id
string
text
string
image
image
230406165/6
Jennifer K. Bertrand and Craig S. Chapman Figure2: Mapping of one representative participant’s raw eye-data to target AOIs. Raw eye-data samples (small circles) were clustered into 4 bins and, based on their spatial configu- ration, bins were assigned to one of the 4 target AOIs (dashed boxes). Individual samples are color-coded by their assigned bin: light green -> NRAOI, dark green -> NLAOI, light blue -> FRAOI and dark blue -> FLAOI. Raw-data bin-cluster cen- troids are shown in corresponding colors as filled circles with black borders. The raw data clearly groups into four clusters but the cluster centroids do not align with the actual task space (skewed down). Our analysis therefore relies on when the eye-data was within each cluster, not the actual space it occupied. [27] and other real-world examples ([49,52]) to segment our data, where thresholds were applied to the cursor and circles’ velocity and AOI-proximity as a means to define the onset and offset of the Reach, Transport and Release phases (see Supplementary Materials for additional details and threshold values). ## Data Analysis 3.8.1 ## Dependent Measures . We had two primary motivations: 1) to determine if the quality of webcam eye-tracking data would be sufficient to explore eye-cursor coordination dynamics for our specific 2D UX context and, if so, 2) to explore, in a preliminary way, if the $\sim$ 500 ms of fixation time around a manual interaction would be preserved in our specific 2D UX context, despite drastic differences in the physics and style of control. With respect to 1), it is a hallmark of eye-hand coordination during object interaction that, even though not directly instructed, participants look almost exclusively at task relevant targets (namely the object they are going to interact with and the locations where they are going to move it to and from). As such, our first dependent measure examines, for each target location in the task (4 total, see subsection 3.3 - Digital Task Layout), the average time spent fixating that location when it was relevant to the current movement (i.e. a Pick-up or Drop-off location) and the average time spent fixating that location when it was irrelevant to the current movement (i.e. one of the two targets on every movement that are not a Pick-up or Drop-off location). AverageFixationDuration(ms). Across a given trial, the average time spent fixating on one of the AOIs (NRAOI, NLAOI, FRAOI or FLAOI, see Figure 1 A) when that AOI was a relevant location (a Pick- up or Drop-off location for the current movement) or not. Across the 8-movement sequence, each AOI was relevant and irrelevant an equal number of times. With respect to 2), fixation time around an interaction consists of two values - how long the eyes are on an object prior to interaction (eye-arrival latency) and how long the eyes linger on an object after an interaction is initiated (eye-leaving latency). Our task involves object manipulations with two interaction events, the Pick-up and the Drop-off. Therefore, we examine the eye arrival and eye leaving latencies for both of these events. Importantly, it has previously been reported that eye arrival and leaving latencies are not absolute ([27]), but can flexibly change based on the demands of the task in general and the durations of each constituent movement and phase in specific. Therefore, to test for these possible within-task adaptations, we also examined eye latencies and each movement in terms of the durations of the Reach, Transport and Release phases. Based on this motivation, we extracted and analyzed the follow- ing measures per movement: PhaseDuration(ms). The time spent in each phase (Reach, Trans- port, Release). Eye-arrivallatencyatPick-upandDrop-off. Eye-arrival latency (EAL) at Pick-up was defined as the difference between Transport start time and the time of the eye’s arrival at the Pick-up location. EAL at Drop-off was defined as the difference between Transport end time and the time of the eye’s arrival at the Drop-off location. Eye-leavinglatencyatPick-upandDrop-off. Eye-leaving latency (ELL) at Pick-up was defined as the difference between Transport start time and the time of the eye leaving the Pick-up location. ELL at Drop-off was defined as the difference between Transport end time and the time of the eye leaving the Drop-off location. 3.8.2 ## Statistical Procedure . Each dependent measure was analyzed in Jamovi (Version 2.2.5; an open-source statistical software) using a two-factor repeated-measure analysis of variance (RMANOVA). If a two-way interaction was revealed from the omnibus RMANOVA, follow-up single-factor RMANOVAs were performed to test the simple main effects of one factor at all levels of the other factor. Sig- nificant main effects were explored with all pairwise comparisons. All reported RMANOVA p-values include a Greenhouse-Geisser correction for violations of sphericity, and all follow-up pairwise comparisons are fully reported in Supplementary Materials (with Bonferroni-corrected p-values). ## RESULTS ## Online, webcam eye-tracking can be a suitable method for quantifying gaze behaviours As described earlier, our data-driven definition of AOIs leaves us vulnerable to circularity in our analyses. To address this poten- tial criticism, here we look at task relevant timing to check if our approach is valid. Since our AOI clustering is agnostic to timing, any effects of spatial gaze distribution across time that match task demands provide solid evidence for the utility of our approach.
230406165/4
Jennifer K. Bertrand and Craig S. Chapman Figure1: A) The digital task layout. The full screen (800 x 450 pixels) is depicted, where a purple Home area and a Restart button are located in the bottom right corner outside the black-bordered task area. The target Areas of Interest (AOIs) are split into those at the top (blue, Far Left and Right) and those at the bottom (green, Near Left and Right). Participants dragged-and-dropped circles (white, one Near, one Far) between target AOIs. B) A still image of the real-world Cup-Transfer Task we based our digital object interaction task on (from [27]). C) The task sequence. A trial involved 8 Moves, with the order and direction of circle movements shown with numbered arrows. The cursor started in the Home area to begin the trial and returned to the Home area after Moves 2, 4, 6, and 8. D) The segmentation of a single object interaction (i.e. one Move) into its events and phases. The top of the panel shows the velocity traces of the circle (white) and cursor (black). The onset of the circle movement was detected as a combination of these velocities and the distance between the cursor and the relevant target AOIs (bottom panel). Together, this defined the Pick-up and Drop-off events, which in turn defined the Reach (green), Transport (light blue) and Release (dark blue) phases. on these key time points. Our 8-movement sequence (see Figure 1 C) is as follows: with the cursor always beginning at Home, Move 1 was a Pick-up of the Near circle at NRAOI, with its Drop-off at NLAOI. Immediately after, the Far circle was picked up from FRAOI and was transported to its Drop-off at FLAOI (Move 2). The cursor then returned to the Home position (as was required after every two object interactions). Moves 3 and 4 were a reflection of the first movements: the Far circle was picked up from FLAOI and moved back to FRAOI, and the Near Circle was picked up from NLAOI and moved back to NRAOI. After the cursor returned to Home, the Far circle was transported from FRAOI to FLAOI (Move 5), and then the Near circle moved from NRAOI to NLAOI (Move 6). This pattern was again reflected after the cursor visited Home, with Move 7 the pickup of the Near circle from NRAOI to drop off at NLAOI, and Move 8 the pickup of the Far circle from FLAOI to FRAOI. The cursor returned to Home to end the trial. ## Procedure Prolific (www.prolific.co) participants were provided a study link and a detailed study description that included an estimate of the study’s duration (1 hour), the hardware requirements, and instruc- tions for avoiding technical complications (included in Supplemen- tary Materials). Clicking the study link launched the full-screen Labvanced window and requested webcam device permission. Par- ticipants failing to meet the minimum requirements would receive an error or warning message immediately. Barring no issues, par- ticipants would first read a consent form and provided they gave their informed consent, would then answer a brief demographic and hardware survey. Next, participants would proceed self-paced through the task instructions. Following online research best practices (e.g. [14]), we developed extensive instructions including an instructional video (see Supplementary Materials), and gave task directions in a way that required participant engagement. Participants were informed that the task was a screen-based version of a real-world task. They were shown a picture of the real-world task (see Figure 1 B) and told that the circles in their task were to be thought of as two- dimensional cups. Finally, participants were encouraged to use favourable lighting conditions and were instructed about the use of a virtual chinrest feature, strategies previously shown to improve webcam eye-tracking data quality [45]. A 5 minute Labvanced eye-tracking calibration followed the instructions, and participants were required to repeat the calibration if the predicted gaze error exceeded 7% of the screen size. Lastly, participants completed one guided (click-through) practice trial, and then a second unguided practice trial with time-delayed hints (i.e. only shown if participant paused for three seconds or longer). Participants could repeat the unguided practice trial as many times as they wanted to ensure they understood the prescribed sequence of movements (1.2 unguided practice trials completed on average). Participants would then complete the 50 self-paced experimental trials. Every 10 trials, they would receive an update about how many trials they had completed. A brief, 7-point eye-tracking re- calibration was performed every 5 trials, enabling the use of Lab- vanced’s adaptive drift correction feature. After completing the 50 trials, participants were offered a long-form text input field to provide study feedback (if any) and thanked for their time. The study then concluded, with the browser exiting fullscreen mode, and participants receiving compensation via Prolific. The entire experimental procedure, as a Labvanced study, can be accessed via the link in Supplementary Materials.
230406165/10
Jennifer K. Bertrand and Craig S. Chapman our task-relevant AOIs were static and spatially distributed. Im- portantly, this means this approach will not be as successful or even possible for tasks with dynamic AOIs or AOIs that are close together. We further discuss these and other important limitations in Section 6 - Limitations and Future Directions, below. Our investigation of gaze distribution (see Figure 3 for a temporal assessment, and Supplementary Materials for a spatial assessment) gave us sufficient confidence in the quality of the eye-tracking data to pursue our primary question of whether or not digital eye- cursor coordination in this task would follow similar patterns to real-world eye-hand coordination. Examining visuomotor coordi- nation in tasks designed to promote natural, self-paced behaviours requires the task first be broken into its constituent interactions and then that those interactions be broken into the discrete phases of interaction. Adopting the segmentation strategy employed by Lavoie et al. [27] we identified an important distinction between real and digital object interactions: despite both forms of movement being self-paced, digital objects are transported in about 200 ms (see Figure 4), 4 to 5 times faster than real objects are moved in the real world. Given these movements, it was impossible that the exact pattern of eye-hand coordination observed in the real world would be preserved during digital interactions. That is, during real- world interactions, when transporting an object between locations, the eye will “look ahead” to the drop-off site about 500 ms before the hand and object arrive. But, as just explained, during digital interactions, the entire transport lasts about 200 ms, meaning the eye cannot look ahead in the same way. Remarkably, our results suggest that the visuomotor system pre- serves the overall interaction eye-dwell time of at least 500 ms by flexibly adapting the pattern of fixations. Specifically, we quantified eye-cursor latencies around both the Pick-up (cursor clicked to start dragging) and Drop-off (release of cursor click to stop dragging) events. At each event, we calculated how long the eye was looking at the location prior to the event (eye arrival latency, or EAL) and how long the eye remained looking at the location after the event (eye leaving latency, or ELL). As depicted in Figure 4, for a digital Pick-up, the pattern of gaze is almost identical to a real-world inter- action: the eyes arrive at the location 400-500 ms before and stay for about 100 ms after. Given the speed of the Transport phase, eye latencies during digital Drop-off are significantly different from the real world. The eyes arrive around 100 ms before the event, but surprisingly, linger for 400-500 ms after the digital object has been released. With consideration to the preliminary nature of our inves- tigation, we take this as the most important finding in our study: gaze allocation during a digital, self-paced, drag-and-drop object interaction flexibly adapts to the drastically different mechanics of movement to ensure about 500 ms of visual information is acquired from the beginning and end of each interaction movement. Thus, this perfectly aligns with the take-home message of real-world interactions, but the route by which it is achieved is significantly different. Given this initial evidence for the persistence of $\sim$ 500 ms of eye fixation towards objects we interact with across real and digital domains (for at least this specific context), what insights might this offer for UX design principles? First, it suggests that as particular digital experiences are being designed, there may be a fundamental lower limit on the pacing with which interactions can occur while respecting the natural cadence of visuomotor coordination. For example, a drag-and-drop movement will take at least one second if executed at a natural speed. A designer wanting to push these movements to be faster might consider rearranging the target lo- cations to make them spatially contiguous, possibly allowing the Pick-up and Drop-off information to be gathered by a single fixation. On the flip side, a designer requiring particularly precise cursor interactions might consider spatially separating the targets, letting the extra distance provide additional time for targeting fixations to occur within the natural rhythms of the task. In both of these cases, these preliminary findings support the notion that a design achieves maximum efficiency not by being the fastest, but rather, by aligning with the demands of a visuomotor system that evolved to optimally coordinate movements [6] on its own time scale. A second design principle which is more indirectly revealed via this introductory study is the role that feedback plays in facilitating successful interactions. Unlike in the real world where a person using their body to interact with an object typically receives haptic feedback about the interaction, digital interactions rely almost ex- clusively on visual feedback for confirmation that an interaction is proceeding successfully. In the current study we attempted to boost the visual cues associated with interaction by changing the visual properties of targets based on the cursor position. But, we believe there is more work to be done exploring how additional modalities could move digital interactions toward their real-world counterparts. For example, adding sound cues relevant to interac- tion, or even more sensitive and dynamic visual cues on objects that are successfully being interacted with may liberate the eyes to move in an even more natural fashion. Some work in high leverage interactions like laparoscopic surgery [37] has shown the utility of this approach. As real and digital worlds become less siloed and share elements across spaces (e.g. virtual and augmented reality), we predict that these mixed interactions will also obey the $\sim$ 500 ms of viewing time and so too will benefit from the exploration of multimodal feedback cues. ## LIMITATIONS AND FUTURE DIRECTIONS Our goal was to measure more ecologically-valid user experiences from a diverse population of participants using their own digital devices in familiar environments. Thus, we collected eye-tracking data from webcams in remotely recruited participants. This meant we sacrificed some experimental control and introduced more eye- tracking noise, leading to a number of notable limitations. Here we describe some of these limitations and for each, offer a future direction for how to test and improve the study. First, we provide no formal validation of the accuracy of the webcam eye-tracking system. While we attempted to quantify the functional accuracy in both time (see Figure 3) and space (see Sup- plementary Materials) a future approach would be to conduct the same experiment under controlled lab conditions while simultane- ously recording both webcam and lab-grade, high-resolution eye- tracking data. Of course, a shift to the lab would also remove some of the environmental confounds of remote participation (lighting, hardware differences etc.) and would therefore provide a best-case measure of the magnitude of accuracy difference between webcam and lab-grade systems. Thus, given the known accuracy reduction
230406165/1
# Dynamics of eye-hand coordination are flexibly preserved in eye-cursor coordination during an online, digital, object interaction task Jennifer K. Bertrand Craig S. Chapman Do patterns of eye-hand coordination observed during real-world object interactions apply to digital, screen-based object interac- tions? We adapted a real-world object interaction task (physically transferring cups in sequence about a tabletop) into a two-dimensional screen-based task (dragging-and-dropping circles in sequence with a cursor). We collected gaze (with webcam eye-tracking) and cursor position data from 51 fully-remote, crowd-sourced participants who performed the task on their own computer. We applied real-world time-series data segmentation strategies to resolve the self-paced movement sequence into phases of object interaction and rigor- ously cleaned the webcam eye-tracking data. In this preliminary investigation, we found that: 1) real-world eye-hand coordination patterns persist and adapt in this digital context, and 2) remote, online, cursor-tracking and webcam eye-tracking are useful tools for capturing visuomotor behaviours during this ecologically-valid human-computer interaction task. We discuss how these findings might inform design principles and further investigations into nat- ural behaviours that persist in digital environments. eye-tracking, cursor-tracking, quantitative methods, eye-cursor coordination, object interaction Jennifer K. Bertrand and Craig S. Chapman.2023. # Dynamics of eye-hand coordination are flexibly preserved in eye-cursor coordination during an online, digital, object interaction task.In Proceedingsofthe2023CHIConfer- ## INTRODUCTION At the core of all well-considered user experiences is the user them- selves. So-called human-centered designs incorporate behaviour, cognition, and perception into their product. An early example is the psychophysical mapping of human sensitivity to flickering light. Over hundreds of years, scientists learned that a light display re- freshing at a minimum of 30 Hz appeared continuous to the human eye - setting the benchmark for early computer screens. Of course, perceiving the world is only part of a user’s experi- ence - they also interact with their environment to achieve their goals. Like the display refresh-rate example, principles of human interaction can dictate good design. Fitts’ Law [11] is one such find- ing that is now adopted as a design principle in human-computer interaction (HCI) [30,46]. In this seminal work, Fitts quantified the speed-accuracy tradeoff for movements of different amplitudes (how far you need to move) to targets of different sizes [11]. Put suc- cinctly, he showed a lawful relationship whereby larger amplitude movements and smaller targets both result in longer movement times. These real-world findings have since been explored in depth in an HCI context [30], informing and assessing the design of two- [31] and three-dimensional [15] pointing devices, the soft (virtual) QWERTY keyboard [32,51], and the properties and placement of interactive web elements [22,29,33,42]. Critically, user experience (UX) is best thought of as a dynamic cycle of perception and action whereby the information we need to guide our upcoming actions is informed by where we look; how we move then shapes the environment causing changes in the perceptual experience. Consider, for example, the coordinated effort required of the visual and motor systems to safely pick up a cup full of hot coffee. Before any movement, the eyes will fixate the drink, leading the action of the hand by about 500 milliseconds. Seamlessly, as soon as the mug is grasped, the eyes will move to look at the next object for action, like a sugar packet, well before the hand is finished moving the hot drink [26]. These patterns of visuomotor coordination are ubiquitous and stereotypic, appearing in human and non-human primates [1,36] alike. Here we ask, in the same way real-world findings have informed computer screen and keyboard design, can principles of real-world eye-hand coordination help inform UX design for digital interactions? To approach this question, we used an online, eye-tracking- enabled platform (Labvanced; [10]) to create a screen-based version of a real-world object interaction task [27]. Instead of moving cups to targets on a table, crowdsourced participants (N = 51) dragged circles to targets on their computer screen while we recorded cursor movements and webcam eye-gaze coordinates. We had two primary motivations - first, to explore if the quality of the webcam gaze data (and subsequent processing procedures) would be sufficient
230406165/3
than the cursor. However, the occurrence of or need for coordina- tion varied by target-type: the gaze led the cursor $\sim$ 30% of the time for ‘TitleBar’ clicks, yet ‘List’ clicks were gaze-first more than 85% of the time [28]. In these unconstrained tasks we find additional evidence to support that the eyes lead the cursor, but also see im- plications for the limitation of this approach when faced with real digital interaction complexity. In the current study, we attempt to strike a balance between ecological validity and experimental con- trol, focusing on a prescribed digital interaction sequence without imposing any constraints on how (e.g. where to look, how fast to move etc.) the sequence should be completed. ## Webcam Eye-tracking One limitation of almost all of the aforementioned studies is that they occur in the lab. A principle of human-centered design is not only to focus on the user but also to consider the environment and context in which their experience is happening (ISO 9241-210:2010). One recent technological advancement that might make it possible to study visuomotor coordination in more authentic environments (e.g. users in their own homes on devices they regularly use) is to use webcam data to derive estimates of screen-based gaze be- haviour. However, webcam eye-tracking has struggled to establish its utility as a research tool. The reasons are numerous: webcam eye-tracking has a much slower sampling rate ($\sim$ 10 Hz compared with 100+ Hz in lab [5,14,45]), many users are unable to partic- ipate due to a slow internet connection or insufficient hardware [5,14], uncontrolled lighting can significantly decrease data quality [12,45,54] and, even with optimal conditions, extensive time must be spent calibrating the system (up to 50% of the study duration as in [45]). Despite these limitations, recent advances, especially in using machine learning to predict gaze location (e.g. Labvanced v2 High Sampling Mode eye-tracking; [10]), offer a path forward, especially where spatial accuracy is the most important feature of the data [45,53]. Further, researchers have shown that focusing on fixations to the most relevant areas (i.e. areas of interest, or AOIs) is a reasonable approach for eye-tracking data [19]. Therefore, a major motivation of the current study was to explore whether state- of-the-art webcam eye-tracking algorithms [10] combined with participation criteria (e.g. processing speed) and AOI-based cluster- ing and analyses would provide sufficient data quality to explore eye-cursor coordination. ## METHODS ## Participants 51 adults provided their informed consent to participate in the ex- periment. Of these, 14 participant datasets were rejected for unsal- vageable eye data, and 8 participant datasets were rejected for low trial count (<50%) after removing trials with procedural or technical errors (see subsection 3.6 - Data Processing and Supplementary Materials for complete data cleaning procedure). The remaining 29 participants (12 female, 1 undisclosed gender; Age: M = 27.07, SD = 10.75) were 26 right-hand users and 3 left-hand users. All partici- pants had no prior knowledge about the experiment or its objective. All experimental proceedings were approved by the University of Alberta’s Research Ethics Board (Pro00087329) and were performed in accordance with relevant guidelines and regulations. All par- ticipants were recruited using the online crowdsourcing platform Prolific (www.prolific.co) and were paid for their time (6 GBP per hour, $\sim$ $ 10 CAD per hour). ## Materials All participant data was collected online using Labvanced [10], a browser-based Javascript experimentation platform. The Labvanced platform offers built-in webcam eye-tracking (Labvanced v2 High Sampling Mode eye-tracking [10]) and can record the position of the cursor across time. It was necessary to impose some minimum requirements to achieve stable data collection: only laptop (n = 19) or desktop (n = 10) computers with an audio output (headphones or speakers); only Mac (n = 3), Windows (n = 26) or Linux (n = 0) operating systems and Chrome browser; a webcam with a minimum resolution of 1280 x 720 pixels; a landscape-oriented screen with a minimum of 600 x 600 pixels (Mode = 1920 x 1080 px) and a system (including internet connection) capable of collecting at least 10 samples per second of the head’s position for optimal eye-tracking precision (M = 14.5 Hz, SD = 4.3 Hz). ## Digital Task Layout We modeled our digital task layout (see Figure 1 A) to mirror Lavoie et al.’s Cups Task apparatus [27] (see Figure 1 B), which featured a short-walled table-top surface with a midline partition, two cups, 4 AOIs, a Home area, and a fixation sphere. In the real world, partici- pants stood next to the counter-height apparatus, looking down on the surface. Thus, we designed our screen-based version to appear like a flat, bird’s eye view of the real-world task. All real-world elements were proportionally scaled to a 800 x 450 pixel frame in Labvanced (and later, automatically scaled by Labvanced to each participant’s screen resolution). Like the real-world version, the Far Left and Right AOIs (FLAOI/FRAOI) were colored blue, the Near Left and Right AOIs (NLAOI/NRAOI) were green, and the Home area was purple. The real-world white paper cups filled with white beads were modeled as white circles. As a proxy for haptic feedback, we designed the circles and Home area to be responsive to cursor hover by darkening in color whenever the cursor landed within their bor- ders. Finally, select elements were introduced to the digital task space in service of loosely replicating the real-world, experimenter- guided experience: a restart button was always available should the participant realize they made a movement sequence error, text appeared with instructions if the participant took more than three seconds to move into their starting position, and a highlighted bor- der appeared around the Home area to mark the important start and end events of the trial. ## Task Our task was an adaptation and extension of an established object interaction task from the real world [27]. We doubled the number of object interactions within a sequence, and transformed the task to a digital, screen-based version. Critically, Lavoie’s real-world task was designed to be segmented in time and space to allow for an examination of eye-hand coordination measures around the critical Pick-up and Drop-off interaction events [27]. By adopting a similar structure, our analysis also relies on segmentation centered
230406165/9
at the Pick-up location earlier - this kind of within-trial flexibility is also observed in the real world [27]. Full reporting of the pairwise comparisons is available in Supplementary Materials. Together, these EAL findings suggest: 1) similar to real-world interactions, during digital Pick-up interactions the eyes arrive about 400-500 ms before the cursor starts to move the object, and 2) unlike real-world interactions, during digital Drop-off interactions the eyes only arrive about 100-200 ms before the dragged object, reflecting the stark differences in the duration of digital versus physical object Transport. For ELL there were significant main effects for both Movement (F (2.93, 82.01) = 22.7, p <.001) and Interaction Site (F (1, 28) = 340.2, p <.001), and also a significant two-way interaction (F (4.62, 129.38) = 15, p <.001). Follow-up simple main effect RMANOVAs compared Pick-up and Drop-off ELLs across the 8 movements and both were significant (Pick-up: F (2.88, 80.7) = 8.2, p <.001; Drop-off: F (3.6, 100.91) = 24.3, p <.001). Despite statistical differences, the timing of the eye leaving the Pick-up location is quite stable and short, ranging from $\sim$ 35-135 ms. Where Pick-up ELL does vary, it appears to change as a function both of the length of the upcoming transport and as a push-pull with the preceding eye arrival latencies. As an example, Movement 6 has a relatively long Transport phase and comparatively short preceding EAL - this results in it having the longest Pick-up ELL. Full reporting of the pairwise comparisons is available in Supplementary Materials. The most surprising result from our study, and what stands out as the biggest difference from real-world eye-hand coordination, is how long our participants spend looking at an object after they have dropped it off. Here, ELL at Drop-off exceeds 400 ms in all cases and is often more than 600 ms. This is drastically different from the real- world Drop-off ELLs which only range from 140-250 ms [27]. This important finding demonstrates that participants compensate for abbreviated digital Transports by having their eyes remain fixated for longer at the location where the object is dragged to. This prolonged Drop-off ELL also scales with duration of the Release phase, with Movements with longer Release phases also showing the longest ELLs. This relationship suggests that the compensatory prolongation of the Drop-off ELL may in part relate to the planning of the next movement following a Release. Full reporting of the pairwise comparisons is available in Supplementary Materials. Together, our ELL findings further inform the nuances of eye- cursor coordination in digital interactions: 1) like real-world Pick- ups, the eyes wait until the Pick-up happens then quickly leave, and 2) unlike real-world Drop-offs, the eyes dwell at the Drop-off site well beyond the end of the Transport. ## DISCUSSION In this preliminary investigation, we show that eye-cursor coor- dination during a specific form of digital object interaction obeys constraints similar to eye-hand coordination during real-world interactions. Specifically, we find the eye dwells on or near the site of a digital interaction for at least 500 ms, almost identical to the amount of time others report in real-world interactions with physical objects. This initial finding demonstrates the potential util- ity of webcam eye-tracking collected from online, remote, crowd- sourced participants as a tool for capturing rich, meaningful, and ecologically-valid visuomotor data. Our study was designed to make the comparison of digital to real-world interactions as valid as possible. Thus, we adapted a previously reported real-world task [27] where the 500 ms mini- mum dwell time per interaction had previously been quantified. In our digital adaptation of this real-world cup-transfer task, we asked crowdsourced, online participants to perform 50 trials of an 8-movement drag-and-drop sequence (see Figure 1). While perform- ing this digital interaction task, participants’ cursor and gaze posi- tions on the screen were monitored via the Labvanced experiment platform ([10]) using their own computer webcams. As described throughout this study, there are significant challenges to collecting webcam eye-tracking data, and as such, a major objective for this project was to assess its feasibility as a tool for quantifying patterns of dynamic eye-cursor coordination. Tentatively, and with the caveat that substantial preprocessing was required, we believe that the eye data quality in this task was sufficient to explore eye-cursor coordination for this specific digital context. First, as reported by other research groups collecting online data (i.e. [44,45,54]), we experienced high rates of data exclusion (>40% of participants were not included in analysis, predominantly due to eye-data quality issues) even though we imposed restrictions on the hardware and internet connection of eligible participants. Second, for the participants who were included in the analysis, the eye data required extensive processing. This included reducing the spatial dimensionality from the (x,y) coordinate frame of the screen to 4 data-driven AOI bins (see Figure 2 for a representative subject), which were then mapped to the 4 task-relevant interaction loca- tions. Given this approach departs from conventional eye-tracking analysis, we confirmed its sensitivity by testing if the distribution of gaze to each of these task-relevant AOIs followed the predictions imposed by task demand. Specifically, we show that participants fix- ated more on interaction locations that were relevant to the current movement (i.e. the target where you were clicking an object and the target where you were dragging it to) than to locations that were not relevant to that movement (see Figure 3, and Supplementary Materials for the complementary spatial investigation). It is important to acknowledge that our successful collection and preliminary validation of webcam eye-tracking is in and of itself a significant contribution. With the uncontrolled nature of online tasks, and a technology that relies on consumer-grade hard- ware, there are many opportunities for noise or error to prevent successful data collection. We worked hard to minimize dropout due to hardware and internet issues by imposing very clear require- ments, stated during crowdsourcing and checked during the study initialization. Then, we spent considerable time developing clear and transparent instructions to assist with participant retention. We provided detailed information about potential privacy concerns as well as video and interactive walk-through demonstrations of the task to promote participant understanding (see Supplementary Materials). As described above, our data was then processed using a k-means clustering technique. While clustering the eye-data was a successful approach for this study, it succeeded in part because
230406165/8
Jennifer K. Bertrand and Craig S. Chapman Figure4: Interaction Phase durations and Eye Latencies at Pick-up and Drop-off across the 8 movements. Each movement is shown as a horizontal bar centered on the middle of the Transport (light blue) phase. Movements start with a Reach (green) as the cursor moves toward the target circle, transition to a Transport as the circle is moved, and end with a Release (dark blue) as the cursor moves away from the circle. The start of Transport is the Pick-up event and the end of Transport is the Drop-off event. White circles show the time the eye arrives (EAL) before each event while black circles show the time the eye leaves (ELL) following an event. Dark green lines connect the Pick-up EAL and ELL while Dark blue lines connect the Drop-off EAL and ELL. Despite the short Transport phases, the eye latencies adapt to ensure at least 500 ms of fixation time around each interaction event. EAL and ELL we ran an 8 x 2 (Movement x Interaction Site [Pick-up / Drop-off]) RMANOVA. For EAL there were significant main effects of Movement (F (3.43, 96.09) = 14.28, p <.001) and Interaction Site (F (1, 28) = 251.6, p <.001), as well as a significant interaction between the two factors (F (3.97, 111.2) = 8.37, p <.001). Follow-up simple main effect RMANOVAs compared Pick-up and Drop-off EALs across the 8 movements. Drop-off EALs were remarkably consistent, showing no effect of Movement, demonstrating that gaze consistently arrives at a Drop- off location just over 100 ms before the clicked-and-dragged object. Pick-up EALs did show a significant effect of Movement (F (3.87, 108.42) = 27.3, p = <.001) which aligns with the duration of the Reach phase in which the Pick-up occurred. That is, for movements with a longer Reach phase (e.g. Movements 3, 5, 7) the eye arrives
230809896/5
Fig.2. Schematic description of the proposed network.
230809896/7
. Specifically, we design the attention-based cross module to generate a cross-attention matrix as: $$ \begin{split} \tilde{e} = W_{v} \cdot e^{(v)} + W_{t} \cdot e^{(t)} + b_{e}, \\ E = ScaledDot(e^{(f)}, \tilde{e}) = \frac{e^{(f)} \cdot \tilde{e}^{\rightarrowp}}{\sqrt{d}}, \end{split} $$ where $E \in \mathbb{R}^{Nd \times Nd}$ is the cross-attention matrix that corresponds to the scaled- dot similarity. We then apply a 1D convolutional layer to the cross-attention matrix $E$ as: $$ \begin{split} \alpha = Softmax(Conv(E)), \end{split} $$ where $Conv$ is the 1D convolutional layer and $\alpha$ is the cross-attention score matrix. We integrate $\alpha$ and $\tilde{e}$ into a weighted representation $e$ as: $$ \begin{split} e = \alpha \odot \tilde{e}. \end{split} $$ Given a batch of patients, the embedding for them can be written as: $$ \begin{split} e = [e_{1}, e_{2}, \cdots, e_{B}] \in \mathbb{R}^{B \times Nd \times T}, \end{split} $$ where $B$ is the batch size. Since $e$ still takes the form of sequence data, we design an attention layer to generate a series of attention weights $(\beta_{1}, \beta_{2}, \cdots, \beta_{T})$ and reweight these weights to produce an overall feature representation as: $$ \begin{split} \beta = Softmax(e \cdot W_{e} + b_{e}), \\ \bar{e} = \sum^{T}_{t = 1} \beta_{t} \odot e_{t}, \end{split} $$ where $\bar{e} \in \mathbb{R}^{B \times Nd}$ is the new generated patient representation. ## Similar Patients Discovery and Information Aggregation Before con- ducting patient similarity calculation, we encode $X_{base} \in \mathbb{R}^{g}$ as $e_{base} \in \mathbb{R}^{d_{g}}$ and concatenate $e_{base}$ with $\bar{e}$ as: $$ \begin{split} e_{base} = W_{base} \cdot X_{base} + b_{base}, \\ e^{\prime} = Concate(\bar{e}, e_{base}), \end{split} $$ where Concate is the concatenation operation. For the batch of patient representations, the pairwise similarities that corre- spond to any two patient representations can be calculated as: $$ \begin{split} \Lambda = sim(e^{\prime}, e^{\prime}) = \frac{e^{\prime} \cdot e^{\prime}}{(Nd+d_{g})^{2}}, \end{split} $$
230809896/12
Table 1. Performance of our approaches with other baselines on clinical time series imputation and in-hospital mortality prediction. \begin{tabular}{crrcc} \rightarrowprule MIMIC-III/24 hours after ICU admission & \multicolumn{2}{c}{Clinical time series imputation} & \multicolumn{2}{c}{In-hospital mortality prediction} \midrule Metrics & \multicolumn{1}{c}{MAE} & \multicolumn{1}{c}{MRE} & AUROC & AUPRC \midrule GRU-D & \multicolumn{1}{c}{1.3134(0.0509)} & \multicolumn{1}{c}{87.33\%(0.0341)} & 0.8461(0.0051) & 0.4513(0.0124) BRITS & \multicolumn{1}{c}{1.3211(0.0923)} & \multicolumn{1}{c}{87.92\%(0.0611)} & 0.8432(0.0040) & 0.4193(0.0144) %V-RIN & \multicolumn{1}{c}{} & \multicolumn{1}{c}{} & 0.8464(0.0057) & 0.4383(0.0054) GRUI-GAN & \multicolumn{1}{c}{1.6083(0.0043)} & \multicolumn{1}{c}{107.20\%(0.0029)} & 0.8324(0.0077) & 0.4209(0.0280) E$^{2}$GAN & \multicolumn{1}{c}{1.5885(0.0045)} & \multicolumn{1}{c}{105.86\%(0.0032)} & 0.8377(0.0083) & 0.4295(0.0137) E$^{2}$GAN-RF & \multicolumn{1}{c}{1.4362(0.0031)} & \multicolumn{1}{c}{101.09\%(0.0027)} & 0.8430(0.0065) & 0.4328(0.0101) STING & \multicolumn{1}{c}{1.5018(0.0082)} & \multicolumn{1}{c}{102.53\%(0.0047)} & 0.8344(0.0126) & 0.4431(0.0158) MTSIT & \multicolumn{1}{c}{0.3988(0.0671)} & \multicolumn{1}{c}{38.44\%(0.0647)} & 0.8029(0.0117) & 0.4150(0.0165) MIAM & \multicolumn{1}{c}{1.1391(0.0001)} & \multicolumn{1}{c}{75.65\%(0.0001)} & 0.8140(0.0044) & 0.4162(0.0079) Ours & \multicolumn{1}{c}{\textbf{0.3563(0.0375)}} & \multicolumn{1}{c}{\textbf{8.16\%(0.0086)}} & \textbf{0.8533(0.0119)} & \textbf{0.4752(0.0223)} Ours$_{\alpha}$ & \multicolumn{1}{c}{0.3833(0.0389)} & \multicolumn{1}{c}{8.78\%(0.0089)} & 0.8398(0.0064) & 0.4555(0.0139) Ours$_{\beta}$ & \multicolumn{1}{c}{0.4125(0.0319)} & \multicolumn{1}{c}{8.95\%(0.0077)} & 0.8417(0.0059) & 0.4489(0.0182) \midrule eICU/24 hours after eICU admission & \multicolumn{2}{c}{Clinical time series imputation} & \multicolumn{2}{c}{In-hospital mortality prediction} \midrule Metrics & \multicolumn{1}{c}{MAE} & \multicolumn{1}{c}{MRE} & AUROC & AUPRC \midrule GRU-D & \multicolumn{1}{c}{3.9791(0.2008)} & \multicolumn{1}{c}{52.11\%(0.0262)} & 0.7455(0.0107) & 0.3178(0.0190) BRITS & \multicolumn{1}{c}{3.6879(0.3782)} & \multicolumn{1}{c}{48.30\%(0.0726)} & 0.7139(0.0101) & 0.2511(0.0111) %V-RIN & \multicolumn{1}{c}{} & \multicolumn{1}{c}{} & 0.7351(0.0132) & 0.2936(0.0085) GRUI-GAN & \multicolumn{1}{c}{9.1031(0.0130)} & \multicolumn{1}{c}{119.29\%(0.0016)} & 0.7298(0.0094) & 0.3013(0.0141) E$^{2}$GAN & \multicolumn{1}{c}{7.5746(0.0141)} & \multicolumn{1}{c}{99.20\%(0.0018)} & 0.7317(0.0155) & 0.2973(0.0253) E$^{2}$GAN-RF & \multicolumn{1}{c}{6.7108(0.0127)} & \multicolumn{1}{c}{90.38\%(0.0015)} & 0.7402(0.0131) & 0.3045(0.0227) STING & \multicolumn{1}{c}{7.1447(0.0651)} & \multicolumn{1}{c}{93.56\%(0.0083)} & 0.7197(0.0154) & 0.2873(0.0182) MTSIT & \multicolumn{1}{c}{1.6192(0.1064)} & \multicolumn{1}{c}{21.20\%(0.0139)} & 0.7215(0.0071) & 0.2992(0.0115) MIAM & \multicolumn{1}{c}{1.1726(0.3103)} & \multicolumn{1}{c}{15.35\%(0.0406)} & 0.7262(0.0179) & 0.2659(0.0148) Ours & \multicolumn{1}{c}{\textbf{0.5365(0.0612)}} & \multicolumn{1}{c}{\textbf{7.02\%(0.0079)}} & \textbf{0.7626(0.0117)} & \textbf{0.3388(0.0211)} Ours$_{\alpha}$ & \multicolumn{1}{c}{0.6792(0.0716)} & \multicolumn{1}{c}{8.89\%(0.0093)} & 0.7501(0.0143) & 0.3325(0.0151) Ours$_{\beta}$ & \multicolumn{1}{c}{0.5923(0.0514)} & \multicolumn{1}{c}{7.75\%(0.0067)} & 0.7533(0.0104) & 0.3303(0.0175) \midrule MIMIC-III/48 hours after ICU admission & \multicolumn{2}{c}{Clinical time series imputation} & \multicolumn{2}{c}{In-hospital mortality prediction} \midrule Metrics & \multicolumn{1}{c}{MAE} & \multicolumn{1}{c}{MRE} & AUROC & AUPRC \midrule GRU-D & \multicolumn{1}{c}{1.4535(0.0806)} & \multicolumn{1}{c}{86.47\%(0.0482)} & 0.8746(0.0026) & 0.5143(0.0077) BRITS & \multicolumn{1}{c}{1.3802(0.1295)} & \multicolumn{1}{c}{82.21\%(0.0768)} & 0.8564(0.0040) & 0.4445(0.0189) %V-RIN & \multicolumn{1}{c}{} & \multicolumn{1}{c}{} & 0.8715(0.0050) & 0.5177(0.0076) GRUI-GAN & \multicolumn{1}{c}{1.7523(0.0030)} & \multicolumn{1}{c}{104.50\%(0.0018)} & 0.8681(0.0077) & 0.5123(0.0166) E$^{2}$GAN & \multicolumn{1}{c}{1.7436(0.0036)} & \multicolumn{1}{c}{103.98\%(0.0022)} & 0.8705(0.0043) & 0.5091(0.0120) E$^{2}$GAN-RF & \multicolumn{1}{c}{1.6122(0.0027)} & \multicolumn{1}{c}{102.34\%(0.0017)} & 0.8736(0.0031) & 0.5186(0.0095) STING & \multicolumn{1}{c}{1.6831(0.0068)} & \multicolumn{1}{c}{100.46\%(0.0035)} & 0.8668(0.0123) & 0.5232(0.0236) MTSIT & \multicolumn{1}{c}{0.4503(0.0465)} & \multicolumn{1}{c}{30.42\%(0.0314)} & 0.8171(0.0114) & 0.4308(0.0189) MIAM & \multicolumn{1}{c}{1.3158(0.0003)} & \multicolumn{1}{c}{78.20\%(0.0002)} & 0.8327(0.0024) & 0.4460(0.0061) Ours & \multicolumn{1}{c}{\textbf{0.4396(0.0588)}} & \multicolumn{1}{c}{\textbf{6.23\%(0.0073)}} & \textbf{0.8831(0.0149)} & \textbf{0.5328(0.0347)} Ours$_{\alpha}$ & \multicolumn{1}{c}{0.7096(0.0532)} & \multicolumn{1}{c}{8.85\%(0.0066)} & 0.8671(0.0093) & 0.5161(0.0151) Ours$_{\beta}$ & \multicolumn{1}{c}{0.5786(0.0429)} & \multicolumn{1}{c}{7.47\%(0.0056)} & 0.8709(0.0073) & 0.5114(0.0176) \midrule eICU/48 hours after eICU admission & \multicolumn{2}{c}{Clinical time series imputation} & \multicolumn{2}{c}{In-hospital mortality prediction} \midrule Metrics & \multicolumn{1}{c}{MAE} & \multicolumn{1}{c}{MRE} & AUROC & AUPRC \midrule GRU-D & \multicolumn{1}{c}{5.8071(0.2132)} & \multicolumn{1}{c}{44.53\%(0.0164)} & 0.7767(0.0141) & 0.3210(0.0182) BRITS & \multicolumn{1}{c}{5.5546(0.5497)} & \multicolumn{1}{c}{42.59\%(0.0421)} & 0.7285(0.0114) & 0.2510(0.0097) %V-RIN & \multicolumn{1}{c}{} & \multicolumn{1}{c}{} & 0.7541(0.0201) & 0.2955(0.0141) GRUI-GAN & \multicolumn{1}{c}{14.0750(0.0301)} & \multicolumn{1}{c}{107.96\%(0.0021)} & 0.7531(0.0167) & 0.2897(0.0201) E$^{2}$GAN & \multicolumn{1}{c}{12.9694(0.0195)} & \multicolumn{1}{c}{99.47\%(0.0015)} & 0.7605(0.0063) & 0.3014(0.0137) E$^{2}$GAN-RF & \multicolumn{1}{c}{11.8138(0.0161)} & \multicolumn{1}{c}{91.52\%(0.0011)} & 0.7763(0.0057) & 0.3101(0.0125) STING & \multicolumn{1}{c}{12.0962(0.0806)} & \multicolumn{1}{c}{92.79\%(0.0062)} & 0.7453(0.0182) & 0.2805(0.0190) MTSIT & \multicolumn{1}{c}{2.8150(0.2105)} & \multicolumn{1}{c}{21.58\%(0.0161)} & 0.7418(0.0091) & 0.3078(0.0120) MIAM & \multicolumn{1}{c}{2.1146(0.4012)} & \multicolumn{1}{c}{16.23\%(0.0414)} & 0.7574(0.0127) & 0.2776(0.0105) Ours & \multicolumn{1}{c}{\textbf{0.9412(0.0930)}} & \multicolumn{1}{c}{\textbf{7.21\%(0.0071)}} & \textbf{0.7907(0.0123)} & \textbf{0.3417(0.0217)} Ours$_{\alpha}$ & \multicolumn{1}{c}{1.1099(0.1064)} & \multicolumn{1}{c}{8.51\%(0.0081)} & 0.7732(0.0100) & 0.3311(0.0265) Ours$_{\beta}$ & \multicolumn{1}{c}{0.9930(0.0817)} & \multicolumn{1}{c}{7.61\%(0.0062)} & 0.7790(0.0117) & 0.3335(0.0178) \bottomrule \end{tabular}
230809896/2
predictive models and tools using machine learning techniques that would en- able healthcare professionals to make better decisions and improve healthcare outcomes. One of the EHR-based risk prediction tasks is to predict the mortality risk of patients based on their historical EHR data [8,29]. The predicted mortal- ity risks can be used to provide early warnings when a patient’s health condition is about to deteriorate so that more proactive interventions can be taken. However, due to a high degree of irregularity in the raw EHR data, it is challenging to directly apply traditional machine learning techniques to perform predictive modeling. We take the medical records of two anonymous patients from the publicly available MIMIC-III database and present these in Figure 1 as an example. Figure 1 clearly indicates the irregularity problem, including many missing values and varying time intervals between medical records. Fig.1. Illustration of medical records of patients A and B. Most studies have focused on exploiting variable correlations in patient med- ical records to impute missing values and establishing time-decay mechanisms to take into account the effect of varying time intervals between records [1,2,17, 18,23–25,31]. After obtaining the complete data matrices from the imputation task, the complete data matrices are used as input for downstream healthcare prediction tasks [1,2,13,17,18,22,23,27,30,31,35]. Although these studies have achieved satisfactory imputation performance, consideration of using the infor- mation of similar patients on the imputation task, which might lead to improved imputation performance, has not yet been fully experimented. Furthermore, with imputation data, high-quality representation must be applied, as the imputation data may affect the performance of downstream healthcare prediction tasks. Patient stratification refers to the method of dividing a patient population into subgroups based on specific disease characteristics and symptom severity. Patients in the same subgroup generally had more similar health trajectories. Therefore, we propose to impute missing values in patient data using information from the subgroup of similar patients rather than the entire patient population. In this paper, we propose a novel contrastive learning-based imputation- prediction network with the aim of improving in-hospital mortality prediction performance using EHR data. Missing value imputation for EHR data is done by exploiting similar patient information as well as patients’ personal contextual information. Similar patients are generated from patient similarity calculation during stratification modeling and analysis of patient graphs.
230809896/11
## Implementation Details and Evaluation Metrics We implement all approaches with PyTorch 1.11.0 and conduct experiments on A40 GPU from NVIDIA with 48GB of memory. We randomly use 70%, 15%, and 15% of the dataset as training, validation, and testing sets. We train the proposed approach using an Adam optimizer [11] with a learning rate of 0.0023 and a mini-batch size of 256. For personalized patient representation learning, the dimension size $d$ is 3. For similar patients discovery and information aggregation, the initial value of $\varphi$ is 0.56, and the dimension size of $W^{e}_{1}$ and $W^{e}_{2}$ are 34 and 55. For contrastive learning, the value of $\tau$ is 0.07. The dropout method is applied to the final Softmax output layer for the prediction task, and the dropout rate is 0.1. For the imputation task, the dimension size of $W^{(l)}_{v}$, $W^{(l)}_{t}$, $W^{(n)}_{v}$, and $W^{(n)}_{t}$ are 28. The performance of contrastive learning heavily relies on data augmentation. We augment the observed value $v$ with random time shifts and reversion. For example, given the observed value $v = [v_{1}, v_{2}, \cdots, v_{T}]$, we are able to obtain $v_{shift} = [v_{1+n}, v_{2+n}, \cdots, v_{T+n}]$ and $v_{reverse} = [v_{T}, v_{T-1}, \cdots, v_{1}]$ from random time shift and reversion, and $n$ is the number of data points to shift. We use the MAE and MRE scores between predicted and actual values as the evaluation metrics for imputation performance. We use the AUROC and AUPRC scores as the evaluation metrics for prediction performance. We report the mean and standard deviation of the evaluation metrics after repeating all the approaches ten times. ## Experimental Results Table 2 presents the experimental results of all approaches on imputation and prediction tasks from MIMIC-III and eICU datasets. Together these results sug- gest that our approach achieves the best performance in both imputation and prediction tasks. For example, for the clinical time series imputation of MIMIC- III (24 hours after ICU admission), the MAE and MRE of Ours are 0.3563 and 8.16%, smaller than 0.3988 and 38.44% achieved by the best baseline (i.e., MT- SIT). For the in-hospital mortality prediction of MIMIC-III (24 hours after ICU admission), the AUROC and AUPRC of Ours are 0.8533 and 0.4752, larger than 0.8461 and 0.4513 achieved by the best baseline (i.e., GRU-D). As Table 2 shows, the RNN-based approach (i.e., GRU-D and BRITS) out- performs the GAN-based approach (i.e., GRUI-GAN, E $^{2}$ GAN, E $^{2}$ GAN-RF, and STING) in the imputation task. From the prediction results of the MIMIC-III dataset, we can see that the transformer-based approaches (i.e., MTSIT and MIAM) resulted in lower values of AUROC and AUPRC. From the prediction results of the eICU dataset, no significant difference between the transformer- based approach and other approaches was evident. Ours outperforms its variants Ours $_{\alpha}$ and Ours $_{\beta}$. This result confirms the ef- fectiveness of the network construction with enhanced imputation and prediction performance.
