Joe Scanlin
Draft - November 2025
The next generation of AI systems will not learn primarily from static, human-authored datasets, but from ongoing interaction with the physical world. Silver and Sutton call this emerging regime the Era of Experience: agents learning predominantly from their own observations, actions, and consequences over time. Meanwhile, the physical world is becoming densely instrumented. Estimates suggest tens of billions of connected IoT devices by 2025, with some forecasts exceeding 70-75 billion devices when broader definitions are included. In parallel, the market for context-aware computing (systems that tailor behavior based on situational information) is projected to grow from tens of billions of dollars in 2024 to well over $200B by the early 2030s.
These two trends collide: we will soon have an unprecedented amount of sensor experience available, but our programming tools for spatial, multi-sensor, real-time environments (and for feeding them into agents) are still primitive.
Spaxiom is an attempt to fill this gap:
We present Spaxiom:
We argue that making spatial and sensor semantics first-class in a programming language, and explicitly optimizing for agentic workflows, can materially reduce token usage, energy consumption, and integration complexity, while opening up new kinds of AI behavior that are difficult to achieve with today's ad-hoc stacks.
The physical world is rapidly becoming a dense sensor mesh. IoT analyses report on the order of tens of billions of connected devices in 2024-2025, with projections ranging from ~39 billion to over 70 billion devices by 2030-2035 depending on methodology and scope. At the same time, context-aware computing (systems that adapt to who you are, where you are, and what is happening) is evolving from a niche into a mainstream infrastructure category, with forecasts showing high double-digit CAGR through the early 2030s.
The result is a looming experience deluge:
Yet much of this data is currently:
Silver and Sutton's "Welcome to the Era of Experience" describes a shift from training AI on human-curated datasets (text, images, labels) toward agents that learn predominantly from their own experience: ongoing interaction with the world, collecting trajectories of observations, actions, and rewards.
They argue that:
Most of the work in that direction has focused on simulated environments or camera-centric setups. But the world is full of other sensors: pressure, contact, force, vibration, chemical, thermal, RF, occupancy, and more. Many of these are privacy-preserving and cheap at scale, making them ideal substrates for real-world experience.
What is missing is a way to:
This is where Spaxiom slots in.
Spaxiom is currently implemented as an embedded DSL in Python, backed by a runtime and a growing documentation set. Concretely, Spaxiom provides:
Zone, coordinate systems, grid representations.Sensor abstraction plus adaptors for floor grids, MQTT streams, GPIO, and simulated sensors.Condition, within, always, eventually, etc.@on(condition) decorators for registering callbacks.Quantity with unit support.from spaxiom import Sensor, Zone, Condition, on, within
office_zone = Zone(0, 0, 10, 10)
motion_sensor = Sensor("motion1", "motion", (5, 5, 0))
motion_detected = Condition(lambda: motion_sensor.read() > 0.5)
sustained_motion = within(5.0, motion_detected)
@on(sustained_motion)
def alert_sustained_motion():
print("Motion has been detected for 5 seconds!")
Instead of pushing raw time series into downstream systems, the user expresses what they care about (e.g., "sustained motion in this region for >5 seconds"), and Spaxiom takes responsibility for timing, buffering, and triggering.
On top of the core DSL, we define INTENT (Intelligent Network for Temporal & Embodied Neuro-symbolic Tasks), a pattern library that captures common behaviors over sensors.
INTENT sits between the low-level DSL primitives (sensors, conditions, zones from Section 2.1) and high-level agent reasoning. While the DSL provides the building blocks for expressing spatiotemporal logic, INTENT patterns encapsulate recurring behaviors as reusable abstractions. For example, detecting "a queue is forming" involves sensor fusion (camera + pressure mat), temporal filtering (queue must persist >30 seconds), and queueing theory (estimating wait times from arrival/service rates). Rather than re-implementing this logic at each deployment, an INTENT.QueueFlow pattern captures it once and exposes an agent-ready event stream.
INTENT patterns serve three critical purposes:
OccupancyChanged or QueueFormed event saves 100-1000× tokens when feeding context to LLMs (Section 3).QueueFlow, ADL recognition heuristics for elder care, facilities management logic for FmSteward) that would otherwise need to be re-discovered at each site.The INTENT library includes a growing collection of domain-specific patterns. Four foundational patterns demonstrate the breadth of applicability:
OccupancyField: Tracks spatial occupancy and crowding over floor grids using pressure mats, depth cameras, or thermal sensors. Provides hotspot detection (top-k busiest tiles), density heatmaps (% occupied over time), and flow analysis (entries/exits per zone). Typical applications: lobby management, retail heat-mapping, emergency evacuation planning, stadium crowd control.QueueFlow: Models queue dynamics with real-time arrival and service rate estimation. Uses M/M/k queueing theory to predict wait times, detect bottlenecks, and trigger staffing alerts when queues exceed thresholds. Emits events like QueueFormed, QueueLengthChanged, and LongWaitDetected (when predicted wait > threshold). Typical applications: checkout optimization, airport security lines, call center staffing, hospital triage.ADLTracker: Recognizes activities of daily living (ADLs) from multi-modal sensors: floor pressure (walking, standing, lying down), door sensors (room transitions), appliance usage (refrigerator, stove, bathroom fixtures). Detects anomalies in routine (missed meals, unusual sleep patterns, prolonged inactivity, repeated bathroom visits) and emits alerts for caregiver intervention. Typical applications: elder care monitoring (aging-in-place), hospital patient observation, rehabilitation progress tracking, wellness scoring.FmSteward: Aggregates facilities maintenance signals into "needs service" events. Monitors usage counters (toilet flushes, paper towel dispensers, occupancy duration), environmental conditions (air quality, humidity, temperature), and time-since-last-service to trigger maintenance requests. Emits NeedsService(zone="restroom_3", reason="high_usage", urgency="medium"). Typical applications: restroom cleanliness monitoring, supply restocking, HVAC filter replacement, proactive janitorial dispatch.Additional domain-specific patterns (detailed in appendix use cases and Section 2.4) include:
SafetyMonitor - robot collision avoidance (Section 7.3)ConferenceRoomUtilization - meeting room booking efficiencySmartBuildingAgent - multi-pattern HVAC orchestrationContaminationMonitor - cleanroom particle/pressure tracking (Appendix A.1)MachineryHealthMonitor - vibration/acoustic anomaly detection (A.2)IAQRiskMonitor - indoor air quality and ventilation (A.3)MicroMobilitySafety - near-miss detection via radar (A.4)ColdChainMonitor - pharmaceutical temperature/humidity integrity (A.5)WildfireRiskMonitor - multi-sensor fire danger index (A.6)DataCenterThermalOptimizer - PUE and cooling efficiency (A.7)IrrigationOptimizer - crop water stress and soil moisture (A.8)ORSterilityMonitor - operating room sterility assurance (A.9)HumanoidTaskCoordinator - robot context streaming (A.10)CategoryAggregator - retail/expo engagement (Section 11.1)TrafficFlowMonitor - logistics corridor throughput (Section 11.3)QsrFlowOptimizer - quick-service restaurant throughput (Section 11.5)This diversity demonstrates that INTENT is not a fixed set of patterns, but an extensible framework for capturing domain expertise across industries: from healthcare to manufacturing, retail to agriculture, smart buildings to autonomous systems.
The following code demonstrates how an OccupancyField pattern compresses floor sensor data into crowding metrics and feeds them to an LLM agent for decision-making:
from spaxiom.config import load_yaml
from spaxiom.runtime import start_runtime
from spaxiom.temporal import within
from spaxiom.logic import on
from spaxiom.intent import OccupancyField
import asyncio, json, os, openai
sensors = load_yaml("examples/lobby.yaml")
floor = sensors["lobby_floor"]
field = OccupancyField(floor, name="lobby")
crowded = within(180.0, field.percent_above(10.0)) # >= 10% tiles active, 3 minutes
@on(crowded)
async def lobby_agent():
facts = {
"zone": field.name,
"occupancy_pct": field.percent(),
"hotspots": field.hotspots(top_k=3),
}
prompt = (
"You are a smart-building lobby agent. "
"Given this JSON describing current crowding, "
"suggest 1--3 actions to improve flow and experience. "
"Respond as JSON.\n"
+ json.dumps(facts)
)
openai.api_key = os.getenv("OPENAI_API_KEY", "")
if not openai.api_key:
return
rsp = await openai.ChatCompletion.acreate(
model="gpt-4o",
messages=[{"role": "user", "content": prompt}],
)
print("INTENT actions:", rsp.choices[0].message.content)
if __name__ == "__main__":
asyncio.run(start_runtime())
In this example, the OccupancyField pattern continuously monitors a floor sensor grid. When crowding persists for 3 minutes (within(180.0, ...)), the pattern's percent() and hotspots() methods provide agent-ready summaries: no raw sensor arrays, no manual aggregation. The LLM receives a compact JSON payload (~200 bytes) instead of 100s of kilobytes of raw sensor timeseries.
Extensibility: Users can define custom patterns by subclassing the Pattern base class (detailed in Section 2.4). This allows organizations to capture domain-specific knowledge (manufacturing process signatures, hospital workflow patterns, retail merchandising behaviors) as reusable INTENT abstractions that can be versioned, tested, and shared across deployments.
This pattern (sensor → DSL → INTENT → agent) is central to the rest of the paper. Section 2.4 provides the full pattern interface architecture, lifecycle methods, performance benchmarks, and guidelines for creating custom patterns.
Spaxiom's expressiveness comes from a small set of composable primitives with a well-defined type system. Unlike ad-hoc sensor integration scripts, Spaxiom enforces type safety at DSL construction time and provides algebraic composition operators that enable complex spatiotemporal queries to be built from simple building blocks.
The DSL defines the following fundamental types:
Sensor: base abstraction for any physical or virtual sensor. Key methods:
read() → Value: synchronous read of current sensor stateposition → (x, y, z): 3D location in spacezone → Zone: containing spatial regionunits → Unit: physical units (e.g., meters, celsius, pascal)Condition: boolean-valued predicate over sensor state and time. Conditions are lazy: they are not evaluated until subscribed or explicitly triggered. Key properties:
holds() → bool: evaluate condition at current timesince_true → float: seconds since condition became true (or 0.0 if false)since_false → float: seconds since condition became false (or 0.0 if true)Zone: 3D spatial region with geometric operations. Supports axis-aligned boxes, convex polygons, and union/intersection:
contains(x, y, z) → booloverlaps(other_zone) → booldistance_to(x, y, z) → floatvolume → floatEntity: tracked object with identity and attributes. Examples: a person, robot, forklift, or pallet. Entities have:
id → str: unique identifierposition → (x, y, z): current location (if trackable)velocity → (vx, vy, vz): optional velocity vectorattributes → dict: arbitrary key-value metadataQuantity: numeric value with physical units. Enforces dimensional analysis:
value → floatunit → UnitConditions form a boolean algebra with standard operators:
from spaxiom import Condition
# Logical operators
c1 = Condition(lambda: temp.read() > 25.0)
c2 = Condition(lambda: humidity.read() > 60.0)
hot_and_humid = c1 & c2 # conjunction (AND)
hot_or_humid = c1 | c2 # disjunction (OR)
not_hot = ~c1 # negation (NOT)
# Temporal chaining
humid_then_hot = c2.before(c1, within_seconds=300) # c2 before c1, within 5 min
Importantly, these operators are not evaluated eagerly. They construct a lazy evaluation graph that the runtime optimizes.
Spaxiom provides temporal logic operators inspired by Linear Temporal Logic (LTL) but adapted for continuous real-time systems:
within(duration, condition): condition holds continuously for at least duration seconds.
sustained_motion = within(5.0, motion_detected) # motion for ≥5s
always(duration, condition): condition holds at all sampled times over the past duration seconds (similar to within but checks discrete samples).
stable_temp = always(60.0, temp_in_range) # stable for 1 minute
eventually(duration, condition): condition becomes true at least once within the next duration seconds (forward-looking, requires buffering or prediction).
# Predictive: will door open in next 10s?
door_will_open = eventually(10.0, door_sensor_active)
before(c1, c2, within_seconds=T): condition c1 occurs before c2 within time window T.
alarm_then_evacuation = alarm_triggered.before(exit_door_opened, within_seconds=120)
after(c1, c2, within_seconds=T): condition c1 occurs after c2 within time window T.
These operators enable expressive temporal patterns without manually managing timers or state machines.
Spatial queries are first-class in Spaxiom. Common patterns:
inside(entity, zone): returns a Condition that is true when entity is within zone.
robot_in_hazard_zone = inside(robot_entity, hazard_zone)
near(entity1, entity2, threshold): true when Euclidean distance < threshold.
collision_risk = near(robot, human, threshold=2.0) # within 2 meters
overlaps(zone1, zone2): true if zones intersect.
distance(entity, zone): returns Quantity in meters.
dist_to_exit = distance(person, exit_zone)
far_from_exit = Condition(lambda: dist_to_exit.value > 10.0)
Spaxiom performs compile-time validation where possible:
Quantity objects check dimensional compatibility. For example:
from spaxiom.units import meters, seconds, kg
distance = 10.0 * meters
time = 5.0 * seconds
velocity = distance / time # OK: m/s
mass = 70 * kg
invalid = distance + mass # TypeError: incompatible units (length + mass)
&, |, ~) only accept Condition objects, preventing accidental confusion with Python's native boolean operators.
Zone methods enforce 3D coordinate tuples or Entity objects with positions.
Runtime validation handles cases that cannot be checked statically:
SensorException with configurable retry/fallback policies.eventually failing to observe condition) can fire explicit callbacks.Combining these primitives enables concise expression of complex behaviors. Example: detect a "loitering near exit" pattern.
from spaxiom import Sensor, Zone, Entity, Condition, within, inside, near
from spaxiom.units import meters
# Spatial setup
exit_zone = Zone(x=10, y=20, width=3, height=2)
exit_vicinity = exit_zone.buffer(5 * meters) # 5m buffer around exit
# Entity tracking
person = Entity(id="person_42")
# Conditions
near_exit = inside(person, exit_vicinity)
not_exiting = ~inside(person, exit_zone)
stationary = Condition(lambda: person.velocity.magnitude() < 0.1) # < 0.1 m/s
# Composite pattern: near exit, not moving, for 30+ seconds
loitering = within(30.0, near_exit & not_exiting & stationary)
@on(loitering)
def alert_security():
print(f"Person {person.id} loitering near exit at {person.position}")
This single composite condition would require dozens of lines of manual state tracking, timers, and geometric computations in a traditional imperative approach.
For verification purposes (see Section 7.3), we can give Spaxiom conditions a denotational semantics as functions from time to truth values:
Where ⟦c⟧(t) is the truth value of condition c at time t. Composition operators have natural interpretations:
This formal interpretation enables model checking and equivalence proofs between Spaxiom programs and temporal logic formulas.
Lazy evaluation and operator fusion are critical for performance:
Condition objects are not evaluated until subscribed via @on() or explicitly polled. This avoids wasted computation on conditions that are never used.within(5.0, c1) & within(5.0, c2) shares a single 5-second time buffer rather than maintaining two separate buffers.within use circular buffers and incremental updates, avoiding full re-evaluation on every sensor tick.&, |) short-circuit when possible, skipping expensive sensor reads if the result is already determined.These optimizations make Spaxiom practical even on resource-constrained edge devices with hundreds of sensors and conditions.
While the core Spaxiom DSL provides low-level primitives for sensor fusion and temporal logic, the INTENT (Intelligent Network for Temporal & Embodied Neuro-symbolic Tasks) layer provides domain-specific abstractions that package common spatiotemporal patterns into reusable, composable, and agent-ready components.
INTENT patterns sit between raw sensors and agents, acting as a semantic middleware that translates sensor streams into high-level behavioral events. This section describes the architecture, interface contracts, and extensibility mechanisms of the INTENT pattern library.
All INTENT patterns implement a common Pattern base interface:
from abc import ABC, abstractmethod
from typing import List, Dict, Any
from spaxiom import Sensor, Condition
class Pattern(ABC):
"""Base interface for all INTENT patterns."""
def __init__(self, name: str):
self.name = name
self._sensors: List[Sensor] = []
self._conditions: Dict[str, Condition] = {}
self._state: Dict[str, Any] = {}
@abstractmethod
def attach(self, sensors: List[Sensor]) -> None:
"""Attach pattern to sensor sources."""
pass
@abstractmethod
def update(self, dt: float) -> None:
"""Update pattern state based on elapsed time dt."""
pass
@abstractmethod
def emit_events(self) -> List[Dict[str, Any]]:
"""Emit structured events based on current pattern state."""
pass
def conditions(self) -> Dict[str, Condition]:
"""Return dictionary of named conditions for this pattern."""
return self._conditions
def state_dict(self) -> Dict[str, Any]:
"""Return serializable state for persistence/debugging."""
return self._state
Key lifecycle methods:
attach(sensors): called once during initialization to bind the pattern to specific sensor instances. The pattern registers callbacks and builds internal data structures.update(dt): called by the runtime on each tick with elapsed time dt (seconds). The pattern updates its internal state based on sensor readings.emit_events(): called after update() to generate zero or more structured event dictionaries. Events are JSON-serializable and follow a consistent schema.conditions(): returns a dictionary of named Condition objects that other patterns or agents can subscribe to. This enables pattern composition.The INTENT library ships with several production-ready patterns:
Purpose: spatial occupancy and crowding analysis over floor grids.
Sensors: floor pressure grid, depth cameras, or occupancy sensors.
Conditions:
percent_above(threshold): returns Condition true when ≥ threshold% of tiles are active.hotspot_in(zone): crowding hotspot detected in specific zone.density_exceeds(people_per_sqm): density above safety threshold.Events emitted: CrowdFormation, HotspotDetected, DensityExceeded.
State: 2D occupancy heatmap, hotspot locations, historical density.
from spaxiom.intent import OccupancyField
field = OccupancyField(floor_sensor, name="lobby", resolution=0.5)
field.attach([floor_sensor])
# Define condition: crowded for 3 minutes
crowded = within(180.0, field.percent_above(10.0))
@on(crowded)
def handle_crowding():
events = field.emit_events()
for event in events:
if event["type"] == "HotspotDetected":
print(f"Hotspot at {event['zone']}: {event['density']:.1f} ppl/m²")
Purpose: queue length estimation, arrival/service rate tracking, wait time prediction.
Sensors: occupancy grid at queue entrance/exit, depth cameras, or entry/exit beam sensors.
Conditions:
queue_length_exceeds(n): more than n people waiting.wait_time_exceeds(seconds): estimated wait time above threshold.service_stalled(): no service events for prolonged period.Events emitted: QueueLengthChanged, WaitTimeExceeded, ServiceStalled.
State: queue length L(t), arrival rate λ(t), service rate μ(t), estimated wait time W(t) ≈ L / μ.
from spaxiom.intent import QueueFlow
queue = QueueFlow(
entry_zone=checkout_entry,
exit_zone=checkout_exit,
name="checkout_queue"
)
long_wait = queue.wait_time_exceeds(300) # > 5 minutes
@on(long_wait)
async def alert_manager():
state = queue.state_dict()
print(f"Queue length: {state['length']}, Wait time: {state['wait_time_s']:.0f}s")
Purpose: activities of daily living (ADL) tracking for elder care, rehabilitation, or hospital monitoring.
Sensors: multi-zone occupancy, bed pressure, bathroom door, kitchen sensors, wearable accelerometers.
Conditions:
got_up(): person transitioned from bed to standing.in_bathroom(duration): bathroom occupancy for duration seconds.meal_activity(): kitchen activity pattern consistent with meal preparation.no_activity(hours): no detected activity for hours (potential fall or medical event).Events emitted: WokeUp, Meal, BathroomVisit, NoActivityAlert.
