UNIVERSIDAD NACIONAL AUTÓNOMA DE MÉXICO POSGRADO EN CIENCIAS (FÍSICA) INSTITUTO DE FÍSICA SEARCH FOR DARK MATTER, REACTOR AND SUPERNOVA NEUTRINOS USING NOBLE LIQUID AND GERMANIUM DETECTORS TESIS QUE PARA OPTAR POR EL GRADO DE DOCTOR EN CIENCIAS PRESENTA: MARIO ANDRÉS ALPÍZAR VENEGAS TUTOR PRINCIPAL DR. ERIC VÁZQUEZ JÁUREGUI INSTITUTO DE FÍSICA, UNAM COMITÉ TUTOR DR. EDUARDO PEINADO RODRÍGUEZ INSTITUTO DE FÍSICA, UNAM DRA. MYRIAM MONDRAGÓN CEBALLOS INSTITUTO DE FÍSICA, UNAM CIUDAD DE MÉXICO, ABRIL 2025 UNAM – Dirección General de Bibliotecas Tesis Digitales Restricciones de uso DERECHOS RESERVADOS © PROHIBIDA SU REPRODUCCIÓN TOTAL O PARCIAL Todo el material contenido en esta tesis esta protegido por la Ley Federal del Derecho de Autor (LFDA) de los Estados Unidos Mexicanos (México). El uso de imágenes, fragmentos de videos, y demás material que sea objeto de protección de los derechos de autor, será exclusivamente para fines educativos e informativos y deberá citar la fuente donde la obtuvo mencionando el autor o autores. Cualquier uso distinto como el lucro, reproducción, edición o modificación, será perseguido y sancionado por el respectivo titular de los Derechos de Autor. PROTESTA UNIVERSITARIA DE INTEGRIDAD Y HONESTIDAD ACADÉMICA PROFESIONAL (Graduación con trabajo escrito) De conformidad con lo dispuesto en los art́ıculos 87, fracción V del Estatuto General, 68, primer párrafo, del Reglamento General de Estudios Universitarios y 26, fracción I, y 35 del Reglamento General de Exámenes, me comprometo en todo tiempo a honrar a la Institución y cumplir con los principios establecidos en el Código de Ética de la Universidad Nacional Autónoma de México, especialmente con los de integridad y honestidad académica. De acuerdo con lo anterior, manifiesto que el trabajo escrito titulado: SEARCH FOR DARK MATTER, REACTOR AND SUPERNOVA NEUTRINOS USING NOBLE LIQUID AND GERMANIUM DETECTORS que presenté para obtener el grado de Doctorado es original, de mi autoŕıa y lo realicé con el rigor metodológico exigido por mi programa de posgrado, citando las fuentes de ideas, textos, gráficos u otro tipo de obras empleadas para su desarrollo. En consecuencia, acepto que la falta de cumplimento de las disposiciones reglamentarias y normativas de la Universidad, en particular las ya referidas en el Código de Ética, llevará a la nulidad de los actos de carácter académico administrativo del proceso de graduación. Atentamente Mario Andrés Alṕızar Venegas 520462236 (Nombre, firma y Número de cuenta de la persona alumna) 11 V¡O¡I'/UKn.>.D H.rqoro¡ ... l A'I!~U: ''', COORDINACiÓN GENERAL DE ESTUDIOS DE POSGRADO CARTA AVAL PARA DAR INICIO A LOS TRÁMITES DE GRADUACiÓN Universidad Nacional Autónoma de México Secretaria General Coordinación General de Estudios de Posgrado Dr. Alberto Güijosa Hidalgo Programa de Posgrado en Ciencias Físicas Presente Quien suscribe, Dr. Eric Vázquez Jáuregui , tutor(a) principal de Mario Andrés Alpízar Venegas , con número de cuenta 520462236 , integrante del alumnado de Doctorado en Ciencias (Física) de ese programa, manifiesto bajo protesta de decir verdad que conozco el trabajo escrito de graduación elaborado por dicha persona, cuyo título es: SEARCH FOR DARK MATTER, REACTOR AND SUPERNOVA NEUTRINOS USING NOBLE L/QUID AND GERMAN/UM DETECTORS , así como el reporte que contiene el resultado emitido por la herramienta tecnológica de identificación de coincidencias y similitudes con la que se analizó ese trabajo, para la prevención de faltas de integridad académica. De esta manera, con fundamento en lo previsto por los artículos 96, fracción 111 del Estatuto General de la UNAM; 21 , primero y segundo párrafos, 32, 33 Y 34 del Reglamento General de Exámenes y; 22, 49, primer párrafo y 52, fracción II del Reglamento General de Estudios de Posgrado, AVALO que el trabajo de graduación presentado se envíe al jurado para su revisión y emisión de votos , por considerar que cumple con las exigencias de rigurosidad académica previstas en la legislación universitaria. Protesto lo necesario, Ciudad Universitaria, Cd. Mx., a 11 de c8,aT .'. Dr. Eric Vá ; q~ez Jáuregui abril de 202 5 Tutor(a) principal Resumen Esta tesis se centra en dos temas principales: la búsqueda de materia oscura usando el de- tector DEAP-3600 y el estudio de la dispersión elástica coherente neutrino-núcleo (CEνNS), incluyendo análisis con datos experimentales de CONUS+ y simulaciones de señales de su- pernovas. Primero, se presentan las evidencias principales que respaldan la existencia de la materia oscura, describiendo las part́ıculas débilmente interactuantes (WIMPs), los axiones y los neutrinos estériles como candidatos destacados, aśı como los métodos de detección. Seguidamente, se describe en detalle el detector DEAP-3600. Se desarrolló un modelo exhaustivo del fondo de neutrones radiogénicos usando el marco estad́ıstico de razón de verosimilitud perfilada, construyendo y optimizando múltiples funciones de densidad de prob- abilidad. Además, se exploró un enfoque basado en aprendizaje automático para identificar retrocesos nucleares múltiples en interacciones de neutrones, un desaf́ıo para el cual no existe una solución actualmente. Un clasificador tipo árbol de decisión potenciado mostró resulta- dos prometedores para mitigar este fondo en futuros análisis. También se mejoró el modelo de fondo gamma en la búsqueda de axiones solares al incluir emisiones gamma por captura de neutrones en molibdeno y cobalto, incrementando significativamente el valor-p para la hipótesis de sólo ruido de fondo, de 0.5±0.3% a 25±1%. Las búsquedas de WIMPs y axiones solares continúan activamente en la colaboración DEAP. Los resultados del experimento CONUS+ fueron empleados para para restringir parámetros del Modelo Estándar (SM) y teoŕıas más allá del Modelo Estándar (BSM) mediante análisis de optimización con χ2. Este trabajo produjo el primer ĺımite del ángulo de mezcla débil obtenido de interacciones CEνNS con antineutrinos electrónicos de reactor provenientes de una medición con significancia mayor a 3σ: sin2 θW = 0.268 ± 0.047. Adicionalmente, se impuso un ĺımite superior al 90% de confianza para el momento magnético del neutrino: µν ≤ 5.6 × 10−10µB. También se impusieron restricciones sobre interacciones no estándar (NSI), con contornos en el plano ϵuVee –ϵ dV ee comparables a los de COHERENT. Estos resul- tados confirman el potencial de CEνNS para estudiar parámetros electrodébiles como θW y µν , y explorar escenarios BSM como las NSI. De manera complementaria, se simularon señales de CEνNS producidas por el colapso gravitacional de una supernova (SN) usando el software de código abierto EstrellaNueva. Las capacidades de detección de DEAP-3600 y LUX-ZEPLIN (LZ) fueron evaluadas para una SN de 15 M⊙, a distancias de 0.196, 10.0 y 20.0 kpc. DEAP se limita a eventos cercanos, mientras que LZ puede detectar señales en todo el disco galáctico. Por consiguiente, se con- struyeron cotas superiores e inferiores en los conteos de eventos para optimización mediante χ2, obteniendo ĺımites para θW y µν , y analizando el efecto de la distancia. Además, se pro- puso un método para aproximar la composición en sabores del haz de neutrinos, permitiendo estudiar las NSI considerando tanto la jerarqúıa normal (NH) como la invertida (IH). Estos resultados destacan el potencial de detectores de ĺıquidos nobles para explorar propiedades fundamentales del neutrino, estableciendo una base sólida para futuras búsquedas con señales de CEνNS provenientes de supernovas. iv Abstract This thesis focuses on two main topics: the search for dark matter using the DEAP- 3600 detector, and the study of coherent elastic neutrino-nucleus scattering (CEνNS). The latter includes analyses based on experimental data from the CONUS+ collaboration and simulations of supernova neutrino signals. To begin with, the main lines of evidence supporting the existence of dark matter are presented. Weakly interacting massive particles (WIMPs), axions, and sterile neutrinos are described as some of the most prominent candidates. The primary detection methods are also outlined in this work. Subsequently, the DEAP-3600 dark matter detector is described in detail. A comprehen- sive background model for radiogenic neutrons was developed within the Profile Likelihood Ratio statistical framework. Numerous probability density functions were constructed and optimized to characterize this background. Additionally, a machine learning approach was explored to identify multiple nuclear recoils from neutron interactions—a challenge for which no current solution exists. A Boosted Decision Tree classifier achieved promising perfor- mance, demonstrating the potential of machine learning techniques for neutron background suppression in future analyses. The gamma background model for the solar axion search was also improved by incorporating neutron capture gamma emissions from molybdenum and cobalt isotopes, significantly increasing the background-only p-value from 0.5±0.3% to 25±1%. Both the WIMP and solar axion searches are ongoing efforts within the DEAP collaboration. Results from the CONUS+ experiment were used to constrain parameters from the Stan- dard Model (SM) and beyond (BSM) using a χ2-based optimization analysis. This work pro- duced the first CEνNS-derived limit on the weak mixing angle for reactor electron antineutri- nos, from a measurement with a significance greater than 3σ: sin2 θW = 0.268± 0.047, along with a 90% confidence upper limit on the neutrino magnetic moment: µν ≤ 5.6× 10−10µB. Limits on Non-Standard Interactions (NSI) were also obtained, with contours in the ϵuVee –ϵ dV ee plane comparable to those reported by COHERENT. These results show that CEνNS is a powerful tool for probing electroweak parameters such as θW and µν , and for exploring BSM scenarios such as NSI. As part of the CEνNS investigation, signals from core-collapse supernovae (SNe) were sim- ulated using the open-source software EstrellaNueva. The detection capabilities of DEAP- 3600 and LUX-ZEPLIN (LZ) were evaluated for a SN originating from a 15 M⊙ progenitor star at distances of 0.196, 10.0, and 20.0 kpc. While DEAP is limited to nearby events, LZ could detect SN signals across the entire galactic disk. Upper and lower bounds on the expected event counts were established to enable χ2 optimization. Limits on θW and µν were derived, and the effect of SN distance on these constraints was analyzed. Additionally, a method was proposed to approximate the flavor composition of the neutrino burst, enabling the study of NSI for all neutrino flavors considering both the normal (NH) and inverted (IH) hierarchies. The results from this investigation demonstrate the potential of noble liquid detectors to explore fundamental neutrino properties and lay the groundwork for future SNe CEνNS-based searches. v Agradecimientos Agradezco a mis papás, Mario y Eney, por su apoyo incondicional durante todos estos años. Gracias por motivarme a perseguir mis sueños y por haberme criado para ser una persona sincera y trabajadora, como ustedes. A mis hermanos, Meli y Jose, gracias por visitarme y mantenerme cerca a través de sus mensajes; sent́ı su cariño a la distancia. A mi abuelito Mario, gracias por habernos dejado un legado de generosidad y cariño. A Tita, gracias por ser mi segunda mamá; este logro también fue posible gracias a tus enseñanzas de vida. Gracias, profe Eric, por tu esfuerzo y dedicación durante estos seis años de mentoŕıa en la maestŕıa y el doctorado. Tu trabajo siempre fue hecho a conciencia, y priorizando el lado humano de las personas. Gracias, profes Eduardo y Myriam. Valoro mucho las clases que tuve con ambos y les agradezco también por completar mi comité tutor all-star. Extiendo mi agradecimiento a otros profesores de la UNAM que me dieron excelentes cursos y que fueron fundamentales en mi doctorado, especialmente a los profes Alexis Aguilar y Alberto Mart́ın. También agradezco a los profes de la UCR, en especial a Blai Garolera y Alejandro Jenkins: este largo proceso comenzó junto a ustedes. A Alberto Ángeles, gracias por la terapia brindada con tanta dedicación; fue indispensable para sacar esto adelante. A Alice: gracias por tu compañ́ıa en este arduo camino. Tu apoyo, paciencia y amor han sido una gran inspiración durante todo este tiempo. A mis hermanos no de sangre, Alfredo y Marco, gracias por creer en mı́ durante todos estos años. Gracias por las visitas, y las charlas a distancia; sin ustedes esto no habŕıa sido posible. A PalcoVIP: Mauricio, Favián y Pedro, gracias; su compañ́ıa ha estado muy presente a través de nuestras charlas de fútbol y demás asuntos dignos de ser considerados en nuestro foro de discusión. A mi hermana catalana, Nat, gracias por la motivación, los consejos tan atinados, y los ánimos en los momentos cuesta arriba del doctorado. Tu determinación me sirvió de ejemplo para seguir adelante. A Kattia, gracias por estar presente tanto en los tiempos de la UCR como de la UNAM; tu amistad, que se ha mantenido durante los años y la distancia, ha significado mucho para mı́. Hago una mención especial a Rigo, mi gato, por su valioso aporte en demostraciones y tareas de f́ısica. A mis demás amigos, de f́ısica, música, artes marciales, roomies y otras facetas de la vida: gracias. Su amistad fue un aliento a lo largo de este proceso. Este trabajo de tesis fue realizado gracias al Programa UNAM-PAPIIT IN105923. vi Contents 1 Introduction 1 2 Evidence for Dark Matter 3 2.1 Galaxy Velocities in Clusters and Galaxy Rotation Curves . . . . . . . . . . 3 2.2 Anisotropy of the Cosmic Microwave Background Temperature . . . . . . . . 6 2.3 Gravitational Lensing in Galaxy Clusters . . . . . . . . . . . . . . . . . . . . 10 3 Dark Matter Candidates and Methods of Detection 13 3.1 Weakly Interacting Massive Particles . . . . . . . . . . . . . . . . . . . . . . 14 3.2 Axions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15 3.3 Sterile Neutrinos . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16 3.4 Methods of Detection . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16 3.4.1 Production . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17 3.4.2 Direct Detection . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18 3.4.3 Indirect Detection . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19 4 The DEAP-3600 Detector 21 4.1 Description . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21 4.2 Scintillation from Nuclear Recoils in Liquid Argon . . . . . . . . . . . . . . . 23 4.3 Pulse Shape Discrimination . . . . . . . . . . . . . . . . . . . . . . . . . . . 23 4.4 Background Sources and Region of Interest . . . . . . . . . . . . . . . . . . . 25 4.4.1 Electrons and Gamma Rays . . . . . . . . . . . . . . . . . . . . . . . 26 4.4.2 Alpha Particles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27 4.4.3 Neutrons . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27 4.4.4 Region of Interest and Background Estimates . . . . . . . . . . . . . 28 5 Neutron Background in DEAP-3600 31 5.1 The Profile Likelihood Ratio Method . . . . . . . . . . . . . . . . . . . . . . 31 5.2 Radiogenic Neutron Simulations . . . . . . . . . . . . . . . . . . . . . . . . . 33 5.3 PDFs for the Radiogenic Neutrons . . . . . . . . . . . . . . . . . . . . . . . 35 5.3.1 NSCBayes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38 5.3.2 Fprompt . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38 5.3.3 MblikelihoodR . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42 5.4 Neutron Discrimination Using Machine Learning . . . . . . . . . . . . . . . . 47 vii viii CONTENTS 5.5 Formulation as a Machine Learning Problem . . . . . . . . . . . . . . . . . . 48 5.6 Data Collecting and Processing . . . . . . . . . . . . . . . . . . . . . . . . . 49 5.7 Model Training and Evaluation . . . . . . . . . . . . . . . . . . . . . . . . . 50 5.7.1 Training . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50 5.7.2 Evaluation Metrics . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50 5.7.3 Validation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52 5.7.4 Additional Tests . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55 5.8 Findings and Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58 6 Gamma Background for the Search of 5.5 MeV Solar Axions 61 6.1 Production Mechanism . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61 6.2 Interactions with the Detector . . . . . . . . . . . . . . . . . . . . . . . . . . 62 6.2.1 Compton Conversion . . . . . . . . . . . . . . . . . . . . . . . . . . . 62 6.2.2 Decay to 2γ . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62 6.2.3 Inverse Primakoff Effect . . . . . . . . . . . . . . . . . . . . . . . . . 62 6.2.4 Axioelectric Effect . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62 6.3 Signal and its dependence on axion mass . . . . . . . . . . . . . . . . . . . . 63 6.4 Region of Interest for Axion Search . . . . . . . . . . . . . . . . . . . . . . . 63 6.5 Background Sources . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64 6.5.1 γ-rays from Neutron Capture . . . . . . . . . . . . . . . . . . . . . . 65 6.5.2 γ-rays from 208Tl Decay . . . . . . . . . . . . . . . . . . . . . . . . . 65 6.6 Simulations for the Neutron Capture γ-ray Background . . . . . . . . . . . . 66 7 Exploration of Neutrinos in the Standard Model and Beyond 73 7.1 General Aspects about Neutrinos . . . . . . . . . . . . . . . . . . . . . . . . 73 7.2 Coherent Elastic Neutrino-Nucleus Scattering . . . . . . . . . . . . . . . . . 78 7.3 CEνNS Measurement by the CONUS+ Experiment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79 7.3.1 Description of the CONUS+ Detector . . . . . . . . . . . . . . . . . . 79 7.3.2 Results of the CEνNS Measurement by CONUS+ . . . . . . . . . . . 82 7.4 Study of Scenarios in the SM and Beyond . . . . . . . . . . . . . . . . . . . 84 7.4.1 Calculation of the Number of events . . . . . . . . . . . . . . . . . . 85 7.4.2 Weak Mixing Angle . . . . . . . . . . . . . . . . . . . . . . . . . . . . 89 7.4.3 Magnetic Moment of the Neutrino . . . . . . . . . . . . . . . . . . . . 90 7.4.4 Non-Standard Interactions . . . . . . . . . . . . . . . . . . . . . . . . 95 7.5 Neutrino Production in Core-Collapse Supernovae . . . . . . . . . . . . . . . 99 7.6 CEνNS Simulations for DEAP-3600 and LUX-ZEPLIN with EstrellaNueva . 103 7.7 SN Neutrino Parameter Analysis: SM and Beyond . . . . . . . . . . . . . . . 109 7.7.1 Results for the Weak Mixing Angle . . . . . . . . . . . . . . . . . . . 111 7.7.2 Findings for the Neutrino Magnetic Moment . . . . . . . . . . . . . . 112 7.7.3 Implications for Non-Standard Interactions . . . . . . . . . . . . . . . 118 7.8 Conclusions from the Parameter Analysis with SN Neutrinos . . . . . . . . . 120 8 Conclusions 123 CONTENTS ix Appendix A Acronyms and Abbreviations Used 127 Appendix B Supplementary Plots for the CEνNS Studies 131 Bibliography 137 x CONTENTS List of Figures 2.1 Rotation curves of 21 galaxies studied by Vera Rubin. . . . . . . . . . . . . . 4 2.2 Rotation curve with the contributions of its main components. . . . . . . . . 5 2.3 Dipole anisotropy of the CMB. . . . . . . . . . . . . . . . . . . . . . . . . . 7 2.4 Multipole anisotropy of the CMB temperature. . . . . . . . . . . . . . . . . . 7 2.5 CMB power spectrum with a fit for the ΛCDM model. . . . . . . . . . . . . 8 2.6 Relation of the peaks in the CMB power spectrum to the cosmological pa- rameters. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9 2.7 Diagram of light deflection by a massive object. . . . . . . . . . . . . . . . . 10 2.8 Gravitational lensing in Galaxy Cluster CL0024+17. . . . . . . . . . . . . . 11 2.9 Galaxy Cluster 1E 0657–56, known as the “Bullet Cluster”. . . . . . . . . . . 11 2.10 Galaxy clusters as evidence for dark matter: MACS J0025, Abell 520, Abell 2744, and ZwCl 1358+62 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12 3.1 Mass range for various dark matter candidates. . . . . . . . . . . . . . . . . 13 3.2 Comparison of WIMP and dark matter relic densities as a function of WIMP mass. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14 3.3 Axion mass range with constraints. . . . . . . . . . . . . . . . . . . . . . . . 15 3.4 Methods of detection for dark matter. . . . . . . . . . . . . . . . . . . . . . . 17 3.5 LZ spin independet WIMP-nucleon upper limit. . . . . . . . . . . . . . . . . 19 3.6 SM products from the annihilation of WIMPs. . . . . . . . . . . . . . . . . . 20 4.1 Outer structure of the DEAP-3600 detector . . . . . . . . . . . . . . . . . . 22 4.2 Acrylic vessel and inner structure. . . . . . . . . . . . . . . . . . . . . . . . . 22 4.3 Nuclear recoil. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23 4.4 Scintillation signal comparison of ER and NR. . . . . . . . . . . . . . . . . . 25 4.5 NR and ER bands in Fprompt and PE. . . . . . . . . . . . . . . . . . . . . . . 26 4.6 WIMP search region of interest . . . . . . . . . . . . . . . . . . . . . . . . . 29 5.1 Comparison of statistical methods for limit setting. . . . . . . . . . . . . . . 34 5.2 PLR Region of Interest. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34 5.3 Representation of the radiogenic neutron simulations in two stages. . . . . . 35 5.4 Flowchart of the datasets in the neutron simulations. . . . . . . . . . . . . . 37 5.5 Normalized PIFGAR histogram. . . . . . . . . . . . . . . . . . . . . . . . . . 37 5.6 NSC fit for the PIF2 tpbUP dataset. . . . . . . . . . . . . . . . . . . . . . . 39 5.7 RPR fit for the PIF2 rfrUP distribution, in the 90-150 PE bin. . . . . . . . . 41 xi xii LIST OF FIGURES 5.8 RPR N and µ parameter fits for the nominal PIF3 and nominal PIF2 datasets. 43 5.9 RPR λ and σ parameter fits for the rfrUP PIF2 and rfrUP PIF3 datasets. . 44 5.10 MBR fit for the PIF2 tpbUP distribution, in the 120-150 PE bin. . . . . . . 45 5.11 MBR N and θ parameter fits for the rfrUP PIF2 and tpbDOWN PIF3 datasets. 46 5.12 Confusion matrix. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51 5.13 Definition of an ROC curve. . . . . . . . . . . . . . . . . . . . . . . . . . . . 53 5.14 Individual vs boosted decision tree. . . . . . . . . . . . . . . . . . . . . . . . 53 5.15 Confusion matrix for the boosted decision tree in the validation phase. . . . 54 5.16 ROC curve for the BDT in the validation phase. . . . . . . . . . . . . . . . . 55 5.17 Confusion matrix for the boosted decision tree in the testing phase. . . . . . 56 5.18 ROC curve for the BDT in the testing phase. . . . . . . . . . . . . . . . . . 57 5.19 Predicted multiple recoils for the AmBe dataset. . . . . . . . . . . . . . . . . 59 6.1 Feynman Diagrams of the Four Axion-Detector Interactions . . . . . . . . . 63 6.2 Signal of the four axion interactions . . . . . . . . . . . . . . . . . . . . . . . 64 6.3 Total energy deposited in the detector for Compton Conversion and 2γ Decay for different axion masses . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64 6.4 Mechanism of γ-ray production by neutron capture . . . . . . . . . . . . . . 65 6.5 Spectrum generated for γ-rays produced in the decay of 208Tl . . . . . . . . . 66 6.6 Comparison of background models before and after gamma rays from neutron captures in 59Co and Mo isotopes were added. . . . . . . . . . . . . . . . . . 67 6.7 Simulated gamma rays from the steel casing and neck of the detector . . . . 68 6.8 Energy deposited in liquid argon by gammas emitted due to neutron capture in 59Co and 95Mo. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 68 6.9 Acceptance for radial cut at 750 mm for simulated neutrons from the acrylic container. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69 6.10 Preliminary background model for the axion search after blind analysis . . . 70 6.11 Preliminary exclusion curves for gAe . . . . . . . . . . . . . . . . . . . . . . . 70 7.1 Neutrino sources. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74 7.2 Neutrino mass hierarchies. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76 7.3 Feynman diagram for neutrinoless double beta decay. . . . . . . . . . . . . . 77 7.4 Feynman diagram for CEνNS. . . . . . . . . . . . . . . . . . . . . . . . . . . 78 7.5 Representation of CEνNS and enhanced cross-section. . . . . . . . . . . . . . 79 7.6 Use of CEνNS events for probing the neutrino magnetic moment and non- standard interactions parameters. . . . . . . . . . . . . . . . . . . . . . . . . 80 7.7 Setup of the CONUS+ experiment. . . . . . . . . . . . . . . . . . . . . . . . 81 7.8 Production of reactor antineutrinos. . . . . . . . . . . . . . . . . . . . . . . . 81 7.9 HPGe diodes in CONUS+. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 82 7.10 CONUS+ CEνNS signal prediction and data. . . . . . . . . . . . . . . . . . 83 7.11 Trigger efficiency of the CONUS+ diodes. . . . . . . . . . . . . . . . . . . . 86 7.12 Germanium total CEνNS cross sections. . . . . . . . . . . . . . . . . . . . . 87 7.13 Event rates for CONUS+. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 88 7.14 Weak mixing angle fit for CONUS+. . . . . . . . . . . . . . . . . . . . . . . 91 7.15 Neutrino Magnetic moment χ2 profile for CONUS+. . . . . . . . . . . . . . . 92 LIST OF FIGURES xiii 7.16 Neutrino magnetic moment 90% CL. upper limit for CONUS+. . . . . . . . 94 7.17 NSI parameter limits for CONUS+. . . . . . . . . . . . . . . . . . . . . . . . 98 7.18 Diagram of the Supernova stages. . . . . . . . . . . . . . . . . . . . . . . . . 101 7.19 Neutrino Fluences with MSW Effect. . . . . . . . . . . . . . . . . . . . . . . 102 7.20 Diagram of the LZ detector structure. . . . . . . . . . . . . . . . . . . . . . . 104 7.21 Variable neutrino fluxes during a SN. . . . . . . . . . . . . . . . . . . . . . . 105 7.22 Xe and Ar total CEνNS cross sections. . . . . . . . . . . . . . . . . . . . . . 106 7.23 Simulated neutrino fluences by flavor and distances. . . . . . . . . . . . . . . 108 7.24 DEAP-3600 and LZ detection efficiencies. . . . . . . . . . . . . . . . . . . . . 109 7.25 Event rates for LZ and DEAP at various SN distances. . . . . . . . . . . . . 110 7.26 Results for sin2 θW as a function of SN distance. . . . . . . . . . . . . . . . . 113 7.27 SN Results for sin2 θW compared to measurements from other experiments. . 114 7.28 90% CL. upper limits for µν as a function of SN distance. . . . . . . . . . . . 116 7.29 SN results for µν compared to measurements from other experiments. . . . . 117 7.30 NSI parameters 95% CL. contour comparison for DEAP-3600, LZ and CONUS+.121 B.1 Helm form factors for germanium. . . . . . . . . . . . . . . . . . . . . . . . . 131 B.2 Event rates for individual CONUS+ detectors. . . . . . . . . . . . . . . . . . 132 B.3 Helm form factors for Xe and Ar. . . . . . . . . . . . . . . . . . . . . . . . . 133 B.4 DEAP-3600 event rates at 10.0 (left) and 20 kpc (right). . . . . . . . . . . . 133 B.5 NSI parameter 95% CL. contours for DEAP-3600 and LZ at 0.196 kpc (NH). 134 B.6 NSI parameter 95% CL. contours for DEAP-3600 and LZ at 0.196 kpc (IH). 135 B.7 NSI parameter contours, NH vs IH, for LZ and DEAP-3600 at 0.196 kpc. . . 136 xiv LIST OF FIGURES List of Tables 2.1 Current physical densities as cosmological parameters of the ΛCDM model. . 9 4.1 Background Estimates for the latest published WIMP search. . . . . . . . . 29 5.1 PMT array mass, purity and neutron yield used in the simulations. . . . . . 36 5.2 Count of multiple and single nuclear recoils from the PMT neutron dataset. 49 5.3 Metric scores for the BDT in the validation phase. . . . . . . . . . . . . . . . 55 5.4 Metric scores for the BDT in testing. . . . . . . . . . . . . . . . . . . . . . . 57 6.1 Number of events generated per molybdenum isotope. . . . . . . . . . . . . . 69 7.1 Detector data in the CONUS+ CEνNS measurement. . . . . . . . . . . . . . 84 7.2 Natural abundances of the germanium isotopes. . . . . . . . . . . . . . . . . 86 7.3 Events calculated for each detector. . . . . . . . . . . . . . . . . . . . . . . . 89 7.4 Natural abundances of the xenon isotopes. . . . . . . . . . . . . . . . . . . . 106 7.5 SN CEνNS simulated detections for LZ and DEAP-3600 at different distances. 107 7.6 SN CEνNS 1σ intervals for LZ and DEAP-3600 at different distances. . . . . 111 7.7 Results for sin2 θW for the SN study. . . . . . . . . . . . . . . . . . . . . . . 112 7.8 Neutrino magnetic moment upper limits for the SN study. . . . . . . . . . . 115 7.9 Flavor compositions of event counts for DEAP-3600 and LZ. . . . . . . . . . 