ABSTRACT Title of dissertation: SUPERCONDUCTING NANOWIRE SINGLE-PHOTON DETECTORS FOR DARK MATTER DETECTION APPLICATIONS Jamie Shawn Luskin, Doctor of Philosophy, 2024 Dissertation directed by: Dr. Matthew Shaw NASA Jet Propulsion Laboratory Superconducting Nanowire Single Photon Detectors (SNSPDs) are a leading detector technology for time-correlated single-photon counting from the UV to the near-infrared. Due to their unique combination of low energy thresholds and low intrinsic dark count rates, SNSPDs have become attractive as sensors for emerging low-mass dark matter (DM) detection experiments, where they offer the potential to fill existing technology gaps and enable the exploration of previously unconstrained parameter space. One developing DM detection concept, sensitive to MeV-scale DM electron recoils, uses n-type GaAs as a cryo- genic scintillating target instrumented with a large-area SNSPD as the sensor for scintilla- tion photons. This thesis focuses on the development and characterization of SNSPDs with mm2- scale active areas for DM applications in general, and specifically for the detection of scin- tillation light. This work demonstrates the coupling of n-type GaAs with SNSPDs, and the design of novel characterization experiments using optical and energy-tagged X-ray exci- tation to measure the effective light yield and photoluminescence timescales of cryogenic scintillators using SNSPDs. The isotropic nature of scintillation light and the large active areas of the devices studied in this work introduce unique challenges for SNSPD design, nanofabrication, and performance. This thesis provides insights into the current state-of- the art, limitations, and approaches to scaling SNSPDs to cm2-scale active areas for future work. The presented findings advance the status of SNSPDs for DM detection and other emerging High-Energy Physics applications. SUPERCONDUCTING NANOWIRE SINGLE-PHOTON DETECTORS FOR DARK MATTER DETECTION APPLICATIONS by Jamie Shawn Luskin Dissertation submitted to the Faculty of the Graduate School of the University of Maryland, College Park in partial fulfillment of the requirements for the degree of Doctor of Philosophy 2024 Advisory Committee: Professor Yanne Chembo, Chair Dr. Matthew Shaw, Advisor Professor Daniel Lathrop, Dean’s Representative Professor Jeanpierre Paglione Professor Carlos A. Rı́os Ocampo © Copyright by Jamie Shawn Luskin 2024 What we observe is not nature itself, but nature exposed to our method of questioning. - Werner Heisenberg ii DEDICATION This thesis is dedicated to the memory of my grandfather Dr. Jerome Lawrence Luskin, The original Dr. Luskin and my greatest inspiration. iii Acknowledgments As I approach the end of my graduate career, I can’t help but feel overwhelmed with gratitude for the people who have helped make this journey possible, both directly and indirectly. It wouldn’t be right to start this section off any other way than to acknowledge my JPL advisor, Matt Shaw, for offering me the life-changing opportunity to work in our group. I can’t thank you enough for your guidance, support, and kindness throughout the past 4.5 years. I also want to express my sincere appreciation to my UMD advisor, Yanne Chembo, for going above and beyond in providing invaluable guidance throughout this journey. And to the one and only Professor Maria Spiropulu - thank you for making me feel at home at Caltech and the INQNET labs, and for treating me like one of your own. My sincerest gratitude extends to the brilliant JPL SNSPD group members. Andrew Mueller - thank you for being an exceptional office-mate, friend, and colleague and for listening to all of my rants (and taking them seriously no matter how inane or dramatic!). I am constantly amazed by your talents and inspired by the way you think. Ioana Craiciu - you are truly the best, and I appreciate all of your guidance and support during these last few chaotic months more than I can ever express. Professor Boris Korzh - you have had such a profound impact on the way I approach measurements and science. Jason Allmaras - thank you for being an incredible resource any time, any place, no matter what goes wrong. I deeply admire and look up to your technical skills, work ethic, and the way you reason through problems. To the rest of the group- Emma Wollman, Sahil Patel, Emanuel Knehr, Sasha Sypkens, Bruce Bumble, Andrew Beyer, Fiona Fleming, Jasen Zion and former members Ekkehart Schmidt and Gregor Taylor - your contributions and support have been instrumental to my professional development and it is a pleasure to work with you all. A sincere thank you goes out to my collaborators at Fermilab Cristian Peña, Si Xie, and Christina Wang for guiding many of the experiments in this thesis, and at LBNL, Maurice Garcia-Sciveres and Steve Derenzo for the insightful and lively discussions. iv Dr. J - I don’t know that I will ever find the words to convey the impact that you and your class have had on me, but I would not have ended up on this trajectory without you and your philosophy will always be my foundation. The next (long-overdue) pitcher is on me! Keric Morinaga - thank you for sharing your love of yoga with me and for building the most incredible and supportive community at the Touchstone gyms. Your lessons and philosophy have played a pivotal role in keeping me mentally grounded throughout this process. A pleasant surprise during this PhD was getting to meet so many wonderful people at Caltech. I am sincerely thankful to Sam Davis, Iraj Umesh, Preeti Bhat, and all of the INQNET group members for welcoming me and supporting me. Tracy Sheffer - you leave us all feeling so taken care of, and we simply could not do what we do without you! Rahaf Youssef - I immensely appreciate your intellect, brilliant insights, kindness, and friendship throughout the past few years. Claire Ellison, Tommy Sievert, Ray Wynne, and Damian Wilson - thank you for being the infinite source of joy that you are for everyone around you. Lautaro Narvaez - thank you for patiently helping me with anything I’ve ever brought to you. There is truly no problem you can’t solve and your expertise in electronics is truly wizard-level. Thank you to the Hutzler group members for being the best neighbors in Downs-Lauritsen and delightful humans to interact with. A tremendous thank you goes out to the illustrious and incomparable Jason Trevor. Thank you for everything that you do to keep so many groups running and for teaching me all things high-energy physics and beyond. You are an absolute legend, one of my favorite people, and the closest thing to a superhero I’ve ever met. I would like to thank my family for their support throughout my entire (long!) educa- tion. I am so fortunate to have all of you in my corner. To my late grandfathers, Jerome Luskin and Nunzio Brusca - you are profoundly missed and will always be a guiding light v for me. I am also eternally grateful for the incredible Leal family for their love and support during those early years of studying like a maniac. A posthumous thank you is extended to my late great-uncle, Dr. Burton Kleinman, who has always been an immense inspiration to me. I deeply admire how your work transcended the boundaries between mathematics, physics, and chemistry (and I hope I inherited some of those Kleinman genes!). There is a near-unity probability that we had a lot in common and I’ll always wish that we could chat about things. Finally, to my incredible wife, Elaina - I could not have done any of this without your unconditional support. Thank you for believing in me, helping me grow, and for being the best partner in the universe. vi Table of Contents Foreword ii Dedication iii ACKNOWLEDGEMENTS iv 1 Dark Matter 1 1.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1 1.2 Evidence for Dark Matter . . . . . . . . . . . . . . . . . . . . . . . . . . . 2 1.2.1 Galactic Rotation Curves . . . . . . . . . . . . . . . . . . . . . . . 2 1.2.2 Cosmological scale . . . . . . . . . . . . . . . . . . . . . . . . . . 3 1.2.3 Gravitational Evidence . . . . . . . . . . . . . . . . . . . . . . . . 6 1.3 Dark Matter Candidates . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9 1.3.1 Classes of Dark Matter Candidates . . . . . . . . . . . . . . . . . . 10 1.3.2 WIMP Dark Matter . . . . . . . . . . . . . . . . . . . . . . . . . . 11 1.3.3 Axions and Axion-Like Particles (ALPs) . . . . . . . . . . . . . . . 13 1.3.4 Dark Photons . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16 1.3.5 sub-GeV Dark Matter . . . . . . . . . . . . . . . . . . . . . . . . . 17 DM-electron scattering . . . . . . . . . . . . . . . . . . . . . . . . 18 DM-electron absorption . . . . . . . . . . . . . . . . . . . . . . . . 19 1.4 DM Detection Strategies . . . . . . . . . . . . . . . . . . . . . . . . . . . 21 1.4.1 Direct Detection of sub-GeV Dark Matter . . . . . . . . . . . . . . 22 1.4.2 n-type GaAs as a Cryogenic Scintillating Target . . . . . . . . . . . 23 1.4.3 DM Detection concept using n-type GaAs . . . . . . . . . . . . . . 26 1.4.4 Sensor Classes for n-type GaAs and Low Mass DM Targets . . . . . 28 2 Superconducting Nanowire Single Photon Detectors 32 2.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32 2.2 Device Operation Principles . . . . . . . . . . . . . . . . . . . . . . . . . 33 2.3 Detection Metrics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35 2.3.1 Detection Efficiency . . . . . . . . . . . . . . . . . . . . . . . . . 36 vii 2.3.2 Dark Count Rate . . . . . . . . . . . . . . . . . . . . . . . . . . . 37 2.3.3 Timing Resolution . . . . . . . . . . . . . . . . . . . . . . . . . . 38 2.4 Design Tradeoffs for Large-area SNSPDs . . . . . . . . . . . . . . . . . . 39 2.4.1 Active Area vs Timing Performance . . . . . . . . . . . . . . . . . 39 2.4.2 Array Size vs Readout Complexity . . . . . . . . . . . . . . . . . . 40 2.5 Large-active area SNSPDs . . . . . . . . . . . . . . . . . . . . . . . . . . 40 2.5.1 Materials . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41 2.5.2 Device architecture . . . . . . . . . . . . . . . . . . . . . . . . . . 42 Nanowire vs Microwire Detectors . . . . . . . . . . . . . . . . . . 43 2.6 SNSPDs for Dark Matter Detection and HEP Applications . . . . . . . . . 43 3 Large Active Area SNSPD Design and Characterization 46 3.1 Background . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46 3.2 Generation 1 Microwire Arrays . . . . . . . . . . . . . . . . . . . . . . . . 49 3.2.1 Detector Design . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49 3.2.2 Materials . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49 3.2.3 Detector Characterization . . . . . . . . . . . . . . . . . . . . . . . 50 Cryogenic set-up and Readout . . . . . . . . . . . . . . . . . . . . 50 Optical Characterization . . . . . . . . . . . . . . . . . . . . . . . 51 Efficiency Measurements . . . . . . . . . . . . . . . . . . . . . . . 55 Tc Measurement . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57 Timing Resolution Characterization . . . . . . . . . . . . . . . . . 59 3.3 Generation 2 Microwire Arrays . . . . . . . . . . . . . . . . . . . . . . . . 62 3.3.1 Detector Design . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65 3.3.2 Tc Measurement . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67 3.3.3 Optical Characterization . . . . . . . . . . . . . . . . . . . . . . . 68 3.3.4 Timing Resolution Characterization . . . . . . . . . . . . . . . . . 70 4 Backgrounds in Microwire Arrays 77 4.1 Extrinsic Background Mitigation . . . . . . . . . . . . . . . . . . . . . . . 78 4.1.1 Filtering . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78 Dark Packaging . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78 4.2 Intrinsic Backgrounds in Microwire Arrays . . . . . . . . . . . . . . . . . 81 4.3 Scaling with Active Area . . . . . . . . . . . . . . . . . . . . . . . . . . . 81 4.4 Temperature Dependence of Dark Counts . . . . . . . . . . . . . . . . . . 83 5 Optically-excited Scintillation Measurements 87 5.1 Overview of systems measured . . . . . . . . . . . . . . . . . . . . . . . . 88 5.1.1 GaAs Specifications and Analysis . . . . . . . . . . . . . . . . . . 89 5.2 Detector and Sample Information . . . . . . . . . . . . . . . . . . . . . . . 