137
Precise Measurement of the Photon Directional Asymmetry in the ~np Reaction by David Blyth A Dissertation Presented in Partial Fulfillment of the Requirement for the Degree Doctor of Philosophy Approved Apr 2017 by the Graduate Supervisory Committee: Ricardo Alarcon, Chair Kevin Schmidt Joseph Comfort Barry Ritchie ARIZONA STATE UNIVERSITY May 2017

~np d Reaction - ASU Digital Repository · 2017-06-01 · Precise Measurement of the Photon Directional Asymmetry in the ~np!d Reaction by David Blyth A Dissertation Presented in

  • Upload
    others

  • View
    1

  • Download
    0

Embed Size (px)

Citation preview

  • Precise Measurement of the Photon Directional Asymmetry

    in the ~np→ dγ Reaction

    by

    David Blyth

    A Dissertation Presented in Partial Fulfillmentof the Requirement for the Degree

    Doctor of Philosophy

    Approved Apr 2017 by theGraduate Supervisory Committee:

    Ricardo Alarcon, ChairKevin SchmidtJoseph ComfortBarry Ritchie

    ARIZONA STATE UNIVERSITY

    May 2017

  • ABSTRACT

    This work presents analysis and results for the NPDGamma experiment, measuring

    the spin-correlated photon directional asymmetry in the ~np → dγ radiative capture

    of polarized, cold neutrons on a parahydrogen target. The parity-violating (PV) com-

    ponent of this asymmetry Aγ,PV is unambiguously related to the ∆I = 1 component

    of the hadronic weak interaction due to pion exchange. Measurements in the second

    phase of NPDGamma were taken at the Oak Ridge National Laboratory (ORNL)

    Spallation Neutron Source (SNS) from late 2012 to early 2014, and then again in the

    first half of 2016 for an unprecedented level of statistics in order to obtain a mea-

    surement that is precise with respect to theoretical predictions of Aγ,PV = O(10−8).

    Theoretical and experimental background, description of the experimental apparatus,

    analysis methods, and results for the high-statistics measurements are given.

    i

  • DEDICATION

    I dedicate this thesis to my parents, Barry and Joni, for their never-ending support.

    Mahalo.

    ii

  • ACKNOWLEDGMENTS

    I would like to thank my research advisor, Ricardo Alarcon for constantly working

    to provide excellent research opportunities for myself and others. Furthermore, his

    guidance and perspective has been invaluable throughout my time at Arizona State

    University. I would also like to thank the rest of my graduate committee, Joseph Com-

    fort, Barry Ritchie, and Kevin Schmidt for their thorough and constructive feedback

    during my prospectus and leading up to my defense.

    I must give a broad, sweeping thank you to all who contributed to the NPDGamma

    experiment for their incredible work on this experiment. The level of expertise and

    passion exhibited by the senior physicists is inspiring, and I’m glad that I could come

    along for the ride. I would especially like to thank Nadia Fomin, Jason Fry, Vince

    Cianciolo, David Bowman, Seppo Penttila, Chad Gillis, and Latiful Kabir for the

    positive impact that they made on my work.

    Back here in the desert, I couldn’t have remained sane over the years without the

    support of many friends and family members. I would especially like to thank my

    sister Roberta for looking out for me and being someone I can look up to. Brittany,

    Jason, Bob, Chris, and others, thank you for putting up with me and keeping my

    spirits up. Also, many thanks to my labmates who have proven to be not only great

    sounding boards, but also great friends over recent years: Lauren, Jason, Glenn, and

    Eugene.

    iii

  • TABLE OF CONTENTS

    Page

    LIST OF TABLES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . vii

    LIST OF FIGURES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . viii

    CHAPTER

    1 INTRODUCTION . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1

    2 PARITY VIOLATION AS A PROBE OF FUNDAMENTAL PHYSICS . 3

    2.1 NN Weak Interaction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5

    2.2 Experimental Status . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8

    2.3 The NPDGamma Experiment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13

    3 THE NPDGAMMA EXPERIMENTAL APPARATUS . . . . . . . . . . . . . . . . . 17

    3.1 ORNL Spallation Neutron Source . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18

    3.1.1 Fundamental Neutron Physics Beamline . . . . . . . . . . . . . . . . . . . 20

    3.2 Beam Choppers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23

    3.3 Beam Monitors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25

    3.4 Supermirror Polarizer . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27

    3.5 Magnetic Holding Field . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29

    3.6 Spin Rotator . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29

    3.7 Gamma Detector Array . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33

    3.8 Data Acquisition Systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35

    3.8.1 Hydrogen Target DAQ . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36

    3.8.2 n3He Delta-Sigma DAQ . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37

    3.9 lH2 Cryostat . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37

    4 ASYMMETRY MEASUREMENTS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41

    4.1 General Strategy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41

    4.2 Overview of Hydrogen Measurements. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42

    iv

  • CHAPTER Page

    4.3 Background Subtraction Strategy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43

    4.4 Overview of Background Measurements . . . . . . . . . . . . . . . . . . . . . . . . . . 48

    4.5 Gamma Detector Phosphorescence . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50

    4.6 30 Hz Beam Power Modulation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50

    4.7 Dropped Pulses . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52

    4.8 Hydrogen DAQ Transient Response . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54

    4.9 Chopper Phasing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54

    5 MONTE-CARLO SIMULATION . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56

    5.1 Physics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56

    5.2 Geometry . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60

    6 DATA REDUCTION . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65

    6.1 Detector Signal Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65

    6.2 Asymmetry Definition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 70

    6.3 Beam Power Rejection . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76

    6.4 Pedestal Drift Rejection . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 82

    6.5 Geometric Factors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83

    6.6 Grand χ2 Fit . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 84

    7 DATA SELECTION . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 87

    7.1 Manual Batch Selection . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 88

    7.2 Normalization Segments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 88

    7.3 Data Integrity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 92

    7.4 Spin Rotator Sequencing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 93

    7.5 Beam Asymmetry . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 94

    7.6 Analysis of Reduced Data . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 95

    7.7 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 98

    v

  • CHAPTER Page

    8 SYSTEMATIC UNCERTAINTIES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 101

    8.1 Geometric Factors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 103

    8.2 Beam Power Modulation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 107

    8.3 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 109

    9 CONCLUSION . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 111

    9.1 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 111

    9.2 Outlook . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 114

    REFERENCES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 115

    APPENDIX

    A GARDNER TRANSFORM . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 121

    B DERIVATION OF THE GRAND χ2 SOLUTION . . . . . . . . . . . . . . . . . . . . . . 123

    C DIAGONALIZATION OF THE GRAND χ2 FIT . . . . . . . . . . . . . . . . . . . . . . 125

    vi

  • LIST OF TABLES

    Table Page

    2.1 DDH Couplings and Predictions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7

    4.1 2016 Background Targets . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49

    6.1 Beam Power Modulation Sensitivity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81

    7.1 Hydrogen Batch Cut Statistics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 99

    7.2 Background Batch Cut Statistics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 100

    8.1 Systematic Uncertainties From False Asymmetries . . . . . . . . . . . . . . . . . . . . 110

    8.2 Multiplicative Factors and Systematic Uncertainties . . . . . . . . . . . . . . . . . . 110

    vii

  • LIST OF FIGURES

    Figure Page

    2.1 DDH Meson Exchange Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6

    2.2 EFT Contact Terms and Pion Exchange . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7

    2.3 Benchmark Basis for Experimental Constraints . . . . . . . . . . . . . . . . . . . . . . . 11

    2.4 NPDGamma Asymmetry . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13

    2.5 Neutron-Proton Continuum and Bound States . . . . . . . . . . . . . . . . . . . . . . . 15

    3.1 NPDGamma Experimental Apparatus . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17

    3.2 SNS Accelerator . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19

    3.3 Proton Beam Accumulation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20

    3.4 Supermirror Coating . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22

    3.5 FnPB Spectrum . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24

    3.6 Chopped Spectrum at Upstream Beam Monitor . . . . . . . . . . . . . . . . . . . . . . 25

    3.7 Beam Monitor Concept . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26

    3.8 Polarizing Bender . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28

    3.9 RF Spin Rotator Concept . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30

    3.10 Spin Rotator Sequencing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32

    3.11 Gamma Detector Array Layout . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34

    3.12 Detector With lH2 Cryostat/Target and SR Coil Model . . . . . . . . . . . . . . . 38

    3.13 Target Vessel and Ortho/Para Conversion . . . . . . . . . . . . . . . . . . . . . . . . . . . 39

    4.1 Gamma Detector Response . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43

    4.2 8-Pulse Gamma Detector Readout Sequence . . . . . . . . . . . . . . . . . . . . . . . . . 44

    4.3 Prompt vs. Delayed Energy Discrepancy. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45

    4.4 Dedicated Target Decay Spectrum . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46

    4.5 Resolved Prompt vs. Delayed Energy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47

    4.6 Variation in Detector Phosphorescence . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50

    viii

  • Figure Page

    4.7 30 Hz Signal Phase Portrait . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51

    4.8 Accelerator Dropped Pulse . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53

    4.9 Chopper Phasing Oscilloscope View . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55

    5.1 28Al De-excitation Gamma Spectrum . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59