119 xv xvi LIST OF TABLES 1 Introduction The field of particle physics has provided some of the most far-reaching insights in hu- manity’s ongoing pursuit to understand more about the Universe. This pursuit is currently focused on pushing beyond known territory, extending the boundaries of what is understood, while also refining the knowledge we already possess. One of the main driving forces in this pursuit is the effort to explain dark matter. The key observations leading to the conclusion that dark matter must exist include galaxy rotation curves, the anisotropy in the cosmic microwave background, and the matter distribution inferred from gravitational lensing observed in numerous galaxy clusters. Together, these unexplained phenomena point to the existence of a type of matter that must be massive, stable, and non-interacting electromagnetically. This is the focus of discussion in Chapter 2. A plethora of ideas have been proposed to account for these observations. Chapter 3 de- scribes three of the leading candidates: weakly interacting massive particles (WIMPs), QCD axions, and sterile neutrinos. The main detection methods (direct, indirect, and production) are also reviewed. One of the leading initiatives in direct detection is spearheaded by the DEAP-3600 exper- iment. Its detection principles as a liquid argon scintillation detector are described in detail in Chapter 4, along with its primary background sources, such as neutrons and gamma rays, relevant to the search for different dark matter candidates. Radiogenic neutron background is the focus of Chapter 5. The statistical methodology by which background sources are modeled has a far-reaching impact on the sensitivity of experiments seeking rare phenomena such as dark matter interactions. Starting in Section 5.1, the Profile Likelihood Ratio framework is introduced and applied to the characterization of radiogenic neutrons, as part of an ongoing initiative by the DEAP collaboration. Physics research often evolves in symbiosis with technological innovation: advances in one often fuel progress in the other. One such case is the application of machine learning, which has had a profound impact on particle physics. Section 5.4 begins the exploration of a novel classification strategy for mitigating neutron background in nuclear recoil dark matter searches. This technique exploits a currently underutilized discriminant: the difference in scattering behavior between neutrons and nuclei, versus WIMPs and nuclei. The results of this exploratory study are encouraging, and future directions are proposed that may prove valuable for WIMP searches dealing with radiogenic neutrons as a considerable background source. In addition to QCD axions, related theoretical models propose the existence of new parti- cles known as axion-like particles (ALPs), which could be produced in the Sun. DEAP also contributes to the investigation of these models through an ongoing solar axion search. A 1 2 CHAPTER 1. INTRODUCTION contribution to the background model for this effort is described in Chapter 6. Neutrinos offer a unique window into new physics: their non-zero masses—revealed through flavor oscillations—fall outside the explanatory scope of the Standard Model. Moreover, their interactions provide a valuable means of probing and refining our understanding of elec- troweak processes. Chapter 7 explores this potential by analyzing coherent elastic neutrino- nucleus scattering (CEνNS) using reactor neutrinos from the CONUS+ experiment. This analysis is used to probe the weak mixing angle, the magnetic moment of the neutrino, and possible new physics trough Non-Standard Interactions (NSIs). The work presented here forms the basis of the first peer-reviewed publication analyzing CONUS+ data. The fascinating events of core-collapse supernovae also offer a powerful avenue for exploring Standard Model physics and beyond, through CEνNS measurements. Beginning in Section 7.5, simulations are carried out to model signals at DEAP-3600 and LUX-ZEPLIN (LZ), currently the largest operating dark matter detector. Constraints are obtained for the weak mixing angle, the magnetic moment of the neutrino, and NSI parameters. A methodology is proposed to guide future studies aiming to expand on this analysis. This thesis represents a focused effort to contribute meaningfully to the rich tradition of particle physics−striving to understand more about the Universe and expand the boundaries of what is known. 2 Evidence for Dark Matter The evidence supporting the existence of dark matter (DM) is substantial and spans mass scales ranging from galaxies to large scale structures such as galaxy clusters and filaments. While this evidence has been accumulated for decades, it is important to clarify that all of it stems from gravitational interactions. However, this evidence arises from varied and compelling sources, making DM the leading contender to explain these observations. The following sections will review some of the most conclusive observations supporting the existence of DM, including galaxy rotation curves, temperature anisotropies in the cosmic microwave background, and gravitational lensing in galaxy clusters. 2.1 Galaxy Velocities in Clusters and Galaxy Rotation Curves The most compelling source of evidence in favor of the DM hypothesis comes from galaxy velocities in a cluster and galaxy rotation curves. The velocities of galaxies in a galaxy cluster are correlated to the amount of matter in the cluster, and can be used to infer its density. Rotation curves are constructed by measuring the rotational velocity of the stars in a galaxy as a function of their distance to the galactic center. As shall be seen, this method can provide an insight into the galactic mass and how it is spatially distributed. Fritz Zwicky conducted an early example of these studies in 1933 [1, 2]. He measured the velocities of the galaxies in the Coma cluster by using the Doppler red-shift in their spectra, and found that there was a higher dispersion than expected in these velocities. Zwicky determined this dispersion by taking into account the well known cosmological red-shift, this meant that these galaxies were moving faster than expected. He related these velocities to the cluster’s mass via the Virial theorem, and found that the density of the cluster should be approximately 400 times greater than what could be accounted for by its luminous matter, otherwise the galaxies in the cluster would disperse. Zwicky hypothesized that there must be a non luminous but massive form of matter providing the necessary density to hold the cluster together, which he referred to as dunkle materie (dark matter). The next significant hint pointing in the direction of DM comes from studying rotational curves of individual galaxies. This is also achieved by examining the Doppler shifted spectra. When looking at the galaxy edge-on, the Doppler effect will produce a more significant red- shift on the end of the galaxy that spins away from the observer, while the opposite will occur on the other end. By using Newtonian mechanics, the rotational velocity of the stars in a galaxy v(r) can be 3 4 CHAPTER 2. EVIDENCE FOR DARK MATTER determined as a function of their distance to the galactic center. Equating the centripetal force to the gravitational force and rearranging, produces the following relation: v(r) = √ GM(r) r , (2.1) where r is the distance of an arbitrary point to the galactic center, M(r) is the galaxy’s mass as a function of r and G is Newton’s gravitational constant. At the outer regions of the galaxy, the mass density due to the stars in the galaxy is approximately constant. Assuming that stars are the dominant component of the galactic mass results in this proportionality relation: v(r) ∝ r−1/2. (2.2) According to (2.2), the rotational velocities of the stars in a galaxy as a function of their distance to the galactic center should vary as r−1/2, however, this is not what was observed by the American astronomer Vera Rubin in 1980 [3] 1. She, along with her colleagues Kent Ford and Norbert Thonnard, measured the rotational curves of 21 spiral galaxies of type Sc, which have an easily identifiable galactic center, and found that v(r) is approximately constant for the outer regions of the galaxies. Figure 2.1: Rotation curves of the 21 Sc type galaxies studied by Vera Rubin [3]. In order to explain the approximately constant rotation curves seen in 2.1, the term M(r) in Equation 2.2 would have to increase linearly proportional to r. If the majority of the mass in galaxy were due to the stars in it, this would mean that the density of stars would increase as the distance from the galactic center also increases, this contradicts the observations made by Rubin. One possible explanation to this conundrum derives from extending Zwicky’s hypothesis, and proposing that DM exists not only in the scale of galaxy clusters, but also around individual galaxies. This would provide a source of mass that could explain the rotation curves while also accounting for the insufficient amount of luminous matter needed to explain 1Prior to this, Rubin and Ford reported flat rotation curves in observations of the Andromeda galaxy [4]. 2.1. GALAXY VELOCITIES IN CLUSTERS AND GALAXY ROTATION CURVES 5 this phenomenon. Following this idea and imposing the proportionality between M(r) and r, the M(r) term in Equation 2.2 takes the form: M(r) = ∫ V (r) ρ(r′)dV ′ = 4π ∫ r 0 ρ(r′)r′2dr′ ∝ r. (2.3) where spherical symmetry has been made use of, which is consistent with Rubin’s observa- tions. In order for the proportionality relation in Equation 2.3 to hold, the mass density term ρ(r′) must be proportional to r′−2: M(r) = 4π ∫ r 0 ρ(r′)r′2dr′ ∝ ∫ r 0 r′2 r′2 dr′ ∝ r. (2.4) By proposing that the majority of the mass in a galaxy is not due to its stars, but instead to DM distributed in a halo around it as outlined in Equation 2.4, the rotation curves can be explained. This massive DM halo is the main contributor to a galaxy’s rotation curve, as can be seen in Figure 2.2 for the galaxy NGC 6503. Figure 2.2: Rotation curve of the galaxy NGC 6503, with the contributions of its main components: non luminous gas, the galactic disk, and the dark matter halo. [5]. As an example, for our Milky Way galaxy it is estimated that its total mass within a radius of 20 kpc is 1.91+0.18 −0.17 ×1011 solar masses, while the DM halo within that same radius accounts for 1.37+0.18 −0.17 ×1011 of those solar masses [6]. From this, it is derived that the local DM density in the solar system is approximately ρ0 = 0.3 GeV/cm3 [5]. It is important to note that DM is not the only possible explanation for the rotation curves. An alternative approach is given by Milgrom [7], in which he proposes that instead 6 CHAPTER 2. EVIDENCE FOR DARK MATTER of adding a new form of matter, changing the behavior of gravity can also be used to explain the observations. This idea receives the name of Modified Newtonian Dynamics (MoND), and has been considered as DM’s most prominent rival hypothesis. However, recent studies have shown that DM is an overall better fit for rotation curves [8], and that MoND makes some predictions that contradict observations in wide binary star systems [9]. Currently, the hypothesis that galaxies are contained in a halo of massive matter that does not interact with light, DM, is the most favored explanation for the galaxy rotation curves. 2.2 Anisotropy of the Cosmic Microwave Background Temperature The cosmic microwave background (CMB) is a relic radiation that began propagating freely approximately 370 thousand years after the Big Bang [10]. Approximately 10−10 seconds after the Big Bang, the universe was composed of a quark-gluon plasma, as well as electrons and positrons. As the universe expanded and cooled, these quarks eventually formed protons and neutrons in a process known as baryogenesis, while the electrons and positrons annihilated each other, producing photons. This led to a period during which these photons did not propagate freely, as they scattered off the charged particles they encountered: the protons and electrons that survived annihilation. As the temperature of the universe dropped, these protons, neutrons, and electrons reached energies low enough to form helium and hydrogen atoms, which are neutral in electrical charge, making them transparent to the photons. It was then, approximately 370 thousand years after the Big Bang, during a period known as recombination, that these photons decoupled from the thermal equilibrium they were in with the electrons and positrons and began to propagate freely. When emitted, these photons were in the infrared region, but the expansion of the universe has caused them to be redshifted to wavelengths in the microwave part of the spectrum, hence the name given to this radiation [11]. The existence of the CMB was proposed in 1948 by Georgiy Gamow [12], and it was confirmed by Arno Penzias and Robert Wilson through an accidental detection in 1965 [13]. During the initial years after its observation, it was believed that the CMB temperature was isotropic, meaning that the temperature of these photons was the same in every direction. However, this was found not to be true in 1991 when the COBE satellite detected fluctuations in the temperature with respect to the mean of 2.7 K [14]. The largest of these fluctuations is due to the Doppler effect caused by the relative motion of the Earth with respect to the CMB rest frame, this is known as the dipole CMB anisotropy [15] (see Figure 2.3). When looked at with a more precise angular resolution, smaller scale fluctuations become apparent. The variation in temperature can be written as a function of the angular position as seen from Earth, θ and ϕ, in the following manner: ∆T T = ∞ ∑ l=0 l ∑ m=−l almYlm(θ, ϕ). (2.5) In Equation 2.5, the l index determines the order of the angular resolution, which increases with l according to the relation: 2.2. ANISOTROPYOF THE COSMICMICROWAVE BACKGROUND TEMPERATURE7 Figure 2.3: Dipole anisotropy of the CMB produced by the Doppler effect due to the relative motion between the Earth and the CMB rest frame. It produces a variation in temperature with respect to the mean of 2.7 K, represented in color scale [16]. ∆θ = 180◦ l . (2.6) The smaller scale fluctuations seen in Figure 2.4 are originated by the phenomenon of baryon acoustic oscillations in the proton, neutron and electron plasma [10]. The acoustic waves are a product of small scale density variations in the plasma, which can be traced back to curvature fluctuations originated in the inflationary period. In regions where the plasma density was slightly higher, the temperature was also higher, due to the effect of the gravitational compression. As a consequence of this compression, more photon radiation was emitted in those regions, which in turn caused the plasma particles to disperse and the density to decrease. When the density decreased, the particles were drawn together again by gravity and the cycle of rarefaction and compression started over, producing oscillations in the plasma. Figure 2.4: Anisotropy of the CMB temperature up to order l ≈ 2500, the dipole term has been removed for clarity [17]. The speed of propagation of these acoustic oscillations in the plasma, also known as the 8 CHAPTER 2. EVIDENCE FOR DARK MATTER speed of sound (cs), is given by [10]: c2s = δp δρ = c2 3 1 1 + 3ρB/4ργ , (2.7) where p is the pressure of the plasma, ρB and ργ are the baryon and photon density in the medium respectively, and c is the speed of light in a vacuum. When the photons propagated, at the moment of the recombination, they carried this information about the oscillation parameters in their temperature anisotropy. The temperature anisotropy can be studied by examining the power spectrum of the CMB with high angular resolution. This allows for comparison with theoretical models to evaluate how well they explain the observations. Such an approach is demonstrated by the results obtained by Planck [18] in Figure 2.5, which include an expansion up to multipole orders of l ≈ 2500. Figure 2.5: Power spectrum of the CMB temperature obtained by the Planck satellite. Multipolar expansion up to the order l ≈ 2500. In blue, the best fit curve, based in the ΛCDM cosmological model [18]. The power spectrum shows the angular scale, as seen from Earth, of the regions with fluctuations in temperature. The odd peaks correspond to regions of the plasma that were at maximal compression at the moment of recombination, while the even peaks are regions at maximal rarefactions. Measuring the diameter of compression regions in the CMB, cor- responding to the highest peak in Figure 2.5 allows us to verify that the universe is flat. This verification, in turn, makes it possible to estimate the total mass-energy content of the universe, since it must coincide with the critical density required for a flat space. The second peak from left to right, corresponds to regions of maximum rarefaction; comparing its 2.2. ANISOTROPYOF THE COSMICMICROWAVE BACKGROUND TEMPERATURE9 height to that of the first peak allows an estimation of the fraction of the total mass-energy content comprised by baryonic matter [19]. The heights of the remaining peaks give insights on when the radiation epoch of the early universe gave way to the matter dominated epoch. The matter that dominated the mass-energy density of the universe at that time could not have been baryonic matter, or it would have resulted in a higher second peak, making DM an ideal explanation for those peaks. The data points from the CMB spectra are fitted with great precision using the ΛCDM cosmological model, where Λ represents the cosmological constant responsible for the acceleration of the universes expansion, and CDM stands for cold DM. The term cold refers to the DM being non-relativistic, capable of forming structures such as galactic halos. These findings are summarized in Figure 2.6, and the estimated pa- rameter values are shown in Table 2.1. This shows that DM is approximately five times more abundant than baryonic matter, highlighting the importance of studying a form of matter that constitutes such a significant portion of the matter-energy density of the universe. Figure 2.6: Power spectrum of the CMB and relation of its peaks with the cosmological parameters [20]. Parameter Value Contribution to the total mass-energy density (%) ΩΛh 2 (Dark Energy) 0.3107± 0.0082 68.5 ± 3.20 Ωmh 2(Total Matter) 0.143± 0.001 31.5 ± 0.859 Ωch 2 (Cold Dark Matter) 0.120± 0.001 26.5 ± 0.757 Ωbh 2 (Baryonic Matter) 0.0224± 0.0001 4.94 ± 0.122 Table 2.1: Current physical densities (Ωih 2) as cosmological parameters of the ΛCDM model, in terms of the dimensionless Hubble parameter h = 0.674± 0.005 = H0 100kms−1Mpc−1 [18]. 10 CHAPTER 2. EVIDENCE FOR DARK MATTER 2.3 Gravitational Lensing in Galaxy Clusters A consequence of general relativity is that the path followed by light curves significantly near massive objects. This is observed as a deflection of the light received by the observer relative to that emitted from the source and as a magnification of the received image [21]. Detecting these deflected and magnified images indicates the presence of a nearby massive object; however, in some cases, this massive object does not interact electromagnetically. The deflection angle α depends on the mass M of the object and the radial distance d between the light beam and the object, also known as the impact parameter [22], as shown in Figure 2.7. The relationship between these quantities is: α = 4GM c2d . (2.8) In Equation (2.8), G is Newton’s gravitational constant, and c is the speed of light in a vacuum. From this relationship, it can be seen that the deflection effect will be greater as the object’s mass increases and as the light beam passes closer to the object. Figure 2.7: Light beam deflected by an angle α by an object of mass M , with an impact parameter d [23]. Considering the magnification effect, the study of gravitational lensing is divided into strong and weak lensing regimes. By studying images taken by the Hubble Space Telescope, among others, strong gravitational lensing has been found on numerous occasions in regions where there is not enough baryonic matter to account for it [24], such as in the galaxy cluster CL0024+17 [25–27] (see Figure 2.8). Similarly, evidence of DM has been discovered through the study of weak gravitational lensing [24, 28, 29]. A case of particular interest is the cluster 1E 0657–56, known as the “Bullet Cluster”. This cluster is actually the collision of two galaxy clusters, observed using optical methods to study the gravitational lensing effect present [31], as well as by the Chandra X-ray observatory satellite [32]. The gravitational lensing study found where most of the mass is concentrated in this cluster, while X-ray observations identified where the highest density of baryonic matter is located. These results were compared (see Figure 2.9), revealing that the region where the gravitational lensing effect is most pronounced does not coincide with the region containing the baryonic matter. This implies that in the Bullet Cluster, most of the mass does not correspond to baryonic matter but rather to a type of matter that exerts a gravitational lensing effect but does not interact electromagnetically, as expected for DM. Although the Bullet Cluster is the most well-known example, other galaxy cluster collisions showing a similar disparity between the gravitational lensing effect and the observed amount of baryonic matter have also been discovered. Notable examples include MACS J0025 (the 2.3. GRAVITATIONAL LENSING IN GALAXY CLUSTERS 11 Figure 2.8: Galaxy Cluster CL0024+17. On the left, magnified and distorted images due to strong gravitational lensing are visible. On the right, in blue, is a reconstruction of the regions where DM is likely located [30]. Figure 2.9: Galaxy Cluster 1E 0657–56, known as the “Bullet Cluster”. In pink, the regions with high baryonic matter density are shown, and in blue, the regions where the gravitational lensing effect is strongest, possibly due to dark matter [33]. “Baby Bullet”) [24], Abell 520 (the “Train Wreck Cluster”) [34], Abell 2744 (the “Pandora Cluster”) [35], and ZwCl 1358+62 [36] (see Figure 2.10). The observations presented in this chapter represent some of the most compelling evidence suggesting the existence of DM. Although all are gravitational in nature, they arise from a diverse set of measurements involving physical processes across a wide range of scales and cosmic epochs. Theoretical models of DM candidates, along with methods of detection, are described in the following chapter. 12 CHAPTER 2. EVIDENCE FOR DARK MATTER Figure 2.10: Galaxy clusters as evidence for dark matter. Upper left: ZwCl 1358+62, the dark matter map is shown in blue, in pink the ordinary matter [36]. Upper right: Abell 520, hot gas is shown in green, dark matter in blue [37]. Lower left: MACS J0025, showing dark matter in blue and ordinary matter in pink [38]. Lower right: Abell 2744, dark matter is shown in blue, hot gas in red [39]. 3 Dark Matter Candidates and Methods of Detection In order to explain the observations outlined in Chapter 2, any particle that constitutes dark matter (DM) must have the following characteristics: • It must be electrically neutral to avoid any electromagnetic interaction. • It must be massive and non-relativistic to form cosmological structures. • It has to be stable, with a lifetime equal to, or greater than, the age of the universe in order to produce acoustic oscillations in the CMB. No particle in the Standard Model (SM) meets all of these criteria: electrons, muons, taus, quarks, and W± are electrically charged; photons are massless; neutrinos are not massive enough and could not form galactic halos and other structures attributed to DM due to being ultrarelativistic; the Higgs and Z bosons, and the gluons are all ruled out by the stability requirement. It becomes clear that DM must be a gateway to physics beyond the SM. Figure 3.1: Mass range for various dark matter candidates [40]. A wide variety of ideas have been proposed, originating from different theoretical frame- works and addressing various problems. As shown in Figure 3.1, DM candidates cover an extensive range of masses: from 10−6 eV QCD axions, to primordial black holes an order of magnitude more massive than our Sun. In this chapter, the defining properties of some of the most studied candidates will be presented. Specifically, weakly interacting massive particles, axions, and sterile neutrinos, all of which play an important role in the research presented in this work. 13 14 CHAPTER 3. DARK MATTER CANDIDATES AND METHODS OF DETECTION 3.1 Weakly Interacting Massive Particles The most extensively studied DM candidates are the Weakly Interacting Massive Particles (WIMPs). These hypothetical particles are expected to interact gravitationally and, at most, via the weak nuclear force. WIMPs naturally satisfy the conditions required for DM, making them a common subject of theoretical and experimental research [41–44]. A widely studied mechanism by which WIMPs may have been produced is known as thermal freeze-out [45]. In the early universe, when temperatures were much higher than the WIMP rest mass, WIMPs were in thermal equilibrium with the plasma. They were produced and annihilated at equal rates through interactions with SM particles. As the universe expanded and cooled, the temperature of the plasma dropped below the WIMP mass threshold. This caused annihilation interactions to dominate. Once the interaction rate dropped below the Hubble expansion rate, WIMPs decoupled from the thermal equilibrium, and their abundance effectively “froze out”, resulting in a WIMP relic density. This would have occurred approximately 10−10 seconds after the Big Bang, during the electroweak era. For a WIMP mass in the range of 100 GeV - 1 TeV, the predicted relic density matches the currently observed DM density (see Figure 3.2). This coincidence is often referred to as the WIMP “miracle”, and it has served as strong motivation for WIMP-focused DM investigations. As discussed in Section 2.2, DM left observable evidence of its existence in the baryon acoustic oscillations of the CMB. These anisotropies indicate that DM already existed well before recombination. This is consistent with thermal freeze-out as the proposed produc- tion mechanism for WIMPs, and further supports the ongoing theoretical and experimental interest in WIMPs as viable DM candidates. Figure 3.2: Comparison of WIMP and dark matter relic densities (ΩX and ΩDM respectively) as a function of WIMP mass (mX). Both densities coincide for WIMP masses in the range 100 GeV - 1 TeV [41]. 3.2. AXIONS 15 3.2 Axions Axions1 were not originally proposed as DM candidates. Instead, they arise from a possible solution to the strong CP problem: the question of why CP (conjugation of charge and parity) symmetry violation is not observed in strong interactions. The CP-violating term in the QCD Lagrangian is proportional to Θ, known as the strong CP phase, which can take values between 0 and 2π. In order to avoid observable CP violation, such as a large neutron electric dipole moment, Θ must be tuned to extremely small values of the order of Θ < 10−10 [46]. Therefore, the strong CP problem is formulated as a naturalness, or fine-tuning, problem related to the value of Θ. The solution proposed by Roberto Peccei and Helen Quinn [47] involves introducing a new global symmetry to the SM, known as the Peccei-Quinn (PQ) symmetry, U(1)PQ. When spontaneously broken, this symmetry produces a pseudo-Goldstone boson, known as the axion. The axion field dynamically relaxes the CP violating phase Θ to zero, thus eliminating the need for fine-tuning. The mass of the axion (ma) is given by [48]: ma = (5.70± 0.07)µeV ( 1012GeV fa ) , (3.1) where fa is the axion decay constant. As shown in Figure 3.3, the axion mass is limited by cosmological and astrophysical observations to a range of 10−11 < ma[eV ] < 10−2 [49]. Axions require additional considerations related to the breaking of the Peccei–Quinn (PQ) symmetry [50], specifically regarding whether the symmetry is spontaneously broken during inflation and whether it remains broken afterward. Regardless of whether both conditions are met or at least one is violated, axions can still account for all or a significant fraction of the currently observed DM. This motivates their study and experimental search as one of the leading DM candidates. Figure 3.3: Masses below 10−22 eV, above 10−2 eV, and within the shaded region are ex- cluded. [49]. Constraints from fuzzy dark matter and black hole spins assume negligible axion self-interactions. The SN 1987A limit applies only to QCD axions. Regions beyond “The QCD axion” correspond to axion-like particles (ALPs) discussed in Chapter 6. 1The viable dark matter candidate axions described here are often referred to as QCD axions, due to their role as part of a proposed explanation to the strong CP problem. 16 CHAPTER 3. DARK MATTER CANDIDATES AND METHODS OF DETECTION 3.3 Sterile Neutrinos Neutrino oscillation between the flavor states (e, µ, τ) provides unequivocal evidence of neutrino masses (discussed in Section 7.1). There is no explanation within the SM for the mechanism by which neutrinos acquire mass; representing one of the most compelling indications for the need of new physics. One possible explanation for neutrino masses is the see-saw mechanism [51–53]. In this framework, right-handed neutrinos are introduced; they must be neutral under both electric and weak interactions, which is why they are referred to as sterile neutrinos. The most general renormalizable mass terms in the lepton sector, including the sterile neutrinos, consist of a Dirac mass term that couples left and right-handed states, and a Ma- jorana mass term for the right-handed neutrinos. When the electroweak symmetry breaking occurs, the Higgs field acquires a vacuum expectation value of ⟨H⟩ ≈ 247 GeV, and neutrinos propagate as mass eigenstates that diagonalize the mass matrix. Consider, for simplicity, one generation: mν(1gen) = ( 0 m m M ) , (3.2) where m represents the Dirac term mass, and M the Majorana mass. This matrix is diago- nalized by the physical masses expressed as: mν(1gen)1,2 = √ m2 + 1 4 M2 ± 1 2 M. (3.3) In the limit where M >> m, the solutions can be approximated by mheavy >> mlight as: mheavy ≈ M, mlight ≈ m2 M ≈ m2 mheavy . (3.4) Equation 3.4 shows an inversely proportional relation between mheavy and mlight, which is the reason for the name of the see-saw mechanism. By assuming a Yukawa coupling in the range of 10−8 − 10−5 and taking the light neutrino mass as mlight ≈ 0.05 eV (inferred from atmospheric neutrino oscillations [54]), the corre- sponding heavy neutrino masses, mheavy, are estimated to lie between 1 keV and 1 TeV, approximately in the electroweak scale. Sterile neutrinos are compelling DM candidates due to their minimal interactions with SM particles, potential masses and stability. Additionally, sterile neutrinos could play a role in leptogenesis [55–57], potentially providing insights into the matter-antimatter asymmetry observed in the universe. These characteristics make sterile neutrinos an interesting subject in the search for DM and physics beyond the SM. 3.4 Methods of Detection The DM candidates discussed in the previous sections are well-motivated by theoretical arguments. However, like all proposals for new physics, their existence must ultimately be 3.4. METHODS OF DETECTION 17 confirmed through experimental detection. Figure 3.4 illustrates the possible interactions between SM particles and DM involved in the three main detection strategies: production, direct detection, and indirect detection. These methods are outlined in this section. Figure 3.4: Methods of detection for dark matter. SM denotes a Standard Model particle, χ a dark matter particle. Production: SM+ SM → χ+ χ+ SM. Direct detection: χ+ SM → χ+ SM. Indirect detection: χ+ χ → SM + SM. 3.4.1 Production The production approach aims to detect DM by creating it through interactions between SM particles. Such processes could occur through interactions of the following form: SM + SM → χ+ χ+ SM. (3.5) In collider experiments, DM particles (χ) are hypothesized to be produced through inter- actions among SM particles. These DM particles, being weakly interacting, would not be detected directly. Instead, their presence would be inferred from the detection of SM particles resulting from subsequent decay processes, alongside a measurable imbalance in transverse momentum, referred to as missing transverse momentum, which corresponds to the unde- tected χ particles. The main effort in DM production is carried out by the European Council for Nuclear Research (CERN) at the Large Hadron Collider (LHC) in Geneva, Switzerland. These searches are led by the ATLAS and CMS collaborations. So far, no conclusive evidence of any deviation from the SM has been observed in proton-proton collisions at a center-of-mass energy of √ s = 13 TeV [58, 59]. Although no DM particles have been detected, these searches 18 CHAPTER 3. DARK MATTER CANDIDATES AND METHODS OF DETECTION have set increasingly stringent limits on various theoretical models. Proposed colliders such as the Future Circular Collider (FCC) [60], and the Circular Electron Positron Collider (CEPC) [61] are expected to lead the next generation of DM production experiments. 