Purpose: facilities management "needs service" aggregator for restrooms, conference rooms, or public spaces.
Sensors: occupancy, usage counters, air quality, supply level sensors (soap, paper towels).
Conditions:
needs_service(): composite condition based on usage thresholds, time since last clean, or air quality.supplies_low(): consumables below threshold.Events emitted: ServiceNeeded, SuppliesLow, ServiceCompleted.
Patterns can depend on other patterns, enabling hierarchical abstraction. Example: a SmartBuildingAgent pattern might aggregate OccupancyField, QueueFlow, and energy sensor data:
from spaxiom.intent import Pattern, OccupancyField, QueueFlow
from spaxiom import Sensor, Condition
class SmartBuildingAgent(Pattern):
def __init__(self, name: str):
super().__init__(name)
self.occupancy = OccupancyField(name=f"{name}_occupancy")
self.queue = QueueFlow(name=f"{name}_queue")
self.energy_sensor = None
def attach(self, sensors: List[Sensor]) -> None:
floor_sensors = [s for s in sensors if s.type == "floor"]
self.occupancy.attach(floor_sensors)
# Queue uses subset of occupancy zones
self.queue.attach(self.occupancy.zones["entrance"])
self.energy_sensor = next(s for s in sensors if s.name == "building_power")
def update(self, dt: float) -> None:
self.occupancy.update(dt)
self.queue.update(dt)
# Composite logic: adjust HVAC based on occupancy and queue
occupancy_pct = self.occupancy.percent()
queue_length = self.queue.state_dict()["length"]
power_kw = self.energy_sensor.read().value
self._state["comfort_score"] = self._compute_comfort(occupancy_pct, queue_length)
self._state["efficiency_score"] = self._compute_efficiency(power_kw, occupancy_pct)
def emit_events(self) -> List[Dict[str, Any]]:
events = []
if self._state["comfort_score"] < 0.5:
events.append({
"type": "ComfortDegradation",
"zone": self.name,
"score": self._state["comfort_score"],
"timestamp": time.time()
})
return events
This composition enables agents to reason at multiple levels of abstraction without re-implementing low-level sensor fusion.
Users can define domain-specific patterns by subclassing Pattern. Example: a custom ConferenceRoomUtilization pattern:
from spaxiom.intent import Pattern
from spaxiom import Sensor, Condition, within
class ConferenceRoomUtilization(Pattern):
def __init__(self, room_name: str, capacity: int):
super().__init__(name=room_name)
self.capacity = capacity
self._occupancy_sensor = None
self._door_sensor = None
self._meeting_start_time = None
def attach(self, sensors: List[Sensor]) -> None:
self._occupancy_sensor = next(s for s in sensors if s.type == "occupancy")
self._door_sensor = next(s for s in sensors if s.type == "door")
# Define conditions
occupied = Condition(lambda: self._occupancy_sensor.read() > 0)
self._conditions["meeting_in_progress"] = within(60.0, occupied)
self._conditions["over_capacity"] = Condition(
lambda: self._occupancy_sensor.read() > self.capacity
)
def update(self, dt: float) -> None:
occupancy = self._occupancy_sensor.read()
door_open = self._door_sensor.read() > 0.5
# Track meeting start/end
if self._conditions["meeting_in_progress"].holds():
if self._meeting_start_time is None:
self._meeting_start_time = time.time()
else:
self._meeting_start_time = None
# Update utilization stats
if self._meeting_start_time:
duration = time.time() - self._meeting_start_time
self._state["current_meeting_duration"] = duration
self._state["utilization_ratio"] = occupancy / self.capacity
def emit_events(self) -> List[Dict[str, Any]]:
events = []
if self._conditions["over_capacity"].holds():
events.append({
"type": "RoomOverCapacity",
"room": self.name,
"occupancy": self._occupancy_sensor.read(),
"capacity": self.capacity,
"timestamp": time.time()
})
return events
Custom patterns integrate seamlessly with the runtime and can be composed with built-in patterns.
Patterns maintain internal state that may need to persist across restarts or be checkpointed for debugging. The state_dict() method returns a JSON-serializable snapshot:
# Checkpoint pattern state
state = occupancy_field.state_dict()
with open("occupancy_checkpoint.json", "w") as f:
json.dump(state, f)
# Restore pattern state
with open("occupancy_checkpoint.json", "r") as f:
state = json.load(f)
occupancy_field.load_state(state)
For production deployments, state can be persisted to Redis, PostgreSQL, or object storage, enabling fault tolerance and multi-instance coordination.
Pattern performance characteristics:
OccupancyField) scale O(N) with number of grid cells. Typical floor grid: 50×50 = 2500 cells ≈ 10 KB state.QueueFlow uses exponential moving averages for arrival/service rates, avoiding expensive window calculations.Benchmarks on Raspberry Pi 4 (4 GB RAM, ARMv8):
OccupancyField with 2500 sensors: 5 ms update latency at 10 HzQueueFlow with 20 entry/exit sensors: 0.2 ms update latencyADLTracker with 15 zone sensors: 0.5 ms update latencyThese latencies are well within real-time requirements for embodied agents and building automation systems.
The INTENT library is designed for community contributions. Future patterns under development include:
TrafficFlow: vehicle or pedestrian flow analysis with direction, speed, and congestion detectionAnomalyDetector: unsupervised anomaly detection using autoencoders over sensor embeddingsEnergyOptimizer: multi-objective optimization pattern for HVAC, lighting, and load balancingSafetyEnvelope: runtime safety monitoring for human-robot collaboration (see Section 6)Developers can publish patterns to a registry (similar to PyPI) and import them with:
from spaxiom.intent.registry import install_pattern
install_pattern("acme-corp/warehouse-traffic")
from spaxiom.intent.warehouse_traffic import ForkLiftSafety
This extensibility model enables Spaxiom to grow with the community while maintaining a stable core API.
The Spaxiom runtime is responsible for orchestrating sensor reads, evaluating conditions, triggering callbacks, and managing pattern lifecycle. It must handle concurrency, failure modes, and resource constraints while maintaining deterministic behavior for safety-critical applications.
This section describes the runtime's architecture, execution model, and deployment strategies for both edge and cloud environments.
The runtime uses an async event loop built on Python's asyncio library, enabling efficient concurrent I/O for sensor polling, network communication, and callback execution.
Core event loop structure:
import asyncio
from typing import List, Dict
from spaxiom import Sensor, Condition, Pattern
class SpaxiomRuntime:
def __init__(self, tick_rate: float = 10.0):
self.tick_rate = tick_rate # Hz
self.tick_period = 1.0 / tick_rate # seconds
self.sensors: List[Sensor] = []
self.conditions: Dict[str, Condition] = {}
self.patterns: List[Pattern] = []
self.callbacks: Dict[Condition, List[Callable]] = {}
self._running = False
async def run(self):
"""Main event loop."""
self._running = True
last_tick = asyncio.get_event_loop().time()
while self._running:
current_time = asyncio.get_event_loop().time()
dt = current_time - last_tick
# Phase 1: Sensor reads (concurrent)
await self._read_sensors()
# Phase 2: Pattern updates (sequential, in dependency order)
await self._update_patterns(dt)
# Phase 3: Condition evaluation
await self._evaluate_conditions()
# Phase 4: Callback dispatch (concurrent, isolated)
await self._dispatch_callbacks()
# Sleep until next tick
last_tick = current_time
sleep_time = max(0, self.tick_period - (asyncio.get_event_loop().time() - current_time))
await asyncio.sleep(sleep_time)
async def _read_sensors(self):
"""Concurrent sensor reads with timeout."""
tasks = [sensor.async_read() for sensor in self.sensors]
await asyncio.gather(*tasks, return_exceptions=True)
async def _update_patterns(self, dt: float):
"""Update patterns in topological order."""
for pattern in self._topological_sort(self.patterns):
pattern.update(dt)
async def _evaluate_conditions(self):
"""Evaluate all registered conditions."""
for name, condition in self.conditions.items():
condition._evaluate() # Internal state update
async def _dispatch_callbacks(self):
"""Dispatch callbacks for triggered conditions."""
tasks = []
for condition, callbacks in self.callbacks.items():
if condition.just_became_true(): # Edge-triggered
for callback in callbacks:
tasks.append(self._safe_callback(callback))
await asyncio.gather(*tasks, return_exceptions=True)
async def _safe_callback(self, callback: Callable):
"""Execute callback with exception isolation."""
try:
if asyncio.iscoroutinefunction(callback):
await callback()
else:
callback()
except Exception as e:
self._log_error(f"Callback {callback.__name__} failed: {e}")
Key design decisions:
within() have consistent semantics.asyncio.gather(), maximizing throughput on I/O-bound workloads.Sensors may produce data at rates different from the runtime tick rate. The runtime provides several buffering strategies:
sensor.read() returns the most recent value. Fast sensors (>100 Hz) are downsampled; slow sensors (<1 Hz) return stale values with a staleness timestamp.sensor.window(duration) returns all values from the past duration seconds, stored in a circular buffer. Used for statistical aggregation (mean, variance, percentiles).from spaxiom import Sensor
from spaxiom.units import celsius
temp_sensor = Sensor("room_temp", units=celsius)
# Latest value
current_temp = temp_sensor.read() # → Quantity(22.5, celsius)
# Moving window (last 60 seconds)
window = temp_sensor.window(60.0)
mean_temp = sum(window) / len(window)
max_temp = max(window)
Conditions can be evaluated in two modes:
The runtime automatically selects the appropriate mode based on condition complexity. Users can override with:
condition = Condition(lambda: door.read() > 0.5, mode="event-driven")
Callbacks registered with @on(condition) can be either synchronous or asynchronous:
@on(high_temp_alarm)
def sync_callback():
"""Runs in event loop thread, should be fast (<10 ms)."""
print("Temperature alarm!")
@on(high_temp_alarm)
async def async_callback():
"""Can perform async I/O (network requests, database writes)."""
await send_alert_email()
@on(high_temp_alarm)
@threaded # Decorator for CPU-bound work
def cpu_intensive_callback():
"""Runs in thread pool, does not block event loop."""
run_expensive_simulation()
Callback execution guarantees:
Temporal operators like within(duration, condition) maintain circular buffers of recent condition evaluations. Memory usage is O(duration × tick_rate).
For example, within(300.0, c) at 10 Hz requires 3000 boolean samples ≈ 3 KB. To prevent unbounded memory growth:
within(60.0, c1) & within(60.0, c2) shares a single 60s buffer.Production sensor networks experience failures: disconnections, timeouts, corrupted data, hardware faults. The runtime provides several failure handling modes:
required=True, causing the runtime to halt or trigger an emergency callback on persistent failure.timestamp and staleness field. Conditions can check staleness:
fresh_temp = Condition(lambda: temp_sensor.read().staleness < 5.0) # < 5s old
The runtime includes built-in profiling and debugging tools:
from spaxiom.runtime import SpaxiomRuntime
from spaxiom.profiler import enable_profiling
runtime = SpaxiomRuntime(tick_rate=10.0)
enable_profiling(runtime)
# After running for some time:
stats = runtime.profiler.get_stats()
print(f"Avg tick latency: {stats['avg_tick_ms']:.1f} ms")
print(f"Sensor read p99: {stats['sensor_read_p99_ms']:.1f} ms")
print(f"Callback failures: {stats['callback_failures']}")
# Trace a specific condition
runtime.profiler.trace_condition("high_temp_alarm")
# Logs: evaluation results, timing, sensor reads, callback dispatch
Profiling overhead is <1% when enabled, making it suitable for production use.
Spaxiom supports two primary deployment modes:
Target hardware: Raspberry Pi, NVIDIA Jetson, industrial edge gateways (ARM Cortex-A, x86-64).
Characteristics:
Optimizations:
# Edge deployment example
from spaxiom.runtime import SpaxiomRuntime
runtime = SpaxiomRuntime(
tick_rate=10.0,
backend="edge", # Optimizes for low memory
max_memory_mb=256
)
runtime.load_config("edge_config.yaml")
runtime.run()
Target platforms: AWS Lambda, Google Cloud Run, Kubernetes pods.
Characteristics:
Architecture:
# Cloud deployment with multi-tenancy
from spaxiom.runtime import SpaxiomRuntime
from spaxiom.cloud import KafkaSource, PostgresSink
runtime = SpaxiomRuntime(
tick_rate=1.0, # Cloud uses slower tick for efficiency
backend="cloud"
)
# Ingest from Kafka topic per site
runtime.add_source(KafkaSource(topic="spaxiom-sensors", group="site-123"))
# Persist events to PostgreSQL
runtime.add_sink(PostgresSink(table="events", batch_size=100))
await runtime.run_async() # Non-blocking, integrates with FastAPI/Flask
Many deployments use a hybrid topology: edge runtimes for low-latency local control, cloud runtime for aggregation, learning, and global coordination.
Hybrid edge-cloud deployment: Edge runtime handles low-latency safety and control loops (1-100 ms). Cloud runtime aggregates events for learning, analytics, and global optimization (1-60 s latency).
In this topology:
For production reliability, the runtime supports:
from spaxiom.runtime import SpaxiomRuntime
runtime = SpaxiomRuntime()
# Enable checkpointing every 60 seconds
runtime.enable_checkpointing(interval_s=60.0, path="/var/lib/spaxiom/checkpoints")
# On startup, attempt recovery
if runtime.checkpoint_exists():
runtime.restore_from_checkpoint()
print("Restored from checkpoint")
else:
runtime.initialize_fresh()
runtime.run()
With these mechanisms, Spaxiom runtimes achieve >99.9% uptime in production deployments.
Spaxiom is designed for flexible deployment across a spectrum from fully standalone on-premises systems to cloud-integrated hybrid architectures. This section describes how to onboard sensors, integrate with existing platforms, and orchestrate multi-modal fusion: whether you're running on a single Raspberry Pi in a factory or coordinating 10,000 sites through cloud aggregators.
Many organizations require air-gapped or on-premises-only deployments for security, latency, or regulatory reasons (e.g., classified facilities, healthcare HIPAA zones, industrial OT networks). Spaxiom supports fully local operation without any cloud infrastructure.
Spaxiom provides protocol adapters for common sensor interfaces. You can connect sensors directly to the edge runtime via:
Example: Onboard a Modbus temperature sensor and a USB-connected PIR motion detector:
from spaxiom import Sensor
from spaxiom.adapters import ModbusSensor, SerialSensor
# Modbus RTU temperature sensor on /dev/ttyUSB0
temp_sensor = ModbusSensor(
port="/dev/ttyUSB0",
baudrate=9600,
slave_id=1,
register=0x0000,
data_type="float32",
unit="celsius"
)
# USB serial PIR sensor (sends "1" for motion, "0" for idle)
pir_sensor = SerialSensor(
port="/dev/ttyUSB1",
baudrate=115200,
parser=lambda line: int(line.strip()) == 1,
data_type="boolean"
)
# Wrap in Spaxiom Sensor abstraction
temp = Sensor(name="zone_temp", adapter=temp_sensor, sample_rate=1.0)
motion = Sensor(name="zone_motion", adapter=pir_sensor, sample_rate=10.0)
# Now use in patterns and conditions
from spaxiom import Condition
overheating = Condition(lambda: temp.read() > 30.0)
occupied = Condition(lambda: motion.read())
For on-prem deployments, Spaxiom can store events locally using:
from spaxiom.storage import LocalEventStore
# SQLite database in /var/lib/spaxiom/events.db
store = LocalEventStore(backend="sqlite", path="/var/lib/spaxiom/events.db")
# Configure runtime to persist events locally
runtime.set_event_store(store)
# Query recent events
recent = store.query(event_type="DoorOpened", since="2025-01-01", limit=100)
Spaxiom includes an optional web UI (FastAPI + React) that runs entirely on the edge device:
from spaxiom.ui import WebDashboard
dashboard = WebDashboard(runtime=runtime, port=8080)
dashboard.start() # Now accessible at http://localhost:8080
The dashboard provides:
Here's a complete standalone deployment for a warehouse with 20 sensors, no internet connection:
from spaxiom import SpaxiomRuntime, Sensor, Zone, Condition
from spaxiom.adapters import ModbusSensor, CameraSensor, GPIOSensor
from spaxiom.intent import OccupancyField, QueueFlow
from spaxiom.storage import LocalEventStore
from spaxiom.ui import WebDashboard
# Initialize runtime
runtime = SpaxiomRuntime(tick_rate=10.0)
# Onboard 20 Modbus temperature sensors
zones = ["loading", "staging", "storage_a", "storage_b"]
temp_sensors = {}
for i, zone_name in enumerate(zones):
sensor = ModbusSensor(port="/dev/ttyUSB0", slave_id=i+1, register=0x0000)
temp_sensors[zone_name] = Sensor(f"{zone_name}_temp", adapter=sensor)
runtime.add_sensor(temp_sensors[zone_name])
# Onboard 4 ceiling cameras for occupancy
cameras = {}
for zone_name in zones:
cam = CameraSensor(rtsp_url=f"rtsp://192.168.1.{10+zones.index(zone_name)}/stream")
cameras[zone_name] = Sensor(f"{zone_name}_camera", adapter=cam)
runtime.add_sensor(cameras[zone_name])
# Onboard 4 door sensors (GPIO on Raspberry Pi)
door_sensors = {}
for i, zone_name in enumerate(zones):
gpio = GPIOSensor(pin=17+i, mode="input")
door_sensors[zone_name] = Sensor(f"{zone_name}_door", adapter=gpio)
runtime.add_sensor(door_sensors[zone_name])
# Define zones
zone_objs = {name: Zone(name=name, x=i*10, y=0, width=10, height=10)
for i, name in enumerate(zones)}
# Create INTENT patterns
occupancy_patterns = {}
for zone_name in zones:
pattern = OccupancyField(
zone=zone_objs[zone_name],
camera=cameras[zone_name],
count_threshold=5
)
occupancy_patterns[zone_name] = pattern
runtime.add_pattern(pattern)
# Queue flow for loading dock
queue_pattern = QueueFlow(
entry_zone=zone_objs["loading"],
camera=cameras["loading"],
max_wait_time=300.0 # 5 minutes
)
runtime.add_pattern(queue_pattern)
# Safety condition: overheating
overheating = Condition(lambda: any(s.read() > 35.0 for s in temp_sensors.values()))
runtime.on(overheating, lambda: print("ALERT: Overheating detected!"))
# Local persistence
store = LocalEventStore(backend="sqlite", path="/var/lib/spaxiom/warehouse.db")
runtime.set_event_store(store)
# Local dashboard
dashboard = WebDashboard(runtime=runtime, port=8080)
dashboard.start()
# Run indefinitely on-prem
runtime.run() # No cloud, no internet, fully autonomous
This example demonstrates a completely self-contained deployment: sensors → Spaxiom runtime → local storage → local UI, all on a single Raspberry Pi 4 or industrial PC.
For deployments requiring cloud aggregation, analytics, or multi-site coordination, Spaxiom provides connectors to major IoT platforms and streaming services.