    5.2 Prompt Energy Seen in CsI Array From Capture on Aluminum . . . . . . . 59

    5.3 Verification of Neutron Cross-Sections in Al . . . . . . . . . . . . . . . . . . . . . . . . . . 60

    5.4 MC Simulation Geometry Overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61

    5.5 Energy Seen by CsI Array vs. Capture Position . . . . . . . . . . . . . . . . . . . . . . 62

    5.6 Energy Seen by CsI Array vs. Capture Position with 100% Ortho-

    hydrogen . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63

    5.7 Energy Seen by CsI Array vs. Capture Position with 100% Parahydrogen 64

    6.1 Θ for Hydrogen Data . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67

    6.2 Θ for 2016 Measurements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 68

    6.3 Symmetric Component Fourier Spectrum . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71

    6.4 Asymmetric Component Fourier Spectrum . . . . . . . . . . . . . . . . . . . . . . . . . . . 72

    6.5 Analyzer Function for Hydrogen Data . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73

    6.6 Analyzer Function for 2016 Data . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74

    6.7 ζ and Θ− ·X(τ) for 2016 Measurements in Fourier Space . . . . . . . . . . . . . 78

    6.8 ζ for Hydrogen Measurements in Fourier Space . . . . . . . . . . . . . . . . . . . . . . . 79

    6.9 α ∗ γ Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 80

    6.10 Pedestal Drift Rejection . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 82

    6.11 An Example of Geometric Factors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 84

    7.1 Normalization Segmentation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 89

    7.2 η Normalization Integral Factors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90

    7.3 Threshold on Beam Monitor Integral . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91

    ix

  • Figure Page

    7.4 Conditional Beam Monitor Distribution . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 92

    7.5 Threshold on SR RMS Current . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 93

    7.6 Beam Asymmetry Cut . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 94

    7.7 Pair Asymmetry Distributions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 96

    7.8 Diagonalized Grand Fit χ2 Metric . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 97

    8.1 Thermal Equilibrium Parahydrogen Fraction . . . . . . . . . . . . . . . . . . . . . . . . . 105

    8.2 Chlorine Asymmetry Measurement . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 106

    8.3 Pair Asymmetry vs. M1 Asymmetry Regression . . . . . . . . . . . . . . . . . . . . . . 108

    8.4 Fit of Composite Target Geometric Factors to Regression Slopes . . . . . . 109

    9.1 DDH Parameter Space Constraints Including This Work . . . . . . . . . . . . . . 112

    9.2 h1π From Experiment and Theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 113

    x

  • Chapter 1

    INTRODUCTION

    This work focuses on the measurement and analysis of the spin-correlated photon

    directional asymmetry in the ~np → dγ radiative capture of polarized, cold neutrons

    on a liquid parahydrogen target in the NPDGamma experiment. The parity violating

    component of this asymmetry, denoted as Aγ,PV , can be unambiguously related to

    the strength of the ∆I = 1 hadronic weak interaction due to pion exchange, and is

    primarily sensitive to neutral weak currents. Since the first measurements in the work

    of Cavaignac et al. (1977), measurement of the parity-violating asymmetry has gone

    without an improvement in precision for four decades. Their work placed an upper

    limit of

    Aγ,PV (~np) = (0.6± 2.1)× 10−7,

    and measurements by the NPDGamma collaboration (Gericke et al. 2011) at Los

    Alamos National Laboratory yielded

    Aγ,PV = (−1.2± 2.1± 0.2)× 10−7

    for the results of the first phase of the experiment. The goal of the second phase is

    to reach an uncertainty in Aγ,PV of ∼ 1× 10−8 from measurements taken at the Oak

    Ridge National Laboratory spallation neutron source in order to obtain a level of pre-

    cision that begins to challenge theoretical predictions. The NPDGamma experiment

    recently concluded measurements, and this work focuses on the analysis of the data.

    In Chapter 2, a theoretical and experimental background will be developed to

    motivate and support the following work, including a conceptual description of the

    1

  • experiment. The experimental apparatus will be described in Chapter 3, as a neces-

    sary foundation for discussing measurements and data reduction. Chapter 4 describes

    measurements taken and strategies for determining Aγ in the presence of background

    signals. The GEANT4-based Monte-Carlo created for the analysis in this work is

    described in Chapter 5. Chapters 6 and 7 go into detail about employed methods of

    reducing and selecting data into a determination of Aγ while minimizing systematic

    effects. Chapter 8 describes the determination of those systematic effects. Finally, a

    conclusion is given in Chapter 9 along with results of the described analysis and an

    outlook for the impact of those results on the physics community.

    2

  • Chapter 2

    PARITY VIOLATION AS A PROBE OF FUNDAMENTAL PHYSICS

    Lee and Yang (1956) hypothesized that, unlike the strong and electromagnetic in-

    teractions, weak interactions may not conserve the parity of a system, where parity

    refers to the symmetry (or antisymmetry) of the state under inversion of all spatial

    coordinates. This hypothesis was inspired (at least in part) by the so-called τ − θ

    puzzle, which was the observation of the precisely identical masses of two apparently

    different particles. The two particles were deduced to be different due to considera-

    tions of angular momentum and conservation of parity in their decays. In light of this

    puzzle, Lee and Yang examined existing experimental evidence to see if parity non-

    conservation could be ruled out as a possibility in the weak interaction, and found

    that such a violation could not. Their 1956 paper proposed that experimentalists

    should actively seek out proof of parity non-conservation in weak interactions. (Par-

    ity non-conservation is referred to throughout this thesis as parity violation, where PV

    and PC are taken to mean “parity-violating” and “parity-conserving”, respectively.)

    Months later, Wu et al. (1957) found evidence of parity violation in the decays of

    60Co.

    The standard model (SM) of particle physics describes interactions of fundamen-

    tal quarks and leptons arising from a mix of local (gauge) symmetry requirements

    within the Lagrangian. In group-theoretical terms, the symmetries are described as

    SU(3) × SU(2)L × U(1), with representations of the groups corresponding respec-

    tively to the strong interaction color charge, weak isospin, and weak hypercharge.

    The SU(3) symmetry applies only to quark fields giving rise to gluon interactions,

    while the subscript on SU(2)L indicates that the symmetry only applies to left-handed

    3

  • fermions. Because angular momentum is an axial vector, a parity transformation ef-

    fectively maps left- and right-handed particles onto each other, and it is therefore the

    explicit handedness of the SU(2) symmetry that gives rise to a maximal violation of

    parity conservation by the associated vector bosons. Of course, nature does not allow

    such an elegant symmetry of physics to go unspoiled, and within the SM we find

    that the Higgs field, which exhibits a finite vacuum expectation value, dynamically

    breaks the SU(2)L×U(1) symmetry, ultimately yielding a pair of massive, electrically

    charged W± bosons, the neutral Z boson, and the neutral, massless photon from the

    electroweak sector.

    The SU(3) theory of quantum chromodynamics (QCD) at low energies gives rise

    to the confinement of quarks in color singlets, i.e. hadrons. Confinement is believed

    to arise from the effective anti-screening of color charge due to gluon-gluon interac-

    tions. This anti-screening also makes the ground state of QCD unreachable by the

    standard method of finding solutions to the interacting theory by treating interac-

    tions as a perturbation of the free, non-interacting theory. Therefore, connecting the

    high-energy theory of QCD to the reality of the structure of every-day matter is very

    difficult. Limited contact has been seen by the success of lattice QCD, where numer-

    ical calculations are made in discrete spacetime (i.e. on the lattice), in such areas

    as predicting the masses of hadrons. Though weak interactions between fermions of

    the SM are quite well understood, in hadronic systems at low energies they become

    inextricable from non-perturbative strong interactions.

    Parity violation, then, becomes a signature for isolating weak interactions in

    hadronic systems. Constructing pseudoscalar products of cartesian vectors, such

    as the ~σ · ~r beta emission asymmetry observed by Wu et al., targets interferences

    between strong (or electromagnetic) and weak amplitudes. However, significant dif-

    ficulties arise in observing hadronic parity violation since the natural scale of the

    4

  • strong/weak interference is GFF2π ∼ 10−7 (Haxton and Holstein 2013). Nevertheless,

    hadronic PV observables provide the opportunity to probe non-perturbative QCD,

    and measurements of PV effects can serve as inputs to effective Lagrangians and help

    to provide model-independent predictions.

    2.1 NN Weak Interaction

    The theoretical interpretation of observed parity violation in non-leptonic flavor-

    changing decays of mesons and baryons has revealed puzzles that remain unresolved.

    For example, anomalously large PV asymmetries in radiative decays of hyperons, as

    well as the well-known ∆I = 1/2 rule (a phenomenological preference of the I = 1/2

    over I = 3/2 isospin channels) in strangeness-changing (∆S = 1) decays, remain

    unexplained by SM symmetries (Ramsey-Musolf and Page 2006). Whether such

    anomalies are particular to strange quark dynamics, or even perhaps new physics

    which fail to live up to the symmetries of the SM, is still unknown. Suppressing the

    role of the strange quark by further exploration of hadronic weak interactions (HWIs)

    in strangeness-conserving (∆S = 0) processes may provide additional insight. The

    most accessible approach is to study nucleon-nucleon (NN) and nuclear interactions.