3.4.2 Direct Detection Direct detection of DM occurs when it interacts with ordinary matter and the products of that interaction include both DM and detectable SM particles. Direct detection differs from indirect detection in that the interactions occur within the detector itself, whereas indirect detection involves observing the byproducts of interactions that take place elsewhere. Direct detection of DM is a central focus of this work, investigated in Chapters 4, 5 and 6. In direct detection experiments, DM particles are hypothesized to interact with SM particles through scattering processes. The primary types of scattering are elastic and inelastic. In elastic scattering, the total kinetic energy of the system is conserved, meaning the particles involved emerge from the interaction without any change in their internal states. This process can be represented as: χ+ SM → χ+ SM. (3.6) Ongoing direct detection experiments employ a variety of detection media and techniques. The main detectors currently in operation include: • DEAP-3600: a 3.6 tonne liquid argon detector located in SNOLAB, Canada [62]. A detailed description is provided in Chapter 4. • LUX-ZEPLIN: a 7 tonne liquid xenon detector, located at the Sanford Underground Research Facility (SURF) in South Dakota, USA [63]. Studied in Section 7.6. • XENONnT: a 5.9 tonne liquid xenon detector [64]. Located at the Gran Sasso National Laboratory (LNGS) in Italy, it is an upgrade of the XENON1T experiment [65]. • PandaX-4T: a liquid xenon detector with a sensitive target mass of 3.7 tonne [66], situated at the China Jinping Underground Laboratory (CJPL) in Sichuan, China. • SuperCDMS: also located at SNOLAB, this experiment uses germanium and silicon crystals to search for DM [67]. • ADMX: an axion direct detection experiment [68] sited at the Center for Experimental Nuclear Physics and Astrophysics (CENPA) at the University of Washington. ADMX uses a magnetic field to convert axions into detectable microwave photons. • CRESST: also sited at Gran Sasso, this experiment utilizes calcium tungstate (CaWO4) crystals for DM detection [69]. • PICO-40L: a bubble chamber detector hosted at SNOLAB, it uses C3F8 as its target fluid to detect dark matter particles [70]. 3.4. METHODS OF DETECTION 19 There is currently no significant evidence of DM from direct detection experiments. How- ever, these results have allowed researchers to place increasingly stringent upper limits on the interaction cross section between WIMPs and the nucleons of the noble elements used in detectors. The most stringent limit on the spin-independent WIMP-nucleon cross section is presently held by LUX-ZEPLIN (shown in Figure 3.5). Figure 3.5: 90% CL. upper limits on the spin-independent WIMP-nucleon cross section set by LUX-ZEPLIN [71]. 3.4.3 Indirect Detection Indirect detection searches involve detecting SM products resulting from the interaction between two or more DM candidates. These interactions are commonly modeled as annihi- lation or decay processes: χ+ χ → SM + SM. (3.7) While indirect detection does not allow for the direct observation of DM itself, it can distinguish between various candidates based on the products that would result from anni- hilation processes. For WIMPs, the detectable products of annihilation include gamma rays, neutrinos, antiprotons, and antideuterons, as illustrated in Figure 3.6. 20 CHAPTER 3. DARK MATTER CANDIDATES AND METHODS OF DETECTION Figure 3.6: Annihilation of WIMPs at a center-of-mass energy of 100 GeV, and their Stan- dard Model products [72]. Examples of experiments that search for DM through indirect detection include: Fermi space telescope [72], IceCube Neutrino Observatory [73], the VERITAS telescope array in Arizona [74], HAWC Cherenkov Detector [75], Super-Kamiokande Neutrino Observatory [76], and the Pierre Auger Observatory [77]. Indirect detection experiments have not yet provided conclusive evidence of DM annihilation or decay. However, these experiments have been crucial in constraining interaction cross sections and decay rates, providing valuable information for theoretical models. A distinct approach from the previously described detection methods is taken by beam- dump experiments, which combine both production and direct detection mechanisms. In these experiments, DM is both produced and detected within the same setup. A proton beam is directed onto a target, where interactions are expected to produce DM particles with masses in the MeV to a few GeV range. These particles are then searched for via direct detection. One example is the MiniBooNE-DM experiment, which consists of a spherical detector filled with 818 tons of mineral oil [78]. In the following chapters, various contributions to the DEAP-3600 direct DM detection effort are presented. Chapter 5 focuses on the radiogenic neutron background, presenting a framework for its characterization and exploring a machine learning–based approach to neutron discrimination. The gamma background is refined for a solar axion-like particle search in Chapter 6. Lastly, in Chapter 7, DEAP-3600 and LUX-ZEPLIN are the subject of a SM and beyond analysis, using a simulated neutrino signal. 4 The DEAP-3600 Detector 4.1 Description DEAP-3600 is a DM detector, located in SNOLAB, Ontario, Canada. It is kept 2 km underground to provide shielding from muons and other backgrounds, and is operated by the DEAP collaboration [62]. It is mainly used in the search for weakly interacting massive particles (WIMPs) [79–81], however, there are also published results and ongoing efforts to: restrict the non-relativistic effective field theory interactions between WIMPs and argon [82], search for Planck-scale multiple interacting massive particles [83] and neutrinos [84], measure the specific activity of 39Ar in atmospheric argon [85] and measure the scintillation quenching factor of alpha particles in liquid argon [86]. The main principle by which DEAP-3600 can be used for particle detection is scintillation light emission from nuclear recoil interactions in liquid argon, this process will be explained in detail in Section 4.2. The outer structure of the detector (see Figure 4.1) consists of a cylindrical tank of 7.8 m of diameter and height, filled with water kept at a temperature of 12 ◦C [87]. Inside this tank there is a 3.4 m diameter spherical steel shell, which is refrigerated using liquid nitrogen. The spherical shell is supported by a 30 cm diameter tube which holds it from the top of the outer cylinder. This shell is equipped with 48 Hammamatsu photo-multiplier tubes (PMT) used for filtering out muons by their Cherenkov radiation emitted in the water tank. Inside the spherical shell (see Figure 4.2) is an acrylic vessel, which is a sphere of an inner and outer radii of 85 and 90 cm respectively. The liquid argon is kept in this vessel, at a pressure of approximately 1 atmosphere and a temperature between 87 and 85 K. The inner surface of the acrylic vessel is coated with a layer of tetraphenyl-butadiene (TPB) which acts as a wavelength shifter for the scintillation emitted by the argon, shifting it from UV to visible light. In addition to the muon veto PMTs, inside the spherical shell there are 255 Hamamatsu R5912 tubes, which amplify the signal generated by the scintillation light after it has been shifted by the TPB layer, with a typical gain factor of 8.0× 105. 21 22 CHAPTER 4. THE DEAP-3600 DETECTOR Figure 4.1: Outer structure of the DEAP-3600 detector [87]. Figure 4.2: Acrylic vessel and inner structure [81]. 4.2. SCINTILLATION FROM NUCLEAR RECOILS IN LIQUID ARGON 23 4.2 Scintillation from Nuclear Recoils in Liquid Argon As mentioned in Section 4.1, the principle by which DEAP-3600 can be used for DM detection is when a nuclear recoil occurs. This could happen if a WIMP from the DM halo scatters elastically with an argon nucleus in the detector (see Figure 4.3), this is called a nuclear recoil (NR). In this interaction, the WIMP could deposit energy in the nucleus according to Equation 4.1, where µ is the reduced mass of the nucleus and the WIMP, v is the relative velocity between both in the center of mass frame and mN is the mass of the nucleus [88]. For a 100 GeV WIMP, the energy deposited in an argon nucleus via elastic scattering is low, of the order of a few tenths of keV [43]. Figure 4.3: Nuclear recoil produced by a WIMP scattering elastically on a target nucleus [89]. ER = µ2v2(1− cosθR) mN . (4.1) When an argon atom gains energy via elastic scattering, it is left in an excited state (Ar∗) [90]. This excited atom, called exciton, might then combine with another argon atom, resulting in the formation of an excimer (Ar2), a short-lived bound state. The excimer subsequently separates, emitting scintillation light in the ultraviolet part of the spectrum, which peaks at 128 nm [91]. This process is outlined in Equation 4.2. The energy of the scintillation UV light corresponds to 9.8 eV, which is lower than the energy required to excite the ground state electrons in an argon atom to the first excited state, approximately 12 eV [92], meaning that liquid argon is transparent to its own scintillation light. This light is then shifted to the visible part of the spectrum by the TPB layer in the inner surface of the acrylic vessel and its signal is amplified by the multiplier tubes, which allows for the nuclear recoil to be detected. Ar + energy → Ar∗ Ar∗ +Ar → Ar2 Ar2 → Ar + Ar + γ(128nm) (4.2) 4.3 Pulse Shape Discrimination The scintillation process described in Section 4.2 can also be triggered by scattering in- teractions with the electron cloud, which can occur between the electrons in the argon atom 24 CHAPTER 4. THE DEAP-3600 DETECTOR and other electrons, or gamma rays. In this case the process is called electron recoil (ER). An ER results in scintillation light with the same peak wavelength, making it difficult to differentiate from an NR. There is, however, one key difference between NR and ER that allows for discrimination between both: a significant fraction of the scintillation light from an NR is emitted in a shorter time duration than in an ER. This is explained by the fact that NR interactions are more prone to form excimers in the singlet spin coupling state, while ERs form triplet spin state excimers with a higher probability. The singlet has an approximate lifetime of 7 ns, and the triplet one of 1600 ns [93]. This difference is exploited in a method called pulse shape discrimination (PSD), by defin- ing an appropriate time span during which to register the fraction of the light captured by the PMTs and comparing that to the approximate duration of the event. Following this idea, a ratio can be defined in which the resulting fraction has a large enough difference between nuclear recoils and electron recoils that allows for reliable discrimination between both. This ratio receives the name Fprompt, and Equation 4.3 shows how it is defined mathematically [81]. The shorter (prompt) time window has a duration of approximately 88 ns, while the longer one lasts for over 10 µs. Fprompt is a ratio of the voltage gathered by the PMTs during these time windows; it takes a typical value of ≈ 0.7 for nuclear recoils, and ≈ 0.3 for electron recoils (see Figure 4.4). This way, nuclear and electron recoil processes can be differentiated from each other. Fprompt = ∑60ns t=−28ns PE(t) ∑10µs t=−28ns PE(t) . (4.3) The parameter space formed by Fprompt and PE, the amount of photoelectrons captured by the PMTs that account for the voltage in the signal, allows for a clear visual representation of NRs and ERs in two separate bands (see Figure 4.5). If the expected signal is a nuclear recoil, the electron recoil band can be discarded as background, and vice-versa. An additional feature of the Fprompt-PE parameter space is allowing for background mitigation based on the energy of the interaction, because processes with higher energies will produce a higher value of PE. The methods by which backgrounds are mitigated are discussed more in detail in Section 4.4. 4.4. BACKGROUND SOURCES AND REGION OF INTEREST 25 Figure 4.4: Scintillation signal comparison of electron recoils (up) and nuclear recoils (down). The Y axis represents the voltage of the signal captured by the PMTs, the X axis is the time during which the signal was detected. The prompt time duration is represented in yellow, while the longer duration is in the light blue area. The nuclear recoils have a higher fraction of their total signal in the yellow-to-blue ratio than the electron recoils [94]. 4.4 Background Sources and Region of Interest As was discussed in Section 4.2, the energy deposited by a WIMP in a nuclear recoil is low, of the order of a few tens of keV, this means that minimizing background interactions is of paramount importance for a DM direct detection experiment. The main background sources that can trigger a scintillation process in liquid argon are: electrons, gamma rays, alpha particles and neutrons. Not all of these backgrounds will trigger a nuclear recoil, making PSD an effective method in mitigating the backgrounds that produce electron recoils (see Section 4.3 for a detailed discussion of PSD). In this section, a description of each of the main background sources, and the physical processes from which they arise, will be made, accompanied with an argument on how each source can be mitigated to possibilitate a WIMP search. 26 CHAPTER 4. THE DEAP-3600 DETECTOR Figure 4.5: NR and ER bands in the Fprompt and PE space [95]. 4.4.1 Electrons and Gamma Rays The dominant source of electrons are the β-decays occurring in the 39Ar isotope present in the LAr. Electrons also originate from 42Ar and its daughter 39K, which have a lower activity but dominate the background spectrum for energies higher than 2.6 MeV [96]. These isotopes are of cosmogenical origin, and are present in any argon sample extracted from the atmosphere. The gamma rays mainly originate from decays of 232Th, 238U and 40K, all of which are present in the structural components of the detector. Gammas can also be produced as emis- sions from neutron captures, which can arise from (α, n) and spontaneous fission reactions in other components of the detector that act as neutron sources, such as the borosilicate glass in the PMTs. Both gamma rays and electrons can result in scintillation light emission if they interact with the 40Ar, however, these backgrounds do so via electron recoil. This means that the scintillation emitted from these interactions will occur during a longer time, and thus, can be reliably identified using the PSD method described in Section 4.3. An additional electromagnetic background process induced by electrons and gamma ray interactions is the Cherenkov radiation that can be emitted in the PMT glass and light guides. This radiation has a significant UV component which can be detected by the PMTs as scintillation light, resembling the light emitted from a nuclear recoil. PSD can also be employed to mitigate these backgrounds, but in this case the prompt fraction of light in the nuclear recoils is lower in comparison. Due to the fact that Cherenkov radiation emission occurs in very short time frames, < 1 ns, these events have an Fprompt value of ≈ 0.9, making it significantly higher than ≈ 0.7 for the NR band [81]. This allows for the successful identification of these processes as not NR in origin. 4.4. BACKGROUND SOURCES AND REGION OF INTEREST 27 4.4.2 Alpha Particles Alpha particles in DEAP-3600 originate mainly from decays of 222Rn, 220Rn y 210Po [95], which are present in the outer components of the detector, such as the piping and the AV. An example of one of the background alpha decays is illustrated in Equation 4.4, the isotope 222 84 Rn decays into 218 82 Po and an alpha particle represented by a helium nucleus. 222 84 Rn →218 82 Po +4 2 He (4.4) Alpha particles interact with the LAr via NR [97], however, as can be seen in Figure 4.5, the alphas produced in the decay of 222Rn, 210Po, 218Po and 214Po, although reconstructed in the NR band, are registered at much higher energies than what an NR by a WIMP from the galactic halo would produce. This allows for these alphas to be discriminated as background even-though they also generate NR scintillation. However, alpha particles that deposit only a fraction of their energy can produce detection signals that overlap with the WIMP region of interest. This region is the part of the Fprompt and PE parameter space designated for WIMP interactions, as described in Section 4.4.4. Most of these alpha particles are emitted by decays that occur in the AV, the TPB coating and the neck of the detector. These are all outer components (see Figure 4.2), and the background effects of these alphas can be mitigated by reducing the fiducial volume of the LAr used in the search. This is done in practice by imposing a radial cut at 750 mm and a cut in the vertical direction (along the neck of the detector) at 550 mm, measured from the geometrical center of the AV. 4.4.3 Neutrons Neutrons can interact with argon atoms through NR, and this interaction may occur within the same energy range as an NR caused by a WIMP. There are two main categories of processes from which neutrons can originate as a background source: cosmogenic and radiogenic. In the present Section both of these types of processes will be examined, and the methods for their mitigation will be described. Cosmogenic neutrons are induced by atmospheric muons in (µ, n) reactions. The con- tribution of these neutrons as a background source is greatly reduced by the 2 km rock shielding of the nickel mine at which DEAP-3600 is located, which is equivalent to 6 km of water. An additional layer of mitigation is provided by the cylindrical water tank in which the detector is submerged [87]. If any atmospheric muons penetrate to that depth, they can still be mitigated by the Cherenkov radiation emitted in the water tank, which is detected by the external PMTs (see Figure 4.1). Radiogenic neutrons arise from the inherent radioactivity in the materials from which the detector was built. The most significant contributions are those of (α, n) reactions and spontaneous fission. Approximately 70% of the background interactions from radiogenic neutrons are produced in (α, n) reactions from alpha decays of 238U and 232Th [98]. Isolating radiogenic neutrons is a difficult task, since they produce NR interactions that overlap with the energy ranges at which WIMP recoils occur, and can’t be discriminated by Cherenkov light like the cosmogenic variant of neutrons. The focus for this background source shifts to estimating its contribution and subtracting the estimated value from the events 28 CHAPTER 4. THE DEAP-3600 DETECTOR detected. This is achieved with the use of Monte Carlo simulations, which are performed using software such as Reactor Analysis Tool (RAT [99]) and NeuCBOT [100], which are widely used tools for simulations and analysis at the DEAP collaboration. 4.4.4 Region of Interest and Background Estimates The Fprompt and PE parameter space is used to define a region of interest (ROI) in which the WIMP interactions can be identified after mitigating the background sources described in the previous sections (see Figure 4.6). The main criteria for defining the ROI is restricting it to the NR band and the energy range at which WIMPs can be reliably singled out from all the background sources. In the latest published WIMP search analysis [81], these criteria were used to further refine the background mitigation in the ROI without significantly decreasing the WIMP acceptance: • In the range of 95-160 PE, the curve is defined such that the number of expected electron recoil events is less than 0.05. • The curve from 160-200 PE is defined for a constant 1% loss of nuclear recoil acceptance. • The upper curve spanning the range 95-200 PE delineates a constant 30% loss of nuclear recoil acceptance, ensuring that the expected number of alpha particle events in the region of interest is less than 0.5. • The upper limit for PE is given by the kinematics of the WIMP-argon nucleus inter- action, for WIMPs in the mass range of 100 GeV. For higher PE values, the noise produced by alpha particles and neutrons is expected to be greater, while the fraction of expected WIMPs would not significantly increase. In the same analysis, an estimated contribution of events in the ROI was made for each of the background sources, which yielded the results shown in Table 4.1. The DEAP-3600 detector plays a central role in this work. Its radiogenic neutron back- ground is the subject of investigation in Chapter 5, where the neutron contribution to the background model is characterized, and a novel mitigation approach involving machine learn- ing is explored for future implementation. A contribution to the search for MeV-scale so- lar axion-like particles using DEAP-3600 is presented in Chapter 6. Lastly, in Chapter 7, DEAP-3600’s sensitivity to coherent elastic neutrino-nucleus scattering from supernovae is examined, along with potential implications for the Standard Model and beyond. 4.4. BACKGROUND SOURCES AND REGION OF INTEREST 29 Figure 4.6: Region of interest (in black) defined for the WIMP search [81]. The Y axis represents the Fprompt parameter and the X axis the photoelectrons, and its equivalent nuclear recoil energy in keV. Background Source NROI ERs from βs and γs 0.03 ± 0.01 Cherenkov from βs and γs <0.14 Neutrons (Radiogenic) 0.10+0.10 −0.09 Neutrons (Cosmogenic) <0.1 Alphas from AV Surface <0.08 Alphas from AV Neck 0.49+0.27 −0.26 Total 0.62+0.31 −0.28 Table 4.1: Background Estimates for the latest published WIMP search, separated by each source and the contribution to the ROI in number of events (NROI) [81]. 30 CHAPTER 4. THE DEAP-3600 DETECTOR 5 Neutron Background in DEAP-3600 Dark matter direct detection experiments aim to observe very rare signal events, while accounting for commonly occurring background events. This requires a background model capable of providing sufficient sensitivity to establish upper limits or discovery contours, depending on the case. The following chapter presents this work’s contribution to modeling and mitigating the radiogenic neutron background in DEAP-3600. Radiogenic neutrons constitute the most challenging background to reduce, as neutron-induced nuclear recoils are indistinguishable from WIMP interactions when relying solely on PSD, as described in Section 4.4.3. The radiogenic neutron background is modeled within a Profile Likelihood Ratio frame- work. By accurately characterizing the radiogenic neutron contribution, the analysis en- hances the experiment’s ability to distinguish potential DM signals from background-induced events. This contribution is part of the collaboration’s ongoing effort to perform a DM search with a longer livetime than that of the most recently published WIMP search [81]. Additionally, beginning in Section 5.4, a machine learning method to further mitigate ra- diogenic neutron backgrounds is introduced. This work constitutes an initial exploration into distinguishing neutrons from WIMPs by leveraging differences in the likelihood of observing multiple versus single nuclear recoils. 5.1 The Profile Likelihood Ratio Method The Profile Likelihood Ratio (PLR) method is a statistical framework widely used in DM detection experiments [54, 101]. It enables a simultaneous fit of signal and background events, while allowing for systematic uncertainties to be parametrized as nuisance parameters. As an example to illustrate the PLR method [102], consider a histogram measuring a variable x, the expected value of the number of events in the ith bin of the histogram (E[ni]) is written as: E[ni] = µsi + bi, (5.1) where µ is a parameter that characterizes the signal strength. The quantities si and bi are the mean number of signal and background events in the ith bin, which are written in terms of a probability density function as: si = stot ∫ i fs(x,θ), 31 32 CHAPTER 5. NEUTRON BACKGROUND IN DEAP-3600 bi = btot ∫ i fb(x,θ); (5.2) where fs and fb are the probability density functions (PDFs) of the variable x for signal and background events. The set of nuisance parameters characterizing the PDFs are given by θ, stot and btot are the total number of signal and background events in the histogram, and the integrals ∫ i fs(x,θ) and ∫ i fb(x,θ) are the probabilities that the signal and background events are found in the ith bin. The likelihood of observing a signal strength parameter µ, given a set of parameters θ, is L(µ,θ), and the profile likelihood ratio is defined as: λ(µ) = L(µ, θ̃(µ)) L(µ̂, θ̂) . (5.3) The numerator in the PLR definition denotes the likelihood of observing µ given θ̃, the set of parameters that maximizes L for that specific µ; that is, a conditional maximum of L. The denominator represents the global maximum of L, where the likelihood is maximized unconditionally for a signal parameter µ̂ and a set of nuisance parameters θ̂. The quantity λ(µ) can take values between 0 and 1, where values close to 1 indicate a good agreement between the measured data and the hypothesized value of µ 1. A test statistic that depends only on the signal strength parameter µ can be constructed from λ(µ) in Equation 5.3 for two different scenarios: imposing an upper limit (κµ), and calculating the significance of a discovery (κ0). For the case of imposing an upper limit, µ̂ takes lower values than µ, and the exclusion test statistic is defined as follows: κµ =    −2ln(λ) for µ̂ ≤ µ, 0 for µ̂ > µ. (5.4) Given that 0 ≤ λ ≤ 1, this implies that κµ ≥ 0. The minimum value, κµ = 0 corresponds to µ̂ = µ, this is the case in which the data is the most compatible with the signal hypothesis. As κµ increases, the data becomes more compatible with the background-only hypothesis. For an observed test statistic κµ,obs, a signal p-value is defined to test Hµ, the signal hypothesis, as follows: pµ = ∫ ∞ κµ,obs f(κµ|Hµ)dκµ, (5.5) where f(κ|Hµ) denotes the PDF for κ assumingHµ. The p-value pµ represents the probability of measuring data of equal or greater incompatibility with Hµ. The hypothesis Hµ is rejected if pµ is below a certain threshold; for example: if pµ ≤ 5%, Hµ is rejected at 95% CL. In the case of calculating the significance of a discovery, the discovery test statistic is defined as: 1By Neyman-Pearson’s lemma [103], the PLR is the most powerful test statistic when testing a null and alternate hypothesis, H0 and H1 respectively, in terms of observed data x that can reject H0. It can also be written as λ = L(H1|x) L(H0|x) . 5.2. RADIOGENIC NEUTRON SIMULATIONS 33 κ0 =    −2ln(λ(0)) for µ̂ ≥ µ, 0 for µ̂ < µ, (5.6) where κ0 ≥ 0, and λ(0) is the PLR for the background-only hypothesis H0. The value of most compatibility with H0 is κ0 = 0, while increasing values indicante less compatibility with H0. The null p-value, to test H0, is given by: p0 = ∫ ∞ κ0,obs f(κ0|H0)dκ0. (5.7) f(κ|H0) is the PDF for κ given the background-only hypothesis H0. This p-value defines the probability of observing a test statistic of greater incompatibility with H0. Analogous to the first case, if p0 is below certain threshold, H0 is rejected with the complementary CL. Due to its statistical advantages, the DEAP collaboration is actively working to implement the Profile Likelihood Ratio (PLR) method [90]. One key reason this method has become the standard across DM experiments is that, when the background model provides sufficient sensitivity, it outperforms other statistical approaches — as illustrated in Figure 5.1. This improved sensitivity enables the use of an expanded region of interest (ROI), as shown in Figure 5.2. Compared to the ROI described in Section 4.4.4, the upper bound of the region is a curve that extends up to values of 0.76 Fprompt, significantly expanding the area of the parameter space analyzed. In the following sections, this work’s contribution to modeling the radiogenic neutron background within the PLR framework is presented. 5.2 Radiogenic Neutron Simulations The simulated neutrons were generated as a collaborative work using the C++ based software Reactor Analysis Tool (RAT) [99], which combines features from Geant4 [106] and ROOT [107]. RAT is widely used in the DEAP collaboration for data generation and analysis. It allows for the simulation of physical processes ocurring in an accurate model of the DEAP-3600 detector. The simulations were performed in clusters to which the DEAP collaboration is granted access by the Canadian organization The Alliance [108]. The radiogenic neutrons simulated correspond to (α, n) reactions from decays in the lower part of the 238U chain (from 226Ra to 206Pb). Two systematic uncertainties were taken into account in these simulations: the scattering length of the TPB wavelength shifter and the refractive index of the liquid argon (LAr) in the detector, both of which play a key role in the propagation of scintillation light to the PMTs. A total of five datasets were generated: one using the nominal detector settings, and four representing the upper and lower bounds of variation for each of the two systematics. In order to optimize the usage of computing resources, the simulations were divided into two stages (illustrated in Figure 5.3): 1. A total of 9.96 × 108 nominal neutrons simulated isotropically from the borosilicate glass in the PMT array. These neutrons were simulated without the optical physical 34 CHAPTER 5. NEUTRON BACKGROUND IN DEAP-3600 Figure 5.1: Comparison of exclusion curves for the median spin independent WIMP-nucleon scattering cross section σmed as a function of WIMP mass, using the Poisson, Maximum Gap [104], and PLR methods [105]. The limits were imposed with a simulated dataset consisting of 10 background events and 1 signal event, as a function of different threshold energies. Figure 5.2: Comparison of the ROI used in the most recent dark matter published search [81] (described in Section 4.4.4), and the proposed ROI for the PLR analysis [90] . 