89 5.2.1 System A . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 89 Device packaging and Hybridization . . . . . . . . . . . . . . . . . 90 5.2.2 System B . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90 viii Device packaging and Hybridization . . . . . . . . . . . . . . . . . 90 5.3 Calibration of SNSPDs for Scintillation Light . . . . . . . . . . . . . . . . 91 5.4 Scintillation timescale measurements . . . . . . . . . . . . . . . . . . . . . 94 5.4.1 Measurement setup . . . . . . . . . . . . . . . . . . . . . . . . . . 96 5.4.2 Results and exponential fitting . . . . . . . . . . . . . . . . . . . . 97 5.5 Effective Light-yield Measurements . . . . . . . . . . . . . . . . . . . . . 98 5.5.1 Measurement Setups . . . . . . . . . . . . . . . . . . . . . . . . . 99 5.5.2 System A effective light yield . . . . . . . . . . . . . . . . . . . . 100 5.5.3 System B Effective Light Yield . . . . . . . . . . . . . . . . . . . . 100 5.6 Future work towards Absolute Light Yield Measurements . . . . . . . . . . 101 6 X-ray Excited Scintillation Measurements 103 6.1 Compton Scattering . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 105 6.2 X-ray Source . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 107 6.3 Calibration of HPGe Detectors . . . . . . . . . . . . . . . . . . . . . . . . 108 6.4 Timescale Measurements . . . . . . . . . . . . . . . . . . . . . . . . . . . 111 6.4.1 Background Measurements of GaAs-SNSPD System . . . . . . . . 113 6.4.2 CAEN Digitizer Readout . . . . . . . . . . . . . . . . . . . . . . . 114 6.4.3 Source Effect on GaAs-SNSPD System . . . . . . . . . . . . . . . 115 6.5 Compton Scattering Analysis and Challenges . . . . . . . . . . . . . . . . 119 6.5.1 Poisson Statistics and Mean Photon Number Fitting . . . . . . . . . 122 6.5.2 Understanding the Observed Distribution . . . . . . . . . . . . . . 124 6.5.3 Understanding System Losses and Secondary X-ray Generation . . 128 7 Designs and Challenges of Improved Microwire Arrays 132 7.1 Optical Stack Design for Microwire Arrays . . . . . . . . . . . . . . . . . 133 7.1.1 Rigorous Coupled Wave Analysis . . . . . . . . . . . . . . . . . . 136 7.1.2 Angled-Incidence RCWA Simulations . . . . . . . . . . . . . . . . 136 7.1.3 DBR Optical Stack Design for Broad-Angle Efficiency . . . . . . . 138 Angular Distribution of Light from GaAs . . . . . . . . . . . . . . 140 DBR Design . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 148 7.2 Fabrication Challenges and Steps Forward . . . . . . . . . . . . . . . . . . 149 8 Microwire Arrays in Other Experiments 155 8.1 Detection of High Energy Particles with Large Area SNSPDs . . . . . . . . 155 8.1.1 Detector Design and Measurement Infrastructure . . . . . . . . . . 156 8.1.2 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 156 8.2 Broadband Reflector Experiment for Axion Detection . . . . . . . . . . . . 158 8.2.1 Detection Concept and Experiment Design . . . . . . . . . . . . . 158 8.2.2 RCWA Efficiency Simulations . . . . . . . . . . . . . . . . . . . . 160 8.2.3 Future Work . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 162 9 Conclusions 164 ix 9.1 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 164 9.2 Future Work and Outlook . . . . . . . . . . . . . . . . . . . . . . . . . . . 165 A Procedure for IV Curve Analysis with DC Coupled Cryogenic Amplifiers 166 B n-type GaAs Sample Information 171 B.1 Sample A - More Information . . . . . . . . . . . . . . . . . . . . . . . . . 171 B.2 Sample B - More Information . . . . . . . . . . . . . . . . . . . . . . . . . 172 C Publications 174 Bibliography 175 x Chapter 1: Dark Matter 1.1 Introduction Elucidating the particle nature of dark matter (DM) is one of the largest unsolved prob- lems in fundamental physics. Though DM has evaded direct detection to date, indirect probes of its gravitational interactions with celestial bodies at multiple scales indicate that it comprises nearly 80% of the matter in the Universe and is necessary to explain large-scale structure formation. The most mature and large-scale experimental DM search programs have focused on the Weakly Interacting Massive Particle (WIMP) paradigm. Due to a lack of conclusive de- tection signal, new theoretical frameworks have emerged in recent years, and there is now strong motivation to explore a broader set of candidate particles experimentally. Hidden sector DM candidates are neutral under Standard Model (SM) forces and charged under new forces that can weakly couple to ordinary matter via portal interactions. The range of potential mass of hidden sector DM is broader than that of the WIMP DM. Importantly, the mass range below a GeV, the so-called sub-GeV DM contains theoretically well-motivated targets that are either below the noise floor or not accessible in WIMP search experimental programs and thus require the development of novel experimental approaches. Increasing the sensitivity of direct detection experiments to lower masses of DM requires the develop- ment of new target materials sensitive to proposed interactions and new low-noise sensors with low-energy thresholds. As these technologies develop, sub-GeV DM can be explored 1 substantially with smaller-scale table-top direct detection pilot experiments. This chapter serves as a brief introduction to DM. Section 1.2 briefly summarizes the experimental evidence for DM, Section 1.3 presents an overview of relevant DM candi- date particles, and Section 1.4 introduces DM detection strategies to motivate the work contained in this PhD thesis. 1.2 Evidence for Dark Matter Evidence for the existence of DM has steadily increased over the last several decades. Observational support includes its gravitational effects, such as galaxy rotation curves, gravitational lensing, and the large-scale structure of the universe, all of which cannot be accounted for by visible matter alone. Further evidence comes from cosmological observa- tions consistent with predictions from the Lambda cold dark matter model (ΛCDM), which serves as the standard model of Big Bang cosmology. 1.2.1 Galactic Rotation Curves The earliest evidence for a cold, non-luminous matter pervading the Universe came from the study of galactic rotation curves. These were first measured in the Coma Cluster by Franz Zwicky in the 1930s [1] and later in the 1970s by V. Rubin [2] in the Andromeda galaxy and others [3]. Obtaining a rotation curve entails measuring the rotational velocity of galaxy or galaxy clusters via Doppler shifts and subsequently plotting this quantity with respect to the distance from the center of the galaxy. The rotational velocity of a galaxy can be expressed using Newtonian mechanics as νrot(r) = √ GM(r) r , (1.1) 2 where M is the enclosed mass, r is the radial distance from the galactic center, and G is the gravitational constant. Assuming that visible matter comprises the entirety of the mass distribution, νrot should decrease as 1√ r2 for distances beyond the visible galactic disc. In practice, the measured vrot curves flatten at large r (see figure 1.1), implying a constant mass distribution that can’t be explained by visible matter alone. Measured galactic rotation curves are all consistent with a spherically symmetric mass distribution that increases linearly with r [4]. On the contrary, the visible mass, consti- tuted by stars and gases which can absorb and emit light, is concentrated in the inner disk. Zwicky, in addition to pioneering these early measurement techniques, coined the term ”dunkle materie” (dark matter) to refer to the non-luminous component of the mass distri- bution. Other proposed explanations for this discrepancy include modifications to gravity in- cluding Modified Newtonian Gravity [5]; however, these lack consistency with phenomena at other scales [6] that are explained in the following sections. 1.2.2 Cosmological scale The Cosmic Microwave Background (CMB) is the background radiation in the Uni- verse comprised of the photons that first decoupled from matter shortly after the Big Bang. It was initially believed to be isotropic until faint (µK-scale) anisotropies in temperature and polarization were discovered in the early 1990s. The origins of these anisotropies are understood to be from the matter-energy density of the early universe. Namely, it is hypothesized that oscillations in the baryon-photon fluid under the influence of the gravita- tional potential due to local wells and peaks of cold dark matter produced the temperature anisotropies, which now manifest in the CMB power spectrum as acoustic peaks [9]. The CMB has a blackbody spectrum with a temperature of 2.73 K, peaked at λ = 3 Figure 1.1: Rotation curve measured for the galaxy NGC 6503 (black dots) and fit to DM-halo model shown with the solid black line. The dashed black line is the Keplerian prediction based on visible matter. The dotted-dash is the DM-halo rotation curve that is required to explain the flatness of the curve at large r. Figure adapted from ref. [7] 4 Figure 1.2: The CMB map as seen by NASA’s COBE (top)and WMAP (middle) missions and ESA’s Planck mission (bottom). Figure adapted from [8]. 5 1.06mm. Atmospheric water molecules absorb this wavelength of light, making it amenable to study via ground and space-based missions (see figure 1.2 for examples). Missions such as COBE [10], WMAP [11], and Planck [12] have directly measured the tempera- ture anisotropies. The power spectrum of these temperature fluctuations offers a means to validate cosmological models that make predictions about this spectrum. Lambda-Cold Dark Matter (ΛCDM), a mathematical model of the Big Bang serving as the cosmological standard model, contains three main components describing the baryonic matter, dark matter, and dark energy contents of the Universe. As shown in figure 1.3, the ΛCDM model prediction fits the CMB anisotropy data with high statistical significance. This alignment reinforces the robustness and accuracy of the ΛCDM model and the impor- tance of dark matter in our understanding of the Universe and its evolution. The baryonic and non-baryonic matter densities extracted from the 2018 Planck col- laboration are Ωbh2 = 0.0224 and ΩDMh2 = 0.120, where h is the dimensionless Hubble constant [12]. The ratio between dark matter and baryonic matter densities is given by ΩDM/Ωb ≈ 0.120/0.0224 ≈ 5.4, indicating that dark matter is approximately five times more abundant than ordinary matter. 