230809896/6
and $\delta^{(n)}_{i, t}$ can be written as: $$ \begin{split} \delta^{(l)}_{i, t} = \begin{cases} \delta_{i, t}, & if\ m_{i, t - 1} = 1 \\ \delta_{i, t} + \delta^{(l)}_{i, t - 1}, & otherwise \end{cases} \end{split} $$ $$ \begin{split} \delta^{(n)}_{i, t} = \begin{cases} \delta_{i, t + 1}, & if\ m_{i, t + 1} = 1 \\ \delta_{i, t + 1} + \delta^{(n)}_{i, t + 1}, & otherwise \end{cases} \end{split} $$ Let $v^{(l)}$ and $v^{(n)}$ denote two neighboring value matrices, the observed values of the last time and next time. $v^{(l)}$ and $v^{(n)}$ can be written as: $$ \begin{split} v^{(l)}_{i, t} = \begin{cases} v_{i, t - 1}, & if\ m_{i, t - 1} = 1 \\ v^{(l)}_{i, t - 1}, & otherwise \end{cases} \end{split} $$ $$ \begin{split} v^{(n)}_{i, t} = \begin{cases} v_{i, t + 1}, & if\ m_{i, t + 1} = 1 \\ v^{(n)}_{i, t + 1}, & otherwise \end{cases} \end{split} $$ where $v^{(l)}_{i, t}$ and $v^{(n)}_{i, t}$ are the values of the i-th variable of $v_{t}^{(l)}$ and $v_{t}^{(n)}$. Let $D = \{(X_{p}, y_{p})\}^{P}_{p = 1}$ denote the EHR dataset with up to $P$ labeled sam- ples. The p-th sample contains a multivariate time series $X_{p}$ consisting of the physiological variables, and a binary label of in-hospital mortality $y_{p} \in \{0, 1\}$. Let $X_{base} \in \mathbb{R}^{g}$ denote the patient-specific characteristics (i.e., age, sex, ethnic- ity, admission diagnosis) with up to $g$ dimension. ## Personalized Patient Representation Learning Given an input multivari- ate time series/a single patient data $X = \{(f_{i}, v_{i}, t_{i})\}^{N}_{i = 1}$, the embedding for the i-th triplet $e_{i} \in \mathbb{R}^{d}$ is generated by aggregating the feature embedding $e_{i}^{(f)} \in \mathbb{R}^{d}$, the value embedding $e_{i}^{(v)} \in \mathbb{R}^{d \times T}$, and the time interval embedding $e^{(t)}_{i} \in \mathbb{R}^{d \times T}$. The feature embedding is similar to the word embedding, which allows features with similar meanings to have a similar representation. Particularly, the value embedding and time interval embedding are obtained by separately implement- ing a multi-channel feed-forward neural network (FFN) as: $$ \begin{split} e_{i, 1}^{(v)}, \cdots, e_{i, T}^{(v)} = FFN_{i}^{(v)} (v_{i, 1}, \cdots, v_{i, T}), \\ e_{i, 1}^{(t)}, \cdots, e_{i, T}^{(t)} = FFN_{i}^{(t)} (\delta_{i, 1}, \cdots, \delta_{i, T}). \end{split} $$ Through the processes above, we are able to obtain $e^{(f)} \in \mathbb{R}^{Nd}$, $e^{(v)} \in \mathbb{R}^{Nd \times T}$, and $e^{(t)} \in \mathbb{R}^{Nd \times T}$, which are fed into the attention-based cross module to generate an overall representation. Note that $e^{(f)} \in \mathbb{R}^{Nd}$ is expanded into $e^{(f)} \in \mathbb{R}^{Nd \times T}$
230809896/4
Another line of related work is based on the generative adversarial network (GAN) architecture, which aims at treating the problem of missing data impu- tation as data generation. The intuitions behind GAN can be seen as making a generator and a discriminator against each other [6]. The generator generates fake samples from random ’noise’ vectors, and the discriminator distinguishes the generator’s fake samples from actual samples. Examples of research into GAN-based imputation methods include GRUI-GAN [17], E $^{2}$ GAN [18], E $^{2}$ GAN- RF [40], and STING [25]. These studies take the vector of actual samples, which has many missing values, use a generator to generate the corresponding imputed values and distinguish the generated imputed values from real values using a discriminator. Several studies have evaluated the effectiveness of applying transformer-based imputation methods to EHR data. Examples of representative studies include MTSIT [37] and MIAM [13]. The MTSIT is built with an autoencoder archi- tecture to perform missing value imputation in an unsupervised manner. The autoencoder architecture used in MTSIT includes the Transformer encoder [32] and a linear decoder, which are implemented with a joint reconstruction and im- putation approach. The MIAM is built upon the self-attention mechanism [32]. Given EHR data, MIAM imputes the missing values by extracting the relation- ship among the observed values, missingness indicators (0 for missing and 1 for not missing), and the time interval between consecutive observations. ## Method ## Network Architecture The architecture of the proposed network is shown in Figure 2. ## Data Representation We represent a multivariate time series $X$ with up to $N$ variables of length $T$ as a set of observed triplets, i.e., $X = \{(f_{i}, v_{i}, t_{i})\}^{N}_{i = 1}$. An observed triplet is represented as a $(f, v, t)$, where $f \in F$ is the variable/feature, $v \in \mathbb{R}^{T}$ is the observed value, and $t \in \mathbb{R}^{T}$ is the time. We incorporate a masking vector $m_{i}$ to represent missing values in $v_{i}$ as: $$ \begin{split} m_{i, t} = \begin{cases} 1, & if\ v_{i, t}\ is\ observed \\ 0, & otherwise \end{cases} \end{split} $$ Let $\delta \in \mathbb{R}^{N \times T}$, $\delta^{(l)} \in \mathbb{R}^{N \times T}$, and $\delta^{(n)} \in \mathbb{R}^{N \times T}$ denote three time interval matrices. $\delta_{t}$ is the time interval between the current time $t$ and the last time $t-1$. $\delta^{(l)}_{i, t}$ is the time interval between the current time $t$ and the time where the i-th variable is observed the last time. $\delta^{(n)}_{i, t}$ is the time interval between the current time $t$ and the time where the i-th variable is observed next time. $\delta^{(l)}_{i, t}$
230809896/10
Given the final representation $e^{*}$ and the embedding matrix $e^{c}$, we use a fully connected layer to impute missing values as: $$ \begin{split} \hat{v} = e^{*} \cdot W^{v}_{1} + e^{c} \cdot W^{v}_{2} + b_{v}. \end{split} $$ The objective loss is the summation of the mean square error and the unsu- pervised contrastive loss with a scaling parameter $\lambda$ to control the contribution of each loss as: $$ \begin{split} \mathcal{L}_{MSE} = \frac{1}{P} \sum_{p = 1}^{P} (m_{p} \odot v_{p} - m_{p} \odot \hat{v}_{p})^{2}, \\ \mathcal{L} = \lambda \cdot \mathcal{L}_{MSE} + (1 - \lambda) \cdot \mathcal{L}_{UC}. \end{split} $$ ## Experiments ## Datasets and Tasks We validate our approach on the MIMIC-III $\footnote{https://mimic.physionet.org}$ and eICU $\footnote{https://eicu-crd.mit.edu/}$ datasets. We conduct clinical time series imputation and in-hospital mortality experiments based on the data from the first 24/48 hours after admission. Detailed information on both datasets can be found in the literature [9] and [28]. The source code of our ap- proach and statistics of features are released at https://github.com/liulab1356/CL- ImpPreNet. ## Baseline Approaches We compare our approach with GRU-D [2], BRITS [1], GRUI-GAN [17], E $^{2}$ GAN [18], E $^{2}$ GAN-RF [40], STING [25], MTSIT [37], and MIAM [13] (see related work section). We feed the output of GRUI-GAN, E $^{2}$ GAN, E $^{2}$ GAN-RF, STING, and MTSIT into GRU to estimate in-hospital mortality risk probabilities. Moreover, the regression component used in BRITS is integrated into GRU-D and MIAM to obtain imputation accuracy. Besides, two variants of our approach are as follows: Ours $_{\alpha}$: We do not perform graph analysis-based patient stratification mod- eling. Ours $_{\beta}$: We omit the contrastive learning component. All implementations of Ours $_{\alpha}$ and Ours $_{\beta}$ can be found in the aforementioned Github repository.
230809896/1
# Contrastive Learning-based Imputation-Prediction Networks for In-hospital Mortality Risk Modeling using EHRs Yuxi Liu (cid:66) 1, Zhenhao Zhang 2, Shaowen Qin 1, Flora D. Salim 3, and Antonio Jimeno Yepes Abstract. Predicting the risk of in-hospital mortality from electronic health records (EHRs) has received considerable attention. Such pre- dictions will provide early warning of a patient’s health condition to healthcare professionals so that timely interventions can be taken. This prediction task is challenging since EHR data are intrinsically irregular, with not only many missing values but also varying time intervals be- tween medical records. Existing approaches focus on exploiting the vari- able correlations in patient medical records to impute missing values and establishing time-decay mechanisms to deal with such irregularity. This paper presents a novel contrastive learning-based imputation-prediction network for predicting in-hospital mortality risks using EHR data. Our approach introduces graph analysis-based patient stratification modeling in the imputation process to group similar patients. This allows informa- tion of similar patients only to be used, in addition to personal contex- tual information, for missing value imputation. Moreover, our approach can integrate contrastive learning into the proposed network architecture to enhance patient representation learning and predictive performance on the classification task. Experiments on two real-world EHR datasets show that our approach outperforms the state-of-the-art approaches in both imputation and prediction tasks. Keywords: data imputation · in-hospital mortality · contrastive learn- ing. ## Introduction The broad adoption of digital healthcare systems produces a large amount of electronic health records (EHRs) data, providing us the possibility to develop
230809896/3
Contrastive learning has been proven to be an important machine learn- ing technique in the computer vision community [12]. In contrastive learning, representations are learned by comparing input samples. The comparisons are made on the similarity between positive pairs or dissimilarity between nega- tive pairs. The main goal is to learn an embedding space where similar samples are put closer to each other while dissimilar samples are pushed farther apart. Contrastive learning can be applied in both supervised [10,33,39] and unsuper- vised [14,15,26] settings. Motivated by the recent developments in contrastive representation learn- ing [34,36,38], we integrate contrastive learning into the proposed network ar- chitecture to perform imputation and prediction tasks. The benefit of incorpo- rating contrastive learning into the imputation task is that such an approach can enhance patient representation learning by keeping patients of the same strati- fication together and pushing away patients from different stratifications. This would lead to enhanced imputation performance. The benefit of incorporating contrastive learning into the prediction task is improved predictive performance of the binary classification problem (i.e., the risk of death and no death), which is achieved by keeping the instances of a positive class closer and pushing away instances from a negative class. Our major contributions are as follows: - – To the best of our knowledge, this is the first attempt to consider patient similarity via stratification of EHR data on the imputation task. – We propose a novel imputation-prediction approach to perform imputation and prediction simultaneously with EHR data. – We successfully integrate contrastive learning into the proposed network architecture to improve imputation and prediction performance. – Extensive experiments conducted on two real-world EHR datasets show that our approach outperforms all baseline approaches in imputation and predic- tion tasks. ## Related Work There has been an increased interest in EHR-based health risk predictions [5, 16,19–21]. It has been recognized that EHR data often contains many missing values due to patient conditions and treatment decisions [31]. Existing research addresses this challenge by imputing missing data and feeding them into the supervised algorithms as auxiliary information [7]. GRU-D [2] represents such an example. The GRU-D is built upon the Gated Recurrent Unit [4]. GRU-D proposes to impute missing values by decaying the contributions of previous observation values toward the overall mean over time. Similarly, BRITS [1] in- corporates a bidirectional recurrent neural network (RNN) to impute missing values. Since the incorporated bidirectional RNN learns EHR data in both for- ward and backward directions, the accumulated loss is introduced to train the model.
230809896/9
as the positive pairs and the patient representations with different labels as the negative pairs. For the imputation task, we augment the standard mean squared error loss with the unsupervised contrastive loss [3]. We treat a single patient representation and its augmented representations as positive pairs and the other patient representations within a batch and their augmented representations as negative pairs. The formula can be written as: $$ \begin{split} \mathcal{L}_{SC} = - \sum_{i = 1}^{B} \frac{1}{B_{y_{i}}} log \frac{\sum_{j = 1}^{B} \mathbbm{1}_{[y_{i} = y_{j}]} exp(sim(e^{*}_{i}, e^{*}_{j}) / \tau)} {\sum_{k = 1}^{B} \mathbbm{1}_{[k \neq i]} exp(sim(e^{*}_{i}, e^{*}_{k}) / \tau)}, \\ \mathcal{L}_{UC} = - log \frac{exp(sim(e^{*}_{i}, e^{*}_{j}) / \tau)} {\sum^{2B}_{k = 1} \mathbbm{1}_{[k \neq i]} exp(sim(e^{*}_{i}, e^{*}_{k})/\tau)}, \end{split} $$ where $B$ represents the batch size; $\mathbbm{1}_{[\cdot]}$ represents an indicator function; $sim(\cdot)$ represents the cosine similarity measure; $\tau$ represents a hyper-parameter that is used to control the strength of penalties on negative pairs; $B_{y_{i}}$ is the number of samples with the same label in each batch. ## Imputation and Prediction Tasks For the prediction task, we feed $e^{*}$ into a softmax output layer to obtain the predicted $\hat{y}$ as: $$ \begin{split} \hat{y} = Softmax(W_{y} \cdot e^{*} + b_{y}). \end{split} $$ The objective loss is the summation of cross-entropy loss and the supervised contrastive loss with a scaling parameter $\lambda$ to control the contribution of each loss as: $$ \begin{split} %\hat{y} = Softmax(W_{y} \cdot e^{*} + b_{y}), \\ \mathcal{L}_{CE} = - \frac{1}{P} \sum^{P}_{p=1} (y_{p}^{\rightarrowp} \cdot log(\hat{y}_{p}) + (1 - y_{p})^{\rightarrowp} \cdot log(1 - \hat{y}_{p})), \\ \mathcal{L} = \lambda \cdot \mathcal{L}_{CE} + (1 - \lambda) \cdot \mathcal{L}_{SC}. \end{split} $$ For the imputation task, we take the neighboring observed values (of each patient) as inputs to incorporate patient-specific contextual information. The process of embedding used by $v^{(l)}$ and $v^{(n)}$ can be written as: $$ \begin{split} e_{i}^{(v), (l)} = FFN_{i}^{(v), (l)} (v_{i}^{(l)}), e_{i}^{(t), (l)} = FFN_{i}^{(t), (l)} (\delta_{i}^{(l)}), \\ e_{i}^{(v), (n)} = FFN_{i}^{(v), (n)} (v_{i}^{(n)}), e_{i}^{(t), (n)} = FFN_{i}^{(t), (n)} (\delta_{i}^{(n)}), \\ \tilde{e}^{(l)} = W_{v}^{(l)} \cdot e^{(v), (l)} + W_{t}^{(l)} \cdot e^{(t), (l)} + b_{e}^{(l)}, \\ \tilde{e}^{(n)} = W_{n}^{(v)} \cdot e^{(v), (n)} + W_{t}^{(n)} \cdot e^{(t), (n)} + b_{e}^{(n)}, \\ e^{c} = Concate(\tilde{e}^{(l)}, \tilde{e}^{(n)}), \end{split} $$ where $\tilde{e}^{(l)}$ and $\tilde{e}^{(n)}$ are the representations of $v^{(l)}$ and $v^{(n)}$ after embedding. The embedding matrix $e^{c}$ is obtained by concatenating $\tilde{e}^{(l)}$ and $\tilde{e}^{(n)}$.
230809896/8
where $sim(\cdot)$ is the measure of cosine similarity and $\Lambda \in \mathbb{R}^{B \times B}$ is the patient similarity matrix. Moreover, we incorporate a learnable threshold $\varphi$ into the patient similarity calculation to filter out similarities below the threshold. The similarity matrix can be rewritten as: $$ \begin{split} \Lambda^{\prime} = \begin{cases} \Lambda, & if\ \Lambda\ >\ \varphi \\ 0, & otherwise \end{cases} \end{split} $$ We take into account the batch of patients’ representations as a graph to aggregate the information from similar patients, where the similarity matrix $\Lambda^{\prime}$ is the graph adjacency matrix. We apply graph convolutional layers to enhance the representation learning as: $$ \begin{split} \hat{e} = [\hat{e}_{1}, \hat{e}_{2}, \cdots, \hat{e}_{B}]^{\rightarrowp} = GCN(e^{\prime}, \Lambda^{\prime}) \\ = ReLU(\Lambda^{\prime} ReLU(\Lambda^{\prime} \cdot e^{\prime} W^{e}_{1}) \cdot W^{e}_{2}), \end{split} $$ where $\hat{e}$ is the aggregated auxiliary information from similar patients. A note of caution is due here since we ignore the bias term. We replace $e^{\prime}$ in Eq. (15) with $e^{\prime\prime}$ for the imputation task. By doing so, the output of graph convolutional layers can take the form of sequence data. Particularly, $e^{\prime\prime}$ is obtained by concatenating $e$ and $e_{base}$, where $e_{base} \in \mathbb{R}^{d_{g}}$ is expanded into $e_{base} \in \mathbb{R}^{d_{g} \times T}$. Through the processes above, we are able to generate $e^{\prime}$ / $e^{\prime\prime}$ and $\hat{e}$ represen- tations for the batch of patients. The $e^{\prime}$ / $e^{\prime\prime}$ refers to the patient themselves. For an incomplete patient $p$ (i.e., the patient data has many missing values), we generate the missing value representations with $\hat{e}$. For a complete patient, we augment $e^{\prime}$ / $e^{\prime\prime}$ with $\hat{e}$ to enhance the representation learning. We design an attention-based fusion module to refine both $e^{\prime}$ / $e^{\prime\prime}$ (the two representations used in prediction and imputation tasks) and $\hat{e}$. Since imputation and prediction tasks involve the same process of modeling, we take the prediction task as an example. The two weights $\gamma \in \mathbb{R}^{B}$ and $\eta \in \mathbb{R}^{B}$ are incorporated to determine the importance of $e^{\prime}$ and $\hat{e}$, obtained by implementing fully connected layers as: $$ \begin{split} \gamma = Sigmoid(e^{\prime} \cdot W_{\gamma} + b_{\gamma}), \\ \eta = Sigmoid(\hat{e} \cdot W_{\eta} + b_{\eta}). \end{split} $$ A note of caution is due here since we keep the sum of $\gamma$ and $\eta$ must be 1, i.e., $\gamma + \eta = 1$. We achieve this constraint by combining $\gamma = \frac{\gamma}{\gamma + \eta}$ and $\eta = 1- \gamma$. The final representation $e^{*}$ is obtained by calculating $\gamma \cdot e^{\prime} + \eta \cdot \hat{e}$. ## Contrastive Learning We integrate contrastive learning into the proposed network architecture to perform imputation and prediction tasks. For the pre- diction task, we augment the standard cross-entropy loss with the supervised contrastive loss [10]. We treat the patient representations with the same label
200603094/20
Proof. Since the abstract compression of $M_2(V)$ by $p \oplus q$ is an operator system, the direct sum $$ \rho = \bigoplus \phi, $$ where the direct sum is taken over all ucp $\phi: M_2(\mathcal{V})/ J_{p \oplus q} \rightarrow M_n$ and all $n \in \mathbb{N}$, is a unital complete order embedding (see [3] as well as Chapter 13 of [15]). Replacing each $\phi$ with a corresponding $\psi$ as in Proposition 5.8 we obtain a unital completely positive map $\rho'$ of $M_2(\mathcal{V})/ J_{p \oplus q}$ into $B(H)$ mapping $p \oplus 0 + J_{p \oplus q}$ to a projection. In fact, $\rho'$ is a complete order embedding. To see this, suppose that $$ \begin{pmatrix} a & b \\ b^* & c \end{pmatrix} + M_n(J_{p \oplus q}) \in M_n(M_2(\mathcal{V}) / J_{p \oplus q}) $$ is non-positive. Then it is necessarily the case that $$ \begin{pmatrix} a & 0 & 0 & b \\ 0 & 0 & 0 & 0 \\ 0 & 0 & 0 & 0 \\ b^* & 0 & 0 & c \end{pmatrix} + M_{2n}(J_{p \oplus q}) $$ is also non-positive. It follows that there exists a unital completely positive map $\phi: M_2(\mathcal{V}) / J_{p \oplus q} \rightarrow M_n$ such that $$ \phi_{2n} \left(\begin{pmatrix} a & 0 & 0 & b \\ 0 & 0 & 0 & 0 \\ 0 & 0 & 0 & 0 \\ b^* & 0 & 0 & c \end{pmatrix} + M_{2n}(J_{p \oplus q}) \right) $$ is a non-positive matrix (for example, see the proof of Theorem 13.1 in [15]). It follows from Proposition 5.8 that the corresponding map $\psi: M_2(\mathcal{V}) / J_{p \oplus q} \rightarrow M_{n'}$ has the property that $$ \psi \begin{pmatrix} a & b \\ b^* & c \end{pmatrix} $$ is non-positive. We conlcude that $\rho'$ is a unital complete order embedding. We complete the proof by precomposing $\rho'$ with the complete order embedding $\pi_p$ so that $\pi = \rho' \circ \pi_p$ is the desired unital complete order embedding. Theorem 5.10 shows that when $p \in \mathcal{V}$ is an abstract projection, we can build a complete order embedding of $\mathcal{V}$ into $B(H)$ mapping $p$ to an “ honest ” projection. Of course a given operator system may contain many abstract projections, and a representation making $p$ into a projection may not map other abstract projections to projections. The next theorem shows that there is always one complete order embedding of $\mathcal{V}$ which maps all abstract projections to concrete projections. The reader should compare the following with Theorem 2.10. Theorem5.11. Let $\mathcal{V}$ be an operator system, and suppose that $p \in \mathcal{V}$ is an abstract projection. Then $p$ is a projection in its C*-envelope $C_e^*(\mathcal{V})$. Proof. Suppose that $p$ is an abstract projection in $\mathcal{V}$, and let $j:\mathcal{V} \rightarrow C_e^*(\mathcal{V})$ denote the inclusion map. By Theorem 5.10, there exists a unital complete order embedding $\phi: \mathcal{V} \rightarrow B(H)$ with the property that $\phi(p)$ is a projection. Let $\mathcal{A} := C^*(\phi(\mathcal{V}))$. By the universal property of the C*-envelope, there exists a $*$ -epimorphism $\pi: \mathcal{A} \rightarrow C_e^*(\mathcal{V})$ satisfying $\pi(\phi(x)) = j(x)$ for all $x \in \mathcal{V}$. Consequently $$ j(p) = \pi(\phi(p)) = \pi(\phi(p)^2) = \pi(\phi(p))^2 = j(p)^2. $$ Since $j(p) = j(p)^*$, we conclude that $j(p)$ is a projection in $C_e^*(\mathcal{V})$.
200603094/18
corner, and the non-zero entries of $$ \phi \widehat{\begin{pmatrix} 0 & b \\ 0 & 0 \end{pmatrix}} $$ lie in its upper right $m' \times (n-m)$ corner. Indeed, when these statements hold, the map $\psi$ is simply the compression of the matrix $$ \begin{pmatrix} \phi \widehat{\begin{pmatrix} a & 0 \\ 0 & 0 \end{pmatrix}} & \phi \widehat{\begin{pmatrix} 0 & b \\ 0 & 0 \end{pmatrix}} \\ \phi \widehat{\begin{pmatrix} 0 & 0 \\ b^* & 0 \end{pmatrix}} & \phi \widehat{\begin{pmatrix} 0 & 0 \\ 0 & c \end{pmatrix}} \end{pmatrix} $$ to the $(m' + n - m) \times (m' + n - m)$ submatrix upon which it is supported, followed by conjugation by the invertible matrix $$ \begin{pmatrix} I_m & & & & & & & \\ & x_{m+1}^{-1/2} & & & & & & \\ & & \ddots & & & & & \\ & & & x_{m'}^{-1/2} & & & & \\ & & & & y_{m+1}^{-1/2} & & & \\ & & & & & \ddots & & \\ & & & & & & y_{m'}^{-1/2} & \\ & & & & & & & I_{n-m'} \end{pmatrix}. $$ We first consider the coset of the matrix with $b$ in its upper right corner and zeroes elsewhere, where $b \in \mathcal{V}$. Since $\widehat{p \oplus q}$ is an order unit for $M_2(\mathcal{V}) / J_{p \oplus q}$, we may assume (by rescaling $b$ if necessary) that $$ \begin{pmatrix} p & b \\ b^* & q \end{pmatrix} \in C(p \oplus q). $$ This implies that $$ \begin{pmatrix} p & 0 & 0 & b \\ 0 & 0 & 0 & 0 \\ 0 & 0 & 0 & 0 \\ b^* & 0 & 0 & q \end{pmatrix} \in C(p_{2} \oplus q_{2}). $$ The complete positivity of $\phi$ implies that $$ \begin{pmatrix} \tilde{P} & \phi \widehat{\begin{pmatrix} 0 & b \\ 0 & 0 \end{pmatrix}} \\ \phi \widehat{\begin{pmatrix} 0 & 0 \\ b^* & 0 \end{pmatrix}} & \tilde{Q} \end{pmatrix} \geq 0. $$ The claim follows. Next we consider $\phi(a \oplus 0)$ for $a = a^*$. By again rescaling $a$ as necessary, we may assume that $$ \begin{pmatrix} p \pm a & 0 \\ 0 & q \end{pmatrix} \in C(p \oplus q). $$ By the definition of $C(p \oplus q)$ and compressing to the upper left corner, this implies that for every $\epsilon > 0$ there exists a $t > 0$ such that $p \pm a + \epsilon p + tq \geq 0$. Using the definition of $C(p \oplus q)$ again, we see that this implies $$ \begin{pmatrix} p \pm a & 0 \\ 0 & 0 \end{pmatrix} \in C(p \oplus q). $$
200603094/17
We conclude that the matrix in line (15) is positive. It follows that $\begin{pmatrix} p & p \\ p & p \end{pmatrix} \in J_{p \oplus q}$ and thus $\pi_p$ is not a complete order embedding (in fact, it is not injective). Thus, similar to the case for C*-algebras, we see that for an operator system $\mathcal{V}$, $p$ is a an abstract projection if and only if $q :=e-p$ is an abstract projection, and every non-trivial abstract projection in $\mathcal{V}$ has norm 1. Theorem 5.3, together with Corollary 4.10 and Lemma 5.2 imply that the map $\pi_p$ is a complete order isomorphism onto its range whenever $p$ is a projection in a concrete operator system. In other words, every concrete projection is an abstract projection. It remains to show that every abstract projection is a concrete projection under some complete order embedding of its containing operator system. We proceed by first showing that matrix-valued ucp maps on $M_2(\mathcal{V})/ J_{p \oplus q}$ can be modified to build new matrix-valued ucp maps sending $p \oplus 0 + J_{p \oplus q}$ and $0 \oplus q + J_{p \oplus q}$ to projections. Proposition5.8. Suppose that $\mathcal{V}$ is an operator system and that $p$ is an abstract projection in $\mathcal{V}$. Let $q = e - p$. Then for every ucp map $\phi: M_2(\mathcal{V})/ J_{p \oplus q} \rightarrow M_n$ there exists a $k \in \mathbb{N}$ and a ucp map $\psi: M_2(\mathcal{V})/ J_{p \oplus q} \rightarrow M_{k}$ such that $\psi(p \oplus 0 + J_{p \oplus q})$ and $\psi(0 \oplus q + J_{p \oplus q})$ are projections and satisfying the property that $$ \phi_{2n} \left(\begin{pmatrix} a & 0 & 0 & b \\ 0 & 0 & 0 & 0 \\ 0 & 0 & 0 & 0 \\ b^* & 0 & 0 & c \end{pmatrix} + M_{2n}(J_{p \oplus q}) \right) = \begin{pmatrix} \phi_n \left(\begin{pmatrix} a & 0 \\ 0 & 0 \end{pmatrix} + M_n(J_{p \oplus q}) \right) & \phi_n \left(\begin{pmatrix} 0 & b \\ 0 & 0 \end{pmatrix} + M_n(J_{p \oplus q}) \right) \\ \phi_n \left(\begin{pmatrix} 0 & 0 \\ b^* & 0 \end{pmatrix} + M_n(J_{p \oplus q}) \right) & \phi_n \left(\begin{pmatrix} 0 & 0 \\ 0 & c \end{pmatrix} + M_n(J_{p \oplus q}) \right) \end{pmatrix} \geq 0 $$ if and only if $\psi_n \left(\begin{pmatrix} a & b \\ b^* & c \end{pmatrix} + M_n(J_{p \oplus q}) \right) \geq 0$ for all $a,b,c \in M_n(\mathcal{V})$. Proof. To simplify notation, we will let $\hat{x}$ denote the coset $x + M_n(J_{p \oplus q})$ for each $x \in M_n(\mathcal{V})$ throughout the proof. Suppose $\phi: M_2(\mathcal{V})/ J_{p \oplus q} \rightarrow M_n$ is ucp. Since $\phi(\widehat{p \oplus 0}) = I_n - \phi(\widehat{0 \oplus q})$, we have that $\phi(\widehat{p \oplus 0})$ commutes with $\phi(\widehat{0 \oplus q})$. Thus we may find a common orthonormal basis for $\mathbb{C}^n$ such that $\phi(\widehat{p \oplus 0})$ and $\phi(\widehat{0 \oplus q})$ are both diagonal. By reordering this basis, we may assume $$ \phi(\widehat{p \oplus 0}) = \tilde{P} = \begin{pmatrix} I_m & & & & \\ & x_{m+1} & & & \\ & & \ddots & & \\ & & & x_{m'} & \\ & & & & 0_{n-m'} \end{pmatrix}, \quad \phi(\widehat{0 \oplus q}) = \tilde{Q} = \begin{pmatrix} 0_m & & & & \\ & y_{m+1} & & & \\ & & \ddots & & \\ & & & y_{m'} & \\ & & & & I_{n-m'} \end{pmatrix} $$ where $0 \leq m \leq m' \leq n$ and $x_i + y_i = 1$ for each $i=m+1,\dots, m'$ and $x_i, y_i \in (0,1)$. Define rectangular matrices $$ V = \begin{pmatrix} I_m & & & & 0 & \dots & 0 \\ & x_{m+1}^{-1/2} & & & & & \\ & & \ddots & & & & \\ & & & x_{m'}^{-1/2} & 0 & \dots & 0 \end{pmatrix} \in M_{m',n}, \quad W = \begin{pmatrix} 0 & \dots & 0 & y_{m+1}^{-1/2} & & & \\ & & & & \ddots & \\ & & & & & y_{m'}^{-1/2} \\ 0 & \dots & 0 & & & & I_{n-m'} \end{pmatrix} \in M_{n-m,n}. $$ Thus $V \tilde{P} V^* = I_{m'}$ and $W \tilde{Q} W^* = I_{n-m}$. We may now define $\psi: M_2(V) / J_{p \oplus q} \rightarrow M_{m' + n - m}$ via $$ \psi \begin{pmatrix} a & b \\ c & d \end{pmatrix} = \begin{pmatrix} V & 0 \\ 0 & W \end{pmatrix} \begin{pmatrix} \phi \widehat{\begin{pmatrix} a & 0 \\ 0 & 0 \end{pmatrix}} & \phi \widehat{\begin{pmatrix} 0 & b \\ 0 & 0 \end{pmatrix}} \\ \phi \widehat{\begin{pmatrix} 0 & 0 \\ c & 0 \end{pmatrix}} & \phi \widehat{\begin{pmatrix} 0 & 0 \\ 0 & d \end{pmatrix}} \end{pmatrix} \begin{pmatrix} V^* & 0 \\ 0 & W^* \end{pmatrix}. $$ Then $\psi$ is ucp, $\psi(\widehat{p \oplus 0}) = I_{m'} \oplus 0_{n-m}$ and $\psi(\widehat{0 \oplus q}) = 0_{m'} \oplus I_{n-m}$. It remains to check the final statement of the proposition. To show this, it suffices to show that the non-zero entries of $\phi(\widehat{a \oplus 0})$ lie in its upper left $m' \times m'$ corner, the non-zero entries of $\phi(\widehat{0 \oplus c})$ lie in its lower right $(n-m) \times (n-m)$
200603094/19
By the positivity of $\phi$, this means that $$ \tilde{P} \pm \phi \widehat{\begin{pmatrix} a & 0 \\ 0 & 0 \end{pmatrix}} \geq 0. $$ It follows that the non-zero entries of $\phi(\widehat{a \oplus 0})$ lie in its upper left $m' \times m'$ corner as claimed. A similar proof shows that $\phi(\widehat{0 \oplus c})$ has its non-zero entries in its lower right $(n-m) \times (n-m)$ corner whenever $c = c^*$. We prove one final lemma before arriving at the main result of this section. This lemma ensures that the map $\pi_p$ in Definition 5.4 is unital. Lemma5.9. For any positive contraction $p$ in an operator system $\mathcal{V}$ we have $$ \begin{pmatrix} p & p \\ p & p \end{pmatrix} + J_{p \oplus q} = \begin{pmatrix} p & 0 \\ 0 & 0 \end{pmatrix} + J_{p \oplus q} $$ and $$ \begin{pmatrix} q & q \\ q & q \end{pmatrix} + J_{p \oplus q} = \begin{pmatrix} 0 & 0 \\ 0 & q \end{pmatrix} + J_{p \oplus q} $$ where $q = e - p$. Consequently the map $\pi_p$ is unital. Proof. It suffices to show that $$ \pm \begin{pmatrix} 0 & p \\ p & p \end{pmatrix}, \pm \begin{pmatrix} q & q \\ q & 0 \end{pmatrix} \in C(p \oplus q). $$ Let $\epsilon > 0$. Then $$ \begin{aligned} \begin{pmatrix} 0 & p \\ p & p \end{pmatrix} + \epsilon \begin{pmatrix} p & 0 \\ 0 & q \end{pmatrix} + \frac{1}{\epsilon} \begin{pmatrix} q & 0 \\ 0 & p \end{pmatrix} & = & \begin{pmatrix} \epsilon p & p \\ p & (1+\frac{1}{\epsilon}) p \end{pmatrix} + \begin{pmatrix} \frac{1}{\epsilon} q & 0 \\ 0 & \epsilon q \end{pmatrix} \\ & = & \begin{pmatrix} \epsilon & 1 \\ 1 & 1+\frac{1}{\epsilon} \end{pmatrix} \otimes p + \begin{pmatrix} \frac{1}{\epsilon} & 0 \\ 0 & \epsilon \end{pmatrix} \otimes q \geq 0. \end{aligned} $$ Also, $$ \begin{aligned} \begin{pmatrix} 0 & -p \\ -p & -p \end{pmatrix} + \epsilon \begin{pmatrix} p & 0 \\ 0 & q \end{pmatrix} + (1+\frac{1}{\epsilon}) \begin{pmatrix} q & 0 \\ 0 & p \end{pmatrix} & = & \begin{pmatrix} \epsilon p & -p \\ -p & \frac{1}{\epsilon} p \end{pmatrix} + \begin{pmatrix} (1+\frac{1}{\epsilon}) q & 0 \\ 0 & \epsilon q \end{pmatrix} \\ & = & \begin{pmatrix} \epsilon & -1 \\ -1 & \frac{1}{\epsilon} \end{pmatrix} \otimes p + \begin{pmatrix} 1 + \frac{1}{\epsilon} & 0 \\ 0 & \epsilon \end{pmatrix} \otimes q \geq 0. \end{aligned} $$ The proof that $$ \pm \begin{pmatrix} q & q \\ q & 0 \end{pmatrix} \in C(p \oplus q) $$ is similar. For the final statement, we observe that $$ \pi_p(e) = \begin{pmatrix} e & e \\ e & e \end{pmatrix} + J_{p \oplus q} = \begin{pmatrix} p & p \\ p & p \end{pmatrix} + J_{p \oplus q} + \begin{pmatrix} q & q \\ q & q \end{pmatrix} + J_{p \oplus q} = \begin{pmatrix} p & 0 \\ 0 & q \end{pmatrix} + J_{p \oplus q}. $$ Noting that $p \oplus q + J_{p \oplus q}$ is the unit of $M_2(\mathcal{V})/ J_{p \oplus q}$ concludes the proof Theorem5.10. Suppose that $\mathcal{V}$ is an operator system and that $p \in \mathcal{V}$ is an abstract projection. Then there exists a unital complete order embedding $\pi: \mathcal{V} \rightarrow B(H)$ such that $\pi(p)$ is a projection in $B(H)$.