5.3. PDFS FOR THE RADIOGENIC NEUTRONS 35 processes, such as scintillation light emission and recollection. The neutrons from this stage that reach the outer AV radius (901.449 mm) are counted, and their energy and momentum is captured to be used as input for stage 2. 2. An energy spectrum and angular distribution were obtained from the energies and momenta of the stage 1 neutrons that reached the outer radius of the AV. Using these distributions, five neutron datasets were generated at the outer AV radius: one corre- sponding to the nominal configuration and four representing the systematic variations. The number of simulated neutrons for each dataset (approximately 7.25×105) matched the number of stage 1 neutrons that reached the outer AV radius. In this second stage, optical processes were enabled, modeling the scintillation light collected by the detec- tor. Figure 5.3: Amplified view of the internal components of DEAP-3600 [81], showing the two stages of the radiogenic neutron simulations. Stage 1 (blue): 9.96× 108 neutrons simulated isotropically from the PMT borosilicate glass. Stage 2 (red): 7.25× 105 neutrons simulated for each dataset (nominal and 4 systematics), from the outer AV radius (901.449 mm), with the energy spectrum and momentum distribution captured from the stage 1 neutrons. This is exemplified for 1 PMT, the simulations accounted for the entirety the PMT array. Additional properties of the radiogenic neutrons generated in the PMT array are listed in Table 5.1. 5.3 PDFs for the Radiogenic Neutrons The goal of this analysis was to produce PDFs for the nominal and systematic radiogenic neutron datasets using functions that accurately represent the histogram data, without being more complex than necessary. The purpose of these PDFs was to characterize the neutron 36 CHAPTER 5. NEUTRON BACKGROUND IN DEAP-3600 PMT Array Parameter Value Mass 188.96± 3.78 kg [109] Purity (lower 238U) 175.64± 4.93 ppb [110] Yield (lower 238U) (132.0± 26.4)× 10−12 neutrons g×s×ppb [111] Table 5.1: PMT array mass, purity and neutron yield used in the simulations. The yield for the lower part of the 238U chain was taken from existing data calculated using the SOURCES4C software [111]. background distribution for the PLR analysis, analogous to fb(x,θ) in Equation 5.2. The variables considered for describing the background sources in the PLR analysis were: • NSCBayes (NSC): a measure of the energy captured from the interactions. It is mea- sured in qPE (labeled as PE for simplicity), which correspond to the pulse charge of an event collected by a PMT, divided by the average single-photoelectron charge in the PMT. • Pulse Index First Gaseous Argon (PIFGAR): the number of PMTs oriented toward the LAr volume that registered light from an event before any PMTs in the gaseous argon (GAr) region detected a signal. • MblikelihoodR (MBR): reconstructed radial position of the event. • Fprompt (RPR): a measure of the fraction of the scintillation light emitted during the prompt window (described in Section 4.3). It is used to discriminate nuclear and electron recoils. Referred to as Rprompt (RPR) in the most recent analyses. These variables were used to construct histograms, and the PDFs were fitted to the his- tograms (see Figure 5.4). The steps taken for constructing the PDFs are summarized as follows: 1. After applying cuts focused on selecting regions of parameter space where radiogenic neutrons reconstruct, the five datasets were divided in two subsets each according to their value of pifgar. Events with PIFGAR ≤ 2 were grouped in the PIF2 set, while all other events (PIFGAR > 2) were categorized as PIF3. The number of events in each of these pifgar categories was approximately equal. 2. A histogram of NSC was constructed for each PIFGAR set, in the range of 90 ≤ PE ≤ 990 which corresponds approximately to a nuclear recoil energy range of 52 ≤ keVnr ≤ 570. These histograms were then fitted, resulting in 10 PDFs for NSC (two PIFGAR subsets times 5 datasets of nominal + 4 systematics). 3. RPR and MBR histograms for each PIFGAR subset were subsequently fitted for spe- cific PE bins. This resulted in 14× 2× 5 = 140 fits for RPR, and 23× 2× 5 = 230 fits for RPR. 4. The parameters of the functions used for fitting the RPR and MBR histograms were later fitted as a function of the PE bin occupied by each respective histogram. 5.3. PDFS FOR THE RADIOGENIC NEUTRONS 37 Figure 5.4: Flowchart of the datasets in the neutron simulations. Five datasets, account- ing for nominal and the four systematic variations: the refractive index of the LAr (rfrUP and rfrDOWN), and the scattering length of the TPB wavelength shifter (tpbUP and tpb- DOWN). Each dataset was divided into two subsets: PIFGAR ≤ 2 and > 3 (labeled pif2 and pif3, respectively). The NSC PDFs were constructed for each of these two subsets. RPR and MBR histograms were fitted on specific PE bins, and the parameters of these PDFs were fitted as a function of the corresponding PE values, for all subsets. Figure 5.5: Normalized PIFGAR histogram for the nominal settings. The dataset was divided into two subsets: PIFGAR ≤ 2 and PIFGAR > 3. 38 CHAPTER 5. NEUTRON BACKGROUND IN DEAP-3600 A normalization convention was adopted by the collaborators studying the different back- ground sources in this analysis. The convention is to normalize the histograms to 1 event per liveday. For the radiogenic neutrons, the normalization weights (w) were obtained in the following manner: f = mPMT × pPMT × yPMT = 378.5 days−1, teq = 9.96× 108/f = 2.63× 106 days, w = 1/teq = 3.80× 10−7 days−1. (5.8) where mPMT , pPMT and yPMT correspond respectively to: the mass, purity and yield of the PMT array (shown in Table 5.1), and the time to which the simulations from stage 1 are equivalent to is teq. The corresponding PDFs and fitted parameters, refined through several iterations, are evaluated using the reduced χ2 metric provided by the ROOT package MINUIT [112], and are detailed in the following sections. The normalized histogram for PIFGAR is presented in Figure 5.5. 5.3.1 NSCBayes The NSC histograms as a function of qPE (labeled PE here for simplicity) were fitted with a sigmoid function of the corresponding form: f(PE) = N 1 + exp ( −PE−xs ts ) · exp ( −PE θ ) θ . (5.9) This function has four parameters: • The norm factor N , adjusts the maximum value of the sigmoid. • The midpoint of the sigmoid is characterized by xs. • The parameter ts describes the steepness of the sigmoid curvature. • The mean of the exponential term is described by θ. Accounting for all five datasets and the two PIFGAR subdivisions, a total of 4 × 5 × 2 = 40 parameter values were produced for the NSC histograms. An example of one of these histograms is presented in Figure 5.6. 5.3.2 Fprompt The Fprompt (labeled as Rprompt in the most recent analysis) histograms were obtained including the full 90-990 PE range, divided in 13 NSC bins of width 60 PE from 90 to 900 PE, and a final 90 PE bin to 990. This was motivated in obtaining sufficient statistics for each PE bin, while providing enough data points for the parameter fits. The Fprompt values considered are in the range 0.5-1.0, corresponding to the nuclear recoil band described in Section 4.3. 5.3. PDFS FOR THE RADIOGENIC NEUTRONS 39 Figure 5.6: NSC fit for the stage 2 neutrons (AV) of the pif2 tpbUP dataset. The parameter values are indicated in the legend, along with the reduced χ2 value. Fit status = 0 is an indicator provided by Minuit of a successful fit convergence. As an additional check, the histogram and the sigmoid area are seen to agree to within 0.5%. 40 CHAPTER 5. NEUTRON BACKGROUND IN DEAP-3600 The histograms resemble a gaussian distribution with a left skew. The function that performed best achieving a satisfactory goodness of fit, while trying to use only as many parameters as necessary, was a gaussian with an error function term, presented as follows: f(x) = N σ √ 2π · exp ( −1 2 ( x− µ σ )2 ) · ( 1 + erf ( λ(x− µ) σ √ 2 )) . (5.10) This function uses four parameters: • The norm factor N . • The mean of the gaussian distribution µ. • The standard deviation σ. • The λ parameter of the error function term. A negative λ results in a left skew, while a positive λ produces a right skew. A value of λ = 0 recovers a gaussian curve depending only on the three parameters described previously. The Accounting for all five datasets, the two PIFGAR subdivisions, and the 14 PE bins, a total of 4× 5× 2× 14 = 560 parameter values were produced for the RPR histograms. An example of one of these fits is shown in Figure 5.7. The four parameters in the curve defined in Equation 5.10 were plotted separately as a function of the PE bin of their respective histogram. As an additional consistency check for the RPR and MBR distributions, a second normalization factor was applied to each histogram to ensure that its integral matched the integral of the corresponding PE bin in the respective systematic and PIFGAR datasets. The parameters N and µ were fitted using a six parameter three-part piecewise function: a quadratic section, followed by a cubic portion, and ending with another quadratic. Conti- nuity conditions on the function and its first derivative fully determine the cubic component. This piecewise function is defined as follows: f(x) =            A1(x−X1) 2 + Y1, for x < X1 ax3 + bx2 + cx+ d, for X1 ≤ x < X2 A2(x−X2) 2 + Y2, for x ≥ X2, (5.11) The points (X1, Y1) and (X2, Y2) are the coordinates for the transition points, which coincide with the vertexes of the quadratic segments; A1 and A2 are additional parameters for the quadratic terms. The coefficients a, b, c, and d are defined by the continuity conditions as given below: 5.3. PDFS FOR THE RADIOGENIC NEUTRONS 41 Figure 5.7: RPR fit for the PIF2 rfrUP distribution, in the 90-150 PE bin. The best fit parameter values are included, along with the reduced χ2 for goodness of fit evaluation. The fit function, histogram and PE integrals are also shown. The histogram was normalized to ensure coincidence with the PE bin integral. 42 CHAPTER 5. NEUTRON BACKGROUND IN DEAP-3600 a = −2Y1 + 2Y2 X3 1 − 3X2 1X2 + 3X1X2 2 −X3 2 , (5.12) b = 3X1Y1 − 3X1Y2 + 3X2Y1 − 3X2Y2 X3 1 − 3X2 1X2 + 3X1X2 2 −X3 2 , (5.13) c = −6X1X2Y1 + 6X1X2Y2 X3 1 − 3X2 1X2 + 3X1X2 2 −X3 2 , (5.14) d = X3 1Y2 − 3X2 1X2Y2 + 3X1X 2 2Y1 −X3 2Y1 X3 1 − 3X2 1X2 + 3X1X2 2 −X3 2 . (5.15) Examples of parameter fits for N and µ are shown in Figure 5.8. The remaining parameters, λ and σ, were fitted using a linear and a quadratic function, respectively, and an example of these fits can be seen in Figure 5.9. Accounting for the number of parameters, the 5 datasets and the PIFGAR subsets, a total of 6 × 5 × 2 = 60 parameters were produced for N and µ, 2 × 5 × 2 = 20 for λ, and 3× 5× 2 = 30 for σ. 5.3.3 MblikelihoodR The NSC range for the MBR histograms was reduced to 90-990 PE, and the radial positions included were in the range 0-720 mm, this was decided to favor consistency in the fits without compromising the integrity of the most relevant radial positions in the distributions. The NSC range was distributed in 17 bins with a width of 30 PE from 90 to 600, followed by 5 bins of 60 PE up to 900, and one last 90 PE bin up to 990. The function used for the MBR histograms is an exponential as written below: f(x) = N θ exp (−x θ ) . (5.16) This exponential function has two parameters: • The norm parameter N . • The mean, given by θ. The total number of parameters produced for the MBR histogram fits is determined by the parameters in the function, the nominal and systematics datasets, the PIFGAR subsets and the number of PE bins, respectively: 2× 5× 2× 23 = 460. The MBR histogram fit for the tpbUP PIF2 dataset in the 120-150 PE bin is presented in Figure 5.10. The norm parameter N was fitted with a two segment piecewise function defined as: f(x) =      A+ s x−B , for 0 < x < x0 m · x+ ( A+ s x0−B −m · x0 ) , for x ≥ x0. (5.17) 5.3. PDFS FOR THE RADIOGENIC NEUTRONS 43 Figure 5.8: RPR N (top) and µ (bottom) parameter fits for the nominal PIF3 and nominal PIF2 datasets, respectively. Both parameters were fitted using the piecewise function defined in Equation 5.11. The best fit values and the reduced χ2 are shown in the legends. The blue dashed lines denote the limits of the 90-990 PE range. N Rprompt Para meter Fit nom (pif > 2) z 0.004 e- quad-cube-<¡u.ad +1 XI • 2.23 1e+02 ± J.898e_Ol se- '(1 • 3.277e-ú3 ± , .2 12e-04 X2 . 7.464e+02 ± 4.64Be_Ol L '(2 . L248e-{)3 ± , .226e-04 0.003 'V -, 1 Al . -1.62Be-Ú8 ± 2.370e-0lI +T A2 . -7.S53e-Q9 ± 6.52Se-09 sF- \'" Chi21NDF _ 10.469 / 8 Fit status . O 0.00 0.002 0.00 2e- 0.001 Se- "- + 0.00 'e- 1 ~ Se- j 0.000 100 200 300 400 soo 600 700 BOO 900 1000 PE mu Rprompt Parameter Fit nom (pif <= 2) 2e- ~ , , Xl .2.'".-Se I ~ § O., o., o. +1 ' e- 1+ 'f-tt , 9e- -~ 0.7 0.7 , e- -r- 0.7 7e- 0.7 Se- 100 200 300 400 soo 600 700 BOO 900 1000 PE 44 CHAPTER 5. NEUTRON BACKGROUND IN DEAP-3600 Figure 5.9: RPR λ (top) and σ (bottom) parameter fits for the rfrUP PIF2 and rfrUP PIF3 datasets, respectively. The parameter λ was fitted linearly, while a quadratic function was used for σ. The best fit values and the reduced χ2 are shown in the legends. sf- of- Sf- of- Sf- -, of- -, Sf- • 0.09 E º' • 0.08 0.07 0.06 0.05 0 .04 0.03 0.02 lambda Rprompt Parameter Fit rfrUP (pif <= 2) -h -L ~ , , Lineal F ~ pO • · 1.127 • • 00 ~ 3. 192.·0' p l ·· M ll . · 05~7 . r.oo. · ().I Chi2;N OF _ 16.376 113 100 200 300 400 500 600 700 800 900 1000 PE sigma Rprompt Parameter Fit rfrUP (pif > 2) d- + -- Polyoorn;aI Fit pO . 9.81Mf1-02 i 3.005&-03 pi • -2033<1-04 i ' .B3I)e·05 p2 • 1 .433f1-07 i L71l5e-08 Chi2lNOF • 19000 / 11 + 100 200 300 400 500 600 700 800 900 1000 PE 5.3. PDFS FOR THE RADIOGENIC NEUTRONS 45 Figure 5.10: MBR fit for the PIF2 tpbUP distribution, in the 120-150 PE bin. The best fit parameter values are included, along with the reduced χ2 for goodness of fit evaluation. The fit function, histogram and PE integrals are also shown. The histogram was normalized to ensure coincidence with the PE bin integral. 46 CHAPTER 5. NEUTRON BACKGROUND IN DEAP-3600 Figure 5.11: MBR N (top) and θ (bottom) parameter fits for the rfrUP PIF2 and tpbDOWN PIF3 datasets, respectively. The parameter N was fitted with the piecewise function defined in Equation 5.17, while θ was fitted linearly. The best fit values and the reduced χ2 are shown in the legends. The blue dotted lines represent the PE=90 lower bound of the NSC region. z O. " O. 2e- De- -O. 2e- -O. ' e- -O. 6e- o - 2 De- -4 De- De- - 8 De- N mblikelihoodR Parameter Fit rfrUP (pif <= 2) ~ , r 200 400 600 l Asymptotic t li near A = 6.92OO-08 ± 2.485e-08 B = 5O.493 ± 5.783e-02 s = -1 .63273e-05 ± 5.1 76e-05 x_O = 231 .606 ± 2.OO2e+OO m = 6.5376ge-13 ± 8.084e-1 4 Chi2INDF = 9,270 / 18 Fit status = O BOO 1000 PE Theta mblikelihoodR Parameter Fil tpbDOWN (pif > 2) Lineal Fit 00 _ · 1.07&. 02 ~ • . 0820 _00 o l .4 . 7060 · 0 2 ~ 3 .10 '0 · 1.r.l Chi2;N DF _ 3.3.687 130 Fit . tatus _ o t -1-ht t f :: ttt t t - ~ - 10 - 12 - 14 De- - 16 O 200 300 400 500 600 700 BOO 900 PE 5.4. NEUTRON DISCRIMINATION USING MACHINE LEARNING 47 The segment to the left of the transition point x0 is an asymptotic function, with a horizontal asymptote in A as x → ∞, and a vertical one in x = B. To the right of x0, the function becomes linear, ensuring that the transition point is part of the line. The parameter θ was fitted linearly. N and θ fits for the rfrUP PIF2 and tpbDOWN PIF3 datasets, respectively can be seen in Figure 5.11. The total number of parameters for the MBR parameter fits is given by: the number of parameters in the function, the 5 datasets and the PIFGAR subsets, respectively, resulting in 4× 5× 2 = 40 parameters for N , and 2× 5× 2 = 20 parameters for θ. Taking all variables into account, the total number of parameters involved in the his- togram fits is 40 + 560 + 460 = 1060, corresponding to NSC, RPR and MBR, respectively. Additionally, the number of parameters from the parameter fits rises to 110+60 = 170 from RPR and MBR. The best fit values from the parameter fits for all variables in the RPR and MBR fits were written to a ROOT file, along with the values from the NSC fits and the PIFGAR histogram shown in Figure 5.5. This ROOT file serves as input to the PLR analysis framework, along with analogous files provided by collaborators working on other primary background sources, including alpha decays originating in the detector neck, the AV surface, and copper dust present in the LAr. This is an ongoing project within the DEAP collaboration, aimed at conducting a WIMP search using the PLR method on a dataset with a livetime of 388 days. 5.4 Neutron Discrimination Using Machine Learning As discussed in 4.4, radiogenic neutrons cannot be filtered by using the PSD method detailed in 4.3 because they also interact via nuclear recoil with the 40Ar nuclei. This means that they will lie in the same Fprompt band that WIMPs would, at overlapping energies. There is, however, one key difference between the interactions of WIMPs and neutrons that might help in mitigating this background source: neutrons have a higher probability of multiple interactions with the argon nuclei. This argument can be made in terms of the mean distance traveled by a particle through matter without having an interaction, the mean free path (λ), which is given by [113]: λ = 1 nσ (5.18) where n is the density of target nuclei per unit volume and σ is the interaction cross section. The numerical values for the atomic density of liquid argon at 87 K and 1 atm, taken from [114], and the upper limit for the WIMP-nucleon scattering cross section for a 100 GeV WIMP imposed by DEAP [81], yield the following approximate upper limit for the mean free path of WIMPs in liquid argon: λWIMP > 1 2.11× 1022 atoms cm3 · 3.9× 10−45cm2 = 1.2× 1022cm = 1.3× 104ly. (5.19) Doing a similar exercise for the neutrons, with the value for the scattering cross section stated in [115], results in this mean free path: 48 CHAPTER 5. NEUTRON BACKGROUND IN DEAP-3600 λneutron = 1 2.11× 1022 atoms cm3 · 4.21× 10−25cm2 = 1.13m. (5.20) The difference is staggering, and given that the radius of the acrylic vessel in the DEAP 3600 detector is approximately 85 cm [87], it becomes evident that the probability of a WIMP scattering more than once with an 40Ar nucleus can be reliably neglected when compared to the same interaction for a neutron. This means that, in principle, recognizing interactions with multiple nuclear recoils can be used as a reliable method for identifying neutrons in the signal seen at the detector. Developing an additional method to mitigate this background could potentially lead to more stringent limits or a higher probability of detecting new phenomena, making this research topic of significant importance. Currently, no established method exists for achieving this. This study aims to serve as an initial step for this purpose, exploring one of the possible directions towards this goal. 5.5 Formulation as a Machine Learning Problem In physical terms, the problem can be stated as finding a method for identifying multiple nuclear recoils in the LAr volume of the detector, which, by the argument presented in Section 5.4, can be reliably discarded as candidates for WIMP-nucleus scattering interactions, being identified instead as caused by background sources such as radiogenic neutrons. Given that there is no straightforward way to determine if the light captured by the PMTs comes from multiple nuclear recoils caused by a single neutron, this problem doesn’t have a practical analytical solution. However, it can be addressed using machine learning methods, which do not require an explicit analytical relationship between the independent and target variables to be provided. Instead, these methods rely on extracting patterns and information directly from the provided dataset, making them particularly suited to problems where no practical analytical formulation exists. Because real detector data cannot currently provide information on the number of nuclear recoils caused by neutrons, the analysis relies on simulated data as its primary input source. Simulations have been extensively used by the DEAP collaboration to estimate background contributions in previous analyses [81, 95, 116]. These simulations, generated using the RAT software [99], accurately model the detector’s geometry and the relevant physical processes, making them a reliable source for this study. An additional advantage of using simulations is that information on the physical trajectory followed by the neutron is available, which makes it possible to count the number of scat- terings between a given neutron and argon nuclei. This provides a ground truth for tagging multiple nuclear recoil interactions produced by radiogenic neutrons, as well as additional useful information such as the proportion of these neutrons that scatter with more than one argon nucleus. The fact that ground truth is accessible makes this problem well-suited for analysis with supervised machine learning algorithms. Since the task can be reduced to determining whether an interaction is either a single or a multiple nuclear recoil, a binary classification supervised learning algorithm is the most appropriate choice. These algorithms are trained 5.6. DATA COLLECTING AND PROCESSING 49 on data labeled as either true or false in a binary target class, and can then be used to classify unlabeled data into one of these two categories. 5.6 Data Collecting and Processing The primary sources for this analysis were simulated radiogenic neutron datasets. The dataset used for model training was a dataset of 100 million neutrons generated isotropically from the PMT array; an additional dataset of 10 thousand neutrons generated from the AV (see 4.1) was used for model evaluation. Both of these datasets consist of neutrons emitted from radioactive decays in the lower part of the U238 decay chain, with the spectra for their corresponding materials (borosilicate glass and acrylic, respectively). These Monte Carlo datasets contain information about the trajectory, position, and physical processes undergone by the neutrons. An additional dataset was used for model validation, consisting of real data collected when an Americium-Beryllium (AmBe) neutron source was deployed. A ROOT script was written that accesses the trajectories of the neutrons, identifies the ones that have penetrated to the LAr volume, and counts the number of scattering interac- tions with 40Ar nuclei. This allowed for the creation of a binary feature that assigned the value of 1 to neutrons with more than 1 nuclear recoil, and 0 to those with exactly 1. Neu- trons that did not produce any nuclear recoils were not used for this analysis, since ultimately they wouldn’t be detected by the PMTs. This binary feature was used as the ground truth that served as a flag for training various supervised machine learning algorithms. The other features in the dataset contain information such as: the time interval between 2 events, the reconstructed position of the events, Fprompt (see Section 4.3), the number of pulses registered in a specific time window, and the pulse charge divided by the average single photo-electron charge for each PMT for each event. From this part of the analysis, it was derived that approximately 80% of the neutrons that scatter with an argon nucleus do so multiple times (see Table 5.2). This finding fur- ther highlights the reach of accurately identifying multiple nuclear recoils, since this could potentially be related to 80% of the neutrons that are detected in the PMTs. Number of Neutrons Percentage of Total Multiple Nuclear Recoils 7754 80.43 Single Nuclear Recoil 1887 19.57 Table 5.2: Count of multiple and single nuclear recoils from the PMT neutron dataset. An imbalanced training dataset, such as the one used here, where the classes are approx- imately in an 80-20 distribution, can introduce significant biases in the model and make predictions unreliable [117, 118]. There are two main methods for addressing class imbal- ance: oversampling and undersampling [119]. Oversampling involves increasing the number of instances in the minority class, while undersampling reduces the number of samples in the majority class. Both methods were applied and compared in this study. Oversampling was performed using the method of Synthetic Minority Oversampling Tech- nique (SMOTE) from the imbalanced-learn Python library [120], which generates synthetic 50 CHAPTER 5. NEUTRON BACKGROUND IN DEAP-3600 samples for the minority class in order to balance the dataset. Undersampling was imple- mented by randomly selecting an equal number of multiple recoil samples to match the number of single recoil samples. The best results were achieved with the dataset balanced through undersampling; consequently, all the results presented in this chapter were generated using this dataset. 5.7 Model Training and Evaluation 5.7.1 Training The dataset outlined in Section 5.6 was used to train and evaluate several binary classi- fication algorithms, using an 80-20 train-validation split. This process consists of splitting the dataset into 2 subsets by random sampling; one containing 80% of the samples, used for training the different classification algorithms, and a second one with the remaining 20%, used for validating and comparing the models. The splits are performed several times, in order to produce an average of the validation metrics for all the splits. This is called cross- validation [121], and it is done to ensure that the results obtained are not the product of a specific train-validation split, but rather an actual feature of the data. The multiple recoil flag was removed from the training dataset before the different models were trained, and then these were evaluated by using them to make predictions on the validation dataset, and comparing these predictions with the flags containing the ground truth for this dataset. The models trained and evaluated with this method include: decision tree, support vector classifier, random forest, boosted decision tree, and fully connected neural network. 5.7.2 Evaluation Metrics The primary evaluation metrics for a classification algorithm are the counts of correctly and incorrectly classified samples [122]. For the case of binary classification, the classes are referred to as positive and negative. The count of samples predicted correctly as positive is called true positive (TP), while the count of samples incorrectly classified as positive is referred to as false positive (FP). The same logic applies to the negative labeled samples, in which the count of correctly classified negative samples is called true negative (TN), and the samples incorrectly classified as negative are called false negative (FN). These metrics can be visualized in a confusion matrix (see Figure 5.12). A variety of additional model evaluation metrics are derived from the confusion matrix [123]. The most relevant for this analysis are listed below. • Accuracy: Accuracy = TP + TN TP + TN + FP + FN (5.21) Accuracy is the fraction of the correctly labeled samples out of the totality of the dataset. The best possible accuracy value is 1, the worst possible is 0. 5.7. MODEL TRAINING AND EVALUATION 51 Figure 5.12: Binary classification metrics true positive (TP), false positive (FP), true neg- ative (TN) and false negative (FN) visualized in a confusion matrix. The true values are arranged in rows, while the predicted labels are represented in the columns of the matrix. • Precision: Precision = TP TP + FP (5.22) The precision metric quantifies how many of the positive predictions made by the model were correct. The best precision score attainable by a model is 1, while the worst possible is 0. • True positive rate (TPR), also known as recall or acceptance: TPR = TP TP + FN (5.23) TPR represents the fraction of samples correctly labeled as positive out of all the positive instances in the dataset. The best possible TPR is 1, the worst is 0. • False positive rate (FPR): FPR = FP FP + TN (5.24) FPR is the fraction of incorrectly classified negatives out of all the negatives in the dataset. The best possible FPR is 0, the worst is 1. • Specificity, also known as background rejection rate (BRR): BRR = TN FP + TN = 1− FPR (5.25) 52 CHAPTER 5. NEUTRON BACKGROUND IN DEAP-3600 BRR measures the performance of the model in minimizing the incorrect classification of instances as positive. The extreme values of this metric are 1 (best possible) and 0 (worst possible). • F1 score (F1): F1 = 2× Precision× Recall Precision + Recall = 2TP 2TP + FP + FN (5.26) F1 is the harmonic mean between precision and recall. It measures how precise the model is in classifying instances as positive, while still being able to correctly classify as many positive instances as possible. The best F1 score is 1, the worst is 0. • Receiver operating characteristic (ROC) curve, and area under the curve (AUC): Classification models predict a probability of a sample belonging to the positive class, usually the threshold for labeling a sample as positive is a probability greater than 1/2. The ROC curve evaluates the model’s performance across all possible thresholds by plotting the TPR against the FPR. A quickly increasing ROC curve that approaches the top left corner of the plot indicates a strongly performing classification model that maximizes TPR while minimizing FPR. The model’s performance is compared with a baseline that represents a classifier that predicts either class with equal probability. The area under the ROC curve (AUC) provides additional information on the effectiveness of the classifier: an AUC of 1 indicates perfect classification, an AUC of 1/2 is that of the random classifier, and values below that indicate a model that systematically misclassifies. The ROC curve and the AUC are illustrated in figure 5.13. 