1.2.3 Gravitational Evidence Gravitational lensing measurements have also provided a compelling set of evidence for the existence of dark matter. Gravitational lensing is a phenomenon whereby a massive celestial body causes significant perturbations in the space-time metric such that light is bent from a distant source as it travels toward an observer. Most lines of sight that are ac- cessible are in the weak gravitational lensing regime, where distortions can’t be measured from a single source. In composite lensed images, mass distributions can be reconstructed through correlations in the observed ellipticity of the distorted objects. In these measure- 6 Figure 1.3: Temperature power spectrum of CMB measured by Plank mission – red dots , and the λCDM model prediction is shown in blue. Figure taken from ref. [12] ments, the reconstructed mass causing the lensing phenomenon exceeds the visible mass of the object by a factor of tens to hundreds, pointing to the presence of dark matter. A well-known example is the bullet cluster [13], which was formed through the colli- sion of two galaxy clusters roughly 3.8 billion years ago. The Hubble space telescope was used to obtain weak gravitational lensing measurements, from which the mass distribution was derived. The luminous mass distribution was also extracted from X-ray spectra of the cluster measured by the Chandra x-ray observatory. The mass reconstructed from lensing is a factor of 200 more abundant than the luminous mass extracted from the spectroscopic measurements. As shown in figure 1.4, there is a clear offset between the visible center of mass and the center of mass reconstructed from weak gravitational lensing. This off- set arises because dark matter only interacts conventionally through gravity and not via traditional electromagnetic forces. Consequently, during the collision, dark matter passed through largely unaffected, while the visible matter, which interacts electromagnetically, experienced drag and slowed down, leading to the observed separation between the two. 7 Figure 1.4: Composite image taken from ref. [13]of the galaxy cluster 1E 0657- 56, also known as the “Bullet Cluster.” Pink depicts the hot gas measured by Chandra X-ray observatory and includes most of the normal baryonic matter - Blue depicts the concentration of mass determined by weak gravitational lensing 8 1.3 Dark Matter Candidates It has has been conclusively shown that none of the particles in the Standard Model can account for the properties of dark matter, as they either interact too strongly with light or do not have the necessary mass and behavior. DM thus requires extensions beyond the Standard Model of particle physics, which describes the known fundamental particles and forces. This has led to the development of numerous theoretically well-motivated dark matter candidate particles. Constraints on these candidates are typically calculated one model at a time, but it is possible that dark matter is composed of multiple species of particles. This collection of particles is often referred to as the dark sector. The dark sector may involve complex interactions, similar to the visible sector, but these would be extremely weak or invisible to current detection methods, making them difficult to study directly. There are specific properties that dark matter models must satisfy based on obser- vational evidence. First, dark matter does not emit light, meaning any viable candidate must have an extremely weak or vanishing electric charge and electric or magnetic dipole moments. As a result, dark matter cannot radiate energy electromagnetically or ther- malize through photon emission, making it approximately dissipationless. Additionally, dark matter is known to be collisionless—if collisions occurred, dark matter halo profiles would be spherical, but observations show that DM halos are clustered, with prolate shapes in their centers and triaxial structures toward their outskirts [14]. Finally, dark matter is non-relativistic (or ’cold’), as indicated by the formation of large-scale cosmic struc- tures [15, 16]. Dark matter candidates are discussed extensively in reviews [17–20]. The following sections contain first an overview of the general classes of DM candidates, WIMP dark matter for context regarding the previous work in the community, and finally sub-GeV 9 dark matter for its relevance to the context of the work contained in this PhD thesis. 1.3.1 Classes of Dark Matter Candidates The search for DM has expanded beyond Weakly Interacting Massive Particles (WIMPs) and now includes a wide range of alternative candidates. These candidates can be broadly classified into two categories: wavelike (or ultralight) and particle-like dark matter. The classification depends on the de Broglie wavelength, λdB, which is given by: λdB = h p where h is Planck’s constant, and p is the momentum of the particle. This wavelength provides insight into whether the particle behaves more like a classical particle or exhibits wave-like properties. Wave-like behavior occurs when the de Broglie wavelength is com- parable to or larger than the characteristic scale of the system, for example the the average inter-particle separation in a galaxy like the Milky Way. Particle-like behavior is exhibited when the de Broglie wavelength is much smaller than the relevant physical scales, causing the dark matter to behave classically with trajectories similar to standard particles. These two classes differ fundamentally in their mass, behavior, and detection strategies. In general, wavelike dark matter consists of particles with extremely low masses. These candidates include bosonic fields such as scalars or pseudoscalars [21]. Axions are a prominent example, originally introduced to resolve the strong CP problem in quantum chromo-dynamics (QCD) [22]. Their mass range is typically 10−22 eV to 10−5 eV, and they are promising dark matter candidates due to their ability to form a coherent oscillat- ing field. Axions can be detected through haloscope experiments like ADMX [23], where they convert into photons in the presence of strong magnetic fields. Other novel detec- tion techniques involve nuclear magnetic resonance (NMR) [24] and condensed matter 10 systems [25]. Dark Photons are another example of wave-like dark matter, arising from a hidden U(1) gauge symmetry [26, 27]. With masses less than 1 eV, dark photons can oscillate with ordinary photons, leading to observable signatures. Detection methods of- ten involve photon mixing experiments, such as cavity searches or precise atomic clock measurements. Particle-like dark matter candidates arise in extensions of the Standard Model and can range from keV to solar mass scales. Sterile Neutrinos [28] are one such candidate, with masses in the range of keV to MeV. Unlike active neutrinos, sterile neutrinos do not interact via the weak force, making them challenging to detect. Additionally, models proposing a Dark Sector suggest the existence of dark baryons, analogous to protons and neutrons but confined to a dark sector. This dark sector may involve its own set of forces and particles, weakly interacting with Standard Model particles. Detection could arise from, missing energy signatures in collider experiments, dark sector particles decaying into visible matter, or direction detection searches [29–31]. 1.3.2 WIMP Dark Matter Of the possible DM candidates, the (particle-like) Weakly Interacting Massive Particle (WIMP) paradigm has been the most tested experimentally over the past several decades. Broadly, a WIMP is a proposed fundamental particle which interacts via gravity and an- other force(s) that is non-vanishing in strength at the weak-scale or below, with a mass range from a few GeV/c2 to several TeV/c2. The primary WIMP production mechanism is thermal freeze out. In the early hot and dense Universe, WIMPs interacted more frequently with ordinary matter due to the high temperature. As the Universe expanded and cooled, the interactions between WIMPs and other particles became less frequent due to lower particle densities. Eventually the expan- 11 sion rate of the universe exceeded the annihilation rate of the dark matter, resulting in a reduced number of WIMPs and leaving a stable relic density. Through observational evidence from cosmological surveys and insights from particle physics experiments, the relic abundance of DM has been determined with high precision. Any DM theory must account for this precise relic abundance in order to be considered viable. The relic density of WIMP DM can be obtained from the Boltzman equation and expressed as: Ωχ ≈ 3×10−27 cm3 s−1 ⟨σAν⟩ (1.2) where ⟨σAν⟩ is the thermally averaged annihilation cross section for DM and h is the dimensionless Hubble parameter. A particle with electroweak-scale interactions naturally leads to a relic abundance that is consistent with the relic abundance derived from cosmo- logical constraints. This has been referred to as the “WIMP miracle”. WIMP experiments are generally designed to detect nuclear recoils caused by interac- tions between WIMPs and atomic nuclei. The interaction cross-section for WIMPs with nuclei is typically larger than that with electrons. Thus, the recoil energy deposited in a nucleus from a WIMP collision is generally more significant than the energy deposited in an electron collision. Targets designed to probe WIMP interactions include noble liquids such as xenon and argon. The LZ experiment is one of the largest-scale WIMP searches. The characteristic nuclear recoil energy in liquid xenon is on the order of 10 keV, which is comparable to the recoil energy from radiogenic backgrounds. Because of this, one of the primary design considerations in WIMP experiments is the minimization of radiogenic backgrounds. Another challenge facing WIMP experiments is the the intrinsic scaling of the interac- 12 tion rate. The number of WIMP interactions, nDM, is given by the relation: nDM = ρDM MDM (1.3) where ρDM is the dark matter density and MDM is the mass of the dark matter particle. As MDM increases, nDM decreases, making the interaction rate lower. For this reason, WIMP experiments need to have extremely large target volumes to significantly probe the parameter space. For example, the latest iteration of the LZ experiment features a target mass of 7 tons of liquid xenon. 1.3.3 Axions and Axion-Like Particles (ALPs) The concept of axions emerged in the 1970s as a solution to the strong CP problem in quantum chromodynamics (QCD). The strong CP problem refers to the apparent absence of a significant violation of the CP symmetry (combined Charge Conjugation and Parity) in strong interactions, despite the fact that the QCD Lagrangian allows for a potential CP- violating term: LQCD = θ 32π2 GµνG̃µν (1.4) where Gµν is the gluon field strength tensor, and G̃µν is the axion field strength tensor. The parameter θ is expected to be a small number, but its value is not constrained by theory, and large values would induce strong CP-violation effects that are not observed experimentally. In 1977, Peccei and Quinn proposed a solution to this problem by introducing a new symmetry (Peccei-Quinn symmetry) that spontaneously breaks, leading to what became known as the axion. The axion is a light scalar particle that provides a dynamic mechanism 13 to relax the value of the θ -term to near zero, effectively solving the strong CP problem. Axion-like particles (ALPs) are generalizations of the axion. These are hypothetical light scalar particles that do not necessarily arise from the Peccei-Quinn mechanism but share similar properties, such as interactions with photons and other particles. The axion field a(x) can be described by the following Lagrangian density: Laxion = 1 2 ∂µa∂ µa− 1 2 m2 aa2 +gaγγaFµν F̃µν (1.5) Here, ma is the axion mass, Fµν is the electromagnetic field strength tensor, and F̃µν is its dual. The coupling constant gaγγ describes the interaction strength of the axion with two photons. This term allows axions to couple to photons and can lead to observable effects, such as axion-photon conversion in external magnetic fields. Axion-like particles have similar interactions, but their Lagrangian can also include interactions with other particles, depending on their origin: LALP = 1 2 ∂µa∂ µa− 1 2 m2 aa2 +gaγγaFµν F̃µν +gaψψaψ̄ψ (1.6) where ψ represents a fermion field, and the second term introduces the possible inter- action of ALPs with fermions. RF resonant-cavity experiments, such as ADMX and HAYSTAC (Haloscope At Yale Sensitive To Axion CDM), aim to detect axion dark matter by exploiting axion-photon con- version in a strong magnetic field. The axion signal is expected to manifest as a narrow- band electromagnetic wave at a frequency determined by the axion mass, resonantly en- hanced by a tunable microwave cavity. ADMX and HAYSTAC focus on axion masses in the range ∼ 1–40 µeV, corresponding to frequencies in the GHz range. HAYSTAC has probed masses in the region of ∼ 4–20 µeV, complementing ADMX at slightly higher fre- quencies. These experiments contribute to systematically excluding portions of the viable 14 axion parameter space. Resonant-cavity experiments face several limitations that constrain their ability to ex- plore the full axion parameter space. These setups are inherently narrow-band, sensitive to a small range of axion masses at a time. Scanning a broader range requires precise tuning of the cavity frequency, which is time and resource intensive. Additionally, the resonant frequency of a cavity is determined by its dimensions, making it challenging to probe very small axion masses (< 1 µeV) due to the prohibitively large cavity sizes required, or very large axion masses (> 40 µeV), where resonant enhancement becomes inefficient. Noise further limits sensitivity, with thermal noise dominating at lower frequencies and quan- tum noise at higher frequencies, though the HAYSTAC collaboration has made significant progress in overcoming the latter through quantum squeezing techniques. Finally, the weak axion-photon coupling strength (gaγγ ) necessitates long integration times to detect potential signals, adding to the challenges. Despite significant progress by experiments like ADMX [23, 32–36] and HAYSTAC [37, 38], large regions of the axion mass parameter space remain unexplored. The low-mass range (< 1 µeV) and high-mass range (> 40 µeV) are particularly challenging due to the constraints of cavity design and sensitivity. Exploring these regions will require innovative approaches and new technologies, as resonant-cavity designs are fundamentally limited in their scalability. Developing broadband haloscopes is essential to overcome these chal- lenges, enabling simultaneous searches over wide mass ranges and reducing the time re- quired to explore the uncharted axion landscape. The Broadband Reflector Experiment for Axion Detection (BREAD) collaboration introduces an innovative approach to search for axion and dark photon dark matter across a wide range of masses, spanning approximately µeV to eV, without requiring tuning. The experiment utilizes a cylindrical dish resonator designed to fit within solenoidal magnets. In this setup, axions (in the presence of the mag- netic field) or dark photons interact with the metallic surface of the dish, converting into 15 photons independent of their mass. These photons are subsequently directed by a parabolic focusing reflector onto a highly sensitive single-photon counting detector. For the pilot dark photon experiment within the BREAD program, the large-area mi- crowire arrays developed and characterized in this thesis are currently being integrated as the single-photon detector. Chapter 8 describes the characterization and integration of these sensors into the experiment and plans for the future. 