak is a "gamma-flash" which ends a micropulse and signifies the arrival of gammas from the spallation target. Since the fastest neutrons arrive first, the low-energy part of each micropulse is on the right. Once the energy and time axes are calibrated, the next stage of the analysis is to gate the ADC spectra on the information in the TDC spectra, to give a neutroninduced energy spectrum for a given neutron energy bin. Since the distance between the fission chamber and the array is known, time of flight analysis can be used to produce gamma ray spectra gated on neutron energy. This is the basis for future high-level analysis, and extracting peak yields from gated spectra is discussed in Section 4.7. 4.3 Neutron Flux After the event files are processed, information from both fission chambers are stored in separate matrices for processing into a neutron flux. For neutron energies up to 2 - 3 MeV, the 235 U foil is used because the 23 8 U (n, f) cross-section is small and the results unreliable. For higher neutron energies, the neutron flux is reconstructed from the 238 U fission chamber instead. Since many of the regions of interest in this experiment are higher than 2 MeV, the 23 8U fission chamber is used almost exclusively. Fission chamber events are stored in a two-dimensional matrix sorted by time of arrival and energy. Both axes must then be calibrated and processed. When summed along the energy axis, the result is a time of flight (TOF) spectrum, which shows the time distribution of all fission chamber events. An example is shown in Figure 4. A small y-flash to the left of the large peak is caused by photons from the spallation target arriving, and signals the start of the micropulse to the data acquisition system. In the figure below, it occurs at channel 118. 0 oE 0 Time Axis (channels) Figure 4: FC TOF: Time Axis Each count is actually a fission event. This time of flight spectrum shows the temporal distribution of fission events over the time span of a micropulse. 1200 o 1000 E 800 U ' 600 E - 400 200 U Energy Axis (channels) Figure 5: FC PH: Energy Axis Each count is actually a fission event. This energy spectrum shows the energy distribution of measured fission events. The lower peak is from a-detection. When summed along the time axis, the result is a pulse height (PH) spectrum, which shows the energy distribution of all events. An example is shown in Figure 5. The small peak on the left comes from a-particle detection, which must be excluded in the analysis, and the peak on the right comes from (n, f) events. To make a corrected TOF spectrum which includes only counts from fission events, the left peak is excluded with a cut. More detail is given in [33]. In the figure above, it was made at channel 237. The matrix is folded along the energy axis from the a-cut to the high end of the energy spectrum. Using the distance between the spallation target and the fission chamber, the time of arrival gives a measure of neutron energy as described above in Section 3.2. Then, using experimentally measured 235,238 U fission cross-sections from the literature, the number of counts, as well as the foil thickness, is turned into a neutron flux. Since the apparatus resolution is 15ns, the flux for a given neutron energy must be binned into an energy region that corresponds to that time bin. Because of the functional form of energy depending on velocity, a 15ns time bin corresponds to a larger energy bin at higher neutron energies than lower one. A measurement of neutron flux derived in this manner is shown in Figure 6. a) C 0 L. C x ULL Neutron energy (MeV) Figure 6: Neutron Flux Source This shows the neutron flux on target as measured by the during the natural copper runs. 238U fission chamber Due to limitations in the data acquisition system, not every measured fission event is recorded in more detail. Every event detected by the system but not written in detail to disk is recorded in a special scaler spectrum. By summing this spectrum and comparing it with the number of recorded events, an apparatus livetime L can be derived. With this in mind, the flux through a given neutron bin can be calculated in the following manner: I (E,) = (Ets (counts/MeV) LB(E,) (4) In equation 4, Icts is the number of fission counts reported by the analysis software for a particular neutron energy bin E, and L is the fractional livetime for the fission chamber. To normalize the results per MeV, we divide by the size of the neutron energy bin B(E,). 4.4 Efficiency Calibration Two known-activity radioactive sources were used to calibrate the detector array. A sealed 226 Ra source in secular equilibrium with some of its shorter-lived daughter products provided for useful calibration points as high as 2.5MeV. The gamma emitters in this chain were used as calibration point sources. This source was calibrated to 1.86 x 106 Bq (decays/second) on 1/2/02. The part of the decay chain in secular equilibrium is 226Ra - a -222 Rn,--- 218 Po -- a 214 Pb -t3 -214 Bi (5) A 152Eu source with two decay modes (5 2 Eu --+ 1 2Sm by electron capture with a branching ratio of .7210, 152Eu --- 152Gd by beta decay with a branching ratio of .2790) provided for calibration points as low as 186-keV. This source was calibrated in 1979 and calculated to have an activity of 1.49 x 105 Bq on 7/13/07. The decay energies and conversion coefficients were taken from [17] and the branching ratios for all isotopes were taken from [11] and [16]. The exact calibration lines and branching ratios used in this calibration are shown in Tables 4 and 3. Each active detector was calibrated separately, with the planar detectors only using calibration lines up to 1MeV, to investigate the spread in the absolute efficiency between the detectors. In addition, an efficiency calibration was derived treating the sum of the detector spectra as a single spectrometer. Though we used the efficiency of the array in our final calculations, the individual detector efficencies are needed in a later analysis of systematic error. The following formula was used to derive the efficiency e of an individual detector for an incident gamma ray of energy y: c(-) = 47r- r-t-L (6) Above, Iy is the measured intensity of the line, fit with gf3 (this piece of software is introduced in Section 4.1 and discussed in detail in Section 4.7) from the calibration spectra, in counts, r is the rate in decays per second, t is the runtime in seconds, and L is the data acquisition system livetime. This is described in more detail below. The major statistical uncertainty in the yield is a function of the intensity of background around the calibration peaks. We did not estimate uncertainty in radioactivity rate or recorded runtime, and the uncertainty reported in determining the livetime was not significant. Due to limitations in the data acquisition system, not every detector energy deposition is recorded in more detail. Every event detected by the system but not written in detail to disk is recorded in a special scaler spectrum. By summing this spectrum and comparing it with the number of recorded events, a livetime L for this specific detector can be derived. For the calibration runs, the runtime t is not simply equal to the logbook runtime. During the 226 Ra and 152Eu calibration runs, the neutron beam was off. Instead of the normal macro- and micropulse information, an artificial electronic system was fed into the data acquisition system. A 37.9MHz pulser took the place of the micropulse structure of the beam. When the pulsar is on, the daq records all events to the "beam on" matrix, and supresses all "beam off" events with a 1:8 ratio. Thus, the actual runtime must be adjusted for this. Table 3: 152Eu and calibration source lines Energy (keV) 121.7817 (3) 244.6975(8) 295.9392(17) 344.2785(12) Decay Branching ratio (%) 152Eu _- 152Sm 28.41(13) 5 2 1 Eu 152 GSm 7.55(4) 152 Eu 152SSm 152 Eu 152 Gd 367.7887(16) 152Eu 152 Gd 411.1163(11) 152 Eu 152Gd 2.237(10) 443.96(4) 152 Eu 152 Sm 3.125(14) 152Eu 152 Sm 0.407(?) 688.670(5) 152Eu 152 Sm 0.834(?) 778.9040(18) 152Eu 152Gd 152EEu 152Sm 12.96(6) 4.241(23) 152Sm 14.62(6) 152Sm 0.647(?) 152Sm 13.40(6) 152Sm 1.415(9) 152Gd 1.632(9) 152Sm 20.85(9) 488.6792(20) 867.378(4) 964.079(18) 1005.272(17) 1112.074(4) 1212.948(11) 1299.140(10) 1408.006(3) 152Eu 152Eu 152Eu 152EU 15 2 Eu - 0.440(?) 