200603094/5
Theorem2.10 ([1]). Given an operator space $\mathcal{E}$. the following are equivalent for an element $u \in \mathcal{E}$: $u$ is - a unitary in $\mathcal{E}$. 2. There exists a TRO $Z$ containing $\mathcal{E}$ completely isometrically such that $u$ is a $C^*$ -unitary in $Z$. $u$ is a $C^*$ -unitary in the ternary envelope $T(\mathcal{E}).$ Remark2.11. Throughout the rest of the manuscript there will be results where we will be considering $\mathcal{V}$ as already sitting in some $B(H)$. In particular, we will consider projections in a concrete operator system in Section 3 and projec- tions in an abstract operator system in Section 4. Thus we will interchange between the use of $e$ or $I$ as the Archimedean order unit dependent on whether our operator system is abstract or concrete. The Archimedean order unit will always be clear from context. ## Concrete compression operator systems In this section we present the motivation for Sections 4 and 5. Starting with a concrete operator system $\mathcal{V} \subset B(H)$ we show that an element $p \in \mathcal{V}^+$, which is also a projection on $B(H)$ induces a natural collection of cones which correspond to the hermitian elements of $\mathcal{V}$ whose compression by $p$ is positive. These cones form a matrix ordering and in particular, will form a proper matrix ordering on a certain quotient $*$ -vector space. It will follow that image of $p$ in the quotient is an Archimedean matrix order unit for the space. Let $\mathcal{V} \subset B(H)$ be an operator system and let $p \in \mathcal{V}$ be a projection (as an operator in $B(H)$). Then letting $p_n:= I_n \otimes p$ we consider the collection of sets $\{C(p_n)\}_n$ where for each $n$ \begin{align*} C(p_n) = \{x \in M_n(\mathcal{V})_h: p_nxp_n \in B(H^{n})^+\}, \end{align*} where $H^{n}$ denotes the $n$ -fold Hilbertian direct sum. We will show that the sequence $\{C(p_n)\}_n$ is a (not necessarily proper) matrix ordering on $\mathcal{V}$. Proposition3.1. Let $\mathcal{V} \subset B(H)$ be an operator system and suppose that $p \in \mathcal{V}$ where $p$ is a projection in $B(H)$. The sequence of sets $\{C(p_n)\}_n$ is a matrix ordering on $\mathcal{V}.$ Furthermore if $p \leq q \leq I$ then $q$ is an Archimedean matrix order unit for $(\mathcal{V}, \{C(p_n)\}_n).$ Proof. By definition $C(p_n)^* = C(p_n)$ for all $n \in \mathbb{N}.$ The compression by the projection $p$ is a linear map and thus $\lambda C(p_n) \subset C(p_n)$ for all $n \in \mathbb{N}, \lambda >0$ and $C(p_n) + C(p_n) \subset C(p_n).$ Finally, let $\alpha \in M_{n,m}, x \in C(p_n).$ Then \begin{align*} p_m\alpha^*x\alpha p_m = p_m\left(\left[\sum_{kl} \overline{\alpha}_{ki}x_{kl}\alpha_{lj} \right]_{ij} \right)p_m = \left[p\left(\sum_{kl} \overline{\alpha}_{ki}x_{kl}\alpha_{lj}\right)p \right]_{ij} = \left[\sum_{kl}\overline{\alpha}_{ki}px_{kl}p\alpha_{lj} \right]_{ij} = \alpha^*(p_nxp_n)\alpha, \end{align*} and $\alpha^*(p_nxp_n)\alpha \in B(H^m)^+.$ This proves the first statement. We now show that if $p \leq q \leq I$ then $q$ is an Archimedean matrix order unit for the pair $(\mathcal{V}, \{C(p_n)\}_n).$ We first make some observations. It is immediate that given the projection $p \in \mathcal{V}$ then $p$ is an Archimedean matrix order unit for operators of the form $pzp, z \in \mathcal{V}.$ Simply notice that for any $n \in \mathbb{N}$ and $x \in M_n(\mathcal{V})$, then $I_n \geq p_n$ and if for $\epsilon >0$ we have $\epsilon p_n + p_nxp_n \in B(H^n)^+$ then \begin{align*} \epsilon I_n + p_nxp_n \geq \epsilon p_n + p_nxp_n \in B(H^n)^+. \end{align*} Since $I$ is an Archimedean matrix order unit for the operator system $B(H)$ we have $p_nxp_n \in B(H^n)^+$ and thus $x \in C(p_n).$ By the assumption that $p \leq q$ it follows that $p \leq pqp$. (In fact since $q \leq I$ we have $pqp = p$ since it also follows $p - pqp = p(I-q)p \geq 0$). Let $x = x^* \in \mathcal{V}$ and let $r>0$ such that $rp - pxp \geq 0.$ Then $$ \begin{aligned} p(rq -x)p = rpqp - pxp \geq rp - pxp \geq 0. \end{aligned} $$
200603094/7
for all $i,j,k,$ which in turn yields that $pa_{ijk}p = 0$ and thus $a_{ijk} \in C(p) \cap -C(p).$ This then yields \begin{align*} x = \sum_{ijk} \ket{i}\bra{j} \otimes \lambda_ka_{ijk} \in M_n(J_p), \end{align*} finishing the proof. Lemma3.5. Given an operator system $\mathcal{V} \subset B(H)$ with projection $p \in \mathcal{V}$, then for all $n \in \mathbb{N}$, $J_{p_n} \subset M_n(\mathcal{V})$ is $*$ -closed and if $\alpha \in M_{n,m}$ then $\alpha^*J_{p_n}\alpha \subset J_{p_m}.$ In particular $p_nJ_{p_n}p_n \subset J_{p_n}.$ Proof. Let $x \in J_{p_n}$. Then $x = \sum_{ijk} \ket{i}\bra{j} \otimes \lambda_k a_{ijk}$ with, as we saw in the proof of Lemma, $a_{ijk} \in C(p) \cap -C(p)$ for all $i,j,k,$ and therefore $a_{ijk}^*= a_{ijk}.$ Thus $x^* = \sum_{ijk} \ket{j}\bra{i} \otimes \overline{\lambda}_ka_{ijk} \in J_{p_n}.$ It is immediate that for any $x \in J_{p_n}$ that $p_nxp_n \in J_{p_n}$ since if we write $x = \sum_k \lambda_k x_k,$ with $x_k \in C(p_n) \cap -C(p_n)$ then $p(px_kp)p = px_kp \in B(H^n)^+.$ This holds for all $k$ and thus $p_nJ_{p_n}p_n \subset J_{p_n}.$ Finally if $x \in J_{p_n}$ and $\alpha \in M_{n,m}$ we have \begin{align*} \pm p_m\alpha^*x\alpha p_m = \alpha^*(\pm p_nxp_n)\alpha \in B(H^m)^+. \end{align*} Thus, $\alpha^*x\alpha \in J_{p_m}.$ We now wish to consider the vector space $\mathcal{V} / J_p$ where $\mathcal{V} \subset B(H)$ is an operator system, and $p \in \mathcal{V}$ is a projection in $B(H)$. The vector space operations and involution are defined in the natural way and are well-defined by Lemma 3.5. We consider the family $\{\widetilde{C}(p_n)\}_n$ where for each $n \in \mathbb{N}$ we have $$ \begin{aligned} \widetilde{C}(p_n):= \{(x_{ij} + J_p) \in M_n(\mathcal{V}/ J_p): x = (x_{ij}) \in C(p_n)\}. \end{aligned} $$ Note that $p \notin J_p,$ for $p$ a nonzero projection in $\mathcal{V}$. This is immediate since if $p = \sum_k \lambda_k c_k, \pm c_k \in C(p) \cap -C(p)$ then it would follow that $$ \begin{aligned} p = ppp = p(\sum_k\lambda_kc_k)p = \sum_k \lambda_k pc_kp, \end{aligned} $$ and since $\pm pc_kp \in B(H)^+$ for all $k$ implies $pc_kp = 0$. Thus, $p = 0.$ Theorem3.6. The triple $(\mathcal{V}/ J_p,\{\widetilde{C}(p_n)\}_n,p+ J_p)$ is an operator system. Proof. We begin by showing that $\{\widetilde{C}(p_n)\}_n$ is a proper matrix ordering. Let $(x_{ij} + J_p) \in M_n(\mathcal{V}/ J_p).$ If $\lambda >0$ then $\lambda (x_{ij} + J_p) = (\lambda x_{ij} + J_p)$ and $\lambda x \in \lambda C(p_n) \subset C(p_n)$, thus $\lambda (x_{ij} + J_p) \in \widetilde{C}(p_n).$ If $(x_{ij} + J_p), (y_{ij} + J_p) \in \widetilde{C}(p_n)$ then \begin{align*} (x_{ij} + J_p) + (y_{ij} + J_p) = (x_{ij} + y_{ij}) + J_p \in \widetilde{C}(p_n) \end{align*} since $x+ y \in C(p_n).$ If $m \in \mathbb{N}$ and $(x_{ij} + J_p) \in \widetilde{C}(p_n)$ and $\alpha \in M_{n,m}$ then \begin{align*} \alpha^*(x_{ij} + J_p)\alpha = \sum_{ijkl} \ket{i}\bra{j} \otimes (\overline{\alpha}_{ki}x_{kl}\alpha_{lj} + J_p) = (\sum_{kl}\overline{\alpha}_{ki}x_{kl}\alpha_{lj} + J_p)_{ij}, \end{align*} and $\alpha^*x\alpha \in C(p_m)$ which implies $\alpha^*(x_{ij} + J_p)\alpha \in \widetilde{C}(p_m).$ If $(x_{ij} + J_p) \in \widetilde{C}(p_n)$ then $(x_{ij} + J_p)^* = (x_{ji}^*+ J_p)$ and since $x \in C(p_n)$ implies $x^* \in C(p_n)$ which implies $(x_{ij} + J_p)^* = (x_{ji}^*+ J_p) \in \widetilde{C}(p_n).$ Finally, suppose that $(x_{ij} + J_p) \in \widetilde{C}(p_n) \cap - \widetilde{C}(p_n).$ This implies that $\pm x \in C(p_n).$ Then it follows $x \in J_{p_n} = M_n(J_p)$ by Lemma and thus $x_{ij} \in J_p$ for all $i,j.$ In other words $(x_{ij} + J_p) = (0 + J_p)$ implying that our cones are proper, and thus the matrix ordering $\{\widetilde{C}(p_n)\}_n$ on $\mathcal{V}/ J_p$ is proper. It remains to show that $p + J_p$ is an Archimedean matrix order unit. Consider
200603094/12
We now briefly return to the structural properties induced by projections in a concrete operator system $\mathcal{V} \subset B(H).$ The following lemma will motivate Definition 4.6 below for a matrix ordering $\{C(p_n)\}$ when $p$ is only a positive contraction in an abstract operator system. Lemma4.5. Let $\mathcal{V} \subset B(H)$ be an operator system and suppose that $p \in \mathcal{V}$ is a projection. Then for any $x \in \mathcal{V}$ with $x = x^*$, we have that $pxp \geq 0$ in $B(H)$ if and only if for every $\epsilon > 0$ there exists a $t > 0$ such that $$ x + \epsilon p + t (I - p) \geq 0. $$ Proof. First assume that for every $\epsilon > 0$ there exists a $t > 0$ such that $y = x + \epsilon p + t (I - p) \geq 0$. Then the compression of $y$ by $p$ is positive. Hence $pyp = pxp + \epsilon p \geq 0$. Since this holds for all $\epsilon > 0$ and since $p$ is an Archimedean order unit for the set of operators of the form $pzp$ (see Proposition 3.1) it follows that $pxp \geq 0$. Now assume that $pxp \geq 0$, and let $\epsilon > 0$. Let $q = I - p$. It follows that if we write $H = pH \oplus qH$, we see that (by exercise 3.2(i) in [15]) an operator $T$ is positive if and only if $pTp \geq0$, $qTq \geq 0$, and for every $h \in pH$ and $k \in qH$ we have that $$ \left\lvert \langle pTqk | h \rangle \right\rvert^2 \leq \langle pTp h | h \rangle\langle qTq k | k \rangle. $$ Now choose $t > \|x\|$ such that $\epsilon(t - \|x\|) > \|x\|^2$ and consider $T = x + \epsilon p + tq$. Then $qTp = qxp$, $pTp = pxp + \epsilon p \geq 0$, and $qTq = tq + qxq \geq tq - \|x\|q \geq 0$. Moreover $$ |\langle pTq k | h \rangle|^2 \leq \|x\|^2 \|h\|^2 \|k\|^2 \leq \epsilon(t - \|x\|) \|h\|^2 \|k\|^2 \leq \langle pTp h | h \rangle\langle qTq k | k \rangle, $$ since $\epsilon \|h\|^2 = \langle \epsilon p h | h \rangle \leq \langle pTp h | h \rangle$ and $(t-\|x\|)\|k\|^2 = \langle (tq - \|x\|q) k | k \rangle \leq \langle qTq k | k \rangle$. So $x + \epsilon p + tq \geq 0$. This lemma thus relates positivity of the compression by $p$ to positivity in the operator system $\mathcal{V}.$ This motivates us to make the following definition. Definition4.6. Let $(\mathcal{V},\{C_n\}_n, e)$ be an operator system, and suppose that $p \in \mathcal{V}$ with $0 \leq p \leq e$, i.e., let $p \in \mathcal{V}$ be a positive contraction of $\mathcal{V}$. For each $n \in \mathbb{N}$ and let $p_n= I_n \otimes p$. We define the positive cone relative to $p_n$, denoted $C(p_n)$, to be $$ \begin{aligned} C(p_n) :&= \{x \in M_n(\mathcal{V}) : x = x^*, \text{for all} \epsilon > 0 \text{there exists} t > 0 \text{such that} x + \epsilon p_n + t(e_n - p_n) \in C_n \}. \end{aligned} $$ An immediate consequence of Lemma 4.5 is that if a positive contraction $p \in \mathcal{V}$ is a projection, then for each $n \in \mathbb{N}$ the positive cone relative to $p_n$ becomes $$ \begin{aligned} C(p_n) = \{x \in M_n(\mathcal{V}): x=x^*, p_nxp_n \in B(H^{n})^+\}, \end{aligned} $$ We now prove a similar string of results mirroring those of Section 3. Proposition4.7. Let $\mathcal{V}$ be an operator system and $0 \leq p \leq e$. Then the sequence $\{C(p_n)\}_n$ is a matrix ordering for $\mathcal{V}$, and $p$ is an Archimedean matrix order unit for $\{C(p_n)\}_n$. Proof. We first check that $\{C(p_n)\}_n$ is a matrix ordering. As noted in the preliminaries, it suffices to check that For each $n,m \in \mathbb{N}$, $C(p_n) \oplus C(p_m) \subseteq C(p_{n+m})$, and for each $n,m \in \mathbb{N}$ and $\alpha \in M_{n,m}$, $\alpha^* C(p_n) \alpha \subseteq C(p_m)$. Suppose $x \in C(p_n)$ and $y \in C(p_m)$. Let $\epsilon > 0$. Then there exist $r_1, r_2 > 0$ such that $x + \epsilon p_n + r_1(e_n-p_n) \geq 0$ and $y + \epsilon p_m + r_2(e_m - p_m) \geq 0$. It follows that $(x\oplus y) + \epsilon p_{n+m} + \max(r_1,r_2) (e_{n+m} - p_{n+m}) \geq 0$. Now, let $\epsilon > 0$ and suppose that $x \in C(p_n)$. Also assume $\alpha \neq 0$. Then there exists $r > 0$ such that $$ x + \frac{\epsilon}{\|\alpha\|^2} p_n + r(e_n - p_n) \geq 0. $$
200603094/14
Proof. By Proposition 4.7 we already have that $\{C(p_n)\}_n$ is a matrix ordering and $p$ is an Archimedean matrix order unit for the pair $(\mathcal{V}, \{C(p_n)\}_n).$ By Proposition, we deduce that $(\mathcal{V} / J_p, \{\tilde{C}(p_n)\}_n, p + J_p)$ is an operator system. Moreover, this operator system is non-trivial by Proposition 4.8, since $\alpha_m(p)=1$. Combining Theorem 4.9, Lemma 4.5 and Theorem 3.6 yields the following key observation. Corollary 4.10. Suppose that $\mathcal{V} \subset B(H)$ is an operator system and that $p \in \mathcal{V}$ is a projection in $B(H)$. Then the abstract compression $(\mathcal{V}/J_p, \{\tilde{C}(p_n)\}_{n \in \mathbb{N}}, p + J_p)$ is completely order isomorphic to the concrete compression $p \mathcal{V} p$. Corollary 4.10 justifies the terminology we will use in the following definition. Definition4.11. Given an operator system $\mathcal{V}$ and a positive contraction $p \in \mathcal{V}$ such that $\alpha_m(p) = 1$ then we call the operator system $({\mathcal{V}/J_p}, \{\widetilde{C}(p_n)\}_{n \in \mathbb{N}}, {p + J_p})$ the abstract compression operator system and denote it by $\mathcal{V}/ J_p.$ In the next section of the paper we will make use of the structure of the abstract compression operator system $M_2(\mathcal{V}) / J_{p \oplus q}$ where $p$ is a positive contraction and $q = e - p$. We denote the positive cones relative to the positive contraction $p \oplus q$ by $\{C((p \oplus q)_n)\}_n$ where for each $n \in \mathbb{N}$ $$ \begin{aligned} C((p \oplus q)_n) = \{x \in M_{2n}(\mathcal{V}): x=x^*, \forall \epsilon >0 \,\, \exists t >0 \,\,\text{such that}\,\, x + \epsilon((p\oplus q)_n) + t((q\oplus p)_n) \in C_{2n}\}, \end{aligned} $$ and $J_{p \oplus q} = \Span C(p \oplus q) \cap -C(p \oplus q).$ The concrete analogue of the following corollary was stated in Corollary 3.14. Corollary 4.12. Given an operator system $\mathcal{V}$ and a finite family of positive contractions $\{p^i\}_{i=1}^N \subset \mathcal{V}^+$ such that $\alpha_m(p^i) = 1$ for some $i$, with $q^i = e - p^i$ for all $i,$, let $P = \oplus_i p^i,$ and $Q = \oplus q^i.$ Then $M_{2N}(\mathcal{V})/ J_{P \oplus Q}$ is an operator system. ## Projections in operator systems In this section we will develop an abstract characterization for projections in operator systems. We start with the following useful observation. Lemma5.1. Let $\mathcal{V} \subset B(H)$ be an operator system and suppose that $p \in \mathcal{V}$ is a projection (when viewed as an operator on $H$). If $q = I - p$ and $a,b,c \in \mathcal{V}$ with $a^* = a$ and $b^*=b$, then $pap + pcq + qc^*p + qbq \in B(H)^+$ if and only if $$ \begin{pmatrix} a & c \\ c^* & b \end{pmatrix} \in C(p \oplus q). $$ Proof. First suppose that $$ \begin{pmatrix} a & c \\ c^* & b \end{pmatrix} \in C(p \oplus q). $$ Then by Lemma 4.5 we know that $$ \begin{pmatrix} p & 0 \\ 0 & q \end{pmatrix} \begin{pmatrix} a & c \\ c^* & b \end{pmatrix} \begin{pmatrix} p & 0 \\ 0 & q \end{pmatrix} \geq 0. $$ Conjugating this matrix by the scalar matrix $\begin{pmatrix} 1 \\ 1 \end{pmatrix}$ yields the expression $pap + pcq + qc^*p + qbq$ and hence $pap + pcq + qc^*p + qbq \geq 0$. Now suppose that $pap + pcq + qc^*p + qbq \geq 0$. Again, by Lemma 4.5 it suffices to prove that the operator $$ T = \begin{pmatrix} p & 0 \\ 0 & q \end{pmatrix} \begin{pmatrix} a & c \\ c^* & b \end{pmatrix} \begin{pmatrix} p & 0 \\ 0 & q \end{pmatrix} $$
200603094/2
is a complete order isomorphism, where $q = e - p$ and $e$ is the unit of $\mathcal{V}$ (Definition 5.4). Our main result is that a positive element $p \in \mathcal{V}$ is an abstract projection if and only if there exists a Hilbert space $H$ and a unital complete order embedding $\pi: \mathcal{V} \rightarrow B(H)$ such that $\pi(p)$ is a projection (Theorem 5.10), establishing a one-to-one correspondence between abstract and concrete projections. Similar to a result of Blecher and Neal appearing in [1,Lemma2.3], we then prove in Theorem 5.11 that $p \in \mathcal{V}$ is an abstract projection if and only if $p$ is a projection in the C*-envelope $C_e^*(\mathcal{V}).$ These observations lead us quickly to a new characterization of the set of quantum commuting correlations $C_{qc}(n,k)$ in terms of abstract operator systems and abstract projections. Specifically, we show that $\{p(a,b|x,y)\} \in C_{qc}(n,k)$ if and only if $p(a,b|x,y) = \phi(Q(a,b|x,y))$ where $\phi$ is a state on an operator system $\mathcal{V}$ spanned by abstract projections $\{Q(a,b|x,y)\}$ satisfying certain operator non-signalling conditions, namely that $$ \sum_{a,b} Q(a,b|x,y) = e $$ for all $x,y$ (where $e$ is the unit of $\mathcal{V}$) and that the marginal operators $$ E(a|x) := \sum_b Q(a,b|x,y) \quad \text{and} \quad F(b|y) := \sum_b Q(a,b|x,y) $$ are well defined (Theorem 6.3). We conclude this introduction by pointing out some related results in the literature. The idea of abstractly char- acterizing certain types of operators in operator spaces was investigated extensively by Blecher and Neal [1, 2], who studied the abstract (linear-metric) structure of an operator space that contains a unit. More precisely, given a pair $(\mathcal{E}, u)$ where $\mathcal{E}$ is an abstract operator space and $u \in \mathcal{E}$, they characterized when there exists a completely isometric embedding of $\mathcal{E}$ into $B(H)$ such that $u$ is mapped to a unitary. They also showed that given a unitary $u \in \mathcal{E}$, then $u$ is necessarily a unitary in the ternary envelope $T(\mathcal{E})$ of $\mathcal{E}$. Other operator system characterizations for correlation sets can be found in the literature, for example in Theorem 3.1 of [13] and Theorem 2.4 of [7]. These papers characterize correlations as having the form $p(a,b|x,y) = \phi(E_{x,a} \otimes E_{y,b})$ where $\{E_{x,a}\}_a$ are projection-valued measures spanning a certain canonical operator system $S$ and $\phi$ is a state on $S \otimes_t S$, where $\otimes_t$ denotes the various operator system tensor products of [10] depending on which type of correlation one intends to construct. Our results differ in that we do not appeal to the hierarchy of operator system tensor products or make use of any canonical operator system. We do not, however, have any result analogous to theirs for quantum approximate or local correlations. Our paper is organized as follows. In Section 2, we provide some elementary background in operator systems and operator spaces, including some preliminary remarks that will be useful throughout the paper. In Section 3 we study the compression of a concrete operator system $\mathcal{V}$ by a projection $p \in \mathcal{V}$, i.e., the operator system $p \mathcal{V} p \subseteq B(pH)$. The observations of this section motivate the results of Section 4, where we provide an abstract definition for the compression of an abstract operator system by a positive contraction $p$ which is not a priori a projection. In Section 5, we combine the results of Section 3 and Section 4 and some additional observations to prove the main results of the paper. We conclude with Section 6 where we present our applications to the theory of quantum correlation sets. ## Preliminaries Though we assume some familiarity with general operator system and operator space theory we will review some definitions and constructions that appear throughout the manuscript. Definition2.1. Given a Hilbert space $H$ a concrete operator system is a self-adjoint unital subspace $\mathcal{V}$ of $B(H)$. An abstract operator system is defined to be the triple $({\mathcal{V}}, \{C_n\}_{n \in \mathbb{N}}, {e})$ where $\mathcal{V}$ is a complex vector space with a conjugate- linear involution $*$ (i.e. a $*$ -vector space), $\{C_n\}_n$ is a proper matrix ordering on $\mathcal{V}$ and $e$ is an Archimedean matrix order unit. By a matrix ordering on $\mathcal{V}$, we mean a sequence $\{C_n\}_n$ which satisfies the following two properties: $C_n \subset M_n(\mathcal{V})_h$ is - a cone whose elements are invariant under the involution $*$; 2. Given any $n,m \in \mathbb{N}$ and $\alpha \in M_{n,m}$ then $\alpha^*C_n\alpha \subset C_m.$
200603094/16
We call $0$ the zero projection and necessarily $\alpha_m(0) = 0.$ Remark5.5. Let $\mathcal{V}$ be an operator system. Then $e \in \mathcal{V}$ is an abstract projection. First note that $q = e - e = 0$ and $\alpha_m(e) = 1$. Finally we consider $\pi_e: \mathcal{V} \rightarrow M_2(\mathcal{V})/ J_{e \oplus 0}$. $\pi_e$ is completely positive and if $(\pi_e)_n(x) \in \widetilde{C}_{2n}(e \oplus 0)$ then $\begin{pmatrix} x & x \\ x & x \end{pmatrix} \in C_{2n}(e \oplus 0)$and thus for all$\epsilon >0$there exists$t >0$such that$\begin{pmatrix} x & x \\ x & x \end{pmatrix} + \epsilon(e \oplus 0)_n + t(0 \oplus e)_n \in C_{2n}.$Compression to the (1,1) corner implies$x + \epsilon e \in C_n$for all$\epsilon >0$and therefore$x \in C_n$. Thus, we have$\pi_e$is a complete order embedding. Thus,$e$ is an abstract projection. Remark5.6. It necessarily follows that $p$ is an abstract projection in an operator system $\mathcal{V}$ if and only if $q = e - p$ is an abstract projection. Indeed, the assertion that $\pi_p$ is a order embedding is equivalent to the assertion that $x \in C_n$ if and only if $\begin{pmatrix} 1 & 1 \\ 1 & 1 \end{pmatrix} \otimes x \in C_{2n}((p \oplus q)_n)$. In other words, $x \in C_n$ if and only if for every $\epsilon > 0$ there exists $t > 0$ such that $$ \begin{pmatrix} 1 & 1 \\ 1 & 1 \end{pmatrix} \otimes x + \epsilon (p \oplus q) \otimes I_n + t (q \oplus p) \otimes I_n \in C_{2n}. $$ Conjugation by the unitary matrix $\begin{pmatrix} 0 & 1 \\ 1 & 0 \end{pmatrix} \otimes I_n$ shows that for every $\epsilon > 0$ there exists $t > 0$ such that $$ \begin{pmatrix} 1 & 1 \\ 1 & 1 \end{pmatrix} \otimes x + \epsilon (q \oplus p)_n + t (p \oplus q)_n \in C_{2n}. $$ Consequently $\begin{pmatrix} 1 & 1 \\ 1 & 1 \end{pmatrix} \otimes x \in C_{2n}((q \oplus p)_n)$ if and only if $x \in C_n$. The reverse implication is the same argument. Remark 5.7. The requirement that $\alpha_m(p) = 1$ is clearly necessary for $p$ to be a projection. However, even when $\alpha_m(p)<1$ it is possible for $\alpha_m(q) = 1$ so that $M_2(V)/J_{p \oplus q}$ is non-trivial. It is instructive to see what happens in this case. So suppose $\alpha_m(p) < 1$, or equivalently $\|p\|<1$. Then for every $\epsilon < 1$ and for sufficiently large $t > 0$ we have $$ -\begin{pmatrix} p & p \\ p & p \end{pmatrix} + \epsilon \begin{pmatrix} p & 0 \\ 0 & q \end{pmatrix} + t \begin{pmatrix} q & 0 \\ 0 & p \end{pmatrix} = \begin{pmatrix} te + (\epsilon - t - 1)p & -p \\ -p & \epsilon e + (t - \epsilon - 1)p \end{pmatrix} \geq 0. $$ To see that this is true, first observe that because $\|p\| < 1$, we can choose $t$ sufficiently large so that $$ p \left(\frac{1 - \epsilon + t}{t} \right) \leq e \text{and} \|p\|^2 \leq t \epsilon \left(1 - \frac{\|p\|(1 - \epsilon + t)}{t} \right). $$ The bound on $\|p\|^2$ holds because $$ \left(1 - \frac{\|p\|(1 - \epsilon + t)}{t} \right) \rightarrow 1 - \|p\| $$ as $t \rightarrow \infty$, and thus $$ t \epsilon \left(1 - \frac{\|p\|(1 - \epsilon + t)}{t} \right) \rightarrow \infty $$ as $t \rightarrow \infty$. Now let $\phi:V \rightarrow B(H)$ be any positive unital map. Then for any $h,k \in H$ we have $$ \begin{aligned} |\langle \phi(p) h, k \rangle|^2 & \leq & \|p\|^2 \|h\|^2 \|k\|^2 \\ & \leq & t \epsilon \left(1 - \frac{\|p\|(1 - \epsilon + t)}{t} \right) \|h\|^2 \|k\|^2 \\ & = & (t - \|p\|(1 - \epsilon + t)) \|h\|^2 (\epsilon \|k\|^2) \\ & \leq & (t - \|p\|(1 - \epsilon + t)) \|h\|^2 (\epsilon \|k\|^2 + \langle (t - \epsilon - 1)\phi(p) k| k \rangle) \\ & \leq & (t \|h\|^2 - t\langle \frac{1 - \epsilon + t}{t} \phi(p) h| h \rangle) (\epsilon \|k\|^2 + \langle(t - \epsilon - 1)\phi(p) k| k \rangle) \\ & = & \langle (t I + (\epsilon - t - 1)\phi(p) h| h \rangle \langle (\epsilon I + (t - \epsilon - 1) \phi(p)) k| k \rangle \end{aligned} $$
200603094/11
Proof. We first show that $M_n(J) \subseteq J_n$. It is necessary and sufficient to show that for every $j,k \leq n$ and $x \in J$ we have $\ket{j} \bra{k} \otimes x \in J_n$. Let $x \in J$. By the previous lemma, $x = a + ib$ for $a,b \in C_1 \cap -C_1$. Then $\ket{j} \bra{j} \otimes a, \ket{j} \bra{j} \otimes b \in C_n \cap -C_n$ since $\{C_n\}$ is a matrix cone. Hence $\pm \ket{j} \bra{j} \otimes x \in J_n$. Similarly, $(\ket{j} + \ket{k})(\bra{j} + \bra{k}) \otimes x \in J_n$ and $(\ket{j} + i \ket{k})(\bra{j} -i\bra{k})\otimes x \in J_n$. So $$ ((\ket{j} + \ket{k})(\bra{j} + \bra{k}) - \ket{j}\bra{j} - \ket{k}\bra{k})\otimes x = (\ket{j}\bra{k} + \ket{k}\bra{j}) \otimes x \in J_n $$ and $$ i((\ket{j} + i \ket{k})(\bra{j} -i\bra{k}) - \ket{j}\bra{j} - \ket{k}\bra{k}) \otimes x = (\ket{j}\bra{k} - \ket{k}\bra{j}) \otimes x \in J_n. $$ It follows that $\ket{j} \bra{k} \otimes x \in J$. Next we show $J_n \subseteq M_n(J)$. Let $x = (x_{jk}) \in J_n$. Then it is necessary and sufficient to show that $x_{jk} \in J$ for every $j,k$. By the previous lemma, $x = a + ib$ for $a,b \in J_n$, so it suffices to show that $a_{jk}$ and $b_{jk}$ are elements of $J$. We will show that $a_{jk} \in J$, and the proof for $b_{jk}$ is identical. Since $a \in C_n \cap -C_n$ and $\{C_n\}_n$ is a matrix cone, we have $a_{jj} = \bra{j} a \ket{j}, a_{kk} = \bra{k} a \ket{k} \in J$. Also $(\bra{j} + \bra{k})a(\ket{j} + \ket{k}) = a_{jj} + a_{jk} + a_{kj} + a_{kk} \in J$, and $(\bra{j} + i \bra{k})a(\ket{j} - i \ket{k}) = a_{jj} - ia_{jk} + ia_{kj} + a_{kk} \in J$. It follows that $a_{jk} + a_{kj}$ and $a_{jk} - a_{kj}$ are elements of $J$ and hence $a_{jk} \in J$. Similarly $b_{jk} \in J$ and hence $x_{jk} \in J$. Lemma allows us to identify the vector spaces $M_n(\mathcal{V}/J)$, $M_n(\mathcal{V}) / M_n(J)$ and $M_n(\mathcal{V}) / J_n$. Define $$ \tilde{C}_n := \{(x_{ij} + J) \in M_n(\mathcal{V}/ J): x=(x_{ij}) \in C_n\}. $$ Lemma4.3. The sequence $\{\tilde{C}_n\}$ is a matrix ordering on $\mathcal{V}/J$. Proof. That $\lambda \widetilde{C}_n \subset \widetilde{C}_n$ and $\widetilde{C}_n + \widetilde{C}_n$ are immediate. To check compatibility, fix $n \in \mathbb{N}$ and let $\alpha \in M_{n,m}$ and $(x_{ij} + J) \in \widetilde{C}_n$. Then \begin{align*} \alpha^*(x_{ij}+J)\alpha &= \sum_{i,j}\ket{i}\bra{j} \otimes \sum_{k,l}\overline{\alpha}_{ki}(x_{kl}+J)\alpha_{lj} \\ &= \sum_{i,j}\ket{i}\bra{j} \otimes (\sum_{k,l}\overline{\alpha}_{ki}x_{kl}\alpha_{lj}+J) \\ &= (\sum_{k,l}\overline{\alpha}_{ki}x_{kl}\alpha_{lj} + J)_{ij}, \end{align*} and $\alpha^*x\alpha \in \alpha^*C_n\alpha \subset C_m.$ This finishes the proof. Suppose $\{C_n\}_n$ is a (not necessarily proper) matrix ordering on a $*$ -vector space $\mathcal{V}$. Recall that $e \in \mathcal{V}$ a matrix order unit for $(\mathcal{V},\{C_n\}_n)$ if for every $x \in M_n(\mathcal{V})$ with $x^*=x$, there exists $r > 0$ such that $x + re_n \in C_n$. If $x + \epsilon e_n \in C_n$ for all $\epsilon > 0$ implies that $x \in C_n$, we call $e$ Archimedean. Suppose that $\mathcal{V}$ is a $*$ -vector space with matrix ordering $\{C_n\}_n$ and an Archimedean matrix order unit $e$. Then $(\mathcal{V}/J, \{\tilde{C}_n\}_n, e + J)$ is an operator system. Proof. By Lemma 4.3 we need only show that the family $\{\widetilde{C}_n\}_n$ is proper and that $e + J$ is an Archimedean matrix order unit. If $(x_{ij}+ J) \in \widetilde{C}_n \cap -\widetilde{C}_n$ then $\pm x \in C_n$ which implies $x \in J_n = M_n(J)$ by Lemma. Thus, $x_{ij} \in J$ for all $i,j$ and therefore $(x_{ij} +J) = (0 + J).$ It remains to show that $e+J$ is an Archimedean matrix order unit. Let $(x_{ij} + J) \in M_n(\mathcal{V}/ J)$ be $*$ -hermitian and choose $r>0$ such that $re_n -x \in C_n$. Then \begin{align*} r(I_n \otimes (e+ J)) - (x_{ij}+J) = (I_n \otimes (re + J)) - (x_{ij}+J)= (r(e_n)_{ij} + J)- (x_{ij}+J) = (r(e_n)_{ij} - x_{ij} + J) \in \widetilde{C}_n \end{align*} since $re_n -x \in C_n.$ Finally, if $\epsilon I_n\otimes(e+J) + (x_{ij} + J) \in \widetilde{C}_n$ for all $\epsilon>0$ then $(\epsilon(e_n)_{ij}+x_{ij} + J) \in \widetilde{C}_n$ for all $\epsilon>0$ and by definition $\epsilon e_n + x \in C_n$ giving us $x \in C_n$ and thus $(x_{ij}+J) \in \widetilde{C}_n$. This finishes the proof.
200603094/6
Thus $rq - x \in C(p)$ implying $q$ is an order unit. If for all $\epsilon >0, \epsilon q + x \in C(p)$ then $\epsilon pqp + pxp \geq 0$ for all $\epsilon >0$ and thus \begin{align*} \epsilon p + pxp \geq \epsilon pqp + pxp \geq 0, \forall \epsilon >0, \end{align*} implying $pxp \geq 0$ (and thus $x \in C(p)$) by our earlier remarks on $p$, and thus $q$ is an Archimedean order unit. The same observations hold for $p_n$ and $q_n$ for all $n \in \mathbb{N}$ and thus the same method will show that $q$ is an Archimedean matrix order unit for $(\mathcal{V}, \{C(p_n)\}_n)$. The next corollary, which follows from Proposition 3.1, states that the cones $C(p_n)$ are closed in the order seminorm induced by the projection $p$. Corollary3.2. Given an operator system $\mathcal{V} \subset B(H)$ and projection $p \in \mathcal{V}$ then if $\{C(p_n)\}_n$ denotes the matrix ordering as in Lemma 3.1 then for each $n \in \mathbb{N}$, $C(p_n)$ is closed in the order seminorm. Proof. For this proof let $\alpha_n: M_n(\mathcal{V}) \rightarrow [0,\infty)$ denote the order seminorm defined for each $n \in \mathbb{N}$ by $$ \alpha_n(x):= \inf \{r >0: \begin{pmatrix} rp_n & x \\ x^* & rp_n \end{pmatrix} \in C(p_{2n})\}. $$ Let $\{x_i\}_{i \in I} \subset C(p_n)$ denote a net such that $x = \alpha_n$ - $\lim_i x_i$ ($x$ is the limit of $x_i$ relative to the norm $\alpha_n$). $x$ is necessarily $*$ -hermitian. Let $r >0$ and let $i_o \in I$ such that $\alpha_n(x_i -x) < r$ for $i_o \preceq i.$ It follows that $$ \begin{pmatrix} rp_n & x - x_i \\ x - x_i & rp_n \end{pmatrix} \in C(p_{2n}) $$ and thus compression by $\begin{pmatrix} 1 \\ 1 \end{pmatrix}$implies$rp_n + x -x_i \in C(p_n)$and therefore$rp_n + x \in C(p_n)$. Since$p$is an Archimedean matrix order unit we have$x \in C(p_n)$ as desired. Remark 3.3. Used in Proposition 3.1 we make the point of saying that if $\varphi: \mathcal{V} \rightarrow B(H)$ is the map defined by $\varphi(v) = pvp$ then $\varphi_n(v) = p_nvp_n,v \in M_n(\mathcal{V}).$ First note that compression by the projection is linear completely positive. Notice, \begin{align*} p_nvp_n = (I_n \otimes p)(\sum_{ij}\ket{i}\bra{j} \otimes v_{ij})(I_n \otimes p) = \sum_{ij} \ket{i}\bra{j} \otimes pv_{ij}p = \sum_{ij}\ket{i}\bra{j} \otimes \varphi(v_{ij}) = \varphi_n(v), \end{align*} Lemma3.4. Given an operator system $\mathcal{V} \subset B(H)$ let $p \in \mathcal{V}$ be a projection as above. Let $J_p = \Span C(p) \cap -C(p),$ and $J_{p_n} = \Span C(p_n) \cap -C(p_n).$ Then $M_n(J_p) = J_{p_n}$. Proof. If $x \in M_n(J_p)$ then we write $x = \sum_{ij} \ket{i}\bra{j} \otimes x_{ij}, x_{ij} \in J_p.$ For each $i,j$ we then write $x_{ij} = \sum_k \lambda_{ijk} a_{ijk}$ where $\lambda_{ijk} \in \mathbb{C}, a_{ijk} \in C(p) \cap -C(p).$ We then see \begin{align*} p_nxp_n = p_n(\sum_{ijk} \lambda_{ijk} \ket{i}\bra{j} \otimes a_{ijk})p_n = \sum_{ijk} \lambda_{ijk} \ket{i}\bra{j} \otimes pa_{ijk}p = 0, \end{align*} since $\pm pa_{ijk}p \in B(H)^+$ for all $i,j,k$ and therefore is equal to 0. Thus $x \in C(p_n) \cap -C(p_n)$ which yields the first inclusion. Conversely, let $x \in J_{p_n}$ and write $x = \sum_k \lambda_k x_k, \lambda_k \in \mathbb{C}, x_k \in C(p_n) \cap -C(p_n).$ We write $$ x_k = \sum_{ij} \ket{i}\bra{j} \otimes a_{ijk} \in M_n(\mathcal{V}). $$ By the assumption that $p_nx_kp_n = 0$ we see that \begin{align*} p_nx_kp_n = \sum_{ij} \ket{i}\bra{j} \otimes pa_{ijk}p = 0, \end{align*}
200603094/4
Here $\{\ket{i}\bra{j}\}_{i,j \leq n}$ denotes the matrix units of $M_n$ and $\{\ket{k}\bra{l}\}_{k,l \leq m}$ denotes the matrix units of $M_m.$ We see $$ \begin{aligned} \sum_{ij}\ket{i}\bra{j} \otimes A_{ij} = \sum_{ij} \ket{i}\bra{j} \otimes \sum_{kl} \ket{k}\bra{l} \otimes a_{ijkl} = \sum_{kl} \ket{k}\bra{l} \otimes \sum_{ij} \ket{i}\bra{j} \otimes a_{ijkl} = \sum_{kl} \ket{k}\bra{l} \otimes B_{kl} = B, \end{aligned} $$ where $B_{kl} = \sum_{ij} \ket{i}\bra{j} \otimes a_{ijkl},$ and $B \in M_m(M_n(\mathcal{V})).$ This canonical map is a $*$ -isomorphism between the ambient $C^*$ -algebras, i.e., betweeen $M_n(M_m(B(H)))$ and $M_m(M_n(B(H)))$. One may also view this using the commutativity of the operator system minimal tensor product, i.e., $M_n(M_m(\mathcal{V})) \simeq M_n \otimes_{\text{min}} M_m \otimes_{\text{min}} \mathcal{V} \simeq M_m \otimes_{\text{min}} M_n \otimes_{\text{min}} \mathcal{V} \simeq M_m(M_n(\mathcal{V})).$ Given an operator system $\mathcal{V}$ consider now a pair $(\kappa, \mathcal{A})$ where $\kappa: \mathcal{V} \rightarrow \mathcal{A}$ is a unital complete order embedding, and $\mathcal{A}$ is a C*-algebra. We will call the pair a C*-extension of $\mathcal{V}$ if $\mathcal{A}$ is generated by the image $\kappa(\mathcal{V})$ as a C*-algebra. In particular, there exists a minimal such extension satisfying a universal property. Given an operator system $\mathcal{V}$ and two C*-extensions, $(\kappa_1, \mathcal{A}_1), (\kappa_2, \mathcal{A}_2)$, we say that the extensions are $\mathcal{V}$ -equivalent if there exists a $*$ -isomorphism $\pi: \mathcal{A}_1 \rightarrow \mathcal{A}_2$ such that $\pi \kappa_1 = \kappa_2.$ Theorem2.6 (Arveson-Hamana). Given an operator system $\mathcal{V}$ then there exists a C*-extension $(\kappa,A)$ satisfying the following universal property: given any other C*-extension $(j,B)$ of $\mathcal{V}$ then there exists a unique $*$ -epimorphism $\pi: B \rightarrow A$ such that $\pi \circ j = \kappa.$ Definition2.7. Given an operator system $\mathcal{V}$ then the C*-envelope of $\mathcal{V}$ will be any C*-extension satisfying Theorem 2.6. Such a C*-extension is unique up to $\mathcal{V}$ -equivalence and we denote it by $C_{\text{e}}^*(\mathcal{V}).$ Given a Hilbert space $H$ a concrete operator space is a closed subspace $\mathcal{E} \subset B(H)$. An abstract operator space will be the pair $(\mathcal{E}, \{\alpha_n\}_n)$ where $\mathcal{E}$ is a linear space and $\{\alpha_n\}_n$ is a sequence of matrix norms on $\mathcal{E}$, that is $\alpha_n: M_n(\mathcal{E}) \rightarrow [0,\infty)$ for all $n \in \mathbb{N}$, satisfying Ruan’s axioms. This is to say that the following two properties are satisfied: $\alpha_{m+n}(x \oplus y) = \max\{\alpha_m(x), \alpha_n(y)\}$ for - all $x \in M_m(\mathcal{E}), y \in M_n(\mathcal{E})$; $\alpha_m(axb) \leq \left\lVert a \right\rVert \alpha_m(x) \left\lVert b \right\rVert$ for all $a,b \in M_m.$ When no confusion will arise we will simply denote an operator space by $\mathcal{E}.$ Given a linear map $\varphi: \mathcal{E} \rightarrow \mathcal{F}$ between operator spaces, then we say $\varphi$ is completely bounded if $\left\lVert \varphi \right\rVert_{\text{cb}}:= \sup_n \left\lVert \varphi_n \right\rVert < \infty$, and $\varphi$ will be called completely isometric if $\varphi_n$ is an isometry for all $n \in \mathbb{N}$. We identify two operator spaces if there exists a completely isometric bijection between the two. This is to say that $\mathcal{E}$ is completely isometric to $\mathcal{F}$ if there exists a linear map $\varphi: \mathcal{E} \rightarrow \mathcal{F}$ which is invertible and both $\varphi$ and $\varphi^{-1}$ are completely contractive. We will thus say that $\mathcal{E}$ is completely isometric to $\mathcal{F}.$ Due to a result of Ruan there is a one-to-one correspondence between concrete and abstract operator spaces. Theorem2.8 ([18]). Given an abstract operator space $\mathcal{E}$ then there exists a Hilbert space $H$ and a concrete operator space $\mathcal{F} \subset B(H)$ such that $\mathcal{E}$ and $\mathcal{F}$ are completely isometric. Conversely, any concrete operator space is an abstract operator space. It is natural to consider the structural properties of an operator space that contains a “ unit ”. Definition2.9 ([1]). A unital operator space is a pair $(\mathcal{E}, u)$ where $\mathcal{E}$ is an operator space and $u \in \mathcal{E}$ is such that there exists a completely isometric embedding $\varphi: \mathcal{E} \rightarrow B(H)$ such that $\varphi(u) = I_H.$ We will call $u$ a unitary of $\mathcal{E}$ if there exists a completely isometric embedding which maps $u$ to a unitary of some $B(H)$. Analogous to the discussion preceding Theorem 2.6 and the theorem itself, we may talk about ternary extensions of unital operator spaces. Consider a ternary ring of operators (TRO) $Z \subset B(K,H)$. This is to say that $Z$ is a closed subspace of $B(K,H)$ and $ZZ^*Z \subset Z.$ Then an element $u \in Z$ is called a C*-unitary if for all $z \in Z$ we have $zu^*u = z$ and $uu^*z = z.$ Given two TROs $Z$ and $W$, a ternary morphism is a linear map $\phi:Z \rightarrow W$ such that $\phi(xy^*z) = \phi(x)\phi(y)^*\phi(z)$ for each $x,y,z \in Z$. Given an operator space $\mathcal{E}$, a ternary envelope for $\mathcal{E}$ is a TRO $T(\mathcal{E})$ generated by $\mathcal{E}$ satisfying the following universal property: if $Z$ is some other ternary envelope generated by $\mathcal{E}$, then there exists a ternary morphism $\pi: Z \rightarrow T(\mathcal{E})$ extending the identity map on $\mathcal{E}$. The existence of the ternary envelope was established in [6].
200603094/15
is positive. To this end, let $h,k \in H$. Define $h_1 = ph$ and $h_2 = qk$. Note that $\langle h_1 | h_2 \rangle= 0$ since $p$ and $q$ are orthogonal projections. Let $\tilde{h} = h_1 + h_2$. Then $$ \begin{aligned} \langle T (h \oplus k) | (h \oplus k) \rangle & = & \langle \begin{pmatrix} a & c \\ c^* & b \end{pmatrix} \begin{pmatrix} h_1 \\ h_2 \end{pmatrix} | \begin{pmatrix} h_1 \\ h_2 \end{pmatrix} \rangle \\ & = & \langle (pap + pcq + qc^*p + qbq) \tilde{h} | \tilde{h} \rangle \geq 0. \end{aligned} $$ We conclude that $$ \begin{pmatrix} a & c \\ c^* & b \end{pmatrix} \in C(p \oplus q). $$ In Section 3, we observed that $\{C(p_n \oplus q_n)\}$ was a matrix ordering on $M_2(\mathcal{V})$. This is due to Lemma 3.9 which shows that $C((p \oplus q)_n)$ can be identified with $C(p_n \oplus q_n)$ via the canonical shuffle map. The next Lemma is an abstract variation on the same result. Lemma5.2. Let $\phi:M_n(M_2(\mathcal{V})) \rightarrow M_2(M_n(\mathcal{V}))$ denote the canonical shuffle map. Then $\phi(C((p \oplus q)_n) = C(p_n \oplus q_n)$ and hence $x \in C((p \oplus q)_n)$ if and only if $\phi(x) \in C(p_n \oplus q_n)$. Proof. As noted in Remark 2.5, the canonical shuffle can be written as $\phi = \psi \otimes id: M_n(M_2) \otimes \mathcal{V} \rightarrow M_2(M_n) \otimes \mathcal{V}$ where $\psi: M_n(M_2) \rightarrow M_2(M_n)$ is a $*$ -isomorphism and $id: \mathcal{V} \rightarrow \mathcal{V}$ is the identity map. Hence it is a complete order embedding. The statement follows from the observation that $$ \phi(x + \epsilon(p \oplus q)_n + t(q \oplus q)_n) = \phi(x) + \epsilon p_n \oplus q_n + t q_n \oplus p_n. $$ Our abstract characterization for projections is based on the following Theorem. Theorem5.3. Let $\mathcal{V} \subset B(H)$ be an operator system and $p \in \mathcal{V}$ be a projection. Set $q = I-p$. Then for every $n \in \mathbb{N}$ and $x \in M_n(\mathcal{V})$ we have that $x \in C_n$ if and only if $$ \begin{pmatrix} x & x \\ x & x \end{pmatrix} \in C(p_n \oplus q_n). $$ Proof. Suppose that $x \in C_n$ which implies that $p_nxp_n, q_nxq_n \geq 0.$ Applying Lemma 5.1 we have $\begin{pmatrix} x & x \\ x & x \end{pmatrix} \in C(p_n \oplus q_n)$if and only if$p_nxp_n + p_nxq_n + q_nxp_n +q_n xq_n \geq 0$. It follows \begin{align*} p_nxp_n + p_nxq_n + q_nxp_n +q_n xq_n = p_nx(p_n + q_n) + q_nx(p_n+q_n) = (p_n +q_n)x(p_n+q_n) = x \geq 0. \end{align*} This proves one direction. Conversely, suppose that $\begin{pmatrix} x & x \\ x & x \end{pmatrix} \in C(p_n \oplus q_n).$ Once again by Lemma this implies \begin{align*} 0 \leq p_nxp_n + p_nxq_n + q_nxp_n + q_nxq_n = (p_n + q_n)x(p_n+ q_n) = x. \end{align*} Definition5.4. Let $(\mathcal{V}, \{C_n\}_n, e)$ be an abstract operator system and suppose that $p \in \mathcal{V}^+ \backslash \{0\}$ with $p \leq e$ and $\alpha_m(p) = 1$. Set $q = e - p$. We call $p$ an abstract projection if the map $\pi_p: \mathcal{V} \rightarrow M_2(\mathcal{V})/ J_{p \oplus q}$ defined by $$ \pi_p: x \mapsto \begin{pmatrix} x & x \\ x & x \end{pmatrix} + J_{p \oplus q} $$ is a complete order isomorphism onto its range.
200603094/10
Given an operator system $\mathcal{V} \subset B(H)$ with projections $p,q \in \mathcal{V}$ then $\{C({p_n \oplus q_n})\}_n$ is a matrix ordering on $M_2(\mathcal{V}).$ Proof. This proof is the same as Lemma 3.10 but we will make some remarks for completeness. Once again, $*$ -closed, and positive homogeniety of the family is immediate. Let $\varphi: M_2(\mathcal{V}) \rightarrow B(H^2)$ once again denote the linear map given by $x \mapsto (p\oplus q)x(p\oplus q).$ Then compatibility is immediate from Lemma 3.9. This finishes the proof. Theorem3.13. Given an operator system $\mathcal{V} \subset B(H)$ and a projection $p \in \mathcal{V},$ with $q = I-p$ then $$ (M_2(\mathcal{V}) / J_{p_\oplus q}), \{\widetilde{C}(p_n \oplus q_n)\}_n, (p\oplus q) + J_{p \oplus q}) $$ is an operator system. Proof. The fact that $J_{p \oplus q}$ is a subspace along with Lemma 3.12 immediately gives that the family $\{\widetilde{C}(p_n \oplus q_n)\}_n$ is a matrix ordering. Given any $(x_{ij})_{ij} \in M_n(M_2(\mathcal{V}))$ one can show that for each $i,j$ that $$ \begin{aligned} x_{ij} = \begin{pmatrix} x_{ij11} & x_{ij12} \\ x_{ij21} & x_{ij22} \end{pmatrix} = \begin{pmatrix} x_{ij11} & 0 \\ 0 & x_{ij22} \end{pmatrix} + \begin{pmatrix} 0 & x_{ij12} \\ x_{ij21} & 0 \end{pmatrix} \in J_{p \oplus q} + J_{p \oplus q} \subset J_{p \oplus q}. \end{aligned} $$ Thus $\widetilde{C}(p_n \oplus q_n) \cap - \widetilde{C}(p_n \oplus q_n)$ has trivial intersection which implies the family $\{\widetilde{C}(p_n \oplus q_n)\}_n$ is proper. An exercise in matrix mechanics shows that $p \oplus q + J_{p \oplus q}$ is an Archimedean matrix order unit. This finishes the proof. Corollary3.14. Consider a family of projections $\{p^i\}_{i=1}^N \subset \mathcal{V},$ and let $q^i = I - p^i$ for all $i$. Let $P := \oplus_i p^i, Q:= \oplus_i q^i$. Then \begin{align*} M_{2N}(\mathcal{V}) / J_{P \oplus Q} \end{align*} is an operator system. Proof. This is immediate since $P, Q \in M_N(\mathcal{V})$ are projections (as operators in $B(H^N)$) with of course $I_N - P = \oplus_i I - p^i = \oplus_i q^i = Q,$ and after making the identification $M_{2N}(\mathcal{V}) = M_2(M_N(\mathcal{V})).$ Thus we see \begin{align*} (P\oplus Q)M_{2N}(\mathcal{V})(P \oplus Q) = (P\oplus Q)M_{2}(M_N(\mathcal{V}))(P \oplus Q) \end{align*} is an operator system by Theorem 3.13. ## Abstract compression operator systems Motivated by Section 3 we now wish to consider compression operator systems in an abstract sense. We begin by considering a $*$ -vector space with structure similar to that of an operator system, but lacking proper cones. We will show that a natural quotient of such a $*$ -vector space is in fact an operator system. We will then consider a particular case of a $*$ -vector space with such a matrix ordering induced by positive contractions. Finally we will show that this structure coincides with the structure of an operator system compressed by a projection, as studied in Section 3. Let $\mathcal{V}$ be a $*$ -vector space, and let $\{C_n\}_n$ be a matrix ordering on $\mathcal{V}$. For every $n \in \mathbb{N}$, let $J_n := \Span C_n \cap -C_n$. Lemma4.1. Suppose that $x \in J_n$. Then $x = a + ib$ for some $a,b \in C_n \cap -C_n$. Proof. If $x \in J_n$, then $x = \sum_k \lambda_k c_k$ where each $c_k \in C_n \cap -C_n$ and each $\lambda_k \in \mathbb{C}$. Assume $\lambda_k = r_k + is_k$ for $r_k,s_k \in \mathbb{R}$. Then $x = (\sum_k r_k c_k) + i(\sum_k s_k c_k)$. So $a = (\sum_k r_k c_k) \in C_n \cap -C_n$ and $b = (\sum_k s_k c_k) \in C_n \cap -C_n$. For convenience, set $J := J_1$. Compare the next result with Lemma 3.4. Lemma4.2. For every $n \in \mathbb{N}$, $M_n(J) = J_n$.