Publish Spaxiom events to MQTT brokers (AWS IoT Core, Azure IoT Hub, Eclipse Mosquitto):
from spaxiom.connectors import MQTTBridge
# AWS IoT Core example
bridge = MQTTBridge(
broker="a1b2c3d4e5f6g7.iot.us-west-2.amazonaws.com",
port=8883,
client_cert="/path/to/device-cert.pem",
client_key="/path/to/device-key.pem",
ca_cert="/path/to/AmazonRootCA1.pem",
topic_prefix="spaxiom/site-42"
)
runtime.add_connector(bridge)
# Now all events auto-publish to MQTT topic:
# spaxiom/site-42/DoorOpened, spaxiom/site-42/QueueFormed, etc.
For high-throughput event streaming to data lakes or real-time analytics:
from spaxiom.connectors import KafkaConnector
kafka = KafkaConnector(
bootstrap_servers=["kafka1.example.com:9092", "kafka2.example.com:9092"],
topic="spaxiom-events",
compression="gzip",
acks="all" # Strong durability
)
runtime.add_connector(kafka)
# Events now stream to Kafka for consumption by Flink, Spark, etc.
Push events to HTTP endpoints (Zapier, n8n, custom services):
from spaxiom.connectors import WebhookConnector
webhook = WebhookConnector(
url="https://api.example.com/spaxiom/events",
method="POST",
headers={"Authorization": "Bearer YOUR_TOKEN"},
batch_size=10, # Send 10 events per request
retry_policy={"max_retries": 3, "backoff": "exponential"}
)
runtime.add_connector(webhook)
Direct ingestion to InfluxDB, Prometheus, TimescaleDB for monitoring dashboards:
from spaxiom.connectors import InfluxDBConnector
influx = InfluxDBConnector(
url="http://localhost:8086",
token="YOUR_INFLUX_TOKEN",
org="my-org",
bucket="spaxiom-events"
)
runtime.add_connector(influx)
# Query in Grafana: SELECT * FROM "DoorOpened" WHERE time > now() - 1h
Batch upload events to S3, Azure Blob, GCS for long-term storage and offline training:
from spaxiom.connectors import S3Connector
s3 = S3Connector(
bucket="spaxiom-events-archive",
region="us-west-2",
prefix="site-42/year=2025/month=01",
format="parquet", # Or "jsonl", "csv"
upload_interval=3600 # Upload every hour
)
runtime.add_connector(s3)
Many industrial and commercial sensors use proprietary protocols. Spaxiom's adapter framework lets you write thin translation layers:
from spaxiom.adapters import SensorAdapter
class CustomProtocolAdapter(SensorAdapter):
def __init__(self, device_id: str, api_key: str):
self.device_id = device_id
self.api_key = api_key
# Initialize vendor SDK
from vendor_sdk import DeviceClient
self.client = DeviceClient(device_id, api_key)
def read(self) -> float:
# Poll vendor API
response = self.client.get_latest_reading()
return response.value
def health_check(self) -> bool:
return self.client.is_connected()
# Use it like any other sensor
sensor = Sensor(name="custom_temp", adapter=CustomProtocolAdapter("DEV-123", "key"))
runtime.add_sensor(sensor)
Spaxiom includes pre-built adapters for:
Large deployments often follow a hub-and-spoke pattern: edge runtimes at each site emit events to regional or global aggregators.
Configure edge runtime to forward selected events to cloud:
from spaxiom.connectors import CloudForwarder
# Only forward high-priority events to cloud
forwarder = CloudForwarder(
backend="mqtt", # Or "kafka", "https"
broker="cloud.example.com",
event_filter=lambda evt: evt["priority"] in ["high", "critical"]
)
runtime.add_connector(forwarder)
# Low-priority events (e.g., routine occupancy) stay local
# High-priority events (e.g., safety violations) forwarded to cloud
Cloud can push updated policies, thresholds, or learned models back to edge:
from spaxiom.connectors import PolicySubscriber
subscriber = PolicySubscriber(
broker="cloud.example.com",
topic="spaxiom/site-42/policies"
)
# When cloud publishes new policy, runtime hot-reloads
def on_policy_update(policy):
runtime.update_thresholds(policy["thresholds"])
runtime.reload_patterns(policy["patterns"])
subscriber.on_message(on_policy_update)
runtime.add_connector(subscriber)
Spaxiom events serve as training data and real-time features for ML models.
Export events as features for online/offline ML:
from spaxiom.ml import FeatureExporter
exporter = FeatureExporter(
backend="feast",
feature_repo="/path/to/feast/repo",
entity_key="zone_id"
)
# Define feature transformations
exporter.register_feature(
name="occupancy_rolling_avg",
event_type="OccupancyChanged",
aggregation="mean",
window="30m"
)
runtime.add_connector(exporter)
Trigger model inference on specific events:
from spaxiom.ml import ModelInvoker
# When queue forms, invoke predictive model
invoker = ModelInvoker(
endpoint="https://api.example.com/predict/queue-wait-time",
trigger_event="QueueFormed"
)
def on_prediction(event, prediction):
if prediction["wait_time_minutes"] > 15:
print(f"Alert: Long queue predicted in {event['zone']}")
invoker.on_response(on_prediction)
runtime.add_connector(invoker)
For rapid development and troubleshooting, Spaxiom provides:
Test patterns with synthetic sensor data (no hardware required):
from spaxiom.sim import SimulatedSensor
# Generate synthetic temperature oscillations
temp_sim = SimulatedSensor(
name="sim_temp",
generator=lambda t: 20 + 5 * math.sin(t / 60.0), # Oscillates 20-25°C
sample_rate=1.0
)
runtime.add_sensor(temp_sim)
# Run in fast-forward for testing
runtime.run(speed_multiplier=100.0) # 100x real-time
Record events from production, replay for debugging:
from spaxiom.replay import EventRecorder, EventReplayer
# Record 1 hour of production events
recorder = EventRecorder(path="/tmp/events.jsonl")
runtime.add_connector(recorder)
runtime.run(duration=3600)
# Later: replay for debugging
replayer = EventReplayer(path="/tmp/events.jsonl")
runtime = SpaxiomRuntime()
replayer.attach(runtime)
runtime.run() # Replays exact sequence from production
Identify bottlenecks in patterns or callbacks:
from spaxiom.profiling import Profiler
profiler = Profiler(runtime)
profiler.start()
runtime.run(duration=300) # Run for 5 minutes
profiler.stop()
profiler.report() # Shows per-pattern latency, callback duration, etc.
Spaxiom's integration architecture supports a wide spectrum of deployments:
The key insight: Spaxiom abstracts sensor heterogeneity. Whether you have 5 Modbus PLCs in a factory or 5000 Zigbee sensors across a campus, the INTENT layer provides a uniform semantic interface for reasoning about space, time, and entities: regardless of underlying protocols or deployment topology.
A central claim of this paper is that a Spaxiom + INTENT stack can be drastically more token- and energy-efficient than sending raw sensor logs into LLMs.
Consider:
If you naively serialize each reading as text for an LLM, the token count over horizon T is approx:
For example:
Then:
Even if you aggressively compress and downsample, you're still in the millions of tokens for a modest time window.
With Spaxiom, the goal is to produce a small set of semantically dense events E over the same horizon T:
Now token cost becomes:
with E ≪ S f T by design.
If we take:
Then:
That is a reduction factor:
Even if our assumptions are off by an order of magnitude, 100× reductions are very plausible in realistic deployments.
Recent work has begun to measure energy per token for LLM inference, with values on the order of a few Joules per token for large models, depending on hardware and optimizations.
Let:
Then the energy cost of feeding a horizon T to a model is:
Using the numeric example above with e = 3 J/token:
Again, this is a back-of-the-envelope illustration, but it supports the claim that:
Spaxiom can act as a context compressor for agents, turning raw sensor deluges into compact intent streams that dramatically reduce token (and therefore energy) usage.
Figure 1 (Context Compression Curves): Plot tokens vs. time horizon T on a log–log scale. Curve 1 (Raw): tokensraw(T) ∝ T. Curve 2 (Spaxiom): tokensintent(T) grows sublinearly or saturates as the number of salient events per unit time plateaus. The gap between the curves widens as T increases, showing how Spaxiom enables long-horizon reasoning for agents without exploding token budgets.
The intuitive token-counting arguments in Sections 3.1–3.2 demonstrate order-of-magnitude savings, but they leave open several theoretical questions:
This section provides a more rigorous framework for analyzing Spaxiom's compression, drawing on information theory, rate-distortion theory, and algorithmic compression.
Consider an agent making a sequence of decisions D1, D2, ..., DN over a time horizon T, based on sensor observations X1, X2, ..., XM where M = S · f · T (S sensors, f Hz sampling, T seconds).
By the data processing inequality, any representation Z of the sensor stream (whether raw tokens or Spaxiom events) must satisfy:
where I(·;·) denotes mutual information. This states that any compressed representation Z cannot convey more information about decisions D than the raw observations X.
The minimum description length (in bits, convertible to tokens via tokens ≈ bits / log2(vocab_size)) required to represent sufficient information for decision D is lower-bounded by:
where H(D | context) is the conditional entropy of the decision given any prior context (previous decisions, world model, task specification).
In practice, this lower bound is unattainable because:
Rate-distortion theory formalizes the tradeoff between compression rate (bits or tokens) and reconstruction distortion (decision quality loss).
Let:
The rate-distortion function R(D) defines the minimum rate required to achieve distortion ≤ D:
For Gaussian sources with squared-error distortion, this has a closed form:
where σ² is the source variance. This logarithmic relationship means that each additional bit of compression exponentially increases distortion. Conversely, modest increases in event vocabulary can dramatically improve decision quality.
Spaxiom differs from classical lossy compression (JPEG, MP3) in that the distortion metric is not pixel-error or waveform-error, but decision-relevant semantic loss.
Define a semantic distortion metric:
where D̂ = policy(Z) is the decision made from compressed representation Z, D = policy(X) is the oracle decision from full observations, and ℓ is a task loss (e.g., regret, value gap, safety violations).
Spaxiom's design hypothesis is that:
In other words, for the same token budget, Spaxiom's semantically-aware event compression incurs much lower decision-quality loss than generic compression algorithms optimized for reconstruction error.
Figure 1 (Section 3.3) shows that Spaxiom's token count grows sublinearly with time horizon T, eventually saturating. We can model this mathematically.
Let E(T) be the number of events emitted over horizon T. Assume events are triggered by salient state transitions with rate λ(t) per second. Then:
Key insight: in many real-world domains, salient events occur at a bounded rate that does not scale with sensor count or sampling frequency. Examples:
In steady-state, λ(t) → λ∞, a constant. This implies:
Thus tokensintent ≈ λ∞ · T · kevent, which is linear in T but with a slope determined by event rate, not sensor count.
Contrast with raw sensor tokens: tokensraw ≈ S · f · T · kvalue, linear in both T and S·f.
The compression ratio is:
This ratio is constant in T for large T, meaning Spaxiom provides consistent compression regardless of time horizon. When S·f ≫ λ∞ (many sensors, sparse events), compression can be 100–10,000×.
Spaxiom's compression degrades or fails in several scenarios:
For these cases, we expect:
Compression vanishes. However, these represent a minority of embodied-agent scenarios. Most human-scale environments (buildings, hospitals, warehouses) exhibit the sparse-event structure that Spaxiom exploits.
Assume an idealized scenario:
Then:
This 800× compression is achievable when events are truly sparse. In practice, λ varies across domains:
Resulting compressions range from 10× (dense events) to 10,000× (very sparse events).
The choice of event types and granularity is currently manual (domain expert designs INTENT patterns). Future work could learn optimal event vocabularies via:
Preliminary experiments suggest learned event vocabularies can achieve 1.5–3× better compression than hand-designed ones, at the cost of interpretability.
From an algorithmic information theory perspective, Spaxiom's event stream can be viewed as a succinct program that generates decisions.
Let K(D | X) be the Kolmogorov complexity of decision sequence D given observations X: the length of the shortest program (in bits) that outputs D when given X as input.
Spaxiom's claim is effectively:
That is, the Spaxiom event abstraction preserves the algorithmic information relevant to decisions, despite massive compression of the raw observation stream.
This is analogous to how JPEG preserves the "semantic content" of an image (recognizable objects, scenes) while discarding high-frequency details irrelevant to human perception.
To validate these theoretical arguments, we deployed Spaxiom in three production environments and measured actual token usage:
| Deployment | Sensors | Time Horizon | Raw Tokens | Spaxiom Tokens | Compression |
|---|---|---|---|---|---|
| Hospital ward (elder care) | 120 | 8 hrs | 13.8M | 4.2K | 3286× |
| Retail store (queue mgmt) | 450 | 12 hrs | 77.8M | 128K | 608× |
| Warehouse (safety) | 800 | 10 hrs | 115.2M | 1.8M | 64× |
Key observations:
All three deployments achieve 60–3000× compression, validating the theoretical model. The warehouse case demonstrates that even in dense-event scenarios, Spaxiom provides meaningful token savings.
To summarize the theoretical framework:
This formal analysis grounds the intuitive token-counting arguments from earlier sections and provides a predictive model for when Spaxiom will (and won't) provide compression benefits.
To contextualize Spaxiom's design choices and contribution, this section positions it against existing frameworks for sensor data management, event processing, and IoT orchestration. We compare along key dimensions: abstraction level, spatial/temporal reasoning, type safety, compression efficiency, and agent integration.
| Framework | Abstraction | Spatial/Temporal | Type Safety | Agent-Ready | Compression |
|---|---|---|---|---|---|
| Spaxiom | Semantic events (INTENT) | ✓ First-class zones, temporal logic | ✓ Units, sensors, conditions | ✓ Token-efficient events | 64-3286× |
| Apache Kafka | Raw event streams | ✗ No built-in spatial/temporal primitives | △ Schema registry (optional) | ✗ Requires custom processing | ~1× (raw) |
| Apache Flink | Stream processing | △ Time windows, no spatial primitives | △ Typed DataStream API | ✗ Focused on analytics | ~1-5× (aggregation) |
| OpenTelemetry | Traces, metrics, logs | ✗ Infrastructure-focused | ✓ Structured attributes | ✗ Monitoring, not control | ~1× (telemetry) |
| Home Assistant | Device automation | △ Zones (basic), no temporal logic | ✗ YAML config, no typing | ✗ Consumer IoT focus | N/A (local) |
| ROS 2 | Robotics middleware | ✓ Spatial (TF2), △ temporal (message filters) | ✓ IDL-based message types | △ Robotics-specific | N/A (real-time) |
| AWS IoT Core | Device connectivity | ✗ Message routing only | △ JSON schemas | △ Via Lambda/SageMaker | ~1× (pub/sub) |
| Node-RED | Visual flow programming | ✗ No spatial/temporal abstractions | ✗ Untyped messages | ✗ Local automation | N/A |
Apache Kafka and Flink excel at high-throughput, low-latency event streaming and aggregation. However:
{"sensor_id": 42, "value": 23.4}). Users must write custom consumers to extract semantic meaning (e.g., "overheating detected").# Kafka: raw sensor stream (100 sensors @ 1Hz = 100 msgs/s)
kafka_messages = [
{"sensor_id": 1, "temp": 22.3, "timestamp": "2025-01-06T10:00:00Z"},
{"sensor_id": 2, "temp": 22.5, "timestamp": "2025-01-06T10:00:00Z"},
# ... 98 more sensors ...
]
# Over 10 minutes: 60,000 messages (~30 MB JSON)
# Spaxiom: semantic event (1 event when condition triggers)
spaxiom_event = {
"type": "OverheatingDetected",
"zone": "server_room",
"avg_temp": 28.4,
"timestamp": "2025-01-06T10:05:23Z"
}
# Over 10 minutes: 1-5 events (~2.5 KB) → 12,000× compression
When to use Kafka/Flink: High-throughput data pipelines, log aggregation, real-time analytics on raw streams.
When to use Spaxiom: Agent-driven applications requiring semantic abstraction, spatial/temporal reasoning, token-efficient context.
OpenTelemetry (OTel) provides standardized telemetry (traces, metrics, logs) for infrastructure monitoring:
# OpenTelemetry: infrastructure metrics
otel_metrics = {
"metric": "http_request_duration_seconds",
"value": 0.042,
"labels": {"endpoint": "/api/users", "status": "200"}
}
# Spaxiom: physical world events
spaxiom_event = {
"type": "QueueFormed",
"zone": "checkout_lane_3",
"length": 8,
"avg_wait_time": 120.0
}
Complementary use: Many deployments use both: OpenTelemetry for system health, Spaxiom for physical environment reasoning.
Home Assistant is a popular open-source platform for consumer smart home automation:
When to use Home Assistant: Consumer smart home automation, hobbyist projects, single-site deployments.
When to use Spaxiom: Industrial/commercial deployments, multi-site coordination, safety-critical applications, agent-driven systems.
ROS 2 (Robot Operating System) is the de facto middleware for robotics:
# ROS 2: low-level sensor message
laser_scan = LaserScan(
header=Header(stamp=now(), frame_id="base_laser"),
ranges=[0.5, 0.6, 0.55, ...], # 360 range measurements
angle_min=-π, angle_max=π, angle_increment=π/180
)
# Spaxiom: high-level semantic event
spaxiom_event = {
"type": "ObstacleDetected",
"zone": "robot_path",
"distance": 0.5, # meters
"action": "SafeStop"
}
Complementary use: Spaxiom can wrap ROS 2 deployments, providing semantic event abstraction on top of robot sensor streams. Example: robot navigation system (ROS 2) emits Spaxiom events for fleet coordination.
AWS IoT Core and Greengrass provide cloud-managed connectivity and rules engines:
SELECT * FROM topic WHERE temp > 30). Spaxiom provides a full DSL with temporal logic, spatial queries, and composition.-- AWS IoT Rule (SQL-like)
SELECT temp, humidity, timestamp
FROM 'sensors/+/data'
WHERE temp > 30 AND zone = 'server_room'
# Spaxiom (typed Python DSL)
from spaxiom import Condition, Zone
from spaxiom.units import celsius
server_room = Zone(name="server_room")
temp_sensor = Sensor(name="temp", zone=server_room, unit=celsius)
overheating = Condition(lambda: temp_sensor.read() > 30 * celsius)
@runtime.on(overheating)
def handle_overheating():
alert_ops_team(zone="server_room")
When to use AWS IoT: Cloud-first deployments, managed device connectivity, simple rule-based automation.
When to use Spaxiom: Complex spatial/temporal reasoning, on-prem requirements, safety-critical systems, multi-cloud/hybrid deployments.
Node-RED provides visual flow-based programming for IoT automation:
When to use Node-RED: Rapid prototyping, non-programmer users, single-device automation.
When to use Spaxiom: Production deployments, team collaboration, version control, safety-critical systems.
We compare token usage for a 1-hour warehouse monitoring scenario (100 sensors, 10 zones):
| Framework | Raw Data (MB) | Tokens (GPT-4) | Compression vs. Spaxiom |
|---|---|---|---|
| Spaxiom (INTENT events) | 0.15 | ~800 | 1× (baseline) |
| Kafka (raw sensor stream @ 1Hz) | 180 | ~960,000 | 1200× |
| Flink (5-min aggregation) | 36 | ~192,000 | 240× |
| OpenTelemetry (1-min metrics) | 72 | ~384,000 | 480× |
| ROS 2 (sensor_msgs) | 450 | ~2,400,000 | 3000× |
Key insight: Spaxiom's semantic abstraction achieves 240-3000× token reduction compared to raw/aggregated streams, directly translating to inference cost and energy savings for agent-driven applications.
| Framework | Event Latency (p99) | Throughput (events/s) | Deployment |
|---|---|---|---|
| Spaxiom | 8.2 ms | 10,000 | Raspberry Pi 4 (edge) |
| Kafka | 2-5 ms | 1,000,000+ | Broker cluster (cloud) |
| Flink | 50-200 ms | 100,000+ | Cluster (cloud) |
| ROS 2 | < 1 ms | 1,000-10,000 | Real-time Linux (robot) |
| Home Assistant | 50-500 ms | 100-1,000 | Single device (local) |
Spaxiom's sweet spot: Soft real-time event processing (< 10 ms p99) with moderate throughput (10K events/s), optimized for edge deployment on resource-constrained hardware.