5.7.3 Validation The best performance with the validation dataset was achieved by the boosted decision tree (BDT) model from the Python library scikit-learn [124]. Decision trees split data itera- tively based on specific variables to create a tree-like structure of decisions, classifying data into distinct groups. Starting from a root node, data is divided into branches using rules that optimize separation between categories, ultimately ending in leaves that represent the final predictions. BDTs enhance the basic decision tree by using an ensemble of trees to improve accuracy [125]. Boosting combines multiple individual trees, called weak learners, sequentially, one after the other, into a more accurate model (see Figure 5.14). By assigning more weight to misclassified data in each iteration, boosting algorithms minimize classifica- tion errors. The model was trained with the following hyperparameters: n estimators (the number of trees) set to 100, and learning rate (which controls each tree’s contribution to the final prediction) set to 0.12. BDTs have been used previously by the DEAP collaboration [126], as well as by other highly renowned experiments such as the Large Hadron Collider (LHC) [127–129], the Jiangmen Underground Neutrino Observatory (JUNO) [130] and the Mini Booster Neutrino Experiment (MiniBooNE) at Fermilab [131]. 5.7. MODEL TRAINING AND EVALUATION 53 Figure 5.13: The ROC curve [123] is constructed by calculating the TPR and FPR for all classification thresholds. The green curve represents a good performance, the red one a classification model that consistently misclassifies. The black diagonal line in the middle serves as a baseline and it represents a model that randomly predicts classes with equal probability. The AUC of a robust model is close to 1, while that of the random classifier represented by the diagonal line is 1/2. Figure 5.14: An individual decision tree (left), blue rectangles represent nodes that split into branches and end in green circles representing leaves. A boosted decision tree (right), a sequence of individual decision trees in which each tree is trained on the misclassified instances of the previous tree, creating a more accurate model. 54 CHAPTER 5. NEUTRON BACKGROUND IN DEAP-3600 The best performing BDT developed in this work achieved the results shown in the con- fusion matrix in Figure 5.15. The following evaluation value counts were obtained: • TP = 265. • FN = 97. • FP = 77. • TN = 316. Figure 5.15: Confusion matrix achieved with the BDT in the validation stage of its evalua- tion. From the confusion matrix, it can be seen that the false negative and positive counts are significantly lower than the true counts; this means that the BDT favors the correct classification of multiple and single recoil events. The scores obtained for the metrics derived from the confusion matrix are listed in Table 5.3. The ROC curve for this model is shown in Figure 5.16, with an achieved AUC value of 0.85, significantly higher than the baseline value of 1/2 attainable by random guessing. 5.7. MODEL TRAINING AND EVALUATION 55 Metric Score Accuracy 0.77 Precision 0.78 True Positive Rate 0.73 False Positive Rate 0.20 Specificity 0.80 F1 0.75 Table 5.3: Metric scores derived from the confusion matrix for the BDT in the validation phase. Figure 5.16: ROC curve for the BDT in the validation phase. The diagonal line ascending from left to right represents the performance of a random classifier that predicts each class with equal probability, with an AUC of 1/2. The AUC value achieved by the BDT is 0.85. 5.7.4 Additional Tests The model was subject to testing with two additional datasets: the first one consists of an MC dataset of radiogenic neutrons generated from the AV of the detector, the second one was gathered from real data taken during runs when a neutron source was deployed. The results presented in this section are color coded by dataset: the figures plotted in purple correspond to the AV neutrons MC dataset, and the image in orange to the neutron source runs. For the first test, the trained BDT did predictions on the AV simulated neutrons dataset. This dataset also has information on trajectories and interactions, making it possible to 56 CHAPTER 5. NEUTRON BACKGROUND IN DEAP-3600 compare with the true values for multiple and single nuclear recoils. The results obtained by the BDT for this dataset are shown in the confusion matrix in image 5.17. The evaluation counts obtained in the testing stage are: • TP = 950. • FN = 383. • FP = 106. • TN = 291. Comparing this confusion matrix with the one achieved in the validation phase reveals a preference for false negatives rather than true negatives. It also shows a vast majority of multiple recoils correctly classified. Figure 5.17: Confusion matrix achieved with the BDT in the testing stage of its evaluation. The metrics derived from the confusion matrix are shown in Table 5.4. These show a minor decrease in accuracy, true positive rate, specificity, and F1 score, while maintaining a favorable evaluation overall. There is a significant increase in precision, which can be attributed to the large number of true positives compared to the false positives shown in the confusion matrix. The ROC curve for the AV dataset, shown in Figure 5.18, shows a performance that is significantly better than the random guessing represented by the diagonal dashed line. 5.7. MODEL TRAINING AND EVALUATION 57 Metric Score Accuracy 0.72 Precision 0.90 True Positive Rate 0.71 False Positive Rate 0.27 Specificity 0.73 F1 0.80 Table 5.4: Metric scores derived from the confusion matrix for the BDT in the testing phase. The AUC has a value of 0.80, a minor decrease with respect to the result obtained in the validation phase of 0.85. This can be traced back to a slower growth in TPR for low values of FPR. Figure 5.18: ROC curve for the BDT in the testing phase with the AV neutrons MC dataset. The diagonal line ascending from left to right represents the performance of a random clas- sifier that predicts each class with equal probability, with an AUC of 1/2. The AUC value achieved by the BDT is 0.80. The last test done on the BDT had the purpose of simulating a realistic application scenario, in which predictions made on real data taken by the detector will be the subject of evaluation. To this end, the dataset used for this test consists of six runs of data taken when an Americium-Beryllium (AmBe) neutron source was deployed, resulting in a livetime of 4 days, 6 hours, and 14 minutes. The AmBe source has a neutron emission rate of 4.8 kHz [87], which allows for the assumption that all events recorded in these runs originate 58 CHAPTER 5. NEUTRON BACKGROUND IN DEAP-3600 from neutrons emitted by the calibration source. As previously mentioned, it is currently not possible to distinguish between multiple and single nuclear recoils in real data. This makes truth labels unattainable, rendering the use of a confusion matrix for evaluation impossible. To overcome this obstacle, an estimation of the expected number of multiple recoils can be obtained by assuming that approximately 80% of neutrons that produce a recoil do so multiple times. By taking into account the assumption that all events registered in this runs are neutrons emitted by the AmBe source, a number of multiple recoils can be estimated as 80% of the total events. In order to better emulate a realistic scenario, the events from the AmBe runs were selected by reconstructing inside the PLR ROI (see Section 5.1); no additional cuts were applied to the data. This would be a case in which being able to identify multiple recoils would be a helpful way of mitigating a background source in the ROI. The predictions for this dataset are shown in Figure 5.19. The points correspond to the location in the Fprompt - qPE parameter space of the events classified as multiple recoils by the BDT. An expected number of multiple recoils is shown, making the assumption that this will correspond to approximately 80% of all events registered. The prediction of 104+14 −12 is fairly lower than the expected 263 multiple recoils; however, being able to filter a significant fraction of the events is still helpful in the mitigation of radiogenic neutrons as a background source. 5.8 Findings and Conclusions This study was conducted as an initial step toward developing a new mechanism for mitigating radiogenic neutron backgrounds through multiple recoil discrimination. This is a complex problem for which a reliable solution has not yet been established. As discussed in Section 5.4, taking advantage of the high probability of neutrons causing multiple nuclear recoils could considerably improve background rejection in DM searches, thereby increasing the sensitivity of detectors like DEAP-3600 to potential WIMP signals. While Boosted Decision Trees (BDTs) implemented at the production level in other appli- cations have achieved higher evaluation results than those obtained in this study [126, 127], this work provides a promising foundation towards the possibility of reaching similar levels of reliability in multiple nuclear recoil identification. The classification results obtained were significantly better than random guessing, and an evaluation in a realistic scenario of appli- cation demonstrated the successful identification of a significant fraction of events expected to be multiple recoils (see Figure 5.19). Future efforts in this direction could benefit significantly from incorporating algorithms that leverage spatial information from interaction positions recorded by the detector, rather than relying solely on a variable-based approach. One promising avenue is the use of Graph Neural Networks (GNNs), which are a specialized type of artificial neural network designed to process graph-structured data, where nodes represent entities and edges define their re- lationships [132]. GNNs have been widely used for particle tracking and identification in particle physics experiments such as IceCube [133–135], and the LHC [136, 137]. Their ability to model complex spatial dependencies suggests they could enhance multiple recoil discrimination, potentially improving classification performance in future studies. Continu- 5.8. FINDINGS AND CONCLUSIONS 59 Figure 5.19: The orange points represent the central value of predicted multiple recoils for the AmBe dataset in the PLR region of interest. The expected number of multiple recoils is an approximation based on the assumption that the total number of events corresponds solely of neutrons from the AmBe calibration source. The uncertainties were calculated by determining a standard deviation of the probabilities predicted by the BDT, and then adding and subtracting that deviation from the probabilities, and applying the threshold of 50% for a positive prediction of a multiple recoil. ing this research may yield valuable results for experiments conducting WIMP searches that must address radiogenic neutrons as a significant background source. 60 CHAPTER 5. NEUTRON BACKGROUND IN DEAP-3600 6 Gamma Background for the Search of 5.5 MeV Solar Axions In addition to WIMPs, DEAP-3600 is sensitive to other new physics models. Among these are axion-like particles (ALPs), pseudoscalar bosons that arise from the spontaneous breaking of global U(1) symmetries. While for the QCD axion described in Section 3.2 the product of its photon coupling and mass is essentially fixed through QCD dynamics, ALPs constitute a broader class in which the coupling and mass can vary independently as free parameters [138]. ALPs would interact through electron recoils within the LAr volume in the detector. This chapter describes the potential mechanism for solar ALP production, outlines the interactions that DEAP-3600 could detect, and presents a contribution to refining the gamma-ray background model relevant for the ALP search. 6.1 Production Mechanism In addition to the QCD axions described in Section 3.2, there are other mechanisms through which ALPs would be produced. One such mechanism is the proton-proton fusion chain that occurs in the Sun [139]: p+ p → d+ e+ + νe d+ p →3 He + γ (5.5MeV), (6.1) where the 3He initially remains in an excited state and can emit a 5.5 MeV photon. If solar ALPs exist, the same excited 3He nucleus can instead emit an ALP of the same energy [140]. The reaction is written as follows, where the ALP is denoted by a: d+ p →3 He + a (5.5MeV). (6.2) Unlike the axions described in Section 3.2, ALPs in the MeV mass range would be too short-lived to be viable DM candidates, but can still have a connection to a rich dark sector [141]. Additionally, MeV ALPs might contribute to the anomalous magnetic moment of the electron, and have not been ruled out by the (g − 2)e results [142]. 61 62CHAPTER 6. GAMMABACKGROUND FOR THE SEARCHOF 5.5 MEV SOLAR AXIONS 6.2 Interactions with the Detector The DEAP-3600 detector is sensitive to ALPs through the process of electron recoil with the electron cloud of 40Ar atoms. Electron recoils can be generated by ALPs through four interactions: Compton conversion, decay to 2 γs, inverse Primakoff effect, and axioelectric effect. A description of each mechanism will be provided next. 6.2.1 Compton Conversion Compton Conversion, see Figure 6.1 upper left corner, occurs when an axion and a free electron interact, producing an electron and a photon: a+ e− → γ + e−. (6.3) The signal produced by this interaction would be an electron and a photon with a combined energy of 5.5 MeV. 6.2.2 Decay to 2γ The decay of an axion into two photons, see Figure 6.1 upper right, is described by the equation: a → γγ. (6.4) The signal observed in this decay would be two photons, each with momentum equal to half of the axion’s mass, in the rest frame of the axion. 6.2.3 Inverse Primakoff Effect In the Inverse Primakoff Effect, see Figure 6.1 lower left, an axion interacts with an electron bound to an atom, producing a photon. Unlike in Compton Conversion, the electron remains bound to the atom: a+ Z → γ + Z. (6.5) In this interaction, a photon with an energy of 5.5 MeV would be observed. 6.2.4 Axioelectric Effect The axioelectric effect, see Figure 6.1 lower right, is analogous to the photoelectric effect. In this interaction, the axion deposits its energy on a bound electron causing it to be ejected: a+ e− + Z → e− + Z. (6.6) In the axioelectric effect, an electron with a kinetic energy of 5.5 MeV would be the observed signal. 6.3. SIGNAL AND ITS DEPENDENCE ON AXION MASS 63 Figure 6.1: Feynman diagrams for the four interactions to which DEAP-3600 could be sensitive in the search for 5.5 MeV solar axions. Time is on the horizontal axis, a represents an axion, e an electron or positron, and Z an atom. The wavy lines represent a photon, and the solid lines without a label can represent a quark, electron, or positron. The label eb is used to denote an electron bound to an atom [143]. 6.3 Signal and its dependence on axion mass The signal that would be observed in the detector, in case of axion detection, was modeled using Monte Carlo simulations of each of the 4 interactions to which DEAP-3600 is sensitive. For these simulations, the GEANT4.9.6 software was used within the RAT framework [143]. For each of the four interactions, the energy deposited in the liquid argon was plotted, as shown in Figure 6.2. The Compton Conversion and 2γ Decay interactions are dependent on the axion mass. The spectrum produced by both interactions does not vary significantly for different axion masses, except when the axion mass approaches 5.5 MeV and enters the non-relativistic regime, as shown in Figure 6.3. 6.4 Region of Interest for Axion Search The region of interest for the axion search in DEAP-3600 is defined by the energy range between 5.26 MeV and 5.71 MeV. These correspond to 36000 qPE and 39000 qPE, respec- 64CHAPTER 6. GAMMABACKGROUND FOR THE SEARCHOF 5.5 MEV SOLAR AXIONS Energy (MeV) 0 1 2 3 4 5 6E ve nt s pe r 0. 03 M eV B in / T ot al E ve nt s 5−10 4−10 3−10 2−10 1−10 1 Compton Conversion γAxion Decay to 2 Inverse Primakov Axio-Electric Figure 6.2: Total energy deposited in the detector by each of the four interactions that DEAP-3600 is sensitive to. The axion mass used for Compton Conversion and 2γ Decay is 49 keV [143]. Energy (MeV) 0 1 2 3 4 5 6 P ro ba bi lit y pe r 0. 03 M eV 5−10 4−10 3−10 2−10 1−10 1 = 10 keVAm = 23 keVAm = 36 keVAm = 49 keVAm = 62 keVAm = 75 keVAm = 100 keVAm = 1000 keVAm = 3000 keVAm = 5000 keVAm = 5480 keVAm Energy (MeV) 0 1 2 3 4 5 6 P ro ba bi lit y pe r 0. 03 M eV 3−10 2−10 1−10 = 10 keVAm = 23 keVAm = 36 keVAm = 49 keVAm = 62 keVAm = 75 keVAm = 100 keVAm = 1000 keVAm = 3000 keVAm = 5000 keVAm = 5480 keVAm Figure 6.3: Total energy deposited in the detector for Compton Conversion (left) and 2γ Decay (right). Both produce two primary particles, an electron and a photon for Compton Conversion, and two photons for Decay. The spectrum produced by axions of different masses is shown. For both interactions, the spectrum remains unchanged until the axion mass approaches 5.5 MeV, at which point it becomes non-relativistic [143]. tively, where the qPE variable corresponds to the charge of the recorded pulse divided by the average charge of a photoelectron for each photo-multiplier tube [144]. 6.5 Background Sources Other physical processes that produce electronic recoils in liquid argon, depositing 5.5 MeV of energy, are background sources for the axion search in DEAP-3600. The two main processes that constitute background sources are: γ-rays emitted from neutron capture and 6.5. BACKGROUND SOURCES 65 γ-rays produced by the radioactive decay of 208Tl. 6.5.1 γ-rays from Neutron Capture The main background source for this search is γ-rays emitted from neutron capture. This process occurs when a neutron emitted by a component of the detector, through either (α, n) reactions or spontaneous fission, is later captured by a nearby nucleus, which emits a gamma-ray that deposits its energy in the liquid argon, causing an electron recoil (see Figure 6.4). To mitigate this background source, Monte Carlo simulations of γ-rays emitted from neu- tron capture are carried out, and these are then subtracted from the measured events (see Section 6.6). Figure 6.4: Example of the mechanism for γ-ray production via neutron capture. In this case, a radiogenic neutron emitted by an (α, n) reaction from the glass of the PMTs is captured by a nucleus in the acrylic, which subsequently emits a γ-ray that deposits its energy in the liquid argon. This is the primary background source for the search of 5.5 MeV solar axions [143]. 6.5.2 γ-rays from 208Tl Decay 208Tl is present in some components of the detector, it can produce γ-rays through various decays [145]. To determine whether this is a significant background source, Monte Carlo simulations of this process were carried out for decays occurring in the acrylic container and the liquid argon, the two components from which the largest number of electronic recoils could be generated. It was found that not enough γ-rays with energies near 5.5 MeV are produced by this mechanism to be considered a significant background source in the axion search (see Figure 6.5). 66CHAPTER 6. GAMMABACKGROUND FOR THE SEARCHOF 5.5 MEV SOLAR AXIONS Energy Deposited in LAr (MeV) 0 1 2 3 4 5 ev en ts p er 5 5 ke V b in / M C d at as et 7−10 6−10 5−10 4−10 3−10 2−10 1−10 Tl in the AV208 Tl in the LAr208 Figure 6.5: Spectrum generated by Monte Carlo for γ-rays emitted by the decay of 208Tl, for decays occurring in the acrylic container (black) and liquid argon (green) [143]. 6.6 Simulations for the Neutron Capture γ-ray Back- ground The main contribution presented in this work to the axion search project is focused on the background model caused by gamma rays from neutron capture, detailed in Section 6.5.1. Initially, the background model for the axion search showed a significant deficit compared to the background observed in the data for energy values close to the region of interest specified in Section 6.4 (see upper Figure 6.6). An extensive search was conducted to determine which background sources had not yet been accounted for and might be causing this excess in relation to the model. It was de- termined that the steel, from which the casing surrounding the acrylic container is made, contains significant amounts of 59Co and molybdenum isotopes 92, 94, 95, 96, 97, 98, and 100 in their natural abundance. Since cobalt and molybdenum are present in the steel, simulations were conducted of the gamma rays emitted with the neutron capture spectrum for these isotopes obtained from the International Atomic Energy Agency (IAEA) database [146]. These simulations were carried out using the RAT software [99], specifying that the gamma rays were emitted from the steel present in the spherical shell and the neck of the detector (see Figure 6.7). It was determined that these isotopes could capture neutrons that emit gamma rays with energies in the region of interest for the search, as shown in Figure 6.8. To complete the background model, similar simulations were carried out for 59Co and all molybdenum isotopes. The distribution of simulated events for these was based on their 6.6. SIMULATIONS FOR THE NEUTRON CAPTURE γ-RAY BACKGROUND 67 Figure 6.6: Background model before neutron capture gamma rays in 59Co and Mo isotopes were added (top), with a p-value of 0.5 ± 0.3%. After 59Co and Mo isotopes were added (bottom), the model shows a significant improvement around the region of interest, which covers values between 5.26 MeV (36,000 qPE) and 5.71 MeV (39,000 qPE). The p-value increases to 25± 1% [147]. neutron capture cross-section and their natural relative abundance (see Table 6.1). The total number of events generated, after 1,000 simulation iterations, was 97,165,000. 68CHAPTER 6. GAMMABACKGROUND FOR THE SEARCHOF 5.5 MEV SOLAR AXIONS Figure 6.7: Projection of the spatial distribution for the gamma rays simulated from the steel casing and neck of the detector. The spatial distribution from which the gammas were emitted for neutron capture in 95Mo is used as an example. Figure 6.8: Energy deposited in liquid argon by gammas emitted due to neutron capture in 59Co (right) and 95Mo (left), as an example of the isotopes included for Mo. In both cases, these isotopes deposit energy in the region of interest for the axion search, around 5.5 MeV. Subsequently, the simulated data was included in the background model for the axion search. By including gammas from neutron capture in Mo isotopes and 59Co, the background model improved significantly. The unexplained background near the region of interest was reduced, and the model’s p-value increased significantly from 0.5 ± 0.3% to 25 ± 1% (see Figure 6.6). This allowed the analysis process for the axion search to continue. Additionally, a study of event acceptance for the radial fiducial cut used in the axion search was carried out. This cut was set at 750 mm measured from the center of the acrylic container. To conduct the study, the acceptance was calculated at different radial cuts in terms of energy deposited in the argon, both for simulated data and data obtained from the AmBe calibration source. It was found that the value used for the radial cut provides satisfactory event acceptance while reducing background sources (see Figure 6.9). 6.6. SIMULATIONS FOR THE NEUTRON CAPTURE γ-RAY BACKGROUND 69 Molybdenum Isotope (A) Events per Iteration 92 920 94 1311 95 71443 96 7968 97 9389 98 4880 100 1254 Total 97165 Table 6.1: Number of events generated per iteration for each molybdenum isotope, consider- ing relative abundance and neutron capture cross-section. A total of 1,000 iterations of data generation were performed with this event distribution, resulting in a total of 97,165,000 events. 2 4 6 8 10 mc_edep (MeV) 0.0 0.2 0.4 0.6 0.8 1.0 Ac ce pt an ce Acceptance for mc_edep with cut on mbR at 750 mm Fits for mc_edep in 1.26 - 9.95 MeV y = 0.6149, p = 0.187 y = 0.0113x +0.5517, p = 0.2216 PMT neutrons MC nominal Figure 6.9: Acceptance for the radial fiducial cut at 750 mm from the center of the acrylic container, calculated in terms of energy deposited in liquid argon for neutrons simulated from the photo-multiplier tubes. A blind analysis was performed by a collaborator with the background model obtained after the inclusion of 59Co and Mo isotopes. No excess of events relative to the null hypothesis was found in the region of interest (see Figure 6.10). This allowed exclusion curves to be set for the axion-electron coupling constant gAe, the axion-photon coupling constant gAγ, and the product of the axion-electron coupling constant and the isovector component of the axion-nucleon coupling constant |gAe × g3AN |. The preliminary exclusion curves obtained by the collaboration are more restrictive than those from the BGO experiment [148] and are an order of magnitude above those obtained by Borexino [149]. Figure 6.11 shows the exclusion curve obtained for the axion-electron coupling constant gae, with the Axioelectric Effect and Compton Conversion processes. 70CHAPTER 6. GAMMABACKGROUND FOR THE SEARCHOF 5.5 MEV SOLAR AXIONS Energy (MeV) 5 6 7 8 9 10 C ou nt s / 6 0. 9 ke V / D at as et 1 10 210 Data Null Hypothesis Total Model Ar (n,*) Fe (n,*) Cr (n,*) Ni (n,*) Mn (n,*) Cl (n,*) Mo (n,*) )γC (n, βTl 208LAr )γCo (n,59 5.5 MeV Solar Axion DEAP 0063Preliminary Energy (MeV) 5 6 7 8 9 10 (d at a - M C ) / M C 1.0− 0.8− 0.6− 0.4− 0.2− 0.0 0.2 0.4 0.6 0.8 1.0 68.2% 95.4% 99.7% Figure 6.10: Preliminary background model for the axion search, following the blind search process. It shows no significant excess of events in the region of interest from 5.26 to 5.71 MeV [147]. An earlier version of these results is available in [143]. [MeV]Am 5−10 4−10 3−10 2−10 1−10 1 10 A e g 13−10 12−10 11−10 10−10 9−10 8−10 7−10 6−10 5−10 4−10 DEAP 0063 BGO AEDEAP AE DEAP CC Borexino CC Figure 6.11: Preliminary exclusion curves obtained for the axion-electron coupling constant gAe in the hadronic axion model, for the Axioelectric Effect (AE) and Compton Conversion (CC) processes [147]. An earlier version of these results is available in [143]. 6.6. SIMULATIONS FOR THE NEUTRON CAPTURE γ-RAY BACKGROUND 71 A draft of the Axion Search article with the preliminary results shown here was written and submitted for internal review by the committee overseeing this analysis. The review process continues, with contributions from various collaborators. 