200603094/13
It follows that $$ \alpha^* x \alpha + \frac{\epsilon}{\|\alpha\|^2} \alpha^* p_n \alpha + r \alpha^*(e_n - p_n) \alpha \geq 0. $$ However, since $\alpha^* p_n \alpha \leq \|\alpha\|^2 p_m$ and $\alpha^*(e_n-p_n)\alpha \leq \|\alpha\|^2(e_m-p_m)$ we have $$ \alpha^* x \alpha + \epsilon p_m + (r\|\alpha\|^2)(e_m - p_m) \geq 0. $$ It follows that $\alpha^* x \alpha \in C(p_m)$. We now show that $p$ is an Archimedean matrix order unit for $\{C(p_n)\}_n$. We verify the relevant properties for the case for $n=1$ and for $n >1$ the proofs are similar. Choose $r > 0$ such that $x + re \geq 0$. Let $\epsilon > 0$. Then $$ (x + rp) + \epsilon p + r(e - p) = x + re + \epsilon p \geq 0. $$ It follows that $x + rp \in C(p)$. Finally, assume $x + \delta p \in C(p)$ for all $\delta > 0$ and let $\epsilon > 0$. Then there exists $r > 0$ such that $$ (x + \epsilon/2 p) + \epsilon /2 p + r(e-p) \geq 0. $$ It follows that $x + \epsilon p + r(e-p) \geq 0$. So $x \in C(p)$. We now come to the main results of this section. Similarly to our notation in the Section 3, given an operator system $\mathcal{V}$, and a positive contraction $p \in \mathcal{V}$ we consider the matrix ordering $\{C(p_n)\}_n$ and we let $J_p = \Span C(p) \cap -C(p).$ We recall the following definition. Given an Archimedean order unit space $\mathcal{V}$ then the minimal order norm $\alpha_m$ on $\mathcal{V}$ is defined for $x \in \mathcal{V}$ by $$ \begin{aligned} \alpha_m(x) = \sup \{\left\lvert \varphi(x) \right\rvert: \varphi \in \mathcal{S}(\mathcal{V})\} \end{aligned} $$ where $\mathcal{S}(\mathcal{V})$ denotes the set of states on $\mathcal{V}$. It is not difficult to show that if $\alpha_o: \mathcal{V}_h \rightarrow [0,\infty)$ denotes the order norm induced by $e$ given by $$ \begin{aligned} \alpha_o(x) = \inf\{t>0: te \pm x \in \mathcal{V}^+\}, \end{aligned} $$ then $\alpha_o = \alpha_m$ when restricted to $\mathcal{V}_h.$ We refer the interested reader to [17,Section4] for the details. Proposition4.8. Let $\mathcal{V}$ an operator system and let $p \in \mathcal{V}$ be a nonzero positive contraction. Let $\alpha_m: \mathcal{V} \rightarrow [0,\infty)$ denote the minimal order norm induced by $e$. Then $\alpha_m(p) = 1$ if and only if $p \notin J_p.$ Proof. By the assumption that $p$ is a positive contraction we know that $\alpha_o(p) = \inf\{t>0: te - p \in \mathcal{V}^+\} \leq 1.$ The assumption that $\alpha_m(p) = 1$ implies $\alpha_o(p) = 1$. We first show that $p \notin -C(p)$. Suppose the contrary. Then by definition for all $\epsilon >0$ there exists $t >0$ such that $-p + \epsilon p + t(e-p) \in \mathcal{V}^+.$ In other words it must follow \begin{align*} p \leq \frac{t}{1 + t - \epsilon}e. \end{align*} If $\epsilon < 1$ then for all $t >0$ we have $\frac{t}{1+t-\epsilon} < 1$ which contradicts the assumption that $\alpha_o(p) = 1.$ Thus $p \notin -C(p).$ Now suppose that $p \in J_p$. Since $p^* = p$ we have $p \in C(p) \cap -C(p)$, a contradiction. As in Proposition 4.4 we will define the family of sets $\{\widetilde{C}(p_n)\}_n$ where for each $n \in \mathbb{N}$ we have $$ \begin{aligned} \widetilde{C}(p_n)=\{(x_{ij} + J_p) \in M_n(\mathcal{V}/ J_p): x=(x_{ij}) \in C(p_n)\}. \end{aligned} $$ We now have the abstract analogue to Theorem 3.6. Theorem4.9. Given an operator system $\mathcal{V}$ and positive contraction $p \in \mathcal{V}$ such that $\alpha_m(p) = 1$, the triple $$ ({\mathcal{V}/J_p}, \{\widetilde{C}(p_n)\}_{n \in \mathbb{N}}, {p + J_p}) $$ is a non-trivial operator system.
200603094/21
## Applications to quantum correlation sets We conclude with a brief application of our results to the theory of correlation sets in quantum information theory. We must first briefly recall some definitions. Let $n,k \in \mathbb{N}$ be positive integers. We call a tuple $\{p(a,b|x,y): a,b \in \{1,2,\dots,k\}, x,y \in \{1,2,\dots, n\}\}$ a correlation if $p(a,b|x,y) \geq 0$ for each $a,b \leq k$ and $x,y \leq n$ and if, for each $x,y \leq n$, we have $\sum_{a,b} p(a,b|x,y) = 1$. These conditions ensure that for each choice of $x$ and $y$, the matrix $\{p(a,b|x,y)\}_{a,b \leq k}$ constitutes a joint probability distribution. We say that a correlation $\{p(a,b|x,y)\}$ is non-signalling if for each $x \leq n$ and $a \leq k$ the quantity $$ p_A(a|x) := \sum_b p(a,b|x,y) $$ is well-defined (i.e. independent of the choice of $y$), and that similarly for each $y \leq n$ and $b \leq k$ the quantity $$ p_B(b|y) := \sum_a p(a,b|x,y) $$ is well-defined. We refer to the integer $n$ as the number of experiments and the integer $k$ as the number of outcomes. We let $C_{ns}(n,k)$ denote the set of non-signalling correlations with $n$ experiments and $k$ outcomes. Correlations model a scenario where two parties, typically named Alice and Bob, are performing probabilistic experiments. Suppose Alice and Bob each have $n$ experiments, and that each experiment has $k$ possible outcomes. Then the quantity $p(a,b|x,y)$ denotes the probability that Alice performs experiment $x$ and obtains outcome $a$ while Bob performs experiment $y$ and obtains outcome $b$. Whenever Alice and Bob perform the experiments independently without communicating to one another, the resulting correlation is non-signalling. It is well-known (and easy to see) that the set $C_{ns}(n,k)$ of non-signalling correlations is a convex polytope when regarded as a subset of $\mathbb{R}^{n^2k^2}$ in the obvious way. Let $H$ be a Hilbert space. We call a set $\{P_1, P_2, \dots, P_n\} \subset B(H)$ a projection-valued measure if each $P_i$ is a projection on $H$ and $\sum_i P_i = I$. A correlation $\{p(a,b|x,y)\} \in C_{ns}(n,k)$ is called a quantum commuting correlation if there exists a Hilbert space $H$, a unit vector $\phi \in H$, and projection valued measures $\{E_{x,a}\}_{a=1}^k, \{F_{y,b}\}_{b=1}^k \subset B(H)$ for each $x,y \leq n$ satisfying the conditions that $E_{x,a}F_{y,b} = F_{y,b}E_{x,a}$ for all $x,y \leq n$ and $a,b \leq k$ and $$ p(a,b|x,y) = \langle \phi | E_{x,a}F_{y,b}\phi \rangle. $$ The set of all quantum commuting correlations with $n$ experiments and $k$ outcomes is denoted by $C_{qc}(n,k)$. It is well-known that $C_{qc}(n,k)$ is a closed convex subset of $C_{ns}(n,k)$ and that it is not a polytope for any $n \geq 2$ or $k \geq 2$. If we modify the definition of the quantum commuting correlations by requiring the Hilbert space $H$ to be finite dimensional, we obtain a quantum correlation. The set of quantum correlations with $n$ experiments and $k$ outcomes are denoted by $C_q(n,k)$. The set $C_q(n,k)$ is known to be a convex subset of $C_{qc}(n,k)$. It was shown by William Slofstra in [19] that for some values of $n$ and $k$, $C_q(n,k)$ is non-closed, and hence $C_q(n,k)$ is a proper subset of $C_{qc}(n,k)$. The recent preprint [8] shows that for some ordered pair $(n,k)$ the closure of $C_q(n,k)$ is a proper subset of $C_{qc}(n,k)$. Precisely which ordered pairs $(n,k)$ satisfy this relation remains unknown. One reason questions about $C_q(n,k)$ and $C_{qc}(n,k)$ are difficult to answer is that these sets are defined by applying arbitrary vector states to arbitrary projection-valued measures acting on arbitrary Hilbert spaces. In principle, it may be easier to understand the sets $C_q(n,k)$ and $C_{qc}(n,k)$ if there were an equivalent definition which was independent of Hilbert spaces and Hilbert space operators. The following proposition indicates that such a characterization is, in some sense, possible. Proposition6.1. Let $n$ and $k$ be positive integers. Then the following statements are equivalent. $\{p(a,b|x,y)\} \in C_{qc}(n,k)$ (resp. - $\{p(a,b|x,y)\} \in C_{q}(n,k)$). There exists a (resp. finite dimensional) C*-algebra $\mathcal{A}$, projection valued measures $\{E_{x,a}\}_{a=1}^k, \{F_{y,b}\}_{b=1}^k \subset \mathcal{A}$ for each $x,y \leq n$ satisfying $E_{x,a}F_{y,b} = F_{y,b}E_{x,a}$ for all $x,y \leq n$ and $a,b \leq k$, and a state $\phi: \mathcal{A} \rightarrow \mathbb{C}$ such that $p(a,b|x,y) = \phi(E_{x,a}F_{y,b})$.
200603094/22
- 3. There exists an operator system $\mathcal{V} \subset B(H)$ (resp. for a finite dimensional Hilbert space $H$), projection valued measures $\{E_{x,a}\}_{a=1}^k, \{F_{y,b}\}_{b=1}^k$ for each $x,y \leq n$ satisfying $E_{x,a}F_{y,b} \in \mathcal{V}$ and $E_{x,a}F_{y,b} = F_{y,b}E_{x,a}$ for all $x,y \leq n$ and $a,b \leq k$, and a state $\phi: \mathcal{V} \rightarrow \mathbb{C}$ such that $p(a,b|x,y) = \phi(E_{x,a}F_{y,b})$. Proof. The equivalence of 1 and 3 are obvious, taking $\mathcal{V}$ to be the linear span of the operator products $E_{x,a}F_{y,b}$. The equivalence of 1 and 2 is an application of the GNS construction, taking $\mathcal{A}$ to be the C*-algebra generated by the set $\{E_{x,a}F_{y,b}\}$. In principle, statement 2 of Proposition 6.1 provides an abstract characterization of correlation sets in that it is independent of Hilbert space representation. However, C*-algebras are themselves complex structures, so it is not clear that statement 2 of Proposition 6.1 is a significant improvement over the definitions of $C_{qc}(n,k)$ and $C_q(n,k)$. Moreover, the correlation is generated by applying a state on the C*-algebra $\mathcal{A}$ to operators of the form $E_{x,a}F_{y,b}$, which span only a linear subspace of $\mathcal{A}$. Thus it seems that one can get by with significantly less data than the C*-algebra $\mathcal{A}$ has to offer. These observations make statement 3 of Proposition 6.1 seem more appealing, except that we have insisted that the operator system $\mathcal{V}$ be concretely represented so that we can enforce the relations $E_{x,a}F_{y,b} = F_{y,b}E_{x,a}$ and that each $E_{x,a}$ and $F_{y,b}$ are projections. Theorem 5.11 provides us with the tools to ensure that $E_{x,a}$ and $F_{y,b}$ are projections in an abstract operator system. It remains to show that the condition $E_{x,a}F_{y,b} = F_{y,b}E_{x,a}$ can also be enforced in an abstract operator system. Definition6.2. Let $n,k \in \mathbb{N}$. We call an operator system $\mathcal{V}$ a non-signalling operator system if it is the linear span of positive operators $\{Q(a,b|x,y) : a,b \leq k, x,y \leq n\} \subset \mathcal{V}$, called the generators of $\mathcal{V}$, with the properties that $\sum_{a,b} Q(a,b|x,y) = e$ for each choice of $x,y \leq n$ and that the operators $$ E(a|x) := \sum_b Q(a,b|x,y) $$ and $$ F(b|y) := \sum_a Q(a,b|x,y) $$ are well-defined (i.e. $E(a|x)$ is independent of the choice of $y$ and $F(b|y)$ is independent to the choice of $x$). We call an operator system $\mathcal{V}$ a quantum commuting operator system if it is a non-signalling operator system with the property that each generator $Q(a,b|x,y)$ is an abstract projection in $\mathcal{V}$. The next theorem justifies the choice of terminology in Definition 6.2. Theorem6.3. A correlation $\{p(a,b|x,y)\}$ is non-signalling (resp. quantum commuting) if and only if there exists a non-signalling (resp. quantum commuting) operator system $\mathcal{V}$ with generators $\{Q(a,b|x,y)\}$ and a state $\phi: \mathcal{V} \rightarrow \mathbb{C}$ such that $p(a,b|x,y) = \phi(Q(a,b|x,y))$ for each $a,b,x,y$. Proof. We first verify the equivalence for non-signalling correlations. Suppose that $\mathcal{V}$ is a non-signalling operator system with generators $Q(a,b|x,y)$. Let $\phi: \mathcal{V} \rightarrow \mathbb{C}$ be a state, and define $p(a,b|x,y) := \phi(Q(a,b|x,y))$. Then for each $x,y,a,b$ we have $\phi(Q(a,b|x,y) \geq 0$ since $\phi$ is positive, and for each $x$ and $y$ we have $$ \sum_{a,b} \phi(Q(a,b|x,y)) = \phi(\sum_{a,b} Q(a,b|x,y)) = \phi(e) = 1. $$ So $\{p(a,b|x,y)\}$ is a correlation. Similarly, the quantities $p_A(a|x)$ and $p_B(b|y)$ are well-defined for each $a,b,x,y$ with $p_A(a|x) = \phi(E(a|x))$ and $p_B(b|y) = \phi(F(b|y))$. So $\{p(a,b|x,y)\}$ is a non-signalling correlation. On the other hand, suppose that $\{p(a,b|x,y)\}$ is a non-signalling correlation. Let $H = \mathbb{C}$ regarded as a one- dimensional Hilbert space. Set $Q(a,b|x,y) = p(a,b|x,y)$ for each $a,b,x,y$. Clearly if $\mathcal{V} = \Span \{Q(a,b|x,y)\} = \mathbb{C} = B(H)$ then $\mathcal{V}$ is a non-signalling operator system with generators $\{Q(a,b|x,y)\}$, since $p(a,b|x,y)$ is a non-signalling correlation. Let $\phi: \mathcal{V} \rightarrow \mathbb{C}$ be the state $\phi(\lambda) = \lambda$. Then $\phi(Q(a,b|x,y)) = p(a,b|x,y)$ for each $a,b,x,y$. We now consider the equivalence for quantum commuting correlations. The forward direction is immediate from Proposition 6.1. We show the converse. Suppose that $\mathcal{V}$ is a quantum commuting operator system with generators $\{Q(a,b|x,y)\}$, and let $\phi: \mathcal{V} \rightarrow \mathbb{C}$ be a state. Since $\mathcal{V}$ is quantum commuting, each $Q(a,b|x,y)$ is an abstract projection.
200603094/1
# An abstract characterization for projections in operator systems Roy Araiza 1 and Travis Russell 2 We show that the set of projections in an operator system can be detected using only the abstract data of the operator system. Specifically, we show that if $p$ is a positive contraction in an operator system $\mathcal{V}$ which satisfies certain order- theoretic conditions, then there exists a complete order embedding of $\mathcal{V}$ into $B(H)$ mapping $p$ to a projection operator. Moreover, every abstract projection in an operator system $\mathcal{V}$ is an honest projection in the C*-envelope of $\mathcal{V}$. Using this characterization, we provide an abstract characterization for operator systems spanned by two commuting families of projection-valued measures. This yields a new characterization for quantum commuting correlations purely in terms of abstract operator systems. ## Introduction Beginning with the work of Choi-Effros in [3], an abstract characterization for self-adjoint unital subspaces of the bounded operators on a Hilbert space was given. More recently the abstract theory of operator systems progressed further with the development of the theory of tensors. In particular it was shown in [10, 11] that if classes of operator systems satisfied certain nuclearity properites then it must follow that $C^*(F_\infty)$ had Lance’s weak expectation property, i.e. particular nuclearity properties of operator systems were proven to be equivalent to Kirchberg’s conjecture. Kirch- berg showed in [12] that if the local lifting property for C*-algebras implied the weak expectation property then Connes’ embedding conjecture, originally appearing in [4], must hold. In [9], [5], and [14], an equivalence between Kirchberg’s conjecture and what is known as Tsirelson’s problem was established. Tsirelson’s problem asks if for all pairs of natu- ral numbers $(n,k)$ the equality $C_{qa}(n,k) = C_{qc}(n,k)$ holds, where $C_{qa}(n,k)$ denotes the closure of the set of quantum correlations, and $C_{qc}(n,k)$ denotes the set of quantum commuting correlations. Introducing new ideas coming from computer science, a recent preprint [8] demonstrates the existence of integers $n$ and $k$ such that $C_{qa}(n,k) \neq C_{qc}(n,k)$, simultaneously refuting the long-standing conjectures of Tsirelson, Kirchberg, and Connes. Sharp estimates on the ordered pairs $(n,k)$ for which $C_{qa}(n,k) \neq C_{qc}(n,k)$ are not known. Estimating these values could shed more light on the failure of the conjectures of Kirchberg and Connes, which would be useful in the study of operator algebras. The purpose of this paper is to better understand the role played by projections in operator systems, since projections play an outsized role in Tsirelson’s problem. Our principle motivation is the hope that new insights about the structure of operator systems may be useful in the study of Tsirelson’s problem. In this paper, we provide an abstract characterization for the set of projections in an operator system. Given an abstract operator system $\mathcal{V}$ and a positive element $p \in \mathcal{V}$ of unit norm (as induced by the order unit and positive cone on $\mathcal{V}$), we consider a collection of cones $\{C(p_n)\}_n$ induced by $p$ and prove that the quotient $*$ -vector space $\mathcal{V}/ J_p$ is an operator system with order unit $p + J_p$ and matrix ordering $\{C(p_n) + M_n(J_p)\}_n$, where $J_p = \Span C(p) \cap -C(p)$ (Theorem 4.9). When $\mathcal{V}$ is a concrete operator system in $B(H)$ and $p$ is a projection, we show that $\mathcal{V} / J_p$ is completely order isomorphic to the compression operator system $p \mathcal{V} p \subseteq B(pH)$ (Corollary 4.10). We call a positive element $p \in \mathcal{V}$ an abstract projection if $p$ has unit norm and the mapping $$ \begin{aligned} \pi_p: \mathcal{V} \rightarrow M_2(\mathcal{V})/J_{p \oplus q}, \quad x \mapsto \begin{pmatrix} x & x \\ x & x \end{pmatrix} + J_{p \oplus q} \end{aligned} $$
200603094/3
A matrix ordering $\{C_n\}_n$ is called proper if in additon $C_n \cap -C_n = \{0\}$ for each $n$. A element $e \in \mathcal{V}_h$ is called an Archimedean matrix order unit for a matrix ordering $\{C_n\}_n$ if given any $x \in M_n(\mathcal{V})_h$ there exists $r>0$ such that if $e_n:= I_n \otimes e,$ then $re_n - x \in M_n(\mathcal{V})^+$ and if $re_n + x \in M_n(\mathcal{V})^+$ for all $r>0$ then $x \in M_n(\mathcal{V})^+.$ When dealing with “ compression ” operator systems beginning in Sections 3 we will be dealing with cones that a priori are not necessarily proper. When it may be unclear from context, we will emphasize when a matrix ordering is not necessarily proper, and otherwise simply use the term matrix ordering. Our use of the term matrix ordering differs from some authors (e.g. [10]) where the term matrix ordering is used synonymously with proper matrix ordering. We note that the family $\{C_n\}_n$ with $C_n \subset M_n(\mathcal{V})_h$ is a matrix ordering if $C_n \oplus C_m \subset C_{m+n}$ and $\alpha^*C_n\alpha \subset C_m$ for all $\alpha \in M_{n,m}, n,m \in \mathbb{N}.$ When no confusion will arise we will simply denote an operator system by $\mathcal{V}$. The morphisms in use between matrix ordered $*$ -vector spaces will be completely positive maps, and the morphisms between operator systems will be the unital completely positive maps. Given a linear map $\varphi: \mathcal{V} \rightarrow \mathcal{W}$ between operator systems, then for each $n \in \mathbb{N}$ there is an induced linear map $I_n \otimes \varphi:= \varphi_n: M_n(\mathcal{V}) \rightarrow M_n(\mathcal{W})$ defined by $\sum_{i,j \leq n} \ket{i}\bra{j} \otimes x_{ij} \mapsto \sum_{i,j \leq n} \ket{i}\bra{j} \otimes \varphi(x_{ij}).$ The map $\varphi_n$ is called the nth amplification of $\varphi.$ Here we have let $\{\ket{i}\}_i$ denote column vectors in $\mathbb{C}^n$ with $1$ in the $i$ th position. We say that $\varphi$ is completely positive if $\varphi_n(M_n(\mathcal{V})^+) \subset M_n(\mathcal{W})^+$ for every $n$, and is a complete order isomorphism if $\varphi$ is invertible with $\varphi$ and $\varphi^{-1}$ both completely positive. A map $\varphi: \mathcal{V} \rightarrow \mathcal{W}$ which is not necessarily surjective is called a complete order embedding if it is completely positive, injective and $\varphi^{-1}$ is completely positive on $\varphi(\mathcal{V}) \subset \mathcal{W}$. We will identify two operator systems $\mathcal{V}$ and $\mathcal{W}$ if there exists a (unital) complete order isomorphism $\varphi$ between the two and we will denote this by $\mathcal{V} \simeq \mathcal{W}.$ A classical result due to Choi and Effros shows that there is a one-to-one correspondence between concrete and abstract operator systems. Theorem 2.2 ([3]). Given an abstract operator system $\mathcal{V}$ then there is a Hilbert space $H$ and a concrete operator system $\mathcal{W} \subset B(H)$ such that $\mathcal{V} \simeq \mathcal{W}$. Conversely, every concrete operator system is an abstract operator system. A well-known fact is that given a matrix-ordered $*$ -vector space, i.e., a pair $(\mathcal{V}, \{C_n\}_n)$ consisting of a $*$ -vector space $\mathcal{V}$ and a matrix-ordering $\{C_n\}_n$, then an element $e \in \mathcal{V}^+$ is an order unit for $\mathcal{V}$ if and only if it is a matrix order unit. We provide a brief proof below for completeness. Lemma2.3 Given an $*$ -vector space $\mathcal{V}$ then $M_n(\mathcal{V})_h = (M_n)_h \otimes \mathcal{V}_h$. Proposition2.4. Given a matrix ordered $*$ -vector space $\mathcal{V}$ then $e \in \mathcal{V}$ is an order unit if and only if it is a matrix order unit. Proof. Of course we need only show the forward direction. Let $x \in M_n(\mathcal{V})_h$ such that $x = \sum_{i \leq n} A_i \otimes x_i \in (M_n)_h \otimes \mathcal{V}_h$. For each $i$ write $A_i = P_i - Q_i$, where $P_i, Q_i \in M_n^+.$ Choose $\lambda > 0$ such that $\lambda e \pm x_i \in \mathcal{V}^+$ for each $i$. We then see \begin{align*} \lambda(\sum_{i \leq n} P_i + Q_i)\otimes e - x = \sum_{i \leq n} P_i \otimes (\lambda e -x_i) + \sum_{i \leq n} Q_i \otimes (\lambda e + x_i) \in M_n(\mathcal{V})^+. \end{align*} Simply choose $\widetilde{\lambda}$ such that $\widetilde{\lambda} I_n \geq \lambda (\sum_{i \leq n} P_i + Q_i)$ which proves the claim. Remark2.5 (TheCanonicalShuffle). Throughout the manuscript we will implement the “ canonical shuffle ” (see [15, Chapter8]). For example, given an operator system $\mathcal{V} \subset B(H)$, we use this in Lemma 3.9 to identify $M_n(\mathcal{V} \oplus \mathcal{V})$ with $M_n(\mathcal{V}) \oplus M_n(\mathcal{V}),$ where the direct sum is in the $\ell_\infty$ sense, i.e., we are realizing $\mathcal{V} \oplus \mathcal{V}$ as the diagonal $2 \times 2$ matrices with entries from $\mathcal{V}$. In particular, consider $A \in M_n(M_m(\mathcal{V}))$. We then write $$ \begin{aligned} A = \sum_{ij}\ket{i}\bra{j} \otimes A_{ij}, A_{ij} \in M_m(\mathcal{V}). \end{aligned} $$ It follows $$ \begin{aligned} A = \sum_{ij}\ket{i}\bra{j} \otimes A_{ij} = \sum_{ij} \ket{i}\bra{j} \otimes \sum_{kl} \ket{k}\bra{j} \otimes a_{ijkl}, a_{ijkl} \in \mathcal{V}. \end{aligned} $$
200603094/9
Let $a_{kl} = \sum_{ij} \ket{i}\bra{j} \otimes a_{ijkl}.$ We finally get \begin{align*} &\ket{1}\bra{1} \otimes \sum_{ij} \ket{i}\bra{j} \otimes pa_{ij11}p + \ket{1}\bra{2} \otimes \sum_{ij} \ket{i}\bra{j} \otimes pa_{ij12}q \ +&\ket{2}\bra{1} \otimes \sum_{ij} \ket{i}\bra{j} \otimes qa_{ij21}p + \ket{2}\bra{2} \otimes \sum_{ij} \ket{i}\bra{j} \otimes qa_{ij22}q \\ &=(p_n\oplus q_n)(\sum_{kl} \ket{k}\bra{l} \otimes a_{kl})(p_n\oplus q_n) \end{align*} which proves the result. Given an operator system $\mathcal{V} \subset B(H)$ with $p \in \mathcal{V}$ a projection. Then $\{C(p_n \oplus q_n)\}_n$ is a matrix ordering on $\mathcal{V} \oplus \mathcal{V}$. Proof. We begin with an observation. By Lemma 3.8 we know $C(p_n) \oplus C(q_n) \subset C(p_n \oplus q_n).$ If we consider $x \oplus y \in M_n(\mathcal{V}) \oplus M_n(\mathcal{V})$ then if $x\oplus y \in C(p_n \oplus q_n)$ we see \begin{align*} (p_n \oplus q_n)(x\oplus y)(p_n \oplus q_n) = \begin{pmatrix} p_nxp_n & 0 \\ 0 & q_nyq_n \end{pmatrix} \in B(H^{2n})^+. \end{align*} Compression by $\begin{pmatrix} 1 \\ 0 \end{pmatrix}$implies that$p_nxp_n \in B(H^n)^+$and therefore$x \in C(p_n).$Similarly, compression by$\begin{pmatrix} 0 \\ 1 \end{pmatrix}$implies that$q_nyq_n \in B(H^n)^+.$Thus$x\oplus y \in C(p_n) \oplus C(q_n).$Thus, we have$C(p_n) \oplus C(q_n) = C(p_n \oplus q_n)$when restricting to elements of the form$x \oplus y.$ $*$ -closed follows by definition. Fix $n \in \mathbb{N}$. It is immediate that given any $\lambda >0$ that $\lambda C(p_n \oplus q_n) \subset C(p_n \oplus q_n)$. Let $a,b \in C(p_n \oplus q_n)$. Since the compression by the projection $p_n\oplus q_n$ is linear, we have that $(p_n\oplus q_n)(a+b)(p_n\oplus q_n) = (p_n\oplus q_n)a(p_n\oplus q_n)+ (p_n\oplus q_n)b(p_n\oplus q_n) \in B(H^{2n})^+ + B(H^{2n})^+ \subset B(H^{2n})^+.$ We now check compatibilty. Let $m \in \mathbb{N}$ with $\alpha \in M_{2n,2m}$ and consider $\alpha^*(x\oplus y)\alpha \in M_2(M_m(\mathcal{V}))$, where $x\oplus y \in C(p_n \oplus q_n)$. Let $\varphi: M_2(\mathcal{V}) \rightarrow B(H^2)$ be defined as in Lemma 3.9. Then \begin{align*} &(p_m \oplus q_m)\alpha^*(x \oplus y)\alpha(p_m \oplus q_m) = \varphi_m(\alpha^*(x \oplus y)\alpha) = \alpha^*\varphi_n(x \oplus y) \alpha \\ &= \alpha^*(p_n \oplus q_n)(x \oplus y)(p_n \oplus q_n)\alpha \in \alpha^* B(H^{2n})^+\alpha \subset B(H^{2m})^+. \end{align*} which proves the result. We consider the family $\{\widetilde{C}(p_n \oplus q_n)\}_n$ where for each $n \in \mathbb{N}$, \begin{align*} \widetilde{C}(p_n\oplus q_n) = \{((x_{ij} \oplus y_{ij})+ J_{p \oplus q}) \in M_{n}((\mathcal{V} \oplus \mathcal{V})/ J_{p \oplus q}): (x\oplus y) \in C(p_n \oplus q_n), x = (x_{ij}), y = (y_{ij})\}. \end{align*} Here we have let $\mathcal{V} \oplus \mathcal{V}$ denote the $*$ -vector subspace of $M_2(\mathcal{V})$ consisting of all diagonal matrices over $\mathcal{V}.$ Given an operator system $\mathcal{V} \subset B(H)$ and a projection $p \in \mathcal{V}$. Let $q = I_H - q.$ Then $(\mathcal{V} \oplus \mathcal{V}) / J_{p \oplus q}$ is an operator system. Proof. By Lemma 3.10 it is immediate to show that $\{\widetilde{C}(p_n \oplus q_n)\}_n$ is a matrix ordering. Another direct consequence of the definition is that given any $((x_{ij} \oplus y_{ij}) + J_{p \oplus q})_{ij} \in C(p_n \oplus q_n) \cap -C(p_n \oplus q_n)$ then $x_{ij} \in C(p) \cap -C(p)$ and $y_{ij} \in C(q) \cap -C(q)$ for all $i,j$ which, along with Lemma 3.8, proves the cones are proper. The claim that $p \oplus q + J_{p \oplus q}$ is an Archimedean order unit follows from the results that $p$ is an Arhimedean order unit for $\{C(p_n)\}_n$, $q$ is an Archimedean order unit for $\{C(q_n)\}_n$ and then applying Lemma 3.8 and the observation in Lemma 3.10. We leave the details to the reader. More generally it will follow that $(p \oplus q)M_2(\mathcal{V})(p \oplus q)$ is an operator system. We begin with a lemma:
200603094/8
a $*$ -hermitian element $(x_{ij} + J_p) \in M_n(\mathcal{V}/ J_p).$ Let $r>0$ be such that $rp_n -x \in C(p_n)$ (see Proposition). It then follows \begin{align*} r(I_n \otimes (p + J_p)) - (x_{ij} + J_p) = (I_n \otimes (rp + J_p)) -(x_{ij} + J_p) = ((r(p_n)_{ij} - x_{ij}) + J_p) \in \widetilde{C}(p_n), \end{align*} since $rp_n - x \in C(p_n).$ Finally if for all $r>0$ we have $(r(p_n)_{ij} + x_{ij} + J_p) \in \widetilde{C}(p_n)$ then $rp_n + x \in C(p_n)$ which implies $x \in C(p_n)$ (see Proposition) and thus $(x_{ij} + J_p) \in \widetilde{C}(p_n)$. This finishes our proof. Definition3.7. Given an operator system $\mathcal{V} \subset B(H)$ with $p \in B(H)$ a projection, we call the set $p\mathcal{V} p$, regarded as linear operators on the Hilbert space $pH$, the concrete compression operator system. Thus, we have seen how to regard the compression of an operator system by a single projection as a quotient of the original operator system. We now wish to explore the compression of the operator system $M_2(\mathcal{V})$ relative to the cones $\{C(p_n \oplus q_n)\}_n$ where $p \in \mathcal{V}$ is a projection with $q = I-p,$ and $$ \begin{aligned} C(p_n \oplus q_n):= \{x \in M_{2n}(\mathcal{V}): x=x^*, (p_n \oplus q_n)x(p_n \oplus q_n) \in B(H^{2n})^+\}. \end{aligned} $$ We first prove the following quick lemma: Lemma3.8. Given an operator system $\mathcal{V} \subset B(H)$ and projections $p,q \in \mathcal{V}$ then $C(p_n) \oplus C(q_n) \subset C(p_n \oplus q_n)$ for all $n \in \mathbb{N}$. Proof. Given $x \in C(p_n)$ and $y \in C(q_n)$ then we see \begin{align*} (p_n\oplus q_n)(x \oplus y)(p_n \oplus q_n) = (p_nxp_n \oplus q_nyq_n) \end{align*} and for any $(\eta_1,\eta_2) \in H^n \oplus H^n$ we see \begin{align*} \langle (p_nxp_n \oplus q_nyq_n)(\eta_1,\eta_2) | (\eta_1,\eta_2) \rangle = \langle p_nxp_n\eta_1 | \eta_1 \rangle + \langle q_nyq_n\eta_2 | \eta_2 \rangle \in \mathbb{R}^{+} \end{align*} and thus $x\oplus y \in C(p_n \oplus q_n)$ which proves the result. We will use the “ Canonical shuffle ” (see Remark 2.5) as presented in the preliminary section. Lemma Given an operator system $\mathcal{V} \subset B(H)$, with projection $p \in \mathcal{V}, q = I -p$, consider the operator $\varphi: M_2(\mathcal{V}) \rightarrow B(H^2)$ defined by $\varphi(x) = (p\oplus q)x(p \oplus q).$ Then $I_n \otimes (p \oplus q) \rightsquigarrow (p_n \oplus q_n)$ under the $*$ -isomorphism $M_n(M_2(\mathcal{V})) \simeq M_2(M_n(\mathcal{V})).$ This is to say that the nth-amplification of $\varphi$ is given by $p_n \oplus q_n$ under the canonical shuffle. Proof. Fix $a \in M_n(M_2(\mathcal{V})).$ We then write $a = \sum_{ij} \ket{i}\bra{j} \otimes a_{ij}, a_{ij} \in M_2(\mathcal{V})$ and see \begin{align*} \varphi_n(\sum_{ij} \ket{i}\bra{j} \otimes a_{ij}) &= \sum_{ij} \ket{i}\bra{j} \otimes \varphi(a_{ij}) = \sum_{ij} \ket{i}\bra{j} \otimes (p \oplus q)(a_{ij})(p\oplus q) \\ &= \sum_{ij} \ket{i}\bra{j} \otimes [\ket{1}\bra{1} \otimes pa_{ij11}p + \ket{1}\bra{2} \otimes pa_{ij12}q + \ket{2}\bra{1} \otimes qa_{ij21}p + \ket{2}\bra{2} \otimes qa_{ij22}q], \end{align*} where $\{\ket{k}\bra{l}\}_{k,l}$ in the bracket above denote the matrix units in $M_2.$ It then follows \begin{align*} &\sum_{ij} \ket{i}\bra{j} \otimes [\ket{1}\bra{1} \otimes pa_{ij11}p + \ket{1}\bra{2} \otimes pa_{ij12}q + \ket{2}\bra{1} \otimes qa_{ij21}p + \ket{2}\bra{2} \otimes qa_{ij22}q] \\ =&\ket{1}\bra{1} \otimes \sum_{ij} \ket{i}\bra{j} \otimes pa_{ij11}p + \ket{1}\bra{2} \otimes \sum_{ij} \ket{i}\bra{j} \otimes pa_{ij12}q \\ +&\ket{2}\bra{1} \otimes \sum_{ij} \ket{i}\bra{j} \otimes qa_{ij21}p + \ket{2}\bra{2} \otimes \sum_{ij} \ket{i}\bra{j} \otimes qa_{ij22}q. \end{align*}
201201456/5
the lifetime is almost constant ($\sim 3.5-5$ Myr) and independent of metallicity (Hurleyetal.2000). On the other hand, the SMS evolves over Kelvin-Helmholtz (KH) timescale (Janka2002) $$ t_{\mathrm{KH}}=6.34 \times 10^8 \left(\frac{M_\ast}{\mathrm{\,\mathrm{M_{\odot}}}} \right)^{-1} \,\text{yr}. $$ We stop our simulation at 5 Myr as we are mostly interested in the early phases of cluster evolution corresponding to the most rapid rates of black hole seeds mass growth, in order to quantify the attainable final black hole masses for each of our simulation models. We do not take into account the change in stellar radius due to stellar evolution of MS stars despite the fact that stellar evolution could play an important role determining the final mass of the SMS (Glebbeeketal.2013; Katz 2019). We also ignore the rotation of the stars which could significantly influence the evolution of the massive stars and hence the final mass of the SMS (Maeder & Meynet 2000; Leigh et al. 2016). Our main goal here is to build a simple model and to better understand the complicated physics involved in the study. ## 2.3 Gas accretion Gas accretion plays a crucial role in our simulations as the stellar masses and radii can increase significantly depending on the accretion model, and hence the number of collisions will be increased. In our model the gas that is being accreted by the stars is assumed to be in a stationary state. We impose momentum conservation during the accretion process, which means the stars will slow down as they gain mass, and fall into the potential well of the cluster. The mass accreted onto the stars is removed from the (static) gas, leading to its depletion and eventually to a transition where the cluster evolution is determined only by N-body dynamics. However, gas motions do not follow the stellar motions and as a result sometimes accretion could add or remove momentum, less or more, respectively. In order to model the gas dynamics we need hydrodynamic simulations. We have considered different accretion scenarios in our work. As a first simplified choice we study the effect of a constant accretion rate three different accretion rates of $10^{-4},\,10^{-5}$ and $10^{-6} \,\mathrm{M_{\odot}}\mathrm{yr^{-1}}$, which roughly correspond to the Eddington and Bondi accretion rates under different assumptions for the mass of the central object. Even though it is highly unlikely that all stars in the cluster will accrete at the same constant rate, and individual star accretion might even be episodic (Vorobyov&Basu2006, 2015), this ad-hoc approximation is a good starting point to understand how the interplay of accretion and collisions works in the cluster. Next we consider the Eddington accretion rate given as $$ \dot{M}_{\mathrm{Edd}}=\frac{4\pi G M_\ast}{\epsilon\kappa c_{\mathrm{s}}}, $$ where $c_{\mathrm{s}}$ is the speed of sound in the gas, $\epsilon$ is the radiative efficiency, i.e. the fraction of the rest mass energy of the gas that is radiated and $\kappa$ is the electron scattering opacity. Using $\kappa=0.4\,\mathrm{cm^2\,g^{-1}}$ and $\epsilon=0.1$, Eq. 10 can be written as $$ \dot{M}_{\mathrm{Edd}}=2.20\times 10^{-8}\left(\frac{M_\ast}{\,\mathrm{M_{\odot}}} \right)\,\,\mathrm{\,\mathrm{M_{\odot}} yr^{-1}}. $$ Finally we consider the Bondi-Hoyle-Lyttleton accretion. Simu- lations (Bonnell et al. 2001) have shown that in larger clusters the stars accrete unequally, with the stars near the core accreting more than those near the outer envelopes of the cluster. This is due to the fact that the gas mass is accumulating near the core due to the cluster potential where it can be accreted by the stars. Even initially uniform clusters show a position dependent accretion as the gas and stars redistribute themselves in the host cluster potential. As a result of the position-dependent accretion, the final configurations show a significant amount of mass segregation. In principle the accretion rate of a star will depend on its cross section $\pi R_{\mathrm{acc}}^2$, where $R_{\mathrm{acc}}$ is the accretion radius, on the gas density $\rho_\infty$, and the relative velocity of the star with respect to the gas $v_\infty$, as $$ \dot{M}=\pi v_\infty \rho_\infty R_{\mathrm{acc}}^2. $$ In the original Hoyle–Lyttleton (HL) problem the accretion ra- dius in the supersonic regime is given as (Hoyle&Lyttleton1939, $$ R_{\mathrm{HL}}=\frac{2GM_\ast}{v_\infty^2}, $$ which leads to the HL accretion rate: $$ \dot{M}_{\mathrm{HL}}=\frac{4\pi G^2M_\ast^2\rho_\infty}{v_\infty^3}. $$ However, Bondi (1952) defined the Bondi radius as $$ R_{\mathrm{B}}=\frac{GM_\ast}{c_{\mathrm{s}}^2}. $$ The flow outside this radius is subsonic and the density is almost uniform, while inside the Bondi radius the gas becomes supersonic. This led Bondi (1952) to propose an interpolated Bondi-Hoyle (BH) formula: $$ \dot{M}_{\mathrm{BH}}=\frac{2\pi G^2M_\ast^2\rho_\infty}{(v_\infty^2+c_{\mathrm{s}}^2)^{3/2}}. $$ Studies have shown that there will be an extra factor of 2 which leads to the final accretion formula given as $$ \dot{M}_{\mathrm{BH}}=\frac{4\pi G^2M_\ast^2\rho_\infty}{(v_\infty^2+c_{\mathrm{s}}^2)^{3/2}}. $$ This accretion rate can be written as in Eq. 2 of Maccarone&Zurek (2012): $$ \dot{M}_{\mathrm{BH}} = 7\times 10^{-9}\ \left(\frac{M_\ast}{\,\mathrm{M_{\odot}}}\right)^2\left(\frac{n}{10^{6}\, \mathrm{cm}^{-3}} \right)^2\left(\frac{\sqrt{c_{\mathrm{s}}^2+v_\infty^2}}{10^6\,\mathrm{cm\ s}^{-1}} \right)^{-3}\mathrm{\,\mathrm{M_{\odot}} yr^{-1}}. $$ A recent study by Kaaz et al. (2019) has shown that the BH ac- cretion rate will depend on the characteristic accretion radius of the cluster $R_{\mathrm{acc}}$ and the mean separation between stars $R_{\bot}$, where $R_{\bot}=R_{\mathrm{cl}} N^{-1/3}$. The average accretion rate of an individual star is given as $$ \langle \dot{M}_{\mathrm{BH}} \rangle=\left\{ \begin{array}{@{}ll@{}} \dot{M}_{\mathrm{BH}}, & \text{when}\ R_{\bot} \gg R_{\mathrm{acc}}, \\ N\times\dot{M}_{\mathrm{BH}}, & \text{when}\ R_{\bot} \leq R_{\mathrm{acc}} \end{array}\right. $$ In our work we consider a BH accretion rate given by Eq. 19, where $\dot{M}_{\mathrm{BH}}$ is given by Eq. 18. ## 2.4 Handling collisions We adopt the sticky-sphere approximation to model collisions be- tween the main sequence stars (see, for example, Leigh & Geller (2012); Leigh et al. (2017). If the distance between the centers of two stars is less than the sum of their radii, we assume that the stars
201201456/7
of collision products, reaching $10^4 \,\mathrm{M_{\odot}}$ already after $0.8$ Myr, and about $3\times10^4 \,\mathrm{M_{\odot}}$ after 5 Myr. The final mass of the MMO is thus very similar to what is found in the case of an accretion rate of $\dot{m}=10^{-5}\,\mathrm{M_{\odot}}\,\mathrm{yr}^{-1}$, however the evolution is considerably accelerated, thus reaching the final mass earlier. The result indicates that the mass of the MMO does not become much larger than about $10\%$ of the initial cluster mass. However, this not a fundamental limit. If all the gas gets accreted (which was seen in some of our simulations), the MMO can gain more mass via accretion and then collisions will only increase it more. We verified that for all our models the free-fall time of the gas is shorter than the gas depletion time, i.e., all the gas will be replenished and stars will keep accreting. 2 Dynamically high accretion rates will push clusters into core collapse on very short timescales compared to the rate from two-body relaxation. ## 3.2 Dependence on the physical recipe for the accretion rate As a next step, we explore the dependence of the evolution in the NSCs for different physical assumptions regarding the accretion rate, i.e. going beyond the simplified assumption of a constant accretion rate. In Fig. 6, we show the expected time evolution for the case of Eddington accretion scenario given by Eq. 11. In this case, we find that the evolution is comparable to what we found for a constant accretion rate of $\dot{m}=10^{-6}\,\mathrm{M_{\odot}}\,\mathrm{yr}^{-1}$. The gas mass decreases by about $15\%$ over 5 Myr, while the stellar mass increases by the same amount. The evolution of the Lagrangian radii is rather stable, with the $10\%$ and $50\%$ radii slightly contracting, and the $90\%$ radius moderately expanding. The first collision occurs after about $1.5$ Myr, and about $15$ collisions are reached after $5$ Myr. The growth of the MMO occurs most rapidly after about $3.5$ Myr, and reaches about $900 \,\mathrm{M_{\odot}}$. The results show that relevant numbers of collisions could occur, which can lead to a SMS mass of $\sim 10^3\,\mathrm{M_{\odot}}$ in the Eddington case. As a more optimistic scenario, we also consider the case of Bondi accretion given by Eq. 18. The results are shown in Fig. 7. As the Bondi accretion rate $\dot{M}_{\mathrm{BH}}\propto M_\ast^2$, this recipe has a strong impact on the evolution of the MMO. Initially the evolution takes place more slowly, due to an initially lower accretion rate, but is rapidly enhanced at late times as the mass of the accretor grows. As seen in the evolution of the gas and stellar mass, for a long time the accretion rates are lower than in the Eddington scenario, and only after about $4.7$ Myr the evolution becomes very rapid, and accelerates so much that the gas becomes fully depleted within a short time, while the stellar mass increases in the same way. The first collision occurs after about $1.5$ Myr and in total about 20 collisions occur before the accretion becomes strongly accelerated, with the MMO reaching about $200 \,\mathrm{M_{\odot}}$ within the first 4 Myr. During the accelerated phase of the evolution, more than 120 additional collisions occur, and the final mass reaches about $10^5 \,\mathrm{M_{\odot}}$. While the Lagrangian radii are very stable for the first $4.7$ Myr, they react to the extreme evolution occurring subsequently, with the $10\%$ Lagrangian radius strongly decreasing due to the formation of the MMO, and the same slightly later even for the $50\%$ radius. The $90\%$ radius shows first a minor decrease at the time when the evolution accelerates and subsequently expands. The evolution in this scenario is sufficiently extreme that our model assumptions will break down early on due to a lack of gas replenishment, so this part of our results needs to be regarded with caution. The evolution of both the average accretion rate in the cluster and the maximum accretion rate is shown for the models with Eddington and Bondi accretion in Fig. 8 (in the case of constant accretion rates, the plots would be trivial). The average accretion rates are almost constant as a function of time and increase only at late times. In the Eddington scenario, this increase is hardly visible by eye, while it is much more pronounced in the case of Bondi accretion, due to the steep increase of the accretion rate of the MMO, which then affects also the calculation of the average. We indeed find that the Bondi accretion rate for the MMO increases by more than two orders of magnitude, while it is about an order of magnitude in the case of Eddington accretion. ## 3.3 Collision vs accretion time scale To determine the respective relevance of collisions vs accretion for a single object, we aim in the following at a systematic comparison of the collision timescale ($t_{\mathrm{coll}}$) and accretion timescale ($t_{\mathrm{acc}}$). To evaluate collision timescale for the MMO, we adopt the formulation of Leighetal. (2017) for a system consisting of heavy particles with mass $m_{\mathrm{A}}$ and light particles with mass $m_{\mathrm{B}}$: $$ t_{\mathrm{coll}}=\frac{G^3 m_{\mathrm{A}}^{1/2}\bar{m}M_{\mathrm{cl}}^{13/2}}{12\sqrt{2}m_{\mathrm{B}}^{3/2}N_{\mathrm{B}}(R_{\mathrm{A}}+R_{\mathrm{B}})^2|E|^{7/2}}, $$ where $\bar{m}$ is the average mass of all particles, $R_{\mathrm{A}}$ and $R_{\mathrm{B}}$ the radii of particles of type A and B, $N_{\mathrm{B}}$ is the number of particles of type B, and $|E|$ the total energy of the stars in the cluster, which we evaluate using the virial theorem and taking into account the gravitational potential from gas and stars via $$ |E|=\frac{GM_{\mathrm{tot}}M_{\mathrm{cl}}}{2R_{\mathrm{cl}}}, $$ with $M_{\mathrm{tot}}=M_{\mathrm{cl}}+M_{\mathrm{g}}$ being the total mass. It is important to note that eq. 21 does not incorporate the effect of accretion on the collisions probability, while our numerical experiments in section 3.1 establish that accretion enhances collisions. The collision timescale employed here for collisions with the MMO can in this sense be regarded as an upper limit. As demonstrated by Barreraetal. (2020), in the absence of accretion, the prescription works best for lower particle numbers and less extreme mass ratios. Once one object becomes more massive than the rest, quasi-Keplerian motions and a loss cone formalism start to become more relevant. We do not account for that here and neglect these effects in the context of the collision timescale. The accretion timescale of the MMO is given by $$ t_{\mathrm{acc}}=\frac{M_{\mathrm{max}}}{\dot{m}}, $$ with $M_{\mathrm{max}}$ being the mass of the MMO and $\dot{m}$ is given by constant, Eddington and Bondi accretion rates. We plot the theoretically expected ratio of these timescales in Fig. 9, considering the cases of a constant accretion rate of $10^{-5}\,\mathrm{M_{\odot}}\mathrm{yr}^{-1}$, as well as Eddington accretion and Bondi accretion, as a function of the mass of the MMO. The color bar represents different ratios of gas to cluster mass. We consider $m_{\mathrm{A}}=20\,\mathrm{M_{\odot}}$, $\bar{m}=22\,\mathrm{M_{\odot}}$ and $N_{\mathrm{B}}=5000$ (motivated from the initial conditions of our simulations). The radii $R_{\mathrm{A}}$ and $R_{\mathrm{B}}$ are calculated using eq. 6 and 7. In the considered parameter space covering several orders of mag- nitude, the ratio $\frac{t_{\mathrm{coll}}}{t_{\mathrm{acc}}}\gg 1$, suggesting that accretion dominates over collisions in this regime. For $t_{\mathrm{coll}}\approx t_{\mathrm{acc}}$, we would require extremely 2 If the free-fall time is longer than the gas depletion time, then there will be a rapid initial phase of accretion, followed by very little accretion while the gas slowly sinks down the cluster potential, eventually increasing again once the gas density gets high enough.