1.3.4 Dark Photons Dark photons are hypothetical particles that are introduced as the gauge bosons (force carriers) of a new U(1) symmetry associated with DM interactions. The idea behind dark photons is that they could mediate interactions between dark matter particles and possibly visible matter, but with a very weak coupling to the standard model particles. This idea is motivated by the dark sector in theoretical physics, which posits the existence of a hidden sector of particles that interact very weakly with ordinary matter. In its theoretical framework, a dark photon can mix with the standard photon, leading to the possibility of dark matter-photon interactions. This interaction is particularly interesting for explaining anomalies observed in certain experiments, such as the g-2 anomaly in the muon magnetic moment, and for providing a candidate for dark matter. The Lagrangian for the dark photon can be written as: Lγ ′ =−1 4 FµνFµν − 1 4 F ′ µνF ′µν + ε 2 FµνF ′µν +m2 γ ′A ′ µA′µ (1.7) where Fµν is the field strength tensor for the standard photon, F ′ µν is the field strength tensor for the dark photon, and ε is a small parameter that represents the mixing between the standard photon and the dark photon. The term m2 γ ′ corresponds to the mass of the dark photon, and the mixing term enables the possibility for photon-dark photon oscillations. 16 The interaction between the dark photon and standard model particles is typically mod- eled as a kinetic mixing term: Lint = εFµνF ′µν (1.8) This term leads to weak couplings between ordinary matter (via the electromagnetic interaction) and dark matter (via the dark photon). 1.3.5 sub-GeV Dark Matter The absence of a DM signal in existing WIMP direct detection experiments and the lack of supporting non-Standard Model physics in collider experiments have led to interest in the development of other candidates for DM. The lack of sizable interaction between DM and ordinary matter motivate the hypothesis that dark matter is neutral under all standard model forces but interacts through new forces that to-date have been undiscovered. This is referred to as hidden sector DM. In several frameworks, the sub-GeV scale arises naturally and via production mechanisms beyond the standard freeze-out. Each mechanism implies targets in the DM parameter space that are accessible experimentally. The DM relic density can be accounted for by mechanisms that describe portal interactions between hidden sector forces and ordinary, for example via an exchange of a dark photon. Importantly, the parameter space for these models is unexplored experimentally and also, in part, beyond the scope of astrophysical bounds. The higher mass region of the hidden sector parameter space is well-probed by WIMP search experiments (few GeV to 100 TeV). The lower region of the hidden sector DM parameter space below a few GeV is not well-probed by WIMP search experiments and requires new detection strategies. This is referred to as sub-GeV dark matter. One of the challenges in probing light dark matter via traditional nuclear recoil methods 17 lies in the kinematics; for lighter masses of DM, the total kinetic energy transferred to the nucleus during a recoil is suppressed. Low nuclear recoil energies are a challenge to detect and necessitate sensors with drastically reduced energy thresholds and background rates. Another avenue to search for DM is through interaction with electrons in target ma- terials as opposed to nuclei. This is more kinematically favorable, as a higher fraction of kinetic energy can be transferred from a light DM particle to an electron via inelastic scattering processes, for example. Due to the inverse relationship between interaction rate and the DM mass shown in Equation 1.3, interactions can be probed with smaller target volumes and exposure for lighter dark matter masses. For this reason, pilot programs for sub-GeV DM searches are able to utilize table-top sized experiments. The measurable signatures of hidden sector DM are dictated by the type of new force coupling the DM to ordinary matter. As are described in benchmark models [39], a new force can be mediated by scalar of vector bosons which couple to leptons or hadrons. The bench-marked coupling models are described in detail in this reference for hidden sector DM in general, and also the low-mass (sub-GeV) parameter space for hidden sector dark matter . For concreteness, sub-GeV DM that interacts with electrons through a dark-photon mediator is described in the following section. DM-electron scattering Using the notation in reference [40], a DM particle χ can couple to the standard model via a U(1) gauge boson A ′ (dark photon). A′ kinetically mixes with the SM U(1) field via the following interaction: L ⊃ ε 2cosθ Fµν Y F ′ µν (1.9) where θ is the mixing angle, ε is the coupling strength, and Fµν Y and F ′ µν are the field 18 strengths. The scattering cross-section can be written as σe = 16πµ2 χeαε2αD (m2 A′ +α2m2 e) 2 ≈ {16πµ2 χeαε2αD m4 A′ , mA′ ≫ αme 16πµ2 χeαε2αD (αm4 e) , mA′ ≪ αme (1.10) where µχe is the DM-electron reduced mass and α = g2 D (the U(1)D coupling. The corresponding form factor can be written as FDM(q) = m2 A′ +α2m2 e m2 A′ +q2 ≈ {1, mA′ ≫ αme α2m2 e q2 , mA′ ≪ αme (1.11) where q is the DM-electron momentum transfer. The DM form factor describes how the interaction between dark matter particles and ordinary matter varies with the momentum transfer in a scattering event which which is essential for calculating event rates in dark matter detection experiments and understanding the interaction mechanisms. In a DM- electron scattering interaction, the DM can transfer at most its kinetic energy 1/2mA′ν2, where ν = 10−3c is the galactic DM velocity and c is the speed of light. The parameter space for both regimes is shown in figure 1.6. DM-electron absorption Dark matter (DM) candidates such as axion-like particles (ALPs) and dark photons can interact with ordinary matter via absorption, especially through interactions with bound electrons. The absorption of dark photons A′ can be modeled as the absorption of a non- relativistic particle with mass m with coupling ε to electrons [26, 27]. In absorption other DM transfers its entire rest mass energy mA′c2. 19 The absorption of dark photons or axion-like particles occurs when these particles inter- act with electrons bound in atoms. The process is commonly referred to as the axio-electric effect for axions and axion-like particles. This interaction leads to the absorption of the dark-matter particle, which transfers its energy to an electron, causing it to be ejected from its atomic orbit. For a dark photon, this effect arises due to kinetic mixing with the photon field, while for axions, the interaction is based on the axion’s coupling to the electromag- netic field. For dark photons, the absorption cross section can be expressed as: σγ ′(E) = ε2 m4 γ ′ ( E2 ∆E2 +E2 ) (1.12) where ε is the kinetic mixing parameter, mγ ′ is the mass of the dark photon, E is the energy of the incoming dark photon, and ∆E is the energy width of the absorption process, typically related to the material’s band gap or the electron binding energy. For axion-like particles (ALPs), the absorption cross section is given by: σa(E) = g2 aγγ m4 a ( E2 ∆E2 +E2 ) (1.13) where gaγγ is the axion-photon coupling constant, ma is the mass of the ALP, and the other terms are analogous to those for dark photons. Current direct detection experiments such as XENON100 [41–43], CDMSlite [44], and SuperCDMS [45–47] are sensitive to these types of interactions, especially with electron recoil events. The cross section can be probed by observing the energy deposited by dark- matter particles in the form of electron recoils, allowing for the constraining of the coupling constants (ε for dark photons, gaγγ for ALPs) in regions of parameter space that were previously unexplored. However, the sensitivity of these experiments to lower mass range for dark-matter parti- 20 cles, especially below a few MeV, remains poorly probed. This region is difficult to explore due to the relatively low recoil energies associated with light dark-matter candidates, which are often below the detection thresholds of existing experiments. Moreover, many direct detection experiments have better sensitivity at higher masses (e.g., GeV-scale DM), leav- ing a significant gap in their ability to probe lighter candidates such as dark photons or ALPs with small masses. To overcome these limitations, future experiments are under development with the goal of lowering the energy thresholds and increasing the sensitivity of detectors. These next- generation experiments aim to probe the unexplored regions of parameter space where dark photons and axions may exist. Theoretical studies have already begun to investigate the potential for detecting bosonic DM absorption in materials with lower energy thresholds, such as semiconductors and superconducting materials. These materials probe potentially weakly coupled interactions of low-mass dark photons and axions, and thus could enable the detection of these particles in new mass ranges [48, 49]. 1.4 DM Detection Strategies As shown schematically in figure 1.5, there are the three general categories of exper- iments that provide complementary insight into the nature of dark matter. Indirect dark matter detection strategies focus on the measurement of the products of dark matter inter- actions (i.e. gravitationally) rather than the dark matter itself. Accelerator experiment aim to produce dark matter directly or via decays of heavier new particles. Direct DM detection experiments aim to directly measure the collisions between dark matter and a target ma- terial in terrestrial apparatuses. Importantly, the verification of an excess as a dark matter signal would require concurrent study and verification between all of the above channels as well as consistency with expected daily and annual modulation of the signal [50, 51]. 21 Figure 1.5: Schematic showing the possible channels of dark matter detection. Indirect DM detec- tion involves searching for the standard model (SM) products of dark matter interactions and con- straining annihilation cross sections. Collider searches seek to produce DM through high-energy collisions. Direct Detection probe interactions of DM with a target material through a measurable signal of the energy deposition. Direct detection experiments search for DM interactions with nuclei or electrons in target materials. Large-scale, mature experiments have placed exclusions on much of the DM pa- rameter space above 1 GeV. With the theoretical developments and evidence from collider experiments that DM might have a lighter mass (>GeV) than once previously thought, it is important to devise novel detection strategies to explore the sub-GeV DM parameter space. 