72CHAPTER 6. GAMMABACKGROUND FOR THE SEARCHOF 5.5 MEV SOLAR AXIONS 7 Exploration of Neutrinos in the Stan- dard Model and Beyond 7.1 General Aspects about Neutrinos In the early 1900s, it was believed that the process of β decay violated the conservation of energy and momentum. This conclusion was drawn from the observation that the spectrum of the resulting electron, inferred from Equation 7.1, was continuous. If energy were accounted for by only the resulting nucleus and electron, the electron spectrum would have been mono- energetic [150]. (A,Z) → (A,Z± 1) + e∓ +missing energy and momentum. (7.1) To solve this problem, in 1930 Wolfgang Pauli proposed the existence of a spin 1/2 particle with a neutral electrical charge and a small mass, that would carry the unaccounted for energy and momentum [151]. In 1933 Fermi developed an explanation for β decay using this particle, which he named neutrino [152]. The neutrino was conclusively discovered in 1956 using a scintillation detector at a reactor in the Savannah River nuclear plant [153]. Neutrino interactions are chiral in nature; only left-handed neutrinos and right-handed antineutrinos participate in weak interactions [138]. This chiral property is intrinsic to the Standard Model (SM) and applies to neutrinos in their leptonic flavor states: electron (νe), muon (νµ), and tau (ντ ). These flavors arise from the different processes that produce the neutrinos and antineutrinos, for example: n → p+ e− + ν̄e, (7.2) π+ → µ+ + νµ, (7.3) and τ− → µ− + ν̄µ + ντ . (7.4) Neutrinos can be produced in several sources such as the Sun, nuclear reactors, and particle accelerators. The different neutrino sources, as well as their energy and flux, are visualized in Figure 7.1. However, neutrinos do not propagate as pure flavor states. Instead, each flavor eigenstate is a quantum superposition of three mass eigenstates, labeled simply as 1, 2, and 3 [155]. This 73 74CHAPTER 7. EXPLORATION OF NEUTRINOS IN THE STANDARDMODEL AND BEYOND Figure 7.1: Flux vs neutrino energy for different sources for neutrinos [154]. became evident when attempts were made to measure solar neutrinos. Electron neutrinos are produced in various steps of the pp chain that occurs in the sun, which starts with the fusion of two protons and results in the production of two alpha particles [156]. The results of the first experiment to measure solar neutrinos were reported in 1968 [157]. The experiment used chlorine to detect νe via the interaction: νe + 37 Cl →37Ar + e−, (7.5) but was only able to account for approximately 1/3 of the expected neutrinos. This discrep- ancy became known as the solar neutrino problem. The resolution of this problem came about when it was considered that neutrinos oscillate between three different flavor states, two of which could not be detected by inverse beta decay in chlorine. The confirmation of solar neutrino oscillations was achieved at the Sudbury Neutrino Observatory (SNO) in 2001 [158], when a detection mechanism using heavy water was able to detect all three flavors. This mechanism consisted of the following processes: νe + d → p+ p+ e−, (7.6) ν + d → n+ p+ ν, (7.7) and ν + e → ν + e, (7.8) where ν represents a neutrino of any flavor, and d the deuterium present in heavy water. When combining their results with those reported previously by Super-Kamiokande [159], 7.1. GENERAL ASPECTS ABOUT NEUTRINOS 75 SNO was able to account for 100% of the expected solar neutrino flux, concluding that 35% consists of νe, while the remaining 65% is made up of νµ or ντ . The probability of neutrino oscillation can be obtained by studying a description of this phenomenon in terms of the superposition of quantum states [156]. Assuming, for simplicity, only two different flavors (νe and νµ), and propagation in a vacuum, the mass eigenstates can be written as: |ν1⟩ = cos θ |νµ⟩ − sin θ |νe⟩ , (7.9) and |ν2⟩ = sin θ |νµ⟩+ cos θ |νe⟩ , (7.10) where the choice of cosine and sine for the coefficients facilitates the normalization of the quantum states, and θ plays the role of a mixing angle. The time evolution of these states can be described using a plane wave solution (ℏ = c = 1): |ν1(t)⟩ = e−iE1t |ν1(t = 0)⟩ , (7.11) |ν2(t)⟩ = e−iE2t |ν2(t = 0)⟩ . (7.12) Assuming that the initial state is an electron neutrino, that is |ν1(t = 0)⟩ = − sin θ |νe⟩ and |ν2(t = 0)⟩ = cos θ |νe⟩, the time dependencies become: |ν1(t)⟩ = − sin θe−iE1t |νe⟩ , (7.13) |ν2(t)⟩ = cos θe−iE2t |νe⟩ . (7.14) Using Equations 7.9, 7.10, 7.13 and 7.14, the time dependence for the muon neutrino state is written as follows: |νµ(t)⟩ = − sin θ cos θe−iE1t |νe⟩+ cos θ sin θe−iE2t |νe⟩ . (7.15) The transition amplitude from electron to muon neutrino is given by: Aµe = ⟨νe|νµ(t)⟩ = sin θ cos θ ( −e−iE1t + e−iE2t ) , (7.16) and the transition probability is: Pνe→νµ = |Aµe|2 = sin2 θ cos2 θ ( 2− ei(E2−E1)t − ei(E1−E2)t ) = [ sin(2θ) sin ( (E2 − E1)t 2 )]2 1. (7.17) 1Identities used: cosx = − eix+e−ix 2 , sin2 θ cos2 θ = sin2(2θ) 4 and 1−cosx = 2 sin2 x 2 , where x = (E1−E2)t. 76CHAPTER 7. EXPLORATION OF NEUTRINOS IN THE STANDARDMODEL AND BEYOND By taking the ultrarelativistic limit E ≈ p ( 1 + m2 2p2 ) , the approximation E1 − E2 ≈ ∆m2 2E is made, where ∆m2 = m2 2 −m2 1. Substituting in 7.17 and recovering the factors of c and ℏ, the transition probability is written as follows: Pνe→νµ = [ sin(2θ) sin ( ∆m2c3 4ℏE L )]2 , (7.18) where E is the neutrino energy, and L = ct is the distance referred to as the baseline. The transition probability is maximized for Lmax = 2πℏE ∆m2c3 , and the distance at which the neutrino oscillates back to its original flavor state, known as the oscillation length, is LO = 4πℏE ∆m2c3 . Equation 7.18 reveals that for oscillations to occur, the masses m1 and m2 must be differ- ent; and at most one of them can be 0, but not both. The phenomenon of neutrino oscillation is unequivocal evidence that neutrinos have a mass, but only their squared difference can be known by studying oscillations. Two different compatible orderings can be inferred for the differences of the squared masses from oscillation data [138]: normal hierarchy (NH) and inverted hierarchy (IH), visualized in Figure 7.2. In the NH, the masses follow the ordering m1 < m2 << m3, while in the IH the ordering is reversed to m3 << m1 < m2. The neutrino mass ordering remains an open question, and experiments such as the Jiang- men Underground Neutrino Observatory (JUNO) [160] and the Deep Underground Neutrino Experiment (DUNE) [161] are expected to provide new insights in the near future. Figure 7.2: Possible hierarchies for the neutrino masses [162]. Note: ∆m2 atm is equivalent to ∆m2 32 and ∆m2 sol to ∆m2 21. When generalizing to the three flavor and mass fields, the relationship between them is described by the Pontecorvo–Maki –Nakagawa–Sakata (PMNS) matrix [163, 164]: 7.1. GENERAL ASPECTS ABOUT NEUTRINOS 77   νe νµ ντ   =   c12c13 s12c13 s13e −iδ −s12c23 − c12s23s13e iδ c12c23 − s12s23s13e iδ s23c13 s12s23 − c12c23s13e iδ −c12s23 − s12c23s13e iδ c23c13     eiα1/2ν1 eiα2/2ν2 ν3   , (7.19) where cij is defined as cosθij and sij as sinθij [162]. In the PMNS matrix, there are three mixing angles: θ12, θ13, and θ23, one Dirac phase δ, and, if there is a Majorana component to the neutrino masses, two additional phases α1 and α2 [46]. These phases relate to a currently open question about whether neutrinos are Dirac particles (with distinct neutrino and antineutrino states) or Majorana particles (where neutrinos are their own antiparticles). If neutrinos are Majorana in nature, this would lead to lepton number violating processes such as neutrinoless double-beta decay (0νββ) [165], illustrated in Figure 7.3. As of the writing of this work, neutrinoless double-beta decay has not been detected. Among the experiments searching for neutrinoless double-beta decay are: Large Enriched Germanium Experiment for Neutrinoless ββ Decay (LEGEND) [166], The Kamioka Liquid-scintillator Anti-Neutrino Detector (KamLAND-Zen) [167], Sudbury Neutrino Observatory + (SNO+) [168], Cryogenic Underground Observatory for Rare Events (CUORE) [169], Enriched Xenon Observatory (EXO) [170] and Germanium Detector Array (GERDA) [171]. Figure 7.3: Left: Feynman diagram for double beta decay with two electron antineutrinos produced (2νββ), the lepton number is conserved. Right: Feynman diagram for neutrinoless double beta decay (0νββ). Two neutrons decay into protons and electrons, without neutrinos in the final state. This process violates the conservation of lepton number, and would require neutrinos to be their own antiparticle, referred to as Majorana particles. The origin of neutrino masses is not explained within the SM, meaning that extensions are required to account for this property. One prominent extension is the seesaw mechanism [46], which introduces heavy, sterile neutrinos with right chirality that possibly only interact via the gravitational force. These right-handed neutrinos couple to the left-handed neutrinos via Yukawa interactions, leading to a mass matrix that, upon diagonalization, results in one very light mass eigenstate (corresponding to the observed light neutrino) and one very heavy 78CHAPTER 7. EXPLORATION OF NEUTRINOS IN THE STANDARDMODEL AND BEYOND mass eigenstate (the sterile neutrino). This mechanism naturally explains why neutrino masses are much smaller than those of other fermions and also introduces sterile neutrinos as a viable DM candidate (see Section 3.3). 7.2 Coherent Elastic Neutrino-Nucleus Scattering The coherent elastic neutrino-nucleus scattering (CEνNS) is a neutral current interaction of the form: ν + A → ν + A, (7.20) that is mediated by a Z0 boson (see Figure 7.4). It was first proposed in 1974 by D. Freedman [172], and was first measured in 2017 by the COHERENT collaboration [173]. Figure 7.4: Feynman diagram for CEνNS. The coherent elastic scattering interaction between a neutrino (ν) and an atomic nucleus A is mediated by a Z0 boson. The property of coherence in CEνNS refers to the fact that the nucleus interacts as a whole rather than as individual nucleons [174]. This occurs when the wavelength λ of the momentum transferred during the scattering is comparable to, or larger than, the nuclear radius R: λ ≳ R, which translates to a condition on q, the transferred momentum: q ≲ 1 R . As a result, all nucleons contribute constructively to the scattering process, leading to an enhanced cross-section compared to other neutrino processes (see Figure 7.5). The SM cross section for CEνNS is given by [175]: dσ dT SM ≈ G2 F 2π MNQ 2 w ( 2− MNT E2 ) 2, (7.21) where MN is the nuclear mass, GF the Fermi constant, E the incident neutrino energy, and T is the recoil energy. The weak charge is: Qw = ZgVp FZ(q 2) +NgVn FN(q 2), (7.22) 2Equation 7.21 is an approximation obtained by taking the limit T << E in the cross section given by [176]: dσ dT SM = G2 F 2π MNQ2 w ( 2− 2T E + ( T E )2 − MNT E2 ) . 7.3. CEνNS MEASUREMENT BY THE CONUS+ EXPERIMENT 79 Figure 7.5: Left: representation of CEνNS and subsequent detection mechanisms. Right: comparison of the total cross-section of CEνNS and other neutrino couplings: neutrino- electron scattering, charged-current interaction (CC), inverse beta decay (IBD) and neutrino- induced neutron (NIN) [173]. with Z and N representing the numbers of protons and neutrons respectively, FZ and FN are the proton and neutron form factors as function of the transferred momentum q = √ 2mT , where m is the nuclear mass, and gVp = 1 2 − 2 sin2 θW , and gVn = −1/2, (7.23) are the proton and neutron couplings respectively; where θW is the weak mixing angle, also known as the Weinberg angle. CEνNS measurements can be used to probe into the weak mixing angle, the magnetic moment of the neutrino µν , and parameters of non-standard interactions (NSI) in extensions of the SM, as will be seen in Section 7.4. Figure 7.6 illustrates potential CEνNS measurement outcomes for different values of NSI parameters and µν . 7.3 CEνNS Measurement by the CONUS+ Experiment 7.3.1 Description of the CONUS+ Detector The CONUS+ experiment aims at the detection of CEνNS between electron antineutrinos and germanium nuclei in high-purity germanium (HPGe) crystals [179]. It is located in a nuclear power plant in Leibstadt, Switzerland; the HPGe crystals are placed at a distance of 20.7 m from the 3.6 GW reactor core (see Figure 7.7). The reactor building provides an average of 7.4 meter water equivalent (m.w.e.) overburden. The production of electron antineutrinos at a nuclear reactor occurs through chains of β decays of the form n → p + e− + ν̄e, originating from the fission of 235U (see Figure 7.8), 238U, 239Pu and 241Pu. Nuclear reactors offer considerable advantages as a neutrino source for experiments like CONUS+ compared to other sources [180]: 80CHAPTER 7. EXPLORATION OF NEUTRINOS IN THE STANDARDMODEL AND BEYOND • The flux is considerably higher than that attainable at an accelerator. • The cost is significantly lower than operating an accelerator, as commercial reactors are already in operation for energy production. • The flux consists exclusively of electron antineutrinos, which considerably facilitates the determination of the flux uncertainty. • Reactors can be temporarily shut down, providing a valuable opportunity to study experimental backgrounds. CONUS+ reuses the four HPGe diodes from the CONUS experiment [181], which were refurbished and optimized for achieving a minimum ionization energy threshold of 160 eVee (see Figure 7.9). When energy is deposited in a Ge nucleus by a neutrino in a CEνNS interaction, the nucleus recoils and moves through the crystal lattice, losing energy primarily through ionization and heat production, also known as quenching effects. The ionization events generate electron-hole pairs, which drift in opposite directions under the influence of an electric field applied across the diode, producing a measurable electrical signal [182]. The fraction of the nuclear recoil energy (T ) that generates ionization after taking into account the quenching effects, known as the ionization energy (K), is modeled by: K = kg(ϵ) 1 + kg(ϵ) · T, (7.24) where ϵ = 11.5Z−7/3T and g(ϵ) = 3ϵ0.15 + 0.7ϵ0.6 + ϵ, with Z = 32 for Germanium, and k = 0.162± 0.004 is determined by experimental measurements [183]. The remaining recoil energy dissipates as heat and is therefore not detected. Figure 7.6: Left: nuclear recoil event rate of CEνNS in a germanium target for the SM (solid black), and different values of µν normalized by the Bohr magneton µB: blue, green and red for µν µB = 2.2× 10−12, 2.9× 10−11 and 3.0× 10−10, respectively. The dashed line represents a typical background level [177]. Right: a deficit of CEνNS events with respect to the SM is shown in pink, according to the shown non-standard interaction parameters [178]. 7.3. CEνNS MEASUREMENT BY THE CONUS+ EXPERIMENT 81 Figure 7.7: Setup of the CONUS+ experiment inside the reactor building at the Leibstadt nuclear power plant [179]. Figure 7.8: Production of electron antineutrinos trough chains of β decays from the fission of 235U in nuclear reactors. In average, 6 ν̄e are produced per fission in a nuclear reactor. The letter ν in this Figure represents ν̄e [180]. As can be seen in Equations 7.22 and 7.23, the CEνNS cross section scales as N2, where N is the number of neutrons in the nucleus. This makes large nuclei attractive targets for these experiments; however, the uncertainty in quenching parameters tends to increase with heavier nuclei. As a result, the medium-sized Germanium nucleus serves as a suitable 82CHAPTER 7. EXPLORATION OF NEUTRINOS IN THE STANDARDMODEL AND BEYOND balance, providing a detectable CEνNS signal while maintaining manageable uncertainties in quenching factors [184]. The main background sources for this experiment include neutrons and electromagnetic cascades originating from cosmic muons, as well as γ-rays produced in the decay of neutron- capture-induced isotopes. To mitigate these background sources, an extensive characteriza- tion campaign was conducted, incorporating Monte Carlo simulations and in situ measure- ments during both reactor on and off periods. The background characterization efforts at the CONUS+ experimental location included a detailed study of neutron fluence using Bonner Sphere spectrometry, γ-ray spectroscopy with HPGe detectors, and muon flux measurements with liquid scintillators [185]. These efforts led to the development of a background model that enables the detection of CEνNS at sub-keV energy thresholds. Figure 7.9: The four HPGe diodes used in the CONUS+ experiment, their height and diameter is 62 mm [179, 182]. 7.3.2 Results of the CEνNS Measurement by CONUS+ The CONUS+ collaboration reported the first evidence of CEνNS from reactor antineu- trinos measuring 395 ± 106 events, with a statistical significance of 3.7σ [181]. This result is consistent with the predicted number of 347 ± 59 events from calculations, as shown in Figure 7.10. The data was gathered between November 2023 and July 2024, including reactor on and off periods. This resulted in a combined exposure for the reactor on period of 327 kg·days, which was obtained using 3 out of the 4 HPGe detectors (Conus-4 was not included in the analysis due to significant instabilities in the event rate). The flux was determined to be 7.3. CEνNS MEASUREMENT BY THE CONUS+ EXPERIMENT 83 Figure 7.10: The top image shows the predicted CEνNS signal as a black line with the uncertainty in blue, the black squares are the data points after background subtraction, and the red dotted lines indicate the energy threshold of each of the detectors used. The bottom left figure shows the agreement between data and background model for an extended ionization energy range. In the bottom right is a histogram of the data points with a gaussian fit in red [181]. 1.5× 1013 ν̄es −1cm−2. The respective energy threshold, mass, livetime, and the detected and predicted signals for each of the detectors are shown in Table 7.1. An ionization energy window of 160 - 800 eVee was defined as the energy region of interest. In this region, a good agreement between the background model and data was achieved. This model includes contributions from cosmogenic neutrons and muons, as well as subdominant inputs from radiogenic activity from isotopes present in the detector materials. The back- ground model was validated using data from the reactor off period; which, together with a rigorous characterization of the detectors’ trigger efficiency (see Figure 7.11), enabled a satisfactory extraction of the neutrino signal using a PLR method. 84CHAPTER 7. EXPLORATION OF NEUTRINOS IN THE STANDARDMODEL AND BEYOND Detector Energy Threshold (keVee) Mass (kg) Livetime (d) Conus-2 180 0.95 ± 0.01 117 Conus-3 160 0.94 ± 0.01 110 Conus-5 170 0.94 ± 0.01 119 Total - 2.83 ± 0.02 - Table 7.1: Data for the detectors used in the CONUS+ CEνNS measurement [181]. The trigger efficiency for each detector is shown in Fig 7.11. 7.4 Study of Scenarios in the SM and Beyond The CEνNS results published by CONUS+ [181] provide a framework for studying pa- rameters of the SM and its extensions. In this work, a χ2 optimization approach is employed to evaluate and constrain the weak mixing angle θW , the neutrino magnetic moment µν , and the non-standard interaction (NSI) parameters ϵuVee and ϵdVee . The χ2 function employed is defined as: χ2 = ( Nmeas − (1 + α)Npred(X) σstat )2 + ( α σα )2 , (7.25) where Nmeas is the number of events measured by CONUS+, and the statistical uncertainty is given by σstat = √ Nmeas. Npred(X) is the number of events calculated in this study, being evaluated as a function of the parameter X (the weak mixing angle, the neutrino magnetic moment, or the NSI parameters), which is to be fitted in the optimization. Lastly, α is a nuisance parameter over which the χ2 function is marginalized. The uncertainty in α accounts for the dominant sources of systematic uncertainties reported by CONUS+: • Energy threshold: 14.1%, • Quenching parameter: 7.3%, • Reactor neutrino flux: 4.6%, • CEνNS cross section: 3.2%, • Active germanium mass: 1.1%, • Trigger efficiency: 0.7%, • Likelihood fit, fit method, background model, and non-linearity: 21.8%. Consequently, the uncertainty in the nuisance parameter α is: σα = √ 0.1412 + 0.0732 + 0.0462 + 0.0322 + 0.0112 + 0.0072 + 0.2182 = 0.276. (7.26) In the following sections, the details and results of this study, which has been published in the journal Physical Review D [186], will be presented. 7.4. STUDY OF SCENARIOS IN THE SM AND BEYOND 85 7.4.1 Calculation of the Number of events The predicted number of events, Npred, was calculated using the following equation: Npred = tNt ∫ Tmax Tmin ϵ(T )dT ∫ Emax Emin(T ) dσ dT (T,E)Φ(E)dE, (7.27) where: • E is the energy of the incident neutrino, • T is the recoil energy of the nucleus 3, • t is the data taking livetime, • Nt is the target number given by Nt = mtNA A ; where mt is the target mass, NA is Avogadro’s number and A is the target mass number, • ϵ(T ) is the detector efficiency as a function of the recoil energy, • dσ(T,E) dT is the differential CEνNS cross section as specified in Equation 7.21, and • Φ(E) is the incident neutrino flux. The Huber-Mueller spectrum for reactor antineu- trinos [187, 188] was used in this study. The limit of integration Emin(T ) = T+ √ T (T+2m) 2 is the minimum neutrino energy required to produce a recoil energy T , where m is the mass of the recoiling nucleus. Emax is the maximum energy with which neutrinos are produced at the source. Tmin is the energy threshold, a property specific for each detector, and Tmax is the upper limit for the energy region of interest. The steps taken to determine the predicted number of events for this study are described in this section. To properly account for the differences in energy threshold, mass, livetime, and efficiency, Equation 7.27 was applied separately to each HPGe diode. The efficiencies as a function of ionization energy used for each one are shown in Figure 7.11; and the masses, livetimes, and energy thresholds are presented in Table 7.1. Additionally, the calculations for each detector included iterations over all five germanium isotopes (with mass numbers 70, 72, 73, 74, and 76), to obtain results with a better accuracy. The relative abundances used are presented in Table 7.2. The number of neutrons in each isotope has an effect on the target number, the recoil energy, and the cross section trough the form factor and the weak charge. The Helm form factor was used for this work [191], it is given by: F (q2) = 3j1(qR0) qR0 e −(qs)2 2 , (7.28) 3For simplicity, Equation 7.27 is shown solely in terms of the nuclear recoil energy T . However, as mentioned in Section 7.3.1, only a fraction of T is detected, the ionization energy K; the remainder dissipates as heat. The efficiency and the integral with respect to T are actually applied with respect to K. The relationship between T and K is given by Equation 7.24. 86CHAPTER 7. EXPLORATION OF NEUTRINOS IN THE STANDARDMODEL AND BEYOND Figure 7.11: The trigger efficiency of each diode, as a function of ionization energy [179]. The Conus-4 diode was not included in the analysis due to an unstable event rate [189]. The efficiencies were interpolated as 0 for values less than the minimum ionization energy shown in the image, this has no effect in the calculation of the number of events due to the minimum energy shown being lower than the energy thresholds of the detectors. Germanium Isotope (A) Relative Abundance (%) 70 20.57 72 27.45 73 7.75 74 36.50 75 7.73 Table 7.2: Natural abundances of the germanium isotopes, used in the calculation of the target number for each isotope in each detector [190]. 7.4. STUDY OF SCENARIOS IN THE SM AND BEYOND 87 where j1 is the first order Bessel spherical function, s = 0.9 fm is the surface thickness parameter, and R0 = √ c2 + 7π2 3 a2 − 5s2 is the effective radius of the nucleus, with c = 1.23A1/3 − 0.60 fm and a = 0.52 fm [192]. The dependence of A in the parameter c causes significant differences for distinct isotopes. The form factors as a function of T for each germanium isotope are shown in Appendix B, Figure B.1. The total cross section further illustrates the difference caused by the isotopes. It is related to the differential cross section shown in Equation 7.21 by: σ(E) = ∫ Tmax 0 dσ dT (E, T )dT, (7.29) where Tmax = 2E2 2E+m . The total cross sections for all germanium isotopes in the neutrino energy range of 0-10 MeV are shown in Figure 7.12. An increase in the cross section with a higher neutron count is observed, as expected according to the discussion in Section 7.3.1. Figure 7.12: Total CEνNS cross sections for all germanium isotopes, calculated for the neutrino energy range of 0-10 MeV, the typical energy range for reactor neutrinos [181]. The differential event rate is the last step prior to obtaining the number of events. It is 88CHAPTER 7. EXPLORATION OF NEUTRINOS IN THE STANDARDMODEL AND BEYOND defined as: dN dT (T ) = tNt ∫ Emax Emin(T ) dσ dT (T,E)Φ(E)dE, (7.30) and is interpreted as the number of events in each ionization energy bin. As shown in Equation 7.30, the target number and livetime play a significant role in the differential event rate. The differential event rates for each isotope, adding together the contributions of all three diodes, are shown in Figure 7.13. The contributions for each individual detector can be seen in Appendix B, Figure B.2. Figure 7.13: Event rates for each germanium isotope as a function of ionization energy. The event rates and the number of events shown are the total contributions of all detectors: CONUS-2, 3, and 5. The number of events is calculated as the integral of the event rate with respect to the ionization energy ϵ(K), applying Equation 7.24 to account for quenching effects, and the efficiencies shown in Figure 7.11. This was calculated for each individual detector, taking into account the different energy thresholds, and the upper limit for the energy region of interest of 800 eVee, as discussed in Section 7.3.1. Lastly, the individual contributions for each detector were added to obtain the overall result. The number of events calculated for for Conus-2, 3, and 5 was 84, 135, and 111, respectively; bringing the total number to 330 7.4. STUDY OF SCENARIOS IN THE SM AND BEYOND 89 Detector Events Calculated Signal Detected Signal Predicted in This Work by CONUS+ by CONUS+ Conus-2 84 69 ± 47 96 ± 16 Conus-3 135 186 ± 66 135 ± 23 Conus-5 111 117 ± 75 116 ± 20 Total 330 395 ± 106 347 ± 59 Table 7.3: Events calculated for each detector and the total number of events. Consistent with the number of events calculated and measured by CONUS+ [181]. events (see Table 7.3). These results are consistent with both the measurements and the calculations done by the CONUS+ collaboration. 7.4.2 Weak Mixing Angle The weak mixing angle (θW ) is a parameter in the SM, proposed in the electroweak theory [193–195]. It receives the name of mixing angle due to its role in the relation between the photon (Aµ) and Z0 boson (Zµ) with the third component of the weak isospin triplet (W 3 µ) and the weak hypercharge boson (Bµ) [46]: Aµ = sin θWW 3 µ + cos θWBµ, Zµ = cos θWW 3 µ − sin θWBµ. (7.31) It also relates the masses of the W± and Z0 bosons through MW = MZ cos θW ; and the weak isospin SU(2)L and hypercharge U(1)Y couplings, g and g′ respectively, via: sin2 θW = g′2 g2 + g′2 . (7.32) The weak mixing angle is not constant but varies as a function of the transferred momentum q, and determining its value for a given energy scale is an important precision test of the SM [196]. CEνNS allows to probe this parameter at low energies, at which the form factor is close to unity. As discussed in Section 7.2, the weak mixing angle is present in the CEνNS cross section via the term gVp = 1 2 − 2 sin2 θW , in the weak charge. This characteristic of the CEνNS cross section was employed to find a best fit value of the weak mixing angle using the CONUS+ results. The fitting process was done using the χ2 function defined in Equation 7.25, with X = sin2(θW ) as the parameter to be fitted, calculating the predicted number of events, as de- scribed in the Section 7.4.1, for each iteration. The minimization was performed using imi- nuit, a Python implementation of the Minuit2 C++ library [197]. The procedure consisted in minimizing the χ2 function for sin2(θW ) values in the [0, 1] interval, varying the nuisance parameter α in the interval [−1, 1]. A minimization convergence was successfully achieved for the best fit value sin2 θW = 0.268± 0.047, where the uncertainty interval corresponds to 1σ. 90CHAPTER 7. EXPLORATION OF NEUTRINOS IN THE STANDARDMODEL AND BEYOND This result is shown in Figure 7.14, as a function of the energy scale. Additional results from a variety of processes and experiments are included for comparison: • Atomic parity violation (APV) [138, 198], • elastic electron-proton scattering (Qweak) [199], • Møller scattering (SLAC E158) [200], • electron-deuteron deep inelastic scattering (eDIS) [201], • neutrino-electron scattering (XENONnT) [202], • CEνNS from spallation neutron source (COHERENT CsI & LAr) [203], • and CEνNS from solar neutrinos (PANDAX-4T & XENONnT) [204]. The CONUS+ results for sin2 θW are consistent with all the other experiments considered within the 1σ uncertainty interval. The results from experiments using electrons (APV, Qweak and eDIS) have a significantly lower uncertainty than the experiments using neutri- nos (CONUS+, XENONnT, PandaX-4T and COHERENT). This work imposed the first CEνNS-derived limit on the weak mixing angle for reactor electron antineutrinos, from a measurement with a significance greater than 3σ. 7.4.3 Magnetic Moment of the Neutrino The magnetic moment is the most well studied of the electromagnetic properties of the neutrino [205–217]. In the SM, neutrinos are considered as massless; however, as discussed in Section 7.1, neutrinos must be massive in order to explain flavor oscillation. In extensions of the SM that account for neutrino masses, these acquire a magnetic moment as well. As an example, in a minimal extension of the SM that considers right handed neutrinos, a diagonal magnetic moment for the Dirac neutrino arises as a product of loop corrections involving W bosons and charged leptons. The magnetic moment produced by this model has the characteristic of being proportional to the neutrino mass in the diagonal elements of the neutrino mass eigenstates [218, 219]: µD ν,ii = 3eGFmν,i 8π2 √ 2 ≈ 3.2× 10−19 (mν,i 1eV ) µB, (7.33) where µD ν,ii denotes the diagonal magnetic moment of the Dirac neutrino, mν,i represents the mass of the ith neutrino mass eigenstate, and µB is the Bohr magneton. For Majorana neutrinos, only non diagonal magnetic moments will be non-zero (µM ν,ij ̸= 0 for i ̸= j). The magnetic moment of the neutrino can be probed using CEνNS measurements trough its cross section contribution, given by [203]: dσ dT µν = πα2 EMZ 2µ2 ν m2 e ( 1 T − 1 E + T 4E ) F 2(q2), (7.34) 7.4. STUDY OF SCENARIOS IN THE SM AND BEYOND 91 10 5 10 4 10 3 10 2 10 1 100 101 [GeV] 0.05 0.10 0.15 0.20 0.25 0.30 0.35 0.40 0.45 0.50 sin 2 W APV Qweak SLAC E158 eDIS CO NU S+ XE NO Nn T e- re co il Pa nd aX -4 T 8 B XE NO Nn T 8 B CO HE RE NT (th is wo rk ) Figure 7.14: Weak mixing angle as a function of the energy scale µ, in the modified minimal subtraction renormalization scheme (MS). The fit achieved in this work for the CONUS+ data is shown in red [186]. The SM prediction is shown in black. Other results are also presented for comparison [138, 198–204]. 92CHAPTER 7. EXPLORATION OF NEUTRINOS IN THE STANDARDMODEL AND BEYOND where µν is the neutrino magnetic moment normalized by the Bohr magneton, and αEM is the fine structure constant. The total CEνNS cross section including the magnetic moment contribution is: dσ dT = dσ dT SM + dσ dT µν (7.35) with dσ dT SM , the SM cross section, given by Equation 7.21. In this analysis, and in all instances in which the mixing angle was not the parameter to be optimized, the value used was sin2 θW = 0.2386 [220, 221]. A similar approach to the one outlined in Section 7.4.2 was used to impose an upper limit on the neutrino magnetic moment. χ2 optimization was used, with the cross section from Equation 7.35 implemented in Equation 7.27 for the calculation of the number of events, and µν as the parameter to be minimized using iminuit. The main difference was that rather than pursuing an absolute minimum, a χ2 profile was produced, in order to impose an upper limit. In this procedure, the χ2 function (equation 7.25) returns a value by optimizing the parameter α for µν values in the interval [0, 4.0×109] and subtracting the minimum for each iteration; the χ2 values are subsequently plotted as a function of µν (see Figure 7.15). This allowed to impose a 90% CL. upper limit of µν ≤ 5.6× 10−10µB. Figure 7.15: Neutrino Magnetic moment χ2 profile for CONUS+, the dashed line marks the 90% CL. upper limit, at µν ≤ 5.6× 10−10µB. The limit to the neutrino magnetic moment imposed in this analysis is shown in Figure 7.16. Also shown in the same Figure are results from direct DM detection experiments using elastic neutrino-electron scattering [222], and CEνNS limits from 8B solar neutrinos [223]; 7.4. STUDY OF SCENARIOS IN THE SM AND BEYOND 93 and from a combined CsI and LAr detection using a spallation neutron source by the CO- HERENT collaboration [203]. Additional results from experiments using HPGe detectors with a reactor as a neutrino source by TEXONO [224] and GEMMA [225] are also shown, including a previous limit imposed by the CONUS experiment at the Brokdorf site [226]. This is a more stringent limit due to a larger energy window and a longer exposure than that of the data used in this analysis. A limit for accelerator neutrinos by the Liquid Scin- tillator Neutrino Detector (LSND) collaboration is also included [227]. Other experiments included that imposed limits with data from solar neutrinos are: SuperKamiokande [228], and Borexino [229]. The result for the neutrino magnetic moment achieved in this analysis is more stringent than those obtained by COHERENT and the CEνNS direct DM detection experiments, and is also competitive when compared to the accelerator neutrino results by LSND. TEXONO, GEMMA and CONUS produced more stringent limits with reactor antineutrinos. The most stringent limits are imposed by the direct DM detectors using electron-neutrino scattering. 