201201456/12
Figure7. Same as in Figure 3, but assuming instead Bondi accretion. the cluster will rather accelerate the accretion process due to the steep dependence on mass, thereby leading to an even more rapid growth via accretion. For the Eddington scenario and in case of a constant accretion rate, the shortening of collision timescale in comparison is more relevant, and in these cases there is a clear correlation between having a more compact cluster and a more massive central object. ## 3.6 Effect of mass loss during collisions Finally, the possible effect of mass loss during collisions is explored for different scenarios and the results are shown in Fig. 13, consid- ering different accretion scenarios, all for our reference IMF with masses between $10-100\,\mathrm{M_{\odot}}$. The mass loss during collisions is particularly relevant in the case of a constant accretion rate, where we consider the reference case of $\dot{m}=10^{-5}\,\mathrm{M_{\odot}}\mathrm{yr}^{-1}$. Both in the cases without mass loss as well as in K15, we find almost the same final mass of the MMO of about $3\times10^4\,\mathrm{M_{\odot}}$. This shows that most of the collisions occur at very high mass ratios, where K15 suggests the mass loss to have little impact. However, if this assumption is not correct or rather an under- estimate, our calculation with a mass loss fraction of $3\%$ shows that the final mass could be considerably reduced by roughly a factor of 10, consistent with similar results by AlisterSegueletal. (2020) for primordial protoclusters. The reason is that in this case, particularly for high mass ratios during the collisions, the mass loss can become comparable to the mass that is added during the collision, or poten- tially even larger, and then compensate for part of the accretion. The final mass is similar but a bit further reduced in case of a $5\%$ mass loss fraction, reflecting the flattening that was also found by Alister Segueletal. (2020) when plotting final masses as a function of the mass loss fraction. In the case of Eddington accretion, we first have to note that here the results depend somewhat sensitively on when the first collision occurs, as subsequently the accretion rate will be enhanced, thereby providing a somewhat statistical element that depends on the nonlin- ear evolution. For this reason, the simulation with the K15 obtains the highest mass of a few times $10^3\,\mathrm{M_{\odot}}$, a very similar value as the simulation with no mass loss. In case of a $3\%$ mass loss fraction,
201201456/14
Figure9. The ratio of collision over accretion timescales, $t_{\mathrm{coll}}/t_{\mathrm{acc}}$, as a function of stellar mass for the case of a constant accretion rate of $10^{-5}\,\mathrm{M_{\odot}}\mathrm{yr}^{-1}$ (top panel), Eddington accretion rate (mid panel) and Bondi accretion rate (bottom panel). We show the results for clusters sizes of $0.3$ pc (left), $1$ pc (middle) and $3$ pc (right), assuming different ratios of gas to cluster mass. The expected ratio $t_{\mathrm{coll}}/t_{\mathrm{acc}}\gg 1$, suggesting that accretion dominates over collisions in this regime. that the collision timescale is much greater when calculated for the MMO. However, we emphasize that the collision rates we employed for this estimate did not account for the potentially enhanced col- lision probability in the presence of accretion, and our simulations show that the number of collisions increases for larger accretion rates. It further must be noted that when evaluating the timescale to have collisions between any two objects within the cluster, the colli- sion timescale can become comparable to or shorter than accretion timescale. When considering the effects of an IMF, the “ Spitzer In- stability ”(Spitzer1969) will lead to a shorter relaxation time in the core and an increased collisional cross section, both of which will increase the collision rate. We therefore expect collisions to play an important relevant role for the growth of the most massive object, consistent with what we find in the simulations as well. As the gas is depleted, collisions start to take over as the dominant mass growth mechanism for the MMO. This is only accelerated by the effects of cluster contraction induced by the presence of the gas reservoir and the dissipational effects it introduces to the distribution of relative particle velocities and masses (Leighetal.2014). We explored the dependence of our results on the details of the IMF, considering Salpeter-type IMFs with different upper-mass lim- its. For the parameter space we explored, when considering constant accretion rates, the results were insensitive to such a change, while in the case of Eddington accretion, a larger upper-mass limit favours larger final masses. In the case of Bondi-accretion, the time of occur- rence of the first collision is more important than an initially larger mass, particularly when the mass change due to a collision is larger than the mass change due to a variation of the upper mass cutoff of the IMF. We further explored the dependence on the central density of the cluster by varying its radius and keeping the total cluster mass constant. For models with constant accretion rates, a more compact cluster significantly favors the formation of a supermassive star, with the final mass varying by more than an order of magnitude, for variations of cluster radii from $0.3$ pc to 5 pc, as found by Giersz
201201456/2
teri2017; Toyouchietal.2019; Takeoetal.2020). Unlike the previous scenario, a temporary non-availability of gas is less of a problem in this scenario, as it can be compensated via highly efficient accretion at other times. However, ensuring a sufficiently high streaming ef- ficiency of the gas from the large-scale reservoir down to horizon scales is difficult to achieve because accretion flows onto seed BHs might be reduced by their own radiation (Milosavljevićetal.2009; Sugimuraetal.2018; Reganetal.2019) and due to the inefficient gas angular momentum transfer (Inayoshietal.2018a; Sugimuraetal. 2018). Another alternate and more optimistic scenario is the direct col- lapse of black holes (DCBH) (Oh&Haiman2002; Bromm&Loeb &Omukai2020; Luoetal.2020). The seed mass produced by this mechanism is $\sim 10^{4-5}\,\,\mathrm{M_{\odot}}$, sufficiently high for the seed BH to grow to a $10^9\,\,\mathrm{M_{\odot}}$ black hole until $z\sim 7$. The basic idea is that a gas cloud in a massive halo with virial temperature $T \gtrsim 8000$ K and no ${\rm H_2}$ molecules or metals present inside the gas cloud can collapse with- out fragmentation directly into a SMS which can collapse into seed black holes of similar mass after their lifetime (Umedaetal.2016). To prevent the formation of ${\rm H_2}$ molecules, a very high background UV radiation flux is required (Latifetal.2015; Wolcott-Greenetal. 2017). Another key requirement for this scenario are large inflow rates of $\sim 0.1\,\mathrm{M_{\odot}}\text{yr}^{-1}$ which can be obtained easily in metal free halos (Agarwaletal.2012; Latifetal.2013; Shlosmanetal.2016; waletal.2019; Latifetal.2020). As a result of the high accretion rate, the SMS becomes inflated and the effective temperature of the SMS drops to several 1000 K (Hosokawaetal.2012, 2013; Schleicher etal.2013; Woodsetal.2017; Haemmerléetal.2018). Therefore, the radiative feedback becomes inefficient and the rapid accretion flow continues, allowing the SMS to reach the mass of $\sim 10^{4-5}\,\mathrm{M_{\odot}}$. Re- cent numerical studies by Latifetal. (2020) and Reganetal. (2020a) have confirmed such large inflow rates lasting for millions of years required to form the DCBHs. However, the presence of even tiny amounts of metals can trigger strong fragmentation through metal line cooling (Omukai et al. 2008; Dopcke et al. 2011; Latif et al. 2016; Corbett Moran et al. 2018; Chon & Omukai 2020) and re- duce the infall rate. Even in metal enriched halos ($Z < 10^{-3}\,Z_\odot$) where fragmentation takes place, the central massive stars could be fed by the accreting gas and grow supermassive (Chon & Omukai 2020). Above this threshold value, the gas fragments into lower mass stars and a single object could not be formed. Alternatively, the re- quirement of a high infall rate could be met by dynamical heating during rapid mass growth of low-mass halos in over-dense regions at high redshifts (Wise et al. 2019), or by massive nuclear inflows in gas-rich galaxy mergers (Mayeretal.2015). Reganetal. (2020b) have shown that atomic cooling haloes with higher metal enrichment ($Z > 10^{-3}\,Z_\odot$) can also be possible candidates for the formation of SMSs in the early universe if the metal distribution is inhomoge- neous. As another caveat, we note that even tiny amounts of dust can initiate fragmentation through enhanced cooling. While the presence of such a strong background radiation is rare, it may be possible to find such conditions (Visbal et al. 2014; Dijkstra et al. 2014; Re- gan et al. 2017). Wise et al. (2019) and Regan et al. (2020c) have recently studied atomic cooling haloes which satisfy the above re- quirements. Wiseetal. (2019) have shown that the requirement of the high UV flux can be significantly reduced in halos with rapid merger histories (Inayoshi et al. 2018b) and in halos located in regions of high baryonic streaming velocities (Naozetal.2013; Hiranoetal. 2017). In realistic scenarios, there will always be at least some frag- mentation happening even if $\mathrm{H}_2$ cooling is suppressed. Using high resolution numerical simulations Latifetal. (2013) have found that fragmentation occurs even in the atomic-cooling regime in 7 out of 9 simulations, unless a subgrid model is added which can provide ad- ditional viscosity in the simulation. The final mass of the collapsed object may further be limited by X-ray feedback (Aykutalp et al. 2014). Thus, it motivates to explore other mechanisms of formation of SMBH seed at high redshift. A third possibility is formation of massive black hole seeds in dense stellar clusters either via runaway collisions of stars, leading to the formation of a supermassive star (SMS) (PortegiesZwart& etal.2020) or via the mergers of stellar-mass black holes with other black holes or stars (Gierszetal.2015; Hasteretal.2016; Rizzuto etal.2021). In dense high redshift dense stellar clusters, black hole seeds of masses $\sim 10^{2-4}\,\mathrm{M_{\odot}}$ could be formed (Greeneetal.2020) which could grow either via tidal capture and disruption events (Stone et al. 2017; Alexander & Bar-Or 2017; Sakurai et al. 2019) or via enhanced accretion and collisions (Vesperinietal.2010; Leighetal. 2013a; Boekholtetal.2018). A recent study by Tagawaetal. (2020) has shown that seeds of masses $\sim 10^{5-6}\,\mathrm{M_{\odot}}$ can also be produced via frequent stellar mergers. Even though the SMSs are one of the most promising progenitors of the observed high redshift quasars, this hypothesis still requires observational verification. Upcoming next generation telescopes e.g. James Webb Space Telescope (JWST), Euclid and WFIRST will be able to detect SMSs at $z\sim 6-20$ et al. 2020; Martins et al. 2020). The number of SMSs per unit redshift per unit solid angle, which are expected to be detected at a redshift $z$, is given by: $$ \frac{dn}{dzd\Omega}=\dot{n}_{\mathrm{SMS}}t_{\mathrm{SMS}}r^2\frac{dr}{dz}, $$ where $\dot{n}_{\mathrm{SMS}}$ is the SMS formation rate per unit comoving volume, $t_{\mathrm{SMS}}$ is the average lifetime of a SMS as given by eq. 9, and $r(z)$ is the comoving distance to redshift $z$ given by: $$ r(z)=\frac{c}{H_0}\int_0^z \frac{dz^\prime}{\sqrt{\Omega_m(1+z^\prime)^3+\Omega_\Lambda}}. $$ However, the actual number of SMSs that will be detected will depend on the sensitivity of the instruments (see discussion below). At present the $\dot{n}_{\mathrm{SMS}}$ is very poorly constrained. SMSs could be observed both to be cool and red or hot and blue (Suraceetal.2018, 2019). Whether an observed SMS will be red or blue depends on how quickly and persistently the SMS is growing, whether it is rapidly accreting gas from an atomic-cooling halo or growing from runaway collisions in a stellar cluster. For rapidly- accreting SMSs, if the star is growing with a rate of $10^{-3}\,\mathrm{M_{\odot}}\mathrm{yr}^{-1}$ (Haemmerléetal.2018) or less, or if accretion is halted for longer than the thermal timescale of the envelope (Sakurai et al. 2015), then the SMS will contract to the main sequence and becomes a hyper-luminous Pop III star and will be observed as blue, implying a spectral temperature of $10^4$ K or higher. For a higher accretion rate $\gtrsim 10^{-3}\,\mathrm{M_{\odot}}\mathrm{yr}^{-3}$ the SMS will be red and bloated (Omukai& Palla 2003; Hosokawa et al. 2012, 2013; Haemmerlé et al. 2018), and the atmospheric temperature will be less than $10^4$ K, implying a negligible amount of UV radiation. Haemmerléetal. (2018) have
201201456/11
Figure6. Same as in Figure 3, but assuming instead Eddington accretion. the increase in mass. The mass of the MMO correlates well with the compactness of the cluster in the cases considered here. In the Eddington accretion scenario where at least initially the accretion rate tends to be lower, the evolution depends more strongly on the size of the cluster, and for $R_{\mathrm{cl}}=2$ pc or more, no significant growth of the mass of the MMO is found. In the case of $R_{\mathrm{cl}}=1$ pc, the central mass starts growing after about 3.5 Myr, reaching about $800 \,\mathrm{M_{\odot}}$. In the case of more compact clusters, the evolution starts earlier, even within the first Myr in case of a $0.3$ pc cluster, with the MMO reaching a final mass of about $2\times10^4 \,\mathrm{M_{\odot}}$. The growth overall is less steep in this scenario, as the Eddington accretion rate is initially lower, thus leading to a more gradual increase in total stellar mass and the contraction of the cluster occurs more gradually. Again, the mass of the MMO correlates well with the compactness of the cluster for these configurations. In the case of Bondi accretion, the behavior becomes somewhat more extreme. It is not directly related to the initial $R_{\mathrm{cl}}$, but depends more on when the first collision happens to occur, thereby producing an object of larger mass, to which the Bondi rate reacts very sensi- tively. The occurrence of this first collision will statistically vary for different cluster sizes, and in fact Fig. 12 shows no clear dependence of the overall evolution on the $R_{\mathrm{cl}}$. In all cases, the evolution initially occurs very slowly, though at different times for the different simu- lations, a sudden acceleration occurs once the Bondi rate becomes sufficiently large. The final mass is always about $10^5 \,\mathrm{M_{\odot}}$, forming in between 2 and 5 Myr for all cases considered here. In this case, the evolution is mostly driven by accretion, with collisions not playing a relevant role. In the cases considered here, the cluster size has the lowest impact in the case of the Bondi accretion rate, due to the very steep depen- dence on the mass of the MMO in this scenario. For more moderate mass dependencies, the cluster size is however found to be highly rel- evant, and can strongly affect the resulting evolution. Understanding the physical mechanism of accretion along with its dependence of mass will thus be crucial to determine the relevance of this parameter. While a more compact cluster in principle shortens the timescale for collisions, we also note that the latter is not necessarily the dominant effect. Particularly in the case of Bondi accretion, early collisions in
201201456/6
Figure2. Evolution of the Lagrange Radii for a test run without accretion with $N=5000$ particles, $M_{\mathrm{cl}}=M_{\mathrm{g}}=1.12\times 10^5 \,\mathrm{M_{\odot}}$, $R_{\mathrm{cl}}=R_{\mathrm{g}}=1$ pc, assuming a Salpeter IMF with mass range $10-100 \,\mathrm{M_{\odot}}$. have merged and we replace them with a single object whose mass is equal to the sum of the masses of the colliding stars. We consider the collision product to be a MS star and the radius of the object to be determined by the $M_\ast-R_\ast$ relation described in Eqs. 6 and 7. We assume that linear momentum is conserved during the collision. However, studies have shown that the mass is not necessarily con- served during the collision of stars (Sillsetal.2002; Dale&Davies 2006; Tracetal.2007). AlisterSegueletal. (2020) have shown that the final mass of the colliding objects could change a lot depending on the mass loss rate. Glebbeek et al. (2013) have shown that the mass loss rate depends on the type of stars colliding. In our study, all the stars are MS stars and we have assumed both a mass loss recipe as given by Eq. 3 of Katzetal. (2015), $$ \Delta M=\text{min}\left[0.062\frac{M_2}{0.7M_1},0.062\right](M_1+M_2), $$ where $M_1$ and $M_2$ are the masses of the colliding stars, as well as constant mass loss rates of $3\%$ and $5\%$. Henceforth, we refer to Eq. 20 as K15. ## 2.5 Stability of numerical setup While our work is based on the setup previously tested and employed by Boekholt et al. (2018) and Alister Seguel et al. (2020), we test the stability pursuing a reference run where both the accretion and collisions are deactivated, to show that our setup corresponds to an overall stable initial condition. For this setup, we employ $N=5000$ particles, $M_{\mathrm{cl}}=M_{\mathrm{g}}=1.12\times 10^5 \,\mathrm{M_{\odot}}$, $R_{\mathrm{cl}}=R_{\mathrm{g}}=1$ pc, and we assume a Salpeter IMF with masses between $10-100 \,\mathrm{M_{\odot}}$. As accretion and collisions are switched off, we find the gas and stellar mass to be consistent. The time evolution of the Lagrangian radii which is the radius of an imaginary sphere around the centre of the stellar cluster containing a fixed fraction of its mass, is shown in Fig. 2. ## 3 RESULTS The main results of our simulations are presented in this section. We start by presenting our reference run in section 3.1. The dependence on the prescription for the accretion rate is explored in section 3.2, and an analysis in terms of collision and accretion time scales is provided in section 3.3. The influence of the IMF is examined in section 3.4, the dependence on the cluster radius in section 3.5, and the implications of mass loss are finally examined in section 3.6. The summary of the initial conditions and results is presented in table 1. ## 3.1 Reference run We start by describing our reference run with $N=5000$, $M_{\mathrm{cl}}=M_{\mathrm{g}}=1.12\times 10^5 \,\mathrm{M_{\odot}}$, $R_{\mathrm{cl}}=R_{\mathrm{g}}=1$ pc, assuming a Salpeter IMF within a stellar mass interval of $10-100 \,\mathrm{M_{\odot}}$ and a constant accretion rate of $\dot{m}=10^{-6}\,\mathrm{M_{\odot}}\,\mathrm{yr}^{-1}$. The result of the simulation is shown in Fig. 3. In this configuration, accretion occurs rather gradually, and the gas mass decreases only by about $20\%$ over 5 Myr, while the stellar mass increases by the same amount. As shown in Leigh et al. (2014) and as reflected in the evolution of the Lagrangian radii, the inner part of the cluster, containing $10\%$ of the mass, goes through contraction and even the radius containing $50\%$ of the mass shows slight contraction, while the outer $90\%$ radius is moderately expanding. The first collision in the cluster occurs after about $1$ Myr and in total about $20$ collisions occur over 5 Myr, where the time interval between the collisions decreases at later times, potentially due to the larger cross section of the central massive object. It is important to say that not all these collisions are with the central object, but nevertheless they lead to the formation of more massive objects within the star cluster, and particularly in the last 1.5 Myr the collision products eventually merge with the central object, thereby contributing to its growth. We now explore how these results depend on the adopted value of the accretion rate. We first note that if we decrease the accretion rate by a factor of $5$, no collisions occur and the results are consistent with a simulation where the accretion and collisions are switched off. On the other hand, if the accretion rate is increased to a value of $\dot{m}=10^{-5}\,\mathrm{M_{\odot}}\,\mathrm{yr}^{-1}$, it has a considerable impact on the evolution of the cluster (see Fig. 4). First, we note in this case that the gas mass will be fully depleted after slightly more than 2 Myr, and the stellar mass therefore reaches a total value of $2.24\times10^5 \,\mathrm{M_{\odot}}$. Stellar collisions are considerably enhanced. They occur relatively early on and increase more rapidly after about 1.5 Myr, reaching a total of about $500$ collisions within 5 Myr. The mass of the MMO reaches about $3\times10^4 \,\mathrm{M_{\odot}}$ in the same time, predominantly driven by the merger of collision products. The $10\%$ Lagrangian radius undergoes initially moderate contraction as before. It however decreases rapidly after $3$ Myr, as the accretion rapidly accelerates core collapse; our highest accretion rates achieve core collapse within one million years (see below). The $50\%$ Lagrangian radius slightly decreases for the first $1.3$ Myr and subsequently shows moderate expansion. The trend is very similar for the $90\%$ Lagrangian radius. This trend has been seen in previous simulations (e.g. Leighetal.2014), especially after the formation of a massive object in the center (Boekholtetal.2018; AlisterSegueletal.2020). As an even more extreme case, we consider the evolution for an accretion rate of $\dot{m}=10^{-4}\,\mathrm{M_{\odot}}\,\mathrm{yr}^{-1}$, as shown in Fig.. In this case, the gas mass is fully depleted after $\sim 0.2$ Myr. Collisions are considerably enhanced and grow rapidly from about $0.2$ Myr onwards, reaching a total number of about 550 collisions after 5 Myr. The mass of the MMO increases very rapidly due to mergers
201201456/4
mass growth due to accretion and collisions. Properly accounting for all of these physical processes provides a significant challenge with respect to the modeling. Moreover, it is computationally costly to model clusters consisting mostly of massive main sequence (MS) stars, due to the large number of stars. We use the Astrophysical MUlti-purpose Software Environment (AMUSE) (Portegies Zwart lan2018) to model the nuclear clusters and include all the physics mentioned above. AMUSE is very efficient in including new physics such as the mass-radius relation, accretion physics, collisional dy- namics, and coupling all of these to existing N-body codes. In this section, we present and discuss all the physical processes that we in- clude in our simulation and how we couple them into the numerical model. ## 2.1 Initial Conditions We model the NSCs with MS stars embedded in a stationary gas cloud. For simplicity, we assume that the stars and the gas are equally distributed, i.e., both the gas mass ($M_{\mathrm{g}}$) and gas radius ($R_{\mathrm{g}}$) are equal to the mass ($M_{\mathrm{cl}}$) and radius ($R_{\mathrm{cl}}$) of the stellar cluster. Both the cluster and gas follow a Plummer distribution with the same char- acteristic Plummer radius (Plummer 1911). The Plummer density distribution is given as $$ \rho(r)=\frac{3M_{\mathrm{cl}}}{4\pi b^3}\left(1 + \frac{r^2}{b^2}\right)^{-\frac{5}{2}}, $$ where $M_{\mathrm{cl}}$ is the mass of the cluster and $b$ is the Plummer length scale or the Plummer radius that sets the size of the cluster core. The Plummer length scale of the cluster (or the gas) is equal to one fifth of $R_{\mathrm{cl}}$ (or $R_{\mathrm{g}}$). We introduce a cut-off radius, equal to five times the Plummer radius, after which the density is set to zero in order to create a stable cluster with finite total mass and to make sure each star initially is within the gas cloud. The initial parameters for the numerical setup are $M_{\mathrm{cl}}$, $R_{\mathrm{cl}}$, $M_{\mathrm{g}}$, $R_{\mathrm{g}}$, and the number of stars $N$. For the initial mass of the stars we consider a Salpeter initial mass function (IMF)) given by: $$ \xi(m)\Delta m = \xi_0 \left(\frac{m}{\,\mathrm{M_{\odot}}}\right)^{-\alpha}\left(\frac{\Delta m}{\,\mathrm{M_{\odot}}}\right), $$ with three different mass ranges, $10\,\mathrm{M_{\odot}}-100\,\mathrm{M_{\odot}},\,10\,\mathrm{M_{\odot}}-120\,\mathrm{M_{\odot}},\,10\,\mathrm{M_{\odot}}-150\,\mathrm{M_{\odot}}$, assuming a power-law slope of $\alpha = 2.35$. Since our main goal is to explore the interplay between accretion physics and the dynamics of the runaway collisions producing a massive object, we do not take into account complicating factors such as the effect of cluster rotation or an initial binary fraction, which would also be highly uncertain in the context considered here, while significantly increasing the computational expense of our sim- ulations. The gravitational interactions between the stars are modelled us- ing the N-body code ph4 based on a fourth-order Hermite algorithm. The gravitational effect of the gas cloud is included via an analytical background potential coupled to the N-body code using the BRIDGE method (Fujii et al. 2007). This allows us to incorporate the grav- itational forces felt by the stars both due to the gas and the stellar component. In other words, the motions of the stars are controlled by the total combined potential of the gas and stars. As explained in more detail in the subsequent sections, we consider direct accretion onto stars but do not factor in gas dynamical friction. This is because in previous works it has been shown to be inefficient, although over sufficiently long timescales it can contribute to cluster contraction (Leighetal.2013b, 2014). $M_\ast-R_\ast$ relations for MS stars from different studies. For masses in the range $10-120\,\mathrm{M_{\odot}}$ we adopted the Demircan&Kahraman (1991); Schalleretal. (1992) data. The green dashed lines show the data from Bond et al. (1984). Even though the model is for really massive stars of mass $\gtrsim 10^4\,\mathrm{M_{\odot}}$, we can see it fits perfectly for masses $\gtrsim 50\,\mathrm{M_{\odot}}$. ## 2.2 Mass-radius relation The mass-radius ($M_\ast-R_\ast$) relation of the stars will play an important role in determining the number of collisions via the collisional cross section. Since the runaway collisions have to start before the most massive stars in the cluster turn into compact remnants, it is justified to consider all stars to be initially on the MS. In Fig. 1 we show the $M_\ast-R_\ast$ relations obtained from different studies (Bondetal.1984; Demircan&Kahraman1991; Schalleretal.1992). The Demircan& Kahraman (1991) and Schalleretal. (1992) data points match quite well below $M_\ast\lesssim 50\,\mathrm{M_{\odot}}$, while the Bondetal. (1984) and Schaller etal. (1992) data points match quite well above $M_\ast\gtrsim 50\,\mathrm{M_{\odot}}$. We therefore combine these relations and adopt a $M_\ast-R_\ast$ relation as outlined below. The $M_\ast-R_\ast$ relation for stars more massive than $M_\ast\gtrsim 100\,\mathrm{M_{\odot}}$ is of crucial importance, as it determines the size of the runaway collision product. Unfortunately, the radius of stars with $M_\ast\gtrsim 100\,\mathrm{M_{\odot}}$ is poorly understood. SMSs may develop a very extended envelope of very low density (Ishiietal.1999; Baraffeetal. 2001). However, this extended envelope contains only a few percent of the total stellar mass, so it may play a negligible role in collisions. Moreover, such extended envelopes are not present in low metallicity main sequence stars (Baraffeetal.2001), which are our main focus in this paper. Overall, the relation we employ is given as $$ \begin{aligned} \frac{R_\ast}{\,\mathrm{R_{\odot}}}&=&1.60\times\left(\frac{M_\ast}{\,\mathrm{M_{\odot}}} \right)^{0.47}\,\,\,\, \mathrm{for}\,\, 10\,\mathrm{M_{\odot}}\lesssim M_\ast<50 \,\mathrm{M_{\odot}},\\ \frac{R_\ast}{\,\mathrm{R_{\odot}}}&=&0.85\times\left(\frac{M_\ast}{\,\mathrm{M_{\odot}}} \right)^{0.67}\,\,\,\, \mathrm{for}\,\, 50\,\mathrm{M_{\odot}}\lesssim M_\ast, \end{aligned} $$ $$ \begin{aligned} \frac{R_\ast}{\,\mathrm{R_{\odot}}}&=&1.60\times\left(\frac{M_\ast}{\,\mathrm{M_{\odot}}} \right)^{0.47}\,\,\,\, \mathrm{for}\,\, 10\,\mathrm{M_{\odot}}\lesssim M_\ast<50 \,\mathrm{M_{\odot}},\\ \frac{R_\ast}{\,\mathrm{R_{\odot}}}&=&0.85\times\left(\frac{M_\ast}{\,\mathrm{M_{\odot}}} \right)^{0.67}\,\,\,\, \mathrm{for}\,\, 50\,\mathrm{M_{\odot}}\lesssim M_\ast, \end{aligned} $$ where Eq. 6 is adopted from Bondetal. (1984) and Eq. 7 is adopted from Demircan&Kahraman (1991). One key quantity in our work is the stellar lifetime. The MS lifetime is given as (Harwit1988) $$ \tau_{\mathrm{MS}}=10^{10}\left(\frac{\mathrm{\,\mathrm{M_{\odot}}}}{M_\ast} \right)^{2.5}\,\text{yr}. $$ For a $M_\ast\sim 20 \,\mathrm{M_{\odot}}$ it is $\sim$ 5.6 Myr. However, for $M_\ast\gtrsim 40\,\mathrm{M_{\odot}}$,
201201456/15
Summary of the results \begin{tabular}{cccccccc} \hline \hline $\dot{m}_{\mathrm{acc}}(\,\mathrm{M_{\odot}}\mathrm{yr}^{-1})$ &$\Delta M(\,\mathrm{M_{\odot}})$ & IMF($\,\mathrm{M_{\odot}}$) & $M_{\mathrm{cl}}=M_{\mathrm{g}}\,(\,\mathrm{M_{\odot}})$ & $R_{\mathrm{cl}}=R_{\mathrm{g}}\,(\mathrm{pc})$ & $N$ & $N_{\mathrm{coll}}$ & $M_{\mathrm{max}}$\hline $10^{-4}$ & None & 10-100 & 1.10$\times 10^5$ & 1.0 & 5000 & 563 & 3.16$\times 10^4$ $10^{-5}$ & None & 10-100 & 1.10$\times 10^5$ & 1.0 & 5000 & 490 & 2.90$\times 10^4$ $10^{-6}$ & None & 10-100 & 1.10$\times 10^5$ & 1.0 & 5000 & 20 & 1.29$\times 10^4$ $10^{-5}$ & None & 10-120 & 1.14$\times 10^5$ & 1.0 & 5000 & 440 & 2.88$\times 10^4$ $10^{-5}$ & None & 10-150 & 1.22$\times 10^5$ & 1.0 & 5000 & 404 & 2.84$\times 10^4$ $\dot{m}_{\mathrm{BH}}$ & None & 10-100 & 1.10$\times 10^5$ & 1.0 & 5000 & 153 & 1.12$\times 10^5$ $\dot{m}_{\mathrm{BH}}$ & None & 10-120 & 1.14$\times 10^5$ & 1.0 & 5000 & 152 & 1.19$\times 10^5$ $\dot{m}_{\mathrm{BH}}$ & None & 10-150 & 1.22$\times 10^5$ & 1.0 & 5000 & 178 & 1.26$\times 10^5$ $\dot{m}_{\mathrm{Edd}}$ & None & 10-100 & 1.10$\times 10^5$ & 1.0 & 5000 & 15 & 9.06$\times 10^2$ $\dot{m}_{\mathrm{Edd}}$ & None & 10-120 & 1.14$\times10^5$ & 1.0 & 5000 & 31 & 2.23$\times 10^3$ $\dot{m}_{\mathrm{Edd}}$ & None & 10-150 & 1.22$\times 10^5$ & 1.0 & 5000 & 23 & 2.64$\times 10^3$ $10^{-5}$ & None & 10-100 & 1.10$\times 10^5$ & 0.3 & 5000 & 1018 & 5.33$\times 10^4$ $10^{-5}$ & None & 10-100 & 1.10$\times 10^5$ & 0.6 & 5000 & 691 & 3.91$\times 10^4$ $10^{-5}$ & None & 10-100 & 1.10$\times 10^5$ & 2.0 & 5000 & 246 & 1.55$\times 10^4$ $10^{-5}$ & None & 10-100 & 1.10$\times 10^5$ & 5.0 & 5000 & 28 & 2.08$\times 10^4$ $\dot{m}_{\mathrm{BH}}$ & None & 10-100 & 1.10$\times 10^5$ & 0.3 & 5000 & 454 & 1.10$\times 10^5$ $\dot{m}_{\mathrm{BH}}$ & None & 10-100 & 1.10$\times 10^5$ & 0.6 & 5000 & 426 & 1.28$\times 10^5$ $\dot{m}_{\mathrm{BH}}$ & None & 10-100 & 1.10$\times 10^5$ & 2.0 & 5000 & 28 & 9.68$\times 10^4$ $\dot{m}_{\mathrm{BH}}$ & None & 10-100 & 1.10$\times 10^5$ & 5.0 & 5000 & 24 & 9.44$\times 10^4$ $\dot{m}_{\mathrm{Edd}}$ & None & 10-100 & 1.10$\times 10^5$ & 0.3 & 5000 & 409 & 2.19$\times 10^4$ $\dot{m}_{\mathrm{Edd}}$ & None & 10-100 & 1.10$\times 10^5$ & 0.6 & 5000 & 107 & 6.83$\times 10^3$ $\dot{m}_{\mathrm{Edd}}$ & None & 10-100 & 1.10$\times 10^5$ & 2.0 & 5000 & 12 & 2.88$\times 10^2$ $\dot{m}_{\mathrm{Edd}}$ & None & 10-100 & 1.10$\times 10^5$ & 5.0 & 5000 & 0 & 1.11$\times 10^2$ $10^{-5}$ & K15 & 10-100 & 1.10$\times 10^5$ & 0.3 & 5000 & 471 & 2.59$\times 10^4$ $10^{-5}$ & $0.3\%$ & 10-100 & 1.10$\times 10^5$ & 0.6 & 5000 & 272 & 2.69$\times 10^4$ $10^{-5}$ & $0.5\%$ & 10-100 & 1.10$\times 10^5$ & 2.0 & 5000 & 234 & 1.62$\times 10^4$ $\dot{m}_{\mathrm{BH}}$ & K15 & 10-100 & 1.10$\times 10^5$ & 0.3 & 5000 & 86 & 1.10$\times 10^5$ $\dot{m}_{\mathrm{BH}}$ & $0.3\%$ & 10-100 & 1.10$\times 10^5$ & 0.6 & 5000 & 79 & 2.95$\times 10^4$ $\dot{m}_{\mathrm{BH}}$ & $0.5\%$ & 10-100 & 1.10$\times 10^5$ & 2.0 & 5000 & 76 & 1.76$\times 10^4$ $\dot{m}_{\mathrm{Edd}}$ & K15 & 10-100 & 1.10$\times 10^5$ & 0.3 & 5000 & 33 & 1.68$\times 10^3$ $\dot{m}_{\mathrm{Edd}}$ & $0.3\%$ & 10-100 & 1.10$\times 10^5$ & 0.6 & 5000 & 16 & 6.34$\times 10^2$ $\dot{m}_{\mathrm{Edd}}$ & $0.5\%$ & 10-100 & 1.10$\times 10^5$ & 2.0 & 5000 & 11 & 2.56$\times 10^2$ \hline \end{tabular} $\dot{m}_{\mathrm{acc}}(\,\mathrm{M_{\odot}}\mathrm{yr}^{-1}$): accretion model, $\Delta M(\,\mathrm{M_{\odot}})$: Mass loss due to collision, IMF($\,\mathrm{M_{\odot}}$): Initial Mass Function used in the models, $M_{\mathrm{cl}}=M_{\mathrm{g}}\,(\,\mathrm{M_{\odot}})$:Initial mass of the cluster (= mass of the gas), $R_{\mathrm{cl}}=R_{\mathrm{g}}\,(\mathrm{pc})$: Initial radius of the cluster (=radius of the gas), $N$:Initial number of particles, $N_{\mathrm{coll}}$: Total number of collisions, $M_{\mathrm{max}}$: Final mass of the MMO in the cluster etal. (2015) for clusters without gas. A very similar dependence is found for Eddington accretion. It is interesting to note that a central massive object of mass $\sim 10^4 \,\mathrm{M_{\odot}}$ could be formed for the case of Eddington accretion if the cluster has central radius $\sim 0.3 pc$ or less. In the Bondi case, the final mass is independent of the size of the cluster, but the required timescale to form the central massive object can vary substantially depending on when the first collision occurs. A very relevant uncertainty in these models is the mass loss during collisions. We compared simulations employing different mass loss recipes, considering the prescription by Katzetal. (2015), where the effect of the mass loss is effectively negligible, as it becomes very small when the mass ratio between the collision partners is high. On the other hand, we noticed that a fixed mass loss of $3\%$ or $5\%$ during collisions has a more pronounced effect and can change the final mass by around an order of magnitude, as the mass loss may then become comparable to the mass gained via the collisions. It will thus be important to understand the expected mass loss from stellar models, as noted in the case of completely primordial clusters by AlisterSegueletal. (2020). As mentioned already in the introduction, the future will provide relevant opportunities for hopefully observing and probing the for- mation of very massive objects in the early Universe. From the work pursued here (and considering the results of other studies), we ex- pect that SMSs formed via moderate accretion rates of less than $10^{-3} \,\mathrm{M_{\odot}}$ yr $^{-1}$, the SMSs will appear with blue colours and high luminosities, while for larger accretion rates or in the regime when collisions occur extremely frequently, the SMSs may be bloated up and preferentially show red colours. It needs to be determined via future observations if such a bimodality actually occurs, or if there is a predominant way in which these types of objects may actually be found. Understanding the evolution of very massive metal poor stars in general will be important for further progress on the topic, as the lifetime of these stars can be a relevant limiting factor for the models. For a 20 $\,\mathrm{M_{\odot}}$ single star it is roughly 5.6 Myr. Above $\sim 40\,\mathrm{M_{\odot}}$, the lifetime is almost constant ($\sim 3.5-5$ Myr) (Hurleyetal.2000), while SMSs evolve over a timescale (given by eq. 9) ranging from $6$ to $0.06$ Myrs for SMS masses of $10^{3-5}\,\mathrm{M_{\odot}}$. These results may significantly change in the presence of rotation (e.g. Leigh et al. 2016), which can act as a stabilizing factor and an accelerator for the collision rate by reducing the relative velocities between objects due to more correlated motions in rotating cluster environments. It may also affect the accretion rate of gas onto stars, when the first colli- sions actually occur and even the underlying stellar evolution of the collision products. We conclude that it will be important to investi-
201201456/10
Figure5. Same as in Figure 3, but assuming instead $\dot{m}=10^{-4} {\,\mathrm{M_{\odot}}\mathrm{yr^{-1}}}$. Finally, in the case of Bondi accretion, the dependence on the initial mass and the resulting behavior is even more extreme. Due to the steep dependence of the rate on the accretor mass, the central mass remains low for extended periods of time, then suddenly accelerates in a run-away fashion, as seen in the previous subsection. When the upper-mass cutoff of the stellar mass is higher, this rapid acceleration occurs earlier. For the cases considered here, a final mass of $10^5 \,\mathrm{M_{\odot}}$ is reached in all cases, only at different times. We note again that the latter corresponds to a rather extreme assumption, so this part of the results needs to be regarded with caution. It is interesting to note that a more moderate increase of the mass of the MMO is found at earlier times when the Bondi accretion rate is low and a larger higher-mass cutoff in the IMF will favor this growth. Similar results were also found by Leighetal. (2013a). ## 3.5 Dependence on cluster radius To also understand how the different accretion mechanisms operate for different central densities of the cluster, we vary $R_{\mathrm{cl}}$ and thus the central density of the cluster. Specifically, $R_{\mathrm{cl}}$ has been varied between $0.3$ and $5$ pc, leading to an evolution as shown in Fig. 12 for the different scenarios. Depending on $R_{\mathrm{cl}}$, the initial central stellar densities range from $10^5\,\mathrm{M_{\odot}}\,\mathrm{pc}^{-3}$ up to $10^9\,\mathrm{M_{\odot}}\,\mathrm{pc}^{-3}$, and due to subsequent evolution and depending on accretion scenarios, even core densities of $10^8\,\mathrm{M_{\odot}}\,\mathrm{pc}^{-3}$ to $10^{12}\,\mathrm{M_{\odot}}\,\mathrm{pc}^{-3}$ can be reached. For a constant accretion rate of $\dot{m}=10^{-5}\,\mathrm{M_{\odot}}\,\mathrm{yr}^{-1}$, we find a con- siderable dependence on $R_{\mathrm{cl}}$, with collisions in principle occurring in all cases, but their number depending significantly on the size of the cluster (see table 1). For a large cluster with $R_{\mathrm{cl}}=5$ pc, the MMO still grows to about $2000 \,\mathrm{M_{\odot}}$, with most of the growth occurring af- ter $3.5$ Myr due to the collisions. Even for $R_{\mathrm{cl}}=2$ pc, the growth starts considerably earlier, with a steep rise in the mass of the MMO after $2$ Myr, and reaching a final mass of more than $10^4 \,\mathrm{M_{\odot}}$. For the most compact cluster with $R_{\mathrm{cl}}=0.3$ pc, rapid growth starts after $0.5$ Myr and the final mass reaches about $5\times10^4 \,\mathrm{M_{\odot}}$. The growth of the central object in these cases is mostly due to collisions, driven by the increase of the central density in the cluster, driven by contraction to
201201456/13
Figure8. Average accretion rate and maximum accretion rate in the cluster as a function of time, for the models assuming Eddington or Bondi accretion. We assume an IMF with stellar masses of 10-100 $\,\mathrm{M_{\odot}}$. the final mass is reduced to about $600\,\mathrm{M_{\odot}}$, and about $250\,\mathrm{M_{\odot}}$ for a $5\%$ fraction. The latter reflects roughly the expected trend regarding the final masses, with some super-imposed scatter due to the time of when the first collision occurs in the models. Finally, in the case of Bondi accretion, in principle a similar be- havior is found, only with an even steeper increase of the final masses at late times, and a more flat evolution initially, as also found before in our exploration of Bondi accretion. The highest final mass is at- tained here in the case with no mass loss, with about $10^5\,\mathrm{M_{\odot}}$ after about 4 Myr. Using K15, a very similar final mass is obtained after almost 5 Myr. The curves with $3\%$ and $5\%$ mass loss fractions behave similarly, starting to increase more significantly after about 3 Myr, reaching a maximum close to $10^5\,\mathrm{M_{\odot}}$ after about 4 Myr, while sub- sequently showing a decrease of their mass once the gas reservoir is exhausted and accretion stops, while mass loss still continues due to collisions. The final mass after 5 Myr is thus between $2-4\times10^4\,\mathrm{M_{\odot}}$ in these cases. ## 3.7 Caveats One of the main caveats of this work is that the simulation setup we employed here is highly idealised and not based on cosmological initial conditions, and also not solving the hydrodynamic equations. Hence, the processes in real cosmological systems could be different if there are differences in the structure, or if the accretion recipes that we explored here do not represent the type of accretion that will occur in these halos. The main goal of this work was to build a simplified model that allows us to study a large part of the parameter space, and to study the evolution of the cluster in that parameter space over timescales of a few Myr, which would not be possible if we were to include the full hydrodynamics. If we were to do so, then in addition to the the cooling, the chemistry of the gas and feedback processes due to the stars would need to be modeled in more detail. Recent studies by Kroupaetal. (2020) and Natarajan (2021) discuss in more detail about gas accretion onto black holes inside NSCs. The results of this work motivate us to run more simulations and will be able to guide future work to explore parts of the parame- ter space that can be regarded as particularly promising. It will be particularly important to understand which of the accretion recipes employed here, if any, provides an accurate prescription that will be consistent with results from real hydrodynamical simulations. We want to point out that sticking to only the first few Myr of evolution of the cluster minimizes concerns related to numerical issues i.e. if we were to re-run the simulations, we are confident that we would get the same results over the relevant timescales we consider. Over a longer timescale we would expect more divergence in the results of identical simulation setups. In addition, we note that other compli- cating factors such as mass loss due to stellar winds, stellar evolution, rotation etc. are currently not included in our simulations. Particu- larly, it is very uncertain how collisions affect stellar evolution, and the mass loss through winds for stars more massive than $1000$ M $_\odot$ is not well known. These provide very relevant uncertainties that need to be studied in more detail to provide reliable conclusions regarding collision-based formation mechanisms of SMBH seeds. ## 4 SUMMARY AND DISCUSSION In this work, we explored possible formation scenarios of SMSs in low-metallicity NSCs where mass loss from winds should be mini- mized. We explored how the interaction of collisions and accretion affects the formation of a central massive object for different physi- cally motivated accretion scenarios: considering accretion rates that are constant in space and time, the Eddington accretion rate and the Bondi accretion rate. Our fiducial model consists of a cluster with $N=5000$, $M_{\mathrm{cl}}=M_{\mathrm{g}}=1.12\times 10^5 \,\mathrm{M_{\odot}}$, $R_{\mathrm{cl}}=R_{\mathrm{g}}=1$ pc, assuming a stellar mass distribution with Salpeter IMF with mass range $10-100 \,\mathrm{M_{\odot}}$. Then we studied different accretion scenarios and also the effect of varying different parameters e.g. the upper mass limit of the Salpeter IMF, the central densities of the cluster and the inclusion of mass loss due to collisions on the final mass of the MMO in the cluster. In the case of a constant accretion rate, the number of collisions depends significantly on the magnitude of the accretion rate. Our simulations show that central massive objects of $\sim 10^{3-5} \,\mathrm{M_{\odot}}$ may form for accretion rates of $10^{-6}-10^{-4} \,\mathrm{M_{\odot}}$ yr $^{-1}$. If the SMSs formed in this way can directly collapse into seed black holes with similar mass, they can grow into the billion solar-mass black holes observed at $z\gtrsim 6$. In general, we find that the problem is quite sensitive to the initial conditions and the assumed recipe for the accretion, due to the highly chaotic and non-linear nature of the problem. The dominant factors that determine the mass growth are the mass of the MMO and the gas reservoir. Eddington accretion produces a central massive object of a few times $10^3 \,\mathrm{M_{\odot}}$ in our reference scenario, while Bondi accretion robustly produces final masses $\sim10^5 \,\mathrm{M_{\odot}}$. It is however important to note that the Bondi scenario assumes $\dot{M}_{\mathrm{BH}}\propto M_\ast^2$ and, as a result, the accretion rate remains low until a critical accretor mass is achieved leading to runaway growth. Only after a sufficiently long time of a few million years, the mass of the central object grows very rapidly. It is thus important to stress that such a scenario could only work if the gas is not previously expelled due to feedback, and if stellar evolution does not restrict the evolution to shorter timescales. In addition, Bondi accretion in general may be an extreme assumption, and we caution that some of the assumptions in our model will break down during the rapid phase of Bondi accretion due to the very rapid gas depletion. We compared the collision and accretion timescales computed both analytically and from our simulations. Our calculations suggest
201201456/1
# Formation of supermassive black hole seeds in nuclear star clusters via gas accretion and runaway collisions Arpan Das, $^{1}$ ★ Dominik R. G. Schleicher, $^{2}$ Nathan W. C. Leigh $^{2,3}$ and Tjarda C. N. Boekholt $^{4}$ $^{1}$ Department of Physics and Astronomy, University of Western Ontario, London, Ontario N6A 3K7, Canada $^{2}$ Departamento de Astronom í a, Facultad Ciencias F í sicas y Matem á ticas, Universidad de Concepci ó n $^{3}$ Department of Astrophysics, American Museum of Natural History, Central Park West at 79th Street, New York, NY 10024, USA $^{4}$ Rudolf Peierls Centre for Theoretical Physics, Clarendon Laboratory, Parks Road, Oxford, OX1 3PU, UK More than two hundred supermassive black holes (SMBHs) of masses $\gtrsim 10^9\,\mathrm{M_{\odot}}$ have been discovered at $z \gtrsim 6$. One promising pathway for the formation of SMBHs is through the collapse of supermassive stars (SMSs) with masses $\sim 10^{3-5}\,\mathrm{M_{\odot}}$ into seed black holes which could grow upto few times $10^9\,\mathrm{M_{\odot}}$ SMBHs observed at $z\sim 7$. In this paper, we explore how SMSs with masses $\sim 10^{3-5}\,\mathrm{M_{\odot}}$ could be formed via gas accretion and runaway stellar collisions in high-redshift, metal-poor nuclear star clusters (NSCs) using idealised N-body simulations. We explore physically motivated accretion scenarios, e.g. Bondi–Hoyle–Lyttleton accretion and Eddington accretion, as well as simplified scenarios such as constant accretions. While gas is present, the accretion timescale remains considerably shorter than the timescale for collisions with the most massive object (MMO). However, overall the timescale for collisions between any two stars in the cluster can become comparable or shorter than the accretion timescale, hence collisions still play a crucial role in determining the final mass of the SMSs. We find that the problem is highly sensitive to the initial conditions and our assumed recipe for the accretion, due to the highly chaotic nature of the problem. The key variables that determine the mass growth mechanism are the mass of the MMO and the gas reservoir that is available for the accretion. Depending on different conditions, SMSs of masses $\sim10^{3-5} \,\mathrm{M_{\odot}}$ can form for all three accretion scenarios considered in this work. Keywords: cosmology: theory — dark ages, reionization, first stars — black hole physics — galaxies: high-redshift — quasars: supermassive black holes — galaxies:star clusters: general ## 1 INTRODUCTION The observation of more than two hundred SMBHs with masses ${\rm \gtrsim 10^9\,\mathrm{M_{\odot}}}$ at redshift $z\, {\rm \gtrsim 6}$ 2019; Onoueetal.2019; Vitoetal.2019) has challenged our gen- eral understanding of black hole growth and formation. How these massive objects formed and grew over cosmic time is currently one of the biggest puzzles in astrophysics (Volonteri2010, 2012; Latif& et al. 2019; Latif & Schleicher 2019). Our current understanding is that the initial populations of black holes seeds were formed at $z \sim 20-30$ (Barkana&Loeb2001), and then they rapidly grow to their final masses by gas accretion and mergers (Dayaletal.2019; Pacucci&Loeb2020; Pianaetal.2021). In depth reviews about the formation and growth of SMBHs in the early universe can be found in Inayoshietal. (2019); Latif&Schleicher (2019); Haemmerléetal. (2020). The discovery of billion solar mass black holes at a time when the Universe was less than 1 Gyr old has created the so called “ seeding ” problem. In the standard Eddington limited accretion scenario a black hole of mass $M_\bullet$ grows exponentially over time, $$ M_\bullet(t)=M_{\bullet,0}\exp\left(\frac{1-\epsilon}{\epsilon}\frac{t}{t_{\mathrm{S}}} \right), $$ where $M_{\bullet,0}$ is the initial mass of the black hole or the ‘seed’ mass, $\epsilon$ is the radiation efficiency with standard value 0.1 (Shapiro2005) and $t_{\mathrm{S}}$ is the Salpeter time-scale, $t_{\mathrm{S}}=450$ Myr. In the standard light seeding models the SMBH starts growing from a Pop III stellar remnant of mass $M_\bullet\lesssim 100\,\mathrm{M_{\odot}}$ 2016). To reach a mass of $\gtrsim 10^9\,\mathrm{M_{\odot}}$ at $z\gtrsim6$ from a $100\,\mathrm{M_{\odot}}$ Pop III star, the growth process requires continuous Eddington accretion (or nearly continuous Bondi accretion (e.g. Leigh et al. 2013a)). However this is extremely difficult to achieve as the radiative and kinetic feedback from the stellar winds slow down the accretion of the gas (Tanaka&Haiman2009; Reganetal.2019). One possible solution to this problem could be galaxy mergers (Capeloetal.2015; Volonteri et al. 2016; Pacucci & Loeb 2020), which could lead to the mergers of massive BHs. However, during mergers the large recoil induced by gravitational wave emission can unbind the merger remnants from the shallow potential wells of their host galaxies (Haiman2004). An alternate solution relies on super-Eddington accretion (Volon- ★ E-mail: [email protected]
201201456/3
argued that the critical value decreases below $10^{-2}\,\mathrm{M_{\odot}}\mathrm{yr}^{-1}$ if the mass of the SMS is $\gtrsim 600\,\mathrm{M_{\odot}}$. For SMSs growing from runaway collisions in a cluster, it is assumed that a SMS persistently growing via collisions will similarly remain red and bloated as long as it is steadily bombarded (Reinosoetal.2018). The deposition of the kinetic energy causes the star to bloat post-collision, which then radiates away on a Kelvin-Hemholtz timescale (Eq. 9). However, if after coalescence the star is allowed to thermally relax, then it will contract to become blue. If the star goes through complete thermal relaxation, it will become a very hot source of ionization (Woods etal.2020). The relevant timescales both for collisions and accretion depend on the environment, but will be of order of one million years or so (e.g. Boekholt et al. 2018); and also the Kelvin-Helmholtz timescale will be of similar magnitude. However the timescale is expected to depend on environmental conditions e.g. the number density of stars (Leighetal.2017) and may also be enhanced in the presence of accretion. If detected via JWST or other missions, the star may be either in the hot or blue state depending on its environment. In general, however, the SMS phase is expected to be short lived, so the most likely observable would be a SMBH in a star cluster, potentially still with gas in its environment. Using existing data such as the Hubble Ultra Deep Field, it is currently not possible to detect either of these SMSs, as the AB magnitude limit is 29 at 1.38 $\mu m$, well below that expected for either type of SMS even at $z \sim 6$. However, with JWST which has NIRCam AB magnitude limits of 31.5, it will be possible to detect cooler, redder SMSs at $z\sim 18-20$ and hotter, bluer SMSs at $z\lesssim 13-10$ due to quenching by their accretion envelopes (Suraceetal.2019). Even though current Euclid and WFIRST detection limits (26 and 28, respectively) are well below the H band magnitudes of both stars at $z \sim 6-20$, both of them would be able to detect blue SMSs as even a slight amount of gravitational lensing will boost the fluxes of these blue SMSs above the detection limits of these two missions. Martinsetal. (2020) found some important spectral features in the observational properties of SMSs with mass range $\sim 10^3-5\times 10^4\,\,\mathrm{M_{\odot}}$. They have computed the spectra of SMSs for non-local thermal equilibrium spherical stellar atmosphere models. According to their model, cool SMSs with effective temperatures of $\sim\,10^4$ K will exhibit a Balmer break in emission, which is not expected for normal stars. Hotter SMSs with effective temperatures of $\gtrsim4\times10^4$ will exhibit a Lyman break in emission. However, the resonant scattering of Ly $\alpha$ photons by the neutral IGM at $z > 6$ will be a great challenge for observations. It is important to note that the detections of SMSs discussed in Suraceetal. (2018, 2019) are for high accretion rate $\sim 1\,\mathrm{M_{\odot}}\mathrm{yr^{-1}}$. Also, it will be difficult to observe the hotter blue supergiants at high redshifts due to the quenching effect mentioned above. Moreover, we do not expect the observations to measure anything related to SMSs directly. It would purely be a color measurement, which would be possible even if the entire cluster is unresolved and observed as a point-source. As our measurement capabilities increase with the upcoming telescopes to probe higher redshifts, a simple color measurement could be quite useful and informative, purely because these processes we quantify would yield a color inconsistent with a single burst of star formation. So potentially one can just compare to stellar population synthesis models. What future telescopes can do is still an open question. As another potential caveat, we note the importance to distinguish SMS signatures from other types of stars in massive star clusters, such as extended horizontal branch stars or A-type, which could even be brighter than blue stragglers in the same clusters (Leighetal.2016). Another challenge related to the observation of blue SMSs will be to distinguish them from hot blue dark stars powered by dark matter annihilation (e.g. Freeseetal.2010, 2016; Suraceetal.2019). One way to distinguish them is using the prominent continuum absorption features on the spectra redward of Ly $\alpha$ in the rest frame of the SMS due to the high accretion rate, which are absent from the spectra of blue dark stars. A very prominent Ly $\alpha$ line is found in blue SMS spectra due to pumping of the accretion envelope by high-energy UV photons from the star. These spectral features will be really important to distinguish blue SMSs from hot dark stars of similar mass. A dark star will not have these spectral features due to the absence of a dense accretion shroud. Studies have found that many galaxies harbor massive NSCs (Car- et al. 2016) with masses of $\sim 10^{4-8}\,\mathrm{M_{\odot}}$. Many of these galaxies host a central SMBH (Kormendy&Ho2013). Interestingly, studies have found correlations between both the SMBH mass and the NSCs mass with the galaxy mass (Ferrareseetal.2006; Rossaetal.2006; Leighetal.2012; Scott&Graham2013; Sethetal.2020). In many galaxies NSCs and SMBHs co-exist (Seth et al. 2008; Graham & Spitler 2009; Georgiev et al. 2016; Nguyen et al. 2019). Galaxies like our own (Schödel et al. 2014), M31 (Bender et al. 2005) and M32 (Nguyenetal.2018) host a SMBH and a NSC in the center. It therefore makes sense to consider a link between NSCs and SMBHs. Studies have shown that SMBH seeds could be formed inside NSCs via runaway tidal encounters (Stone et al. 2017) or from core col- lapse and stellar collisions (Devecchi & Volonteri 2009; Devecchi etal.2010; Daviesetal.2011) or by gas inflows (Lupietal.2014). Recent studies by Kroupaetal. (2020) and Natarajan (2021) investi- gated black hole accretion inside NSCs. In depth reviews about NSCs can be found in Neumayeretal. (2020). In this paper, we explore high-redshift, metal-poor NSCs as the possible birthplaces of SMSs via gas accretion and runaway stellar collisions. Low metallicity may favor the formation of very dense clusters, as fragmentation will occur at higher density, while the mass loss from winds becomes negligible (Vinketal.2001), thereby making this formation channel more effective. In addition, in low metallicity environoments the gas will be warmer, thereby contribut- ing to higher accretion rates through a higher speed of sound. 1 Here we explore how runaway collisions between stars and gas accretion onto the stars can lead to the formation of a SMS in such clusters. We use N-body simulations to model runaway collisions and gas accretion in dense NSCs. We describe our simulation setup in Sec. 2. Our results are presented in Sec. 3 and the final discussion is given in Sec. 4 along with a summary of our main conclusions. Similar work had been done by Boekholtet al. (2018); Reinoso etal. (2018) for Pop III stellar clusters. The most important distinc- tions between our work and the previous works are that we study clusters with higher initial stellar mass ($10^5\,\mathrm{M_{\odot}}$), higher number of stars (5000), a Salpeter type initial mass function (IMF), and more sophisticated and physically motivated accretion scenarios. We fur- ther provide a more detailed analysis here on the comparison of the timescales for collisions and accretion. ## 2 SIMULATION SETUP The complicated physical processes operating together in stellar clus- ters include gravitational N-body dynamics, gravitational coupling between the stars and the gas, accretion physics, stellar collisions and 1 In the regime where self-gravity regulates fragmentation and at least the initial masses of the clumps, the accretion rate can be estimated as the Jeans mass over the free-fall time, leading to a dependency on sound speed cubed.