1.4.1 Direct Detection of sub-GeV Dark Matter Many new experiments spanning a diversity of fields from atomic [52] and molecu- lar [53] physics to condensed matter physics [54], [55], [56], [57] have been proposed to probe sub-GeV DM via both electron and nuclear recoils. Figure 1.6 shows the approxi- 22 mate mass range of sub-GeV regime, along with some of the proposed experiments that are under development. WIMPAxion DM (a) ����� �$"0- сѷ ��-$*0. (�/ -$�'. !*- � / �/$)" .0�Ҋ� � ��-& (�// - �- $) 0. җ.*'$� '$) .Ҙ *- 0)� - �*).$� -�/$*) !*- /# ) �- / -( *- '*)" / -( җ'*)"Ҋ��.# � �)� .#*-/Ҋ��.# � '$) .Ѷ - .+ �/$1 '4Ҙѵ �# #*-$5*)/�' .��' $)�$��/ . /# (�..Ҋ. ).$/$1$/4 -�)" ѵ �- �$/ѷ �ѵ �..$"ҝ�/*)4 �-**& �)$1 -.$/4Ѹ ���ҝ��-$) ��$) .*Ҋ��'' � �*.*)$� ��-& (�// -Ѷ .0�# �. � ��-& +#*/*)Ѷ �) �3$*)'$& +�-/$�' Ѷ *- � .��'�- +�-/$�' Ѷ 2$/# (�.. . �*2) /* р ( �ѵ ) +�-/$�0'�-Ѷ $) .$'$�*)Ѷ � �*.*)$� ��-& (�// - +�-/$�' �. '$"#/ �. р �ҝc2 ��) � ��.*-� � җ$)./ �� *! .��// - �ҘѶ � '$1 -$)" /* /# ' �/-*) �) ) -"4 �*-- .+*)�$)" /* /# +�-/$�' ҁ. (�.. �)� 3�$/$)" /# ' �/-*) ��-*.. /# �-4./�' ��)� "�+ѵ �*( /�-" / (�/ -$�'. !*- � / �/$)" .0�Ҋ� � ��-& (�// -Ѷ .0�# �. '$,0$� 3 )*)Ѷ '$,0$� �-"*)Ѷ .$'$�*)Ѷ �)� " -(�)$0(Ѷ #�1 � ) 0. � +- 1$*0.'4 !*- � �� . �-�# .ѵ �/# - +-*+*. � /�-" / (�/ -$�'. 2*0'� � ) 2 /* /# ��-& (�// - . �-�# �*((0)$/4ѵ �# . $)�'0� ) 2 .�$)/$''�/*-.Ѷ .0+ -�*)�0�/*-.Ѷ .0+ -!'0$� # '$0(Ѷ �$-�� (�/ -$�'.Ѷ +*'�- (�/ -$�'.Ѷ �$�(*)�Ѷ .$'$�*) ��-�$� Ѷ �-*(�/$� *-"�)$� /�-" /.Ѷ �)�(*' �0'�- "�. . җ�$"ѵ сҘѵ �#�/ҁ. ) � �Ѷ #*2 1 -Ѷ $. � !0'' � .�-$+/$*) *! /# .$")�' +-*�0� � �4 � ��-& (�// - +�-/$�' $)/ -��/$)" 2$/# /# . (�/ -$�'. $) / -(. *! /# +�-/$�' ҁ. +-*+ -/$ .ѵ ) ���$/$*)Ѷ /# . .$")�'. ) � /* � ��'$�-�/ � 3+ -$( )/�''4ѵ �#$. 2*-&Ѷ 2#$�# $. $) .*( ��. . �'- ��4 0)� -2�4Ѷ - ,0$- . � %*$)/ Ȃ*-/ � /2 ) - . �-�# -. $) +�-/$�' Ѷ �./-*+�-/$�' Ѷ �*)� ). � (�// -Ѷ �)� �/*($�Ҋ(*' �0'�-Ҋ*+/$��' +#4.$�.ѵ �-*($.$)" � / �/*-. �) Ȃ �/$1 . �-�# !*- '*2Ҋ(�.. ��-& (�// - - ,0$- . � � / �/$*) ( �$0( �)� � . ).*- 2$/# �/ ' �./ /#- +-*+ -/$ .ѵ �$-./Ѷ /# ( �$0( #�. /* - ��/ /* /# � +*.$/$*) *! /$)4 �(*0)/. *! ) -"4ѵ � �*)�Ѷ /# ( �$0((0./ � .��'��' /* � '�-" )0(� - *! �/*($� ҂/�-" /.҃ /* $)�- �. /# �#�)� *! ��+/0-$)" � -�- $)/ -��/$*) 2$/# � ��-& (�// - #�'* +�-/$�' ѵ �$)�''4Ѷ /# . ).*- �)�( �$0((0./ #�1 �) 3,0$.$/ '4 '*2 -�/ *! � / �/*-Ҋ$)�0� � ���&"-*0)�. җ҂��-& �*0)/.҃Ҙѵ �#4.$�$./. #�1 (�� )*-(*0. +-*"- .. *1 - /# +�./ ! 2 4 �-. $) �0$'�$)" /# !$-./ " ) -�/$*) *! 0'/-�. ).$/$1 � / �/*-. /#�/ �- ��' /* . �-�# !*- ��-& (�// - 2$/# (�.. . �. '*2 �. р � �ҝc2Ѷ �0/ (0�# - (�$). /* � �*) ѵ �# !$-./ � (*)./-�/$*) *! . ).$/$1$/4 /* .$)"' ' �/-*). ��( 2$/# ��/� !-*( � -*0"#'4 рпҊ&" /�-" / �*).$./$)" *! '$,0$� 3 )*)Ѷ 2#$�# #�. �) 3�$/�/$*) ) -"4 *! �-*0)� рс � ҙтѶ рчҚѵ �*/# �-"*) �)� '�-" - 3 )*)Ҋ��. � � / �/*-. #�1 .0�. ,0 )/'4 .#*2) $(+- ..$1 . ).$/$1$/4 /* � -/�$) ��-& (�// - ��)�$��/ . �. '$"#/ �. �-*0)� ф � �ҝc2 ҙршѶ спҚѶ �0/ !0-/# - +-*"- .. $. �0-- )/'4 #�(+ - � �4 ���&"-*0)�.Ѷ �. 3+'�$) � � '*2ѵ �) � 1$� /#�/ #�. . ) $)�- �$�' +-*"- .. - � )/'4Ѷ $)�'0� . � .+ �$�' ҂�&$++ -҃ �#�-" Ҋ�*0+' � � 1$� җ���Ҙ ҙсрѶ ссҚѶ �*).$./$)" *! ($''$*). *! .$'$�*) +$3 '. /#�/ �- рф (( × рф (( × хцф (( $) .$5 ѵ �#$. � 1$� $. ��' /* +- �$. '4 ( �.0- /# )0(� - *! ' �/-*). $) ��# +$3 'Ѷ +-*1$�$)" .0Ȃ$�$ )/ . ).$/$1$/4 /* .+*/ � �*''$.$*) � /2 ) � �ҝc2 ��-& (�// - �)� � .$'$�*) �-4./�' /#�/ +-*(*/ . �) ' �/-*) /* /# �*)�0�/$*) ��)� җ�$"ѵ тѶ ' ȅҘѵ ���# �&$++ - ��� #�. � (�.. *! *)'4 ��*0/ с "ѵ � !$-./ ��-& (�// - . ).$/$1$/4 / ./ 2$/# � .�$ )� Ҋ"-�� ��� #�. � ) �*)�0�/ � �4 /# ����� �*''��*-�/$*) �/ � -($'�� ҙссҚ җ. �4)*+.$.ѷ ��-& ��// - � / �/*- �-*1 . $/. � ).$/$1$/4Ҙѵ �# ) 3/ ./ +Ѷ +'�)) � !*- спсрѶ $. /* $)./�'' � -*0"#'4 рппҊ" � / �/*-Ѷ �*).$./$)" *! ��*0/ фп �&$++ - ���.Ѷ �/ ��)���ҁ. 0)� -"-*0)� '��*-�/*-4 ��� ��ѵ �# ��� �Ҋ� �*''��*-�/$*) $. �'.* +'�))$)" /* $)./�'' � рҊ&" � / �/*- �/ /# �*��) �)� -"-*0)� ��*-�/*-4 $) �-�)� ҙстҚѶ �)� /# �� � +�-/( )/ *! �) -"4 җ���Ҙ $. !0)�$)" /# �о� *! � (0'/$Ҋ&" � / �/*-Ѷ ��'' � �.�0-�Ѷ 2#$�# �*0'� /�& ��/� '�/ - /#$. � ��� ҙсуҚѵ �# �*)./-0�/$*) *! /# . '�-" Ҋ(�.. � / �/*-. $. �#�'' )"$)"ѵ �# ���. ) � /* � .0Ȃ$�$ )/'4 '*2 �*./Ѷ /# - ��Ҋ*0/ ' �/-*)$�. (0./ - �� *0/ (�)4 ���. �/ *)� �)� 2$/# '*2 )*$. Ѷ �)� /# -��$*��/$1 ���&"-*0)�. (0./ � �*)/-*'' � /* $)�- �$�' +- �$.$*)ѵ +#4.$�.ѵ�+.ѵ*-" Ҟ Ҷ спсп �( -$��) �#4.$��' �*�$ /4 Ҟ � � (� - ррѶ спсп Ҟ �#4.$�. ртѶ рцс Ҟ �� ѷ рпѵррптҝ�#4.$�.ѵртѵрцс ��" т (b) sub-GeV DM Figure 1.6: (a) Parameter space of dark matter (DM) mass showing the sub-GeV regime between WIMP and axion DM predictions. (b) Experiment proposals for sub-GeV dark matter searches. Materials in use are indicated by solid lines, while those under consideration for the near-term and long-term are shown by solid-dashed and light-dashed lines, respectively. The horizontal scale indicates the mass-sensitivity range. Figure adapted from ref. [58]. 1.4.2 n-type GaAs as a Cryogenic Scintillating Target Scintillators are a widely adapted experimental target for direct dark matter searches [59]. In direct DM searches, when dark matter particles interact with electrons in a scintillating material, they deposit energy, causing the material to emit photons through radiative de- 23 Acceptor Energy Level Wavelength Shallow defects 1.44 eV 861 nm Ionized boron on arsenic site 1.33 eV 930 nm Si˙GaV˙Ga complex 1.16 eV 1069 nm Si˙GaV˙GaSi˙Ga complex 0.93 eV 1333 nm Table 1.1: Radiative acceptors in n-type GaAs and their associated energy levels and wavelengths. cays. These photons are then detected by sensitive detectors, providing a measurable signal that can be analyzed to infer the presence of dark matter. As a novel cryogenic scintillator, n-type Gallium Arsenide (GaAs) features a band gap of 1.52 eV and desirable properties as a target for the MeV/c2 mass range, which falls within the poorly explored sub-GeV regime [60]. An n-type GaAs target coupled to a single photon detector is the system of interest that this work will focus on. N-type GaAs doped with silicon (Si) and boron (B) was first proposed as a cryogenic scintillator for the detection of sub-GeV DM via elec- tron recoils due to its high luminosity and the commercial availability of large, high-purity crystals[61–63]. Scintillation photons are emitted from n-type GaAs via the following pro- cess: (1) Electronic excitation leaves one or more holes in the valence band (2) Acceptors trap the valence band holes (3) The acceptor holes recombine radiatively with delocalized donor band electrons and scintillation photons are emitted The emission wavelengths are thus dependent on acceptor energy levels. Known accep- tors and their associated energy level/corresponding scintillation wavelength are listed in table 1.1 [64]. Scintillation in the 930 nm band is the most dominant. Figure 1.7 adapted from ref [62]shows the x-ray excited luminescence spectra of n-type GaAs doped with silicon and boron measured with silicon and InGaAs CCDs at 10K, where the contribu- tions from these different radiative decay processes can be seen. The relative intensities of 24 > REPLACE THIS LINE WITH YOUR PAPER IDENTIFICATION NUMBER (DOUBLE-CLICK HERE TO EDIT) < 3 doped GaAs crystals. But in the framework of this work, such features will not be considered. For simplicity, these emission peaks were collected in one group – band A. The nature and specific of these short wavelength emissions can explains the position and asymmetric shape of band A in GaAs:Si and GaAs:(Si, B) (see Fig. 2). As can be seen from the experimental and literature data, the shape, intensity and peak position of the bands in 800-900 nm region in GaAs crystals drastically depends on purity, type of dopants and defects [11-15]. And it also depends, as shown in Reference [13], on the type of conductivity. The transitions from shallow silicon donors to boron acceptors on an arsenic site leads to emissions at 930 nm (band B) [10, 16, 17]. The energy relaxation on gallium vacancy-donor center leads to emission at ~1070 nm [18]. This center (band C) is a gallium vacancy (VGa) bound to a donor (like Si, Ge, Sn, C, S, Se, Te). The peak position of band C depends on the type of activator impurity. The nature of the broad band D at ~1335 nm is associated with defects or complex centers like (SiGa + VGa + SiGa) or (SiGa + SiAs) and other [11, 12]. It is important to note that the formation of long-wavelength centers at ~1377 nm was observed in undoped GaAs crystals with neutron dose. The intensity of band D increases with neutrons dose rate increasing, which indicates the important role of vacancies in the formation of this emission center [12]. B. Light Yield Estimation The photocurrent of all samples under X-ray was measured using the InGaAs or Si photodiodes with or without filters and shown on Fig. 3. Table III compiles the total light output and the contribution of each emission band separately which were obtained in this work based on PIN detectors photocurrent. The obtained results indicate a very high efficiency of the scintillation process in activated crystals. It is seen that the LY of the GaAs:(Si, B) is very high: 125 ph/keV. The Si doped GaAs crystal has LY=71 ph/keV. It is important to note that about 65% of the total output of activated GaAs is contained in the region of 1000-1700 nm, i.e. bands C and D. This feature is dictated by the specificity of the radiative relaxation of energy in the co-doped GaAs crystal. The results show how important the addition of the acceptor boron is. Also, this indicates the need to use photo detectors that are sensitive in a fairly wide range (like Ge, InGaAs). Silicon-based photodiodes are only partially suitable as photodetectors, since 35 to 60% of the light is lost in case of GaAs:(Si, B) and GaAs:Si respectively. It should be noted that the experimental approach used to determine the light output does not take into account the refractive index. GaAs has a refractive index of 3.55 at 930 nm and 3.4 at 1300 nm. Since the scintillation light is emitted isotropically, only a small fraction is able to exit the crystal directly. In this experiment, scattering within the crystal or on the roughened surface allowed a large fraction of the light to exit. A coating with a graded index of refraction will allow Fig. 2. The X-ray luminescence spectra of undoped GaAs, GaAs:Si and GaAs:(Si, B) crystals obtained by Si (left column) and InGaAs (right column) CCDs at 10K. TABLE II THE PARAMETERS OF RADIATIVE CENTERS IN GAAS CRYSTALS Band λ, nm E, eV Comments A 800-870 1.42-1.54 Transition from CB to VB, exciton and shallow levels (defects) B 933 1.33 The boron in arsenide site (BAs center) C 1050-1090 1.14-1.18 Center like (SiGa + VGa) D 1305-1335 0.93-0.95 Complex center like (SiGa + VGa + SiGa) or (SiGa + SiAs) Fig. 3. The InGaAs (Top) and Si (Bottom) PIN photodiodes current dependence on X-ray filament current with/without the different shortpass (FES) and longpass (FEL) optical filters obtained on the GaAs:(Si, B) (Left) and GaAs:Si (Right) sample at 10 K. Figure 1.7: X-ray excited luminesence spectra of n-type GaAs doped with silicon and boron mea- sured with a silicon CCD (left) and an InGaAs CCD (right) showing the near-infrared emission bands. Figure adapted from ref.