94CHAPTER 7. EXPLORATION OF NEUTRINOS IN THE STANDARDMODEL AND BEYOND e eff Neutrino flavor 10 12 10 11 10 10 10 9 10 8 Ne ut rin o m ag ne tic m om en t [ B ] CONUS+ (this work) TEXONO CONUS GEMMA COHERENT LSND DM (8B) DM (e recoil) KAMLAND BOREXINO Figure 7.16: 90% CL. upper limit for the neutrino magnetic moment normalized by the Bohr magneton, for the CONUS+ results [186]. Also shown in the same Figure are: direct DM detection experiments using elastic neutrino-electron scattering (e-recoil) [222] and CEνNS limits from 8B solar neutrinos [223], COHERENT [203], TEXONO [224], GEMMA [225], CONUS [226], LSND [227], KamLAND [228], and Borexino [229]. 7.4. STUDY OF SCENARIOS IN THE SM AND BEYOND 95 7.4.4 Non-Standard Interactions Within the SM, neutrinos participate in weak interactions and undergo flavor oscillations due to their small but nonzero mass. Various Beyond Standard Model (BSM) frameworks propose additional Non-Standard neutrino Interactions (NSI), which can modify neutrino propagation and interactions with matter. NSI typically arise from SM extensions that introduce new mediators, such as additional U(1) gauge symmetries with a Z ′ gauge boson [175, 230–234]. As a result, NSI provide a promising window for probing new physics. The effective neutral current NSI Lagrangian for neutrinos in matter (considering only vector interactions), takes the form [235, 236]: LNSI = −2 √ 2GF ϵ fV ℓℓ′ ( ν̄αγ ρPLνβ )( f̄γρf ) , (7.36) where: • GF is the Fermi constant, • ϵfVℓℓ′ is the parameter describing the strength of the NSI, • f denotes a first generation SM fermion (e, u, d), • and the indices ℓ and ℓ′ represent the neutrino flavors (e, µ, τ). The SM CEνNS weak charge, Equation 7.22, is generalized to include NSI in the following way (assuming that flavor changing parameters are negligible, i.e ϵfVℓℓ′ ≈ 0 for ℓ ̸= ℓ′ [232, 237, 238]): QNSI w,ℓ = Z ( gVp + 2ϵuVℓℓ + ϵdVℓℓ ) FZ(q 2) +N ( gVn + ϵuVℓℓ + 2ϵdVℓℓ ) FN(q 2). (7.37) Notice the dependence of ℓ in the NSI weak charge, indicating that the weak charge—and consequently, the differential cross section are now flavor-dependent: dσ dT ℓ = G2 F 2π MN(Q NSI w,ℓ )2 ( 2− MNT E2 ) . (7.38) The number of predicted events for each neutrino flavor can be calculated with a slightly modified version of Equation 7.27: Npred,ℓ = tNt ∫ Tmax Tmin ϵ(T )dT ∫ Emax Emin(T ) dσ dT ℓ (T,E)Φℓ(E)dE, (7.39) where the flavor dependence is made explicit in the cross section and the flux. Following this reasoning, the total number of events is the sum over the distinct flavors: Npred = ∑ ℓ Npred,ℓ. (7.40) As discussed in Section 7.3, the antineutrino flux produced at a nuclear fission reactor consists entirely of electron antineutrinos. This simplifies the analysis of NSI parameters for 96CHAPTER 7. EXPLORATION OF NEUTRINOS IN THE STANDARDMODEL AND BEYOND CONUS+ data, as all CEνNS events originate from a single neutrino flavor, eliminating the need for flavor-dependent calculations, i.e. QNSI w,ℓ = QNSI w,e and Npred = Npred,e. The NSI parameters in the CONUS+ results were analyzed using a χ2 optimization ap- proach, similar to the procedures employed for the weak mixing angle and the neutrino magnetic moment. However, significant differences were introduced to reduce computational time, requiring specific simplifications. The χ2 function was evaluated over a 400×400 grid of ϵuVℓℓ and ϵdVℓℓ values, spanning the interval [−1, 1]. To expedite the process, the nuissance parameter α was analitically minimized. This was achieved by taking the derivative of the χ2 function, equating to 0 and solving for α̂, the minimum value of α, as follows: dχ2 dα = 1 σstat ( 2αN2 pred + 2N2 pred − 2NpredNmeas ) + 2α σ2 α = 0, α̂ = σ2 α [ NpredNmeas −N2 pred N2 predσ 2 α + σ2 stat ] 4. (7.41) To further optimize the analysis, Npred was obtained using a scaling method based on the previously obtained SM event count. This approach exploits the fact that the Helm form factor (equation 7.28) takes the same form for the proton and neutron terms; as a result the weak charge squared term of the cross section can be written as follows: ( QNSI w,ℓ )2 = [ Z ( gVp + 2ϵuVℓℓ + ϵdVℓℓ ) +N ( gVn + ϵuVℓℓ + 2ϵdVℓℓ )]2 F (q2)2. Consequently, the weak charge term can be factored out of the integral in Equation 7.39, allowing for the scaling to be implemented: Npred,ℓ = (QNSI w,ℓ )2tNt ∫ Tmax Tmin ϵ(T )dT ∫ Emax Emin(T ) G2 F 2π MN × 1× ( 2− MNT E2 ) Φℓ(E)dE, = (QNSI w,ℓ )2tNt ∫ Tmax Tmin ϵ(T )dT ∫ Emax Emin(T ) G2 F 2π MN × Q2 w Q2 w × ( 2− MNT E2 ) Φℓ(E)dE, = (QNSI w,ℓ )2 Q2 w tNt ∫ Tmax Tmin ϵ(T )dT ∫ Emax Emin(T ) G2 F 2π MN ×Q2 w × ( 2− MNT E2 ) Φℓ(E)dE, Npred,ℓ = (QNSI w,ℓ )2 Q2 w Npred,ℓ(SM); (7.42) Npred,e = ( QNSI w,e )2 1959.3 330. (7.43) In the last step, the SM number of events was substituted (Npred(SM) = 330, as shown in Table 7.3); as well as the SM weak charge for the HPGe detectors (Q2 w = 1959.3). This approach led to the following flavor-dependent expression for the χ2 function: 4The second derivative confirms that α̂ corresponds to a minimum: d2χ2 dα2 = 2 [ N2 pred σ2 stat + 1 σ2 α ] > 0. 7.4. STUDY OF SCENARIOS IN THE SM AND BEYOND 97 χ2 NSI,ℓ = ( Nmeas,ℓ − (1 + α̂)Npred,l(X) σstat )2 + ( α̂ σα )2 . (7.44) This method eliminates the need to recalculate the convolution in Equation 7.39 at ev- ery iteration by replacing it with the calculation of the weak charge, which is significantly less computationally demanding. Moreover, this procedure does not introduce any loss of generality in the analysis. Additionally, to further minimize computing time, a parallel computation approach was implemented using the Python library multiprocessing [239]. This strategy allowed simulta- neous evaluations of different sets of parameter values, significantly improving efficiency in the use of computation resources. A 95% CL. limit contour was imposed in the ϵuVee − ϵdVee plane (see Figure 7.17). The results indicate that, along the allowed combinations of parameter values, the ϵuVee = ϵdVee = 0 point—which corresponds to the scenario in which the measurements are consistent with exclusively SM contributions, is within the contour. Furthermore, the limits imposed in this analysis are consistent with those from a similar study with data by the COHERENT experiment [238], which used CsI, LAr and HPGe detectors. The analysis presented here, which evaluates the weak mixing angle, the magnetic moment of the neutrino, and the NSI parameters for the CONUS+ results, was published in the journal Physical Review D [186], achieving to be the first analysis of the CONUS+ data to appear in a peer-reviewed publication. 98CHAPTER 7. EXPLORATION OF NEUTRINOS IN THE STANDARDMODEL AND BEYOND CO N U S+ (this w ork) CO H EREN T (CsI+ LAr+ G e) -1.00 -0.75 -1.00 -0.50 -0.25 0.00 0.25 0.50 0.75 1.00 -0.75 -0.50 -0.25 0.00 0.25 0.50 0.75 1.00 dV eeε u V e e ε Figure 7.17: 95% CL. limits for the NSI parameters ϵdVee and ϵuVee [186], compared with limits imposed with COHERENT measurements [238]. 7.5. NEUTRINO PRODUCTION IN CORE-COLLAPSE SUPERNOVAE 99 7.5 Neutrino Production in Core-Collapse Supernovae An additional case of interest for the study of neutrinos are those produced at a Supernova (SN). As a complement to the reactor antineutrinos analysis, in the following sections a study using simulations of SN neutrinos is presented, as well as the signals in the noble liquid detectors DEAP-3600 and LUX-ZEPLIN. This study aims to serve as the initial steps on evaluating the weak mixing angle, neutrino magnetic moment, and NSI parameters, for a signal of SN neutrinos. A Core-collapse SN, also known as a Type II SN, occurs at the end of a massive star’s life cycle. At this stage, the star’s core can no longer sustain nuclear fusion reactions to counteract gravitational forces, due to the accumulation of iron; this leads to a rapid collapse of the stellar core. This event occurs when the mass of the progenitor star is approximately between 8 and 30 M⊙ [240]. Core-collapse supernovae (SNe) are among the most energetic events in the universe, releasing approximately 1053 erg (1046 J) of energy, with 99% of this energy emitted as ∼ 1058 neutrinos of all flavors, with average energies ranging from 10 to 20 MeV [241, 242]. The core-collapse process of a SN can be summarized in five main stages, during which various neutrino production mechanisms occur (see Figure 7.18). 1. In the last stages of its life cycle, a massive star develops an onion-like layered structure with heavier elements like oxygen, silicon, and iron, concentrated towards the core. 2. As nuclear fusion becomes unsustainable due to the accumulation of iron, the in- ward gravitational pressure overcomes the outward thermal pressure. The iron-rich core, a “white dwarf”, remains stable due to electron degeneracy pressure until the Chandrasekhar limit is reached (∼1.4-1.5 M⊙) [243, 244]. At this critical point, the degeneracy pressure is overcome, triggering electron capture processes that produce electron neutrinos. Simultaneously, the photo-disintegration of iron nuclei within the core further reduces pressure support, rendering the core gravitationally unstable and leading to collapse. During this phase, electron captures continue, generating a signifi- cant flux of electron neutrinos. This process is known as neutronization, during which the lepton number in the core decreases, resulting in a neutron-rich environment. 3. As the core reaches nuclear densities, it undergoes a sudden rebound, launching a shock wave. This event, known as the core bounce, produces a burst of electron neutrinos. This is the first observable neutrino signal from the SN. 4. After the prompt electron neutrino burst is emitted, the outward moving shock wave is stalled due to the infalling material from the outer layers, in whats known as the accretion phase. This results in an increase of the core’s temperature, which is now effectively a proto-neutron star. The core’s temperature rises significantly, leading to the production of both electron neutrinos and electron antineutrinos through charged- current interactions: e− + p → n + νe and e+ + n → p + ν̄e. Additionally, muon and tau neutrinos, and their antineutrinos, are also produced trough neutral current processes such as: electron-positron annihilation e− + e+ → ν + ν̄, nucleon–nucleon bremsstrahlung N + N → N + N + ν + ν̄, and plasmon decay γ∗ → ν + ν̄, where N 100CHAPTER 7. EXPLORATION OF NEUTRINOS IN THE STANDARDMODEL AND BEYOND represents a proton or a neutron, and ν and ν̄ denote neutrinos and antineutrinos of all flavors. This phase lasts until the stalled shock wave emitted in the bounce is revived by neutrino heating, leading to the explosion of the SN. 5. The final stage of the SN is the cooling phase, during which neutrino emission continues, via direct Urca processes [245, 246] ((A,Z − 1) → (A,Z) + e− + νe and (A,Z) + e− → (A,Z − 1) + ν̄e), as well as trough the continuation of the neutral current processes from the previous phase. During this phase, the proto-neutron star gradually cools down, stabilizing as a cold (T ∼1 MeV) neutron star. The flavor oscillations of neutrinos are modified in a SN due to their interactions with elec- trons. As neutrinos travel through regions with varying electron densities, this phenomenon, known as the Mikheyev–Smirnov–Wolfenstein (MSW) effect [248, 249], plays a significant role in the flavor composition of the neutrino flux. This phenomenon arises because electron neutrinos experience coherent forward scattering via charged-current weak interactions with electrons in matter, effectively modifying their propagation characteristics. Such interactions can lead to resonant flavor conversion that occur when the effective mass difference between neutrino flavors, modified by their interactions with electrons in matter, matches the local electron density. At this resonance point, the probability of a neutrino changing from one flavor to another is significantly enhanced, leading to substantial alterations in the flavor composition of the neutrino flux as it propagates outward from the supernova core. The flavor conversions favored by the MSW effect in SNe depend on the neutrino mass ordering hierarchy [250]. In the NH, the resonance conditions favor the conversion of electron neutrinos to other flavors; conversely, in the IH, electron antineutrinos are more likely to undergo resonant conversion (see Figure 7.19). Although the MSW effect alters the flavor composition of the neutrino flux, it does not change the total number of neutrinos. The galactic average rate for core-collapse SNe in the Milky Way is approximately 1.63 events per century; which is equivalent to a time between SNe of ∼61 years [252]. However, this rate is not uniform throughout the galaxy. A more localized study estimates a SN rate of 5-6 times the galactic average, at a distance of 600 pc from the Sun [253]. This elevated rate in the solar neighborhood is attributed to a higher density of massive stars in nearby star-forming regions, which are progenitors of core-collapse SNe. The most recent record of a core-collapse SN in the vicinity of the Milky Way is SN1987A, which occurred at approximately 51 kpc from the Sun, in the Large Magellanic Cloud [254]. It was discovered on February 23, 1987, and was visible to the naked eye. Recent observations with the James Webb Space Telescope have provided compelling evidence for the existence of a neutron star within the remnants of SN1987A [255]. Notably, several neutrino observatories made detections two to three hours before the visible light from SN1987A reached Earth: • Kamiokande II detected 12 neutrinos [256]. • Irvine-Michigan-Brookhaven (IBM) detected 8 neutrinos [257]. • Baksan Underground Scintillation Telescope detected 5 neutrinos [258]. 7.5. NEUTRINO PRODUCTION IN CORE-COLLAPSE SUPERNOVAE 101 Figure 7.18: Stages of a Core-collapse SN [242, 247]. A late-stage massive star with onion-like structure of heavy elements concentrated towards the core (top left). Nuclear fusion becomes unsustainable due to the accumulation of iron; the gravitational pressure is balanced by the electron degeneracy pressure of the “white dwarf” until in reaches the Chandrasekhar mass limit (∼1.4-1.5 M⊙) [243, 244]. The degeneracy pressure is overcome, and neutrinos are emitted trough electron capture, triggering neutronization. Photo-disintegration of the iron nuclei further enables the core collapse (bottom left). As the iron core reaches nuclear densities, it rebounds, launching a burst of νe (bottom right). In the accretion phase, the SN shock wave is stalled due to the infall of material (top right). Core temperature rises, leading to the production of νe and ν̄e in e− and e+ captures. Neutrinos of all flavors (ν) are also produced in neutral current processes such as: pair annihilation, nucleon–nucleon bremsstrahlung and plasmon decay. The shock wave is revived by neutrino heating, and the SN explosion is launched. Lastly, the cooling phase (center), during which neutrino emission of all flavors continues via direct Urca processes [245, 246], and the continuation of the neutral current processes from the previous phase. The remnant of the SN is a neutron star. 102CHAPTER 7. EXPLORATION OF NEUTRINOS IN THE STANDARDMODEL AND BEYOND Figure 7.19: Simulated SN neutrino fluences (denoted as λ in this Figure) as a function of the neutrino energy E. The Figure compares different cases of the adiabatic MSW effect [251]: (a) MSW effect not considered (νx refers to the sum of νµ, ντ and the respective antineutrinos), (b) MSW effect considered assuming the NH neutrino mass ordering, and (c) MSW effect considered assuming the IH neutrino mass ordering. 7.6. CEνNS SIMULATIONS FOR DEAP-3600 AND LUX-ZEPLINWITH ESTRELLANUEVA103 These detections provided observational support for theoretical supernova models, confirming that approximately 99% of the collapse energy is emitted as neutrinos of all flavors [259]. Additionally, the observations coincided with predictions estimating the emission of around 1058 neutrinos, each with average energies in the range of a few tens of MeV [260]. Having established the importance of neutrino astronomy and the effectiveness of reactor neutrino CEνNS experiments in probing new physics, this study serves as an initial effort in constraining the weak mixing angle, the neutrino magnetic moment, and non-standard neutrino interaction (NSI) parameters using simulated SN neutrinos. The following sections outline a methodology and show potential implications for future research involving SN neutrinos, with CEνNS as the detection channel. 7.6 CEνNS Simulations for DEAP-3600 and LUX-ZEPLIN with EstrellaNueva The simulations for this analysis were done using EstrellaNueva [251], an open-source software designed to study the interactions of neutrinos emitted by SNe 5. It provides a computational framework to calculate the cross sections, fluences and the expected inter- action rates of SN neutrinos in various detector materials, incorporating multiple detection channels, including CEνNS, and allowing for custom target compositions. EstrellaNueva implements the analytical SM CEνNS cross section (equation 7.21), and various nuclear form factors, including Helm (equation 7.28). The adiabatic MSW effect can be incorporated into the simulations in either mass ordering hierarchy; both of which are considered in the relevant analysis in this work. Additionally, EstrellaNueva allows users to set SN distances from Earth, and choose from various progenitor star models. These features make it a valuable tool for studying SN neutrinos in detectors, like DM experiments, which are particularly sensitive to CEνNS interactions. Following the approach taken in [262], three SN distances were selected to represent dif- ferent astrophysical scenarios: regions close to the Sun, a medium-range galactic distance, and areas near the edge of the Milky Way disk. These distances, as measured from Earth, were chosen as follows: • 0.196 kpc (640 ly): the approximate distance to Betelgeuse, a red supergiant with a mass of ∼15 M⊙ [263]. • 10.0 kpc (32600 ly): roughly equivalent to the distance to the center of the Milky Way. • 20.0 kpc (65200 ly): a distance approaching the opposite outer limits of the galactic stellar disk. Additionally, the chosen SN model uses the equation of state for hot dense matter proposed by Latimer and Swesty [264], with a nuclear incompressibility of 220 MeV (LS220-s15.0 EoS). Developed by the Core-Collapse Modeling Group at the Max Planck Institute for As- trophysics, this model was implemented in EstrellaNueva [251, 265]. It assumes a progenitor star of 15 M⊙, coinciding with the approximate mass of Betelgeuse. 5EstrellaNueva Version 1.1 was used for this work [261]. 104CHAPTER 7. EXPLORATION OF NEUTRINOS IN THE STANDARDMODEL AND BEYOND The targets chosen for this study were two noble liquid DM detectors: DEAP-3600 (see Section 4.1), and LUX-ZEPLIN (LZ), the largest detector currently in operation [266]. LZ is a scintillation detector, consisting of a liquid xenon (LXe) time projection chamber (TPC) with an active mass of 7 tonne (see Figure 7.20). It makes use of the LXe scintillation properties that enable pulse shape discrimination, allowing for the distinction between nu- clear and electron recoils. LZ is located underground at the Sanford Underground Research Facility, at Lead, South Dakota, shielded by an overburden of 4300 m water-equivalent [63]. It started taking data on December, 2021; and currently holds the record for the most strin- gent limit on spin-independent WIMP-nucleon cross section, at 2.1×1048cm2 at the 90% CL. for a WIMP mass of 36 GeV/c2 [71] (shown in Figure 3.5). Figure 7.20: Diagram of the LZ detector structure [63]. 1) The LXe TPC, monitored by 2 arrays of PMTs. 2) Double-walled vacuum insulated titanium cryostat, surrounded by an outer detector consisting of an organic liquid scintillator loaded with gadolinium (GdLS). 3) Array of 8” PMTs monitoring the GdLS. 4) Water shield. 5) High voltage connection for the cathode. 6) Conduit for neutron calibration sources. The procedure for calculating the number of events in this analysis closely resembles the one outlined for the CONUS+ detector in Section 7.4.1; however, there is a significant difference related to the fluxes. In the CONUS+ analysis, the assumption was made that the neutrino flux remained approximately constant during reactor on periods, this allowed for the flux and the livetime to be treated as independent. In the case of a neutrino burst in a SN, this assumption no longer holds, as the neutrino flux varies heavily during the different stages of the SN (see Figure 7.21). For this reason, instead of using the flux and the livetime separately to calculate the number of events, as in Equation 7.27, the fluence (φD(t, E)) was used, which relates to the flux as follows6: 6In the case of a time independent flux and a constant distance to the source, this reduces to ϕ(t, E) = ∫ t2 t1 Φ(E)dt = ∆tΦ(E), recovering Equation 7.27. 7.6. CEνNS SIMULATIONS FOR DEAP-3600 AND LUX-ZEPLINWITH ESTRELLANUEVA105 φD(t, E) = ∫ t2 t1 ΦD(t, E)dt (7.45) where t1 and t2 denote the starting and ending times of the SN neutrino detection, and the dependence on the SN distance has been made explicit in the subindex D. Figure 7.21: Simulation of the fluxes of distinct neutrino species during a SN, as a function of time measured after the bounce [267]. The neutrino denoted as νµ represents the sum of all the remaining neutrino species (νµ, ν̄µ, ντ and ν̄τ ) After this adjustment, the equation for the number of events is: Npred,D = Nt ∫ Tmax Tmin ϵ(T )dT ∫ Emax Emin(T ) dσ dT (T,E)φD(t, E)dE, (7.46) where the livetime is implicit in the fluence. This equation is used to calculate the number of events for each detector, at a distance D = (0.196, 10.0, 20.0) kpc. The CEνNS cross section is given by Equation 7.21, the total cross sections (equation 7.29) for the Xe isotopes and 40Ar are shown in Figure 7.22. The form factors as a function of the recoil energy for Xe and Ar are shown in Appendix B, Figure B.3. The other elements needed for calculating the number of events are: the number of targets, efficiency functions, fluences, energy thresholds, and the energy resolution functions. The number of targets was determined using the total active noble liquid masses for each detector, 3.3 and 7 tonne, for DEAP and LZ, respectively. In the case of DEAP, the LAr composition was assumed to be 100% 40Ar, while for the LXe in LZ, the natural relative abundances of the Xe isotopes were used, as shown in Table 7.4. 106CHAPTER 7. EXPLORATION OF NEUTRINOS IN THE STANDARDMODEL AND BEYOND Figure 7.22: Total CEνNS cross sections for all Xe isotopes and 40Ar. Xenon Isotope (A) Relative Abundance (%) 124 0.095 126 0.089 128 1.910 129 26.401 130 4.071 131 21.232 132 26.909 134 10.436 136 8.857 Table 7.4: Natural abundances of the xenon isotopes, used in the calculation of the target number for each isotope in LZ [268]. 7.6. CEνNS SIMULATIONS FOR DEAP-3600 AND LUX-ZEPLINWITH ESTRELLANUEVA107 The recoil energy detector efficiencies used were the trigger efficiencies for each detector (see Figure 7.24), assuming that no further cuts would be needed for the detectors in the case of a highly time-localized and signal-dominant event such as a SN neutrino burst. The fluences were obtained from the EstrellaNueva simulations (see Figure 7.23), which cover a range of incident neutrino energies up to 100 MeV. The energy threshold for LZ was found to be 1 keV [269], while a threshold of 52 keV was used for DEAP-3600 [80]. Lastly, the energy resolution functions model the fraction of the nuclear recoil energy that is collected by the PMTs, analogous to the quenching function in Section 7.4.1. For both detectors, the energy resolution was modeled as a gaussian function, with a standard deviation that varies as a function of the recoil energy. The standard deviation for DEAP is [81]: σDEAP (T ) = √ 1.4× 10−3T + 4.0× 10−7T (7.47) where T denotes the recoil energy function. The energy resolution implemented for LZ was a combination of a linear segment, and a fit of the general form σLZ = A√ T , where the constant A was determined by the resolution values specified in [270]. The resulting function is represented as follows: σLZ(T ) =    4.7×10−6 √ T for T ≤ 3.5 keV, 3.0× 10−2T for T > 3.5 keV. (7.48) A gaussian sampling was applied to the event rates with the resolution functions as the standard deviation for each detector, resulting in the event rates shown in Figure 7.25. The total number of events at a given distance was obtained using Equation 7.46, and adding the contributions for all isotopes in each noble liquid target material. The results are presented in Table 7.5. A steep decline in the number of events as a function of distance is observed, corresponding to an inverse square law in the neutrino fluence emitted by the SNe. LZ is able to detect a significant signal across the entire galactic disk, at all distances probed; while DEAP-3600 is only able to gather a number of events greater than 1 at 0.196 kpc. This can be attributed mainly to the CEνNS cross section for 40Ar being approximately one order of magnitude lower compared to the same cross section for the Xe isotopes, as shown in Figure 7.22. The significantly higher target mass of LZ, nearly double that of DEAP-3600, is also a significant factor. In the following section these results are used to show potential implications for the weak mixing angle, the neutrino magnetic moment and the NSI parameters. Detector Distance [kpc] 0.196 10.0 20.0 LZ 104809 40 10 DEAP 468 <1 <1 Table 7.5: Simulated SN neutrino CEνNS at different distances for LZ and DEAP-3600 detectors. 108CHAPTER 7. EXPLORATION OF NEUTRINOS IN THE STANDARDMODEL AND BEYOND Figure 7.23: Fluences simulated with EstrellaNueva. Top: flavor composition of the fluences at 0.196 kpc, accounting for the MSW effect assuming the NH ordering. Bottom: total fluences for each distance considered. 6 5 .......... 4 N I E u r-i I 3 > Q) ::E S-2 1 o 1e13 (\ I \ \ , \ A \ ~ '- ~ I " t " ~s 1'--- o 20 40 E [MeV] ~---" ............. ,/ ' :-- " 60 1 1 SN Distance = 0.196 kpc t- -- Ve -- Vil -- VT --- Ve --- Vil --- vT -- Total 80 100 .L .1 Total Fluences at SN Distances [kpc] t- - , 1- - 1 ' ... ".-------+------------+------------1 : "" 0.196 10 20 J , " , , , " , " " ", 1011, - -+----------r----------r-------- ~~ --------_r--------~--- 1 ""'" N " : ... " :' /'-'-'-. ...... ~, :'/' ....... , ". "" : . ' '''-. __ ----=....0.-:-+-__________ +--________ --+-_____ '-0,.------__ +--_1 :/ ' ..... " .... ,~ 1 ", I E u r-i I 109 > Q) ::E ". ; '. ", ' .. "" S- 107 ". ", ..... . I ·'·,·1········ .. "" ..... . 105, - -+ ----------+-----------r----------r----------r---- ·, ~ -- ~ --- 1 I ". I '. O 20 40 60 80 100 E [MeV] 7.7. SN NEUTRINO PARAMETER ANALYSIS: SM AND BEYOND 109 Figure 7.24: Detection efficiencies for DEAP-3600 (left) [81] and LZ (right) [270]. For both detectors, the trigger efficiency curves were used (the light blue curve labeled “high upper bound” for DEAP-3600, and the blue “Trigger” curve for LZ). 7.7 SN Neutrino Parameter Analysis: SM and Beyond This section presents a proposed methodology for utilizing the previously discussed results to derive implications for the weak mixing angle, the neutrino magnetic moment, and the NSI parameters. To this end, the approach described in Sections 7.4.2, 7.4.3 and 7.4.4 was adapted, with modifications that account for the physical differences between reactor and SNe neutrinos, and the simulated nature of the data. This analysis represents an early effort in using DM detectors and SN neutrinos to investigate scenarios in the Standard Model and beyond, laying the ground work for future studies. Equation 7.25 was implemented for a χ2 optimization approach, under the assumption that for a highly time-localized SN neutrino burst, the background contribution can be neglected. This approach requires predicted and measured numbers of events, the predicted events are the results presented in Table 7.5. To emulate the role of measurements, a 1σ interval is constructed for each predicted result, by assuming a Poisson statistical distribution to calculate upper and lower bounds, as follows: N+ = Npred + √ Npred, N− = Npred − √ Npred. (7.49) The upper and lower bounds for Nmeas, denoted as N+ and N− respectively, are presented in Table 7.6. Under these circumstances, the χ2 function for a given distance and detector takes the following form: χ2 b = ( Nmeas,b − (1 + α)Npred(X) σstat,b )2 + ( α σα )2 , (7.50) 110CHAPTER 7. EXPLORATION OF NEUTRINOS IN THE STANDARDMODEL AND BEYOND Figure 7.25: Event rates for CEνNS at various SN distances. DEAP-3600 and LZ at 0.196 kpc (top left and right respectively), and LZ at 10.0 kpc (bottom left) and 20.0 kpc (bottom right). The event rates for DEAP at 10.0 and 20.0 kpc can be seen in Appendix B, Figure B.4. 7.7. SN NEUTRINO PARAMETER ANALYSIS: SM AND BEYOND 111 Detector Distance [kpc] 0.196 10.0 20.0 Npred N+ N− Npred N+ N− Npred N+ N− LZ 104809 105133 104485 40 46 34 10 13 7 DEAP 468 490 446 <1 <1 <1 <1 <1 <1 Table 7.6: Predicted and measured events with 1σ intervals for SN neutrino CEνNS at different distances for LZ and DEAP-3600 detectors. where the subindex b represents an upper or lower bound, thus Nmeas,b = (N+, N−) and σstat,b = ( √ N+, √ N−). This approach resulted in sets of 2 minima for the optimized physical parameters, one for each of the bounds. These minima were combined, or a conservative value was taken, depending on the specific circumstances and practicality associated with each parameter. The nuisance parameters in this analysis are conservative values that account for the main astrophysical uncertainty sources, and detector efficiencies: • Distance to SN: 25% [271], • Helm form factor: 5% [272], • Neutrino flux for all 6 species: 10% [273], • DEAP-3600 efficiency: 10% [81], and LZ efficiency: 15% [270]. The uncertainty in the nuisance parameter α for each detector is the following: σα,DEAP = √ 0.252 + 0.052 + 0.102 + 0.102 = 0.29, (7.51) σα,LZ = √ 0.252 + 0.052 + 0.102 + 0.152 = 0.31. (7.52) σα,LZ is slightly higher due to LZ’s efficiency, and both have similar values to the uncertainty used for the CONUS+ study, shown in Equation 7.26. 