201201456/9
Figure4. Same as in Figure 3, but assuming instead $\dot{m}=10^{-5} {\,\mathrm{M_{\odot}}\mathrm{yr^{-1}}}$. presence of an IMF will lead to mass segregation as massive stars tend to sink to the core through dynamical friction where they may eventually decouple from the remainder of the cluster. This is also known as “ Spitzer Instability ”(Spitzer1969). The Spitzer instability will lead to a shorter relaxation time in the core and an increased collisional cross section, both of which will increase the collision rate. What we can derive from the considerations in this subsection, is that collisions overall will be relevant and frequently occur, which may in turn may even accelerate the accretion process, due to the mass dependence in the Eddington and Bondi rate, and in particular early collisions may then determine the point when accretion starts to become more efficient. ## 3.4 Effect of the IMF To understand how the results so far depend on additional assump- tions, we explore how the results for the different accretion rates change using different assumptions regarding the IMF. We generally assume a Salpeter IMF with a lower-mass cutoff of $10 \,\mathrm{M_{\odot}}$, and vary the upper mass limit in the initial stellar mass distribution, consider- ing $100$, $120$ and $150 \,\mathrm{M_{\odot}}$. In the first case with a constant accretion rate of $\dot{m}=10^{-5}\,\mathrm{M_{\odot}}\,\mathrm{yr}^{-1}$, Fig. shows that such a variation has only a very moderate impact. The intrinsic scatter, when varying the initial conditions, means that the evolution at late times basically cannot be uniquely distinguished between the different cases, but it is likely that overlap in the parameter space of the results will occur when considering a larger set of initial conditions. For the case of Eddington accretion, we see very clearly that the MMO in our simulations grows more rapidly where the IMF extends to a higher mass. This is expected due to the dependence of the Eddington accretion rate on the mass of the accretor, accelerating both the accretion of the MMO as well as the conversion of gas mass into stellar mass in general. We note that the first collisions tend to occur earlier in the case of a larger initial mass, and we find a difference in the final mass by about a factor of 3 comparing the most extreme cases. In case of mass-dependent accretion recipes, the growth of the MMO is thus further enhanced if the initial IMF already extends to higher masses.
201201456/8
Figure3. Simulation with $N=5000$, $M_{\mathrm{cl}}=M_{\mathrm{g}}=1.12\times 10^5 \,\mathrm{M_{\odot}}$, $R_{\mathrm{cl}}=R_{\mathrm{g}}=1$ pc, assuming a Salpeter IMF between $10-100 \,\mathrm{M_{\odot}}$ and an accretion rate of $\dot{m}=10^{-6} {\,\mathrm{M_{\odot}}\mathrm{yr^{-1}}}$. The top left panel shows the evolution of gas and stellar mass as a function of time, the top right panel the evolution of the Lagrangian radii, the bottom left panel the total number of collisions as a function of time, and the bottom right panel the evolution of the mass of the MMO as a function of time. compact clusters with $R_{\mathrm{cl}}\ll0.1$ pc or the mass of the MMO would have to be much less than $100\,\mathrm{M_{\odot}}$. So in principle a situation can occur for which collisions will dominate over accretion, but it will be for a small part of the parameter space. We find a very similar result if we plot the ratio of these timescales directly from the con- ditions obtained in the simulations, which we plot in Fig. 10. Since our simulations have an initial range of masses between 10-100 $\,\mathrm{M_{\odot}}$ we assumed the most massive particle to be the type A and consider the rest of the particles as type B. The ratio of the timescales is again considerably larger than $1$. In principle this suggests that accretion should be more impor- tant than collisions for the clusters considered here. However, it is important to note from the results in section 3.1 that only a mod- erate number of collisions occurs for an accretion rate of accretion rate $\sim 10^{-6}\,\mathrm{M_{\odot}}\mathrm{yr}^{-1}$ (and we checked that no collisions occur for a low accretion rate $\sim 2\times10^{-7}\,\mathrm{M_{\odot}}\mathrm{yr}^{-1}$), but a relevant number of collisions occurs for a high accretion rate of $\sim 10^{-5}\,\mathrm{M_{\odot}}\mathrm{yr}^{-1}$ or $10^{-4}\,\mathrm{M_{\odot}}\mathrm{yr}^{-1}$. It is thus important to realize that the collision proba- bility is affected by accretion, an effect which is not incorporated into the formula above for the collision timescale. This is compatible with the work of Davisetal. (2010), who have shown that the number of stellar collisions should be $\propto N^{5/3}\dot{M} ^{2/3}$, where in their work $\dot{M}$ is the accretion rate onto the cluster, which is however reflected in accretion onto individual stars. We also emphasize that collision timescale in Eq. 21 is for a single object. To obtain the timescale for the collision of any two objects in the cluster, the collision timescale adopted here should be multiplied with $2/(N-1)$ (Leighetal.2017), bringing both timescales much closer together, or allowing collision timescale to be shorter for part of the parameter space. Hence, stellar collisions will be strongly favoured in clusters with high $N$ such as in all our models. However, it is important to note that this collision probability was computed using a simplified model where all the stars in the cluster are of the same mass, in contrast to a realistic IMF adopted in our models. The
220406532/5
The Heliospheric Ambipolar Potential Inferred from Sunward-Propagating Halo Electrons Figure2. Top: Sunward-directed field-aligned eVDF cuts (pitch angle bin $\theta$ $\in$ $[0^\circ,15^\circ]$). The different lines describe the average PAD measured at different distances ($r_1$, $r_2$,... $r_N$ = $r_8$), which are labeled by color. The legend shows the range of $r$ averaged in each distance bin. The colored sections of the plot indicate the halo regime used for the 1D fitting analysis (section 3.1). The continuation of the PAD averages outside our domain of interest (black) are not considered in our analysis. Bottom: The same averaged cuts, focusing only on the halo regime. It can be observed that the halo has a similar shape at each distance, up to an energy shift. - interpolation calculation. This $\sigma_{\Delta K, j}$ depends on the uncertainties in $f_A$ and $f_B$, the sampled energies of the spectra, and the intrinsic energy uncertainty of the SPAN-E detector (7%). (iii) The “ weighted mean ”(e.g., Barlow1989) of the energy difference $\Delta K_j$ is computed, which in the context of our Liouville mapping (10) is simply the energy shift $e\Delta \phi$: $$ e\Delta\phi(r_A, r_B) = \frac{\sum_{j=1}^{N} (\Delta K_j / \sigma^2_{\Delta K,j})}{ \sum_{j=1 }^{N} (1 / \sigma^{2}_{\Delta K,j})}. $$ Figure 3 illustrates the 1D Liouville mapping described above, for a particular case. The data in the lower panel are derived from the red ($r_1\approx$ 0.2 AU) and dark blue ($r_6\approx$ 0.5 AU) curves of Fig. 2. Taking these two energy spectra as $f_A$, $f_B$ respectively, we show the energy differences $\Delta K_j$ between $f_B$ and the interpolated spectrum $f_{int}$, in Figure 3. We apply the 1D Liouville mapping to the distance-averaged eVDFs to compute the change in potential $\Delta \phi$ (eq. 11). We use the most field-aligned cut of the eVDF (representing $0^\circ$ $<\theta<$ 15 $^\circ$). The eVDF at the sunward-most distance (at $r_A=r_1$ $\approx$ 0.2 AU) is used as the reference ($f_A$). This is compared to the corresponding $\theta\approx0^\circ$ cut ($f_B$) for each other distance $r_2$, $r_3$,... $r_N$, giving an estimate of $e\Delta \phi$ at each distance. The results are shown in Table 1. We find that the potential decreases by $\sim$ 250 V from 0.18 to 0.79 AU. It should be noted that the formula for the energy shift, eq. 10, only works for exactly field-aligned particles ($\theta=0^\circ$). Since an ESA can only sample a finite range of pitch angles, the 1D method has some systematic error because pitch angle broadening from moment conservation is not accounted for. This error manifested as a negative correlation between $f$ and $\Delta K$ (somewhat visible in Fig. 3). This error was generally small ($\sim$ 10%), because the average PADs deviate only slightly from isotropy. The 2D method, described below, is resilient against this error. ## 3.2 2D Liouville Mapping The 2D mapping is performed by fitting a function $f(r, K, \theta)$ of the form (6) to the full (0 $^\circ$ < $\theta$ <90 $^\circ$) sunward PADs. This single fit func- tion describes the eVDF at all distances observed. We will assume that at the outermost distance $r_N$ the logarithm of the distribution $f(r_N, K, \theta)$ can be matched to a 2D polynomial: $$ \ln f^\star(K, \theta) = \ln f(r_N, K, \theta) = \sum_{i=0}^D \sum_{j=0}^D A_{ij} \theta^i K^j. $$ $$ \begin{split} &\ln f(r, K, \theta) = \\ & \sum_{i=0}^D \sum_{j=0}^D A_{ij} \bigg(\sin^{-1} \sqrt{\frac{B(r_N)K \sin^2 \theta}{B(r)(K - e\Delta\phi(r,r_N))}} \bigg)^i \Big(K - e\Delta\phi(r,r_N)\Big)^j , \\ \end{split} $$ In the empirical formula (13), the terms are given in the following units: [f] = m $^6$ s $^{-3}$, [$K$]=eV, [$e\Delta\phi$]=eV, [$\sin^{-1}(x)$]=deg. so that the fit coefficients $A_{ij}$ are unitless. We fit a single 3D function (13) simultaneously to the N distance- averaged SPAN-E PADs, as measured at the distances ${r_1,r_2,..., r_N}$. We apply a 4th-degree polynomial (D=4) to match the data. To eval- uate the magnetic field $B(r)$ as it appears in eq. 13, we use PSP’s FIELDS fluxgate magnetometer data (averaged by distance bin, see Table 1). There are $(D+1)^2 + N-1$ fit parameters in total: the co- efficients $A_{ij}$ and potentials $\phi_k \equiv \Delta \phi(r_1, r_k)$ for $k=2,3,...N$ (note $\phi_1$ =0). The latter can be substituted into eq. 13 using the identity: $$ \Delta \phi(r_k, r_N) = \phi_N - \phi_k, $$
220406532/7
The Heliospheric Ambipolar Potential Inferred from Sunward-Propagating Halo Electrons Figure4. Halo electron pitch angle distributions ($0^\circ<\theta<90^\circ$, bin width $\Delta \theta$ =15 $^\circ$) measured by SPAN-E, averaged at $N$ =8 distance $r_1$, $r_2$,... $r_8$. Each panel represents a different distance, and different cuts of constant energy are distinguished by color. We considered PAD data in the energy range [391,935] eV and phase space densities within [2e-20,1e-17] m $^{-6}$ s $^3$. A single non-linear least squares fit is performed to all data shown (fit parameters listed in Tables 1, 2). The fit uses a flexible model equation (13), a 2D polynomial that is constrained to satisfy Liouville’s theorem. The electron energization $e\Delta\phi$ manifests in these plots as a vertical translation in phase space density of a given energy cut (color) between distances (panels).
220406532/2
consideration of instabilities that may affect the core (e.g., Schroeder etal.2021), which may even be generated by a resonant interaction with the deficit electrons themselves (Berčičetal.2021a). Thus the authors of B21 caution that their approach is “ simplified and includes strong assumptions ”, and may cause either a systematic overestima- tion or underestimation of the potential depending on how the method is applied. The core deficit in particular is not clearly described theo- retically, as this subtle feature is simply identified with the “ electron cutoff ” (a sharp discontinuity) in the sunward eVDF that is seen in exospheric models. Without more detailed predictions, the deficit can only be measured with a heuristic approach. The core perturbation has only been detected in 57% of eVDFs at heliospheric distances 20 – 85 $R_S$. This fractional occurrence decreases at the larger radial distances occupied by most spacecraft. In the current work, we develop a method that complements the pioneering study of B21 by considering the effect of the large-scale potential on a different part of the electron distribution: the suprather- mal “ halo ” population. The halo is identified in solar wind eVDFs as a nearly-isotropic tail at energies $\sim$ 100-1000 eV (e.g., Feldmanetal. 1975; Pilipp et al. 1987). As the halo energies are comparable to predictions for the inner heliospheric potential, this potential should profoundly affect the eVDF. In the present work we neglect the processes that generate the halo, but this topic deserves some review. Notably, the apparent growth of the halo at the expense of the anti-sunward suprathermal “ strahl ” population may imply that the halo is locally formed in the inner heliosphere by scattered strahl electrons (e.g., Maksimovic et al. 2005; Štverák et al. 2009). This has led to significant theoretical development, focused on the resonant interaction of electrons with the whistler and fast-magnetosonic whistler (FM/W) modes (e.g., etal.2022; Tangetal.2022). Observations have struggled to confirm these theories. Notably, whistlers are practically absent (occurrence rate $<$ 0.1%) during PSP perihelion passes (Cattelletal.2022). Addi- tionally, the eVDFs sampled by Helios and PSP are stable with respect to the oblique FM/W mode (Jeongetal.2022a). Theoretical calcula- tions show that at $r\lesssim$ 1 AU the strahl is stable to whistler fluctuations (Horaites et al. 2018b; Schroeder et al. 2021) and should be unaf- fected by whistler turbulence in the inner heliosphere (Boldyrev& Horaites2019). High-resolution measurements of the strahl at 1 AU confirm that “ anomalous diffusion ”, e.g. from whistler waves, is not required to explain the strahl angular widths at resolvable energies $\lesssim$ 300 eV Horaites et al. (2018a, 2019). Similar results were found from simulations at distances $r\lesssim$ 20 $R_S$ (Jeongetal.2022b), which showed that the strahl is adequately described by Coulomb collisions near the corona. This all suggests that a mechanism besides local wave particle scattering may account for the halo’s presence in the inner heliosphere. Such theories have been proposed (e.g., Leub- Scudder2019), though no consensus has emerged. Leaving the halo’s precise origin aside, we here adopt a very simple model. We assume that the sunward-moving halo electrons originated from some heliocentric distance beyond the view of PSP ($>$ 0.8 AU), as it is natural for highly energetic sunward-moving particles to orig- inate from larger distances. From the considerations detailed above, we are motivated to neglect scattering effects. The halo electrons will then be treated in the context of steady-state collisionless theory. Neglecting both Coulomb collisions and wave-particle interac- tions, we will apply Liouville’s theorem, which states that the eVDF is conserved along the electron trajectories. The electrostatic field essentially causes a shift in the halo energy spectra, and this energy shift directly corresponds with the ambipolar potential itself. We will therefore perform a “ Liouville mapping ”(e.g., Schwartzetal.1998; Lefebvre et al. 2007) to infer the potential from the eVDFs as ob- served by Parker Solar Probe in the inner heliosphere 0.18 – 0.79 AU (39 – 170 $R_S$). We will demonstrate that a single Liouville mapping can be applied to accurately model the halo eVDFs over a broad range of distances. Moreover, we will show that the ambipolar potential inferred from the eVDFs is exactly that required to accelerate the solar wind protons over these distances. This creates a coherent physical picture in which the same electric field causes two independently measurable trends as heliocentric distance increases: the halo loses energy while the bulk proton flow accelerates. ## 2 THEORY The Vlasov equation describes the evolution of the distribution func- tion $f(\mathbf{x}, \mathbf{v}, t)$ for a particle species in the absence of diffusion: $$ \dfrac{\partial f}{\partial t} + \mathbf{v}\cdot \nabla_x f + \mathbf{a} \cdot \nabla_v f = 0, $$ where $\mathbf{a}(\mathbf{x}, t)$ represents the acceleration due to the forces (Lorentz, gravitational) that act on the particles. Equation 1 neglects both Coulomb collisions and wave-particle interactions. The general so- lution, known also as Liouville’s Theorem, states that $f$ is constant along the particle trajectories: $$ f(\mathbf{x}(t), \mathbf{v}(t), t) = f(\mathbf{x}_0, \mathbf{v}_0, t_0), $$ In eq. 2, the coordinates $\mathbf{x}(t)$, $\mathbf{v}(t)$, $t$ describes the time-dependent motion of an arbitrary particle with initial position $\mathbf{x}_0 = \mathbf{x}(t_0)$ and velocity $\mathbf{v}_0 = \mathbf{v}(t_0)$, under the acceleration $d\mathbf{v}/dt =\mathbf{a}(\mathbf{x}(t), t)$. Liouville’s Theorem (2) is especially useful when the particle trajectories $\mathbf{x}(t)$, $\mathbf{v}(t)$ observe some constants of motion ($c_1$, $c_2$,...). In such a case it is possible to recast $f$ in terms of those constants (e.g., Pierrard&Lemaire1996), so that the function $f=\tilde f(c1, c2, ...)$ is a solution to the steady-state ($\partial f / \partial t$ =0) Vlasov equation (1). Let us consider a steady-state solar wind in the ecliptic, where the magnetic field $B$, ambipolar potential $\phi$, and the halo electron distribution $f$ vary spatially with the heliocentric distance $r$. For halo electrons in the solar wind, the relevant constants of motion are the total energy $\mathcal{E}$ and magnetic moment $M$: $$ %\mathcal{E} = \frac{m v^2}{2} - e \phi, \mathcal{E} = K - e \phi(r), $$ $$ %M = \frac{m v^2 (1 - cos^2 \theta)}{2 B}, M = \frac{K \sin^2 \theta}{B(r)}, $$ In eqs. 3 - 4, variables $K$ and $\theta$ represent the electron kinetic energy ($K=m_ev^2/2$) and pitch angle with respect to the magnetic field, respectively. We also introduce $e$, the modulus of the elementary charge. Note that in eq. 3 the Sun’s gravitational potential is ignored, as the gravitational energy of an electron at 1 AU is $\ll$ 1 eV. In the solar wind we expect the large scale gradients $d B/dr<$ 0 and $d\phi/dr<$ 0. Consequently, as a halo electron travels towards the Sun its pitch angle and kinetic energy will both tend to increase as the
220406532/6
Figure3. Top: In the 1D method (section 3.1), interpolation between two sunward-directed $\theta$ $\approx$ $0^\circ$ halo cuts at fixed phase space density (y-axis) yields a set of measurements $\Delta K_j$ of the energy shift (dashed lines). These are used to estimate the potential difference $\Delta\phi(r_A, r_B)$ between the distances $r_A$, $r_B$. Bottom: An example of the 1D method applied to our PAD data, showing measurements of the energy shift $\Delta K_j$ (blue dots) between the $r_1$ $\approx$ 0.2 AU and $r_6$ $\approx$ 0.5 AU cuts. A weighted average of the $\Delta K_j$ (eq. 11) yields the potential difference $\Delta\phi(r_1, r_6)$ =174.8 $\pm$ 10.4 eV. The final results for the 1D method potentials, $\Delta \phi(r_1,r_k)$ for each $r_k$, are shown in Table. A nonlinear least-squares fit is performed using the Levenberg- Marquardt algorithm, as implemented in the MPFIT software (Mark- wardt2009). The fit is seeded with initial guesses for the coefficients $A_{ij}$, which are generated by directly fitting the polynomial (12) to the PAD measured the outer distance $r$ = $r_N$. Through trial and er- ror we found that an optimal goodness-of-fit could be achieved by considering only energies in the range [391,935] eV. This removes $<$ 20% of our halo PAD data and has minimal impact ($\lesssim$ 10%) on the fit parameters. The final fit parameters $\phi_k$ and $A_{ij}$ are listed in Tables 1, 2 respectively. As can be seen in Table 1, the 1D and 2D methods yield similar estimates of the potential $\Delta \phi$. The SPAN-E PAD data are plotted in Figure 4 and compared with our fit results. In each panel of the plot we show energy cuts of the halo PAD (0 $^\circ$ $<$ $\theta$ $<$ 90 $^\circ$) averaged at a different distance. The fit to eq. (13), shown with solid lines in the plot, accurately describes all data shown in the 8 panels. We calculate a reduced chi-squared value of $\chi^2 / \nu=1.10$ ($\nu=$ 172 degrees of freedom, p-value), which is statistically significant. It is especially noteworthy that our fit successfully matched the averaged PADs even though they have very small error bars. As our model function (13) is constrained to comply with Liouville’s theorem, we may infer that Liouville’s theorem is consistent with the halo PAD data. To further investigate the significance of our results, we checked whether the collisionless model (13) could represent the strahl (which is superficially similar to the halo) equally well. This amounted to repeating our 2D fit method, but redefining the pitch angle: ${\theta\rightarrow (180^\circ-\theta)}$. This way, the fit domain $\theta\in[0^\circ,90^\circ]$ corresponded to the anti-sunward half of the eVDF. The converged fit (not shown) clearly did not match the strahl data, and resulted in an unacceptable reduced chi-squared value: $\chi^2/\nu$ =20.1 ($\nu$ =171 degrees of freedom, p-value $\lll$ 1). Of course, it is well-known that scattering (Coulomb collisions and/or wave-particle interactions) is required to explain the strahl angular width (e.g., Lemons & Feldman 1983), so it is expected that the model function (13) cannot explain the radial varia- tion of the strahl. But this illustrates that Liouville’s theorem is highly restrictive — it cannot be satisfied by arbitrary eVDF data. ## 3.3 Proton Acceleration The large-scale ambipolar potential is known to accelerate the solar wind proton flow. In the approximation that the potential in the eclip- tic is cylindrically symmetric and the protons flow radially outward, a change in potential corresponds directly to a change in radial flow speed. If the ambipolar potential changes with distance as we have inferred from the halo eVDFs (Sections 3.1, 3.2), then the proton flow energy should change by the same magnitude (up to a small correction due to gravity). To test this hypothesis, we examine the measurements of the pro- ton flow speed made by SWEAP’s SPC Faraday cup. We use the Level 3 velocity data, where for each SPC ion distribution a single 1-dimensional Maxwellian fit was applied in the inertial RTN frame. From this fit we can extract the radial ("R") component of the proton velocity $v_p$. We then compute the kinetic energy of the proton flow: $K_p = m_p v_p^2 / 2$, where $m_p$ is the proton mass. The measurements of $K_p$ are averaged into 4-hour intervals and subsequently averaged by distance, as was done with the electron PADs. This gives a proton energy measurement $K_p(r_k)$ at each distance $r_k$ (see Table 1). Because the protons (here treated as a mono-energetic beam) con- serve their total (kinetic+potential) energy, the change in electric potential $\Delta \phi$ is given by the formula: $$ e \Delta \phi(r_1, r_k) = \Delta K_p(r_k) + \Delta \Phi_G(r_k), $$ where we have defined $\Delta K_p(r_k)$ and $\Delta \Phi_G(r_k)$ as the changes in
220406532/4
referencing the FIELDS fluxgate magnetometer (Baleetal.2016). Each PAD has 12 evenly spaced pitch-angle bins (15 $^\circ$ wide each) and 32 energy bins logarithmically spaced over a range $\sim$ 2 – 1800 eV. Such PADs are appropriate for the study of solar wind halo, as the halo electrons are gyrotropic and have typical energies 100-1000 eV. We considered available Level 3 SPAN-E observations ($\sim$ $10-20$ second cadence) between the dates October 31, 2018 and De- cember 31, 2020. This time range is centered approximately around the solar minimum associated with the onset of Solar Cycle 25. During this time period the PADs are only available irregularly, but as the spacecraft completed multiple orbits there is sufficient coverage of the distances 0.18-0.79 AU (i.e. each distance is sam- pled by a few orbits). We remove data associated with co-rotating interaction regions (CIRs) and stream interaction regions (SIRs) en- countered by PSP, as indexed by Allenetal. (2021). Coronal mass ejections (CMEs) encountered by PSP are also removed using the HELIO4CAST ICMECAT list (Möstl et al. 2020). Guided by the criteria in Nieves-Chinchilla et al. (2018), it contains only ICME events that show clear signatures of magnetic obstacles. This helps for instance to avoid contamination by the “ counterstreaming ” strahl, which is a sunward-directed electron beam that is frequently seen during the passage of CMEs (Goslingetal.1987, 1992). Measure- ments taken while SPAN-E’s mechanical attenuator was deployed were also excluded from our data set, because the action of the at- tenuator reduces the ESA signal enough that the halo distribution can become contaminated with noise (Whittleseyetal.2020). As the attenuator was usually deployed during PSP’s perihelia passes, it is infeasible to investigate the halo at distances $r\lesssim$ 0.18 AU. In this study we compare electron data with proton data measured by SWEAP’s Solar Probe Cup (SPC) (Caseetal.2020). Therefore we require SPC data (i.e., the COHO hourly data product) to be available within 30 minutes of each SPAN-E PAD. We make a rough correction to isolate the slow wind, which has a different origin and composition than the “ fast wind ” (e.g., Verscharen et al. 2019a). After removal of transient events, we excluded any data for which the radial solar wind (proton) velocity exceeded the 95th percentile at that distance (e.g., McGregoretal.2011; Larrodera&Cid2020)— this mitigated the fast wind peak in the “ bimodal ” distribution of proton bulk speeds. As we are interested in finding the variation of the potential over large scales via a Liouville mapping, we compute the average PAD at different distances. We bin the PADs into N=8 logarithmically spaced distances 0.18-0.79 AU to compute the average PAD for each bin. Each average PAD is associated with a nominal distance $r_k$, which are indexed sequentially ($r_1$ < $r_2$ <...< $r_N$). The PADs, which are supplied by SPAN-E in terms of differential energy flux, are converted into phase space density (units m $^{-6}$ s $^3$) and are removed of outliers (0.01% of data) before averaging. We do not correct for the floating spacecraft potential as this should have a negligible effect ($\sim$ 2 eV) on the measured energies (Baleetal.2020). Before averaging, we re-orient the Level 3 PADs according to the 1-minute FIELDS magnetometer data so that $\theta$ =0 $^\circ$ corresponds with sunward-moving field-aligned electrons. We then test for the strahl by checking that the phase space density in the 486 eV channel is greater for the strahl direction ($180^\circ$) than for the sunward direc- tion ($0^{\circ}$) — any data that do not meet this test are removed ($<$ 5%). This process separates the anti-sunward strahl population ($\theta\approx 180^\circ$) from our measurements ($\theta<90^\circ$) of the sunward-oriented halo. The occasional “ sunward strahls ” associated with magnetic switchbacks Macneiletal. (2020) are consequently removed as well. In the anal- ysis presented here, we consider only energies for which the phase space density at $\theta$ $\approx$ 0 $^\circ$ was in the range [2e-20,1e-17] m $^{-6}$ s $^3$. This yields spectra with energies 200-1000 eV, avoiding the low-energy core as well as the high-energy “ superhalo ”(Lin1998; Wangetal. 2012). Technically, we pre-averaged the PADs into 4-hour long intervals before computing the average over distances. This helped to reduce bias that might arise due to uneven time cadence, and smoothed over short small-scale variability (e.g., Gazisetal.1999). We chose 4 hours as a pre-averaging interval because this corresponds with the eddy turnover time in the slow solar wind (Weygand et al. 2013), i.e. the time-scale over which successive measurements become un- correlated. We assume that 4-hour averages of solar wind data are independent and identically distributed (i.i.d.). This allows error bars of the distance-averaged PADs to be computed from the standard deviation of the mean. We expect i.i.d. is a coarse approximation as PSP measurements are affected by many processes on different timescales, e.g. turbulence (hours or less), data gaps (days), and solar rotation (weeks). The solar cycle, which occurs over an 11-year pe- riod, is unlikely to affect our analysis since our data are all measured near solar minimum. Energy cuts of the resulting distance-averaged PADs are shown in Figure 2, for different distances. The energy cuts of the distribution are taken from the most field-aligned sunward pitch angle bin of the PAD (0 $^\circ$ < $\theta$ <15 $^\circ$). It can be seen that at any given energy the phase space density of the halo decreases with the heliocentric distance $r$. Conversely, as it is relevant for our analysis, we note that for a fixed phase space density the energy decreases with distance. This is just the expected signature given by Liouville’s theorem for electrons moving in a potential gradient $d\phi/dr<0$. We performed a Liouville mapping using two different methods, which we will call “ 1D ” and “ 2D ”. These methods yield similar esti- mates of the potential $\phi(r)$, but the 1D technique is more illustrative and simple to implement while the 2D method is more general as it accounts for the angular variation of the eVDF. ## 3.1 1D Liouville Mapping A changing potential causes a corresponding shift in the eVDF. In the 1D Liouville mapping, we consider only the electrons that propagate towards the Sun with pitch angles nearly aligned with the magnetic field (i.e., $\theta \approx$ 0 $^\circ$). We may approximate that the kinetic energies of such electrons change, but their pitch angles stay field-aligned. It may be readily shown that as a consequence of Liouville’s theorem (6), the 1D energy cuts $f_A(K)=f(r_A,K,0^\circ)$ and $f_B(K)=f(r_B,K, 0^\circ)$ should look identical up to an energy shift due to the potential dif- ference $\Delta\phi$: $$ f_A(K) = f_B(K - e\Delta\phi(r_A,r_B)), %f_B(K) = f_A(K + e\Delta\phi(r_A,r_B)), $$ so the shift between two energy spectra gives the potential $\Delta \phi$. We apply the technique described in Horaites et al. (2021) to measure the energy shift between two 1D eVDF cuts. Given two discrete 1D energy spectra $f_A(K)$, $f_B(K)$ measured by SPAN-E at respective distances $r_A$, $r_B$, the following procedure is applied: - (i) A linear piecewise interpolation is performed on the spectrum $f_B$ to yield $f_{int}$. The interpolation energies $K_{int, j}$ are chosen so that each of the N data points in $f_{int}$ has a matching point in $f_A$ with the same phase space density, at energy $K_{A,j}$. (ii) For each such matched pair, the energy difference ${\Delta K_j = K_{A,j} - K_{int,j}}$ is computed. The uncertainty of this difference, $\sigma_{\Delta K, j}$, is computed by propagating errors through the
220406532/1
# The Heliospheric Ambipolar Potential Inferred from Sunward-Propagating Halo Electrons Konstantinos Horaites, $^{1}$ ★ Stanislav Boldyrev $^{2,3}$ $^{1}$ Department of Physics, University of Helsinki, Helsinki, Finland $^{2}$ Department of Physics, University of Wisconsin – Madison, Madison, WI 53706, USA $^{3}$ Center for Space Plasma Physics, Space Science Institute, Boulder, CO 80301, USA We provide evidence that the sunward-propagating half of the solar wind electron halo distribution evolves without scattering in the inner heliosphere. We assume the particles conserve their total energy and magnetic moment, and perform a “ Liouville mapping ” on electron pitch angle distributions measured by the Parker Solar Probe SPAN-E instrument. Namely, we show that the distributions are consistent with Liouville’s theorem if an appropriate interplanetary potential is chosen. This potential, an outcome of our fitting method, is compared against the radial profiles of proton bulk flow energy. We find that the inferred potential is responsible for nearly 100% of the proton acceleration in the solar wind at heliocentric distances 0.18-0.79 AU. These observations combine to form a coherent physical picture: the same interplanetary potential accounts for the acceleration of the solar wind protons as well as the evolution of the electron halo. In this picture the halo is formed from a sunward-propagating population that originates somewhere in the outer heliosphere by a yet-unknown mechanism. Keywords: solar wind – plasmas ## 1 INTRODUCTION Electrons in the solar wind occupy different collisional regimes, de- pending on their energy. The typical frequency of Coulomb collisions experienced by a test particle is known to fall off precipitously with the particle’s speed (as $v^{-3}$). At low energies $\lesssim$ 10 eV, collisions help to shape the electron velocity distribution function (eVDF) into a Maxwellian “ core ”. But the so-called “ suprathermal ” electrons, which have speeds much greater than the thermal speed, largely ig- nore their binary electrostatic interactions with nearby charged parti- cles. It is frequently appropriate to treat the suprathermal electrons as “ collisionless ”, so that their motion is only guided by the collective fields in the plasma. The influence of these fields on the particles can be loosely separated into two categories: the acceleration by large-scale fields and scattering by wave-particle interactions. The rapid escape of electrons to large heliospheric distances es- tablishes an equilibrium where the near-sun environment has a slight excess of positive charge. The resulting large-scale “ ambipolar ” or “ polarization ” electric field radially accelerates the solar wind ions, which then drag the electrons along. This field was not considered explicitly in the earliest hydrodynamic models of the solar wind (e.g., Parker1958). But it was gradually incorporated into the theory, where for instance it features prominently in so-called “ exospheric ” or kinetic models (e.g., Sen1969; Jockers1970; Lemaire&Scherer 1971). In these kinetic models, particles abruptly become collision- less above the “ exobase ” (e.g., Zouganelis et al. 2004; Boldyrev etal.2020). The estimated polarization potential between the exobase ($\sim$ 0.01-0.05 AU) and 1 AU is on the order of 100-1000 eV. Despite the general acknowledgment of a large-scale electro- static field’s presence in the solar wind, the magnitude of this field ($E$ $\sim$ $10^{-9}$ V/m) is far too small to be measured directly by spacecraft. However, recent progress has been made by considering the impact of the electric potential on the eVDF. In Berčičetal. (2021b), abbre- viated here as B21, the ambipolar potential was indirectly inferred from eVDFs measured by the Parker Solar Probe (PSP) SPAN-E electron instrument (Whittlesey et al. 2020). These measurements are based on two signals associated with the large-scale potential: 1) the “ deficit ” of sunward-moving electrons in the Maxwellian core (Halekasetal.2020, 2021), and 2) the “ breakpoint energy ”(Scudder &Olbert1979; Bakraniaetal.2020) that delineates the core from the suprathermal electrons. In B21, this theory was applied to PSP data to infer a parallel electric field $E_\parallel$ with a typical magnitude $\sim$ $10^{-9}$ V/m and scaling with heliocentric distance $E_\parallel\propto r^{-1.69}$. As reported, a solar wind accelerated by such a field would have 59% of the energy (77% of the speed) predicted by exospheric models at 45 solar radii ($R_S$). The remaining energy however presents a significant gap between these observations and theory. Although the results described above demonstrate a strong con- nection between the potential and the electron distribution, some lim- itations of the core deficit method can be recognized. The detailed physics governing the Maxwellian core distribution is complex, as the solar wind electrons are weakly collisional (Knudsen number $\gtrsim$ 0.01), and exhibit large plasma gradients (e.g., Baleetal.2013). Such a kinetic regime is difficult to treat with standard methods (e.g., Spitzer & Härm 1953; Gurevich & Istomin 1979). This is before ★ E-mail: [email protected]
220406532/3
The Heliospheric Ambipolar Potential Inferred from Sunward-Propagating Halo Electrons Figure1. As a consequence of Liouville’s theorem, an initial halo electron distribution at position $r_B$ will distort as the particles migrate towards the Sun along a field line (blue). This process is shown schematically. The green wedge-shaped region of the velocity distribution $f_B$ (imagined in $v_\parallel$, $v_\perp$ space) maps to the distribution $f_A$. The mirror force causes broadening, while the changing potential causes the electrons to gain energy. The red regions of $f_B$ represent particles that will be reflected by the magnetic field before arriving at $r_A$. The white semicircular vacancies in $f_A$, $f_B$ are the domain of the collisional core distribution, not considered here. particle attempts to conserve its values of $\mathcal{E}$ and $M$. The phase space density is conserved along the electron trajectories, so the sunward- moving half of the (gyrotropic) distribution function evolves with distance as shown in Fig. 1. Note that the 2D velocity space of the distribution $f$ may be writ- ten in terms of many sets of variables, e.g., $\{v_\perp$, $v_\parallel\}$, ${\{v,~\cos\theta\}}$, or ${\{K,~\theta\}}$. In the context of Liouville’s theorem it is often conve- nient to use $\{\mathcal{E}, M\}$. For the purpose of comparison with spacecraft measurements, in this work we will primarily use $\{K, \theta\}$. As discussed above, conservation of $\mathcal{E}$, $M$ leads to a general so- lution to the steady-state ($\partial f/\partial t$ =0) Vlasov equation (1) for the halo electrons: $$ f=\tilde f(\mathcal{E}, M). $$ Given such a function $\tilde f$, the solution $f(r, K, \theta)$ can be written im- mediately by substituting the definitions (3), (4) into (5): $$ f(r,K,\theta) = \tilde f\Big(\hspace{0.1cm} K - e \phi(r),\hspace{0.2cm}\frac{K \sin^2 \theta}{B(r)} \hspace{0.1cm}\Big). $$ In terms of the variables $r$, $K$, $\theta$, Liouville’s theorem requires that any solution for $f$ should be of the form (6), which just restates that $f$ is a function of the total energy and the magnetic moment. For simplicity we will only consider sunward-directed electrons, i.e. $\theta\in$ [0 $^\circ$, 90 $^\circ$]. This avoids accounting for the phase-space redundancy $\sin^2\theta = \sin^2(180^\circ - \theta)$, which would complicate the notation. We also treat the “ passing ” and “ reflected ” populations (e.g., Lefebvre etal.2007) equally, as the distinction is not of great importance here. Frequently, one wishes to map the eVDF from one location to another. Let us suppose that at a given position $r_\star$ the functional form of the eVDF, $f^\star(K, \theta)$, is known: $$ f^\star (K, \theta) \equiv f(r_\star,K,\theta). $$ $$ \begin{split} f(r,&K,\theta) = \\ & f^\star\bigg(\hspace{0.1cm}K - e\Delta\phi(r,r_\star)\hspace{0.1cm}\text{{\huge ,}}\hspace{0.2cm}\sin^{-1} \sqrt{\frac{B(r_\star)K \sin^2 \theta}{B(r)(K - e\Delta\phi(r,r_\star))}}\hspace{0.1cm} \bigg), \end{split} $$ where we have introduced the notation $\Delta\phi(r_A, r_B)$ to represent the potential difference between distances $r_A$ and $r_B$, $$ \Delta \phi(r_A, r_B)\equiv \phi(r_A)-\phi(r_B). $$ It can be verified by inspection that the mapping formula (8) has the form (6) required by Liouville’s theorem and satisfies the boundary condition (7). The arguments of the 2D function $f^\star$ in (8) are the formulas for how K, $\theta$ of a given electron respectively map across distance under the constraints $\mathcal{E}$ =const., $M=$ const. Technically, our Liouville mapping is derived for eVDFs measured simultaneously along the same field line, but we ignore this detail by assuming the solar wind in the ecliptic is steady-state and cylindrically symmetric. From equations 6, 8 it is readily seen that if the halo eVDFs are measured at two different distances along with the local magnetic field $B$, the only unknown variable is the potential $\phi$. With an accurate determination of $\phi$, a function of the form (6) can be found that matches two (or more) such eVDFs. Such a “ Liouville mapping ” may be accomplished with statistical methods (see sections 3.1, 3.2) and constitutes a measurement of the potential. In section 3 we apply a Liouville mapping to the average eVDFs measured by PSP, to infer the potential $\phi(r)$ in the inner heliosphere. ## 3 OBSERVATIONS Our primary data set comes from the SPAN-E electron experiment, a pair of electrostatic analyzers (ESAs) onboard the PSP satellite (Whittleseyetal.2020). SPAN-E is part of the Solar Wind Electrons Alphas and Protons (SWEAP) investigation (Kasper et al. 2016). Each SPAN-E ESA has a nominal field of view (FOV) of 240 $^\circ$ $\times$ 120 $^\circ$, but the instruments are arranged in a complementary fashion so that together they observe $>$ 90% of the sky after considering physical ob- structions such as the spacecraft heat shield. We studied the Level 3 pitch angle distribution (PAD) data, which were generated by col- lating the 3D eVDFs measured by the two ESAs into 2D ($K,~\theta$) distributions. The pitch angle $\theta$ is determined in the Level 3 data by
220406532/9