[62]. the emission bands depend on the dopant and free carrier concentrations, as do the decay timescales. One of the most desirable features of n-type GaAs compared to other cryogenic scintil- lators is its lack of afterglow [60], the emission of scintillation photons at long time scales after a particle interaction has occurred. In order to avoid these afterglow effects in GaAs, the samples must be doped above the so called Mott transition [65]. In cooling the crystals from room temperature to below 1 K, some of the free carriers bind to individual silicon atoms at an energy level 2.3 meV below the conduction band, and repulsion then forces the additional electrons into the next higher available energy level, which is in the conduc- tion band. Electrons in the conduction band remain highly mobile down to 0 K, and they serve two purposes: (1) annihilation of any meta-stable radiative states that would result in unwanted afterglow, and (2) efficient combination with electron holes to maximize the prompt scintillation light (the desired signal). To demonstrate this, figure 1.8 adapted from reference [63] shows the intensity of the scintillation light emitted from a GaAs sample at 10K of sample before, during, and after 25 as the system returns to its ground state. A closer look at the GaAs emission curve [Fig. 8(b)] shows that it is indistinguish- able from the instrumental background. The apparent absence of thermoluminescence in n-type GaAs can be explained by the annihilation of metastable radiative states by the delocalized conduction band electrons. A search of the literature found only one paper describing thermally stimulated luminescence from a conductive semiconductor.32 This paper described thermolumi- nescence from an n-type ZnSe crystal using UV (365 nm) exci- tation. This was later explained as the absorption of the UV exciting radiation (attenuation length 10!4mm) by a layer of non-conductive surface defects.33 IV. CONCLUSIONS (1) Electron excitations "1.44 eV can produce 1.33 eV photons. (2) The most luminous GaAs sample studied (#13316) has an observed scintillation luminosity of 43 photons/keV, below the theoretical limit of about 200 photons/keV. (3) Ionization from background radiation (e.g., muons and gamma rays) is not expected to cause afterglow, pro- vided that the free carrier concentration is above the Mott transition, where even near 0 K conduction band electrons are not frozen out and can annihilate any meta- stable radiative states. These measurements show that n-type GaAs is a promis- ing cryogenic scintillator for DM particle detection in the MeV/c2 mass range in that it can be grown as large, high- quality crystals, and has good scintillation luminosity, a threshold sensitivity at the 1.52 eV bandgap, and potentially no afterglow. However, more work is needed to optimize the doping concentrations, to reduce hole traps that compete with boron acceptors, and to develop anti-reflective coatings FIG. 7. (a) Optical emission of sample 13 316 at 10 K before, during, and after exposure to a 50 keVp X-ray beam, measured in 2 s time bins. (b) Enlargement of emission around the time the beam was turned off. FIG. 8. (a) Thermally stimulated luminescence of GaAs sample #13330 and crystals of NaI(Tl), NaI, and CsI as a function of temperature after a 30-min expo- sure to a 50 keVp X-ray beam at 10 K. All crystals were similar in size. The vertical scale is the same for all samples. (b) Comparison between the GaAs curve and the instrumental background with vertical scale enlarged 20x. 114501-5 Derenzo et al. J. Appl. Phys. 123, 114501 (2018) 09 N ovem ber 2024 19:55:39 Figure 1.8: Figure adapted from ref [63] showing (a) the normalized intensity of emitted scintillation light before, during, and after irradiation with 50 keVp X-ray beam and (b) an inset showing the time during which the shutter closed. Of note is the sharp return to a baseline around 0 intensity after irradiation. a 600 second exposure to the 50 keVp X-ray beam. The scintillation intensity is cleanly correlated with the opening and closing of the excitation beam shutter with no lingering emission after irradiation, as shown by the return to a baseline around 0 intensity. 1.4.3 DM Detection concept using n-type GaAs In the detection concept that is the focus of this work, photons are emitted from n- type GaAs after excitation by DM-electron scattering or absorption of DM by an electron and subsequent radiative recombination. The near-infrared scintillation photons are then detected by an SNSPD. The signal associated with a DM – electron interaction depends on the mass of the DM particle and the interaction. In a DM-electron scattering event, for the case that the DM particle is heavier than an electron (the case relevant for n-type GaAs as a target material), the maximum energy transferred is the kinetic energy of the DM. Given a galactic DM velocity vχ = 10−3 c, where c is the speed of light, a 10 MeV/c2 DM particle transfers 26 Figure 1.9: Schematic depiction of the detection concept: a DM particle scatters off (or absorbs onto) an electron in n-type GaAs, exciting it to a higher-energy level; scintillation photons are emitted through radiative recombination processes; these are subsequently detected by the SNSPD array 27 Ee ≤ 1 2 mχv2 χ ≲ 1 2 (10MeV)2 (10−3c )2 ≲ 5eV [39]. (1.14) A 5 eV energy deposit into n-type GaAs leads to the emission of 3–4 scintillation pho- tons given its 1.52 eV band-gap. In a DM-electron absorption event, the DM transfers all of its rest mass to the electron. The energy transferred to an electron is thus given by simply the DM mass [39]: Ee ≈ mχc2 χ ≈ mχ . (1.15) In principle, using n-type GaAs as a target, bosonic DM (dark photons, scalars, and axion- like particles) can be probed via absorption onto electrons in n-type GaAs for DM masses down to 1 eV [48]. Figure 1.10 shows projections for GaAs as a scintillating target for dark photon absorp- tion with sensitivity down to one or more photons. Figure 1.11 shows projected sensitivities of the dark matter - electron scattering cross section as a function of dark matter mass for GaAs targets, given exposures of 1g-month to 10kg-years for two different form factors. 1.4.4 Sensor Classes for n-type GaAs and Low Mass DM Targets The sensor requirements for the scintillation photons emitted from n-type GaAs are high sensitivity in the near-infrared, low dark count rate, and cryogenic operation. Given this, superconducting devices are natural candidate to develop detection concept. It should be noted that the development of SNSPDs for this detection concept will extend to many other experiments in the dark matter detection landscape in the future. SNSPDs are typically operated at sub-kelvin cryogenic temperatures and have been 28 J H E P 0 6 ( 2 0 1 7 ) 0 8 7� �� ��� ��� ��� ��­�� ��­�� ��­�� ��­�� ��­�� ��� [��] ϵ �� �� �� � � �� ��� �� �� ������� ��������� ����������� ��� ������� ��������� �� ��������� �� Figure 3. Constraints (shaded regions) and prospective sensitivities (solid colored lines) for axion- like particle (ALP) dark matter (left) and dark-photon (A′) dark matter (right), assuming that the ALP/A′ constitutes all the dark matter. Colored regions show constraints from XENON10, XENON100, and CDMSlite, as derived in this work, as well as the DAMIC results for A′ from [85]. Shaded bands around XENON10 and XENON100 limits show how the bound varies when changing the modeling of the secondary ionization in xenon. Deep- and light-purple solid lines show projected 90% C.L. sensitivities for SuperCDMS SNOLAB HV using either Ge (20 kg-years) or Si (10 kg- years) targets, respectively. Yellow, orange, and green solid lines show projected sensitivities for hypothetical experiments with the scintillating targets CsI, NaI, and GaAs, assuming an exposure of 10 kg-years. All projections assume a realistic background model discussed in the text, but zero dark counts to achieve sensitivity to low-energy electron recoils. In-medium effects are included for all A′ constraints and projections. Shaded gray regions show known constraints from anomalous cooling of the Sun, red giant stars (RG), white dwarf stars (WD), and/or horizontal branch stars (HB), which are independent of the ALP or A′ relic density. Also shown (left) are the combined bounds from XENON100 [22], EDELWEISS [21], CDMS [20], and CoGeNT [19]; and (right) a bound derived in [35] based on XENON100 data from 2014 [22]. Shaded orange region in left plot is consistent with an ALP possibly explaining the white dwarf luminosity function. – 12 – Figure 1.10: Constraints (shaded regions) and projected sensitivities (solid colored lines) for dark photon dark matter. The yellow, orange, and green solid lines show projected sensitivities for hypo- thetical experiments with the scintillating targets CsI, NaI, and GaAs, assuming an exposure of 10 kg-years. Figure adapted from ref. [48]. 29 � �� ��� ��� ��-�� ��-�� ��-�� ��-�� ��-�� ��-�� ��-�� ��-�� ��-�� ��-�� ��-�� ��-�� ��-�� ��-�� ��-�� ��-�� ��-�� �χ [���] σ � [� � � ] ������� � � � � � ��� � ���� ���� � � � � �- � � �� � � � �� ������� ����@� ������ ������� ����@� ���� � � � � � �� ��� ��������� ���=� ����� �� ��- ��������� � ��-� ��� ����� ��� �-������ ����� � �-������ ����� �@��� �� ��� ��� ��� ��� ��� ��-�� ��-�� ��-�� ��-�� ��-�� ��-�� ��-�� ��-�� ��-�� ��-�� ��-�� ��-�� ��-�� ��-�� ��-�� ��-�� �χ [���] σ � [� � � ] ������� ��� ��� �� ��� ��� ����� ����� � ����� ����� ���� ��@� ����� � ��� �-� ��� ����� ��� ���@ ��� �� ���=(α��/�) � ��� � �������� ������@����� ��� �� �� ��-� ������ �� � � �-�� �� ��� �� ��� �-� ���� � ��� �� � � -�� ���� Figure 1.11: Projected sensitivities (blue lines) to dark-matter-electron scattering mediated by a heavy mediator (left) or light mediator (right) of a GaAs target assuming for different exposures (1g-month, 100g-month, 1 kg-year, 10 kg-year). We assume zero background events for events with one or more photons, a radiative efficiency of 1, and a photon detection efficiency of 1. Ex- isting constraints are shown in gray from SENSEI, DAMIC at SNOLAB, XENON10, XENON100, XENON1T, DarkSide-50, and CDMS-HVeV [66–75]. Orange regions labelled “Key Milestone” are from [76]. demonstrated with near-unity system detection efficiency in the NIR and ultra-low dark count rates. The main challenge with SNSPDs in the context of instrumenting large-scale scintillating targets is scaling the active area of SNSPDs to the cm2-regime while preserving these detection metrics to the greatest extent possible. As an alternative sensor, TES-based technologies are under development to instrument n-type GaAs. TES are somewhat maturely developed for nuclear recoil DM experiments. In large collaborations such as SuperCDMS [45], large-area cryogenic detectors, with a to- tal area of 45cm2, utilizing TES-based athermal phonon sensors have achieved 4 eV energy sensitivity [45]. Given that σE ∝ √ A, where A is the total instrumented area, resolutions of 600meV in 1cm2 detectors are feasible given the current state-of-the-art. Extending ather- mal phonon sensors to lower mass electron recoil DM searches will entail design features such as lowering the critical temperature (Tc) of the TES and optimizing sensor geometry 30 to improve athermal phonon collection efficiency. This would enable further enhancements in sensitivity and expand the instrumented active area per channel beyond 1cm2. In GaAs, both phonons and NIR photons are produced after excitation. The TESSER- ACT collaboration is developing both nuclear recoil and electron recoil experiments with the approach of exploring multiple target materials and developing multiple detection strate- gies. One of the developing ER experiments involves n-type GaAs instrumented with TES- based athermal phonon sensors. The prototype design includes a 1 cm3 GaAs target in- strumented with athermal phonon sensors on its surface, and inside an optical cavity with a 1 cm3 x 1 mm thick Ge crystal that serves as an infrared photon collector [77]. TES readout offers the possibility to use the same detector to measure both scintillation and phonons in coincidence which can give background discrimination capabilities. The main challenge to TES readout for low energy threshold DM experiments is the low energy excess (LEE), the observed increase in event rates or signals at low energy thresholds that are not accounted for by known backgrounds or expected detector noise. Nearly all experiments instrumented with phonon sensors with sufficiently low energy thresholds see a this steeply rising event rate [78–81]. Efforts to understand the origin of [82] and mitigate the LEE are ongoing. 