26.58(12) 0.845(?) Table 4: 226 Ra and calibration source lines Energy (keV) Parent 186.211(13) 241.997(3) 295.224(2) 351.932(2) 609.312(7) 665.453(22) 226Ra 768.356(10) 786.1(4) 806.174(18) 934.061(12) 964.08(3) 1120.287(10) 214Bi 1155.19(2) 1238.110(12) 1280.96(2) 1377.669(12) 1385.31(3) 1401.50(4) 1407.98(4) 1509.228(15) 1583.22(4) 1661.28(6) 1729.595(15) 1764.494(14) 1847.420(25) 2118.55(3) 214Bi 2204.21(4) 2293.40(12) 2447.86(10) 2694.7(2) 2769.9(2) 214Bi 214pb 214pb 214pb 214Bi 214Bi 214Bi 214Bi 214Bi 214Bi 214Bi 214Bi 214Bi 214Bi 214Bi 214Bi 214Bi 214Bi 214Bi 214Bi 214Bi 214Bi 214Bi 214Bi 214Bi 214Bi 214Bi 214Bi Branching ratio (%) 3.56(6) 7.43(11) 19.3(2) 37.6(4) 46.1(5) 1.46(3) 4.94(6) 0.31(9) 1.22(2) 3.03(4) 0.362(17) 15.1(2) 1.63(2) 5.79(8) 1.43(2) 4.00(6) 0.757(18) 1.27(2) 2.15(5) 2.11(4) 0.690(15) 1.15(3) 2.92(4) 15.4(2) 2.11(3) 1.14(3) 5.08(4) 0.305(9) 1.57(2) 0.031(2) 0.025(2) Treating the array as a single spectrometer, an efficiency calibration curve was fit to either set of data points. The fit was performed in log-energy, log-efficiency space using a sixth-order polynomial and uncertainty in each data point as uncertainty for the fit. As suggested by [14] and [3], a high-order polynomial is the standard method for calibrating solid-state Ge detectors, using a high-enough order to fully specify the shape of the efficiency curve. For the 152Eu calibration source, these fits had ten degrees of freedom with a typical reduced x 2 value of 60/29. For the 22 6Ra calibration source, these fits had twenty-four degrees of freedom with a typical reduced x 2 value of 82/20. In this case, the 47r factor on equation 6 was not included, since, by assumption, GEANIE had close to total angular coverage. The uncertainties in the efficiency curve polynomial coefficients were given by the least-squares fit parameters. 0o U U 500 1000 1500 2000 Energy (keV) Figure 7: Radium-226 Calibration Source Spectrum 2500 U 4UU ZO IUUU bUU bUU LUU i4UU ItUU Energy (keV) Figure 8: Europium-152 Calibration Source Spectrum efficiency calibration curve I{I · · 1000 1500 -3.0 -_Q.• 1 0 500 2000 2500 gamma energy (kev) Figure 9: GEANIE Array Absolute Energy Efficiency Fit 4.5 Gamma-ray Attenuation A correction had to be made for the attenuation of excited gamma rays travelling through the target material. Since the target was millimeters wide with the area of a few centimeters, the effect was more pronounced in detectors facing the "edges" of the target. An estimation of the gamma-ray attenuation was obtained by calculating the attenuation resulting from a gamma-ray being created at any depth within the target, and then integrating along this path. We call the median distance seen by created gamma-rays the "effective thickness" seen by that detector. The intensity I (as a ratio with the original intensity lo) of gamma radiation of energy E that remains after travelling a distance t through a material with mass attenuation coefficient p/p and density p is given by the National Institute of Standards and Technology [24]: I A -= exp(-- p-t) (7) o0 P To approximate the effect of the attenuation if the gamma ray is created anywhere along the path, we integrate equation 7 above from t = 0 to t = 2x, where x is the effective thickness. The result is I 10 1 - exp(-2 • -p -x) 0P 2. p P x (8) In general, p/p varies with energy. NIST also keeps a list of p/p coefficients for natural elements at certain energies E [23], measured in cm 2/g A linear extrapolation was used to approximate the coefficient for intermediate energies. For the copper target used in this experiment, p = 8.96 g/cm3 . The distance x seen by an individual detector can be calculated from its (0, ¢) coordinates. Table 5: Coaxial detector attenuation coefficients Detector # Distance thru target Detector 6 Detector 7 Detector 12 Detector 14 Detector 15 Detector 16 Detector 17 Detector 19 Detector 22 Detector 24 Detector 25 4.1mm 4.0mm 0.9mm 0.8mm 1.4mm 1.4mm 3.3mm 4.1mm 3.8mm 4.2mm 1.4mm Ey = 500-keV E, = 1-MeV E = 2-MeV p/p = .08362 .746 .754 .938 .940 .904 .905 p/p = .05901 .810 .817 .956 .957 .931 .931 .844 .810 .825 -.806 .928 p/p = .04205 .860 .865 .968 .969 .950 .950 .885 .860 .871 .856 .948 .788 .746 .764 .741 .899 This table gives the effective distance for each active coaxial detector and, using equation 8, the attenuation coefficient calculated for that detector at the given gamma-ray energy. p/p is in units of cm 2 /g, and the attenuation coefficients are unitless. I-- 0 U C a) 1-i 0 4-J r( energy (keV) Figure 10: Detector Attenuation Correction Two attenuation coefficient curves. The y-axis unitless. The higher curve is from a detector which sees a relatively low effective thickness, and the lower curve is from a detector which sees a relatively high effective thickness. For high-energy gamma lines (regardless of detector orientation) or for detectors facing the flat side of the target (regardless of energy), the attenuation I/Io was on the order of > 80%. In other cases, it could be dramatically smaller. These individual attenuation calculations had to be synthesized into a single correction for the entire array F. Under the assumption that the absolute efficiencies of each detector are not significantly different, we take the mean of the attenuation coefficients for each detector to arrive at one for the array. Table 5 gives values of p/p for a number of prominent energy values, the effective thickness of each detector (half the maximum path length), and the attenuation coefficient at that detector at that gamma ray energy. Figure 10 shows attenuation coefficients for detector 6 (positioned almost perpendicular to the sample) and detector 12 (positioned almost in front of the target). 4.6 Neutron-induced Copper Spectrum Figure 11 shows a semi-log plot of the induced copper spectrum seen by the detector array for a wide range of neutron energies. The spectrum is feature-rich, and many prominent transitions that are neutron-induced 63 '65 Cu lines are listed in Tables 6, 7 and 8. We have listed the rough peak intensity over this neutron energy range, uncorrected for efficiency to give a qualitative result for their strength. Lf .4-. C 0 U 0 500 1000 1500 2000 2500 3000 3500 Energy (keV) Figure 11: Neutron-Induced Natural Copper Spectrum, 3 < E, < 30 MeV The energy resolution of the apparatus, obtained by taking the full-width at halfmaximum of gamma peaks in the energy region, is two keV for every MeV, which is a fractional resolution of 0.2%. This applies over the entire detectable energy range. For individual detectors, however, this resolution can be higher or lower. For solidstate, enriched germanium detector of this type, 0.2% is a good result. We use this expected full-width at half-maximum when investigating apparatus sensitivities to peaks hidden by background in Section 4.7. Table 6: Neutron-induced Energy (keV) 255.0(13) 312.4(6) 365.2(4) 414.3(4) Counts 4 10 104 439.7(7) 449.93(5) 105 105 105 105 469.2(4) / 471.0(3) 10 4 499.7(21) me = 510.998 104 106 533.8(6) 584.82(15) 609.5(1) 612.7(8) 10 4 624.3(3) / 625.6(3) 645.4(3) 669.62(5) 685.6(6) / 686.3(2) 694.3(2) 742.25(10) 754.8(8) 765.7(5) 770.6(2) 836.3(2) 852.7(2) 881.0(1) 899.0(4) 924.3(5) 962.06(4) 978.8(3) 991 / 991.9(3) 10 4 104 104 104 105 105 104 105 104 104 104 105 103 105 105 104 106 105 6 x 104 63 '65 Cu Spectrum Identification, 0-1MeV Source Transition (keV) 2533 -+ 2278 65 Cu 2406 -+ 2094 63 Cu 1326 -* 962 63 Cu 2506 -~ 2092 65 Cu 2533 * 2094 63 Cu 1412 -* 962 63 Cu / 65Cu 2677 -- 2207 / 2094 -> 1623 65 Cu 2593 -- 2094 e annihilation -y 63 Cu 2081 - 1547 63 Cu 1547 - 962 65 Cu 1725 * 1115 65 Cu 2094 -* 1481 63Cu / 65Cu 2716 - 2092 / 2107 - 1481 63 Cu 2506 -* 1861 63 Cu 699 -- 0 63 Cu 63 Cu 2547 - 1861 / 2696 - 1861 63 Cu 4156 -> 3461 63 Cu 1412 - 669 63 Cu 2081 -> 1326 63Cu 2092 -+ 1326 65 Cu 770 -4 0 63 Cu 5413 -,4577 65 Cu 1623 -+ 770 63 Cu 2207 -- 1326 63 Cu 1861 -> 962 63 Cu 2336 - 1412 63 Cu 962 -- 0 65 Cu 2094 - 1115 63 Cu / 65Cu 2404 -* 1412 / 2107 - 1115 65 Cu Table 7: Neutron-induced Energy (keV) 1048.8(5) 1077.8(2) 1115.546(4) 1130.7(3) / 1129 1162.6(11) / 1163.7(11) 1178.9(3) 1245.2(2) 1290.0(19) 1327.03(8) 1341.7(6) 1346.4(2) 1350.1(4) 1374.47(13) 1389.66(8) / 1392.55(8) 1412.08(5) 1437.6(5) 1442.7(1) / 1442.2(3) 1547.04(6) 1558.4(3) 1585.4(2) 1624.0(2) / 1623.42(6) 1638(2) 1724.92(6) 1762.4(3) 1827.0(5) 1861.3(3) 1927.2(7) 1964.1(3) 63 '65 Cu Spectrum Identification, 1-2MeV Counts Source Transition (keV) 103 104 105 104 104 104 104 104 6x 105 7 x 103 63CU 2011 -+ 962 2404 -> 1326 5x 2x 8 x 9x 7 x 2x 5x 2x 63Cu 65 Cu 63Cu / 65Cu / 63Cu 63Cu 63Cu 63Cu 65 Cu 63 Cu 63 Cu 8 x 103 63Cu 4 x 104 1 x 104 4 x 104 63Cu 2 x 105 63Cu 103 63Cu 5x 3x 2x 1x 1x 4x 1x 7x 7x 9 x 2x 1x 1x 4 10 63 Cu 63Cu 63Cu / 65Cu 105 63Cu 104 104 104 104 104 65 Cu 63Cu 63Cu / 65Cu 65Cu 65Cu 103 65Cu 103 105 104 104 63Cu 63 Cu 63Cu 65Cu .1115 -- 0 2092 -* 962 / 2678 -* 1547 2278 -- 1115 / 2643 -t 1481 2506 - 1326 2207 -+ 962 2406 - 1115 1326 -+ 0 2011 -* 669 2673 -- 1326 2677 -* 1326 2336 -+ 962 2716 -+ 1326 / 2062 - 962 1412 -- 0 3775 - 2336 2404 -- 962 / 2212 -* 770 1547 - 0 2329 - 770 2547 -* 962 4130 -4 2506 / 1623 -- 0 3120 -+ 1481 1725 -+ 0 2533 - 770 2497 - 669 1861 -+ 0 2888 -+ 962 3079 - 1115 63 '6 5 Cu Table 8: Neutron-induced Energy (keV) 2011.4(5) /2012 2026.8(3) 2062.1(3) 2081.4(3) 2092.6(5) 2107 / 2107.4(2) 2188.0(7) 2212.8(2) 2309.0(3) 2329.0(2) 2336.5(3) 2356(3) 2468(3) 2497.4(4) 2512.0(5) 2536.0(3) 2562.0(7) 2627.7(1) 2696.6(3) 2716.9(4) 2780.3(4) 2806.6(6) 2862.7(2) 2874.4(2) 2889.4(8) 2902.4(2) 3032 3044.6(8) Counts 104 103 103 104 104 104 Spectrum Identification, 2-3MeV Source 63 Cu / 63Cu Transition (keV) 2011 -+ 0 / 2682 -- 669 63 Cu 2696 -+ 669 63 Cu 63 Cu 2062 2082 -+ 0 63 Cu 2092 -- 0 63CU / 65Cu 2776 -* 669 / 2107 103 63 Cu 2857 -- 669 104 65 Cu 2213 --+ 0 103 65 Cu 3079 -+ 770 103 65 Cu 2329 -+ 0 104 63 Cu 2336 -- 0 103 65 Cu 3127 -- 770 104 63 Cu 3428 -* 962 104 63 Cu 2497 --+ 0 103 63Cu 2512 -+ 0 104 63Cu 2536 -- 0 103 104 103 103 63 Cu 3888 -* 1326 63 Cu 3297 -÷ 669 63Cu 2697 -+ 0 63 Cu 2717 -* 103 63 Cu 2780 -+ 0 103 63 Cu 2807 -* 0 103 103 65 Cu 2862 -+ 0 65 Cu 2874 63 Cu 2889 -- 0 103 65 Cu 2902 -- 0 103 63 Cu 3032 -+ 0 103 63 Cu 3043 -~ 0 10 3 -+ 0 0 - 0 4.7 Gamma-ray Yields and Sensitivity Limits All analysis was performed using gf3 [27], which was developed specifically for use with Germanium detectors like the ones used in this experiment. When measuring the yield of a given gamma line in a given neutron bin, two primary methods were used. When the line in question was well resolved and appeared on a mostly flat background, the 'pk' function was used to automatically determine and subtract the background, and sum the remaining counts. This could be quickly automated to fit dozens of lines over many dozens of neutron energy bins. When the peak in question was weak or unresolved, the more careful 'nf' method was used. This procedure performed a least-squares fit for any number of peaks on top of a quadratic background function. Each peak was fit with three components: * a main Gaussian lineshape, which provided for most of the area of the peak * a small skewed Gaussian to model an exponential tail on the low-energy side of the line * a small, decaying step function on the lower-energy side to simulate Compton scattering When examining the regions of interest, we used a different procedure. When there were no detectable peaks in the region of interest, we estimated the sensitivity of our apparatus as follows. We used the resolution of the Germanium detector around the appropriate gamma-ray energy to estimate the width of the ROI. The full width at half maximum varied as 0.2% of the neutron energy, as discussed in Section 4.6. Modeling the background as a Poisson process, we took the square root of the counts in this region to be the standard deviation in the background. Multiplying this by 2, we obtained the 2a sensitivity threshold. With high probability, any peak yields with area smaller than this amount could not be distinguished from background. The livetime L for the gamma-ray spectrometer is described above in Section 4.4. The yield for a given peak was calculates as follows: Iy (En) C= -t (1 - ay) L B(E,) (counts/MeV) (9) In equation 9 above, Ict, is the number of counts for a given gamma peak 7 in a particular neutron energy bin, a, is the internal converstion coefficient for that line, and L is the livetime fraction for the detector array. To normalize the results per MeV, we divide by the size of the neutron energy bin B(E,). In general, the sensitivity threshold was on the order of one millibarn or lower, with a slight dependence on gamma-ray energy. At low gamma ray energies, the region of interest was smaller due to the more precise resolution of the apparatus, but there were a higher number of background counts. When a region of interest contained a feature, we measured its yield as above, and constructed its cross-section as given below. 4.8 Constructing the Integral Cross Section The neutron flux information, efficiency calibration and yield measurements are then all combined into a single differential cross-section. As noted above in Section 2.3, the integral cross-section is approximated by treating the whole detector array as making a single measurement. The three quantities above can be combined to give a measure of "how many instances of the transition occur per incident neutron". We want to turn this into a nuclear science measurement, and express the results in barns. The conversion ratio t is given in Section 3.3. Thus, the integral cross section of a line 3y in a neutron bin E, is given by: N I(y, E,)/(-y) F(y)- (E) = t (10) In equation 10 above, I is the corrected yield of a peak in that neutron energy bin as described in Section 4.7, c is the corrected efficiency at that gamma energy as described in Section 4.4, 1 is the corrected flux through that neutron energy bin as described in Section 4.3, F is the attenuation correction as described in Section 4.5 and t is the target thickness in atoms per barn, which for this experiment is given in equation 2. All other issues and corrections (internal conversion coefficient, livetime corrections, gamma attenuation, etc.) are folded into one of those three categories. N is a factor allowing for an absolute normalization to a reference line. For this experiment, all cross sections were normalized to the prominent and well-measured 1115-keV transition in 63 Cu. The reference value used for normalization is the result of the TALYS code [15], and is shown later in Section 6.1.2. The normalization factor was obtained by taking the ratio between TALYS predictions and experimental results around the threshold of the reaction and performing a least-squares fit. The normalization condition can be written as an equation for N. At any value E,, N is related to the reference value O1 115 at that neutron energy. Ul5u(En) -=N I(1115, E,)/E(1115) t (11) I(1115) - (En) Because of this normalization, there is an alternate way to write equation 10 that relates the cross-section of any measured line to the cross-section of the 1115-keV transition. I(y, E,) c(1115) F(1115) I1(1115, En) E(7) F(7y) Since all of the factors in equation 10 are multiplied or divided, standard error propagation [2] states that the relative error on the final result is the sum of the relative errors of all of the terms added in quadrature. This is discussed in more detail in the following section. When constructing cross-sections for prominent transitions in natural copper, statistics were good enough that the analysis could be performed using 15ns neutron energy bins (the limit on granularity, due to apparatus time resolution). However, when examining the 0v43 regions of interest, statistics were low enough that wider, 150ns bins, had to be used. 5 Error Analysis The experimental uncertainties of these measurements must be discussed. There are four major questions here. What fundamental or convenient necessary approximations and assumptions were made while constructing the cross-sections we present? How do these translate to sources of systematic errors? What sources of statistical error are there and how significant are they? How can we reduce the effects of these factors in later experiments? 