The Heliospheric Ambipolar Potential Inferred from Sunward-Propagating Halo Electrons \begin{tabular}{| c | c | c | c || c | c |} \hline $r$ range $[AU]$ & $r_k$ [AU] & B($r_k$) [nT] & $K_p(r_k)$ [eV] & 1D Method: $\Delta\phi(r_1,r_k)$ [V] & 2D Fit: $\phi_k \equiv \Delta\phi(r_1,r_k)$ [V] \hline $[0.18,0.21]$ & $r_1$ = 0.202 &54.3 $\pm$ 5.5 & 463 $\pm$ 31&-0.00 $\pm$ 0 & 0$[0.21,0.26]$ & $r_2$ = 0.237 &37.9 $\pm$ 0.9 & 499 $\pm$ 11&-18.9 $\pm$ 23.2 & 5.5457373$[0.26,0.31]$ & $r_3$ = 0.288 &26.0 $\pm$ 0.8 & 501 $\pm$ 12&25.56 $\pm$ 13.0 & 49.565330$[0.31,0.37]$ & $r_4$ = 0.343 &19.3 $\pm$ 0.7 & 478 $\pm$ 17&63.26 $\pm$ 13.2 & 100.22172$[0.37,0.45]$ & $r_5$ = 0.417 &12.4 $\pm$ 0.3 & 586 $\pm$ 24&134.6 $\pm$ 9.48 & 146.75335$[0.45,0.54]$ & $r_6$ = 0.498 &10.8 $\pm$ 0.3 & 649 $\pm$ 21&174.8 $\pm$ 10.4 & 185.08152$[0.54,0.65]$ & $r_7$ = 0.600 &8.20 $\pm$ 0.2 & 684 $\pm$ 20&213.9 $\pm$ 8.74 & 233.34487$[0.65,0.79]$ & $r_8$ = 0.712 &6.69 $\pm$ 0.1 & 696 $\pm$ 18&250.7 $\pm$ 7.80 & 278.53871\hline \end{tabular} Table1. Summary of physical parameters. —(cleanupvalues) We divide the range of distances in our data set [0.18,0.79] AU into logarithmically spaced bins, labeled in the column “ r range [AU] ”. We average over the available within each distance bin data to get a nominal position $r_k$, magnetic field $B(r_k)$, and proton bulk flow energy $K_p(r_k)$ — errors of these quantities are calculated as the standard deviation of the mean. Data are only considered into these averages at times when electron PADs are also available. As described in the text, for each distance $r_k$ the potential difference $\Delta \phi(r_1,r_k)=\phi(r_1)-\phi(r_k)$ is computed via the 1D and 2D method. In the 2D method, the potentials $\phi_k$ are fit parameters used in our model (13). \begin{tabular}{| c |} \hline 2D Fit Parameters $A_{ij}$ \hline \end{tabular} \begin{tabular}{| l | l | l | l | l |} \hline $A_{00}$= -36.939956 & $A_{01}$= -0.018878924 & $A_{02}$= 2.7295349e-05 & $A_{03}$= -4.1321067e-08 & $A_{04}$= 2.5958121e-11 % delete the last & above! $A_{10}$= -0.12204923 & $A_{11}$= 0.0013464076 & $A_{12}$= -5.1785826e-06 & $A_{13}$= 8.1686931e-09 & $A_{14}$= -4.5181963e-12 % delete the last & above! $A_{20}$= 0.0021237555 & $A_{21}$= -3.0146726e-05 & $A_{22}$= 1.2182674e-07 & $A_{23}$= -1.9396898e-10 & $A_{24}$= 1.0711158e-13 % delete the last & above! $A_{30}$= 1.4938766e-05 & $A_{31}$= -2.5832770e-10 & $A_{32}$= -1.4007037e-10 & $A_{33}$= 2.5217069e-13 & $A_{34}$= -1.2619921e-16 % delete the last & above! $A_{40}$= -3.3124890e-07 & $A_{41}$= 2.7152059e-09 & $A_{42}$= -9.5596288e-12 & $A_{43}$= 1.4803501e-14 & $A_{44}$= -8.2264039e-18 % delete the last & above! \hline \end{tabular} Table2. Fit coefficients $A_{ij}$ used in the 2D method, eq. (13). Formal errors are very small, so are not reported. If the sunward-moving halo electrons observed 0.18-0.79 AU are not produced locally by wave-particle diffusion, then they must have originated from some larger heliocentric distance. Such a mecha- nism of halo generation has not been deeply explored in the present body of research. However, as suggested in Horaites et al. (2019), if a sunward-moving suprathermal population is formed in the outer heliosphere, the process of magnetic mirroring should cause it to appear nearly isotropic in the inner heliosphere. ## 5 SUMMARY AND CONCLUSIONS Using PSP data, we have shown that sunward-moving halo electrons evolve in accordance to Liouville’s theorem in the inner heliosphere. This provides a very simple description of the halo dynamics. The potential $\phi(r)$ is the only quantity not measured in situ that is needed in order to map the halo eVDF from one location to another. This allowed us to apply the Liouville mapping technique to estimate $\phi$. We have independently measured the energy of the proton bulk flow, and found that the inferred potential has exactly the energy (within $\sim$ 10%) required to accelerate the slow solar wind. Our measurements of the ambipolar potential are similar to those provided by B21, at least at heliocentric distances where both tech- niques were applied. Extrapolating the power laws $\phi(r)\sim r^{\alpha_\Phi}$ pro- vided in B21 to distances $r\gtrsim$ 0.4 AU underestimates the change of potential $\Delta \phi$ compared to our analysis. As our approach is built on Liouville’s theorem, it rests on firmer theoretical ground than the core deficit approach in its current state of development. From an observational standpoint, the two methods are complementary, and may even inform each other. It is not feasible to apply our method at distances $r\lesssim$ 0.2 AU with available data, because of the practi- cal concern of instrument noise at energies $>$ 500 eV when PSP’s mechanical attenuator is engaged. On the other hand, the subtle mea- surements of the core deficit are reported to be infeasible at distances $\gtrsim$ 0.4 AU. A more detailed comparison of these methods is beyond the scope of this paper. We have shown that significant solar wind acceleration occurs be- tween 0.2 – 0.8 AU. Based on a wind speed of $\sim$ 400 km/sec at 1 AU during solar minimum (McGregoretal.2011), we may expect the potential to decrease by another $\sim$ 100 – 200 eV outside of PSP’s aphe- lion, $r_N\sim$ 0.8 AU. This means the halo energy shift caused by the large-scale potential may still be observable at distances $\gtrsim$ $r_N$. Unfor- tunately as the protons reach their asymptotic speed the acceleration will likely become even more difficult to discern in the variable so- lar wind data. It is worth noting as well that at distances $r$ $>$ 30 AU the solar wind actually decelerates, reportedly due to the pickup of interstellar material Elliottetal. (2019). We have assumed that the sunward-moving halo electrons evolve collisionlessly in the inner heliosphere, without experiencing wave- particle interactions or any other effect that could invalidate Liou- ville’s theorem as it is applied here. This might seem like a great leap. Wave-particle interactions are often invoked as a mechanism that could plausibly account for both the halo’s isotropy and its evolution with distance. However, our model also meets these requirements. The isotropy can be explained in terms of the mirror force, which broadens the PAD and also reflects the sunward-moving particles into an anti-sunward distribution. The radial evolution is directly ex- plained, as our fit function has been matched to the halo eVDF at all observed distances (Fig. 4). The Liouville mapping’s effectiveness suggests that the sunward halo did not experience local scattering during the PSP measurements. We additionally infer that if the strahl electrons undergo wave-particle interactions in the inner heliosphere, these interactions do not significantly influence the sunward-moving electrons. The present work applies a simple collisionless model to explain how the average halo eVDFs observed by PSP evolve in the inner heliosphere. Our results also support the basic premise of exospheric theories, that the proton flow speed is dictated by the large-scale po- tentials. However, if this model is correct, we require an explanation for how the seed population of energetic, sunward-moving electrons is formed in the outer heliosphere (or beyond). This highly motivating question can be addressed in future research. ## ACKNOWLEDGEMENTS
220406532/8
200 Figure 5. Top: The 4-hour averages of the radial proton bulk speed $v_p$, as measured by the SPC Faraday Cup, are shown as circles. The red line shows the average $v_p$ binned by distance (at $r_k$). The data show a steady increasing trend, except the dip around $r_4\approx$ 0.34 AU, which we treat as an outlier. Bottom: The change in electric potential $\Delta \phi(r_1,r_k)$ inferred in two different ways: 1) from the change in proton kinetic energy $\Delta K_p$, correcting for gravity (red, eq. 15), and 2) from the electron halo via the 1D (green) and 2D (blue) methods. The two independent estimations of the potential agree within the error bars. The change in potential is given with respect to the minimum distance $r_1=$ 0.202 AU. Two lines are plotted 0.1-0.4 AU, representing the variation of the potential $e\Delta\phi(r_1,r)$ derived from B21, for comparison (see text). kinetic and gravitational potential energy, relative to the innermost distance $r_1$: $$ \Delta K_p(r_k) \equiv K_p(r_k)-K_p(r_1), $$ $$ \Delta \Phi_G(r_k) \equiv G m_S m_p\Big(\frac{1}{r_1} - \frac{1}{r_k}\Big). $$ In eq. (17), $G$ is Newton’s constant and $m_S$ is the solar mass. When the gravitational potential is small, we see from (15) that a change in electric potential directly causes the same amount of proton accel- eration: $\Delta K_p \approx \Delta \phi$. Note though that for protons the gravitational energy cannot be completely neglected, as $\Delta \Phi_G \approx$ 40 eV across the distances 0.18-0.79 AU. In Fig. 5, we compare the change in potential $e\Delta \phi(r_1, r_k)$ as it is calculated independently in two ways: 1) from the proton kinetic energy $K_p$ (corrected by gravity, eq. 15) and 2) from the halo eVDFs (1D and 2D methods). The two estimates agree well at all distances 0.18-0.79 AU, implying a change in potential $\Delta \phi \sim$ 270 eV over the entire interval. This agreement implies that the electric potential in- ferred from the halo eVDFs fully explains proton acceleration over this distance interval. In rough numbers, the change in electric poten- tial $\sim$ 270 eV causes the halo electrons to lose this same amount of ki- netic energy, while the total (electric+gravitational) potential causes the protons to gain $\sim$ 230 eV in bulk flow energy. In terms of velocity, on average the protons start with radial speeds $v_p$ $\approx$ 290 km/sec at 0.18 AU and increase to $v_p$ $\approx$ 360 km/sec at 0.79 AU. ## 4 DISCUSSION It is appropriate to compare our measurements of the potential with the results of B21. In that work, the authors report that the large-scale potential varies as $\phi(r) = \Phi_0 (r/R_S)^{\alpha_\Phi}$ in the interval 0.1 $\lesssim$ $r$ $\lesssim$ 0.4 AU. They report two pairs of fit parameters: {$\Phi_0$ =1556.64 V, $\alpha_\Phi$ =- 0.66} and {$\Phi_0$ =1043.88 V, $\alpha_\Phi$ =-0.55}. From these profiles we cal- culate the change in potential $\Delta\phi(r_1, r)$, which we plot for com- parison as dashed/dotted lines in Figure 5. The B21 profiles agree roughly with our measurements of e $\Delta\phi$ in the interval where both techniques were applied (0.2-0.4 AU). The error bars, however, are comparable to the signal for these small energies ($\lesssim 50$ eV). We note that in our measurements $\Delta \phi$ changes by $>$ 100 volts in the interval [0.4,0.8] AU, while extrapolation of B21 yields only $\Delta \phi$ $\approx$ 30 – 40 volts over this same interval. An increase of hundreds of volts between 0.2 and 0.8 AU is not at all unreasonable in the slow wind. Such an increase may indeed be expected for the protons to accelerate from their modest $\sim$ 300 km/sec speeds observed at 0.2 AU to their typical $\sim$ 1 AU values $\sim$ 400km/sec (McGregoretal.2011). Our results displayed in Fig. 5 indicate that the acceleration of the slow solar wind is almost entirely due to the ambipolar electric field. Gravity is found to have a significant impact as well, and should increase in importance near the Sun (Lamyetal.2003). In the same sense, gravitational forces diminish rapidly with distance and may be entirely neglected in the outer heliosphere. As may be seen from Ulysses measurements (e.g. Štveráketal.2009), the halo itself does evolve slowly in the outer heliosphere. This evolution could be due in part or in whole to the radial variation of the ambipolar potential. The Liouville mapping technique accurately describes the halo eVDF. This implies that halo electrons are not affected by diffusive wave-particle interactions in the inner heliosphere. This is in itself an important result, as many theories of halo generation presuppose some local wave mechanism, that for example could scatter the strahl population into the halo. However, the absence of halo and/or strahl diffusion is consistent with the recent measurements of Cattelletal. (2022); Jeong et al. (2022a) which respectively show that whistler and FM/W waves do not scatter the strahl near the Sun. Our results do not preclude the occasional action of instabilities that derive their free energy from the halo particles, which have been observed e.g., at 1 AU (Tong et al. 2019). But to zeroth order we may infer that the average sunward halo eVDFs observed by PSP are not locally affected by these instabilities.
201011270/5
Figure 2: Elastic pendulum $$ x_{t+\Delta} \approx \frac{-b\Delta}{m}(x_t-x_{t-\Delta})-\frac{k\Delta^2}{m} x_t+2x_t-x_{t-\Delta} $$ ## Canonical weights Three physical constants govern Equation 11: the mass $m$, the spring constant $k$ and the damping coefficient $b$. From now on we will refer to them as canonical weights, to be learned by the solver network: $$ \mathcal{W} = [m,b,k]. $$ We chose the canonical weights of the network to reflect the proper parametriza- tion of our ODE for explanatory reasons only. As we will reveal in the result section, this choice allows us to demonstrate the capabilities of OscillatorNet in an intuitive manner. However if we want the learned values to be expressed in their proper SI units we have to initialize them with values of the correct order of magnitude. In Section 4.1.2 we generalize these results by training more general weights only dependent on the sampling frequency of the input time-series. For two input vectors $\mathbf{x}$ and $\mathbf{x}^{t-\Delta}$ we can express a finite difference step of the network in terms of the canonical weights as: $$ \mathopen{}\mathclose\bgroup\originalleft[ \begin{array}{cccc} \frac{k\Delta^2}{m}& \frac{-b\Delta}{m} \end{array} \aftergroup\egroup\originalright]\cdot \mathopen{}\mathclose\bgroup\originalleft[ \begin{array}{c} \mathbf{x}^t\\ \mathbf{x}^t-\mathbf{x}^{t-\Delta} \end{array} \aftergroup\egroup\originalright] $$ ## Experiments In our experiment we input 60 consecutive real-space trajectory points of a damped harmonic oscillator with mass $m = 1 \si{kg}$, damping coefficient $b=0.8 \si{kg/s}$ and spring constant $k=40 \si{kg/s^2}$ in a 1-layer OscillatorNet (kernel size = 1,
201011270/7
Figure 3: Top panel: Output from OscillatorNet without doubling the number of filters, i.e. the network cannot learn the dissipative term oft he governing equation. Botton panel: The network has enough free parameters to learn the full equation $\ddot{x} = f(x,\dot{x})$. The blue dots represent the output of 1 layer ResNet. Inset: Forecast if the network is trained on less than $1/4 \pi$ of the oscillator signal. ## Coupled harmonic oscillators OscillatorNet generalizes to arbitrary many coupled masses. Let’s take a look at two coupled oscillators (Figure 4), subject to different spring constants and damping coefficients, described by the system of ODEs:
201011270/12
Figure 7: Top panel: Network output of the mapping within the training set. Lower panel: Mapping output for causal padding (blue diamonds) and valid padding (green triangles) and their respective free forecasts in the test set (for clarity only the first ten points are plotted). The orange line is the network output using valid padding but with the inner feedback loop activated. padding during the training of the network, the numerical derivatives calculated on the mapping stencils are affected, as can be seen in the top panel of Figure 7 between timestamp t= $-60$ and t= $-54$. If we set padding to valid, these points are truncated. What is more evident, is that even though if padding is set to valid, the mapping is not learning the correct trajectory for the second mass, it approximates a higher amplitude solution, i.e. all points where the trajectories of the two oscillators time-series intersect, are matched by the mapped trajectory. This result is not surprising since we update only a subset of all canonical weights when we back-propagate the error, the network is free to learn a solution that is
201011270/14
Figure 8: Top panel: Blue diamonds show the quality of the mapping and the impact of zero padding, the green triangles show the mapping with padding set to valid. Botton panel: Free forecast of oscillator 1 on basis of wide mapping kernel with padding set to valid (solid green line), and causal (solid blue line). \begin{tabular}{|c||c|c|c|} \hline \multicolumn{4}{|c|}{\textbf{OscillatorNet, learned parameters with partial input}}\multicolumn{4}{|c|}{\textbf{padding = causal, mapping kernel = 25, stencil order = 5}} \hline & True value &OscillatorNet $\mathcal{V}$&Init.$\mathcal{V}$\hline \hline Masses [\si{kg}] & 1.5 & 1.173 &1 \hline Dampings [\si{kg/s}]& 0.5 & 0.487 & 1\hline Spring constants [\si{kg/s^2}]& 14, 35 & 14.558, 15.091 & 15, 15\hline \end{tabular}
201011270/2
equations relate the canonical coordinates, position and momentum $(\boldsymbol {q},\boldsymbol {p})$, of a system to its total energy, $\mathcal {H}(\boldsymbol {q},\boldsymbol {p}) = T(\boldsymbol {p})+ V(\boldsymbol {q})$ (for a 1-dim. system with one particle of mass $m$, $T= \boldsymbol {p}/2m$ and $V = V(\boldsymbol {q})$). $$ \frac {\mathrm {d} \boldsymbol {p}}{\mathrm {d} t}=-\frac {\partial {\mathcal {H}}}{\partial {\boldsymbol {q}}}\quad ,\quad \frac {\mathrm {d} {\boldsymbol {q}}}{\mathrm {d} t}=+\frac {\partial {\mathcal {H}}}{\partial {\boldsymbol {p}}}. $$ There have been several recent articles that focus on dynamical systems, only subject to conservative forces, in the context of machine learning [1, 2, 3, 4, 5]. The fact that such systems are reducible to first order ODEs qualifies them as interesting examples to be studied with residual type networks. It has been first observed by [6] that a ResNet block can be understood as a difference equation that approximates a first order ODE. Residual Neural Networks [7] are deep neural networks defined by stacking network blocks combined with a residual connection, adding the input of the block to its output (see left panel of Figure 1). If we assume a 1-layer convolutional network with $K$ hidden nodes and $\omega_h$, $h=1,...,K$ filter, applied on a 1- dimensional input $(x_t)_{t=0}^{N}$, this becomes: $$ x_{t+1, h} = x_t + \text{\emph{NL}} \bigg(\sum_{j=0}^{\infty} \omega_{h ,j}^{l=1}x_{t-j}\bigg), $$ or: $$ x_{t+1} = x_{t} + F(x_{t}, \Theta_t) $$ with $\Theta_t$ representing the learned network parameters and NL the chosen non-linearity. If we have more than one layer in the network and $x_l$ represents the hidden state of the $l^{th}$ layer, we can rewrite Equation 3 by introducing a strictly positive parameter $d$, representing the time discretization of a given observable: $d =$ timestep/number of layers. $$ x_{l+1} = x_l+ d\tilde{F}(x_l, \Theta_l). $$ For sufficiently small $d$, the residual block can be interpreted as a forward Euler discretization of a first order differential. In this sense Equation 4 builds the link between the network architecture and the dynamics of the system under observation. The family of problems captured by this formalism is limited to systems that are not subject to dissipation. In this note we are widening the scope and target non-reducible, second-order dynamical systems, i.e. systems that are subject to non-conservative forces as for example friction. We introduce a new network that incorporates the notion of a second-order time derivative, as a function of the systems real space observable and it’s time derivative. We then show that we can learn the systems approximate dynamics from its real space trajectory only. We then further ask the question if we still can learn the governing physics of a
201011270/16
\begin{tabular}{|c||c|c|c|} \hline \multicolumn{4}{|c|}{\textbf{OscillatorNet, learned parameters with partial input}}\multicolumn{4}{|c|}{\textbf{padding = valid, mapping kernel = 1, stencil order = 5}} \hline & True value &OscillatorNet $\mathcal{U}$&Init. $\mathcal{U}$\hline \hline parameter$_a$ $[\si{s}^{-2}]$ & 0.104 & 0.065 &0.066 \hline parameter$_b$ $[\si{s}^{-2}]$& 0.173 & 0.074 & 0.074\hline parameter$_c$ $[\si{s}^{-1}]$& 0.022 & 0.009& 0.066\hline parameter$_d$ $[\si{s}^{-1}]$& 0.022 & 0.022& 0.022\hline parameter$_e$ & 1.4 & 1.999 &2.0\hline \end{tabular} to non-conservative forces. Instead of locally interpolating a time-series, the network introduced learns a prior representing the governing physical laws and is therefore capable to forecast the systems behaviour over a large time-horizon. Moreover we have shown that given the correct parametrization of the network we can approximate the governing equations to high accuracy. We have further investigated the possibility that, given only partial information about a coupled system we can train an internal mapping to retrieve a stable forecast of this partial system over many points.
201011270/11
the mapping and the solver combined. However, since $\mathcal{V} \subset \mathcal{W}$, and we embed only one of the two trajectories in the network, when back-propagating the error not all canonical weights are updated but only the subset $\mathcal{V}$. This limits our capability in learning the physics that govern the system but not necessarily our ability to produce a stable forecast over a large horizon. Figure 6: Schematic of the network - mapping & solver ## Results It is only possible to share the parameters between mapping and solver, if we implement the projection according to Equation 18. In other words, the choice of canonical weights dictates the size of the convolutional kernel that learns the linear combination of $\alpha, \beta$ and $\gamma$. For the canonical weights chosen in our examples the kernel is limited to size $1$. We can choose a wider kernel but have to cut the umbilical cord between solver and mapping. In our experiments we have tried both, a kernel of size $1$ that allows us to share $m_1, b_1, k_1$ and $k_2$, as well as a wider mapping kernel of size $25$, where we update $\mathcal{W}$ only on the solver side. ## Canonical weights The first experiment shows the results of sharing the solver weights with the mapping, implemented according to Equation 18, with a projection kernel of size one. The blue diamonds show the mapping output if we set padding to causal, for the green triangles, padding has been set to valid. If we use zero
201011270/6
channels = 2, i.e. number of free parameters = 3). We can now carry out a single finite difference step while the parameters of the differential equation are automatically learned from the data. If we choose to forecast multiple points (free forecast), each prediction is fed back into the layer, before the network is re-trained to generate the next forecast. Figure 3 shows the results of a free forecast of 60 real-space trajectory points, i.e. except for the training set, no data of the time-series is used to train the network. The top panel shows the networks output with only one filter. We learn the correct frequency but the amplitude deviates increasingly as a function of time, due to the fact that we cannot learn a second inhomogeneous term as function of the oscillators momentum and therefor the total energy is conserved. The lower panel of Figure 3 shows the forecast of OscillatorNet with two filters. In comparison we also show the result of a single ResNet layer (blue dots). The fact that we can forecast multiple time-steps in a stable fashion, without exposing the network to more ground-truth data, indicates that we indeed learn the differential equation. To confirm this claim and to show that we accurately learn the governing physics of the system, we extract the trained canonical weights. A comparison of the true values versus the canonical weights learned by OscillatorNet can be found in Table 1. An impressive feature of the network is it’s ability to approximatively learn \begin{tabular}{|c||c|c|c|} \hline \multicolumn{4}{|c|}{\textbf{Learned Parameters of a single oscillator in SI units}} \hline & True value &OscillatorNet $\mathcal{W}$&Init.$\mathcal{W}$\hline \hline Mass [\si{kg}] & 2 & 2.058& 1\hline Damping [\si{kg/s}]& 1.5 & 1.487&1 \hline Spring constant [\si{kg/s^2}]& 40 & 40.249& 15\hline \end{tabular} the correct parameters even if it is only trained on a very small sample of the time-series. The inset in Figure 3 shows the output of the network if trained on less than a quarter period of the damped oscillator signal. The learned parameters can be found in Table 2. \begin{tabular}{|c||c|c|c|} \hline \multicolumn{4}{|c|}{\textbf{Training on less than} $\mathbf{1/4} \mathbf{\pi}$} \hline & True value &OscillatorNet&Init.$\mathcal{W}$\hline \hline Mass [\si{kg}] & 2 & 1.918& 1\hline Damping [\si{kg/s}]& 1.5 & 1.343&1 \hline Spring constant [\si{kg/s^2}]& 40 & 38.80& 15\hline \end{tabular}
201011270/4
Figure 1: Left: Schematic of the ResNet architecture. Right: Schematic of the OscillatorNet architecture. The yellow arrow indicates the residual connection, the green arrows indicate the differential residual . $$ \ddot{x} = f(x,\dot{x}). $$ ## Learning ODEs from real-space trajectories ## Damped harmonic oscillator The damped harmonic oscillator is a simple dynamical system subject to dissi- pation (see Figure 2), described by the differential equation: $$ \frac{d^2 x}{dt^2} = -\frac{b}{m}\frac{d x}{dt} - \frac{k}{m} x. $$ We replace the derivatives with their central difference approximation and get: $$ \bigg(\frac{x_{t+\Delta}-2x_t + x_{t-\Delta}}{\Delta^2} \bigg) \approx \frac{-b}{m} \bigg(\frac{x_t - x_{t-\Delta}}{\Delta} \bigg) - \frac{k}{m} x_t $$
201011270/15
we choose weights that only carry the units $[1/s]$ and $[1/s^2]$. In terms of the canonical weights, we can express the new set of parameters $\mathcal{U}$ as: $$ \begin{aligned} \text{parameter}_a &=& \frac{\Delta^2 k_2}{m_1}\\ \text{parameter}_b &=& \frac{\Delta^2 k_2}{m_2}\\ \text{parameter}_c &=& \frac{\Delta b_1}{m_1}\\ \text{parameter}_d &=& \frac{\Delta b_2}{m_2}\\ \text{parameter}_e &=&\frac{k_1+k_2}{k_2} \end{aligned} $$ $$ \begin{aligned} \text{parameter}_a &=& \frac{\Delta^2 k_2}{m_1}\\ \text{parameter}_b &=& \frac{\Delta^2 k_2}{m_2}\\ \text{parameter}_c &=& \frac{\Delta b_1}{m_1}\\ \text{parameter}_d &=& \frac{\Delta b_2}{m_2}\\ \text{parameter}_e &=&\frac{k_1+k_2}{k_2} \end{aligned} $$ $$ \begin{aligned} \text{parameter}_a &=& \frac{\Delta^2 k_2}{m_1}\\ \text{parameter}_b &=& \frac{\Delta^2 k_2}{m_2}\\ \text{parameter}_c &=& \frac{\Delta b_1}{m_1}\\ \text{parameter}_d &=& \frac{\Delta b_2}{m_2}\\ \text{parameter}_e &=&\frac{k_1+k_2}{k_2} \end{aligned} $$ $$ \begin{aligned} \text{parameter}_a &=& \frac{\Delta^2 k_2}{m_1}\\ \text{parameter}_b &=& \frac{\Delta^2 k_2}{m_2}\\ \text{parameter}_c &=& \frac{\Delta b_1}{m_1}\\ \text{parameter}_d &=& \frac{\Delta b_2}{m_2}\\ \text{parameter}_e &=&\frac{k_1+k_2}{k_2} \end{aligned} $$ $$ \begin{aligned} \text{parameter}_a &=& \frac{\Delta^2 k_2}{m_1}\\ \text{parameter}_b &=& \frac{\Delta^2 k_2}{m_2}\\ \text{parameter}_c &=& \frac{\Delta b_1}{m_1}\\ \text{parameter}_d &=& \frac{\Delta b_2}{m_2}\\ \text{parameter}_e &=&\frac{k_1+k_2}{k_2} \end{aligned} $$ We initialize the network with the same values as before, for the combined weights, all results can be found in Figure 9 and Tables 8, 9 & 10 \begin{tabular}{|c||c|c|c|} \hline \multicolumn{4}{|c|}{\textbf{OscillatorNet, learned parameters with partial input}}\multicolumn{4}{|c|}{\textbf{padding = causal, mapping kernel = 25, stencil order = 5}} \hline & True value &OscillatorNet $\mathcal{U}$&Init.$\mathcal{U}$\hline \hline parameter$_a$ $[\si{s}^{-2}]$ & 0.104 & 0.054 &0.066 \hline parameter$_b$ $[\si{s}^{-2}]$& 0.173 & 0.074 & 0.074\hline parameter$_c$ $[\si{s}^{-1}]$& 0.022 & 0.002& 0.066\hline parameter$_d$ $[\si{s}^{-1}]$& 0.022 & 0.022& 0.022\hline parameter$_e$ & 1.4 & 1.647 &2.0\hline \end{tabular} \begin{tabular}{|c||c|c|c|} \hline \multicolumn{4}{|c|}{\textbf{OscillatorNet, learned parameters with partial input}}\multicolumn{4}{|c|}{\textbf{padding = valid, mapping kernel = 25, stencil order = 5}} \hline & True value &OscillatorNet $\mathcal{U}$&Init. $\mathcal{U}$\hline \hline parameter$_a$ $[\si{s}^{-2}]$ & 0.104 & 0.062 &0.066 \hline parameter$_b$ $[\si{s}^{-2}]$& 0.173 & 0.074 & 0.074\hline parameter$_c$ $[\si{s}^{-1}]$& 0.022 & 0.048& 0.066\hline parameter$_d$ $[\si{s}^{-1}]$& 0.022 & 0.022& 0.022\hline parameter$_e$ & 1.4 & 1.956 &2.0\hline \end{tabular} ## Conclusion In this work we have introduced a new type of network, capable of solving the time-series patterns that form the solution of second order ODEs, subject
201011270/10
consisting of two coupled masses depicted in Figure 4. If only partial information about the system is available, i.e. just one time-series describing the trajectory of the first oscillator, a stable forecast is not possible (see top panel of Figure 5). Let’s go back to a system of coupled ODEs as in Equation 13. If the boundary conditions are fixed on one side, we can recover the position in time of a second coupled oscillator from the first. If we solve the first equation for $x_2$ and replace the derivative with it’s proper backward difference approximation, we arrive at: $$ x_{2}^t = \frac{1}{k_2}\{\frac{m_1}{\Delta^2}(x_1^t - 2^{t-\Delta}x_1+x_1^{t-2\Delta}) +\frac{b_1}{\Delta}(x_1^t - x_1^{t-\Delta})+k_1x_1^t\}+x_1^t. % $$ If we have more than two coupled oscillators in the system of consideration, the $i^\text{th}$ oscillator position is given by: $$ \begin{aligned} x_{i}^t = \frac{1}{k_i}\{\frac{m_{i-1}}{\Delta^2}(x_{i-1}^t - 2^{t-\Delta}x_{i-1}+x_{i-1}^{t-2\Delta}) \\ +\frac{b_{i-1}}{\Delta}(x_{i-1}^t - x_{i-1}^{t-\Delta}) \},\qquad 2< i\leq N \end{aligned} $$ $$ \begin{aligned} x_{i}^t = \frac{1}{k_i}\{\frac{m_{i-1}}{\Delta^2}(x_{i-1}^t - 2^{t-\Delta}x_{i-1}+x_{i-1}^{t-2\Delta}) \\ +\frac{b_{i-1}}{\Delta}(x_{i-1}^t - x_{i-1}^{t-\Delta}) \},\qquad 2< i\leq N \end{aligned} $$ Equation 17 can be realized with a convolutional network of the form: $$ x_{2}^t = \alpha \cdot \text{Conv}^3(x_1^t, [+1, -2, +1]) + \beta \cdot \text{Conv}^2 (x_1^t, [+1, -1])+ \gamma\cdot x_1^t. $$ We refer to this part of the network as mapping (see Figure 6), where we learn the relation between $x_1$ and $x_2$ ($[x_1,x_2]$ and $x_3$ for 3 oscillators, etc.). The weights of the convolutional kernel in Equation 18 are fixed and represent a stencil on which we perform the numerical derivative. It is important that we use backward finite difference coefficients to respect the causal order of the time- series. The accuracy of this numerical derivative can be increased by changing its order. For example the second order stencil carries the following coefficients: \begin{tabular}{|c|c|c|c| c|} \hline grid position: & -3&-2&-1&0\hline \hline $1^{st}$ deriv.& & 1/2 &-2 &3/2\hline $2^{nd}$ deriv.&-1 &4 &-5&2 \hline \end{tabular} The 3 parameters $\alpha, \beta$ and $\gamma$ in Equation 18 however are trainable by an additional convolutional kernel that acts as a projection operator: $$ \begin{aligned} \alpha = \frac{m_1}{k_2\Delta^2},\quad \beta = \frac{b1}{k_2\Delta},\quad \gamma = \frac{k_1+k_2}{k_2} \end{aligned} $$ Since these parameters are also a combination of canonical parameters that form a subset $\mathcal{V}$ of $\mathcal{W}$, the canonical weights that where introduced in the solver part of the network, it seems natural to share this subset over the whole network (solver and mapping). In other words, during training we update the weights of
201011270/13
\begin{tabular}{|c||c|c|c|c|} \hline \multicolumn{5}{|c|}{\textbf{OscillatorNet, learned parameters with partial input}} \multicolumn{5}{|c|}{\textbf{padding = valid, mapping kernel = 1, stencil order = 5}} \hline & True value &OscillatorNet $\mathcal{V}$&$\mathcal{V}$ with IFL &Init.$\mathcal{V}$\hline \hline $m_1$ [\si{kg}] &1.5 & 1.535& 1.089&1 \hline $b_1$ [\si{kg/s}]& 0.5 & 0.216 & 0.090&1\hline $k_1$, $k_2$ [\si{kg/s^2}]& 14, 35 & 15.000, 15.000 & 15.000, 15.000&15, 15\hline \end{tabular} numerically more stable, given it’s partial degrees of freedom: $m_1$, $b_1$, $k_1$ and $k_2$, while $m_2$ and $b_2$ are fixed at initialization. The lower panel of Figure 7 shows the out-of-sample free forecast based on the above training. Not surprisingly, the highest error is observed when padding is set to causal and therefore affecting the mapping kernel. But even the forecast with padding set to valid, proves to be only stable for the first 4 points before the error becomes too large. The forecast however can be stabilized, if we feed the mapped time-series back into the solver at each free-forecast step. The result of this inner feedback loop (IFL) is the orange curve in the lower panel of Figure 7. If we widen the projection kernel to size 25, we are no longer able to share the weights between solver and mapping, but on the other hand we get more stable results in terms of the quality of the forecast. Figure 8 shows the behaviour of the mapping network in the test and training set (blue diamonds for padding set to causal and green triangles for valid padding). If we take a look at the parameter learned (Table 6), and compare it to Table 3, the results of embedding both trajectories into the network, we can see that we achieve a lower error when mapping to a higher frequency solution. \begin{tabular}{|c||c|c|c|} \hline \multicolumn{4}{|c|}{\textbf{OscillatorNet, learned parameters with partial input}}\multicolumn{4}{|c|}{\textbf{padding = valid, mapping kernel = 25, stencil order = 5}} \hline & True value &OscillatorNet $\mathcal{V}$&Init.$\mathcal{V}$\hline \hline Masses [\si{kg}] & 1.5 & 1.112 &1 \hline Dampings [\si{kg/s}]& 0.5 & 0.908 & 1\hline Spring constants [\si{kg/s^2}]& 14, 35 & 14.959, 14.827& 15, 15\hline \end{tabular} ## Combined weights When we look at the above results, we find that all examples where we use mapping, fail at approximating the second spring constant. The reason lies in the fact that the chosen weights make the problem numerically ill defined. In an attempt to make the network more general and ensure more numerical stability,
201011270/1
# Learning second order coupled differential equations that are subject to non-conservative forces R. A. M ü ller 1, J. Laflamme-Janssen 1, J. Camacaro 1, and C. Bessega 1 In this article we address the question whether it is possible to learn the differential equations describing the physical properties of a dynamical system, subject to non-conservative forces, from observations of its real- space trajectory(-ies) only. We introduce a network that incorporates a difference approximation for the second order derivative in terms of residual connections between convolutional blocks, whose shared weights represent the coefficients of a second order ordinary differential equation. We further combine this solver-like architecture with a convolutional network, capable of learning the relation between trajectories of coupled oscillators and therefore allows us to make a stable forecast even if the system is only partially observed. We optimize this map together with the solver network, while sharing their weights, to form a powerful framework capable of learning the complex physical properties of a dissipative dynamical system. ## Introduction and Review Dynamical systems find manifold applications to processes in everyday life and allow insights into many areas not only of physics, but also mathematics, or theoretical biology. A dynamic system can be described by a mathematical model of a time-dependent process whose further course depends only on its initial state, but not on the choice of the starting point in time. In this work we focus on dy- namical systems that can be described by ordinary differential equations (ODEs). If the system is only subject to conservative forces, i.e. forces that can be derived from a scalar potential, higher order ODEs can be written as first order systems by introducing derivatives as new dependent variables of the system. An example of such a reduction are Hamilton’s equations (Equation 1). These
201011270/3
coupled dynamic system driven by non-conservative forces, if we only partially observe its real-space trajectories? From here the paper is structured in the following way: In Section 2 we introduce the concept of how residual connections can be used to emulate a second-order finite difference step. In Section 3 we put the architecture to the test and Section 4 deals with the question if we can further exploit the mathematical relation between real-space trajectories of coupled systems to approximate the differential equations of such a system if observed only partially. ## Architecture The base architecture of the network proposed, is inspired by a finite difference solver. We are looking for solutions of a second order (linear) problem of the form: $$ \ddot{x}+p(t)\dot{x}+q(t) x = r(t) $$ To be able to feed our network with real-space trajectory coordinates of a dissipative dynamical system, we need to incorporate the notion of a second order derivative into our architecture that we call OscillatorNet. We use the second order central difference approximation, where $\Delta$ represents the time discretization or sampling frequency of a time-series, $$ \ddot{x}_t \approx \frac{x_{t+\Delta}-2x_t+x_{t-\Delta}}{\Delta^2} + \mathcal{O}(\Delta^2), $$ to retrieve the residual architecture as: $$ x_{t+\Delta} = x_t+(x_t-x_{t-\Delta}) - F(x_t, \Theta_t) \cdot \Delta^2 $$ The right panel in Figure 1 shows a schematic of a second order differential block. Additional to the residual connection (yellow arrow) we add a differential residual (green arrows), i.e. we subtract in a separate channel $\mathbf{x}_{t-1}$ from $\mathbf{x}_t$ and then add it together with the residual to the network-block output, therefore implementing a second order finite difference step. The differential residual only adds the terms necessary to approximate a second order differential. The network therefore does not have parameters available to learn dissipation (see top panel of Figure 3). By doubling the number of filters we not only learn the inhomogeneous term as a function of the time series variable, i.e. the position (e.g. the force applied by a spring in an elastic pendulum; see lower panel of Figure 2) but it also becomes possible, to learn a second inhomogeneous term as a function of the derivative of the time series, i.e. the speed (e.g. the friction force acting on an elastic pendulum). This allows the system to numerically approximate the full differential equation generating the data:
201011270/9
Figure 5: Top panel: 60 points free forecast of one trajectory of a coupled harmonic oscillator after the network is trained on only on the trajectory of oscillator 1. Lower panel: 60 points free forecast of the full oscillator system after training on 60 points of each oscillator trajectory. \begin{tabular}{|c||c|c|c|} \hline \multicolumn{4}{|c|}{\textbf{Learned Parameters of a coupled oscillator in SI units}} \hline & True value &OscillatorNet $\mathcal{W}$&Init.$\mathcal{W}$\hline \hline Masses [\si{kg}] & 1.5, 0.9 & 1.514, 0.906 &1, 1 \hline Dampings [\si{kg/s}]& 0.5 , 0.3 & 0.483 , 0.297 & 1, 1\hline Spring constants [\si{kg/s^2}]& 14, 35 & 14.013, 34.472& 15, 15\hline \end{tabular} In a real-world application we might not be able to observe the complete configu- ration space of a coupled dynamical system. Let’s consider again the toy-model
201011270/8
$$ \begin{aligned} \frac{d^2{x}_1}{dt^2} &=& \frac{-b_1}{m_1}\frac{d{x}_1}{dt} - \frac{k_1}{m_1} x_1 +(x_2 - x_1)\frac{k_2}{m_1} \\ \frac{d^2{x}_2}{dt^2} &=& \frac{-b_2}{m_2}\frac{d{x}_2}{dt} -(x_2 - x_1)\frac{k_2}{m_2} \end{aligned} $$ Replacing the derivatives with the respective central difference approximations and solving for $\mathbf{x}^{t+\Delta}$ becomes: $$ \begin{aligned} x_1^{t+\Delta} &\approx&\frac{-b_1\Delta}{m_1}(x_1^t-x_1^{t-\Delta})-\frac{\Delta^2}{m_1}(k_1+k_2)x_1^t +\frac{k_2\Delta^2}{m_1}x_2^t +2x_1^t -x_1^{t-\Delta} \\ x_2^{t+\Delta} &\approx&\frac{-b_2\Delta}{m_2}(x_2^t-x_2^{t-\Delta})-\frac{k_2\Delta^2}{m_2}(x_1^t +x_2^t) +2x_2^t -x_2^{t-\Delta} \end{aligned} $$ The canonical weights of this problem are: $$ \mathcal{W} =[m_1,m_2,b_1,b_2,k_1,k_2] $$ For two input time-series $[\mathbf{x}_1, \mathbf{x}_1^{t-\Delta}$]and $[\mathbf{x}_2,\mathbf{x}_2^{t-\Delta}]$ we can express a finite difference step of the network in terms of the canonical weights as: $$ \mathopen{}\mathclose\bgroup\originalleft[ \begin{array}{cccc} \frac{-\Delta^2}{m_1}(k_1+k_2)& \frac{\Delta^2}{m_1}k_2& \frac{-\Delta b_1}{m1}&0\\ \frac{\Delta^2}{m_2}k_2& \frac{-\Delta^2}{m_2}k_2&0& \frac{-\Delta b_2}{m2} \end{array} \aftergroup\egroup\originalright]\cdot \mathopen{}\mathclose\bgroup\originalleft[ \begin{array}{c} \mathbf{x}_1^t\\ \mathbf{x}_2^t\\ \mathbf{x}_1^t-\mathbf{x}_1^{t-\Delta}\\ \mathbf{x}_2^t-\mathbf{x}_2^{t-\Delta} \end{array} \aftergroup\egroup\originalright] $$ We train OscillatorNet on 60 points of the signal by embedding each trajectory as a separate channel into the network. The network then carries out a single finite difference step for both time-series embeddings, while learning the parameters of the differential equations, before we feed the forecast back into the network and retrain to produce the next forecast. The lower panel in Figure 5 shows the results of the free forecast of 60 points for each trajectory, achieved in this manner. The learned weights can be found in Table 3. ## Exploiting the relation between real-space tra- jectories of a coupled system Figure 4: Damped coupled harmonic oscillator
210812810/20
Figure10: The dependence of $h^2 \Omega^{\rm peak}_{\rm GW}$ on the second Higgs boson mass $m_{h_2}$ with all other parameters fixed. The chosen original FOPT (before parameters’ variation) is denoted by a red circumference. \begin{tabular}{@{}rcccr@{}}\rightarrowprule & \multicolumn{3}{c}{Scenario 1} \cmidrule{2-4} & Parameter & Range & Distribution &\midrule & $m_\text{D1}$ & $[50,\,1000]\,\text{GeV}$ & linear& $m_\text{D2}$ & $[50,\,1000]\,\text{GeV}$ & linear && $\lambda_{\Phi \sigma}$ & $\pm[0.05,\,1]$ & exponential && $\lambda_{\sigma}$ & $[0.05,\,1]$ & exponential &\bottomrule \end{tabular} Table 3: Ranges of the input parameters in the scans for scenario 1. signal-to-noise (SNR) ratio for the phase transition (right) for a mission profile of 3 years. The color grade scale is the same on both plots. The right panel was generated using PTPlot 1.0.1 [34]. The colored isolines display the expected values for the SNR that depend on $T_*$, $g_\ast$ and $v_b$ while the dashed black contour lines represent the shock formation time $\tau_\mathrm{sh}$ (see Eq.(42)). The grey shaded region corresponds to an acoustic period lasting longer than a Hubble time and it is where the sound waves treatment is mostly reliable [30, 35]. For $\tau_\mathrm{sh} \ll 1$, the turbulence effects may become important dumping the acoustic contribution. However, none of our points feature a too small shock formation time. Using the formula for turbulence effects in Refs. [35, 56] for an estimate, we realize that it does indeed have very little impact in the peak position on the left panel. This first plot is just shown for reference. As concluded in the previous sections in the case where the singlet does not acquire a VEV, phase transitions are very weak with peaks amplitudes below $10^{-22}$. We have generated a total of 73 047 points points with a FOPT and several dedicated scans were performed but the trend did not change. Note that the actual role of turbulence is not yet well understood [57 – 59] and further studies are needed for a more reliable calculation of such a component. The SNR contours on the right panel take into account the effect of an increasingly short-lasting shock formation at the cost of a decreasing SNR value. In Fig. 12 we present a similar scatter plot but now for scenario 2. As previously discussed there are good chances of probing this model in some regions of the parameter space. In
210812810/18
A final word regarding the variation of the relevant parameters that measure the strength of the GW. Both $\alpha$ and $\beta/H$ show a shift that is proportional to the shift obtained with the SM masses with $\alpha$ with a positive slope and $\beta/H$ with a negative one. ## Scenario 3 We now move to scenario 3. In Table 2 we present the ranges of variation for the input values of the model. The ranges for the parameters of the potential were already explained. The ranges for the specific parameters of the neutrino masses and Yukawa coupling were chosen such that the Majorana-like Yukawa coupling $Y_\sigma$ is sizeable enough, but still perturbative, in order to modify the thermal coupling $c_\sigma$ in comparison to the singlet model. The parameter $M_n$ establishes a mass scale for the heavy neutrinos not to far from the EW one provided that such a scenario relies on a low-scale inverse seesaw mechanism. This is also convenient for our fixed scale treatment since renormalization group effects in the neutrino sector can be ignored. \begin{tabular}{@{}rcccr@{}}\rightarrowprule & \multicolumn{3}{c}{Scenario 3} \cmidrule{2-4} & Parameter & Range & Distribution &\midrule & $m_{h_2}$ & $[50,\,1000]\,\text{GeV}$ & linear& $v_{\sigma}$ & $[50,\,1000]\,\text{GeV}$ & linear && $\mu_{b}^2$ & $[-500000,\,-1250]\,\text{GeV}^2$ & linear && $\theta$ & $[-\arccos{(0.85)},\, \arccos{(0.85)}]$ & linear && $M_n$ & $[50,\,550]\,\text{GeV}$ & linear& $Y_\sigma$ & $[0.01,\,\sqrt{4 \pi}]$ & exponential &\bottomrule \end{tabular} Table 2: Ranges of the input parameters in the scans for scenario 3. In Fig. 9 we present six points that could be probed by LISA in the scope of scenario 3, identified by a red circle. As for scenario 2, we present in the left column the variation of the peak amplitude with the variation of the Higgs mass in the interval 124.96 GeV to 125.24 GeV ($\pm 1$ standard deviation), while keeping the remaining parameters of the SM constant together with those from the dark sector. Again the top quark mass (middle column) was varied between 172.46 GeV and 173.06 GeV. From all the points above LISA the maximum variation found for the peak amplitude was 150% for the variation with the Higgs mass and 30% for the variation with the top quark mass. This is in line with what was obtained for scenario 2. We end this section showing in Fig. 10 the dependence of $h^2 \Omega^{\rm peak}_{\rm GW}$ on the second Higgs boson mass $m_{h_2}$ with all other parameters fixed. As expected, since we allow for a much larger variation of order 5%, the impact on the amplitude of the primordial GW spectrum is huge spanning several orders of magnitude. This is the general trend for all other points in the scan with strong FOPTs. ## Comparison of the results for three scenarios In this last section we will focus in more detail on the observability of GWs by the LISA experiment starting with scenario 1. In Table 3 we show the range of variation for the parameters in scenario1.