31 Chapter 2: Superconducting Nanowire Single Photon Detectors 2.1 Introduction Since their first demonstration by Golt’sman et al. in 2001 [83], superconducting nanowire single photon detectors (SNSPDs) have emerged as detector for time correlated single pho- ton counting from the UV to the near-infrared. Specialized devices have shown detection efficiency of 98% at 1550 nm [84], timing jitter as low as 2.7 ps [85], intrinsic dark count rates of 6×10−6 counts per second [86], and energy thresholds as low as 42.8 meV (wave- length 29 µm) [87]. Many of these demonstrations have involved detectors that are opti- mized for one performance metric at the expense of others. As a relatively new detector technology, one of the challenges for SNSPDs is to realize devices that perform well across multiple metrics. This continues to develop as our understanding of the fundamental device physics and intrinsic trade-offs becomes more complete. SNSPDs have made their way into a wide variety of applications, including quantum optics [88], [89] quantum information science [90–95], imaging [96, 97] and free-space laser communication [98]. Most of these demonstrations involve fiber-coupled light in- cident on a small, circular active area. Driven by imaging [96] and astronomy[99] ap- plications, the active areas of SNSPDs have grown steadily over the past several years from the order of the cross section of a single-mode optical fiber core (10s of µm2) to ar- rays with active areas in the mm2 regime [100]. Large-scale arrays up to kilopixel [101] and now megapixel-scales [102] have been developed using row-column [103] or ther- 32 mal row-column multiplexing techniques[104] that provide spatial resolution, and time-of- flight readout bus multiplexing for large scale cameras [102]. Recent developments in nanofabrication, device characterization, and modeling of fun- damental device physics have pushed forward the performance of SNSPDs in their his- torical strengths as single photon detectors, revealing novel applications in fundamental physics. The project described in this paper will focus on the development of SNSPDs for dark matter searches using cryogenic scintillating targets. With sufficient research and development of the technology, these SNSPDs have the ability to provide new capabilities for low-mass dark matter searches and across many areas of High Energy Physics. 2.2 Device Operation Principles SNSPDs are constructed from current-carrying nanoscale wires. Typical devices have wire widths from 10s to 100s of nm and thicknesses of 2-10 nm. A common technique to achieve high optical coupling efficiency and absorption is to meander the nanowire over the active area with a high fill factor. SNSPDs are usually operated from 1–4 K depending on the superconducting materials used (WSi, MoSi, NbTiN, NbN, etc.) with a DC bias current applied through a bias tee. In parallel with the nanowire is the readout arm which consists of a load impedance, usually the input to a readout amplifier. The readout portion can be DC or AC coupled. A schematic of the photon detection mechanism in SNSPDs is shown in figure 1. The detection cycle begins with the nanowire in the current-carrying supercon- ducting state. When a photon is absorbed in the cross section of the nanowire, it triggers a cascade of events that ultimately leads to a detection signal. Qualitatively, this cascade can be understood as follows: absorption of a photon leads to the formation of electron-hole pairs, which through a down-conversion process transfer their energy into a hot-spot of ex- cited quasi-particles and high energy phonons. The evolution of this hot-spot leads to the 33 4 Figure 1.1: (a) Schematic of the SNSPD detection process. The device begins in the current-carrying superconducting state (1) until a photon is absorbed (2) and generates a hotspot of excited quasiparticles and high-energy phonons through a process of downconversion shown in (b). The evolution of these quasiparticles leads to the instability of the superconducting state, resulting in the formation of vortices (vortex-antivortex pair generation if the hotspot is in the center of the nanowire or vortex entry from the edge if the absorption is near the edge) as shown in (c). As the vortices move due to the current flow (3), they dissipate energy, leading to the formation of a normal domain across the entire cross-section of the nanowire. Once formed, the normal domain grows along the length of the wire due to Joule heating (4), increasing the impedance of the device. This change in impedance diverts current from the nanowire to the readout circuitry, leading to a voltage transient (center). Once current is diverted to the readout, the nanowire cools, recovers to the superconducting state, and current returns to the device. (d) A typical SNSPD is operated using a bias-tee with the DC port carrying the bias current and the RF port coupled to a low-noise amplifier. Figure 2.1: (a) Schematic depiction of the SNSPD photon detection mechanism reproduced from [105]. The device begins in the current-carrying superconducting state (1) until a photon is absorbed (2) which generates a hot-spot of excited quasi-particles and high-energy phonons through a down-conversion process depicted in (b). The evolution of this hot-spot destabilizes the supercon- ducting state and leads to the formation of vortices, which are depicted in (c). Current flow causes the vortices to move (3), and they dissipate energy, causing the formation of a normal domain across the entire cross-section of the nanowire. The normal domain propagates along the length of the wire (4), which increases the impedance of the device. This change in impedance diverts current from the nanowire to the readout circuitry, leading to a voltage transient (center). Once current is diverted to the readout, the nanowire cools, recovers to the superconducting state, and current returns to the device. (d) A typical SNSPD is operated using a bias-tee with the DC port carrying the bias current and the RF port coupled to a low-noise amplifier. 34 instability of the superconducting state, and vortices form. The dissipative vortex evolution leads to the formation and growth of a normal domain which spreads along the length of the nanowire, increasing the impedance of the device, until the nanowire transitions fully into the normal state. At this time, the bias current gets diverted into the readout electron- ics. The nanowire then cools and recovers to its superconducting state, and its impedance decreases until the bias current returns fully to the device. In typical SNSPDs, the photon absorption and hot-spot formation phase typically occurs on a picosecond timescale, and the recovery phase timescale which is set by the kinetic inductance of the nanowire, is on the order of nanoseconds[105]. Traditional SNSPDs are comprised of nanowires with widths O(100 nm) meandered over the photoactive area and are fabricated using electron-beam lithography (EBL) to achieve the required resolution and optimal performance. Recent advances in the prepara- tion of thin superconducting films have enabled the fabrication of a novel class of single- photon detectors with wire widths of several micrometers. Microwire devices offer many advantages over traditional nanoscale devices for applications such as dark matter searches where large active areas area required, as they are amenable to fabrication via optical lithog- raphy - a more scalable fabrication process than electron beam lithography which is used to fabricate traditional devices [106]. 2.3 Detection Metrics This section reviews some of the detection metrics of SNSPDs, following ref [107] and the current state-of-the-art for each metric. 35 2.3.1 Detection Efficiency The system detection efficiency (SDE) is the probability of registering an electrical out- put signal produced by a photon once the photon has entered into the the photon detection system. The SDE can be thought of as being comprised of multiple components: SDE = ηcoupling ·ηabsorb ·ηinternal ·ηtrigger, (2.1) where ηcoupling is the efficiency of the delivery of light to the active area of the device, ηabsorb is the probability that a photon incident on the detector’s active area gets absorbed into the material, ηinternal is the probability that an absorbed photon results in an observable output pulse, and ηtrigger is the efficiency of the readout electronics in registering the output signal as a detection event. While conceptually useful, it is often difficult to accurately measure each separate com- ponent of SDE. Instead, it is conventionally measured by using the following relation: SDE = Rlight −Rdark Rincident , (2.2) where Rlight is the recorded count rate when the device is under illumination, Rdark is the count rate recorded when the device is not under illumination (background count rate), and Rincident is the incident rate of photons delivered to the detector. The current state-of-the-art result for the highest SDE is 98% for the TE polarization (electric field parallel with the nanowire grating) at 1550 nm in a MoSi SNSPD developed by the NIST group [84]. This was achieved using an optical stack with distributed Bragg reflector (DBR) mirrors and a 50-µm diameter active area. 36 2.3.2 Dark Count Rate A dark count in a photodetector is defined as an output pulse that registers as a detection event in the absence of an incident photon. The dark count rate (DCR) can be expressed as: DCR = Rintrinsic +Rbackground +Relectronic. (2.3) The intrinsic dark counts, Rintrinsic, in SNSPDs result from spontaneous resistive state formation. At the temperatures at which SNSPDs operate (1–4 K), it has been shown that the primary phenomenon underlying intrinsic dark counts is thermally activated vortex- anti-vortex pair unbinding [108] or vortex entry and crossing from the nanowire’s edge. This effect is exacerbated by geometric factors [109, 110], such as constrictions in the nanowire width or sharp bends at the ends of meanders, where it is observed that dark count rates increase [110]. In these regions, increased current density from current crowd- ing [111–114] lowers the barrier for both vortex entry and vortex-anti-vortex unbinding. Background counts, Rbackground, primarily result from spurious photons reaching the detector’s active area, originating from black-body radiation outside or within the cryo- stat. Spurious photons can also arise from Cherenkov radiation in materials surrounding the detector or from transition radiation from charged particles passing through bound- aries between media with different dielectric properties. The background count rate in a given system, Rbackground, can be reduced through appropriate shielding around the detec- tor, the implementation of light-tight packaging, the use of mid-IR absorbing paint on areas surrounding the detector in the cryostat, and spectral filtering to attenuate the black-body radiation reaching the detector. This is a particularly important consideration for rare-event search applications such as dark matter detection. The electronic noise component, Relectronic, arises from electronic noise which causes 37 the readout electronics to register a detection click that did not originate from the detector itself. This can be mitigated by using low-noise cryogenic amplifiers, or if the waveforms of an SNSPD are measured on an oscilloscope, pulse shape discrimination between an actual SNSPD output pulse and a detection event caused by electronic noise can be implemented. The current record for the lowest DCR observed in an SNSPD was measured to be 6× 10−6 cps in a WSi SNSPD with an active area of 400 µmby 400 µm, measured at 300 mK [115]. Chapter 4 of this thesis details the study of the dark count rates in large-area microwire arrays and efforts. 