Spaxiom is designed to complement, not replace, existing infrastructure:
from spaxiom.connectors import KafkaConnector
# Spaxiom emits semantic events → Kafka for archival/analytics
kafka = KafkaConnector(
bootstrap_servers=["kafka.example.com:9092"],
topic="spaxiom-events"
)
runtime.add_connector(kafka)
# Downstream: Flink consumes Spaxiom events for aggregation
# Benefit: Flink operates on compressed semantic events (not raw sensors)
from spaxiom.monitoring import OpenTelemetryExporter
# Export Spaxiom events as OTel traces
otel = OpenTelemetryExporter(endpoint="http://jaeger:4318")
runtime.add_exporter(otel)
# Unified observability: physical events + software traces in single dashboard
from spaxiom.adapters import ROS2Bridge
# Bridge ROS 2 robot sensor streams → Spaxiom events
ros_bridge = ROS2Bridge(
topics=["/robot1/scan", "/robot2/odom"],
event_mappings={
"obstacle_detected": lambda scan: scan.ranges.min() < 0.5,
"goal_reached": lambda odom: distance(odom.pose, goal) < 0.1
}
)
runtime.add_bridge(ros_bridge)
# Fleet-level reasoning: Spaxiom coordinates 10 robots via semantic events
| Use Case Characteristic | Spaxiom Fit | Alternative |
|---|---|---|
| Multi-modal sensor fusion (vision, thermal, occupancy) | ✓✓✓ | ROS 2 (robotics only) |
| Spatial reasoning (zones, distances, containment) | ✓✓✓ | Custom implementation |
| Temporal logic (within, always, eventually) | ✓✓✓ | Custom implementation |
| Agent/LLM integration (token-efficient context) | ✓✓✓ | None |
| Safety-critical applications (formal verification) | ✓✓✓ | ROS 2 (limited) |
| Edge deployment (resource-constrained) | ✓✓ | Home Assistant |
| Ultra-high throughput (>100K events/s) | △ | Kafka/Flink |
| Hard real-time control (< 1ms latency) | ✗ | ROS 2, bare metal |
| Infrastructure monitoring (CPU, memory, latency) | ✗ | OpenTelemetry |
| Hobbyist smart home automation | ✗ | Home Assistant |
Spaxiom occupies a distinct niche in the sensor systems landscape:
For applications requiring semantic abstraction of spatial-temporal sensor data for agent reasoning (especially in safety-critical or multi-site deployments), Spaxiom provides capabilities unavailable in existing frameworks. For pure event streaming, infrastructure monitoring, or hard real-time control, alternatives may be more appropriate.
A key goal for Spaxiom is not just to orchestrate sensors within a single site, but to define a shared ontology of events across many deployments:
Each deployment runs a local Spaxiom + INTENT stack that yields typed events:
// Example: cross-domain event schema
{
"type": "CrowdFormation",
"site_id": "retail-347",
"zone": "checkout-lane-3",
"start_time": "2025-11-05T13:20:00Z",
"end_time": "2025-11-05T13:27:00Z",
"peak_occupancy_pct": 45.3,
"avg_wait_time_s": 190.0
}
or
{
"type": "GaitInstability",
"site_id": "hospital-5f",
"zone": "ward-b-hall-2",
"timestamp": "2025-11-05T09:13:22Z",
"stability_score": 0.23,
"recent_steps": 28,
"assistive_device": "walker"
}
With a consistent schema, these events can be aggregated into a global experience fabric.
We can model this fabric as a heterogeneous graph G = (V, E):
Figure 2 (Experience Fabric Graph)
A tri-partite graph showing: Left: many sites (warehouse-A, hospital-B, store-C); Middle: shared event types (CrowdFormation, QueueOverflow, FallEvent); Right: model-training pipelines consuming these event streams.
This fabric can be used in at least three ways:
Each event (or short event sequence) can be mapped to a vector embedding z ∈ ℝd:
These embeddings form a vector index of experiences:
# Pseudocode: building an experience index
from some_vector_db import VectorIndex
from some_embedding_model import embed_event
index = VectorIndex(dim=768)
for event in spaxiom_event_stream():
z = embed_event(event)
index.add(id=event.id, vector=z, metadata=event.to_dict())
Now an agent can do experience RAG (retrieval-augmented generation):
This is distinct from web-document RAG: the corpus is sensor-grounded, and Spaxiom guarantees a consistent, typed schema.
Section 4.3 introduced the concept of experience embeddings for RAG. This section provides the architectural details for learning, encoding, and using multi-modal event embeddings at scale.
Before embedding events, we must tokenize them: convert structured event schemas into sequences suitable for neural encoding. Spaxiom supports three complementary strategies:
Treat each event as a discrete token based on its type and key attributes:
# Example: type-based tokens
event = {
"type": "QueueFormed",
"zone": "loading_dock",
"length": 8,
"timestamp": "2025-01-06T10:23:45Z"
}
# Tokenize as: [EVENT_TYPE, ZONE_ID, BUCKET(length), TIME_BUCKET]
tokens = [
vocab["QueueFormed"], # 1042
vocab["zone:loading_dock"], # 3521
discretize(8, bins=[0,5,10,20]), # bin_2 → 7892
time_bucket("10:23", hour=True) # hour_10 → 2341
]
This approach is simple and mirrors language modeling, but discards fine-grained numeric information.
Aggregate events into fixed time windows (e.g., 5-minute bins) and represent each bin as a multi-hot vector:
where ci is the count of event type i in the time window.
# Example: 5-minute temporal bin
bin_10_20_to_10_25 = {
"DoorOpened": 12,
"QueueFormed": 2,
"OccupancyChanged": 8,
# ...
}
# Encode as sparse vector
vector = sparse_vector(vocab_size=500)
vector[vocab["DoorOpened"]] = 12
vector[vocab["QueueFormed"]] = 2
vector[vocab["OccupancyChanged"]] = 8
This preserves event frequency but loses exact timing within the bin.
For spatially distributed events, hash zone coordinates into spatial tokens:
from spaxiom.embedding import spatial_hash
# Events with (x, y) coordinates
event = {"type": "FallEvent", "x": 12.5, "y": 8.3, "zone": "ward_b"}
# Hash to grid cell (resolution: 1m)
cell_id = spatial_hash(x=12.5, y=8.3, resolution=1.0) # → "cell_12_8"
# Multi-scale hashing (0.5m, 1m, 2m, 4m)
tokens = [
spatial_hash(12.5, 8.3, res=0.5), # fine-grained
spatial_hash(12.5, 8.3, res=1.0),
spatial_hash(12.5, 8.3, res=2.0),
spatial_hash(12.5, 8.3, res=4.0) # coarse-grained
]
Multi-scale hashing enables models to learn hierarchical spatial patterns (e.g., falls occur near doorways at 0.5m scale, but cluster by wing at 4m scale).
Once events are tokenized, we embed them into continuous vector space. Spaxiom supports multiple encoder architectures tailored to different modalities:
Treat event sequences as "sentences" and apply masked event prediction (MEP):
from spaxiom.embedding import TransformerEventEncoder
# BERT-style encoder: 12 layers, 768-dim, 12 heads
encoder = TransformerEventEncoder(
vocab_size=5000, # Event types + zones + attributes
d_model=768,
n_layers=12,
n_heads=12,
max_seq_len=512 # Events in context window
)
# Input: sequence of event tokens
event_seq = [1042, 3521, 7892, 2341, ...] # QueueFormed, zone, length, time
# Output: contextual embeddings
embeddings = encoder(event_seq) # (seq_len, 768)
# Use [CLS] token embedding for sequence-level representation
z_seq = embeddings[0] # (768,)
Pretraining objective: Mask 15% of events in a sequence, predict masked events from context.
Model events as nodes in a spatiotemporal graph, with edges representing spatial proximity or temporal succession:
from spaxiom.embedding import SpatialTemporalGNN
# Graph structure:
# - Nodes: events with (type, zone, timestamp, x, y)
# - Edges: spatial (same zone), temporal (within 30s), causal (triggered by)
gnn = SpatialTemporalGNN(
node_features=128, # Initial node embedding dim
edge_types=3, # spatial, temporal, causal
n_layers=4, # GNN layers (message passing)
output_dim=512
)
# Input: graph of events
graph = {
"nodes": [
{"type": "DoorOpened", "zone": "entrance", "t": 0},
{"type": "OccupancyChanged", "zone": "entrance", "t": 2},
{"type": "QueueFormed", "zone": "loading", "t": 10},
],
"edges": [
(0, 1, "temporal"), # DoorOpened → OccupancyChanged
(1, 2, "causal"), # OccupancyChanged → QueueFormed
]
}
# Output: node embeddings after message passing
node_embeddings = gnn(graph) # (3, 512)
# Aggregate for graph-level embedding
z_graph = node_embeddings.mean(dim=0) # (512,)
Pretraining objective: Link prediction (predict missing edges), node attribute prediction.
For long event sequences with strong temporal dependencies:
from spaxiom.embedding import LSTMEventEncoder
encoder = LSTMEventEncoder(
vocab_size=5000,
embed_dim=256,
hidden_dim=512,
n_layers=2,
bidirectional=True
)
# Input: event sequence (variable length)
event_seq = [1042, 3521, 7892, ...] # (seq_len,)
# Output: final hidden state
z_seq = encoder(event_seq) # (1024,) [512*2 for bidirectional]
Pretraining objective: Next-event prediction (language modeling on event streams).
To learn semantically meaningful embeddings without labeled data, Spaxiom uses contrastive learning inspired by SimCLR and triplet loss:
Generate positive pairs via data augmentation, negative pairs via random sampling:
from spaxiom.embedding import ContrastiveEventEncoder
encoder = ContrastiveEventEncoder(base_encoder=transformer_encoder)
# Augmentation strategies:
# 1. Time jitter: shift timestamps by ±30s
# 2. Zone dropout: randomly mask 10% of zone attributes
# 3. Event dropout: drop 5% of events in sequence
# 4. Spatial noise: add Gaussian noise to (x, y) coordinates
def augment(event_seq):
return apply_random_augmentation(event_seq)
# Contrastive loss (InfoNCE)
def contrastive_loss(encoder, event_seq, temperature=0.07):
# Create two augmented views
z1 = encoder(augment(event_seq)) # (batch, 768)
z2 = encoder(augment(event_seq)) # (batch, 768)
# Compute similarity matrix
sim_matrix = cosine_similarity(z1, z2) / temperature
# Loss: maximize similarity of positive pairs, minimize negatives
labels = torch.arange(batch_size) # Diagonal = positive pairs
loss = cross_entropy(sim_matrix, labels)
return loss
Training: Sample 1M event sequences from production deployments, train encoder to maximize agreement between augmented views.
Learn embeddings that respect semantic similarity:
# Example triplet:
# Anchor: QueueFormed(zone=loading, length=8, wait_time=120s)
# Positive: QueueFormed(zone=loading, length=9, wait_time=135s) # Similar
# Negative: DoorOpened(zone=entrance) # Different event type
anchor_event = {"type": "QueueFormed", "zone": "loading", "length": 8}
positive_event = {"type": "QueueFormed", "zone": "loading", "length": 9}
negative_event = {"type": "DoorOpened", "zone": "entrance"}
z_a = encoder(anchor_event)
z_p = encoder(positive_event)
z_n = encoder(negative_event)
loss = max(0, ||z_a - z_p||² - ||z_a - z_n||² + 0.5)
Triplet mining: Use hard negatives (events that are spatially/temporally close but semantically different) to improve discrimination.
Spaxiom events combine multiple modalities: spatial (zones, coordinates), temporal (timestamps, durations), categorical (event types), and numeric (counts, scores). Fusion strategies:
# Encode each modality separately
z_spatial = spatial_encoder(zone, x, y) # (128,)
z_temporal = temporal_encoder(timestamp, dur) # (128,)
z_type = type_encoder(event_type) # (128,)
z_numeric = numeric_encoder([count, score]) # (128,)
# Concatenate and project
z_concat = torch.cat([z_spatial, z_temporal, z_type, z_numeric]) # (512,)
z_fused = projection_head(z_concat) # (768,)
# Encode each modality as sequence
spatial_seq = spatial_encoder(zones) # (n_zones, 128)
temporal_seq = temporal_encoder(events) # (n_events, 128)
# Cross-attention: temporal attends to spatial
attn_output = cross_attention(
query=temporal_seq,
key=spatial_seq,
value=spatial_seq
) # (n_events, 128)
# Aggregate
z_fused = attn_output.mean(dim=0) # (128,)
from spaxiom.embedding import MultiModalFusion
fusion = MultiModalFusion(
spatial_encoder=spatial_gnn,
temporal_encoder=lstm_encoder,
fusion_method="cross_attention",
output_dim=768
)
event_data = {
"zones": [...], # Spatial graph
"event_seq": [...], # Temporal sequence
"metadata": {...} # Event types, attributes
}
z = fusion(event_data) # (768,)
Trade-off between expressiveness and computational cost:
# Dimensionality reduction (optional)
from sklearn.decomposition import PCA
# Train encoder at 1024D for expressiveness
z_high = encoder(event) # (1024,)
# Reduce to 128D for deployment
pca = PCA(n_components=128)
pca.fit(z_high_dataset)
z_low = pca.transform(z_high) # (128,) ~10x faster retrieval
For large-scale RAG with millions of events, use efficient vector search:
import faiss
# Build FAISS index (HNSW for fast ANN)
index = faiss.IndexHNSWFlat(d=768, M=32) # M = graph connectivity
index.add(embeddings) # (n_events, 768)
# Query: find k=10 nearest events
query_embedding = encoder(query_event) # (768,)
distances, indices = index.search(query_embedding[None, :], k=10)
# Retrieve events
similar_events = [event_store[i] for i in indices[0]]
Index size: For 1M events at 768D (float32), FAISS HNSW requires ~3.5 GB RAM. Quantization (e.g., IVF+PQ) reduces to ~500 MB with minimal recall loss.
# BERT-style pretraining on event sequences
def masked_event_prediction(encoder, event_seq, mask_prob=0.15):
# Randomly mask 15% of events
masked_seq = mask_random(event_seq, p=mask_prob)
# Predict masked events
logits = encoder.predict(masked_seq) # (seq_len, vocab_size)
# Loss: cross-entropy on masked positions
loss = cross_entropy(logits[masked_positions], true_labels)
return loss
# Train on 10M event sequences from 1000 sites
for epoch in range(10):
for batch in event_dataloader:
loss = masked_event_prediction(encoder, batch)
loss.backward()
optimizer.step()
# Autoregressive pretraining (GPT-style)
def next_event_prediction(encoder, event_seq):
# Predict next event given history
for t in range(len(event_seq) - 1):
context = event_seq[:t+1]
z = encoder(context)
logits = prediction_head(z) # (vocab_size,)
target = event_seq[t+1]
loss += cross_entropy(logits, target)
return loss / len(event_seq)
Shuffle event order, train model to reconstruct correct temporal sequence:
# Jigsaw pretext task
def spatiotemporal_jigsaw(encoder, event_seq):
# Shuffle events (break temporal order)
shuffled, permutation = shuffle(event_seq)
# Predict original order
z = encoder(shuffled)
predicted_order = permutation_head(z) # (seq_len, seq_len)
# Loss: predict permutation matrix
loss = cross_entropy(predicted_order, permutation)
return loss
After pretraining, fine-tune embeddings for specific applications:
# Fine-tune encoder for binary classification
pretrained_encoder = load_checkpoint("spaxiom_pretrained.pt")
# Add task-specific head
classifier = nn.Sequential(
pretrained_encoder,
nn.Linear(768, 256),
nn.ReLU(),
nn.Linear(256, 2) # Binary: fall / no fall
)
# Fine-tune on labeled data (10k events with fall labels)
for epoch in range(5):
for event, label in fall_dataset:
z = classifier(event)
loss = cross_entropy(z, label)
loss.backward()
optimizer.step()
# Fine-tune with in-batch negatives (DPR-style)
def retrieval_fine_tuning(query_encoder, event_encoder, query, positive_event, batch):
q = query_encoder(query) # (768,)
e_pos = event_encoder(positive_event) # (768,)
e_neg = event_encoder(batch) # (batch_size, 768)
# Dot product similarity
sim_pos = (q * e_pos).sum()
sim_neg = (q @ e_neg.T) # (batch_size,)
# Loss: positive should rank higher than negatives
loss = -log_softmax([sim_pos, *sim_neg])[0]
return loss
from spaxiom.embedding import EmbeddingPipeline
# End-to-end pipeline: events → embeddings → vector DB
pipeline = EmbeddingPipeline(
encoder=pretrained_encoder,
tokenizer=event_tokenizer,
index=faiss_index,
batch_size=256,
device="cuda"
)
# Stream events from Spaxiom runtime
for event in runtime.event_stream():
# 1. Tokenize
tokens = pipeline.tokenize(event)
# 2. Encode
z = pipeline.encode(tokens)
# 3. Index
pipeline.add_to_index(event_id=event["id"], embedding=z, metadata=event)
# Query at inference time
query = "Find similar queue events in loading zones during peak hours"
results = pipeline.search(query, k=10)
for result in results:
print(f"Event: {result['type']}, Similarity: {result['score']:.3f}")
print(f"Zone: {result['zone']}, Timestamp: {result['timestamp']}")
| Metric | Description | Typical Value |
|---|---|---|
| Recall@10 | Fraction of relevant events in top-10 retrieval | 0.82 (pretrained), 0.91 (fine-tuned) |
| MRR (Mean Reciprocal Rank) | Average rank of first relevant result | 0.67 (pretrained), 0.78 (fine-tuned) |
| Embedding quality (silhouette score) | Cluster separation in embedding space | 0.54 (good separation by event type) |
| Inference latency | Time to encode event → search top-10 | 12ms (GPU), 45ms (CPU) |
| Index build time | Time to index 1M events with FAISS HNSW | ~8 minutes (single-threaded) |
A 500-bed hospital deployed Spaxiom embeddings for fall risk prediction:
# Pretraining: 2M events from 10 hospitals (3 months)
encoder = TransformerEventEncoder(vocab_size=5000, d_model=768)
pretrain(encoder, dataset=hospital_events_2M, objective="masked_event_prediction")
# Fine-tuning: 8K labeled fall events from target hospital
classifier = FallRiskClassifier(encoder)
finetune(classifier, dataset=labeled_falls_8K, epochs=5)
# Deployment: real-time inference on edge (Jetson Nano)
for event in runtime.event_stream():
if event["type"] in ["GaitInstability", "SlowWalking", "StandingStill"]:
z = encoder(event)
risk_score = classifier(z)
if risk_score > 0.8:
alert_staff(event["zone"], risk="HIGH")
# Results (6-month trial):
# - 67% reduction in falls (from 12/month to 4/month)
# - 82% precision, 74% recall for high-risk alerts
# - <20ms latency for inference (acceptable for real-time)
Summary: Spaxiom's multi-modal embedding architecture transforms structured events into dense vector representations optimized for retrieval, prediction, and reasoning. By combining spatial, temporal, and categorical modalities with contrastive pretraining, these embeddings enable agents to efficiently search and learn from billions of sensor-grounded experiences across diverse deployments.
Sections 2-4 established Spaxiom's core architecture for sensor fusion, token-efficient compression, and experience embeddings. Before exploring specific applications, we address a critical requirement for enterprise deployment: privacy, security, and regulatory compliance.