5.1 Approximations and Systematic Error There are a number of necessary approximations in our model of the experiment made during the process of data analysis. They are discussed in order of descending prominence here, along with potential systematic effects. The angular distribution problem is the most serious. As of this writing, we do not have a rigorous treatment of the problem, and have been using the detector as a full-coverage spectrometer. In fact, as Table 2 shows, detectors tend to fall at specific angles. The worst-case scenario is that the angular distribution happens to have local maxima (or minima) at just the angles that detectors in the array sample. To calculate how bad of an effect this could have, we took typical differential cross-sections for the most common multipolarities [22], and calculated the average integral cross-section seen by sampling at the detector angles given in Table 2. For E2 multipolarities and El + M2 multipolarities, a typical cross-section might look like [22]: E2 : do•(cos 0) = co(1 + 0.5428P2 (cos 0) - 0.3428P4 (cos 0)) da M1 + E2: dO (cos 0) = ao(1 - 0.4282P2 (cos 0) - 0.0490P4 (cos 0)) (13) (14) Taking the average over the ten active coaxial detectors, our array would see 0.9056 0o for the E2 transitions and 1.054co for the E1+M2 transitions. The situation is slightly more complicated than this, however. Individual detectors have different efficiencies and attenuation coefficients, and so this average is idealistic. Furthermore, the typical differential cross-sections were taken from the reactions at threshold, and it is known that anisotropy falls off with neutron energy [19]. Since the transition used for normalization has El + M2 multipolarity, we set the systematic uncertainty from normalization at 5.4%. Since this is the worst-case analysis, this probably overstates the systematic error. Nevertheless, we include it. The gamma ray attenuation problem is the next most prominent. We approximated the attenuation coefficient for the array by averaging the individual coefficients seen by each detector. However, individual detectors have a slightly different efficiency, so this is not absolutely correct. In fact, the standard deviation in the mean absolute efficiency of all the detectors is 16 - 17% over the majority of the energy range. In this manuscript, we are unable to provide a systematic uncertainty associated with the attenuation problem. It is one of the few key issues we are still investigating. However, there are two reasons why this effect falls off at high energies and is thus likely to be less relevant to our region of interest results. First, the attenuation of a given gamma ray through any amount of material decreases asymptotically with increasing neutron energy. Second, the spread of detector efficiency is greater at lower energies because the detectors have different low-energy suppression behavior. Another potential problem is the even distribution of deadtime. In the worst-case scenario, the data acquisition system is less able to write events to disk when many of them happen at once. Thus, at neutron energies that cause a large amount of gamma ray events, the deadtime might be higher. After the deadtime correction was made, the net effect of this would be to supress high cross sections and abnormally raise low cross sections. However, we are not concerned with this for two reasons. First, this is somewhat cancelled out by a similar effect in the fission chamber. Second, when the cross-section is written using the normalization to the 1115-keV line in equation 12, any livetime correction to the corrected flux, yields and efficiency measurements cancels. The promptness of the excitation and decay of the target is an issue. If the excited states of the nuclei have a long half-life, and there is a significant delay between excitation and decay, then transitions measured later would be interpreted by the experimenter as being caused by a later-arriving (i.e. slower) neutron. The overall effect would be to distort the cross-section as a function of neutron energy towards the lower energies. In practice, all of the transitions presented here are produced by levels with half-lives on the order of picoseconds [17]. There is a possibility that some very high-lying levels have a half-life on the order of nanoseconds, and in the cascade of gamma rays down to the ground state. To measure this effect directly, we examined the spectrum at neutron energy levels below threshold for the first excited states in 63,65 Cu. There was no significant signal. Similarly, the time it takes for a created gamma ray to leave the target and enter a detector (- 15cm in just - 0.5ns) is not a significant fraction of the apparatus senstivity. Another issue is the neutron-overlap problem. Specifically, as the highenergy neutrons of the next micropulse are arriving, the slow neutrons from the previous pulse are still hitting the target. However, with a micropulse width of 1.8 ps and a distance of 20.34m between the spallation source and the target, only neutrons with E, = 650keV have not yet hit the target. This is not high enough in energy to activate either first excited state (e.g. the 670-keV level). Another possibility for a small adjustment is the difference between the collision rest frame and the lab frame in which transitions are measured. The conservation of linear momentum in the incident neutron + atom at rest system means that the direction of motion of created gamma rays have some non-zero component along the direction of the beam, on average. A recent publication [25] covers this effect in more detail, giving a correction to the angular distribution caused by this effect. While this effect is prominent in reactions on light nuclei (e.g. In +1 H), it is not significant in relatively large atoms such as 63 '65 Cu, which are many orders of magnitude more massive than the incident neutrons. The velocity of the rest frame (v = .003c even with incident energy E, = 200 MeV) is not fast enough to skew the angular distribution. Because of the neutron arrival resolution time in the experimental apparatus, our measure of the cross-section from inelastic neutron scattering has to be reported in neutron bins that are, at the minimum, 15 ns wide. When the cross-section as a function of energy is slowly-changing, this is a fine approximation. However, this smudging out of the cross section makes it impossible to measure phenomena at specific neutron energies, such as the threshold for a transition. In general, neutron binning distorts the shape of the cross-section. Since this experiment is primarily interested in measuring lines in the regions of interest with very small cross-sections, very large neutron bins are used anyway, and we are not concerned with this effect. The last minor issue is that the experimental target was natural, not isotopic, copper. This had a few implications. First, it was impossible to seperate any overlapping lines in 63Cu and 65 Cu, such as the important 63 Cu 365.2-keV and 65 Cu 366.3-keV lines, and report their cross-sections separately. Secondly, for 63 Cu transitions, it is impossible in theory to seperate the contribution to the cross-section from 63 Cu (n,n) reactions from the contribution from 65 Cu (n,3n) reactions. In practice however, the latter reactions had a much higher-energy threshold, by which point the former crosssection had largely fallen off. Finally, while it was possible to take measured (n,xn) reactions in isotopic copper, take a weighted sum and measure them against our data, the reverse process does not work. This point is mitigated by the fact that the aim of this experiment is to see how natural copper behaves in future experiments: information about individual isotopes is not required. 5.2 Statistical Error Every counting experiment has statistical error. For an experiment measuring nuclear events, the error associated with any number has a Poisson uncertainty. This applies to two key of the total cross-section. Individual peak yields have a roughly square root error. When statistics for a given peak are low (for example, because of a small choice of neutron energy bin or for a rare transition), the relative error can be significant. This kind of error applies both to peak yields used in constructing cross sections as described in Section 4.7 and also to measurements used to provide an efficiency calibration as described in Section 4.4. When incident neutrons cause fission in 235,238 U atoms, the fission chamber registeres a number of counts due to this process. The number of decays measured by the fission chamber is a nuclear measurement, but due to several other issues such as experimental uncertainty in the (n, f) cross-sections, degradation of the fission chamber over time, and apparatus time resolution, the experimentally determined relative error is on the order of 5%. This amount decreases slightly with neutron energy. The error in the efficiency calibration comes from the uncertainty in the fitted efficiency function coefficient, which tends to be very small since the efficiency spectra have good statistics. However, it increases with gamma-ray energy, especially past 2.5-MeV, since there are no efficiency calibration peaks there. Since the efficiency calibration has no neutron-energy dependence, we consider it a source of systematic, instead of statistical, error. Although the livetime correction for the efficiency calibration, yield and fission chamber measurements are significant and important, since they concern such a large number of counts, the statistical error introduced by these corrections is not significant. These two key sources of statistical uncertainty are added in quadrature to come up with the final statistical error in the experimentally measured cross-section. 