7.7.1 Results for the Weak Mixing Angle Following the order in the CONUS+ analysis, the first parameter to be optimized was the weak mixing angle (see Section 7.4.2). The χ2 function with upper and lower bounds yielded two optimal values for sin2 θW . The values were subsequently combined under the assumption of a squared distribution, constructing upper and lower limits by adding or subtracting its uncertainty to the upper and lower optimal values, respectively. The midpoint between these limits was then calculated, representing the central value of the combined 1σ uncertainty interval. The findings are presented in Table 7.7. The values for sin2 θW,+ are higher than those of sin2 θW,− for both detectors at 0.196 kpc and across all distances for LZ; however, the uncertainty intervals make these results consistent with the SM prediction of sin2 θW = 0.238 [220, 221]. As the distance increases, the central values deviate further, increasing for sin2 θW,+ and decreasing for sin2 θW,−, this accounts for the significant increase 112CHAPTER 7. EXPLORATION OF NEUTRINOS IN THE STANDARDMODEL AND BEYOND in the combined uncertainty interval. The uncertainties for sin2 θW,+ and sin2 θW,− increase for more distant SNe, this is expected due to the statistical uncertainty becoming more significant at greater distances. Distance to Supernova [kpc] Detector 0.196 sin2 θW,+ sin2 θW,− sin2 θW,c DEAP 0.245± 0.044 0.232± 0.042 0.239± 0.050 0.239± 0.055 0.238± 0.055 0.238± 0.055 10.0 LZ 0.265± 0.064 0.210± 0.057 0.241± 0.088 20.0 0.289± 0.082 0.179± 0.070 0.240± 0.131 Table 7.7: Upper (sin2 θW,+), lower (sin 2 θW,−) and combined (sin2 θW,c) weak mixing angle intervals for LZ and DEAP-3600. The central values and combined uncertainty intervals are plotted as a function of the SN distance in Figure 7.26. LZ is capable of setting limits for the weak mixing angle for SNe occurring all across the galactic disk. At the shortest distance considered in this study, the combined uncertainty of LZ is slightly larger than that of DEAP, due to a higher uncertainty in the nuisance parameter, that can be traced back to a higher conservative estimate in the efficiency for LZ. The combined uncertainty interval increases with distance, as discussed previously. This can be interpreted as an interval in which the central value for a sin2 θW measurement would be contained, if the difference between the predicted and measured number of SN CEνNS is within 1σ. Lastly, Figure 7.27 compares the sin2 θW results from this study with those from CONUS+ and other experiments. For SNe at distances comparable to Betelgeuse, both detectors demonstrate the capability to constrain sin2 θW with a precision similar to that of CONUS+. At greater distances within the Milky Way, LZ achieves precision comparable to experi- ments utilizing solar neutrinos and neutrino-electron scattering, such as PANDAX-4T and XENONnT. These findings support the feasibility of probing the weak mixing angle with SN CEνNS measurements from DM detectors like LZ and DEAP-3600. 7.7.2 Findings for the Neutrino Magnetic Moment The next parameter to be fitted was the magnetic moment of the neutrino. Similar to the case of the weak mixing angle, two minima per detector were obtained for each distance, accounting for the N+ and N− bounds. These findings are listed in the Table 7.8. LZ sets bounds for all distances evaluated, and as the distance increases the magnetic moment limit for the upper bound (µν [µB]+) increases as well, while the lower bound value (µν [µB]−) decreases. This inverse trend can be explained by referring back to the CEνNS differential cross section with the µν contribution (equations 7.34 and 7.35). The magnetic moment contribution is added to the SM cross section, this implies that a lower number of events than predicted by the SM, as is the case for N−, would require a lower contribution 7.7. SN NEUTRINO PARAMETER ANALYSIS: SM AND BEYOND 113 Figure 7.26: Central values and combined uncertainty intervals for sin2 θW as a function of SN distance. 114CHAPTER 7. EXPLORATION OF NEUTRINOS IN THE STANDARDMODEL AND BEYOND Figure 7.27: Weak mixing angle results for SN simulations, compared to the CONUS+ work described in Section 7.4.2, and other experiments [138, 198–204]. The SM prediction is shown in black. 7.7. SN NEUTRINO PARAMETER ANALYSIS: SM AND BEYOND 115 from the magnetic term, thus resulting in a lower µν . The opposite is true for the case of a measurement resulting in more events than the SM prediction, this implies a stronger magnetic contribution and a larger µν . This relationship enables the magnetic moment derived from the lower bound (µν [µB]−) to set more stringent limits. Distance to Supernova [kpc] Detector 0.196 µν [µB]+ µν [µB]− DEAP 3.7× 10−9 3.4× 10−9 2.0× 10−9 2.0× 10−9 10.0 LZ 2.4× 10−9 1.9× 10−9 20.0 2.8× 10−9 1.9× 10−9 Table 7.8: LZ and DEAP-3600 90% CL. upper limits for the upper (µν [µB]+) and lower (µν [µB]−) bounds of the neutrino magnetic moment. The limits for µν as a function of the SN distance are shown in Figure 7.28. The upper limits were not combined in this analysis. Although a variety of methods exist to achieve this [274, 275], their implementation requires significant computational resources, which were unavailable and beyond the primary scope of this study. Instead, the least stringent limits were chosen as conservative estimates. At 0.196 kpc LZ is able to set a lower limit than DEAP-3600, this is attributed to statistical effects due to LZ detecting more events than DEAP-3600 by a factor of ∼220. The statistical uncertainty is also responsible for the gradual increase in the limit set by LZ with increasing SN distance. The limits set in this study for 0.196 kpc were compared to the results obtained for CONUS+ in Section 7.4.3, and other experiments, this is shown in Figure 7.29. These re- sults are less stringent than those set with the CONUS+ analysis, however, are comparable to those by COHERENT [203] and limits from 8B solar neutrinos [223]. The results pre- sented here indicate that limits imposed on the neutrino magnetic moment via SN neutrinos measurements on DM detectors would be comparable and complementary to studies using other detection mechanisms. The SN distance does not play a significant role in this re- gard, since the difference in the limits is not significant at the scale shown in Figure 7.29. The cross section difference between LXe and LAr, as well as the target mass difference between LZ and DEAP-3600, account for an appreciable difference, however, it is negligible when compared to the differences with limits set by other detection channels such as elastic neutrino-electron recoil ([222] and [229]). 116CHAPTER 7. EXPLORATION OF NEUTRINOS IN THE STANDARDMODEL AND BEYOND Figure 7.28: LZ and DEAP-3600 90% CL. upper limits for µν as a function of SN distance. The limits shown are the most conservative limits imposed by each detector at a given distance. 7.7. SN NEUTRINO PARAMETER ANALYSIS: SM AND BEYOND 117 Figure 7.29: Upper limits at the 90% CL. set for SN simulations. The limits shown here are the results for the SN distance of 0.196 kpc, the variation in the limits due to the distances studied in this work is not significant at this scale. The CONUS+ limit obtained in Section 7.4.3 is also shown. Other limits shown in the same Figure are: direct dark matter detection experiments using elastic neutrino-electron scattering (e-recoil) [222] and CEνNS limits from 8B solar neutrinos [223], COHERENT [203], TEXONO [224], GEMMA [225], CONUS [226], LSND [227], KamLAND [228], and Borexino [229]. 118CHAPTER 7. EXPLORATION OF NEUTRINOS IN THE STANDARDMODEL AND BEYOND 7.7.3 Implications for Non-Standard Interactions The NSI parameters were analyzed last in this study. As discussed in Section 7.4.4, these parameters are flavor-dependent, requiring flavor-specific number of events for the χ2 optimization process. This poses a significant challenge when analyzing NSI parameters with SN neutrinos, as the SN flux, unlike that of reactor neutrinos, is composed of multiple flavors. Furthermore, CEνNS measurements do not allow for the differentiation between neutrino flavors. As a consequence, the results obtained are sensitive to the hierarchy in the neutrino mass ordering trough the influence of the MSW effect. Both hierarchies are considered in this study. One possible approach to estimating the flavor composition of a SN neutrino flux and probing NSI parameters is to use SM predictions to determine the expected flavor ratios in the event count. These ratios can then be applied to the measured event count to infer the flavor composition. EstrellaNueva provides flavor-specific fluences, facilitating this method. The flavor ratios are defined as follows: rℓ = Npred,ℓ Npred , (7.53) where ℓ = (e, µ, τ) is the flavor index. Subsequently, the event count for each flavor is determined as: Nmeas,ℓ = rℓNmeas. (7.54) This approach was used to determine the flavor-dependent upper and lower bounds: N±,ℓ = rℓ ( Npred ± √ Npred ) = Npred,ℓ ± √ Npred,ℓ Npred Npred,ℓ. 7 (7.55) The predicted event count, along with the flavor compositions provided by EstrellaNueva, as well as the upper and lower bounds estimated using the described method, are presented in Table 7.9. Significant differences in the flavor composition were observed for all neutrino flavors; however, these differences were particularly pronounced when comparing e and µ neutrinos at closer distances. This is a direct consequence of hierarhcy dependence in the MSW effect. The NSI contours were obtained by following the procedure detailed in Section 7.4.4, utilizing a 400×400 grid for evaluating the ϵuVℓℓ and ϵdVℓℓ parameter pairs in the χ2 function in Equation 7.44. The contours for the upper and lower bounds were combined by selecting the points in the grid with the lower χ2 value from each bound. This resulted in 95% CL. contours including all the points from both bounds that satisfy the threshold criterion, therefore setting a conservative contour. Another significant difference compared to the CONUS+ contour, is that in this analysis contours can be obtained for all neutrino flavors, thanks to the composition of the SN fluences. 7The total upper and lower bounds are recovered from the flavor-dependent bounds as follows: N± = ∑ ℓ ( Npred,ℓ ± √ Npred,ℓ Npred Npred,ℓ ) = ∑ ℓ Npred,ℓ ± 1√ Npred ∑ ℓ √ N2 pred,ℓ = Npred ± √ Npred. 7.7. SN NEUTRINO PARAMETER ANALYSIS: SM AND BEYOND 119 Distance to Supernova [kpc] Detector Event Count 0.196 e µ τ Total Normal Hierarchy Mass Ordering Npred,ℓ 171 145 152 468 DEAP N+,ℓ 179 152 159 490 N−,ℓ 163 138 145 446 Npred,ℓ 35880 34119 34810 104809 N+,ℓ 35991 34225 34917 105133 N−,ℓ 35769 34014 34702 104485 10.0 Npred,ℓ 14 13 13 40 LZ N+,ℓ 16 15 15 46 N−,ℓ 12 11 11 34 20.0 Npred,ℓ 4 3 3 10 N+,ℓ 5 4 4 13 N−,ℓ 3 2 2 7 0.196 Inverted Hierarchy Mass Ordering Npred,ℓ 145 165 158 468 DEAP N+,ℓ 152 173 165 490 N−,ℓ 138 157 151 446 Npred,ℓ 31184 37042 36583 104809 N+,ℓ 31280 37156 36697 105133 N−,ℓ 31087 36927 36471 104485 10.0 Npred,ℓ 12 14 14 40 LZ N+,ℓ 14 16 16 46 N−,ℓ 10 12 12 34 20.0 Npred,ℓ 3 4 3 10 N+,ℓ 4 5 4 13 N−,ℓ 2 3 2 7 Table 7.9: Flavor compositions of event counts for DEAP-3600 and LZ, for both mass ordering hierarchies. The predicted counts (Npred,ℓ) were calculated with Equation 7.46, using the flavor-specific fluences provided by EstrellaNueva. The upper and lower bounds (N+,ℓ and N−,ℓ) were estimated with the ratio-based method defined in Equation 7.55. 120CHAPTER 7. EXPLORATION OF NEUTRINOS IN THE STANDARDMODEL AND BEYOND The combined contours for LZ and DEAP-3600 at 0.196 kpc are shown in Figures B.5 and B.6, in Appendix B. Due to the small variation in the upper and lower bounds, attributable to the low statistics, the analysis for LZ at 10.0 and 20.0 did not produce meaningful contours. The contours set by DEAP-3600 are narrower than those set by LZ; however, both detectors are capable of probing the NSI parameters for SNe at 0.196 kpc. The electron NSI parameters were also compared with the CONUS+ contour obtained in Section 7.4.4 of this work, this is presented in Figure 7.30. The DM detector contours are more stringent than those set from the CONUS+ analysis. This is explained by the larger relative difference between measured and predicted events for CONUS+, 395 and 330 respectively; while the upper and lower bounds for DEAP and LZ are closer to the predicted values, as shown in Table 7.9. The mass hierarchies also produce a variation in the contours when comparing the DM detectors, this is due to the statistical difference in the electron- flavored fluxes in each mass ordering scenario. A comparison of NH vs IH contours for LZ and DEAP is shown in Figure B.7, in Appendix B. 7.8 Conclusions from the Parameter Analysis with SN Neutrinos This study represents an initial step toward exploring potential implications for SM and beyond parameters through the measurement of SN CEνNS in noble liquid DM detectors. The results achieved support the feasibility of using SN neutrinos to probe the weak mixing angle, the neutrino magnetic moment, and, with some limitations, the NSI parameters. The weak mixing angle and the neutrino magnetic moment can be probed by a detector such as LZ for SNe occurring all across the stellar disk of the Milky Way. A LAr detector, like DEAP-3600, is capable of probing the same parameters at distances near the solar vicinity. The potential values and upper limits obtained by these means are comparable to those set by reactor, spallation neutron sources, and solar neutrino experiments. The NSI analysis faces various limitations, the most significant ones encountered in this study being: the inability of obtaining flavor-specific event counts via CEνNS, the depen- dence on the assumed neutrino mass ordering hierarchy due to the influence of the MSW effect, and the low statistics for SNe ocurring deeper in the galaxy. The approach taken here to approximate the flavor composition of a SN neutrino burst allows for setting NSI contours for all three flavors at closer distances. However, the two mass hierarchy cases remain indis- tinguishable through methods incapable of measuring flavor-specific events, both should be treated as separate scenarios since there are significant differences in the flavor composition for neutrino bursts at short SN distances. Future extensions of this study could incorporate different stellar masses and SN models. The upper and lower bounds could be extended to 2 or 3σ, this would provide further insights into the role of SN neutrinos in future research. The range of considered detectors could also be broadened to include current DM experiments, such as PandaX-4T [66] and XENONnT [276]. Additionally, potential implications could be explored for future multi-ton detectors, such as DarkSide-20k [277], DARWIN/XLZD [278], and ARGO [279]. 1.00 3600 (NH) -——— —3 NH) | ( Ma LZ CONUS+ 0.75 1 o LA LA N o o -0.754 E DEAP- 0.25 - O (LAA -1.00 -0.50 0.25 0.00 0.25 0.50 0.75 ¿dv == 0.75 1.00 In Lz 1H) ( 0 CONUS+ 1.00 0.7543 0.50 7 1 19) % o 0.00 e AAA EU OS AN 3 050) eiii rc ci -0.754 ES DEAP- 1.00 -0.50. 0.25 0.00 0.25 0.50 0.75 ¿dv ee 0.75 -1.00 NA” 7.8. CONCLUSIONS FROM THE PARAMETER ANALYSIS WITH SN NEUTRINOS121 Figure 7.30: 95% CL contours for NSI electron parameters from DEAP-3600 (green), LZ (purple), and CONUS+ (red). The CONUS+ result was obtained in Section 7.4.4. The DEAP-3600 and LZ contours correspond to a SN at a distance of 0.196 kpc. NH mass ordering is shown at the top, IH mass ordering at the bottom image. ~QJ :JQJ Lo ~QJ :JQJ Lo . 0 .75 0.50 _____________________ 1- ________ _ , , ---------------------¡---------------------¡---------, , 0.25 , , , , , , , , , , , , , , , , , , , , , , , , , , , , , ---------------------r--------------------r--------------------r------, , , 0.00 , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , I I I I I I I I I I I I _____________________ 1- ____________________ 1- ____________________ L ____________________ L ___ _ I I I I I I I I -0. I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I ---------------------¡---------------------¡---------------------¡---------------------¡---------------------1"-- I I I I I -0.50 I [[[[I] ~ Z ( ) : l I I -0.75 ' E3 ~~~~~~OO {NH),' ¡....,.. , -l. 0 +------+------+-------I---+-----+-----+--- ~ -----....; ""-'-'""=-===-q -l.00 -0.75 . 0 -0.25 . 0 . 5 0.50 . 0 .75 5 0.25 0.00 -0.25 -0.50 -0.75 ---------------------1---------- , , ---------------------¡---------------------¡---------, , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , ---------------------r--------------------r--------------------r------, , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , E dV ee I I I I I I I I I I I I _____________________ 1- ____________________ 1- ____________________ L ____________________ L ___ _ I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I ---------------------¡---------------------¡---------------------¡---------------------¡---------------------1"-- I I I I I I I I I I I I I I I , , , , , , I [[[[I] LZ (1 ) i i i . 5 1.00 ' E3 ~~~~~~OO (lH) ¡¡" -l.00 +------+------+-------I---+-----+-----+--- ~~ """"--" ..."... l 0 -0.75 .50 -0.25 . 0 . 5 0.50 E dV e . 5 1.00 122CHAPTER 7. EXPLORATION OF NEUTRINOS IN THE STANDARDMODEL AND BEYOND 8 Conclusions Radiogenic neutrons constitute a significant background source that requires dedicated strategies to mitigate their impact on the detector’s sensitivity. Their characterization is carried out within the Profile Likelihood Ratio (PLR) statistical framework. Numerous probability density functions for the radiogenic neutron background in DEAP-3600 were achieved, resulting in a methodical characterization of this background in the most relevant variables for the WIMP search being currently undertaken by the DEAP collaboration. This characterization supports ongoing efforts to implement the PLR method in the WIMP search. A machine learning approach is used to explore an alternative method for the mitigation of radiogenic neutrons. This method exploits the substantial difference in mean free paths between WIMPs and neutrons in argon, leading to a much higher probability of multiple scatters for neutrons than for WIMPs. Neutron simulations were conducted in which the trajectory of the particle and its interactions provided information on the number of nuclear recoils. It was found that approximately 80% of neutrons that scatter off an argon nucleus undergo multiple scatterings. Several binary classification algorithms were trained with the simulated data and evalu- ated using metrics such as the confusion matrix and receiver operating characteristic (ROC) curve. The best performing algorithm was a boosted decision tree (BDT), which has been used previously by other particle physics experiments. The BDT was tested with two simu- lated datasets, producing encouraging results with validation and testing accuracy scores of 77% and 72% respectively. Additional notable evaluations include precision scores of 78% and 90%, and area under the curve (AUC) scores from the ROC curves of 0.85 and 0.80, respectively for the validation and test datasets. An additional test was conducted on physical data gathered by DEAP-3600 when a neutron source was deployed. Using the ratio of multiple to single recoils found in a previous stage of this investigation, an approximate total of 263 multiple recoils were expected for this dataset, the BDT was able to classify 104+14 −12 of the events as multiple recoils. While lower than the expected number of multiple recoils, the ability to filter a significant fraction of events remains valuable for mitigating radiogenic neutrons as a background source. Although the BDT did not reach the evaluation performance seen at production level in other experiments, the results demonstrate a promising starting point for reliable multiple nuclear recoil identification. The classification notably outperformed random guessing, and tests in a realistic application scenario showed successful identification of a significant frac- tion of expected multiple recoil events. Future improvements could come from integrating algorithms that incorporate spatial information from the detector. This line of research could benefit various WIMP search experiments that must address radiogenic neutrons as a 123 124 CHAPTER 8. CONCLUSIONS substantial background source. This work also presents a contribution in the background model for the search of solar axions, also referred to as axion-like particles (ALPs) in DEAP-3600. The results achieved in this work have significantly improved the background model, as evidenced by the change in p-value from 0.5±0.3% to 25±1% after incorporating the Mo and Co isotopes. This p-value indicates that, assuming the background-only hypothesis, there is a 25±1% probability of obtaining results at least as extreme as those observed. The solar axion search is an ongoing effort being undertaken by the DEAP collaboration. The results of the CEνNS measurements reported by the CONUS+ experiment were replicated using the defining parameters of CONUS+. This resulted in 330 events, which is consistent with the observed data. This analysis was used to study SM and BSM parameters using a χ2 optimization approach, where systematic uncertainties were incorporated as nui- sance parameters. This study was published in the journal Physical Review D, achieving to be the first analysis of the CONUS+ data to be accepted for publication in a peer-reviewed journal. The first parameter to be studied was the weak mixing angle (θW ). The best fit value obtained from this analysis is sin2 θW = 0.268 ± 0.047, which is consistent with measure- ments from other experiments within the 1σ interval. This analysis produced the first limits on sin2 θW via the CEνNS process for reactor antineutrinos. The following parameter to be probed was the magnetic moment of the neutrino (µν). This study produced a 90% CL. upper limit of µν ≤ 5.6 × 10−10µB. This analysis sets a more stringent limit on µν than COHERENT and CEνNS-based DM experiments, and is competitive with LSND re- sults. However, more stringent constraints have been obtained by TEXONO, GEMMA, and CONUS using reactor antineutrinos, with the strongest limits coming from direct DM detectors via electron-neutrino scattering. The last scenario to be investigated was the BSM framework of the Non-Standard Inter- actions (NSI). In this investigation, a 95% CL. contour was achieved in the ϵuVee -ϵ dV ee plane, which is comparable to that produced by the COHERENT experiment. The results from the CONUS+ study demonstrate that CEνNS offers a valuable framework for probing electroweak parameters such as θW and µν . It also provides meaningful insights into BSM scenarios, including those involving NSIs. These results can be further improved by reducing both statistical and systematic uncertainties in the measurements. Additionally, a complementary study of CEνNS using simulated SNe neutrinos was con- ducted. CEνNS signals from a 15 M⊙ core-collapse SN at various distances were simulated using the EstrellaNueva software, targeting the noble liquid dark matter detectors DEAP- 3600 and LZ. The signal observed by each detector was calculated considering their energy resolutions and efficiencies. LZ is capable of detecting CEνNS from simulated SNe occurring across nearly the entire galactic disk, whereas DEAP-3600 is limited to nearby events at dis- tances comparable to that of Betelgeuse. Differences in the CEνNS differential cross section between LAr and LXe, along with the target mass, play a significant role in the capability attainable comparing both detectors. Upper and lower event bounds were established using a χ2 method inspired by the CONUS+ study to probe θW , µν , and NSI parameters, with astrophysical and detector systematic un- certainties treated as nuisance parameters. The study of θW yielded results for all distances considered with LZ, where uncertainties increased with distance; and for DEAP-3600 at the 125 closest distance. These results are comparable to those obtained in the CONUS+ study and similar experiments. The exploration of µν produced conservative limits for both detectors at 0.196 kpc, with the sensitivity decreasing for LZ at larger distances. The sensitivities projected for θW , µν are comparable to those from reactor, spallation, and solar neutrino experiments. The NSI analysis faces additional limitations due to the inability to resolve flavor-specific events, the dependence on the assumed mass hierarchy, and reduced statistics for distant SNe. However, a methodology was proposed to study the NSI parameters for all neutrino flavors under both mass orderings by approximating the flavor composition of the SN burst using simulation-based flavor ratios. As a result, exclusion contours similar to those of CONUS+ were obtained at the closest distance evaluated. 126 CHAPTER 8. CONCLUSIONS A Acronyms and Abbreviations Used List of the acronyms and abbreviations most commonly used in this work. ALP: axion-like particle. AmBe: Americium-Beryllium, neutron source. AUC: area under the curve, classification metric. AV: acrylic vessel. BDT: boosted decision tree, classification algorithm. BRR: background rejection rate, classification metric. BSM: beyond Standard Model. CEνNS: coherent elastic neutrino-nucleus scattering. CL: confidence level. CMB: cosmic microwave background. CsI: cesium iodide. DM: dark matter. ER: electron recoil scattering interaction. FN: false negative, evaluation metric. FP: false positive, evaluation metric. FPR: false positive rate, evaluation metric. F1: F1 score, evaluation metric. 127 128 APPENDIX A. ACRONYMS AND ABBREVIATIONS USED GAr: gaseous argon. HPGe: high purity germanium, material used in the crystals of the CONUS+ experiment. IH: inverted hierarchy neutrino mass ordering. LAr: liquid argon. LXe: liquid xenon. LZ: LUX-ZEPLIN, a liquid xenon dark matter detector. MBR: mblikelihoodR, variable for analysis in DEAP-3600, represents the reconstructed ra- dial position of events. MC: Monte Carlo, statistical analysis methods based on simulations. MR: multiple nuclear recoil. NH: normal hierarchy neutrino mass ordering. NR: nuclear recoil scattering interaction. NSC: nSCBayes, a variable that measures of the energy captured from the interactions in DEAP-3600. NSI: non-standard interactions. PDF: probability density function. PE: number of registered photoelectrons. PIFGAR: Pulse Index First Gaseous Argon: a variable of analysis in DEAP-3600 that quan- tifies how many PMTs captured a signal in the liquid argon before any PMTs gathered scintillation light in the gaseous argon region. PLR: profile likelihood ratio. PMT: photo-multiplier tubes. PSD: pulse shape discrimination. qPE: quenched photo-electrons (pulse charge divided by the average single photo-electron charge for each PMT). 129 RAT: Reactor Analysis Tool, simulation and analysis software. ROC: receiver operating characteristic, a graph used as a classification metric. ROI: region of interest. RPR: Fprompt (Rprompt in newer analysis). Variable used for pulse shape discrimination in the DEAP collaboration. SM: Standard Model of particle physics. SN, SNe: supernova, supernovae. SR: single nuclear recoil. TN: true negative, evaluation metric. TP: true positive, evaluation metric. TPB: tetraphenyl-butadiene, wavelength shifter. TPR: true positive rate, evaluation metric WIMP: weakly interacting massive particle. 130 APPENDIX A. ACRONYMS AND ABBREVIATIONS USED B Supplementary Plots for the CEνNS Studies Compendium of relevant plots for the CEνNS analyzes in Chapter 7. Figure B.1: Squared Helm form factors for germanium, as a function of the recoil energy T . 131 132 APPENDIX B. SUPPLEMENTARY PLOTS FOR THE CEνNS STUDIES Figure B.2: Event rates for each individual detector in CONUS+. Conus-2 and 3 are in the top left and right, and Conus-5 and the total event rate in the bottom left and right respectively. 133 Figure B.3: Helm form factors for Xe and Ar, used in the calculation of the number of events for LZ and DEAP-3600. Figure B.4: DEAP-3600 event rates at 10 and 20 kpc. 134 APPENDIX B. SUPPLEMENTARY PLOTS FOR THE CEνNS STUDIES Figure B.5: NSI parameter 95% CL. contours for DEAP-3600 (green) and LZ (purple) at 0.196 kpc. The parameter flavors are identified with distinctive patterns as follows: horizontal lines for e, diagonal lines for µ, and dotted for τ . The NH in the neutrino mass ordering is assumed. L OO LOO 0.75 0.75 0.50 0.50 0.25 0.25 :::'ClJ ::JClJ 0.00 :::'ClJ ::JClJ 0 .00 LV LV -0.25 -0.25 -0.50 -0.50 - 0.75 - 0.75 95% Cl. DEAP-3600 0.196 kpc (N H) - 95% e l. l Z 0.196 kpc (NH) - LOO - LOO -LOO -0.75 -0 .50 -0.25 0.00 0 .25 0.50 0 .75 LOO -LOO -0.75 -0.50 -0.25 0 .00 0.25 0.50 0.75 LOO EdV ee EdV ee LOO LOO 0.75 0.75 0.50 0.50 0.25 0.25 :::'::t :::'::t ::J::t 0.00 ::J::t 0 .00 LV LV -0.25 -0.25 - 0 .50 - 0.50 - 0 .75 - 0 .75 95% el. DEAP-3600 0 .196 kpc (NH) -95% el. l Z 0.196 kpc (NH) -LOO - 1.00 - LOO - 0 .75 - 0 .50 - 0 .25 0.00 0.25 0.50 0 .75 LOO - LOO - 0 .75 - 0.50 - 0.25 0.00 0 .25 0.50 0.75 LOO EdV J1I-l EdV J.1J.1 LOO LOO 0.75 0.75 0.50 0.50 0.25 0.25 :::. .... :::. .... ::J .... 0 .00 ::J .... 0.00 LV LV -0.25 -0.25 -0.50 -0.50 -0 .75 -0.75 95% Cl. DEAP-3600 0.196 kpc (NH) -95% el. LZ 0.196 kpc (NH) - LOO - LOO - l.00 - 0.75 - 0.50 - 0.25 0.00 0.25 0.50 0.75 LOO - l.00 - 0 .75 - 0.50 - 0.25 0.00 0.25 0.50 0.75 l.00 EdV TT EdV TT 135 Figure B.6: NSI parameter 95% CL. contours for DEAP-3600 (green) and LZ (purple) at 0.196 kpc. The parameter flavors are identified with distinctive patterns as follows: horizontal lines for e, diagonal lines for µ, and dotted for τ . The IH in the neutrino mass ordering is assumed. L OO LOO 0.75 0.75 0.50 0.50 0.25 0.25 :::'ClJ ::JClJ 0.00 :::'ClJ ::JClJ 0 .00 LV LV -0.25 -0.25 -0.50 -0.50 - 0.75 - 0.75 95% Cl. DEAP-3600 0.196 kpc (lH) - 95% e l. l Z 0.196 kpc (IH) - LOO - LOO -LOO -0.75 -0 .50 -0.25 0.00 0 .25 0.50 0 .75 LOO -LOO -0.75 -0.50 -0.25 0 .00 0.25 0.50 0.75 LOO EdV ee EdV ee LOO LOO 0.75 0.75 0.50 0.50 0.25 0.25 :::'::t :::'::t ::J::t 0.00 ::J::t 0 .00 LV LV -0.25 -0.25 - 0 .50 - 0.50 - 0 .75 - 0 .75 95% el. DEAP-3600 0 .196 kpc (lH) -95% el. l Z 0.196 kpc (lH) -LOO - 1.00 - LOO - 0 .75 - 0 .50 - 0 .25 0.00 0.25 0.50 0 .75 LOO - LOO - 0 .75 - 0.50 - 0.25 0.00 0 .25 0.50 0.75 LOO EdV J1I-l EdV J.1J.1 LOO LOO 0.75 0.75 0.50 0.50 0.25 0.25 :::. .... :::. .... ::J .... 0 .00 ::J .... 0.00 LV LV -0.25 -0.25 -0.50 -0.50 -0 .75 -0.75 95% Cl. DEAP-3600 0.196 kpc (lH) -95% el. LZ 0.196 kpc (IH) - LOO - LOO - l.00 - 0.75 - 0.50 - 0.25 0.00 0.25 0.50 0.75 LOO - l.00 - 0 .75 - 0.50 - 0.25 0.00 0.25 0.50 0.75 l.00 EdV TT EdV TT 0.75 0.50 0.25 —0.25 0.50 0.75 -1.00 1) DEAP-3600 NH 1 DEAP-3600 IH -1.00 0.75 0.50 0.25 0.00 0.25 0.50 0.75 1.00 dv Ese 0.75 0.50 0.25 —0.25 0.50 0.75 y -1.00 0.75 0.50 NT 0.25 0.00 0.25 0.50 0.75 1.00 dv Ese 136 APPENDIX B. SUPPLEMENTARY PLOTS FOR THE CEνNS STUDIES Figure B.7: 95% CL NSI parameter contours, NH (blue) vs IH (red), for DEAP-3600 (top) and LZ (bottom) at 0.196 kpc. ~