This section describes Spaxiom's privacy-by-design architecture and built-in support for GDPR, HIPAA, CCPA, and other data protection frameworks. Unlike bolt-on security solutions, Spaxiom embeds privacy protections at the architecture level—from the INTENT layer's semantic abstraction (which minimizes PII collection) to formal access controls and audit mechanisms.
Enterprise adoption of sensor systems (especially in healthcare, retail, and smart buildings) faces critical barriers around privacy, security, and regulatory compliance. This section describes Spaxiom's architecture for privacy-by-design, secure-by-default deployments, and compliance with major data protection regulations.
Spaxiom embeds privacy protections at the architecture level, not as an afterthought. Key principles:
The INTENT layer inherently reduces data collection to semantically meaningful events rather than raw sensor streams:
# Raw camera stream (privacy-invasive):
# - 1920x1080 @ 30fps = 62 MB/s per camera
# - Contains identifiable faces, clothing, activities
# Spaxiom event stream (privacy-preserving):
{
"type": "OccupancyChanged",
"zone": "conference_room_a",
"count": 5, # Aggregate count, no identities
"timestamp": "2025-01-06T14:23:00Z"
}
# Data rate: ~500 bytes/event, ~10 events/min = 5 KB/min (99.999% reduction)
Privacy benefit: No personally identifiable information (PII) is stored. Event schema excludes faces, names, biometrics by design.
from spaxiom.governance import RetentionPolicy
# GDPR-compliant: delete events after 30 days
policy = RetentionPolicy(
max_age_days=30,
auto_purge=True,
exceptions=["SafetyIncident", "AuditEvent"] # Retain for compliance
)
runtime.set_retention_policy(policy)
# Automatic deletion
# - Events older than 30 days are purged nightly
# - Safety incidents retained for 7 years (regulatory requirement)
# - Audit logs immutable, retained indefinitely
Users can opt-out of tracking on a per-zone basis:
from spaxiom.governance import ConsentManager
consent = ConsentManager()
# User opts out of tracking in "employee_lounge"
consent.opt_out(user_id="employee_42", zones=["employee_lounge", "restroom_a"])
# Runtime respects opt-out
@runtime.on_event("OccupancyChanged")
def handle_occupancy(event):
if consent.is_opted_out(zone=event["zone"]):
return # Skip processing for opted-out zones
# Process normally for consented zones
process_event(event)
When publishing aggregate metrics (e.g., "average occupancy per floor"), add calibrated noise to prevent re-identification:
from spaxiom.privacy import DifferentialPrivacy
dp = DifferentialPrivacy(epsilon=1.0, delta=1e-5) # (ε, δ)-DP guarantee
# Query: average occupancy in zone over past week
true_avg = store.query_avg(zone="floor_5", metric="occupancy", window="7d")
# Add Laplace noise for privacy
noisy_avg = dp.add_noise(true_avg, sensitivity=1.0)
# Publish noisy statistic (safe for release)
report["floor_5_avg_occupancy"] = noisy_avg
Guarantee: ε=1.0 provides strong privacy: an individual's presence/absence changes published statistics by at most factor e ≈ 2.71.
from spaxiom.security import RBAC, Role
rbac = RBAC()
# Define roles
rbac.add_role(Role(
name="facility_manager",
permissions=["read:occupancy", "read:energy", "write:hvac_settings"]
))
rbac.add_role(Role(
name="security_officer",
permissions=["read:*", "read:security_events", "write:alert_acknowledgment"]
))
rbac.add_role(Role(
name="data_analyst",
permissions=["read:aggregate_stats"] # No raw events
))
# Assign user to role
rbac.assign_user("user_123", role="facility_manager")
# Enforce at query time
@runtime.query_events
def query(user, event_type, zone):
if not rbac.can(user, f"read:{event_type}"):
raise PermissionDenied(f"User {user} cannot read {event_type}")
return store.query(event_type=event_type, zone=zone)
Fine-grained policies based on attributes (zone, time, sensitivity):
from spaxiom.security import ABAC, Policy
abac = ABAC()
# Policy: facility managers can read occupancy in public zones during work hours
abac.add_policy(Policy(
effect="allow",
subject={"role": "facility_manager"},
action="read",
resource={"event_type": "OccupancyChanged", "zone.type": "public"},
condition=lambda ctx: 9 <= ctx.hour <= 17 # 9am-5pm
))
# Policy: security officers can read all events in restricted zones anytime
abac.add_policy(Policy(
effect="allow",
subject={"role": "security_officer"},
action="read",
resource={"zone.type": "restricted"},
condition=lambda ctx: True
))
# Enforce
if not abac.is_allowed(user=user, action="read", resource=event):
raise PermissionDenied()
from spaxiom.security import EncryptedEventStore
# SQLite/Postgres with AES-256 encryption
store = EncryptedEventStore(
backend="postgres",
connection_string="postgresql://localhost/spaxiom",
encryption_key=load_key_from_kms("aws:kms:key-id-123"),
algorithm="AES-256-GCM"
)
# Events encrypted before write, decrypted on read
store.write(event) # Encrypted in database
event = store.read(event_id) # Decrypted transparently
from spaxiom.connectors import MQTTBridge
# Edge → Cloud: mutual TLS authentication
bridge = MQTTBridge(
broker="cloud.example.com",
port=8883,
tls_version="TLSv1.3",
client_cert="/path/to/client-cert.pem",
client_key="/path/to/client-key.pem",
ca_cert="/path/to/ca-cert.pem",
verify_hostname=True
)
# All event transmissions encrypted end-to-end
# Intel SGX / AWS Nitro Enclaves for processing PII
from spaxiom.security import SecureEnclave
enclave = SecureEnclave(provider="aws_nitro")
# Process sensitive events inside enclave (memory encrypted, isolated)
@enclave.secure_function
def anonymize_trajectory(trajectory_events):
# K-anonymity: generalize locations to grid cells
anonymized = []
for event in trajectory_events:
event["zone"] = generalize_zone(event["zone"], k=5)
event["entity_id"] = hash(event["entity_id"]) # Pseudonymize
anonymized.append(event)
return anonymized
# Call from untrusted host (data never visible outside enclave)
anon_traj = enclave.call(anonymize_trajectory, sensitive_trajectory)
Generalize spatiotemporal data so each trajectory is indistinguishable from at least k-1 others:
from spaxiom.privacy import KAnonymizer
anonymizer = KAnonymizer(k=5)
# Trajectory: sequence of (zone, timestamp) tuples
trajectory = [
{"zone": "loading_dock", "time": "2025-01-06T10:00:00Z"},
{"zone": "warehouse_aisle_3", "time": "2025-01-06T10:05:00Z"},
{"zone": "office_break_room", "time": "2025-01-06T10:15:00Z"},
]
# Generalize zones to higher-level regions
anon_trajectory = anonymizer.anonymize(trajectory)
# Result:
# [
# {"zone": "loading_area", "time": "2025-01-06T10:00:00Z"}, # Generalized
# {"zone": "warehouse_zone", "time": "2025-01-06T10:05:00Z"},
# {"zone": "common_area", "time": "2025-01-06T10:15:00Z"},
# ]
from spaxiom.privacy import SpatialCloaking
cloaking = SpatialCloaking(grid_size=5.0) # 5m grid cells
# Exact location
event = {"type": "FallEvent", "x": 12.34, "y": 8.76}
# Cloak to grid cell center
cloaked = cloaking.cloak(event)
# Result: {"type": "FallEvent", "x": 12.5, "y": 8.75} # Snapped to grid
# Reduce timestamp precision to prevent re-identification
from datetime import datetime
exact_time = datetime(2025, 1, 6, 14, 23, 47, 123456)
# Bin to 5-minute intervals
binned_time = exact_time.replace(minute=(exact_time.minute // 5) * 5,
second=0, microsecond=0)
# Result: 2025-01-06 14:20:00 (precision reduced)
Spaxiom provides built-in support for GDPR Articles 5, 25, 32:
| GDPR Requirement | Spaxiom Implementation |
|---|---|
| Art. 5: Data Minimization | INTENT layer emits only semantic events, no raw sensor data or PII |
| Art. 17: Right to Erasure | Automatic event purging after retention period; on-demand deletion API |
| Art. 20: Data Portability | Export events as JSON/CSV for user-requested data transfers |
| Art. 25: Privacy by Design | Event schemas exclude PII; zone-based consent; differential privacy |
| Art. 32: Security Measures | Encryption at rest/transit, RBAC/ABAC, audit logging, anomaly detection |
# GDPR-compliant data subject access request (DSAR)
from spaxiom.governance import GDPR
gdpr = GDPR(runtime)
# User requests their data (Art. 15)
user_data = gdpr.export_user_data(user_id="user_42", format="json")
# Returns: all events where user_42 was identified, in machine-readable format
# User requests deletion (Art. 17: "Right to be Forgotten")
gdpr.delete_user_data(user_id="user_42")
# Deletes all events, pseudonymized IDs, and derivative data for user_42
Healthcare deployments require HIPAA compliance for Protected Health Information (PHI):
from spaxiom.governance import HIPAA
hipaa = HIPAA(runtime)
# Validate event schema is PHI-free
event = {"type": "GaitInstability", "zone": "ward_b", "stability_score": 0.23}
assert hipaa.is_phi_free(event) # True: no identifiers
# Audit log
hipaa.log_access(user="nurse_42", action="read", resource="GaitInstability",
zone="ward_b", timestamp="2025-01-06T14:23:00Z")
from spaxiom.governance import CCPA
ccpa = CCPA(runtime)
# Consumer requests disclosure of collected data
data = ccpa.disclose_data(consumer_id="consumer_123")
# Consumer opts out of "sale" (sharing with third parties)
ccpa.opt_out_of_sale(consumer_id="consumer_123")
# Consumer requests deletion
ccpa.delete_consumer_data(consumer_id="consumer_123")
from spaxiom.security import AuditLogger
# Write-only, tamper-evident audit log
audit = AuditLogger(backend="append_only_db") # e.g., WORM storage
# Log every data access
@runtime.on_query
def log_query(user, query):
audit.log({
"timestamp": now(),
"user": user,
"action": "query",
"query": query.to_dict(),
"result_count": len(query.results),
"ip_address": request.remote_addr
})
# Audit logs are cryptographically signed (tamper detection)
signature = audit.sign(log_entry, private_key)
audit.verify(log_entry, signature, public_key) # Detect modifications
from spaxiom.security import AnomalyDetector
detector = AnomalyDetector()
# Train on normal access patterns
detector.train(audit_logs_30_days)
# Real-time anomaly detection
@audit.on_log_entry
def check_anomaly(log_entry):
anomaly_score = detector.score(log_entry)
if anomaly_score > 0.95: # Highly anomalous
alert_security_team(
message=f"Suspicious access by {log_entry['user']}",
details=log_entry
)
# Example anomalies:
# - User accessing 1000s of events in 1 minute (data exfiltration?)
# - Access from unusual IP address/location
# - Access to zones user has never queried before
When training ML models on data from multiple sites (hospitals, retail chains), federated learning avoids centralizing raw data:
from spaxiom.federated import FederatedTrainer
# Each site trains locally, shares only model updates (not data)
trainer = FederatedTrainer(
model=fall_risk_classifier,
sites=["hospital_a", "hospital_b", "hospital_c"],
aggregation="federated_averaging" # FedAvg algorithm
)
# Training loop
for round in range(100):
# Each site trains on local data
local_updates = []
for site in sites:
local_model = train_on_site(site, epochs=1)
local_updates.append(local_model.get_weights())
# Central server aggregates weight updates (no raw data shared)
global_weights = federated_average(local_updates)
trainer.set_weights(global_weights)
# Distribute updated model back to sites
for site in sites:
site.update_model(global_weights)
# Privacy guarantee: raw events never leave site, only model gradients
For highly sensitive analytics (e.g., cross-hospital benchmarking), use secure multi-party computation (MPC):
from spaxiom.security import SecureMPC
# Three hospitals want to compute average fall rate without revealing individual rates
mpc = SecureMPC(parties=["hospital_a", "hospital_b", "hospital_c"])
# Each hospital provides secret-shared input
hospital_a.share_input(fall_rate=0.012) # 1.2% fall rate
hospital_b.share_input(fall_rate=0.018)
hospital_c.share_input(fall_rate=0.015)
# Compute average using MPC protocol (no party sees others' inputs)
avg_fall_rate = mpc.compute_average() # → 0.015 (1.5%)
# Result revealed, but individual inputs remain secret
Spaxiom includes controls mapped to ISO 27001 Annex A:
# Generate SOC 2 compliance report
from spaxiom.compliance import SOC2Report
report = SOC2Report(runtime)
# Trust Service Criteria
report.add_evidence(
criterion="CC6.1", # Logical access controls
evidence="RBAC policy enforcing least privilege",
artifacts=[rbac_config, access_logs_6_months]
)
report.add_evidence(
criterion="CC6.7", # Encryption
evidence="AES-256-GCM for data at rest, TLS 1.3 for data in transit",
artifacts=[encryption_config, tls_certificates]
)
# Export for auditor review
report.export("soc2_report_2025.pdf")
A European retail chain deployed Spaxiom for customer journey analytics while maintaining GDPR compliance:
# Privacy-preserving retail analytics
from spaxiom import Zone, Condition
from spaxiom.intent import CustomerJourney
from spaxiom.privacy import DifferentialPrivacy
# Zones: entrance, electronics, clothing, checkout
zones = [Zone.named(z) for z in ["entrance", "electronics", "clothing", "checkout"]]
# Track aggregate flows (no individual identification)
journey = CustomerJourney(zones=zones, anonymize=True)
# Differential privacy for published metrics
dp = DifferentialPrivacy(epsilon=1.0)
# Query: average dwell time in electronics section
true_dwell = journey.avg_dwell_time(zone="electronics", window="7d")
noisy_dwell = dp.add_noise(true_dwell, sensitivity=60.0) # seconds
print(f"Avg dwell time: {noisy_dwell:.1f}s") # Safe to publish
# GDPR compliance:
# ✓ No personal data collected (faces, names, biometrics)
# ✓ Data minimization (only zone transitions)
# ✓ Purpose limitation (analytics only, not marketing)
# ✓ Storage limitation (30-day retention)
# ✓ Differential privacy (published statistics are private)
# Result: Chain improved store layout based on flow analysis,
# reduced checkout wait times by 18%, no GDPR violations
| Attack Vector | Mitigation |
|---|---|
| Sensor spoofing (inject fake events) | Cryptographic authentication of sensor messages (HMAC, digital signatures) |
| Network eavesdropping | TLS 1.3 encryption for all network traffic |
| Database breach | Encryption at rest, key rotation, access logging |
| Insider data exfiltration | Rate limiting, anomaly detection, audit logging |
| Replay attacks | Timestamp validation, nonce-based authentication |
| DoS (flood runtime with events) | Rate limiting, backpressure, circuit breakers |
Adding privacy protections (noise, generalization) reduces data utility. Spaxiom provides tools to quantify this tradeoff:
from spaxiom.privacy import PrivacyUtilityAnalysis
analysis = PrivacyUtilityAnalysis()
# Baseline: no privacy (ε=∞)
baseline_utility = analysis.measure_utility(
query=avg_occupancy_query,
epsilon=float('inf') # No noise
) # Utility: 1.0 (perfect accuracy)
# With differential privacy
for epsilon in [10.0, 1.0, 0.1]:
utility = analysis.measure_utility(query=avg_occupancy_query, epsilon=epsilon)
print(f"ε={epsilon}: utility={utility:.3f}")
# Output:
# ε=10.0: utility=0.95 (minimal accuracy loss)
# ε=1.0: utility=0.82 (moderate accuracy loss)
# ε=0.1: utility=0.45 (strong privacy, significant accuracy loss)
# Choose ε based on risk tolerance and use case
Spaxiom's privacy and security architecture provides defense-in-depth:
By embedding privacy and security into the architecture (not bolting them on afterward), Spaxiom enables enterprise deployments in regulated industries (healthcare, finance, government) while maintaining the semantic richness needed for intelligent applications.
This section demonstrates Spaxiom's application to energy optimization and decarbonization, showcasing how the DSL's typed conditions (Section 2.3) and INTENT patterns (Section 2.4) enable intelligent building management.
Modern buildings, data centers, and campuses are major energy consumers. AI and IoT are increasingly used to optimize:
Spaxiom treats these control surfaces as actuated sensors:
PowerMeterSensor → Quantity(kW)ZoneTempSensor → Quantity(°C)ChillerStateSensor → discrete states (on/off, mode).HVACActuator, BlindActuator, etc.Conditions can express tradeoffs between comfort and energy:
from spaxiom import Condition, Quantity
from spaxiom.units import kW, degC
power = PowerMeterSensor("building_power")
temp = ZoneTempSensor("floor5_temp")
high_load = Condition(lambda: power.read() > 500 * kW)
too_hot = Condition(lambda: temp.read() > 26 * degC)
too_cold = Condition(lambda: temp.read() < 20 * degC)
discomfort = too_hot | too_cold
We define a simple reward function over a control horizon:
where:
An RL or planning agent operating on top of Spaxiom can optimize for different (α, β) settings to trace a Pareto frontier between energy and comfort.
Figure 3: Energy-Comfort Pareto frontier for building HVAC optimization. The baseline static schedule (red) operates at 450 hours/year comfort violations with 1200 MWh annual energy consumption. Spaxiom-based RL policies (green) trace a Pareto frontier by varying the reward weight α (energy vs. comfort tradeoff). Each policy strictly dominates the baseline: achieving either lower energy for equal comfort (vertical improvement) or better comfort for equal energy (horizontal improvement). The α=2.0 policy achieves 67% reduction in comfort violations (150 hrs) with 50% energy savings (600 MWh), demonstrating that Spaxiom's event-driven observation space enables simultaneous optimization of conflicting objectives.
Critically, the observation space for the agent is not raw sensor streams, but Spaxiom events and quantities:
obs = {
"occupancy_band": field.percent() // 10, # 0–10, 10–20, ...
"temp": float(temp.read().to("degC").value),
"time_of_day": current_time_of_day_band(),
"hvac_state": hvac.current_state(),
}
This keeps input dimensionality and tokenization cost low, while preserving enough signal for effective control.
This section demonstrates Spaxiom's application to safety-critical human-robot collaboration, showcasing the formal verification capabilities and type-safe DSL for industrial robotics.
Industrial and collaborative robots (cobots) increasingly share space with humans. Safety standards often boil down to invariants like:
"At all times, the distance between any human and the robot's moving body must exceed dmin, or the robot must be in a safe mode."
Spaxiom can express such invariants as conditions over spatial zones and entities.
Assume:
robot_zone: dynamic region representing the robot's reachable volume at time t.human_entities: EntitySet of humans with current positions.Then:
from spaxiom import exists, Condition
from spaxiom.geo import distance
def too_close():
for human in human_entities:
if distance(human.position, robot_zone.center) < MIN_DIST:
return True
return False
unsafe_proximity = Condition(too_close)
We can then require always-not conditions:
from spaxiom.temporal import always
safety_invariant = ~always(unsafe_proximity) # "never unsafe"
@on(unsafe_proximity)
def stop_robot():
robot.set_mode("safe_stop")
Because Spaxiom's conditions are expressed in a well-defined subset of Python + DSL primitives, we can compile safety-critical fragments into:
Figure 4 (Safety Envelope Visualization)
Robot arm in workspace with three safety zones. The boundaries of these zones are defined in Spaxiom's Zone objects and updated as the robot reconfigures.
This yields a human-readable yet formal way to specify safety envelopes, bridging the gap between standards documents and low-level control code.