5.3 Error Propagation The major sources of statistical error come from the counts used to construct the gamma-peak yield and the flux through a given neutron bin. The major sources of systematic error are the uncertainty in the overall normalization due to angular effects and the systematic uncertainty involving the efficiency. Following [2] and equation 10, the final uncertainty at a neutron bin E, for a peak y is given by equation 15 below. + I 4b (E,) )2E ) ( (Ia(E") (En) + N(E,)) + _(E ) (15) To give an illustration of how each relative error term contributes to the final result, and which uncertainties are prominent at which regions of neutron energy, the terms in equation 15 are plotted in Figure 12 for the measured cross-section of the 962-keV transition. At low neutron energies, the uncertainty in the flux tends to dominate. At higher neutron energies, the uncertainty in the peak yields tends to dominate. I .. o 0J w 41-J Cu neutron energy (MeV) Figure 12: Contributions to Error in the 962-keV Cross-Section The highest curve is the sum of all of the error terms. In general, the error from peak yields dominates these measurements. When statistics are low, such as with region of interest-contaminating line cross sections, statistical error from background dominates all other factors. In fact, for the ROIs, statistics are so low that we report results in 150ns-wide bins. When statistics are high, such as with the (n, n') cross sections of prominent copper lines, systematic errors below are responsible for the discreprancy betwen simulation and experiment. 5.4 Future Error Reduction There are two components to fully investigating the angular problem. The first is a theoretical understanding of the cos 0 dependence of the cross-section for given transitions, and how these vary as a function of neutron energy. The second is to understand the systematics and efficiency issues of each individual detector so that differential cross-sections can be constructed. When we attempted to do this in the past, we found that although GEANIE's results for integral cross-sections were mostly faithful given appropriate calibration and normalization, systematic errors and eccentricities between individual detectors were larger than the measured differences in the differential cross-section between them. Both of these are potential tasks to work on in future experiments of this type at this facility. In particular, as individual detector systematics are understood, a measurement of angular dependence becomes more of a possibility. To fully investigate the gamma attenuation issue, a full Monte Carlo N-Particle (MCNP) code simulation of the gamma ray attenuation through the target should be done. This simulation will take into account the geometric details of the target and detector array. One of the advantages of this treatment is that the attenuation and efficiency calibration would take place within a single measurement. To ensure that all transitions under investigation are prompt, a coincidence analysis between transitions between higher energy levels and lower energy levels and between lower energy levels and the ground state should be performed. This way, the connection between which excited energy levels cause which observed transitions can be made explicit. For example, if the second excited ground state has a high halflife, but the first excited state has a very prompt one, then a coincidence measurement between 2 -* 1 and 1 -- gs events can establish what proportion of the 1 --+ gs cross-section is caused by the second ecxited energy level. Unfortunately, the only way to defeat statistical error is with more statistics. Since the statistics in this experiment come from Poisson-process nuclear decay events, the error varies as the square root of the number of counts. For both fission chamber and detector array events, the total number of events varies linearly with the runtime. Since each data point contains contributions from these errors in quadrature, doubling the run time is likely to supress uncertainty by a factor of V2. The unfortunate upside is that to obtain an extra digit of precision in any measurement, the experiment would need to run for ten times as many days. Since this data was obtained with two weeks at the facility, which is only operational for part of the year, this is not feasible. In summary, there are numerous systematics. Some of them can be shown to be insignificant or not applicable, we can put a numerical cap on the effect of some others, and others are still under investigation. Future experimenters can probably make significant improvements to these systematics. As we said above, the statistical errors are unlikely to improve unless run time increases significantly. To that extent, the statistical uncertainty on the data presented in this experiment is likely to be the best that this facility can produce. 6 6.1 Experimental Results natCu (n,xn'y) Cross Sections Below are results for the integral cross section of prominent neutron-induced transitions in natural copper. We present four prominent lines in 63Cu and two in the less-abundant 65 Cu isotope. 6.1.1 63 '65 Cu (n,xn'y) 6 5Cu We present integral gamma-ray production cross-section measurements for (n, n'y) on 63 Cu and (n, n'-) reactions on 65 Cu that result in 63 Cu lines. Understanding the disreprancies between theory and experiment both drives nuclear data simulations and validates region of interest results, presented in the next subsection. Below, the cross-section for the 670-keV first-excited state to ground state transition is shown in Figure 13, the cross-section for the 962-keV second-excited state to ground state transition is shown in Figure 14, the cross-section for the 1327-keV third-excited state to ground state transition is shown in Figure 15 and the crosssection for the 1412-keV fourth-excited state to ground state transition is shown in Figure 16. All of the plots have similar qualitative features. There is a threshold energy equal to the energy of the transition (below this energy, the neutron can't actually excite the nucleus high enough). There is a peak at which the (n, n') reaction is most probable, and then a fall-off. In the case of 63 Cu lines, there are two peaks. The first is from (n, in') reactions with energy 3 - 10MeV neutrons. The second, smaller peak is from (n, 3n') reactions on 65 Cu with energy 12+ MeV. Since more energy is required to knock neutrons out of the nucleus, the threshold for this reaction is higher. The total (n, xn) cross-section is the super-position of these two effects. The plots on left side show the experimentally determined cross-section in blue and the result of a TALYS code simulation in red. The plots on the right side show the relative experimental error in blue and the relative difference between theory and experiment in red. - (Exp - Th) / Th Stat. Error Ii 02 0. .. ,, 63 Cu Figure 13: Calculated and Simulated nat-Cu (n,n') 63-Cu 962-keV transition integral cross section ncident neutron energy (MeV) incident neutron energy (MeV) 670-keV cross-section 0.61 - (Exp - Th)/ Th I - Stat. Error I A\ i; 0.0 10i Incident neutron energy (MeV) Figure 14: Calculated and Simulated Incident neutron energy (MeV) 63 Cu 962-keV cross-section nat-Cu (nn') 63-Cu 1327-keV transition integral cross section Incident neutron energy (MeV) Incident neutron energy (MeV) Figure 15: Calculated and Simulated 63 Cu 1327-keV cross-section nat-Cu (n,n') 63-Cu 1412-keV transition integral cross section Incident neutron energy (MeV) Figure 16: Calculated and Simulated 6.1.2 65 Cu (n,n'7) Incident neutron energy (MeV) 63 Cu 1412-keV cross-section 6 5 Cu Similarly, we present integral gamma-ray production cross-section measurements for (n, n't) on 65 Cu that result in 65 Cu lines. The cross-section for the 1115-keV second-excited state to ground state transition is shown in Figure 17 and the cross-section for the 1481-keV third-excited state to ground state transition is shown in Figure 18. In this case, there is no (n, xn) contribution from isotopes with a higher N number, so the total cross-section is just the result of (n, n) reactions. The 1115-keV line was used to perform an absolute normalization. This is the reason for the agreement between theory and experiment in the threshold region for the reaction. Again, the plots on left side show the experimentally determined cross-section in blue and the result of a TALYS code simulation in red. The plots on the right side show the relative experimental error in blue and the relative difference between theory and experiment in red. Figure 17: Calculated and Simulated 65 Cu 1115-keV cross-section on Incident neutron energy (MeV) Incident neutron energy (MeV) Figure 18: Calculated and Simulated 6.2 65 Cu 1481-keV cross-section OvP,3 Regions of Interest Below are results on the regions of interest for seven neutrinoless double-beta decay isotope candidates. Again, we examined the region around the energy at the reaction endpoint and the region 1022-keV higher because of the possibility of an electron signal from a double-escape peak. In cases of an overlapping line, we present the integral cross-section. All of the lines we observes were known transitions in 63 ,65 Cu. Otherwise, we give the sensitivity limit of the apparatus. A summary of our results is given in Table 9. We do not have results for regions of interest above the upper energy limit of our 0 detector (- 4MeV), for the double-escape peak ROI in ooMo and 15 0Nd and in both 48 Ca regions. For each isotope, we discuss a few current, future and past experiments that use this candidate to search for v0/3. The list is not comprehensive, but rather a sampling of current research. Table 9: Experimental Ovoy Isotope Region of Interest Results ROI Energy 2039-keV 3061-keV 2995-keV 4017-keV ROI Spectrum Result at E, = 8 MeV n/a n/a < 0.39 mbarn looMo 3035-keV 3034-keV line 116Cd 2809-keV 3831-keV 2808-keV line n/a 2530.3-keV 3552-keV 2536-keV line n/a 76 Ge 82Se 130Te 3003-keV line n/a < 0.42 mbarn 4.2 mbarn < 0.29 mbarn 4.6 mbarn 4.4 mbarn < 0.30 mbarn 8.5 mbarn < 0.35 mbarn 2457-keV n/a < 0.39 mbarn 3479-keV 3476-keV line 2.4 mbarn 15 0 Nd 150Nd 3368-keV < 0.38 mbarn n/a 1 J , i A summary of our results. For every isotope which had regions of interest within our visible energy range, we report on any lines that fell within that ROI, and give the measured cross-section at E, = 8 MeV. 13aXe 6.2.1 76 Ge 76Ge is a popular double-beta decay candidate and is used in several current, past and future experiments. One of its advantages is that it can be used as both detector and source in these experiments, two of which we list here. In the present work, we present sensitivity limits on the otherwise-clean 76 Ge regions of interest in Figures 19 and 20. The MAJORANA experiment aims to perform a low-background measurement of neutrinoless double-beta decay in 86% enriched 76 Ge, and refute or confirm the Klapdor-Kleingrothaus claim of OvPP3 in 76 Ge. The experimental site will be held underground, and recent R&D efforts are focused on keeping the germanium crystals in an electropurified copper cryostat. One of the original motivations for this experiment was to make (n, n') background measurements for the MAJORANA collaboration. The GERmanium Detector Array (GERDA) is a 76 Ge experiment currently under construction, which plans to use enriched germanium crystals in a sealed copperplated cryostat filled with a liquid noble gas. Like other experiments outlined in this section, the experimental site would be installed deep underground in the Laboratori Nazionali del Gran Sasso (LNGS), in Italy. 1i0n 00 -.10o Figure 19: 76 Ge Energy (kev) Figure 20: 76 Ge nat-Cu (nn') 2040-keV ROIsensitivity limit 10o Incident neutron energy (MeV) OvPP Q-value ROI and sensitivity limit. Incident neutron energy (MeV) Ov/3 Q-value DEP ROI and sensitivity limit. 6.2.2 82 Se The wide 3004(3)-keV line from the 63 Cu 4416 -4 1412 transition contaminates the region around the 82 Se Q-value, and we give an integrated cross-section for it in Figure 21. The higher ROI is clean, and we give a sensitivity limit in Figure 22. This isotope is one of the ones used by the Neutrino Ettore Majorana Observatory (NEMO) in the current NEMO-3 incarnation of the experiment, which has a movable center into which different sources can be placed. The apparatus is installed below the earth's surface at the Frejus Underground Laboratory in France, at a depth of 4800 meters of water equivalent, and the source is surrounded by a purified copper shield. Recently, the NEMO collaboration has published results using Ikg of 97% enriched s2 Se. nat-Cu (n,n') 63-Cu 3003 keV '' 10c Incident neutron energy (MeV) Energy (kev) Figure 21: 82 Se Ovo0 Q-value ROI and cross-section. nat-Cu (n,n') 4017-keV ROI sensitivity limit Ln C: 0 U _i~ 3970 3980 3990 4000 4010 Energy (kev) Figure 22: 82 Se 4020 4030 4040 101 Incident neutron energy (MeV) 0v3/3 Q-value DEP ROI and sensitivity limits. 1 00Mo 6.2.3 The 3032.1(10)-keV line from the 63 Cu 3032 --+ 0 transition contaminates the region around the 10oMo Q-value, and we give an integrated cross-section for it in Figure 23. The higher ROI, at 4055-keV, was beyond the limits of the experimental facility. The MOlybdenum Observation of Neutrinos (MOON) proposes to study double beta decay using one ton of 85% pure 10 0Mo as the source. In addition to being a O/30 candidate, 100Mo has several attractive properties in neutrino physics such as being a good solar neutrino detector through inverse beta decay. The NEMO-3 experiment above also uses a 98% pure 100Mo foil as a source. nat-Cu (n,n') 63-Cu 3032 keV 4200 4000 4- 3800 3 W 3600 0 3400 0 u 3200 3000 0100 Energy (kev) Figure 23: 6.2.4 100Mo 01 102 Incident neutron energy (MeV) Ov/3 Q-value ROI and cross-section. 116 Cd The 2806.6(6)-keV line from the 63 Cu 2806 --, 0 contaminates the region around the 116 Cd Q-value, and we give an integrated cross-section for it in Figure 24. The higher ROI is clean, and we give a sensitivity limit in Figure 25. The Cadmium-Telluride O-neutrino double-Beta Research Apparatus (COBRA) experiment aims to detect neutrinoless double beta decay using CdZnTe ("CZT") crystals, which contain 11 Cd as well as 30oTe, both beta-decay candidates. The COBRA prototype array of CZT crystals sits inside a a thick natural copper core, making the present work particularlt relevant. The experiment takes place deep underground, in the Laboratori Nazionali del Gran Sasso (LNGS). nat-Cu (n,n') 63-Cu 2808 keV ++ U neutron 0Incident energy (MeV) Incident neutron energy (MeV) Figure 24: 116 Cd Ovil Q-value ROI and cross-section. nat-Cu (n,n') 3831-keV ROIsensitivity limit Incident neutron energy (MeV Incident neutron energy (MeV) Figure 25: "116CdOv/3 Q-value DEP ROI and sensitivity limit. 6.2.5 130Te The 2536.0(3)-keV line from the 63 Cu 2536 - 0 transition contaminates the region around the 130Te Q-value, and we give an integrated cross-section for it in Figure 26. The higher ROI is clean, and we give a sensitivity limit in Figure 27. This isotope is used in the Cryogenic Underground Observatory for Rare Events (CUORE) experiment, which proposes to use a large array of TeO 2 crystals (with 33.8% abundant 130Te) to measure neutrinoless double-beta decay. CUORE will use 741 kg of TeO 2 deep underground in the Laboratori Nazionali del Gran Sasso (LNGS) installation in Italy, at 3400 meters of water equivalent. As mentioned above, the COBRA experiment is searching for OvP/ in 130 Te. nat-Cu (n,n') 63-Cu 2536-key + 65-Cu 2533-keV 4% 4-t 2 -t 0 101 Incident neutron energy (MeV) 13 0Te Figure 26: S-- 2500 OvoP Q-value ROI and cross-section. u target3-30 MeV 1.0 nat-Cu (n,n') 3552-keV ROI sensitivity limit 1 2400 2300 U %lia1~ "¼ S200 - ------ .. 