210812810/17
Figure 8: The dependence of $h^2 \Omega^{\rm peak}_{\rm GW}$ on the Higgs boson mass $m_{h_1}$ (left) and on the top-quark mass (center). In the right plot we present the corresponding variation for the parameters $\alpha$, $\beta/H$ and $T_*$, where the subscript $0$ denotes the central values at $m_{h_1} = 125.1$, the top text is the relative change between the strongest point and weakest point. For each point varied all other model parameters are fixed. The chosen original FOPT (before parameters’ variation) is denoted by a red circumference. All masses are expressed in GeV. uncertainty in the top quark mass of $\pm 1$ standard deviation. The variation is clearly smaller than that for the Higgs mass but it is still relevant in the context of the considered variation in the top quark mass with a maximum of 50%. In conclusion, if a GW signal is detected one must be cautious in taking too strong conclusions about either the model or its parameters since the experimental uncertainty in the SM parameters can still play a significant role. With this in mind, a more precise determination of the Higgs and top quark masses can be rather important in light of a hypothetical discovery of a primordial GW signal and its theoretical interpretations, suggesting a further motivation for lepton colliders in the future.
210812810/19
Figure 9: The dependence of $h^2 \Omega^{\rm peak}_{\rm GW}$ on the Higgs boson mass $m_{h_1}$ (left) and on the top-quark mass (center). In the right plot we present the corresponding variation for the parameters $\alpha$, $\beta/H$ and $T_*$, where the subscript $0$ denotes the central values at $m_{h_1} = 125.1$. For each point varied all other parameters of the models are fixed. The chosen original FOPT (before parameters’ variation) is denoted by a red circumference. All masses are in GeV. In Fig. 11 we present scatter plots for scenario 1 showing the GW peak position as a function of the strength of the phase transitions $\Delta v_{h}/T_*$ in the colour scale (left) and the corresponding
210812810/5
in terms of $h$ - $\sigma_R$ mixing angle $\theta$, while the the DM candidate gets a pseudo-Goldstone mass, $$ \begin{aligned} m_D^2\equiv m_{\sigma_I}^2 = - 2\mu_b^2 \,, \qquad \mu_b^2 < 0 \,. \end{aligned} $$ Finally, scenario 3 is exactly the same as scenario 2 in what concerns the scalar sector and the only difference resides in the fermion content of the model. In particular, three families of right-handed neutrinos $\nu_{1,2,3}^c$ carrying lepton number $L(\nu^c) = -1$ and three families of singlet fermions $S_{1,2,3}$ with the opposite lepton number, i.e. $L(S) = 1$, are introduced such that one can write two additional Yukawa interactions, one of them of the Dirac-like tying together the Higgs, lepton doublets and right-handed neutrinos while the other coupling is of the Majorana-like and ties the two singlet fermions with the complex singlet, $Y_{\sigma_i} S S \sigma$ (for this reason, $\sigma$ is dubbed Majoron in this model) which is invariant under the lepton number $\mathrm{U}_\mathrm{L}$ symmetry provided that $L(\sigma) = -2$. Note that as long as the singlet $\sigma$ develops a VEV, a Majorona mass term of the form $\mu_i S_i S_i$ is induced, with $$ \mu_i = \dfrac{Y_{\sigma_i}}{\sqrt{2}} v_\sigma\,, $$ where, for simplicity of illustration, we assume a flavour diagonal basis. The lepton number symmetry also allows a mass term of form $M_i \nu_i^c S_i$, which is also considered in our numerical analysis. Since the singlet VEV is expected to be generated not far from the EW scale, the model features a low-scale type-I seesaw mechanism. ## Gravitational waves from FOPTs In this section, we will define the physical quantities relevant for understanding the characteris- tics of the GW signals originating from EW FOPTs in the early Universe. A detailed knowledge of the effective scalar potential at finite temperatures $V_{\rm eff}(\phi_{\alpha};T)$ is important in order to obtain the key parameters of the primordial GWs power spectrum. To the one-loop order, the effective potential takes the form [16, 17], $$ V_{\rm eff}(T) = V_0 + V^{(1)}_{\rm CW} + \Delta V(T) + V_{\rm ct}\,, $$ in terms of $V_0$ and $V^{(1)}_{\rm CW}$ being the tree-level (classical) part and one loop Coleman-Weinberg (CW) potential, respectively, and the counterterm potential $V_{\rm ct}$, while finite-temperature cor- rections are denoted as $\Delta V(T)$. The one-loop zero-temperature effective potential is given by the standard formula [18] (in the $\overline{MS}$ scheme and in the Landau gauge) $$ V^{(1)}_{\rm CW} =\frac{1}{64 \pi^2}\sum_{a}n_a m_a^4(h,\phi)\left[\log\frac{m_a^2(h,\phi)}{\mu^2}-C_a\right], $$ where $n_a$ counts the number of degrees of freedom and for a particle of spin $s_a$ is given by $$ n_a=(-1)^{2s_a} Q_a N_a (2s_a+1),\notag% $$ where $N_a$ stands for the number of colours and $Q_a=1,2$ for neutral/charged particles. $m_a(h,\phi)$ correspond to (tree-level) field-dependent masses.
210812810/7
While for scenarios 1 and 2 we have $$ \begin{aligned} & c_h = \frac{3}{16} g^2 + \frac{1}{16} {g'}^2 + \frac12 \lambda_\Phi + \frac{1}{12} \lambda_{\Phi \sigma}+ \frac14 (y_t^2 + y_b^2 + y_c^2 + y_s^2 + y_u^2 + y_d^2) + \frac{1}{12} (y_{\tau}^2 + y_{\mu}^2 + y_{e}^2)\,, \\ & c_\sigma = \frac13\lambda_\sigma + \frac16 \lambda_{\Phi \sigma}\,, \end{aligned} $$ with $g$ and $g'$ the EW gauge couplings and $y_i$ the Yukawa coupling of the SM particle $i$, for the case of scenario 3 the only relevant modification comes from the neutrino sector where $c_\sigma$ receives an additional contribution from the neutrino Yukawa couplings of the form $$ c_\sigma \rightarrow c_\sigma + \dfrac{1}{24} \sum_{i = 1}^6 Y_{\sigma_i}^2\,. $$ The longitudinal modes of the gauge bosons also receive thermal corrections which look like $$ \begin{aligned} && m_{W_L}^2(\phi_{h};T) = m_W^2(\phi_{h}) + \frac{11}{6}g^2T^2\,, \\ && m_{Z_L,A_L}^2(\phi_{h};T) = \frac{1}{2}m_Z^2(\phi_{h}) + \frac{11}{12}(g^2+{g'}^2)T^2 \pm {\cal D} \,, \end{aligned} $$ $$ \begin{aligned} && m_{W_L}^2(\phi_{h};T) = m_W^2(\phi_{h}) + \frac{11}{6}g^2T^2\,, \\ && m_{Z_L,A_L}^2(\phi_{h};T) = \frac{1}{2}m_Z^2(\phi_{h}) + \frac{11}{12}(g^2+{g'}^2)T^2 \pm {\cal D} \,, \end{aligned} $$ with $$ {\cal D}^2 = \Big(\frac{1}{2}m_Z^2(\phi_{h}) + \frac{11}{12}(g^2+{g'}^2)T^2 \Big)^2 - \frac{11}{12} g^2{g'}^2 T^2 \Big(\phi_h^2 + \frac{11}{3}T^2 \Big) \,. $$ The counterterm Lagrangian $V_{\rm ct}$ is given by $$ \begin{aligned} V_{\rm ct}&=\delta\mu_{\Phi}^{2} \Phi^{\dagger} \Phi+\delta\lambda_{\Phi}\left(\Phi^{\dagger} \Phi\right)^{2}+\delta\mu_{\sigma}^{2} \sigma^{\dagger} \sigma+\delta\lambda_{\sigma}\left(\sigma^{\dagger} \sigma\right)^{2} \\ &+\delta\lambda_{\Phi \sigma} \Phi^{\dagger} \Phi \sigma^{\dagger} \sigma+\left(\frac{1}{2} \delta\mu_{b}^{2} \sigma^{2}+\text {h.c.}\right)\, . \end{aligned} $$ Note that we only perform the renormalization of the potential parameters and leave the fields untouched. The counterterms are fixed by imposing that the Coleman-Weinberg potential and counterterm potential should not change the form of the minimum conditions and masses at zero temperature [10, 23] $$ \begin{aligned} \left\langle \frac{\partial V_{\rm ct}}{\partial h_i}\right\rangle = \left\langle-\frac{\partial V^{(1)}_{\rm CW}}{\partial h_i}\right\rangle\,, & & \left\langle \frac{\partial^2 V_{\rm ct}}{\partial h_i\partial h_j}\right\rangle = \left\langle- \frac{\partial^2 V^{(1)}_{\rm CW}}{\partial h_i\partial h_j}\right\rangle\,. \end{aligned} $$ With these conditions, the counterterms for the scalar singlet extension model in the conditions of scenario 1 where $\sigma$ has no VEV at zero temperature, are given by, $$ \begin{aligned} \delta\mu^2_\Phi&=-\frac{3}{2v_h}\frac{\partial V^{(1)}_{\rm CW}}{\partial h} +\frac{1}{2}\frac{\partial^2 V^{(1)}_{\rm CW}}{\partial h^2}\,, & \delta\lambda_\Phi &= \frac{1}{2v_h^3}\frac{\partial V^{(1)}_{\rm CW}}{\partial h}-\frac{1}{2v_h^2}\frac{\partial^2 V^{(1)}_{\rm CW}}{\partial h^2}\,, \\ \delta\mu_\sigma^2&=0\,, & \delta\lambda_\sigma&=0\,,\\ \delta\lambda_{\Phi\sigma}&=-\frac{2}{v_h^2}\frac{\partial^2 V^{(1)}_{\rm CW}}{\partial \sigma_R^2}\,, & \delta\mu^2_{b}&=0\,. \end{aligned} $$
210812810/12
of the type $(v_h^i, 0) \rightarrow (v_h^i, v_\sigma^f)$, where the superscript $i$ and $f$ mean initial and final 1. Clearly, most transitions prefer a final phase with a non-vanishing singlet VEV. Before proceeding to the presentation of the results, we need to discuss the calculation of the inverse time-scale of the phase transition in units of the Hubble parameter $H$, $\beta/H$. The value of $\beta/H$ is obtained from Eq. (35) and it is calculated numerically, resorting to the values of the action provided by the CosmoTransitions code. As can be seen from the expression, the calculation involves the derivative of the action, which means that $\hat{S}_3$ has to be a continuous function in the vicinity of the percolation temperature. This is not the case – the action obtained from CosmoTransitions is irregular and we have devised a method to smoothen the action before performing the derivative. This procedure is described in detail in appendix A together with the estimation of the error in the calculation of $\beta/H$. All results presented, unless otherwise stated, exclude points with an error above 25% (see appendix A for further details). Figure1: The peak-amplitude of the GW signal $h^2 \Omega^{\rm peak}_{\rm GW}$ as a function of the peak frequency $f_{\rm peak}$ in logarithmic scale. The colour bar indicates the strength of the phase transition $\alpha$ (left panel) and the inverse time-scale of the phase transition in units of the Hubble parameter $H$, $\beta/H$ (right panel). The PISCs for LISA BBO and DECIGO are represented with dashed, dot-dashed and dotted lines respectively. In Fig. 1 we present the GW signal $h^2 \Omega^{\rm peak}_{\rm GW}$ as a function of the peak frequency $f_{\rm peak}$ in logarithmic scale. The colour bar shown in the scatter plots represents the strength of the phase transition $\alpha$ (left panel) and the inverse time-scale of the phase transition in units of the Hubble parameter $H$, $\beta/H$ (right panel). From the figure it is clear that only values of $\alpha$ above about $0.1$ may lead to GW signals detectable in the near future 2. As for the inverse time-scale, the points within LISA reach are in the range $34 \leq \beta/H \leq 2257$. The grey curves in these and all remaining plots represent the peak integrated sensitivity curves (PISCs) for sound waves recently derived in Ref. [52]. In Fig. 2 we again present the GW signal $h^2 \Omega^{\rm peak}_{\rm GW}$ as a function of the peak frequency $f_{\rm peak}$ in logarithmic scale but now the colour scale represents $\Delta v_h/T_*$ (left panel) and $\Delta v_\sigma/T_*$ (right panel). As expected the points within reach have large values of $\Delta v_h/T_*$ and of $\Delta v_\sigma/T_*$. A 1 The first pair of values represent the VEVs before the phase transition and second pair are the values after the phase transition. The first term in the pair is the doublet VEV while the second is the real part of the singlet VEV (the imaginary component of the singlet has always zero VEV). 2 To be precise, we found points within LISA reach in the range $0.1\leq \alpha \leq 6.6$.
210812810/14
Figure4: Peak-amplitude of the GW signal $h^2 \Omega^{\rm peak}_{\rm GW}$ as a function of the peak frequency $f_{\rm peak}$ in logarithmic scale with $T_n - T_*$ in the colour bar and given in GeV. ## The dark sector Let us now discuss the impact of the parameters of the dark sector on possible detection of GWs originating from a FOPT. In Fig. 5 we present peak-amplitude of the GW signal as a function of the peak frequency with the portal coupling $\lambda_{\Phi \sigma}$ in the color bar. In the left plot we have set $\lambda_\sigma<1$ and in the right plots the points obey $\lambda_\Phi<1$. We see that there is no trend but it is clear that large values for the peak are obtained with all quartic couplings sufficiently small. Although we could not find in our scan points with all quartic couplings below 1, points with two Figure5: Peak-amplitude of the GW signal $h^2 \Omega^{\rm peak}_{\rm GW}$ as a function of the peak frequency $f_{\rm peak}$ in logarithmic scale with the portal coupling $\lambda_{\Phi \sigma}$ in the color bar. In the left plot we have set $\lambda_\sigma<1$ and in the right plots the point obey $\lambda_\Phi<1$. quartic couplings below 1 and the third one below 2 were common. This is important because as shown in [54] if all couplings are below 1 the model is stable up to Planck scale and if they are all below 2 the model is stable to slightly below or in the GUT scale. The stability study [54] was performed for the complex singlet extension of the SM using the full two-loop renormalization
210812810/24
is reduced to 192 316 and if we further restrict the error $\Delta (\beta/H) < 0.05$, the number of allowed points is reduced to 108 032, that is, only 50% of the points remain. In the plots presented in the paper, all points have $\Delta (\beta/H) < 0.25$. We did not want to further restrict the error because it could be that we were also losing too many good points. Figure14: The peak-amplitude for the GW signal $h^2 \Omega^{\rm peak}_{\rm GW}$ as a function of the peak frequency $f_{\rm peak}$ in logarithmic scale for scenario 2. The scatter plots present, in the colour bar, the strength of the phase transition $\alpha$. In the left plot, there are no restrictions related to the calculation of $\beta/H$, in the middle plot only points with $\Delta (\beta/H) < 0.25$ are accepted, and in the right panel only points with $\Delta (\beta/H) < 0.05$ are accepted. Our method has another benefit, the interpolation of the bounce action provides us with an approximate analytical expression for $\hat{S}_3/T$ which, in turn, gives us an approximate analytical expression for the tunnelling rate $\Gamma(T)$. This allows us to promptly calculate the nucleation and percolation temperatures and because we have 4 samples we can also estimate the error associated with our calculation of the characteristic temperatures.
210812810/2
## Introduction As often happens in physics, there is an apparently strange connection between particle physics and gravitational waves (GWs). Assuming that the Higgs potential at zero temperature is a result of a strong first order phase transition (FOPT) in the Higgs vacuum that occurred in the early Universe, signs of that transition could appear today in the form of primordial GWs and, under particular circumstances, could be detected in a not too distant future [1 – 4]. A strong EW FOPT is considered to be an important prerequisite for the generation of baryon asymmetry in the early Universe as one of the Sakharov conditions [5]. It is well known that the Higgs potential of the Standard Model (SM) cannot provide a FOPT and an extension, even minimal, is needed in order to include this new feature in the model. One of the most simple extensions of the Higgs potential, with the addition of a singlet field (with Isospin and Hypercharge zero), is sufficient to trigger a strong enough FOPT needed for EW baryogenesis [6]. The implications for the detection of GWs in the case of the singlet-extended Higgs potential were first discussed in Ref. [7] while an extension with an arbitrary number of singlets was studied in Ref. [8] (other simple extension like two-Higgs doublet models were discussed in [9 – 11]). These and other works study a connection between the values of the parameters of the scalar potential, the characteristics of the FOPTs (their duration and latent heat) and properties of the associated primordial GW spectrum such as its peak amplitude and frequency. Once the potential is fixed it is possible to search for the regions of the parameter space that would give rise to potentially detectable GW signatures. The relation between the parameters of the Higgs potential and the strength of the GW spectrum is prone to large instabilities that can be traced back to the multidimensional nature of the field content as well as to numerical instabilities in the calculation of the bounce action and its derivative. In fact, the phase transition happens at a very specific point of the parameter space and field configuration and it is possible that a point extremely close to that one will not undergo a phase transition. This fact leads to the following question: would this instability be also reflected on the sensitivity of the results to the SM parameters and if so to what extent? And there is one more question to be asked: how do the codes presently used deal with such instabilities? Hence, one of the goals of this paper is to understand the effect of the precision in the measurement of the SM parameters in the characteristics of the GW spectra. Another interesting point that we will address is if different realisations of the spontaneous symmetry breaking of a given model may lead to a significant difference in the strength of the GW signal. In fact, we may ask ourselves if when we study a specific model, say the complex singlet extension of the SM, and allow for two different patterns of the symmetry breaking, how disparate the corresponding GW spectra can be. Moreover, if the differences are indeed significant, how does the addition of new particles affect the GW spectrum? All these issues are discussed by analysing the FOPTs and their GW signatures in a simple extension of the SM featuring an additional complex scalar singlet in two different phases, as well as in the scope of the Majoron model with inverse seesaw mechanism described in Ref. [12]. The paper is organized as follows. In section 2 we briefly present the potential and the vacuum structure of the model and in section 3 we discuss the relation between GWs and FOPT. In section 4.1 we discuss the detection of GW in the case where the singlet acquires a vacuum expectation value (VEV) at zero temperature followed in section 5 by a discussion on the dependence of GW spectra on the SM parameters for the same model. We then move to a
210812810/23
Hence, we believe that our study delivers two clear messages. First, any such study needs to take into account the precision in the measurements of at least the Higgs mass and the top quark mass. Second, even the same model, if considered in different phases at zero temperature, can exhibit a very distinct behaviour in what concerns the detection of GWs originating from strong FOPTs. Finally, we have discussed technical issues in $\beta/H$ computations and how not taking proper care of the derivative of the action can lead to numerical instabilities and, hence, to wrong results. ## Smoothing the action The value of ${\beta}/{H}$ is given by Eq. (35). As can be seen from this expression, the calculation involves the derivative of the action with respect to the temperature. A possible method to calculate it numerically is to use the Difference Quotient Method (DQM) with the bounce action numerically computed by CosmoTransitions as $$ \frac{\beta}{H} = T_* \left. \frac{d}{d T} \left(\frac{\hat{S}_3(T)}{T}\right) \right|_{T_*} \approx T_* \frac{1}{2 \cdot\Delta T}\left(\left(\frac{\hat{S}_3(T)}{T}\right)\Bigg|_{T = T_*+\Delta T}-\left(\frac{\hat{S}_3(T)}{T}\right)\Bigg|_{T = T_*-\Delta T}\right)\,, $$ where $\Delta T$ is the small step of the DQM. This method correctly calculates ${\beta}/{H}$ for points with a strong GW signal, but for weaker points, numerical errors in the calculation of $\hat{S}_3$ make the DQM not completely reliable. Our solution is to interpolate the action around $T_n$ starting by sampling $N = \{60\,,75\,,90\,,105\}$ bounce actions inside the interval ranging from $\max\{T_n-30,T_n/2\} \text{GeV}$, which should leave enough room to calculate the percolation temperature, up to $T_c-3\text{GeV}$. We do not interpolate exactly up to $T_c$ to prevent numerical instabilities regarding the exis- tence/location of the minimum. Moreover, for the calculation of the percolation temperature, $T_*$, this truncation has a negligible effect since the biggest contribution comes from the epoch around $T_n$. The $4$ independent samples of $N$ actions are linearly distributed inside the mentioned interval, in each sample. If we calculate more than $7$ bounce actions then a degree 6 polynomial in $T$ can be fitted that models $\hat{S}_3/T$. After this procedure the calculation of the derivative is trivial. Applying this method to four independent samples of points allows us to calculate $\beta/H$ four times. We consider the most correct value for $\beta/H$ to be the average between all 4 samples. This allows us to estimate an error $\Delta\left(\beta/H\right)$ for our method which we define as the difference between the biggest $\beta/H$ and the smallest $\beta/H$ divided by two, $$ \begin{aligned} \Delta\left(\beta/H\right) = \frac{\text{max}\{\beta/H\}-\text{min}\{\beta/H\}}{2}\,. \end{aligned} $$ The next question to ask is: which points will we consider as valid when we use this method on the output of CosmoTransitions? In Fig. 14 we now present a scatter plot with the GW signal $h^2 \Omega^{\rm peak}_{\rm GW}$ as a function of the peak frequency $f^{\rm peak}$ for three different levels of constraints set upon the uncertainty of $\beta/H$. On the left panel, a total of 215 065 points are shown which are the ones with no restrictions on $\Delta (\beta/H)$, that is, the original set of points before the smoothing procedure is applied. Once we impose that $\Delta (\beta/H) < 0.25$ (middle panel) the number of points 3 CosmoTransitions is not always able to calculate all of the $N$ bounce actions. When this happens we do the fit with whichever points it managed to calculate.
210812810/16
and as a function of $v_\sigma$ (right). Again no particular pattern emerges although a slight preference for negative values of theta exists. The values of $v_\sigma$ that allow for a strong FOPF are in the range $50.8 \text{GeV} < v_\sigma < 860.1 \text{GeV}$. ## The dependence on the SM parameters ## Scenario 2 In the studies presented in the literature connecting FOPT with the detection of primordial GWs, the role of the SM parameters is never discussed (to the best of our knowledge). So suppose that a GW is detected and that it points to a given class of models. One should then ask: what happens if we vary each of the SM parameters within the experimentally determined error? Will it lead to significant changes in the characteristics of the GW or is it negligible? These are the questions we will answer in this section taking as benchmark scenario 2. Note that this scenario is not only one of the simplest extensions of the SM but it also features couplings of the Higgs boson to the SM particles that are all modified by a common factor $\cos \theta$ that is getting closer and closer to unity. Moreover, as we have seen in the previous section, this is a model that leads to GW signals potentially detectable in not-so-distant future. Scenario 3 will be the subject of the next section. In order to understand the impact of the variation of the SM parameters within their exper- imental errors we have first chosen points that are within LISA reach taking all SM parameters with their central values according to the PDG review [55]. We then varied each of the fermion masses, from the electron to the top quark, the $W$ and $Z$ bosons’ masses and the Higgs mass. The variation of the masses of the SM particles and the calculation of the error in the peak of the GW power spectrum are described in detail in appendix A. We concluded that, provided we would use a smoothed action, the only SM masses capable of inducing a significant shift in the GW peak amplitude and frequency are the top quark mass and even more so the Higgs boson mass. In Fig. 8 we present five points that could be probed by LISA in the scope of scenario 2, identified by a red circle. In the left column we show the variation of the peak amplitude with the variation of the Higgs mass performed from 124.96 GeV to 125.24 GeV, corresponding to an uncertainty in the Higgs mass of $\pm 1$ standard deviation, while keeping the remaining parameters of the SM constant together with those from the dark sector. From all the points above LISA the maximum variation found for the peak amplitude was 250% for the variation with the Higgs mass. The message is very clear and quite striking: the variation of only the Higgs mass within its experimental uncertainty leads to a variation of the peak amplitude (and also in the peak frequency) of up to 250% for the sample obtained. This could mean the following: starting with a central value giving rise to a GW signal detectable by LISA, one could move outside of the LISA sensitivity range just by varying the Higgs boson mass (within the current experimental uncertainty). Hence, the dependence on the Higgs mass is indeed meaningful and must be considered in this type of numerical studies. The other SM parameter that can shift the GW peak amplitude and frequency is the top quark mass (middle panel in Fig. 8). The red circled points again show the central values. We have varied the top quark mass between 172.46 GeV and 173.06 GeV, which corresponds to an
210812810/11
The bubble wall velocity has to be rather large to give rise to detectable GWs spectra although it is quite challenging to provide a precise estimate for it [43, 44]. Our analysis is performed using CosmoTransitions [37], considering the case of supersonic detonations in or- der to maximize the GW peak amplitude where the wall velocity $v_b$ is taken to be above the Chapman-Jouguet limit, $$ v_\mathrm{J} = \dfrac{1}{1+\alpha} \left(c_s + \sqrt{\alpha^2 + \tfrac{2}{3} \alpha}\right)\,. $$ For certain parameter configurations, one also expects sequential phase transition patterns potentially leading to multi-peak GWs spectra [45 – 48]. ## Analysis of Scenario 2 The first goal of this study is to understand in which of the three scenarios a strong FOPT leading to the detections of gravitational waves in future experiments is realisable. We have concluded that only in scenarios 2 and 3 do one finds GW that can be probed by those experiments. Since scenario 3 was already discussed in [12] we will now focus on scenario 2. We will come back to scenario 3 when discussing the variation of the GW peak and frequency with the SM parameters. ## GWs detection In Table 1 we show the ranges of the input parameters in the scans for scenario 2. The range for $\mu_{b}^2$ reflects a variation of the mass of the DM candidate between 50 GeV and 1 TeV. The range of variation of $\theta$ takes into account the LHC Higgs couplings measurements that force the 125 GeV Higgs, dubbed as $h_1$, to be very SM-like with $\cos \theta > 0.85$. We will also show in the plots \begin{tabular}{@{}rcccr@{}}\rightarrowprule & \multicolumn{3}{c}{Scenario 2} \cmidrule{2-4} & Parameter & Range & Distribution &\midrule & $m_{h_2}$ & $[50,\,1000]\,\text{GeV}$ & linear& $v_{\sigma}$ & $[50,\,1000]\,\text{GeV}$ & linear && $\mu_{b}^2$ & $[-500000,\,-1250]\,\text{GeV}^2$ & linear && $\theta$ & $[-\arccos{(0.85)},\, \arccos{(0.85)}]$ & linear &\bottomrule \end{tabular} Table 1: Ranges of the input parameters in the scans for scenario 2. the reach predicted by LISA, by the Deci-Hertz Interferometer Gravitational Wave Observatory (DECIGO) [49] and by the Big Bang Observer (BBO) [50] proposed as a follow-on mission to LISA. There is also a more recent proposal under discussion, the TianQin Observatory [51]. Having set the stage for the calculation of the GW power spectrum we will now examine which parameters play an important role in the detection of GWs in the near future. Let us start by noting that from all the possible 16 phase transition patterns we found that in the considered scenario the ones that are above the LISA line are 148 of the type $(0,0) \rightarrow (0,v_\sigma^f)$, 29 $(v_h^i,0) \rightarrow (v_h^f, v_\sigma^f)$, 14 $(0,v_\sigma^i) \rightarrow (v_h^f, v_\sigma^f)$, 11 $(0,v_\sigma^i) \rightarrow (0, v_\sigma^f)$, 2 $(0,v_\sigma^i) \rightarrow (v_h^f, v_\sigma^i)$, 1 $(0, 0) \rightarrow (v_h^f, 0)$ and 1
210812810/6
One-loop thermal corrections are given by [16] $$ \Delta V(T) = \frac{T^4}{2 \pi^2} \left\{\sum_{b} n_b J_B\left[\frac{m_i^2(\phi_\alpha)}{T^2}\right] - \sum_{f} n_f J_F\left[\frac{m_i^2(\phi_\alpha)}{T^2}\right] \right\}\,, $$ where $J_B$ and $J_F$ are the thermal integrals for bosons and fermions, respectively, provided by $$ \begin{aligned} J_{B/F}(y^2) = \int_0^\infty d x \, x^2 \log\left(1 \mp \exp [- \sqrt{x^2 + y^2}] \right)\,. \end{aligned} $$ For the first non-trivial order of the thermal expansion $\sim(m/T)^2$, $\Delta V(T)$ can be approximated as $$ \begin{aligned} \Delta V^{(1)}(T)|_{\rm L.O.} = \frac{T^2}{24} \left\{{\rm Tr}\left[M_{\alpha\beta}^2(\phi_\alpha) \right] + \sum_{i=W,Z,\gamma} n_i m_i^2(\phi_\alpha) + \sum_{i = \mathrm{f}_i} \frac{n_i}{2} m_i^2(\phi_\alpha) \right\} \,, \end{aligned} $$ where in the last sum we consider all the fermions in the considered models consisting of three generations of quarks and charged leptons for scenarios 1 and 2, as well as six heavy neutrinos in what concerns scenario 3. The first term in Eq. (13) denotes the trace of the field- dependent scalar Hessian matrix $M_{\alpha\beta}^2(\phi_\alpha)$, which is a basis invariant quantity. In the case of the discussed models we have used Eqs. (6) and (7) upon replacing the VEVs by their classical field configurations $v_h \rightarrow \phi_h$ and $v_\sigma \rightarrow \phi_\sigma$. This means that the leading thermal corrections only affect the quadratic terms (in mean-fields) of the scalar potential, preserving the shape of $V_{0}$ and affecting only the masses of the scalar fields. The $n_i$ coefficients in Eq. (13) represent the number of d.o.f for a given particle, as indicated by the sums. In particular, for the SM gauge bosons ($W, Z$ and transversely polarised photon $\gamma$) we have $$ n_W = 6, \qquad n_Z = 3, \qquad n_\gamma = 2 \,, $$ whereas for scalars and the longitudinally polarized photon $(A_L)$ we have $$ n_s = 6, \qquad n_{A_L} = 1\,, $$ while for fermions $$ n_{u,d,c,s,t,b} = 12, \qquad n_{e,\mu,\tau} = 4\,, \qquad n_{N_{1,\ldots,6}} = 2 \,. $$ with $N_{1,\ldots,6}$ denoting the six physical heavy neutrinos in scenario 3. The presence of $T^2$ terms in the thermal expansion suggests the possibility for symmetry restoration at high temperatures. Furthermore, it typically implies the breakdown of perturba- tion theory in a close vicinity of the critical temperature. This must be addressed by means of an all-order resummation procedure via the addition of the so called daisy or ring diagrams [19 – 22]. In practice, this is done by a correction to the potential mass terms $$ \mu_\alpha^2(T) = \mu_\alpha^2 + c_\alpha T^2 \,, $$ where the $c_\alpha$ coefficients can be calculated from Eq. (13) as follows $$ c_\alpha = \dfrac{\delta^2 {\Delta V^{(1)}(T,\phi_h, \phi_\sigma)}|_{\rm L.O.}}{\delta \phi_\alpha^2}\,. $$
210812810/4
- zero temperatures, featured a temperature dependent mixing with the neutral component from the doublet, vanishing as $T\rightarrow 0$. The $U(1)_L \rightarrow \mathbb{Z}_2$ soft breaking term in Eq. (1) provides a pseudo-Goldstone mass to the imaginary part of the EW singlet field. - Scenario 2 – in this case both the doublet and the real component of the gauge singlet acquire VEVs at $T=0$, that is, $\phi_{h,\sigma}(T=0)\equiv v_{h,\sigma}$. One of the CP-even scalar states is identified with the SM-like Higgs boson with a mass of $125$ GeV. The second scalar, that mixes with the SM-like Higgs boson, can be either heavier or lighter than the 125 GeV Higgs boson candidate in this case. The soft breaking term in the potential explicitly breaks $U(1)_L \rightarrow \mathbb{Z}_2$ providing a pseudo-Goldstone mass to the imaginary part of the field $\sigma_I$ known in many contexts as a Majoron. - Scenario 3 – From the point of view of this work, Scenario 3 can be seen as an extension of Scenario 2. The scalar potential is exactly the same but right-handed neutrinos are introduced in the context of an inverse seesaw mechanism. The details of the Majoron model as well as its constraints are discussed in Refs. [12 – 15]. Let us now describe the first two scenarios in more detail. We first note that such scenarios are just two different phases of the same potential at zero temperature. This means that there are conditions that are the same in both cases. The conditions for the potential to be bounded from below read $$ \lambda_\Phi > 0, \quad \lambda_\sigma > 0, \quad \lambda_{\Phi \sigma} > - 2 \sqrt{\lambda_\Phi \lambda_\sigma} \,, $$ and will be imposed in our numerical calculations. Also, in our analysis we impose a conservative perturbativity bound on the quartic couplings, $\lambda_{\Phi \sigma}, \lambda_\sigma < 2 \pi$. In scenario 1, the mass spectrum at zero temperature is just the one of the SM with two new dark scalars. The SM-like Higgs boson has a mass, $m_h\simeq 125$ GeV and emerges entirely from the doublet. This in turn also means that the Higgs couplings to the remaining SM particles are not modified. The DM candidates only couple to the Higgs boson via the portal coupling $\lambda_{\Phi \sigma}$ which can be constrained by measurements of the invisible Higgs decay as well as by direct and indirect DM detection data. In the scalar sector the mass spectrum is given by $$ \begin{aligned} && m_{h}^2 = 2\lambda_\Phi v_h^2 \,, \quad m_{D1}^2 = \mu_\sigma^2 + \mu_b^2 + \frac{\lambda_{\Phi \sigma} v_h^2}{2}\,, \quad m_{D2}^2 = \mu_\sigma^2 - \mu_b^2 + \frac{\lambda_{\Phi \sigma} v_h^2}{2} \,, \end{aligned} $$ for the SM Higgs boson and for the two DM candidates, $D1$ and $D2$, respectively. Which of these two is the stable one at zero temperature depends on sign of $\mu_b^2$ parameter. Indeed, $$ m_{D2}^2 - m_{D1}^2= - 2 \mu_b^2 \,, $$ which means that $D2$ is the DM particle if $\mu_b^2 > 0$, while $D1$ is the DM candidate if $\mu_b^2 < 0$. In practice, it may be possible to consider also a scenario with one stable and one metastable DM candidates. In scenario 2, the main difference is in the particle spectrum. Now the CP-even component of the singlet mixes with the CP-even component from the doublet and only one DM candidate remains. The masses of the CP-even states can be written as $$ \begin{aligned} m_{h_1,h_2}^2=\lambda_\Phi v_h^2 + \lambda_\sigma v_\sigma^2 \mp \frac{\lambda_\sigma v_\sigma^2 - \lambda_\Phi v_h^2}{\cos 2\theta} \,, \end{aligned} $$
210812810/15
group equations. Therefore we believe the values shown for quartic couplings should be stable up to somewhere between the GUT and the Planck scale. Moreover, it is possible that with more time of running we would find points with all quartic couplings below 1. In Fig. 6 we present the peak-amplitude of the GW signal as a function of the peak frequency in logarithmic scale. The left panel of the scatter plot shows the behaviour with the DM mass, $m_{D}$, while the right panel refers to the second Higgs mass $m_{h_2}$. Remember that the range of variation chosen for both masses is between 50 GeV and 1 TeV. The figure shows no particularly interesting pattern with the values for the parameters within the initial chosen range. The allowed values of the parameters for a strong FOPT with points above LISA are $52.0 \text{GeV} < m_{h_2} < 997.2 \text{GeV}$ and $68.2 \text{GeV} < m_{D} < 999.8 \text{GeV}$. Figure6: The peak-amplitude of the GW signal $h^2 \Omega^{\rm peak}_GW$ as a function of the peak frequency $f_{\rm peak}$ in logarithmic scale. The scatter plots present, in the colour bar, the dark matter mass $m_D$ (left panel) and non-SM Higgs $m_{h_2}$ (right panel). Masses are expressed in GeV. Figure 7: Same as Fig. 6 but now as a function of the mixing angle $\theta$ (left) and as a function of $v_\sigma$ (right). In Fig. 7 we show the same set of points but now as a function of the mixing angle $\theta$ (left)
210812810/10
due to the fact that the actual phase transition starts at $T_n < T_c$, a temperature for which the bubble nucleation rate exceeds that of the cosmological expansion, and finishes effectively at $T_\ast < T_n$. Nevertheless, this condition does not necessarily lead to the generation of strong and potentially observable GWs. A sizeable GW signal and small $\beta/H$ needs a large bubble wall velocity $v_b$ and a substantial latent heat release which is related to $\alpha$. In our analysis, we consider only GWs originating from sound shock waves (SW) which are generated by the bubble’s violent expansion in the early Universe. According to the discussion in Ref. [34] their contribution dominates the peak frequency and the peak amplitude in the primordial GW spectrum. Furthermore, bubble wall collision does not give a meaningful con- tribution to GWs as discussed in Refs. [30, 38] while magnetohydrodynamic turbulence of the early Universe plasma is usually not accounted for due to large theoretical uncertainties [34]. The primordial GW signals produced in such violent out-of-equilibrium cosmological pro- cesses as the FOPTs are redshifted by the cosmological expansion and look today as a cosmic gravitational stochastic background. The corresponding power spectrum [31, 32, 39 – 41] $$ h^2 \Omega_{\rm GW}(f) \equiv \frac{h^2}{\rho_c} \frac{\partial \rho_{\rm GW}}{\partial \log f}\,, $$ where $\rho_c$ is the critical energy density today, can be found for various GW frequencies $f$ by multiplying the peak amplitude $h^2 \Omega_\mathrm{GW}^\mathrm{peak}$ by the spectral function and reads $$ h^2 \Omega_\mathrm{GW} = h^2 \Omega_\mathrm{GW}^\mathrm{peak} \left(\dfrac{4}{7}\right)^{-\tfrac{7}{2}} \left(\dfrac{f}{f_\mathrm{peak}}\right)^3 \left[1 + \dfrac{3}{4} \left(\dfrac{f}{f_\mathrm{peak}}\right) \right]^{-\tfrac{7}{2}}\,, $$ where $f_\mathrm{peak}$ is the peak-frequency. Semi-analytic expressions for peak-amplitude and peak- frequency in terms of $\beta/H$ and $\alpha$ can be found in Ref. [34] and can be summarised as follows $$ \begin{aligned} && f_\mathrm{peak} = 26 \times 10^{-6} \left(\dfrac{1}{H R} \right) \left(\dfrac{T_\mathrm{n}}{100} \right) \left(\dfrac{g_\ast}{100~\mathrm{GeV}} \right)^{\tfrac{1}{6}} \mathrm{Hz} \,, \,, \\ && h^2 \Omega_\mathrm{GW}^\mathrm{peak} = 1.159 \times 10^{-7} \left(\dfrac{100}{g_\ast}\right) \left(\dfrac{HR}{\sqrt{c_s}}\right)^2 K^{\tfrac{3}{2}} \qquad \rm{for} \qquad H \tau_\mathrm{sh} = \dfrac{2}{\sqrt{3}} \dfrac{HR}{K^{1/2}} < 1 \,, \\ && h^2 \Omega_\mathrm{GW}^\mathrm{peak} = 1.159 \times 10^{-7} \left(\dfrac{100}{g_\ast}\right) \left(\dfrac{HR}{c_s}\right)^2 K^{2} \qquad \rm{for} \qquad H \tau_\mathrm{sh} = \dfrac{2}{\sqrt{3}} \dfrac{HR}{K^{1/2}} \simeq 1 \,, \end{aligned} $$ $$ \begin{aligned} && f_\mathrm{peak} = 26 \times 10^{-6} \left(\dfrac{1}{H R} \right) \left(\dfrac{T_\mathrm{n}}{100} \right) \left(\dfrac{g_\ast}{100~\mathrm{GeV}} \right)^{\tfrac{1}{6}} \mathrm{Hz} \,, \,, \\ && h^2 \Omega_\mathrm{GW}^\mathrm{peak} = 1.159 \times 10^{-7} \left(\dfrac{100}{g_\ast}\right) \left(\dfrac{HR}{\sqrt{c_s}}\right)^2 K^{\tfrac{3}{2}} \qquad \rm{for} \qquad H \tau_\mathrm{sh} = \dfrac{2}{\sqrt{3}} \dfrac{HR}{K^{1/2}} < 1 \,, \\ && h^2 \Omega_\mathrm{GW}^\mathrm{peak} = 1.159 \times 10^{-7} \left(\dfrac{100}{g_\ast}\right) \left(\dfrac{HR}{c_s}\right)^2 K^{2} \qquad \rm{for} \qquad H \tau_\mathrm{sh} = \dfrac{2}{\sqrt{3}} \dfrac{HR}{K^{1/2}} \simeq 1 \,, \end{aligned} $$ $$ \begin{aligned} && f_\mathrm{peak} = 26 \times 10^{-6} \left(\dfrac{1}{H R} \right) \left(\dfrac{T_\mathrm{n}}{100} \right) \left(\dfrac{g_\ast}{100~\mathrm{GeV}} \right)^{\tfrac{1}{6}} \mathrm{Hz} \,, \,, \\ && h^2 \Omega_\mathrm{GW}^\mathrm{peak} = 1.159 \times 10^{-7} \left(\dfrac{100}{g_\ast}\right) \left(\dfrac{HR}{\sqrt{c_s}}\right)^2 K^{\tfrac{3}{2}} \qquad \rm{for} \qquad H \tau_\mathrm{sh} = \dfrac{2}{\sqrt{3}} \dfrac{HR}{K^{1/2}} < 1 \,, \\ && h^2 \Omega_\mathrm{GW}^\mathrm{peak} = 1.159 \times 10^{-7} \left(\dfrac{100}{g_\ast}\right) \left(\dfrac{HR}{c_s}\right)^2 K^{2} \qquad \rm{for} \qquad H \tau_\mathrm{sh} = \dfrac{2}{\sqrt{3}} \dfrac{HR}{K^{1/2}} \simeq 1 \,, \end{aligned} $$ where $\tau_\mathrm{sh}$ is the fluid turnover time or the shock formation time, which quantifies the time the GW source was active. In these expressions, $c_s = 1/\sqrt{3}$ is the speed of sound, $R$ is the mean bubble separation, $$ K = \dfrac{\kappa \alpha}{1 + \alpha} $$ is the fraction of the kinetic energy in the fluid to the total bubble energy, and $$ H R = \dfrac{H}{\beta} \left(8 \pi \right)^{\tfrac{1}{3}} \max\left(v_b, c_s\right) \,. $$ where $\kappa$ is the efficiency factor that can be found in Ref. [42].
210812810/13
Figure2: The peak-amplitude of the GW signal $h^2 \Omega^{\rm peak}_{\rm GW}$ as a function of the peak frequency $f_{\rm peak}$ in logarithmic scale. The scatter plots present, in the colour bar, $\Delta v_h/T_*$ (left panel) and $\Delta v_\sigma/T_*$ (right panel). clearer picture is obtained when we plot these two variables in the same plot. This is done in Fig. 3 and it shows that it is enough to have a large variation for the singlet, $\Delta v_\sigma/T_*$, to generate observable GW signals. Figure 3: Scatter plot showing $\Delta v_h/T_*$ vs. $\Delta v_\sigma/T_*$ with the strength of the GW signal $h^2 \Omega^{\rm peak}_{\rm GW}$ given by the colour bar. Using our polynomial interpolation of the action and plugging it into Eq. (32), we have observed that the points above LISA have a difference that is below 7 GeV and can even be close to zero in some cases. This can be seen in Fig. 4 where we display the peak-amplitude of the GW signal as a function of the peak frequency with $T_n - T_*$ in the colour bar. It is clear that there is no special trend regarding the difference between the two temperatures in relation to the peak except that above a certain temperature difference there are no points above LISA in our numerical simulations. Note that some of our points, in particular those with larger temperature differences, tend to agree with the calssification in [53] with respect to supercooling scenarios where we find $\alpha \gtrsim 0.1$ within the LISA sensitivity range.
210812810/21
Figure 11: Scatter plots showing the typical strength of the phase transitions $\Delta v_{h}/T_*$ in the colour scale of each plot. On the left panel the position of the GW peak is shown whereas on the right we show the corresponding SNR for a mission profile of three years. The colored lines show the SNR that depends on $T_{*}$, $g_{\ast}(T_{*})$ and $v_{\text{b}}$. The dotted curves are contour lines representing the shock formation time $\tau_{\text{sh}}$ as defined in Eq. (42). The grey shaded region corresponds to an acoustic period lasting longer than a Hubble time and it is where the sound waves treatment is mostly reliable [30, 35]. Figure 12: Scatter plots showing the typical strength of the phase transitions $\Delta v_{h}/T_*$ and SNR for scenario 2. particular, we have found 94 points with SNR larger than 5 out of which 45 feature a SNR above 100. Finally in Fig. 13 we show similar plots but for scenario 3. We found that there was no major difference with the addition of right-handed Majorana neutrinos to the broken phase. We again find 8 points with SNR larger than 5 and two points withs SNR above 100. The sample is 42 509. Further studies of Majoron models in the context of primordial GWs were performed in Ref. [12] and, more recently, in Ref. [60].