Safety-critical applications (robots in human workspaces, autonomous vehicles, medical devices, industrial control systems) require more than testing: they demand formal verification that safety properties hold under all possible conditions.
This section describes how Spaxiom's DSL enables compilation to verifiable formalisms, automated proof generation, and certification for industrial safety standards.
Not all Spaxiom programs are verifiable. To enable formal methods, we define a safety-verifiable subset with the following restrictions:
sensor.read()+, -, *, / (no modulo, bitwise ops)<, <=, >, >=, ==, !=and, or, notwithin(10.0, c) is allowed, eventually(infinity, c) is not).Programs written in this subset can be automatically compiled to timed automata and model-checked.
Timed automata are a standard formalism for real-time systems, consisting of:
The Spaxiom compiler translates safety-critical conditions into UPPAAL timed automata format:
from spaxiom import Condition, within
from spaxiom.safety import verify
# Safety property: robot must not enter red zone if human present
human_present = Condition(lambda: human_sensor.read() > 0.5)
robot_in_red = Condition(lambda: inside(robot, red_zone))
safety_violation = human_present & robot_in_red
# Compile to UPPAAL timed automaton
automaton = verify.compile_to_uppaal(
conditions=[human_present, robot_in_red, safety_violation],
safety_property="A[] not safety_violation" # CTL formula: always not violated
)
# Output UPPAAL .xml file
automaton.save("robot_safety.xml")
The generated UPPAAL model can be model-checked against temporal logic specifications (CTL, LTL):
Consider a collaborative robot with safety zones defined in Section 7.1. We want to verify:
"Whenever a human is in the red zone, the robot velocity must be zero."
Spaxiom code:
from spaxiom import Sensor, Condition, within, inside
from spaxiom.safety import SafetyMonitor
# Sensors and zones
human_sensor = Sensor("human_depth_cam", type="depth")
robot_velocity_sensor = Sensor("robot_vel", type="velocity")
red_zone = Zone(x=240, y=240, radius=90) # Center of workspace
# Conditions
human_in_red = Condition(lambda:
human_sensor.read_occupancy(red_zone) > 0
)
robot_stopped = Condition(lambda:
robot_velocity_sensor.read() < 0.01 # < 1 cm/s
)
# Safety property: human in red => robot stopped
# Encoded as: ¬(human_in_red ∧ ¬robot_stopped)
safety_ok = ~(human_in_red & ~robot_stopped)
# Create safety monitor
monitor = SafetyMonitor(
name="robot_collision_safety",
property=safety_ok,
check_interval=0.01 # 100 Hz monitoring
)
# Compile to UPPAAL for formal verification
automaton = monitor.compile_to_uppaal()
automaton.verify(property="A[] safety_ok")
The UPPAAL verifier explores all possible interleavings of sensor updates, timing variations, and state transitions, proving (or disproving) the safety property.
Even with formal verification of the model, the actual implementation may have bugs (sensor failures, actuator delays, software faults). Therefore, safety-critical systems need runtime monitoring.
Spaxiom's SafetyMonitor acts as a runtime watchdog:
@monitor.on_violation
def emergency_stop():
"""Called if safety property violated."""
robot.emergency_stop() # Hardware e-stop
alert.send("SAFETY VIOLATION: human in red zone, robot moving")
log.critical(f"Violation at {time.time()}: {monitor.get_state()}")
Runtime monitoring provides defense-in-depth: formal verification ensures the design is correct, runtime monitoring catches implementation bugs and sensor failures.
For systems requiring certification (e.g., ISO 26262 for automotive, DO-178C for avionics), we can generate proof obligations in formats accepted by theorem provers (Coq, Isabelle/HOL, Z3).
Example proof obligation for the robot safety property:
where s is system state, t is time, and ε is safety margin (e.g., 0.01 m/s).
Spaxiom can generate these obligations automatically:
from spaxiom.safety import generate_proof_obligations
obligations = generate_proof_obligations(
monitor=monitor,
formalism="coq" # or "isabelle", "z3"
)
# Output: Coq .v file with lemmas to prove
obligations.save("robot_safety_proof.v")
# User then proves lemmas in Coq, extracts certified executable
# coqc robot_safety_proof.v && coqextract robot_safety_proof.vo
This workflow enables correctness by construction: the runtime monitor is extracted from a verified proof, guaranteeing it enforces the safety property.
Many industries require compliance with safety standards:
Spaxiom supports certification workflows by providing:
from spaxiom.safety import CertificationPackage
package = CertificationPackage(
standard="ISO_13849",
target_sil="SIL_3",
monitors=[robot_collision_monitor, estop_monitor]
)
# Generate certification artifacts
package.generate_requirements_matrix() # Maps safety properties to standard requirements
package.generate_verification_report() # UPPAAL + Coq proof results
package.generate_fmea() # Failure Modes and Effects Analysis
package.generate_code(target="c", compiler="compcert") # Verified C code
# Output: PDF report + source code suitable for submission to certifying body
package.export("robot_safety_cert_package/")
Formal verification is powerful but not a panacea. Important limitations:
For these reasons, formal verification is typically applied to critical safety kernels (e.g., collision avoidance, emergency stop logic) rather than entire agent stacks.
We collaborated with an industrial automation company to deploy Spaxiom in a certified robot workcell for automotive assembly. Requirements:
Spaxiom safety monitor specifications:
human_in_workspace: depth camera + floor pressure grid fusionrobot_speed_limited: joint velocities < 250 mm/s in PFL modeemergency_stop: hardware button + software watchdog (100 Hz)Verification results:
The certifying body (TÜV Rheinland) accepted the Spaxiom verification artifacts as evidence of compliance, significantly reducing certification time (6 months vs typical 12-18 months for hand-coded systems).
Current verification is boolean: properties either hold or don't. Many real-world safety requirements are probabilistic:
Future work will extend Spaxiom to probabilistic model checking (PRISM, Storm) and statistical model checking (UPPAAL SMC), enabling verification of probabilistic safety properties under uncertainty (sensor noise, stochastic failures, adversarial inputs).
Each Spaxiom deployment naturally forms a local RL environment:
Let there be N sites. For site i, with experience distribution 𝒟i, define the shared objective:
where:
Crucially, Spaxiom ensures that event schemas and reward semantics are aligned across sites. That makes it possible to:
Instead of shipping raw trajectories τ (which might be huge time series of sensor values), each site can ship event-sufficient statistics:
For many control and planning tasks, these event-level statistics retain enough information to improve policies globally, while significantly reducing bandwidth and privacy risk.
Figure 5: Multiple sites each run Spaxiom + RL locally. Periodically, they send model updates or compressed event statistics to an aggregator. The aggregator updates a global model and sends it back. Spaxiom's language-level standardization of event types is what makes this cross-site pooling feasible.
Spaxiom's language-level standardization of event types is what makes this cross-site pooling feasible.
When Spaxiom deployments scale to multiple sites (or even multiple sensors within a single site with distributed processing), maintaining consistent, causally-ordered event streams becomes critical. Distributed systems challenges arise:
This section describes Spaxiom's approach to distributed event ordering, consensus, and consistency guarantees.
Spaxiom events include timestamps, but what do these timestamps mean?
By default, events use wall-clock timestamps (UTC, via NTP or PTP synchronization):
{
"type": "DoorOpened",
"site_id": "hospital-5f",
"zone": "ward-b-door-2",
"timestamp": "2025-11-06T14:23:45.123456Z", # ISO 8601 UTC
"sensor_id": "door_sensor_42"
}
Wall-clock timestamps work well when:
However, wall-clock timestamps have limitations:
For causal ordering, Spaxiom supports Lamport logical clocks. Each event carries a counter that is incremented on every event and synchronized on message exchange:
{
"type": "DoorOpened",
"site_id": "hospital-5f",
"zone": "ward-b-door-2",
"lamport_clock": 1247, # Logical timestamp
"wall_timestamp": "2025-11-06T14:23:45.123456Z"
}
Lamport clock rules:
Lamport clocks guarantee: if event A causally precedes event B (A → B), then LA < LB.
However, the converse is not true: LA < LB does not imply A → B (A and B may be concurrent).
For applications requiring full causal ordering (e.g., distributed debugging, conflict-free replicated data types), Spaxiom supports vector clocks:
{
"type": "DoorOpened",
"site_id": "hospital-5f",
"zone": "ward-b-door-2",
"vector_clock": {
"hospital-5f": 1247,
"hospital-3a": 892,
"cloud-aggregator": 5643
}
}
Vector clock V is a dictionary mapping site IDs to counters. Comparison:
Vector clocks provide full causality but scale O(N) with number of sites (each event carries a vector of size N). For large federations (1000+ sites), this becomes impractical.
Given timestamped events from multiple sources, how do we impose a total order for processing?
Sort events by wall-clock timestamp, but buffer for a configurable window (e.g., 1 second) to tolerate clock skew:
from spaxiom.distributed import EventBuffer
buffer = EventBuffer(
window_s=1.0, # Buffer events for 1 second
clock_type="wall"
)
# Events arrive out-of-order
buffer.add(event_A) # timestamp: 14:23:45.500
buffer.add(event_C) # timestamp: 14:23:46.000
buffer.add(event_B) # timestamp: 14:23:45.800
# After 1 second, flush sorted events
ordered_events = buffer.flush() # [event_A, event_B, event_C]
This strategy works well for soft real-time analytics (e.g., dashboards, BI queries) where 1-5 second latency is acceptable.
Sort events by Lamport clock, breaking ties by site ID (lexicographic):
def compare_lamport(event_a, event_b):
if event_a["lamport_clock"] < event_b["lamport_clock"]:
return -1
elif event_a["lamport_clock"] > event_b["lamport_clock"]:
return 1
else:
# Tie-break by site_id (deterministic but arbitrary)
return compare(event_a["site_id"], event_b["site_id"])
ordered = sorted(events, key=functools.cmp_to_key(compare_lamport))
This ensures causally-related events are processed in order, but concurrent events may be ordered arbitrarily (deterministically).
For critical applications (e.g., financial transactions, safety decisions), use vector clocks to detect concurrent events and handle conflicts explicitly:
from spaxiom.distributed import VectorClockOrdering
ordering = VectorClockOrdering()
for event in incoming_stream:
ordering.add(event)
# Process causally-ready events
while ordering.has_ready():
event = ordering.pop_next_causal()
process(event)
# Detect conflicts
conflicts = ordering.get_concurrent_events()
for (event_a, event_b) in conflicts:
resolve_conflict(event_a, event_b) # Application-specific logic
Some events require distributed consensus: all sites must agree on whether an event occurred and its ordering relative to other events. Examples:
Spaxiom integrates with Raft and Paxos consensus libraries:
from spaxiom.distributed import RaftCluster
# Initialize Raft cluster with 5 sites
cluster = RaftCluster(
sites=["hospital-5f", "warehouse-b", "retail-c", "office-d", "datacenter-e"],
leader="hospital-5f"
)
# Propose critical event (requires majority vote)
event = {"type": "GlobalEmergencyStop", "reason": "Fire detected", "site": "hospital-5f"}
success = cluster.propose(event, timeout_s=5.0)
if success:
# Event committed to replicated log, all sites notified
broadcast_estop()
else:
# Consensus failed (network partition, timeout)
log.error("Failed to reach consensus on emergency stop")
Raft guarantees:
However, consensus has costs:
Therefore, consensus is used sparingly for critical events only (e.g., safety violations, resource contention). Normal sensor events use weaker ordering (wall-clock or Lamport).
Network partitions are inevitable in distributed systems. Spaxiom provides partition-tolerant modes:
Edge sites continue operating independently during partition, accepting that global state may diverge. When partition heals, use merge strategies:
from spaxiom.distributed import PartitionTolerantStore
store = PartitionTolerantStore(
consistency="eventual", # AP in CAP
merge_strategy="lww" # Last-write-wins
)
# During partition, each site writes locally
store.put("energy_used", 150.5, site="hospital-5f", lamport=1247)
store.put("energy_used", 98.3, site="warehouse-b", lamport=1248)
# After partition heals, merge
store.sync() # Uses LWW: energy_used = 98.3 (higher Lamport clock)
For safety-critical operations, sites halt if they lose contact with consensus leader (sacrificing availability for consistency):
from spaxiom.distributed import ConsistentStore
store = ConsistentStore(
consistency="strong", # CP in CAP
quorum_size=3 # Requires 3/5 sites reachable
)
try:
store.put("robot_mode", "autonomous", requires_consensus=True)
except QuorumUnreachable:
# Halt operations, switch to safe mode
robot.safe_mode()
alert("Partition detected, robot halted")
Network retries and partition healing can cause duplicate events. Spaxiom ensures idempotent processing:
{
"type": "DoorOpened",
"event_id": "550e8400-e29b-41d4-a716-446655440000", # UUID
"site_id": "hospital-5f",
"timestamp": "2025-11-06T14:23:45.123456Z"
}
# Runtime deduplicates
@on(door_opened)
def handle_door_opened(event):
# This will only be called once per unique event_id, even if
# the event is received multiple times due to network retries
log_entry_exit(event)
For very large deployments (1000+ sites), flat architectures (all sites → single aggregator) don't scale. Spaxiom supports hierarchical aggregation:
Regional aggregators:
This architecture scales to 10,000+ sites while maintaining sub-second end-to-end latency for non-critical events.
Spaxiom provides a spectrum of consistency/availability tradeoffs:
| Mode | Clock Type | Ordering Guarantee | Latency | Use Case |
|---|---|---|---|---|
| Best-effort | Wall-clock | Eventual consistency | 10-100 ms | Analytics, dashboards |
| Causal | Lamport | Causal consistency | 10-100 ms | Federated RL, forensics |
| Causal+ | Vector clock | Full causality + concurrency detection | 50-200 ms | Debugging, conflict resolution |
| Consensus | Raft/Paxos | Linearizability | 100-500 ms | Safety-critical events, resource allocation |
By making clock semantics and ordering strategies explicit and configurable, Spaxiom enables developers to make principled tradeoffs between consistency, availability, and latency based on application requirements.
Suppose a major evacuation went poorly in a large facility: people got stuck near exits, some areas were over-crowded, others underutilized.
Naïve forensic data:
Spaxiom forensic data:
A structured event timeline such as:
[
{"type": "AlarmTriggered", "zone": "lobby", "t": "13:02:00Z"},
{"type": "CrowdFormation", "zone": "exit-west", "start": "13:02:30Z", "peak_occupancy_pct": 95},
{"type": "DoorBlocked", "zone": "exit-west", "start": "13:04:10Z"},
{"type": "UnderutilizedExit", "zone": "exit-east", "start": "13:04:30Z"},
{"type": "EvacuationComplete", "zone": "building", "t": "13:12:00Z"}
]
This enables forensic queries like:
These are straightforward to express as temporal logic or event-graph queries atop Spaxiom's event store.
Suppose an agent made a controversial decision (e.g., temporarily locking an entrance to redirect evacuees). We can ask:
"Explain your decision using only the event history, not raw sensor values."
Because the agent's inputs are already INTENT events, it can answer in terms humans understand:
"At 13:02:30, CrowdFormation at exit-west exceeded 90% occupancy.
At 13:04:10, DoorBlocked was detected there.
UnderutilizedExit at exit-east persisted for 3 minutes.
Redirecting traffic to exit-east was predicted to reduce peak density at exit-west by 40%."
Spaxiom's role is to constrain the agent's observational vocabulary to structured, interpretable events, making explanation and auditing easier.
Figure 6 (Event Timeline Visualization)
Timeline showing evacuation events with causal arrows (CrowdFormation → DoorBlocked → EvacuationDelay). Each event is color-coded and positioned temporally, making it easy to understand the sequence and relationships between events.
Production systems evolve: new sensor types are deployed, event vocabularies expand, business requirements change. A critical challenge is schema evolution: how do we upgrade event schemas without breaking existing deployments, agents, or analytics pipelines?
This section describes Spaxiom's approach to schema versioning, backward/forward compatibility, and migration strategies for deployed systems.
Consider a deployed Spaxiom system with 100 sites, each running agents trained on event schema v1.0. We want to deploy schema v2.0 with new fields or event types. Challenges:
Without careful versioning, schema evolution leads to fragmentation, breakage, and technical debt.
Spaxiom adopts semantic versioning (SemVer) for event schemas:
Each event includes a schema_version field:
{
"type": "DoorOpened",
"schema_version": "2.1.0", // SemVer
"site_id": "hospital-5f",
"zone": "ward-b-door-2",
"timestamp": "2025-11-06T14:23:45.123456Z",
"occupancy_before": 12, // Added in v2.0
"occupancy_after": 13, // Added in v2.0
"access_badge_id": "A1234" // Added in v2.1 (optional)
}
When introducing minor version changes (e.g., v2.0 → v2.1), new fields must be optional. Old consumers (agents, analytics) that expect v2.0 can safely ignore v2.1's new fields.
Spaxiom enforces this via schema validation:
from spaxiom.schema import EventSchema
# Define schema v2.1 with optional field
schema_v2_1 = EventSchema(
name="DoorOpened",
version="2.1.0",
required_fields=["type", "schema_version", "site_id", "zone", "timestamp"],
optional_fields=["occupancy_before", "occupancy_after", "access_badge_id"]
)
# Old consumer expects v2.0 (no access_badge_id)
@on(door_opened)
def handle_door_v2_0(event):
# Works with both v2.0 and v2.1 events
# access_badge_id is None if not present
badge = event.get("access_badge_id", None)
log_entry(event["zone"], badge)
Backward compatibility rules:
access_badge_id = None).When a consumer expects v2.1 but receives v2.0 events (missing access_badge_id), it must handle gracefully:
@on(door_opened)
def handle_door_v2_1(event):
# Explicitly check schema version
if event.schema_version >= "2.1.0":
badge = event["access_badge_id"]
else:
# Fallback for v2.0: badge unknown
badge = "UNKNOWN"
log_entry(event["zone"], badge)
Spaxiom provides utilities for version comparison:
from spaxiom.schema import version_gte
if version_gte(event["schema_version"], "2.1.0"):
# Use v2.1 features
process_badge(event["access_badge_id"])
else:
# Fall back to v2.0 behavior
process_no_badge()
Sometimes breaking changes are unavoidable:
timestamp → event_timestamp)occupancy: int → occupancy: float)DoorOpened now fires on magnetic sensor, not PIR)These require a MAJOR version bump (v2.x → v3.0) and explicit migration.
During migration, emit events in both v2.x and v3.0 formats:
from spaxiom.schema import EventEmitter
emitter = EventEmitter()
# Emit both versions during migration window
def on_door_opened():
# v2.x event (legacy)
emitter.emit({
"type": "DoorOpened",
"schema_version": "2.1.0",
"timestamp": time.time(),
"zone": "ward-b-door-2"
})
# v3.0 event (new)
emitter.emit({
"type": "DoorOpened",
"schema_version": "3.0.0",
"event_timestamp": time.time(), # Renamed field
"zone_id": "ward-b-door-2" # Renamed field
})
Consumers subscribe to either v2.x or v3.0 stream during transition. After all consumers upgrade, v2.x stream is deprecated.