1900 1800 3520 3540 3560 Energy (kev) Figure 27: 6.2.6 13 0Te 3580 3600 .,i 10i Incident neutron energy (MeV) I O3/03 Q-value DEP ROI and sensitivity limit. 136 Xe The lower ROI is clean, and we give a sensitivity limit in Figure 27. The 3476-keV line from the 63 Cu 3476 0 transition contaminates the region around the higher 136Xe Q-value, and we give an integrated cross-section for it in Figure 29. - This isotope is used in the Enriched Xenon Observatory (EXO) experiment, which proposes to use over 200kg of 80% enriched 136 Xe cylinders to search for neutrinoless double-beta decay. EXO-200 will be installed underground at the Waste Isolation Pilot Plant (WIPP) in Carlsbad, New Mexico. The Gotthard underground laboratory is also conducting a 136 Xe experiment, using 2 x 105 cm3 of xenon gas at 5 bar, with 62.5% pure 13 6Xe. A copper shield protects the xenon, which serves both as source and detector, so this measurement is particularly relevant. The experiment takes place underground at a laboratory in the Swiss alps, at 3700 meters of water equivalent. rn I nat-Cu (n,n') 2457-keV ROI sensitivity limit 0.8 0.4 o -E z= 0.2 [) o0 i Figure 28: 136 Xe v0/3 l 10 Incident neutron energy (MeV) Q-value ROI and sensitivity limit. nat-Cu (n,n') 63-Cu 3476 keV 101 Energy (kev) Figure 29: 136Xe Ov0/ 6.2.7 Incident neutron energy (MeV) Q-value DEP ROI and sensitivity limit. 15 0Nd There were no identificable features in the region of interest for 5's Nd. The sensitivity limits on the otherwise-clean ROI are shown in Figure 30. The higher ROI, at 4389keV, was beyond the limits of the experimental facility. The SuperNEMO experiment, an improvement of the NEMO-3 experiment listed earlier, proposes to use an 15 0Nd source, but is still under development. The Drift Chamber Beta-ray Analyzer (DCBA) proposes to use a large-magnetic field tracking chamber to search for neutrinoless double-beta decay in 80% enriched 150 Nd. This project is under development in Japan. ~nnn nat-Cu (n,n') 3367-keV ROIsensitivity limit 1.0 -- Cu target 3-30 MeV I 2900 2800 02700 ~2600 ~i-L~.; 2500 2400 ~n i 1 i 2300 3320 3340 3360 3380 nn U. 10o 3400 Energy (kev) Figure 30: 15 0Nd dent neutron energy e) Incident neutron energy (MeV) Ov,0 Q-value ROI and sensitivity limit. 62 7 Conclusion We have presented a quantitative investigation of the neutron-induced gamma ray radiation in natural copper that contaminate regions of interest corresponding to the Q-value and double-escape peak regions in six double-beta decay candidate isotopes. When a neutron-excited 63 '65 Cu transition overlapped with a given region of interest, we calculated the integral cross section over the neutron energy region from 1 to 200 MeV. When the region of interest had no visible features, we reported the limits of our sensitivity. To give our results credibility, we discuss the experimental apparatus in some detail and calculate the (n, xn'y) cross sections for prominent transitions in natCu and compare them with the results of TALYS nuclear simulations. We discussed the efficiency calibration and data analysis process. We investigated sources of systematic error such as the angular problem, overall normalization, deadtime issues, gamma ray attenuation and statistical error. These measurements are important for experimenters planning the next generation of underground double-beta decay experiments. Integral cross section measurements for region of interest contaminants in natural copper, along with an understanding of cosmic ray-induced spallation neutron flux, give important benchmarking data points for Monte Carlo simulations and background studies. Since the regions of interest for many Ov00 candidate isotopes were investigated, these results apply to many experiments under development which take place underground, where neutron flux is a source of background, and with significant natural copper around the experimental site. This analysis can be performed with other detector and shielding materials. Within the last year, data was taken at GEANIE using a natural lead target, and a similar analysis with enriched germanium is being performed in parallel. Our team at Los Alamos has recently proposed beam time at LANSCE for a CZT target. We hope that the results presented here will contribute to the exciting experimental efforts in neutrino physics. References [1] G. Audi, A.H. Wapstra, and G. Thibault. Nucl. Phys. A, 729(337), 2003. [2] P.R. Bevington and D.K. Robinson. Data Reduction and ErrorAnalysis for the Physical Sciences. McGraw-Hill, 2003. [3] C.E. Crouthamel. Applied Gamma-ray Spectroscopy. Pergamon Press, 2nd edition, 1970. [4] S.R. Elliott and J. Engel. Double beta decay. J. Phys G, (30), 2004. [5] K. Eguchi et. al. KamLAND Collaboration. First results from kamland: Evidence for reactor antineutrino disappearance. Phys. Rev. Lett., 90, 2003. [6] Q.R. Ahmad et. al. SNO collaboration. Measurement of the rate of ve + d p + p + e- interactions produced by 8b solar neutrinos at the sudbury neutrino observatory. Physical Review Letters, 87, August 2001. [7] Y. Fukuda et al Super-Kamiokande Collaboration. Evidence for oscillation of atmospheric neutrinos. Phys.Rev.Lett., 81, 1998. [8] N. Fotiades, G. D. Johns, R. O. Nelson, M. B. Chadwick, M. Devlin, W. S. Wilburn, P. G. Young, J. A. Becker, D. E. Archer, L. A. Bernstein, P. E. Garrett, C. A. McGrath, D. P. McNabb, and W. Younes. Measurements and calculations of 238 u(n,xny) partial -y-ray cross sections. Phys. Rev. C, 2004. [9] V.M. Gehman. Physics Reach of the Global Neutrinoless Double-Beta Decay Program and Systematic Uncertainties of the Majorana Project. PhD thesis, University of Washington, 2007. [10] David J Griffiths. Introduction to Quantum Mechanics. Pearson Prentice Hall, 2nd edition, 2005. [11] International atomic energy agency nuclear data centre, 2004. Accessed online at http: //www-nds. iaea. org/xgamma_standards/genergiesb .htm. [12] C L Cowan Jr, F Reines, F B Harrison, H W Kruse, and A D McGuire. Detection of the free neutrino: a confirmation. Science, (124), Jul 1956. [13] H.V. Klapdor-Kleingrothaus, A. Dietz, H.V. Harney, and I.V. Krivosheina. Evidence for neutrinoless double beta decay. Modern Physics Letters A, 16(37), 2001. [14] Glenn F. Knoll. Radiation Detection and Measurement. John Wiley & Sons, 3rd edition, 2000. [15] A.J. Koning, S. Hilaire, and M. Duijvestijn. Talys-1.0 user manual, 2007. Accessed online at http://www. talys .eu/f ileadmin/talys/user/docs/talysl. O.pdf. [16] Lawrence berkeley national laboratory table of radioactive isotopes, 2003. Accessed online at http://ie.lbl. gov/toi/radSearch. asp. [17] Lawrence berkeley national laboratory table of isotopes, 2003. Accessed online at http://ie.lbl. gov/T012003/GammaSearch. asp. [18] Matlab-like plotting library (matplotlib) homepage, 2008. Accessed online at http: //matplotlib. sourceforge .net/. [19] D.P. McNabb, J. A. Becker, D. Archer, L.A. Bernstein, D. M. Drake, P. E. Garrett, G. D. Johns, C. A. McGrath, R. O. Nelson, M. A. Stoyer, W. S. Wilburn, and W. Younes. Neutron-induced partial gamma-ray cross-section measurements with geanie at lansce/wnr. In CAPTURE GAMMA-RAY SPECTROSCOPY AND RELATED TOPICS: 10th InternationalSymposium. AIP Conference Proceedings, 2000. [20] D.-M. Mei, S.R. Elliott, A. Hime, V. Gehman, and K. Kazkaz. Neutron inelastic scattering processes as background for double-beta decay experiments. ?, 2007. [21] D.-M. Mei and A. Hime. Muon-induced background study for underground laboratories. Physical Review D, 2006. [22] L.C. Mihailescu. Neutron (n,xn-y) cross-section measurements for 52Cr, 209Bi and 206,207,208Pb from threshold up to 20 MeV. PhD thesis, University of Bucharest, 2007. [23] Table of x-ray mass attenuation coefficients, 1996. Accessed online at http: //physics. nist. gov/PhysRefData/XrayMassCoef/tab3. html. [24] X-ray mass attenuation coefficients, 1996. Accessed online at http://physics. nist. gov/PhysRefData/XrayMassCoef/chap2. html. [25] Heather Olliver, Thomas Glasmacher, and Andrew Stuchbery. Angular distribution of y rays with intermediate-energy beams. Physical Review C, 68, 2003. [26] Wolfgang Pauli. Dear radioactive ladies and gentlemen, Dec 1930. Letter. [27] D.C. Radford. Notes on the use of the program gf3, 2000. Accessed online at http://radware.phy.ornl.gov/gf3/gf3.html. [28] D.C. Radford. Radare homepage, 2000. Accessed online at http://radware. phy. ornl. gov. [29] S. Rahaman et al. Q value of the 100Mo Double-Beta Decay. B662:111-116, 2008. Phys. Lett., [30] M. Redshaw, E. Wingfield, J. McDaniel, and E. G. Myers. Mass and doublebeta-decay q value of 13 6xe. Phys. Rev. Lett., 98, 2007. [31] Scientific python (scipy) homepage, 2008. scipy.org. Accessed online at http://www. [32] H. Vonach and A. Pavlik. 