    In nuclei, significant enhancement of PV effects can be observed. Treating the HWI

    as a perturbation to the strong interaction, near degeneracies in nuclear states with

    opposite parity can give rise to contributions of the form

    � =〈Ψ′|HW |Ψ〉

    ∆E. (2.1)

    However, observations of parity violation in complex nuclear systems that exhibit such

    near degeneracies are theoretically difficult to interpret (with few exceptions), being

    significantly affected by nuclear structure effects. To minimize theoretical ambiguity,

    observables in two-body or at least few-body systems are sought out despite the

    5

  • π, ρ, ω

    PC(Strong)

    PV(Weak)

    N

    N N

    NDDH

    Figure 2.1: DDH meson exchange model - Desplanques et al. (1980) modeled thenucleon-nucleon hadronic weak interaction by exchange of the three lightest availablemesons, the π, ρ, and ω. At one nucleon vertex, a strong (PC) interaction emits ameson which is absorbed by the second nucleon in a weak (PV) interaction.

    experimental challenges.

    In 1980, Desplanques, Donoghue, and Holstein (DDH) developed a framework

    for analyzing PV NN observables (Desplanques et al.) which remains the default

    benchmark. DDH provide a quark-model based solution where the HWI is assumed

    to be mediated by the exchange of a meson (Figure 2.1). The three lightest available

    mesons are included in the model, and the model is parameterized by a combination

    of the exchanged mesons and the isospin change, leading to a total of six couplings,

    h1π, h0,1,2ρ , h

    0,1ω ,

    where the subscript indicates the exchanged meson and the superscript indicates the

    change in isospin. The “best value” predictions and reasonable ranges provided by

    DDH are listed in Table 2.1, along with slightly newer predictions by Dubovik and

    Zenkin (1986); Feldman et al. (1991). The reasonable ranges are intended to cover

    theoretical uncertainties, but are unfortunately quite large. Because PV observables

    from the exchange of neutral pseudoscalar mesons would also be CP violating (Barton

    1961), such mesons are excluded from consideration. A seventh coupling, termed h1′ρ ,

    6

  • Coupling DDH range DDH ”best” DZ FCDH

    h1π 0.0→ 11 4.6 1.1 2.7

    h0ρ −31→ 11 -11 -8.4 -3.8

    h1ρ −0.4→ 0.0 -0.2 0.4 -0.4

    h2ρ −11→ −7.6 -9.5 -6.8 -6.8

    h0ω −10→ 5.7 -1.9 -3.8 -4.9

    h1ω −1.9→ −0.8 -1.1 -2.3 -2.3

    Table 2.1: DDH couplings and predictions - Listed here are the DDH meson-nucleon coupling parameters along with predicted values. DDH give reasonable rangesfor each parameter as well as a “best value”, also known as a “best guess”. Addi-tionally, slightly more recent estimates from Dubovik and Zenkin (DZ) and Feldmanet al. (FCDH) are listed.

    N

    N N

    N

    π

    N

    N N

    Na) b)

    Figure 2.2: EFT contact terms and pion exchange - In a pionful low-energyEFT, a) the dynamics of quarks and heavy mesons are absorbed into the contactterms, and b) pions are included as dynamical degrees of freedom.

    exists in the original DDH work; however that coupling has since been determined

    to be negligible (Holstein 1981). (Notably, however, Phillips et al. argue that a 1/Nc

    QCD expansion indicates that h1′ρ should be no more suppressed than h

    1π.)

    In more recent years, developments in PV effective field theories (EFTs) have

    allowed a more systematic approach to the analysis of low-energy measurements of

    NN parity violation (Zhu et al. 2005; Ramsey-Musolf and Page 2006; Liu 2007).

    7

  • EFT Lagrangians take advantage of the separation of scales at low energies and are

    expressed as expansions in the ratio of scales. A complete set of terms that obey

    the symmetries of the underlying theories make up the effective theory, where the

    coupling constants that absorb the dynamics of high-energy scales can be determined

    phenomenologically. Where the dynamics of the pion can be effectively integrated out,

    a set of local contact terms remain in PV EFTs as illustrated by the left-hand diagram

    in Figure 2.2. In systems where the dynamics of the pion are important, “hybrid”

    EFTs obtain an additional degree of freedom with an explicit dynamical pion, as

    illustrated by the right-hand diagram in Figure 2.2; such a degree of freedom also

    makes it possible to systematically address two-pion exchange. The “hybrid” (pionful)

    formulations then express the NN PV observables in terms of five contact couplings

    equivalent to the dimensionless Danilov (1965; 1971) parameters that correspond to

    the possible S − P transitions at low energy, as well as an additional parameter C̃π6

    for the dynamical pion, which is proportional to the DDH h1π coupling.

    2.2 Experimental Status

    Starting with the work of Tanner (1957) in search of parity violation in (p, α) reso-

    nances of 19F, many have followed the findings of Wu et al. with PV searches in other

    systems. Though the Tanner experiment lacked the sensitivity to observe PV effects,

    a number of experiments since have measured non-zero effects, and their results have

    begun to paint a picture of the HWI. In this section, a number of experiments with

    theoretical analyses that are considered to be on firm ground will be briefly reviewed

    within the framework of DDH, beginning with an overview of past measurements,

    and followed by a summary considering the combined influence of their results on the

    DDH parameter space.

    PV observables from strict two-body interactions are of particular interest. Two-

    8

  • body PV observables are very elusive, however, as their natural size tends to be

    O(10−7). Nevertheless, non-zero effects have been observed, namely in the analyz-

    ing power for scattering of longitudinally-polarized protons on an unpolarized target,

    AL(~pp). AL(~pp) has been measured at a number of energies (Eversheim et al. 1991;

    Nagle et al. 1978; Balzer et al. 1980; Berdoz et al. 2003), yielding complementary con-

    straints within the DDH model. The measurements at different energies are analyzed

    together, placing key constraints on linear combinations of h0,1,2ρ and h0,1ω . (The ~pp

    results have been summarized by Haxton and Holstein, which also incorporated most

    of the other measurements mentioned in this section.)

    Additionally, a pair of measurements of the circular polarization of photons emit-

    ted in the np → dγ reaction, Pγ(np), was made at the Leningrad WWR-M reactor

    facility (Lobashov et al. 1972; Knyaz’kov et al. 1984). However, the first was deter-

    mined to be in error, while the second only placed an upper limit on Pγ(n) of O(10−7).

    Measurements of Pγ(np) are primarily sensitive to h0,2ρ .

    Finally, the PV directional asymmetry of photons emitted in the ~np → dγ reac-

    tion, Aγ,PV (~np), has been measured in two experiments conducted several decades

    apart. The Aγ,PV (~np) observable is dominated by the long-range coupling h1π. The

    first measurement of Aγ,PV (~np) was made at the Institut Laue-Langevin (ILL) reactor

    by Cavaignac et al. (1977). The second measurement of Aγ,PV (~np) was made by the

    NPDGamma collaboration (Gericke et al. 2011) at the Los Alamos Neutron Science

    Center (LANSCE) spallation source; this latter work was the precursor to measure-

    ments at the Oak Ridge National Lab (ORNL) spallation neutron source (SNS) upon

    which this work is based. The results of these two experiments place the upper limits

    Aγ,PV (~np)|ILL = (0.6± 2.1)× 10−7,

    Aγ,PV (~np)|LANL = (−1.2± 2.1± 0.2)× 10−7.(2.2)

    The next section will cover the Aγ,PV measurement in greater detail.

    9

  • Due to the difficulty in detecting PV observables in the two-body systems, larger

    systems have been studied, some of which are considered to have reliable theoreti-

    cal interpretation. For example, the analyzing power for the scattering of polarized

    protons from an alpha target, AL(~pα), has been measured by Lang et al. (1985) at

    46 MeV with a non-zero result. While uncertainties in the theoretical analysis of such

    a measurement exist, that analysis is nonetheless considered to be on firm ground,

    and provides an important constraint on a linear combination of h1π, h0,1ρ , and h

    0ω.

    Additionally, measurements involving flourine isotopes have been interpreted in the

    DDH model with theoretical uncertainties that are considered to be minimal, due to

    the ability to measure the necessary nuclear matrix elements from beta decay rates in

    isobaric analog nuclei. Searches for circular polarization in photons emitted from the

    decay of an excited state of 18F, Pγ(18F), were undertaken by several groups (Barnes

    et al. 1978; Ahrens et al. 1982; Bini et al. 1985; Page et al. 1987). Measurements

    of this kind are sensitive to isovector couplings (primarily h1π) and are therefore very

    interesting results. The measurements, however, while having reached significant sen-

    sitivities, did not observe parity violation. Finally, parity violation has been observed

    in the directional asymmetry in the emission of photons from an excited state of 19F,

    Aγ(19F), by two groups (Adelberger et al. 1983; Elsener et al. 1987). The Aγ(

    19F )

    observable is sensitive to couplings similar to those of AL(~pα) measurements.

    Figure 2.3 shows the effect of the most constraining classes of experiments on a

    convenient basis of isoscalar and isovector couplings chosen by Haxton and Wieman

    (2001). The one-standard-deviation (1σ) constraints are shown with slight expansion

    due to allowing uncorrelated variations of h1ρ, h2ρ, h

    0ω, and h

    1ω within the DDH reason-

    able ranges, similar to the approach taken by Haxton and Wieman (2001); Haxton

    and Holstein (2013). Additionally, a relatively recent lattice QCD result is shown.