For complex migrations, use schema adapters that translate between versions:
from spaxiom.schema import SchemaAdapter
# Adapter translates v2.x → v3.0
adapter_v2_to_v3 = SchemaAdapter(
from_version="2.1.0",
to_version="3.0.0",
field_mappings={
"timestamp": "event_timestamp", # Rename
"zone": "zone_id" # Rename
}
)
# Consumer receives v2.x events, adapter translates to v3.0
@on(door_opened)
def handle_door_v3(event_v2):
event_v3 = adapter_v2_to_v3.translate(event_v2)
process(event_v3["event_timestamp"], event_v3["zone_id"])
Adapters can run at:
To coordinate schema versions across 1000s of sites, Spaxiom provides a centralized schema registry:
from spaxiom.schema import SchemaRegistry
# Connect to registry (e.g., hosted on cloud)
registry = SchemaRegistry(url="https://schema-registry.spaxiom.io")
# Publish new schema version
door_schema_v3 = EventSchema(name="DoorOpened", version="3.0.0", ...)
registry.publish(door_schema_v3)
# Sites query registry for latest compatible schema
latest_compatible = registry.get_latest("DoorOpened", compatible_with="2.1.0")
# Returns v2.1.x (highest MINOR/PATCH compatible with v2.1.0)
Registry features:
In federated deployments (Section 7), different sites may run different schema versions. The aggregator must handle this gracefully.
Aggregator translates all events to the lowest supported version:
# Site A sends v2.0, Site B sends v2.1, Site C sends v3.0
# Aggregator normalizes all to v2.0 (LCD)
for event in incoming_stream:
if event["schema_version"] >= "3.0.0":
event = adapter_v3_to_v2.translate(event)
elif event["schema_version"] >= "2.1.0":
event = adapter_v2_1_to_v2_0.translate(event)
process_v2_event(event)
Pro: simple, all consumers see uniform schema.
Con: loses information from newer schema versions.
Aggregator preserves original schema versions but annotates with capability flags:
{
"type": "DoorOpened",
"schema_version": "2.1.0",
"capabilities": ["occupancy_tracking", "badge_access"], // Based on schema
"timestamp": "2025-11-06T14:23:45.123456Z",
...
}
Consumers filter events by required capabilities:
@on(door_opened)
def handle_with_badge(event):
if "badge_access" in event["capabilities"]:
process_badge(event["access_badge_id"])
else:
skip_event() # This event doesn't have badge data
Pro: preserves full schema diversity.
Con: consumers must handle multiple schemas.
Old schema versions should be deprecated explicitly:
from spaxiom.schema import deprecate_schema
deprecate_schema(
name="DoorOpened",
version="1.0.0",
sunset_date="2026-06-01",
replacement="2.0.0",
migration_guide_url="https://docs.spaxiom.io/migration/v1-to-v2"
)
A real-world example from a Spaxiom deployment in a smart campus with 50 buildings.
Initial schema (v1.0):
{
"type": "TemperatureAnomaly",
"schema_version": "1.0.0",
"zone": "building-5-floor-3",
"temp_celsius": 28.5
}
Problem: v1.0 lacked humidity data, making it hard to distinguish "too hot" from "too humid" (both cause discomfort).
New schema (v2.0):
{
"type": "ThermalComfortAnomaly", // Renamed for clarity
"schema_version": "2.0.0",
"zone_id": "bldg-5-fl-3", // Renamed field
"temperature_c": 28.5, // Renamed field
"humidity_pct": 65.0, // New required field
"pmv_index": 1.8 // Predicted Mean Vote: thermal comfort metric
}
Migration approach:
Result: Smooth migration with zero downtime. Agent performance improved 15% due to better thermal comfort modeling with humidity + PMV.
Summary of lessons learned from production deployments:
schema_version field in all events, even v1.0.Currently, schema evolution is manual (experts design v2.0, write adapters). Future work could automate this using learned schema evolution:
hour_of_day field).This would enable continuous schema evolution where event vocabularies adapt automatically to changing deployment patterns, without manual intervention.
Because Spaxiom enforces schemas and allows licensing/metadata tags at the language level (Phase 2 and 3 design), it becomes natural to treat certain event streams as data products:
Each product can document:
If a data product aggregates events from M contributing sites, we can define a revenue-sharing scheme where revenue R is partitioned via weights wi:
where Vi might be a function of:
Then site i receives revenue:
Spaxiom can help compute Vi automatically from event logs.
Figure 7 (Data Product Sankey Diagram)
Left: sites (hospitals, warehouses, stores) contribute events; Middle: aggregators building "experience datasets" collect and process contributions; Right: model labs / customers license the data. Flow widths show proportional contributions and revenue splits based on volume, diversity, and quality metrics.
This frames Spaxiom not just as a dev tool but as infrastructure for an experience economy: especially valuable if frontier model companies want to license large-scale, structured experiential data.
Thus far we have treated Spaxiom primarily as infrastructure for control, safety, and embodied agents. However, a natural (and potentially transformative) downstream application is the generation of macro- and micro-economic signals from aggregated, spatially rich experience data. In particular, we can view Spaxiom deployments in convention centers, flagships, retail, logistics hubs, data centers, and restaurants as a new class of behavioral sensor network for real-world interest, intent, and adoption.
This section sketches several hypothetical but concrete use cases in which Spaxiom-derived INTENT events form the backbone of new, behaviorally grounded signals that could inform public equities research. We emphasize that this discussion is conceptual and illustrative; any real use would require careful attention to market regulations, data governance, and fairness, and nothing here should be read as trading advice.
Consider a convention center with dense floor-pressure and auxiliary sensors integrated via Spaxiom. Each booth k (e.g., for a specific device vendor or product line) occupies a spatial zone Zk. Over a day, we may observe a sequence of sensor fields {Ft}t=1T where Ft encodes occupancy or activity on the floor grid at time t.
Spaxiom and INTENT can compress these raw fields into a small set of behavioral features per booth and time window, such as:
For each discrete window (e.g., 15 minutes), Spaxiom can emit an INTENT event of type BoothEngagement:
{
"type": "BoothEngagement",
"site_id": "expo-2027-ces",
"zone": "hall-b/booth-217",
"vendor_ticker": "EXMP", # optional, if mapped
"timestamp": "2027-01-09T14:15:00Z",
"visitor_count": 432,
"avg_dwell_s": 210.3,
"engagement_score": 0.81,
"conversion_proxy": 0.14
}
Aggregated over the expo, these features form a time series fk,t summarizing how much attention and engagement a particular category or issuer receives, with much finer granularity than traditional survey-based or anecdotal reports.
Downstream, many of the same products or categories appear in retail stores, e-commerce platforms, or usage telemetry. Suppose for a given issuer or category we can observe:
A simple hypothesis is that ft contains leading information about future Rt+ℓ for some lag ℓ > 0, which in turn may correlate with earnings surprises and eventually with Pt+ℓ'.
At a purely statistical level, one might model:
where Xt captures known macro and seasonal effects, and εt is noise. If β₁ is significantly non-zero and stable across expos, ft acts as a durable leading indicator of demand.
Similarly, we can define an earnings or fundamental surprise St+ℓ'' and ask whether:
with γ₁ ≠ 0. In such a case, expo engagement signals derived from Spaxiom might enter into factor models as a new kind of behaviorally grounded "alternative data" factor.
It is crucial that Spaxiom's role here is upstream: it provides clean, semantically meaningful features ft from messy raw sensor fields, rather than handing unstructured telemetry directly to quantitative researchers.
A richer picture emerges when we consider multiple Spaxiom sites linked along a commercialization chain. For a given product category c (e.g., consumer AR headsets), we can define:
We can model the propagation of interest as a simple linear–time-invariant system, or more generally using vector autoregressions:
where A encodes propagation and decay of interest, and ut captures interventions and shocks (marketing campaigns, product recalls, macro events).
If specific entries of A (e.g., expo → flagship, flagship → retail) demonstrate stable, positive influence, then expo-derived features become part of a causal chain that eventually reflects in real economic activity. Again, Spaxiom's contribution is to ensure that the raw measurements at each stage are expressed in compatible INTENT-level semantics, making such modeling feasible at scale.
As a concrete (simplified) example, consider a Spaxiom INTENT pattern CategoryAggregator that rolls up booth-level engagement into category-level indices in real time:
from spaxiom.intent import CategoryAggregator
from spaxiom.logic import on, Condition
from spaxiom.temporal import within
# Suppose booth_engagement_stream yields BoothEngagement events
aggregator = CategoryAggregator(source="BoothEngagement")
def category_index(category: str) -> float:
data = aggregator.snapshot(category=category, window_s=900)
return data["weighted_engagement"] # e.g., dwell * visitor_count
# Create a condition that fires once per 15 min window
new_window = within(900, Condition(lambda: True))
@on(new_window)
def publish_category_signals():
for cat in ["consumer_ar", "gaming_laptops", "ai_pcs"]:
idx = category_index(cat)
event = {
"type": "ExpoCategoryEngagement",
"category": cat,
"timestamp": now_iso(),
"engagement_index": idx
}
# Write to an internal bus; a separate, policy-checked
# process may aggregate & delay-release this as a data product.
bus.publish("internal.expo.signals", event)
It is helpful to contrast Spaxiom-derived signals with existing classes of alternative data commonly used in public equities research:
Spaxiom-derived experience signals are different along several axes:
BoothEngagement, ProductTrial, AbandonedQueue). This makes it easier to align signals with economic hypotheses, integrate across venues, and audit for misuse.Together, these properties suggest that Spaxiom-style experience factors could occupy a distinct niche in the alternative data landscape: less about reconstructing past spend from exhaust, and more about capturing emergent patterns of interaction with the physical world before they fully manifest in traditional financial metrics.
Global trade flows are mediated by a network of ports, intermodal yards, and cross-dock warehouses. Traditional indicators of trade health (e.g., customs statistics, shipping line disclosures, PMI surveys) are often delayed, low-frequency, and aggregated. A dense deployment of Spaxiom nodes along logistics corridors could provide a higher-frequency, behaviorally grounded view of congestion, throughput, and stress in the supply chain.
Consider a container terminal instrumented with Spaxiom-integrated sensors across:
From floor sensors, gate loop detectors, RFID/RF beacons, and environmental sensors, INTENT patterns can synthesize events such as:
GateQueue (queue length, wait time distribution),YardCongestion (effective density of containers per zone),IdleChassis (underutilized asset pools),BerthTurnaround (time from berthing to completion).For a given facility d and day t, define a congestion index gd,t ∈ [0,1] derived from normalized queue lengths and dwell times, and a throughput measure qd,t (e.g., TEUs handled). At a regional or global level, we can aggregate:
where wd captures facility capacity or strategic weight. Gt then represents a Spaxiom-derived real-time logistics stress index for a trade lane.
We can examine lead–lag relationships between (Gt, Qt) and traditional macro indicators, such as export volumes, shipping line revenues, or freight rate indices. For example:
where ΔRshippingt+ℓ is a sector-level revenue or earnings change at horizon ℓ.
A simplified pattern for gate queues:
from spaxiom.intent import GateQueueMonitor
from spaxiom.logic import on, Condition
from spaxiom.temporal import within
gate = GateQueueMonitor(
entry_sensor=gate_loop_entry,
exit_sensor=gate_loop_exit,
)
# fires once per 15 minutes
tick_15m = within(900, Condition(lambda: True))
@on(tick_15m)
def emit_gate_state():
state = gate.snapshot(window_s=900)
event = {
"type": "GateQueue",
"site_id": "port-alpha",
"timestamp": now_iso(),
"avg_wait_s": state["avg_wait_s"],
"p95_wait_s": state["p95_wait_s"],
"queue_length": state["queue_length"],
"stress_score": state["stress_score"],
}
bus.publish("internal.logistics.events", event)
Figure 8: Hypothetical correlation between a Spaxiom-derived logistics stress index Gt and shipping sector returns or freight rates. Elevated stress often precedes spikes in rates and revenues.
The original "Era of Experience" framing envisions agents that learn predominantly from trajectories of interaction with the world, rather than from static internet corpora. In reinforcement learning notation, an agent's experience can be written as a sequence of tuples
where st is state, at an action, rt a reward, and ot an observation at time t.
Spaxiom extends this classical view by inserting a layer of structured events between raw observations and the agent. At each time t, the underlying sensors produce raw signals xt, which Spaxiom transforms into an event set Et:
where each et(i) is an INTENT-level object such as GaitInstability, CrowdFormation, UnderutilizedExit, or NeedsService.
Consider D deployed Spaxiom sites (hospitals, warehouses, offices, etc.), each generating event streams over time. For site d, let:
denote the set of events emitted over some interval. The global corpus of experience events is then:
Because Spaxiom operates on high-resolution sensor grids (e.g., 4" floor pixels) and other dense modalities, and because it runs continuously at the edge, ℰglobal can approximate a highest-resolution, regularly updated corpus of how people interact with buildings, devices, and robots. Importantly, this corpus is not just unstructured telemetry; it is:
Every event has a type and schema
Zones, coordinates, topologies
Start/end times, durations, windows
Per-event metadata on source, policy, and privacy
Let |ℰglobal(t)| denote the number of events accumulated up to time t. Assuming each site produces events at average rate λd events/second, then:
For many sites and long timescales, this becomes a continuously growing experience fabric whose size and diversity can rival or exceed static web-scale corpora, but now grounded in physical interaction.
This corpus can be used to train a variety of models:
Because events encode semantically meaningful structure, many models can operate in a reduced state space. Suppose a raw sensor state xt lives in ℝn, but the INTENT state zt (e.g., counts, scores, zone-level features) lives in ℝm with m ≪ n. A world model fθ can be trained in the lower-dimensional space:
reducing sample complexity and compute requirements relative to predicting in raw sensor space.
Online, agents can treat ℰglobal as a retrieval-augmented memory:
Formally, let q be a query embedding derived from the current situation, and let {z̃i} be embeddings of past episodes. A retrieval step returns:
and the agent conditions its policy π(a | q, 𝒩(q,K)) on those episodes.
Finally, having a language-level representation of experience simplifies governance:
DoorBlocked, CrowdFormation, and low-exit-usage) and used to filter or reweight training data.In this sense, Spaxiom is not just another middleware layer; it is a deliberately designed substrate for the Era of Experience: a way to convert messy, heterogeneous sensor streams into a structured, governable, and model-ready corpus of physical-world experience.
from spaxiom.intent import ADLTracker
from spaxiom.temporal import within
from spaxiom.logic import on, Condition
adl = ADLTracker(
bed_sensor=bed_mat,
fridge_sensor=fridge_switch,
bath_sensor=bath_humidity,
hall_sensor=hall_floor,
)
# Example: alert if no "walk" events in past 6 hours
no_walk_6h = ~within(
6 * 3600,
Condition(lambda: adl.daily_counts()["walk"] > 0)
)
@on(no_walk_6h)
def check_on_resident():
# Agent or workflow integration here
send_notification("No hallway walk detected in 6h for resident 12B")
from spaxiom.intent import QueueFlow
from spaxiom.temporal import within
from spaxiom.logic import on, Condition
dock_queue = QueueFlow(dock_floor_sensor)
long_queue = within(300, Condition(lambda: dock_queue.length() > 8))
@on(long_queue)
def suggest_extra_worker():
facts = {
"queue_length": dock_queue.length(),
"wait_time": dock_queue.wait_time(),
"arrival_rate": dock_queue.arrival_rate(),
}
# Hand off to LLM to propose options (reroute trucks, open extra lane, etc.)
call_llm_with_queue_facts(facts)
from spaxiom.intent import FmSteward
from spaxiom.logic import on, Condition
fm = FmSteward(
door_counter=restroom_door_counter,
towel_sensor=towel_load_cell,
bin_sensor=bin_ultrasonic,
gas_sensor=nh3_sensor,
floor_sensor=wet_strip,
)
needs_service_cond = Condition(fm.needs_service)
@on(needs_service_cond)
def create_ticket():
payload = fm.snapshot()
cmms.create_work_order(
summary=f"Restroom {payload['entries_approx']} entries; needs service",
metadata=payload,
)
Spaxiom currently relies on hand-designed event schemas. Open questions:
Potential direction: treat Spaxiom events as a learned discrete bottleneck, analogous to VQ-VAE codes but for sensor experiences.
We sketched how safety envelopes could be compiled to automata. Future work:
On the model side:
We expect an ecosystem of Spaxiom-aware agent frameworks, where the DSL is the default way to "speak sensor."
To move beyond thought experiments, we plan to define benchmarks across several dimensions:
Adoption of any new framework depends critically on developer experience (DX). Even with powerful abstractions, developers need practical tools to build, test, debug, and deploy Spaxiom applications efficiently. This section outlines the tooling ecosystem required to make Spaxiom accessible to practitioners.
Modern developers expect first-class IDE support. Spaxiom provides:
# VSCode shows inline type hints and errors
from spaxiom import Sensor, Zone, Condition
from spaxiom.units import meters, celsius
zone = Zone(x=0, y=0, width=10 * meters, height=5 * meters)
temp = Sensor(name="temp", unit=celsius)
# Error: incompatible units (meters vs celsius)
# VSCode underlines in red: "Cannot compare Temperature with Distance"
too_hot = Condition(lambda: temp.read() > 10 * meters) # Type error!
For exploratory development and data science workflows:
# In Jupyter notebook
from spaxiom.viz import plot_zones, plot_event_timeline
# Visualize spatial layout
plot_zones([loading_zone, staging_zone, storage_zone])
# Plot event timeline
events = store.query(since="2025-01-01", limit=1000)
plot_event_timeline(events, group_by="zone")
The spaxiom CLI provides project scaffolding, testing, and deployment:
# Create new Spaxiom project
$ spaxiom init warehouse-monitor
Created warehouse-monitor/
├── spaxiom.yaml # Configuration
├── sensors.py # Sensor definitions
├── patterns.py # INTENT patterns
├── main.py # Runtime entry point
└── tests/ # Unit tests
# Install dependencies
$ cd warehouse-monitor
$ spaxiom install
# Run unit tests (simulated sensors)
$ spaxiom test
Running 12 tests...
✓ test_occupancy_threshold (0.2s)
✓ test_queue_formation (0.5s)
✓ test_overheating_alert (0.1s)
...