2.3.3 Timing Resolution Because SNSPDs are used for time-correlated single-photon counting, timing resolu- tion is an important detection metric. The timing resolution is typically characterized by the instrument response function (IRF), which is measured by comparing the detector re- sponse to a reference. Typically, the full width at half max (FWHM) of the IRF is taken to be the timing jitter of the device. The measured IRF has contributions from intrinsic and extrinsic IRFs. Extrinsic contributions from electrical and amplifier noise can broaden the IRF [116, 117], and a lot of work has been done in the SNSPD community to engineer low-noise cryogenic amplifiers and readout to minimize this. Intrinsic contributions include contributions from the stochastic nature of energy down-conversion in the superconductor after photon absorption and those from geometric jitter, geometric inhomogeneities, and constrictions in the device. These dominate in large-scale meander arrays. Chapter 3 details these effects and presents jitter measurements in large-area microwire arrays. Sub-3-ps jitter has been demonstrated in specialized 5-µm-long niobium nitride (NbN) devices that were designed to minimize intrinsic (down-conversion) and geometric jit- 38 GND AMP SNSPD photon arrival pulse delays longitudinal geometric jitter time a) b) Figure 2.2: The origin of longitudinal geometric jitter is shown in (a), where the microwave signals associated with photon detection events along different segments of the nanowire have different lengths to travel. This leads to different delay times relative to the photon arrival, shown in (b). ter [85]. 2.4 Design Tradeoffs for Large-area SNSPDs 2.4.1 Active Area vs Timing Performance There is an intrinsic trade-off between the active area and timing performance in SNSPDs. Increasing the active area of an individual pixel also increases the length of the nanowire, which in turn increases the kinetic inductance. The recovery timescale in SNSPDs is gov- erned by the relationship Tr = Lk RL , where Lk is the kinetic inductance and RL is the load resistance. Because of this direct proportionality, the recovery timescale (Tr) will be slower in SNSPDs with longer wire lengths, which will also limit the maximum count rate achiev- able by the detector. Additionally, longer nanowires have greater longitudinal geometric jitter, which arises from propagation delays due to detection events at varying locations along the nanowire’s length. The origin and effect of longitudinal geometric jitter on delay times in the onset of the SNSPD detection pulse are shown in figure 2.2. This effect can be 39 mitigated using differential readout, in which both ends of the nanowire are connected to readout circuits. In this configuration, when a photon is absorbed at a particular position in the nanowire, the counter-propagating positive and negative voltages pulses counter prop- agate, and the arrival times of the signals at the two ends are which are slightly offset from each-other can be analyzed such that the propagation delay effects are canceled out. Larger SNSPD pixel areas have more bends and a greater probability of constrictions occurring along the nanowire length. Constrictions have the effect of locally increasing current density and causing suppressed Iswitch . This prevents operation at higher Ibias, at which the jitter performance is better due to enhanced signal-to-noise ratio (SNR). 2.4.2 Array Size vs Readout Complexity One way to alleviate the constraints described in the previous subsection, where larger SNSPDs have the intrinsic trade-off of lower timing resolution, and a reduced maximum count rate (slower recovery time), is to design an SNSPD array in which multiple pixels operate in parallel. Direct readout of individual pixels increases thermal load in the cryostat and readout complexity. Given this, the mm2-scale arrays studied in this work have been designed in linear array formats. These arrays, with pixel counts of 4, 8, and 16, are feasible for direct readout. The direct, single-ended readout scheme employed in this work has been scaled to 64-pixel arrays [118]; however, in the pursuit of cm2-scale devices with thousands of pixels, this approach isn’t scalable. Multiplexing techniques reduce the number of coax lines required to bias and readout arrays. 2.5 Large-active area SNSPDs In recent years, the active areas of SNSPDs have steadily increased from the scale of a single-mode optical fiber diameter (∼100 µm2) to the mm2 regime [119]. As new applica- 40 tions for SNSPDs that leverage these large active areas emerge, there is a growing interest in exploring novel materials and device architectures. For applications in high energy physics such as dark matter (DM) searches, it is necessary to realize SNSPDs with larger active areas than the current state-of-the-art, by at least an order of magnitude. Scaling devices beyond the mm2 regime remains an open question that presents many challenges in terms of design, fabrication, and readout. 2.5.1 Materials The question of which materials are best for large-active area SNSPD fabrication re- mains an open topic in the community. For scaling to mm2 active areas and beyond with large pixels sizes, it is an open question which approach to take as far as the material selec- tion. The challenge is to yield a thin-film superconductor with high material homogeneity over a large active area. WSi and MoSi are promising materials for large active-area de- tectors since the films are amorphous and are, thus, expected to be very homogeneous over large areas compared to polycrystalline superconductors such as NbN or NbTiN. Producing a thin-film superconductor with a large area is generally easier with an amorphous mate- rial than with a polycrystalline one. Amorphous superconductors lack the grain boundaries present in polycrystalline materials, which can interrupt superconducting pathways and create weak links between grains. This uniformity makes amorphous films less prone to structural defects over large areas, which is beneficial for achieving consistent supercon- ducting properties across the entire film. In polycrystalline films, grain boundaries intro- duce stress and dislocations that are amplified over large areas. Amorphous films, being structurally homogeneous, have fewer internal stresses and are more resilient over larger areas. Additionally, amorphous films are easier to deposit over large areas because their lack of well-defined crystal structure makes them less sensitive to the surface on which 41 they’re deposited. This allows for more flexibility in deposition conditions and substrates, making it simpler to produce large, uniform films without needing the precise conditions required for well-formed crystal growth in polycrystalline films. The substrate versatility of amorphous superconductors was shown with MoSi in reference, wherein MoSi deposited on a passivated GaAs surface was observed to have the same critical temperature as MoSi deposited on silicon, and the detection metrics of the devices fabricated from both of these were very similar [120]. 2.5.2 Device architecture In traditional devices, nanowires are often patterned in a meander geometry to enhance optical absorption. This design allows the nanowire to cover a larger area by folding back and forth in a continuous, closely packed pattern, thereby increasing the likelihood that an incoming photon will interact with the detector. The fundamental unit enabling superconducting meanders is the hairpin bend, which consists of a single 180° turn connecting two wire segments. However, one of the main factors limiting the performance of meander-based SNSPDs is the current crowding ef- fect, particularly in the bending regions [111, 112]. In SNSPDs, current crowding oc- curs when the super-current density increases at sharp bends, corners, or discontinuities in the nanowire. This localized increase in current density can surpass the superconductor’s switching current, Iswitch, causing the SNSPD to transition to the normal state prematurely. As a result, the performance of the detector is degraded. Additionally, the higher current density in the bends lowers the energy barrier for vortex entry, which is a primary cause of dark counts in SNSPDs. Thermally activated vortices can lead to an increase in dark count rate, further degrading performance by decreasing the signal-to-noise ratio [113]. The impact of current crowding becomes more pronounced in 42 meanders with higher fill factors. To mitigate these effects, several strategies can be employed. One approach is to use bend thickening techniques, which help reduce the localized current density in the bending regions [113, 114]. Additionally, the design of the bends can be optimized to minimize cur- rent crowding and the associated enhancement of dark counts. Specifically, sharper bends with smaller inner radii tend to degrade performance, while rounded bends with smoother transitions are more effective [109, 111]. Another technique to improve current crowding effects involves helium ion irradiation of the straight segments of the meander [121]. Nanowire vs Microwire Detectors In 2020, single photon sensitivity in micrometer-wide superconducting wires was first demonstrated in small microstrip devices [122]. Near-IR photons were detected in a device made from a 3.3 nm-thick MoSi film with 2 µm-wide wires. This led to increased interest in developing microwire devices with non-trivial active areas and an exploration of new materials. Near-IR photon sensitivity in wires as wide as 4 µm and with active areas as large as 1 mm2 have since been reported [123], with similar results seen in NbN devices[124]. 2.6 SNSPDs for Dark Matter Detection and HEP Applications SNSPDs are the single-photon detector of choice for quantum information science and technology applications. However, their unique combination of high timing resolution, low dark count rates, and a response over a broad range of wavelengths makes the technology attractive for HEP applications. In particular, SNSPDs can be used in low-mass DM ex- periments where the primary sensor requirements are (1) (vanishingly) low intrinsic dark count rates, (2) a high photon detection efficiency and (3) low energy thresholds for infrared single photon detection. For these metrics, SNSPDs are the highest performing detectors 43 available in separate demonstrations. To mature the technology for DM detection, it is nec- essary to combine these performance metrics into one device and understand the inherent trade-offs. This is a central focus of the work contained within this thesis. Additionally, it is important to investigate the fundamental limits of the aforementioned detection metrics for SNSPDs. There are now multiple pilot and proof-of-concept phase low mass dark matter detec- tion experiments using SNSPDs in the detection scheme. The development of large-active area SNSPDs to instrument n-type GaAs targets is one of the primary focuses of the work contained in this thesis. Light A’ Multilayer Periodic Optical SNSPD Target (LAMPOST) is a dark photon search targeting DM in the eV mass range, via coherent absorption in a multilayer dielectric haloscope [115]. The layered dielectric target is designed to capture the conversion of dark photons to regular photons which are then focused onto an SNSPD. In this detection concept, the sensitivity is ultimately governed by the limits of the SNSPD and in principle as the energy thresholds of SNSPDs are lowered, versions of this exper- iment that probe lower mass dark photon DM can be implemented to constrain more of the parameter space at lower masses. Of interest to this future direction are mid-infrared SNSPDs whose current status and challenges are summarized in the next section. The Broadband Reflector Experiment for Axion Detection (BREAD) experiment fea- tures a cylindrical metal barrel to convert dark matter into photons and parabolic reflectors to focus signal photons onto a photosensor [125, 126]. This science program contains two pilot experiments. GIGABread is the RF pilot axion and dark photon experiment targeting DM in the tens of GHz regime and focuses them onto a coaxial horn antenna. INFRABread is the program’s pilot dark photon search at IR frequencies. The detectors developed and characterized in this thesis are being deployed as the photo sensors for this pilot experiment. More on the current status of this implementation is described in chapter 8. In addition to their use in low mass DM searches, the mm2-scale arrays from this work 44 are also large enough to be potentially interesting technology for broader HEP applications. One of the 8 pixel 2 mm x 2mm SNSPD arrays of this work was deployed in the Fermi- lab Test Beam Facility and exposed to various high energy particles [127]. The results are d