    The result by Wasem (2012) shows uncertainties that are a combination of statistical

    10

  • 1ω-0.18h

    1ρ-0.12h

    1πh

    0.2− 0 0.2 0.4 0.6 0.8 1 1.2 1.46−10×

    )0 ω

    +0.

    7h0 ρ

    -(h

    0.5−

    0

    0.5

    1

    1.5

    2

    2.5

    3

    3.56−10×

    )αp(LAp)p(LAF)18(γPF)19(γA

    LQCD (Wasem)DDHDZFCDH

    Figure 2.3: Benchmark basis for experimental constraints - Shown here is aplot in the style of Haxton and Wieman; Haxton and Holstein showing the 1σ con-straints placed by several classes of experiments considered to be on firm theoreticalground on a convenient DDH parameter basis. Additionally, theoretical predictionsfrom Table 2.1 are shown as points, and a relatively recent lattice QCD result is givenin dashed lines.

    11

  • and systematic. Wasem does not include renormalization of the bare PV operators,

    EFT extrapolation to a physical pion mass (mπ ∼ 389 MeV is used), or disconnected

    diagrams. However, the author expects systematic errors that are small compared

    to the statistical uncertainties. There is disagreement in the theoretical literature

    about whether the isovector couplings, particularly h1π, should be suppressed. The

    Pγ(18F ) constraint shown in Figure 2.3 shows an upper limit that some consider dif-

    ficult to accommodate theoretically. At the same time, recent work by Phillips et al.

    (2015), for example, argues that a 1/Nc expansion of QCD indicates a suppressed

    h1π . (0.8± 0.3)× 10−7, which tends to agree with the Pγ(18F) results. On the other

    hand, Lobov (2002) give an estimate of h1π = 3.4× 10−7 from QCD sum rules that is

    relatively intermediate with respect to predictions by Desplanques et al. and Phillips

    et al..

    Other experimental efforts require some consideration as well. As far as current

    efforts go, the n3He experiment has recently finished taking data at the ORNL SNS,

    and analysis is ongoing. The n3He experiment measures the PV directional asym-

    metry of outgoing particles in the reaction ~n +3 He → p + t + K.E. (Gericke 2016).

    Additionally, a recent upper bound on PV neutron spin rotation (NSR) in 4He has

    been set by Snow et al. (2011) at NIST, and the group intends to improve the ex-

    perimental sensitivity in the near future. Both experiments provide opportunities

    to constrain the HWI through coupling combinations independent of those shown in

    Figure 2.3. Notably, PV nuclear anapole moment measurements of 133Cs (Wood et al.

    1997) and 205Tl (Vetter et al. 1995) are in stark disagreement with the bounds found

    by the experiments shown in Figure 2.3, as well as with each other. Those results

    have not been included here due to their difficult theoretical interpretation.

    12

  • n p

    θγ

    d

    n p

    θγ

    d

    Figure 2.4: NPDGamma asymmetry - An effective parity transformation ismade by alternating the spin of the neutron beam. For θγ the angle between theneutron spin and the emitted photon momentum, the differential cross section sees aspin-correlated modulation of magnitude Aγ,PV cos θγ.

    2.3 The NPDGamma Experiment

    This thesis is particularly concerned with the determination of the Aγ(~np) asym-

    metry (henceforth simply Aγ), by analysis of recent data from the NPDGamma ex-

    periment. Aγ (illustrated in Figure 2.4) is taken to be a general spin-correlated

    asymmetry in the direction of photons emitted from ~np capture. For clarity, Aγ,PV

    and Aγ,PC are used to denote the PV and PC components of the asymmetry. The dif-

    ferential cross section for the emission of a photon into solid angle dΩ can be written

    as

    dΩ∝ 1 + Aγ,PV k̂γ · σ̂n + Aγ,PC k̂γ ·

    (σ̂n × k̂n

    ), (2.3)

    where the PV part comes from a pseudoscalar combination of cartesian vectors, and

    the PC part comes from a scalar combination. Equation 2.3 serves to explicitly

    define the two asymmetries. The goal of the NPDGamma experiment is to measure

    Aγ,PV down to a sensitivity of 1 × 10−8, representing an order of magnitude better

    13

  • sensitivity than has been achieved. Though measurement of parity violation is the

    primary motivation, the PC component will also be extracted. To isolate the physics

    asymmetries and suppress instrumental effects, the asymmetries are extracted by

    rapidly alternating the spin orientation of a polarized neutron beam incident on a

    proton target. The photons emitted from the reaction are detected and their intensity-

    spin correlation as a function of dΩ is measured. From that intensity distribution,

    the orthogonal PV and PC correlations are extracted. In practice, the finite detector

    size must be incorporated, and false asymmetries must be carefully addressed. The

    experimental setup and the data reduction approach used in this work are described

    in chapters 3 and 6, respectively.

    In order to understand the relationship of Aγ,PV to the meson exchange picture,

    let us work from experiment to theory. Considering isospin as a valid symmetry,

    the neutron and proton are different states of the same particle. From Fermi-Dirac

    statistics, the deuteron must have a wavefunction that is antisymmetric under nucleon

    exchange. Furthermore, the deuteron is an isospin singlet since pp and nn bound

    systems are not found. Particle exchange symmetry, then, allows only S states in a

    spin triplet configuration, and P states in a spin singlet configuration. Because the

    3S1, I = 0 state is energetically favorable, the spin triplet state is the ground state

    of the deuteron. The parity of the ground state, being determined entirely by the

    orbital angular momentum in the NN system, is (−1)0 = +1, which is even parity.

    Treating the weak interaction at low energies as a perturbation to the strong

    interaction, as is naturally done by assuming definite parity of the ground state, an

    admixture of opposite-parity (irregular) P states into the (regular) ground state arises.

    Admixtures also arise for the S-wave scattering states, as shown in Figure 2.5 (in the

    style of Byrne 2013). In the figure, unhindered electric and magnetic matrix elements

    are represented by arrows, where the regular transition occurs from the 1S0 partial

    14

  • 1S0 I=1 3P0 I=11P1 I=03P1 I=13S1 I=0

    3S1 I=0 3P1 I=1 1P1 I=0

    m(0)e(1)

    e(0)

    Figure 2.5: Neutron-proton continuum and bound states - HWIs contributeadmixture of opposite-parity states onto low-energy S states. The unhindered tran-sitions are shown with arrows with e(J) and m(J) indicating electric and magnetictransitions starting from a J total spin state.

    wave. According to Gari and Schlitter (1975), the PV asymmetry comes about from

    interference between the m(0) regular transition and the e(1) irregular transitions:

    Aγ,PV = −√

    2e(1)

    m(0). (2.4)

    (Alternatively, circular polarization in photons emitted from unpolarized neutron

    capture comes from interference between m(0) and e(0) transitions.) Thus, we can

    see that Aγ,PV is sensitive to admixtures of3P1 states with I = 1 onto the scattering

    and bound 3S1, I = 0 states, meaning that Aγ,PV is sensitive to ∆I = 1 weak

    interactions.

    If we consider the low-energy charged weak Hamiltonian in current-current form,

    we have

    H cw =GF√

    2J+µ J

    µ+ + h.c., (2.5)

    where

    J+µ = cos θC ūγµ (1 + γ5) d+ sin(θC)ūγµ (1 + γ5) s. (2.6)

    15

  • For ∆S = 0 interactions, a ∆I = 0, 2 component appears multiplied by cos2 θC ,

    while a ∆I = 1 component appears multiplied by sin2 θC . Thus, the Cabibbo angle

    suppresses the charged current in the Aγ,PV observable. The neutral weak current

    gives no such suppression, so the NPDGamma asymmetry is expected to be primarily

    sensitive to neutral currents. The ∆I = 0 admixture dependence also indicates that

    the asymmetry is dominated by the exchange of long-range pions. The calculated

    relationship between the NPDGamma asymmetry and the DDH couplings is (Gericke

    et al. 2011)

    Aγ,PV = −0.1069h1π − 0.0014h1ρ + 0.0044h1ω, (2.7)

    confirming that Aγ,PV is dominated by pion exchange.

    16

  • Chapter 3

    THE NPDGAMMA EXPERIMENTAL APPARATUS

    The NPDGamma experiment consists of a dedicated apparatus operated for a

    high-statistics result at the ORNL SNS. As the most intense pulsed-mode spallation

    neutron source, the ORNL SNS delivers cold neutrons on the Fundamental neutron

    Physics Beamline (FnPB) at a rate high enough to reach or exceed the goal of an

    unprecedented 10−8 measurement of the σ̂n · k̂γ parity-violating correlation in the

    ~np→ dγ reaction: a measurement that has been heavily statistics limited for decades.

    The FnPB was designed to maximize cold neutron fluence and remove line-of-sight

    between the experiment room and the neutron moderator. By removing line-of-sight,

    fast background radiation is minimized (Fomin et al. 2014). The apparatus, illustrated

    in Figure 3.1, polarizes the incoming neutron pulses and alternatingly flips their spin

    on a pulse-by-pulse basis. This alternating spin sequence allows the measurement

    Spin rotatorPolarizerFrom spallationsource and LH2moderator

    n p d

    Chopper1

    Chopper2 Shutter

    Guide

    Lead shieldingParahydrogentarget

    Ortho/paramonitor

    Beam monitor

    "Racetrack" holding field coils

    CsI crystal detectors

    ŷ

    Figure 3.1: NPDGamma experimental apparatus - This figure gives a cartoonoverview of the components of the NPDGamma apparatus. The green bar illustratesthe neutron beam and intensity as it travels from left to right in the ẑ direction. Thespins are polarized up in the ŷ direction, leaving the x̂ direction at beam left into thepage.