20 7,2 08pb(n,xny) reactions for neutron energies 3 to 200 mev. Physical Review C, 1994. [33] S.A. Wender, S. Balestrini, A. Brown, R.C. Haight, C.M. Laymon, T.M. Lee, P.W. Lisowki, W. McCorkle, R.O. Nelson, W. Parker, and N.W. Hill. A fission ionization detector for neutron flux measurements at a spallation source. Nuclear Instruments and Methods in Physics Research A, (336), 1993. [34] P.G. Young, E.D. Arthur, and M.B. Chadwick. Comprehensive nuclear model calculations: theory and use of the gnash code. In A. Gandini and G. Reffo, editors, Proceedings of the IAEA Workshop on Nuclear Reaction Data and Nuclear Reactors: Physics, Design and Safety, 1996. [35] Kai Zuber, editor. Summary of the Workshop on: Nuclear matrix elements for neutrinoless double beta decay. Institute for Particle Physics Phenomenology, May 2005. 8 # # # # Appendix I: geanie.py Source Code This is a set of GEANIE (n,n') measurement-related functions. There are utilities to usefully manage data in the formats used throughout the GEANIE analysis, and also utilities to intelligently construct cross-section measurements. # copyright 2007 Dennis V. Perepelitsa (dvp@mit.edu, dvp@lanl.gov) # import a non-linear least-squared fitting module written by dvp import jlabfit # import numerical analysis and plotting python packages import scipy, pylab # import other useful modules import os # we model the efficiency as a high-order polynomial in log-energy, # log-efficiency space def f6(params, y): """This is a helper function which defines the functional form of the efficiency for a solid-state, high-purity germanium detector. You shouldn't call it directly.""" # this is the energy to which the logarithm of all else is # relative x = scipy.log(y/1000.0) return params[0] + params[l] *x + params[2]*(x**2)\ + params[3]*(x**3) + params[4]*(x**4) + params[5]*(x**5) def writeout(fil, desc, cols): ,I II II This is a simple function that writes out data to a file. It's intended to have an easy interface. fil : filename to write to desc : the first line of the file cols : an array of columns of data II II II f = file(fil,'w') f.write(desc + '\n') for i in range(len(cols[O])): for j in range(len(cols)): f.write('cols[j] [i]' + '1 ) f.write('\n') f.close() def __sum_scaler(scalerfile): """This is a helper function that extracts a scaler from an appropriate scaler *.spe converted to a textfile. You shouldn't call it directly.""" f = file(scalerfile, "r") lines = [] n = [] line = f.readline() # cut the first one line = f.readline() while line != "": for j in line.split(): n += [int(j),] line = f.readline() s =0 for i in range(4096): s += i * n[i] f.close() return s def getlivetime(gmtfile,detarray): """Returns an array of livetimes in a given .gmt for each detector index in the passed array (FC1 and 2 are treated as det# 29 and 30, respectively). Currently ignores livetime uncertainty.""" livetimes = [] for det in detarray: res = os.popen4("echo -e \"id " + 'det' + "\\nws tmp-" +\ 'det' + "\" I rgmt " + gmtfile) for line in res[l].readlines(): pass res = os.popen4("echo -e \"id " + '290+det' + "\\nws tmp-"\ + '290+det' + "\" I rgmt " + gmtfile) for line in res[l].readlines(): pass res = os.popen4("sumspe tmp-" + 'det' + ".spe") # we need the standard output result here for line in res[ll.readlines(): adc = float(line.split() [-1]) res = os.popen4("echo -e \"1\\nn\\ntmp-" + '290+det' +\ ".spe\\n\" I spec_ascii") for line in res[l].readlines(): pass scaler = __sum_scaler("tmp-" + '290+det' + ".txt") print adc, scaler livetimes += [adc/scaler,] res = os.popen4("rm tmp-" + 'det' + ".spe tmp-" + 'det'\ + ".txt tmp-" + 'det+290' + ".spe + tmp-"\ + 'det+290' + ".txt") for line in res[ll.readlines(): pass return livetimes def getefficiency(effparams, effparamrelerr, E): """Returns the calculated efficiency at this value of energy as a tuple with relative uncertainty.""" eff = scipy.exp(f6(effparams,E)) x = scipy.log(E/1000.0) abserrsq = 0 # propagate absolute error term by term for i in range(len(effparams)): abserrsq += (eff**2)*(effparamrelerr[i]**2)*((x**i)**2) return (eff, scipy.sqrt(abserrsq)) def calculate_efficiency(efffile,effsource,decays,livetime,\ plot=False,silent=True): """Takes a source calibration file, a calibration file for this detector, total number of decays and detector deadtime, and returns a set of polynomial coefficients that describe the best fit. """" # read in intensities here fil = file(effsource,'r') sourcelines = fil.readlines() fil.close() fil = file(efffile,'r') caliblines = fil.readlines() fil.close() E = [] intensity = [] intensityrelerr = [] cts = [] ctsrelerr = [] if not len(sourcelines) == len(caliblines): print "Number of source and eff peaks MISFIT" for i in range(len(sourcelines)): x = caliblines[i] .split() # if we didn't see this peak, skip to the next source peak if not int(x[1]) == 0: cts += [float(x[1]),] ctsrelerr += [float(x[2] )/float(x[l]),] x = sourcelines[i] .split() E += [float(x[O]),] intensity += [float(x[1)/100.,] intensityrelerr += [float(x[21 )/float(x[1]) ,] E = scipy.array(E) intensity = scipy.array(intensity) intensityrelerr = scipy.array(intensityrelerr) # canonical use of 4\pi solid angle factor logeff = scipy.log(4*scipy.pi*scipy.array(cts)\ /(decays*(livetime)*intensity)) # perform correct error propagation logeffrelerr = scipy.sqrt(scipy.array(ctsrelerr)**2\ + intensityrelerr**2)/logeff pO = [1.0,1.0,1.0,1.0,1.0,1.0] res = jlabfit.fit(f6,E,logeff,logeffrelerr*logeff,p0,\ plot,None,silent) return (res[O],scipy.array(res[1])/res[0]) def getflux(fluxfile,livetime): """Takes a flux per MeV output file created with fluxperMeV.f and fission chamber deadtime and returns a tuple of flux and flux relative uncertainty.""" flux = [] fluxrelerr = [] fil = file(fluxfile,'r') lines = fil.readlines()[1:] fil.close() for line in lines: x = line.split() flux += [float(x[1]),] fluxrelerr += [float(x[2] )/float(x[1]),] flux = scipy.array(flux)/(livetime) fluxrelerr = scipy.array(fluxrelerr) return (flux, fluxrelerr) def getbins(binfile): """Takes a .bin file output by exbins and returns a tuple of neutron bin mid-points and bin sizes.""" binmid = [1 binwidth = [] fil = file(binfile,'r') lines = fil.readlines()[6:] fil.close() for line in lines: x = line.split() binmid += [float(x[2]),] binwidth += [float(x[1]) - float(x[O]),] binmid = scipy.array(binmid) binwidth = scipy.array(binwidth) return (binmid, binwidth) def get_yields(yieldfile,column,bins,livetime,alpha=O.0): """ Takes a pointer to a file (and which column within said file), a bin data construct, deadtime and internal conversion corrections, and outputs a tuple of yield (per MeV) and relative error therein.""" y = [] yerr = [1 fil = file(yieldfile,'r') lines = fil.readlines()[1:] fil.close() for line in lines: x = line.split() y += [float(x[2*column-1]),] if not float(x[2*column-1]) == 0: yerr += [float(x[2*column])/float(x[2*column-1]),] else: yerr += [0,] # convert to MeV, apply other correctional factors y = (scipy.array(y)/bins[1])/(livetime*(1-alpha)) yerr = scipy.array(yerr) return (y, yerr) def dcs_energy(yields,eff,flux,t): """ Computes a differential cross-section at one angle as a function of incident neutron energy. Requires calculated yield, efficiency and flux output, as well as the thickness in atoms/barn. """ ys = yields[0] yerr = yields[l] e = eff [0] eerr = eff[1] f = flux[O] ferr = flux[1] # vector arithmetic cs = ((ys/f)/e)/t # this is still relative error cserr = scipy.sqrt(yerr**2 + eerr**2 + ferr**2) return (cs, cserr) def getattenuation(rho,detector thickness,mu): """ Computes the attenuation at a given gamma ray energy, averaged over all detectors in an array. Requires the density rho, an array of the effective thicknesses seen by each detector, and the value mu/rho at the desired gamma ray energy.""" # for the copper runs, rho = 8.96 sum = 0.0 for d in detector_thickness: sum += -(scipy.exp(-2*mu*rho*d)-l)/(2*mu*rho*d) return sum/len(detector_thickness)