    17

  • to be insensitive to detector gain, gain drifts, and slow background variations. The

    majority of the neutrons capture on a liquid parahydrogen target, and the emitted

    gammas are detected by an array of 48 thallium-doped CsI scintillation detectors.

    These detectors are arranged symmetrically about the beam axis, and cover a solid

    angle of approximately 3π sr. The symmetry of the detectors allows the cancellation

    of beam power fluctuations that would otherwise complicate error analysis. The

    asymmetry of the target is determined by the correlation of power deposited into the

    CsI detectors with the spin orientation.

    The apparatus was operated for hydrogen target measurements from 2012-2014. In

    2015, it was determined that additional measurements of backgrounds were required,

    and the apparatus was reinstalled for measurements in the first half of 2016. In

    this thesis, these two distinct instances of the experiment are typically referred to as

    hydrogen measurements and 2016 background measurements, respectively.

    3.1 ORNL Spallation Neutron Source

    The SNS is a pulsed neutron source that uses a proton beam to release neutrons

    from Hg nuclei in a process known as spallation. Of critical importance in spallation

    sources is the production of large fluxes of cold, thermal, and epithermal neutrons

    while maximizing the ability to use neutron Time-Of-Flight (TOF) information to

    separate energies. The spallation process is driven by 1 GeV protons impacting the

    mercury target, producing ∼ 30 fast neutrons per incident proton. The fast neutrons

    are then moderated (slowed) by a hydrogenous medium with a particular composition

    and temperature (depending on the beamline) to energies of order meV for neutron

    scattering applications and fundamental neutron studies.

    The proton beam (Figure 3.2) starts with H− ions selected from a plasma (Stockli

    et al. 2010), which are then injected into a 335 m-long linac at 402.5 MHz. The protons

    18

  • from linac

    accumulation ring

    to spallation target

    mer

    ge

    Figure 3.2: SNS accelerator - The SNS proton accelerator consists of a 1 GeV linacwhich produces bunches that “wrap” around an accumulation ring. The accumulatedproton bunch is then extracted to the mercury spallation target. The above figure isa cartoon of the physical layout from an overhead perspective.

    are then accelerated to 1 GeV. A chopper selects a 695 ns macro pulse of protons at

    1.059 MHz (Jones 2016). The chopper configuration is designed to fascilitate the

    “wrapping” of macro pulses around an accumulator ring. The negative charge of the

    accelerated bunch allows a dipole magnet to merge the incoming linac pulse into the

    positively charged stored bunch as shown in Figure 3.3. At the point that the two

    bunches merge, the protons pass through a carbon foil that strips the electrons from

    the H− ion. The accumulator ring stores and shapes the proton bunch, retains the

    695 ns envelope, and accumulates protons over 1060 macro pulses (∼1 ms fill time).

    The ring is filled and subsequently extracted into a spallation target at 60 Hz.

    The spallation target is ≈1.4 m3 of mercury circulating in a loop at ≈325 kg s−1 in

    a stainless steel vessel (Fomin et al. 2014). The flowing mercury along with flowing

    water in an outer vessel cools the target. The cooling is necessary because & 850 kW

    of sustained power is deposited into the mercury target by the accelerated protons.

    19

  • stored

    bunch

    H - bunch

    merged bunch

    dipole magnet

    foil

    Figure 3.3: Proton beam accumulation - The H− bunch from the SNS linacmerges with the accumulation ring using a dipole magnet and a carbon stripping foil.

    An array of moderators sit above and below the spallation target. These moderators

    are viewed by the neutron beamlines.

    3.1.1 Fundamental Neutron Physics Beamline

    The FnPB views the bottom downstream 20 K supercritical parahydrogen neutron

    moderator through a 10 cm (width) by 12 cm (height) curved window. The moderator

    has an average thickness of 55 mm, and is surrounded by a 20 mm light-water premod-

    erator coupled to a beryllium reflector, which serves to contain neutrons and increase

    the flux into the beamline at the expense of a longer time response. Parahydrogen

    moderators are known as non-thermalizing leakage moderators because their moderat-

    ing properties and cross section tend to disappear below energies where the 14.5 meV

    para-ortho upconversion is the primary energy loss mechanism. Below ≈50 meV, nu-

    clear spin coherence begins to remove the upconversion moderating mechanism (Kai

    et al. 2004). This type of moderator has the advantage of producing high peak intensi-

    ties for cold neutrons. While the moderator is designed to maintain reasonable timing

    20

  • characteristics, the sharp bunches delivered to the spallation target are expanded out

    to tens of microseconds by the time the cold neutrons escape the moderator (Iverson

    2016).

    Viewing the moderator is a cold neutron guide that serves multiple purposes

    beyond transporting cold neutrons to the experiment. The guide is 10 cm (width)

    by 12 cm (height) rectangular, consisting of neutron optics which have the ability to

    totally reflect if the transferred neutron momentum is below a threshold. The guide

    is also curved, with a radius and length suited for removing line of sight and higher

    energy neutrons / gamma radiation.

    To understand the neutron optics, for unbound neutrons the time-independent

    Schrödinger equation can be rewritten as a Helmholtz equation,

    ∇2Ψ(~r) + k2(~r)Ψ(~r) = 0. (3.1)

    A solution to this equation is

    k = k0n(~r) (3.2)

    where k0 is the wave number of the neutron in free space, and

    n(~r) =√

    1− V (~r)/E, (3.3)

    as discussed in depth in Byrne (2013). For materials presenting positive bulk nuclear

    potentials to low-energy neutrons, refractive indices less than that of free space are

    typically seen, providing the opportunity for total external reflection of neutrons: a

    useful feature for guiding the particles.

    For neutron wavelengths long compared to the interaction between neutron and

    nuclei within a medium, the neutron wavefunction satisfies equation 3.1 with V (~r) =

    VF , a uniform nuclear potential derived from the Fermi contact pseudopotential given

    by

    VF =2π~2Nbmn

    , (3.4)

    21

  • n1n2

    substrate

    Figure 3.4: Supermirror (SM) coating - A SM consists of a multilayer coatingof alternating refractive indices that decrease in thickness deeper into the coating.This produces a superposition of Bragg reflection peaks which are further broadenedby random variation in the layer thicknesses.

    where N is the mean number of scattering nuclei per unit volume, b is the mean

    scattering length observed under the assumption of an infinitely massive nucleus, and

    mn is the rest mass of the neutron. A neutron in the zero-energy limit, as is applicable

    for a cold neutron source, behaves as if interacting with a continuous medium. Since

    scattering lengths are known experimentally to be typically positive, refractive indices

    will usually be less than one. The result is that a cold neutron incident upon such a

    medium at glancing angles less than

    θc = sin−1(√

    VF/E)

    '√VF/E

    (3.5)

    will be totally externally reflected.

    For natural nickel, a particularly good cold neutron reflector,

    θc,Ni = 1.73 mrad Å−1 × λ0,

    22

  • where λ0 is the free neutron wavelength (Fomin et al. 2014). In order to accept a

    larger phase space, the neutron guide improves upon this benchmark limit of total

    external reflection by using multiple layers of alternating materials on its surface pro-

    ducing carefully-designed Bragg scattering. The layers are of varying thickness in an

    arrangement known as a supermirror (SM), designed to overlap multiple Bragg peaks

    that are themselves broadened by randomization of the layer thickness, effectively

    increasing θc (Figure 3.4). SMs are characterized by their effective θc in a ratio to

    that of single-layer natural nickel in a value known as m. The FnPB uses SM coatings

    within the guides that have m values up to 3.8.

    3.2 Beam Choppers

    The NPDGamma experiment operates by alternating the spin orientation of the

    pulsed neutron beam on a pulse-by-pulse basis. In order to do so, the experiment

    uses an RF spin rotator (section 3.6), which is configured to efficiently flip the spin

    of the pulse for a single neutron wavelength at any given instant in time. As such,

    the neutron beam at any given time must have a narrow spectrum of wavelengths.

    Though parahydrogen moderators are non-thermalizing, such moderators still pro-

    duce a neutron energy distribution that is nearly Maxwellian, with a higher effective

    temperature (Brun 1997). Given the distribution, a long-wavelength tail exists (Fig-

    ure 3.5) that will be overlapped down the beamline by faster neutrons from the next

    spallation pulse.

    On the short-wavelength end of the spectrum, neutrons with wavelengths shorter

    than ∼0.4 Å will rapidly depolarize by up converting the molecular parahydrogen in

    the capture target. The neutrons in this short-wavelength range come primarily from

    leakage of ”slowing-down” neutrons in the moderator (Byrne 2013; Iverson 2016).

    These higher-energy neutrons, then, should also be removed from the beam.

    23

  • Figure 3.5: FnPB spectrum - The neutron spectrum is shown in blue, alongwith a calculation in red. The long-wavelength tail that comes from a near Maxwell-Boltzmann distribution of neutron energies must be truncated to avoid overlap withsubsequent pulses. On the short-wavelength end of the spectrum, the faster neutronsare also truncated to avoid depolarization from up converting parahydrogen in thetarget.

    In order to truncate the neutron spectrum, a pair of choppers synchronized to

    the 60 Hz pulses are located on the beamline at 5.5 m and 7.5 m downstream of the

    moderator (Fomin et al. 2014). The choppers consist of rotating carbon fiber disks

    that are coated in a material containing a concentration of 10B of at least 0.13 g cm−2.

    The upstream and downstream choppers have angles cut out of them that are 131◦

    and 167◦ respectively. The distance from the chopper axes to the center of the neutron

    beam is 25.0 cm.

    The choppers are phased appropriately to work together to eliminate fast neutrons

    and slow neutron overlap. The neutron spectrum seen by the upstream beam monitor

    24

  • wavelength (Å)0 5 10 15 20 25 30 35 40

    Main chopped spectrum ~3-6 Å

    Wraparound neutrons

    Figure 3.6: Chopped spectrum at upstream beam monitor - Shown here isthe neutron spectrum seen by the upstream 3He beam monitor. Due to the 1/v lawfor the capture within the beam monitor, the spectrum seen in this plot is artificiallyweighted by the wavelength.

    described in section 3.3 is shown in Figure 3.6. As shown, small windows of long-

    wavelength overlap, or “wraparound” neutrons can still be seen, though at very small

    levels.

    3.3 Beam Monitors

    Of significant importance to the experiment is the ability to determine the relative

    neutron beam flux (where flux refers to the count per unit time over the entire area of

    the beam) into and out of the target. For that purpose, current-mode neutron detec-

    tors (or beam monitors) were commissioned and installed upstream and downstream

    of the experiment. These detectors have been used to perform neutron polarimetry

    (Musgrave 2013), studies of orthohydrogen concentration within the hydrogen target

    (Barrón-Palos et al. 2011; Grammer 2015), a new parahydrogen cross-section mea-

    surement (Grammer et al. 2015), and analysis of neutron beam flux modulation into

    25

  • Figure 3.7: Beam monitor concept -The NPDGamma neutron beam monitorscontain helium and nitrogen gas to pro-duce ionization from cold neutrons. Thecartoon to the right shows a side-on crosssection of the detector. Three parallel pla-nar electrodes along with the aluminumhousing define the electric field inside theionization chamber. Two of the electrodesare held at positive high voltage, and thecenter electrode is a pseudoground inputto the preamplifier. The aluminum hous-ing is also held at ground, though it is notread out and current to the housing is lost.Ionization comes from the conversion ofneutrons to t+p+K.E., which ionizes he-lium (and nitrogen) gas for downstream(upstream) monitors.

    + +

    Signal

    Beam in

    t p

    the hydrogen and background targets for the purpose of NPDGamma asymmetry

    determination.

    Several such beam monitors were used throughout different stages of the NPD-

    Gamma experiment, but within this document monitors known as M1 and M4 are of

    interest, which were used as the upstream and downstream beam monitors, respec-

    tively. Each monitor is composed of an aluminum ionization chamber approximately

    4 cm thick with a closely-integrated preamplifier, and contain some level of 3He gas

    for the conversion of neutrons into charged particles via the reaction

    n+3 He→ p+ t+ 764 keV,

    which dominates the cross-section for low-energy (below eV) incident neutrons due

    to a nearby resonance of 4He and the suppression of gamma production (Gillis 2006).

    The conceptual operation of the beam monitors can be seen in Figure 3.7. Three

    parallel planar electrodes, along with the aluminum chamber, define the electric field

    within the gas volume. 3He gas at a partial pressure of 0.03 atm, which absorbs

    about 3.7% of 4 Å neutrons, is contained in M1. The overall He pressure in M1

    26

  • is increased to 0.5 atm with 4He, and another 0.5 atm of N2 facilitates ionization

    and dielectric strength (Gillis 2006). M4 contains primarily 3He gas, which tends to

    absorb a significant portion of the remaining neutrons since no neutrons are needed

    downstream (Musgrave 2013).

    3.4 Supermirror Polarizer

    As discussed in 3.1.1, a convenient way of guiding neutrons is through reflection

    at material boundaries. Total external reflection occurs at or below a critical glancing

    angle θc, which is enhanced by layering alternating materials to generate reflections

    caused by interference from the material boundaries at different depths. This multi-

    layer coating, when created in such a way as to spread out the Bragg scattering and

    effectively extend θc, constitutes a SM coating.

    In equation 3.4, the potential due to magnetic interactions is ignored. Such scat-

    tering is incoherent in paramagnetic materials due to the random orientation of atomic

    magnetic moments. However, in ferromagnetic materials, magnetic domains arise be-

    low the Curie temperature that produce magnetic fields much larger than that of any

    individual atom (Byrne 2013). This gives rise to optical birefringence with respect to

    neutron spin, made clear if the neutron refractive index is modified for ferromagnetic

    materials:

    n± =

    √1− VF ± µnB

    E, (3.6)

    where µn is the neutron magnetic moment, B is the magnitude of the magnetic

    field within the ferromagnetic domain, and the upper and lower signs represent the

    refractive indices for neutron spin antialigned and aligned with the domain.

    Combining the concepts of SM coatings and birefringence leads to the possibility

    of a neutron reflector with the ability to select a certain neutron spin orientation

    within a range of neutron momentum transfer if the SM substrate is loaded with

    27

  • Figure 3.8: Polarizing bender - The NPDGamma SMP is a polarizing benderconsisting of an array of curved polarizing blades. Neutrons with the desired spinorientation travel paths between the blades, reflecting along the surfaces. Neutronswith the opposite spin are absorbed into the borofloat glass substrate. The lengthand curvature of the SMP is such that no line-of-sight exists from one end to theother.

    a neutron absorbing isotope. If the SM layers are created from alternating layers

    of ferrous material (with saturated magnetization aligned by an external field) and

    (typically) silicon compounds, the coating can be tuned such that neutrons with one

    spin orientation see alternating refractive indices, exhibiting reflection coefficients

    near unity, while neutrons with the opposite spin orientation see very small changes

    in refractive index, exhibiting correspondingly small reflection coefficients (Petukhov

    et al. 2016). Such a mirror is known as a supermirror polarizer (SMP).

    The SMP used on the FnPB for NPDGamma was comissioned by Swiss Neutron-

    ics, and is a polarizing bender (Figure 3.8) made up of an array of 45 40 cm-long

    thin blades each with a radius of curvature of 9.6 m. Each blade consists of Fe/Si

    SM layer pairs on top of 0.3 mm borofloat glass with enriched 10B (Balascuta et al.

    28

  • 2012; Fry 2015). This SMP has a 10 cm-by-12 cm active cross section (matching that

    of the beamline), and the blades are oriented such that the neutrons are polarized

    in the ŷ direction. The radius of curvature and the length dictate that no neutron

    can pass through the bender without interacting with a polarizing mirror surface.

    To saturate the Fe layers, the SMP is surrounded by 12 pairs of NdFeB permanent

    magnets, producing an applied field within the SMP in excess of 300 G. Outside of

    the SMP, a compensation magnet was installed specifically to cancel the magnetic

    field gradients at the experiment due to the SMP magnets (Balascuta et al. 2012).

    The performance of the polarizer is good, measured at an average of 94% polariza-

    tion at the target for the neutron wavelengths used for asymmetry analysis (Musgrave

    2014). The polarization was measured a number of times throughout the experiment,

    and the results are consistent.

    3.5 Magnetic Holding Field

    As shown in Figure 3.1, after exiting the SMP, the neutrons enter a magnetic

    holding field defined by a set of racetrack coils. There are four such coils stacked

    vertically to produce a field in the ŷ direction. The purpose of this field is two-fold:

    to maintain polarization of the beam as the neutrons traverse the distance to the

    target, and to provide an appropriate Larmor precession for the RF spin rotator

    described in section 3.6. The optimal holding field was determined to be about 9.4 G.

    3.6 Spin Rotator

    The pulse-by-pulse alternating neutron spin flip is achieved in this experiment

    through the application of nuclear magnetic resonance in a device referred to here as

    a resonant RF spin rotator (SR) as termed by Seo et al. (2008). The SR consists of

    a magnet (section 3.5) to produce a uniform, time-independent magnetic field ~B0 in

    29

  • ŷ

    Figure 3.9: RF spin rotator concept - Magnetic fields involved in the spin rota-tion are the holding field caused by the red racetrack coils, and the oscillating fieldof the illustrated coil.

    the direction of the incident neutron polarization surrounding a coil driven by an RF

    signal (in this case at∼30 kHz) that produces a linear oscillating field orthogonal to ~B0

    (see Figure 3.9 for visualization). The time-independent field establishes a resonance

    with respect to the frequency of the oscillating field for flipping the neutron spin.

    The conceptual operation of the SR is most easily understood from a semi-

    classical approach. Consider a neutron propagating in the ẑ direction through a

    time-independent field in the same direction as the neutron polarization, ŷ, as the

    particle begins to propagate through the coil. Also assume for now that the linear

    oscillating field produced by the coil is instead a field with constant magnitude that

    rotates at a given rate in the ẑ, x̂ plane. Throughout the propagation, ~B0 defines

    (along with the gyromagnetic ratio of the neutron, γn) a Larmor precession of the

    neutron at frequency

    ωL = γnB0. (3.7)

    30

  • As the neutron begins to propagate through the rotating field, consider a frame of

    reference that rotates along ŷ at frequency ωL. If the angular frequency of the rotating

    field matches the Larmor frequency, the effective total magnetic field in this frame

    of reference is a constant vector in the plane orthogonal to ŷ. The effect is that the

    expectation value of the neutron spin precesses about this vector at a rate dependent

    on the magnitude of the rotating field. If the field magnitude is tuned such that

    the neutron spin rotates 180◦ by the time the neutron exits the coil, the spin of the

    neutron will be efficiently flipped.

    The probability of spin transition, derived quantum mechanically in Byrne (2013),

    is

    W (t, ω) =(γnB1)

    2 sin2[t2

    √(ω − ωL)2 + (γnB1)2

    ](ω − ωL)2 + (γnB1)2

    , (3.8)

    where t is the time of propagation through the oscillating coil, ω is the frequency of

    the rotating field, and B1 is the magnitude. This has the form of a Breit-Wigner

    resonance multiplied by a function oscillating in time. If we assume that the rotating

    field is perfectly on resonance (ω = ωL), perfect spin-flip efficiency can be obtained

    for

    t =nπ

    γnB1, n odd. (3.9)

    In the actual SR, the field within the coil is, as stated, a linear oscillating field. Such

    a field can be considered a superposition of two rotating fields, one rotating at ω

    and the other at −ω. Since the former is on resonance, and the latter is very far off

    resonance for realistic B0 and B1, equation 3.8 holds approximately true for the actual

    linear oscillating field. The degree to which the effective resonance changes under the

    assumed equivalence of a rotating and linear oscillating field is called the Bloch-Siegert

    shift, and becomes important in experiments that take advantage of the resonance

    to measure magnetic moments. In this experiment, however, the actual value of the

    31

  • (s)t0 0.02 0.04 0.06 0.08 0.1 0.12

    curr

    ent (

    arb)

    0.04−

    0.03−

    0.02−

    0.01−

    0

    0.01

    0.02

    0.03

    0.04

    0 0.02 0.04 0.06 0.08 0.1 0.12

    flux

    (ar

    b)

    0.2

    0.4

    0.6

    0.8

    1

    1.2

    1.4

    1.6

    1.8

    2

    2.2

    Figure 3.10: Spin rotator (SR) sequencing - Shown in blue is the measuredcurrent delivered to the RF SR. The ramping is a 1/TOF function for optimal spin-flip efficiency as the wavelength varies. In red, the approximate neutron flux at the SRis extrapolated by shifting the neutron capture signal for the purpose of visualizingalignment between the neutron flux and SR operation.

    resonance frequency is not important, and parameters are simply tuned to maximize

    the transition probability in equation 3.8.

    Furthermore, at least for the case of a rotating field, the kinetic energy of the

    neutron that undergoes such a rotation remains unchanged in the case where the

    spatial boundaries into and out of the rotating field are infinitely sharp. This has

    been shown to be the case by Golub et al. (1994), even though the potential energy

    of the system changes from a spin flip due to interaction with the time-independent

    field.

    Because the propagation time through the SR coil depends on the neutron energy,

    the SR is driven at a time-dependent amplitude to efficiently flip the full chopped

    neutron spectrum. The propagation time is proportional to the TOF from the moder-

    32

  • ator, so to meet the condition of equation 3.9, the amplitude has the form of 1/TOF.

    The measured current delivered to the RF SR for a set of eight consecutive pulses

    during hydrogen measurements can be seen in Figure 3.10, along with an overlay of

    the neutron flux. In the figure, the ↓↑↑↓↑↓↓↑ spin sequence used for hydrogen mea-

    surements is exhibited, which has the effect of cancelling up to second-order drifts in

    the signal. For 2016 background measurements, this sequence was changed to a 30 Hz

    alternating sequence.

    For this experiment, two different oscillating coils were used, one for the hydrogen

    measurements, and one when the experiment was set back up for background mea-

    surements in 2016. Each coil is contained in a 6061 Al alloy housing with 0.5 mm

    beam windows for hydrogen measurements, and 0.031” for 2016 measurements. Po-

    larimetry was performed to determine the spin-flip efficiency for both measurements,

    yielding a value ≈ 97% for hydrogen measurements (Musgrave 2014), and nearly

    100% for 2016 measurements.

    3.7 Gamma Detector Array

    Gamma rays emitted from the capture of neutrons were detected by using an array

    of 48 thallium-doped CsI scintillation detectors. Each detector is made up of a pair

    of two CsI (Tl) crystals with a total volume of 152 x 152 x 152 mm3, as described

    by Gericke et al. (2005). This volume enables the capture of 84% of the energy of an

    average 2.2 MeV γ-ray incident on the center of the volume, with 11% leaking out

    of the rear face, 3% backscattered from the front face, and the remaining 2% from

    the sides. The detectors are arranged in 4 rings of 12, forming an approximately

    circular-cylindrical array covering a solid angle ≈ 3π sr as seen in Figure 3.11.

    Each scintillator volume comprised of 2 individual crystals is viewed by a single

    Vacuum PhotoDiode (VPD). VPDs were chosen for their large detection area, linear

    33

  • Beamn

    36-4724-3512-230-11

    011

    Figure 3.11: Gamma detector array layout - Illustrated here is the basic layoutand geometry of the CsI gamma detector array. The center-of-capture for the capturetarget is approximately centered in the array, and the polarized neutron beam entersfrom the left. The detectors are numbered in the scheme indicated above, and thecolors of the CsI blocks are illustrative of the effect of an up-down PV asymmetry onthe power deposited into the crystals.

    response, and insensitivity to stray RF fields. High-voltage batteries hold the VPD

    cathodes at a bias ≈− 300 V. The VPD anode currents are amplified by solid state

    electronics that drive coaxial transmission lines carrying the signal away from the

    array. Aluminum housings encase each crystal volume, VPD, and amplifier; the

    housings are electrically isolated from the support structure and each other to avoid

    ground loops. The housings are supported by an aluminum frame covered in 6Li2CO3-

    loaded silicon-matrix sheets. The same sheets are used to shield the inside of the array

    from scattered neutrons.

    The signals that reach the electronics rack where the data acquisition system

    (DAQ) resides are then further amplified by active, 2nd-order, low-pass Bessel filters

    with a cutoff frequency ≈5 kHz. The signals then feed into a variable-gain amplifier

    before being digitized by the DAQ in order to utilize the digitizer’s full bit depth.

    34

  • 3.8 Data Acquisition Systems

    Aside from using different SR coils, the difference between the hydrogen measure-

    ments and the background measurements taken in 2016 is primarily in the data ac-

    quisition system (DAQ). For hydrogen measurements, a set of 16-bit flash ADC VME

    modules were used to digitize signals and pass them on through the VME backplane

    to a CPU module that would reduce the data and pass them on. In the processes

    of planning the reinstallation of the NPDGamma experiment to measure background

    asymmetries, time constraints made reuse of the n3He DAQ desirable, which was at

    the time in active use and could remain at the FnPB for the NPDGamma background

    measurements once n3He was removed. The n3He system satisfied the requirements

    of NPDGamma and provided far less deadtime due to readout. Therefore, 2016 back-

    ground measurements were acquired by using 24-bit ∆Σ ADCs with an effective bit

    depth comparable to the ADCs used for hydrogen measurements.

    In both cases, two electrically-isolated yet synchronized DAQs were in operation

    at the same time. One DAQ is known as “clean”, containing only signals from the

    gamma detectors which are insensitive to SR influence down to far below the target

    10−8 statistical uncertainty. The other DAQ is considered “dirty”, being used to

    directly measure current delivered to the SR coil, as well as other signals to be syn-

    chronized with gamma measurements such as beam monitors. The clean and dirty

    DAQs are only allowed to be connected by optically-isolated digital signals, which

    allows synchronization and data export without the risk of contaminating the clean

    DAQ with SR-related signals.

    The SNS facility provides what are known as T0 signals, synchronized to the

    extraction of proton bunches into the spallation target. The T0 signals are therefore

    synchronized to the arrival of chopped neutrons at the experiment. A set of electronics

    35

  • gate the T0 signals in order to define a “run”, which are each about 7 min long. Many

    such runs are carried out back-to-back for long periods of time. The T0 signals are

    precisely delayed to coincide with the arrival of neutrons, and logically duplicated

    to the clean and dirty DAQs to trigger acquisition of 16 ms of data, leaving 2/3 ms

    before the next trigger for transferring data and avoiding corruption issues. The runs

    are stored in files referenced by unique consecutive numbers starting at zero. The

    same T0 signals are duplicated at the dirty DAQ and used to trigger the driving of

    the SR coil.

    3.8.1 Hydrogen Target DAQ

    At the heart of the DAQ system used for hydrogen measurements is a set of 16-bit

    flash ADCs by Alphi Technology Corp housed in VME crates, one for the clean DAQ

    and one for the dirty. The clean and dirty ADC modules are fed external 50 kHz

    and 62.5 kHz clocks, respectively. Since the dirty DAQ samples the SR current which

    is primarily 30 kHz, it is clocked slightly faster to avoid aliasing. For the case of

    gamma signals, which are already limited to 5 kHz, oversampling at 5× the Nyquist

    frequency pushes roughly 80% of the digitization noise beyond the desired band. A

    clock generator module for each DAQ takes T0 triggers and produces 800 and 1000

    clock edges for the clean and dirty DAQs, respectively. Once eight such sample sets

    are acquired, the ADC modules place an IRQ (interrupt request) on the VME bus to

    be read out.

    An Acromag CPU module for each VME crate listens for the ADC IRQ and

    initiates a transfer for all channels over the backplane. This process takes enough

    time that an entire 9th T0 is suppressed as a “readout pulse”. The CPUs on the

    VME crates bin the samples into 40 time bins per T0 (except for the SR signal which

    is left unbinned). The reduced data for both DAQs is then pushed over optical links

    36

  • to a workstation for monitoring and to be organized into files for transfer to long-term

    storage.

    3.8.2 n3He Delta-Sigma DAQ

    The company D-TACQ was commissioned to provide custom firmware for their

    DAQ appliances used for the n3He experiment. These appliances carry 24-bit ∆Σ

    ADCs and were used for NPDGamma background target measurements in the first

    half of 2016. The custom firmware allows the appliances to be used in a similar way

    to the DAQ for hydrogen measurements, while providing the same data structure

    without a missing readout pulse. To achieve this, the digital filtering required for

    ∆Σ operation was reset on each T0. The appliances generate their own internal

    programmable clocks distributed to all channels. The clean and dirty DAQs for 2016

    measurements were again electrically isolated.

    3.9 lH2 Cryostat

    For the np → dγ reaction, a liquid molecular parahydrogen target is optimal

    because such a target does not depolarize cold neutrons below 14.5 meV. The pri-

    mary target for the NPDGamma experiment is a 16 L volume of liquid hydrogen with

    a maximum parahydrogen concentration of 99.985%. The target is contained and

    cooled by a cryostat made of 6061 Al alloy to ∼ 15.5 K. To reach and maintain a

    near-thermal equilibrium para concentration, the hydrogen is filled and continuously

    circulated through a catalyst to facilitate ortho to para downconversion. The hydro-

    gen pressure vessel is surrounded by 6LiF-loaded polymer-matrix neutron absorber,

    thermal shielding, and a vacuum chamber. Surrounding the vacuum chamber is the

    ∼ 3π CsI gamma detector array, as shown in Figure 3.12.

    The hydrogen pressure vessel is a cylindrical volume with elliptically domed up-

    37

  • AD

    CB

    Figure 3.12: Detector with lH2 cryostat/target and SR coil model cutaway- Shown here is a cutaway representation of the geometry created for the GEANT4MC simulation in the hydrogen-target configuration. The beam enters from right toleft, first through the RF SR coil (A). It then enters the main vacuum chamber (B)of the lH2 cryostat through a double window with an intermediate vacuum volume,where the cylindrical extrusion of the vacuum chamber is centered in the CsI array(C). The incident and scattered neutron beam is controlled by LiF absorber sheetsarranged to limit gamma-emitting capture to the hydrogen contained in the pressurevessel (D) and small amounts of aluminum in the flight path.

    and down-stream walls, and has a thickness at the center of the upstream wall of

    0.063”, which the beam incident on the hydrogen must pass through. The vessel

    has a length of approximately 30 cm, and an ID of 10.5”. The thickness of the

    cylindrical side wall and downstream wall is 0.12” and 0.13”, respectively. The vessel

    has ports at the top and bottom of the downstream wall for vent operation and

    circulation through the ortho/para converter (OPC), as shown in Figure 3.13. A

    99.999% Al, high-conductivity thermal bus connects the pressure vessel to a feedback-

    controlled temperature. The lines connected to the vent and OPC are stainless steel

    with relatively low thermal conductivity to allow temperature gradients.

    In order to maintain a high fractional parahydrogen concentration, liquid hydrogen

    38

  • Tliq.

    A

    C

    E

    B

    D

    TOPC

    T evap.

    Figure 3.13: Target vessel and ortho/para conversion - A simplified schematicof the target is shown on the left, and engineering drawings of the same componentsare shown on the right. In the schematic, each feedback-controlled fixed temperatureis labeled with a dot. Volume A is the hydrogen target which is connected to fill/ventline B. Hydrogen vaporizes above the target volume and is reliquified in liquifactionchamber C, which is thermally bound to A by the five-nines Al thermal bus D. Liquidhydrogen flows down into OPC chamber E, and then back into A.

    above the vent port is slowly vaporized and subsequently liquified over the OPC,

    enforcing continuous circulation. The OPC is made of Fe(OH)3 powder and exhibits

    high magnetic field gradients for the hydrogen molecules that flow through it, enabling

    a spin flip of one of the ortho molecule nuclei. The estimated time for recirculation

    of the target is about one day (Santra et al. 2010). A schematic of the components

    involved in the circulation can be seen in Figure 3.13.

    A number of studies of transmission through the target were performed throughout

    the experiment using M1 and M4, which yield information about the O/P concen-

    tration within the target. Firstly, exponential convergence of O/P concentration can

    be seen in the transmission through the target within the first few days after the

    vessel is filled from room temperature gas (as shown in FIG. 4 of Grammer et al.

    (2015)). Secondly, flow rates were changed at certain points of operation by altering

    temperatures indicated in Figure 3.13 and observing power delivered to heaters used

    39

  • to control the temperature. In principle, differing flow rates will allow the target to

    reach different equilibrium levels of parahydrogen due to the opposition of occasional

    up conversion in the vent tube and in the walls. However, no statistically signifi-

    cant changes were detected, allowing only upper limits to be established. The O/P

    concentration is believed to be very close to thermal equilibrium.

    40

  • Chapter 4

    ASYMMETRY MEASUREMENTS

    Measurements made with the NPDGamma apparatus for the purpose of determining

    the asymmetry Aγ are described in this chapter along with general measurement

    strategies, and features in the data pertinent to the data reduction (Chapter 6) and

    selection (Chapter 7). To start, a general strategy for determining the asymmetry is

    given.

    4.1 General Strategy

    The apparatus with the hydrogen target (described in section 3.9) installed pro-

    duces a signal in the CsI detectors due to deposition from gammas of the 2.2 MeV

    deuteron binding energy. Additional background is also deposited from radiative cap-

    ture on, and radioactive decay of, apparatus materials in the beam path and in the

    path of scattered neutron flux. The prompt background from radiative capture is

    nearly perfectly correlated in time with the signal, and reasonably correlated in the

    space of gamma detector number. Furthermore, because of the current-mode oper-

    ation of the gamma detector array, the gamma energy cannot be used to select the

    np → dγ reaction. The fractional prompt background power deposition into the de-

    tectors, for all significant sources, must be determined along with their asymmetries.

    The fractional contribution is determined by Monte-Carlo (MC) and radioactivity

    of the materials following a long beam-on-target history. A number of background

    target configurations are defined in order to determine or limit their contribution to

    the asymmetry seen by the gamma detectors.

    For each target configuration, raw asymmetries for each detector are determined

    41

  • (for derivation, see section 6.2). These asymmetries have contributions from multiple

    materials in general. Each contribution is the product of the physics asymmetry for

    that material multiplied by polarization and geometric factors. Geometric factors

    include the finite structure of the beam, effective solid angle of the detector, spatial

    distribution of the material in question, an other effects. The geometric factors and

    depolarization in the target are determined via MC simulation. The incident po-

    larization is measured periodically. The validity of this approach will be tested by

    applying this method to the measurement of an asymmetry for chlorine, which is very

    large and easily measured due to significant enhancement.

    In order to estimate the statistical variances in the determination of raw asym-

    metries, the variances are determined internally by effectively performing many mea-

    surements and examining the distribution. For calculation of Aγ, raw asymmetries

    for all detectors and target configurations are combined with measured and calculated

    polarization and geometric factors in a linear set of equations, in an application of

    maximum likelihood with all physics asymmetries as parameters. Statistical variances

    are propagated from raw asymmetries to the uncertainty in Aγ.

    4.2 Overview of Hydrogen Measurements

    Production measurements of the asymmetry Aγ with the hydrogen target in place

    began at the end of November 2012 with a proton accelerator beam power of 850 kW,

    and ended at the end of March 2014 with a beam power that ramped up to over 1 MW.

    In that time, approximately 200 days of accumulated usable beam time were recorded

    in the data stream. 50,620 runs were collected, of which 47,159 were chosen for the

    analysis in this thesis. The runs are organized into groups of data that are referred to

    in this thesis as batches. The hydrogen target remained operational throughout this

    time, except when the vessel was vented and subsequently refilled over the significant

    42

  • t (ms)20− 0 20 40 60 80 100 120 140 160 180

    ampl

    itude

    (arb

    )

    5−10

    4−10

    3−10

    2−10

    A

    B

    C

    D

    E

    Figure 4.1: Gamma detector response - Plot of the average digitized waveformfor a typical detector during a special 1 Hz operation of the accelerator. Feature A isa flash of photons from the accelerator preceding the arrival of neutrons. Feature Bis known as the leading chopper edge, and C is a pair of Al Bragg edges. Feature Dis phosphorescence from the CsI following the trailing cho