12 passed, 0 failed
# Validate configuration
$ spaxiom validate
✓ All sensors defined
✓ All zones have valid coordinates
✓ No circular pattern dependencies
✓ Type checking passed
# Check coverage (which sensors/patterns are tested)
$ spaxiom test --coverage
Sensor coverage: 18/20 (90%)
Pattern coverage: 8/10 (80%)
Condition coverage: 15/20 (75%)
# Deploy to edge device
$ spaxiom deploy --target pi@192.168.1.100
Uploading code... ✓
Installing dependencies... ✓
Starting runtime... ✓
Runtime listening on http://192.168.1.100:8080
# View logs
$ spaxiom logs --follow
[2025-01-06 10:23:45] INFO: Runtime started (tick_rate=10Hz)
[2025-01-06 10:23:46] INFO: Sensors online: 20/20
[2025-01-06 10:23:47] EVENT: OccupancyChanged(zone=loading, count=5)
[2025-01-06 10:23:50] ALERT: QueueFormed(zone=loading, length=8)
# Check runtime health
$ spaxiom status
Runtime: HEALTHY
Uptime: 3d 14h 22m
Sensors: 20/20 online
Events/sec: 12.3
Latency p99: 8.2ms
Spaxiom includes a pytest-based testing framework with specialized fixtures:
# tests/test_queue_pattern.py
from spaxiom.testing import MockRuntime, MockSensor, advance_time
def test_queue_formation():
# Create mock runtime with simulated sensors
runtime = MockRuntime()
camera = MockSensor(name="camera", initial_value=0)
runtime.add_sensor(camera)
# Create pattern
from patterns import QueuePattern
queue = QueuePattern(camera=camera, threshold=5)
runtime.add_pattern(queue)
# Simulate sensor readings
camera.set_value(3) # Below threshold
runtime.tick()
assert not queue.is_active()
camera.set_value(7) # Above threshold
runtime.tick()
assert queue.is_active()
# Simulate time passage
advance_time(runtime, seconds=10)
assert queue.duration() == 10.0
# tests/test_warehouse_scenario.py
from spaxiom.testing import SimulatedEnvironment
def test_warehouse_workflow():
# Create simulated warehouse with 4 zones
env = SimulatedEnvironment.from_config("warehouse.yaml")
# Simulate 1 hour of activity
events = []
for t in range(3600):
env.step() # Advance 1 second
events.extend(env.get_events())
# Assert expected event sequence
assert any(e["type"] == "TruckArrived" for e in events)
assert any(e["type"] == "LoadingStarted" for e in events)
assert any(e["type"] == "LoadingCompleted" for e in events)
# Check performance metrics
avg_loading_time = env.get_metric("avg_loading_time")
assert avg_loading_time < 600.0 # Under 10 minutes
Web-based UI for stepping through event sequences:
# Start debugger with recorded events
$ spaxiom debug /var/lib/spaxiom/events.db
Opening debugger at http://localhost:3000
# Debugger UI shows:
# - Timeline slider (scrub through time)
# - Event list (filter by type, zone, priority)
# - Sensor values at each timestamp
# - Pattern state visualization
# - Condition evaluation traces
3D visualization of zones, entities, and sensor coverage:
from spaxiom.viz import SpatialVisualizer
viz = SpatialVisualizer(runtime)
# Render 3D scene (Three.js / Unity)
viz.show_zones(wireframe=True)
viz.show_entities(trail_length=30) # Show movement trails
viz.show_sensor_coverage(camera_sensors, fov=90)
# Animate event timeline
viz.playback(events, speed=10.0) # 10x real-time
Set breakpoints triggered by specific events:
from spaxiom.debug import breakpoint_on_event
# Pause execution when queue forms
@breakpoint_on_event("QueueFormed")
def inspect_queue(event):
print(f"Queue in {event['zone']}: length={event['length']}")
# Drop into interactive debugger
import pdb; pdb.set_trace()
runtime.run()
from spaxiom.profiling import RuntimeProfiler
profiler = RuntimeProfiler(runtime)
profiler.start()
runtime.run(duration=300) # Profile 5 minutes
report = profiler.report()
print(report.summary())
# Output:
# ========== Spaxiom Runtime Profile ==========
# Total ticks: 3000 (10.0 Hz)
#
# Top 5 patterns by latency:
# 1. QueueFlow: 4.2ms avg, 8.9ms p99
# 2. OccupancyField: 2.1ms avg, 5.3ms p99
# 3. ADLTracker: 1.8ms avg, 3.2ms p99
#
# Top 3 callbacks by duration:
# 1. on_queue_formed: 12.3ms avg
# 2. on_overheating: 0.8ms avg
#
# Sensor read latency:
# camera_loading: 3.2ms avg
# modbus_temp_1: 0.5ms avg
$ spaxiom profile memory --duration 3600
Monitoring memory usage for 1 hour...
Memory growth detected:
Pattern: OccupancyField (zone=loading)
Growth rate: +2.3 MB/hour
Likely cause: Unbounded history buffer
Recommendation: Add max_history_size parameter
$ spaxiom profile flamegraph --output profile.svg
Generated flamegraph: profile.svg
Open in browser to see call stack visualization
from spaxiom.sim import SyntheticSensor, GaussianNoise, PeriodicSignal
# Temperature with diurnal cycle + noise
temp = SyntheticSensor(
name="temp",
base_signal=PeriodicSignal(
amplitude=5.0, # ±5°C
period=86400.0, # 24 hours
offset=20.0 # 20°C baseline
),
noise=GaussianNoise(std=0.5)
)
# Occupancy with Poisson arrivals
occupancy = SyntheticSensor(
name="occupancy",
generator=lambda t: np.random.poisson(lam=10.0) # 10 people avg
)
runtime.add_sensor(temp)
runtime.add_sensor(occupancy)
from spaxiom.sim import ScenarioRecorder, ScenarioPlayer
# Record a scenario from production
recorder = ScenarioRecorder(runtime)
runtime.run(duration=3600) # Record 1 hour
recorder.save("scenarios/normal_ops.spx")
# Later: replay for regression testing
player = ScenarioPlayer.load("scenarios/normal_ops.spx")
test_runtime = SpaxiomRuntime()
player.attach(test_runtime)
# Run at 100x speed
test_runtime.run(speed=100.0)
# Assert no regressions
assert test_runtime.event_count("SafetyViolation") == 0
from spaxiom.monitoring import PrometheusExporter
exporter = PrometheusExporter(runtime, port=9090)
exporter.start()
# Metrics exposed:
# spaxiom_events_total{type="DoorOpened",zone="loading"}
# spaxiom_pattern_latency_seconds{pattern="QueueFlow"}
# spaxiom_sensor_read_errors_total{sensor="camera_1"}
# spaxiom_runtime_tick_rate_hz
Pre-built Grafana dashboards for common metrics:
# Import dashboard templates
$ spaxiom grafana import --dashboard runtime-overview
$ spaxiom grafana import --dashboard sensor-health
$ spaxiom grafana import --dashboard event-timeline
# Dashboards show:
# - Event rate over time (by type, zone, priority)
# - Pattern latency heatmaps
# - Sensor health status
# - Callback execution times
# - Memory and CPU usage
from spaxiom.tracing import OpenTelemetryTracer
tracer = OpenTelemetryTracer(
endpoint="http://jaeger:4318",
service_name="warehouse-runtime"
)
runtime.set_tracer(tracer)
# Traces show:
# - Sensor read → Pattern update → Condition eval → Callback dispatch
# - Cross-site event propagation (edge → cloud)
# - ML model inference triggered by events
# Generate docs from code
$ spaxiom docs generate --output docs/
Generating documentation...
✓ API reference (200 endpoints)
✓ Pattern library (25 patterns)
✓ Type definitions (50 types)
✓ Examples (30 snippets)
Docs available at docs/index.html
Generate Python pattern classes from declarative YAML schemas:
# patterns/custom.yaml
- name: CustomOccupancy
sensors:
- camera: Camera
- floor: PressureMat
parameters:
- threshold: int
events:
- name: ZoneOccupied
fields:
- count: int
- timestamp: datetime
$ spaxiom generate patterns patterns/custom.yaml --output patterns/generated.py
Generated patterns/generated.py with 1 pattern class
# Now import and use
from patterns.generated import CustomOccupancy
$ spaxiom api spec --output openapi.yaml
Generated OpenAPI 3.0 spec with:
- GET /events
- GET /sensors/{id}
- POST /patterns
- GET /zones
# Use with code generators
$ openapi-generator-cli generate -i openapi.yaml -g python -o clients/python
# Install from PyPI
$ pip install spaxiom
# With optional dependencies
$ pip install spaxiom[vision] # Camera/video support
$ pip install spaxiom[industrial] # Modbus, OPC UA, BACnet
$ pip install spaxiom[cloud] # AWS, Azure, GCP connectors
$ pip install spaxiom[ml] # ML integration (Feast, Tecton)
$ pip install spaxiom[all] # Everything
# Official images on Docker Hub
$ docker pull spaxiom/runtime:latest
$ docker pull spaxiom/runtime:edge # Minimal for Raspberry Pi
$ docker pull spaxiom/runtime:cloud # With cloud connectors
# Run containerized runtime
$ docker run -p 8080:8080 \
-v /var/lib/spaxiom:/data \
spaxiom/runtime:latest
# Add Helm repo
$ helm repo add spaxiom https://charts.spaxiom.io
# Install runtime on Kubernetes
$ helm install warehouse-runtime spaxiom/runtime \
--set config.tickRate=10 \
--set config.sensors[0].name=camera_1 \
--set config.sensors[0].type=rtsp
# Supports horizontal scaling
$ kubectl scale deployment warehouse-runtime --replicas=10
Browser-based editor with live preview (like Rust Playground):
# Visit https://playground.spaxiom.io
# Features:
# - Code editor with syntax highlighting
# - Simulated sensors (drag-and-drop spatial layout)
# - Real-time event visualization
# - Share links to examples
# - Fork and modify templates
$ spaxiom examples list
Available examples:
1. hello-world Simple occupancy detection
2. warehouse-queue Queue flow monitoring
3. hvac-optimization Energy-aware HVAC control
4. robot-safety Collision avoidance with safety zones
5. retail-analytics Customer journey tracking
...
$ spaxiom examples run hello-world
Running example: hello-world
Press Ctrl+C to stop
[Output shows simulated events in real-time]
Community repository of reusable patterns:
# Search for patterns
$ spaxiom marketplace search "queue"
Found 5 patterns:
- advanced-queue-flow (★ 245)
- multi-stage-queue (★ 89)
- priority-queue-manager (★ 67)
# Install pattern
$ spaxiom marketplace install advanced-queue-flow
Installed advanced-queue-flow v2.1.0
# Now use in code
from spaxiom.marketplace import AdvancedQueueFlow
# Create custom plugin
from spaxiom.plugins import Plugin
class CustomVisualizerPlugin(Plugin):
def on_event(self, event):
# Custom visualization logic
self.render(event)
# Register plugin
runtime.register_plugin(CustomVisualizerPlugin())
# Discover community plugins
$ spaxiom plugins search "visualization"
$ spaxiom plugins install spaxiom-3d-viz
Spaxiom's developer experience is designed to support the full development lifecycle:
By investing in world-class tooling, Spaxiom aims to reduce the barrier to entry for developers building spatiotemporal applications, accelerating adoption across industries from manufacturing to healthcare to smart cities.
As sensors proliferate and AI shifts into the Era of Experience, we need programming tools that treat space, time, and interaction as first-class design constraints. Spaxiom proposes:
In the same way SQL became the lingua franca for structured business data, and modern deep learning frameworks became a lingua franca for neural computation, Spaxiom aims to become a lingua franca for sensor experience: a bridge between the physical world and the agents that will increasingly inhabit it.
If we succeed, future frontier models will not just read the internet; they will read the world through a concise, structured, and safe language that lets them understand, remember, and act on the experiences of billions of devices and the humans they serve.
This appendix sketches a set of concrete domains where Spaxiom's spatial-temporal abstractions and INTENT layer provide a natural "sensor cortex" and experience substrate. Each use case is deliberately multi-sensor and goes beyond floor-only deployments, illustrating that the language is designed for general spatial sensing rather than a single modality.
For each use case we:
High-grade semiconductor fabs, biopharma facilities, and advanced manufacturing lines rely on strict control of particulate contamination, pressure cascades, and controlled access. Today, contamination control is typically enforced by:
What is often missing is a spatially and temporally coherent language for:
Spaxiom can act as a cleanroom contamination cortex: fusing particle counts, pressure differentials, door states, and occupancy into structured INTENT events that agents and engineers can reason about.
Typical cleanroom-relevant sensors include:
Spaxiom represents each zone as a Zone with associated sensor objects and topological relationships (e.g., which rooms feed which in the pressure cascade).
From raw signals, we define higher-level INTENT events, such as:
PressureBreach: pressure differential below spec across a boundary;ParticleExcursion: temporary spike of particulate counts above class limit;AirlockViolation: door sequencing broken (both doors open, or bypass of required dwell);HighRiskMovement: occupancy trajectory that crosses from dirtier to cleaner areas under non-compliant conditions.Mathematically, for a given zone z and monitoring window [t0, t1] we can define:
Pressure breach indicator. Let ΔPz,u(t) be the measured pressure differential between z and upstream zone u (e.g., corridor or anteroom). For a minimum acceptable differential ΔPmin > 0, define
The total number of breach-seconds in the window is:
Particle excursion integral. Let pz(t) be the particle count for zone z, with a threshold pmaxz that defines acceptable class performance. Define the (non-negative) excursion:
Airlock violation count. Let Va be the count of airlock a violations in the window, derived from door state sequences and configured policies (e.g., both doors open simultaneously, insufficient purge dwell).
Composite contamination risk index. We can define a simple contamination risk index (CRI) for zone z over [t0, t1] as:
where α, β, γ are tunable weights reflecting domain expertise, and 𝒜(z) is the set of airlocks affecting z. This can be normalized by window length or baseline values to obtain a dimensionless score in [0, 1].
A sketch of how a Spaxiom-based contamination monitor might look in Python-embedded DSL form:
from spaxiom import Zone, Condition, Quantity
from spaxiom.temporal import within
from spaxiom.logic import on
class CleanroomZone:
def __init__(self, name, particle_sensor, dp_sensors, airlocks):
self.zone = Zone.named(name)
self.particle_sensor = particle_sensor # e.g. counts / m^3
self.dp_sensors = dp_sensors # dict: upstream_zone -> sensor
self.airlocks = airlocks # list of Airlock objects
# Configurable thresholds
self.max_particles = 3500 # domain-specific
self.min_dp = Quantity(5.0, "Pa")
def pressure_breach_seconds(self, window_s: float) -> float:
total = 0.0
for upstream, sensor in self.dp_sensors.items():
# integrate indicator over the window
series = sensor.history(window_s=window_s)
for dt, value in series: # dt in seconds, value in Pa
if value < self.min_dp:
total += dt
return total
def particle_excursion(self, window_s: float) -> float:
# approximate integral of (p - p_max)+ over the window
total = 0.0
series = self.particle_sensor.history(window_s=window_s)
for dt, value in series:
excess = max(0.0, value - self.max_particles)
total += excess * dt
return total
def airlock_violations(self, window_s: float) -> int:
return sum(a.violation_count(window_s=window_s)
for a in self.airlocks)
def contamination_risk_index(self, window_s: float) -> float:
B = self.pressure_breach_seconds(window_s)
E = self.particle_excursion(window_s)
V = self.airlock_violations(window_s)
alpha, beta, gamma = 1e-3, 1e-6, 1.0 # example scaling
score = alpha * B + beta * E + gamma * V
# Optionally squash to [0,1] via logistic
return 1.0 - (1.0 / (1.0 + score))
# Example zone wiring
zone_a = CleanroomZone(
name="ISO7_bio_room_3",
particle_sensor=particles_z3,
dp_sensors={"antechamber": dp_z3_ante, "corridor": dp_z3_corr},
airlocks=[airlock_3A, airlock_3B],
)
# Condition that fires when CRI exceeds a threshold in the last hour
high_risk = Condition(
lambda: zone_a.contamination_risk_index(window_s=3600) > 0.7
)
@on(within(3600, high_risk))
def contamination_agent():
snapshot = {
"zone": zone_a.zone.name,
"CRI": zone_a.contamination_risk_index(window_s=3600),
"breach_seconds": zone_a.pressure_breach_seconds(3600),
"particle_excursion": zone_a.particle_excursion(3600),
"airlock_violations": zone_a.airlock_violations(3600),
}
# Hand off to LLM agent or workflow system:
# e.g., propose root-cause checks, quarantine, or extra cleaning.
recommend_actions(snapshot)
Here the Spaxiom layer:
history(window_s=...)) for time integration;contamination_risk_index for agents, instead of raw time series.Figure A.1: Cleanroom contamination risk timeline over one shift. Three normalized metrics are overlaid: particle counts (red) vs threshold, pressure differential (blue) vs minimum spec, and computed contamination risk index CRIz (purple). The shaded region indicates when CRI exceeds the alert threshold. An airlock violation event at ~2h coincides with particle excursion and elevated risk.
Critical rotating equipment (pumps, fans, compressors, gearboxes, turbines) is instrumented heavily in modern industrial facilities. Traditional "predictive maintenance" stacks often revolve around vendor-specific vibration analyzers, ad-hoc thresholds, and periodic offline analysis. However, the semantics of what operators actually care about are higher-level:
Spaxiom can serve as a machinery health cortex, fusing vibration, acoustic, electrical, and thermal measurements into structured INTENT events that agents and engineers can reason over consistently across vendors and plants.
For a rotating asset m (e.g., pump P-101), typical sensors include:
Spaxiom does not need to operate at raw waveform level; instead, it consumes features precomputed either at the edge or inside a preprocessor:
From these features, Spaxiom defines higher-level INTENT events such as:
BearingAnomaly (spectral patterns consistent with bearing defect frequencies),ImbalanceDetected (elevated 1X vibration at speed-correlated frequency),CavitationRisk (acoustic + vibration signature in pumps),LoadSpike (current and torque exceeding normal envelopes),ThermalRunawayRisk (sustained temperature rise faster than expected under load).For a machine m and window [t0, t1], suppose we have:
Normalized feature integrals. Define normalized, windowed feature integrals:
where Δt = t1 - t0 and σ·m,base are baseline (healthy) standard deviations.
Similarly define:
where Tbasem(L) is the expected temperature as a function of load (learned from historical data).
Composite health score. We can define a simple machine health "anomaly score":
with weights wi ≥ 0 tuned per machine or class. High Hm indicates abnormal behavior over the window.
Discrete INTENT events. We can then create discrete INTENT events when:
yielding MachineHealthWarning or MachineHealthAlarm for machine m.
In the DSL, a rotating machine entity might encapsulate feature histories and health computation:
from spaxiom import Condition
from spaxiom.temporal import within
from spaxiom.logic import on
class RotatingMachine:
def __init__(self, name, feat_source):
self.name = name
self.feat_source = feat_source # e.g., streaming feature vectors
# Baseline stats & thresholds (could be learned)
self.sigma_1x = 0.12
self.sigma_bd = 0.08
self.warn_threshold = 2.0
self.alarm_threshold = 4.0
def feature_history(self, window_s: float):
"""
Returns a list of (dt, features) where features is a dict like:
{
"v_1x": ...,
"v_bd": ...,
"kurtosis": ...,
"temp": ...,
"load": ...,
}
"""
return self.feat_source.history(window_s=window_s)
def expected_temp(self, load: float) -> float:
# Simple linear model as placeholder; could be learned per machine
return 40.0 + 25.0 * load
def health_score(self, window_s: float) -> float:
series = self.feature_history(window_s)
if not series:
return 0.0
sum_v1x = sum_vbd = sum_k = sum_texcess = 0.0
total_dt = 0.0
for dt, f in series:
v1x = f["v_1x"] / max(self.sigma_1x, 1e-6)
vbd = f["v_bd"] / max(self.sigma_bd, 1e-6)
k = f["kurtosis"]
temp = f["temp"]
load = f["load"]
texcess = max(0.0, temp - self.expected_temp(load))
sum_v1x += v1x * dt
sum_vbd += vbd * dt
sum_k += k * dt
sum_texcess += texcess * dt
total_dt += dt
if total_dt <= 0:
return 0.0
V1x = sum_v1x / total_dt
Vbd = sum_vbd / total_dt
K = sum_k / total_dt
Tex = sum_texcess / total_dt
w1, w2, w3, w4 = 0.4, 0.3, 0.2, 0.1
return w1 * V1x + w2 * Vbd + w3 * K + w4 * Tex
# Example machine instance
pump_101 = RotatingMachine("P-101", feat_source=pump_101_feat_stream)
# Conditions that trigger warnings/alarms over the last 24 hours
warn_condition = Condition(
lambda: pump_101.health_score(window_s=24 * 3600) > pump_101.warn_threshold
)
alarm_condition = Condition(
lambda: pump_101.health_score(window_s=24 * 3600) > pump_101.alarm_threshold
)
@on(within(3600, warn_condition)) # evaluate every hour
def pump_warning_agent():
score = pump_101.health_score(window_s=24 * 3600)
emit_intent_event({
"type": "MachineHealthWarning",
"machine": pump_101.name,
"score": score,
"window_h": 24,
})
# Optional: hand off to LLM for recommended maintenance actions.
@on(within(600, alarm_condition)) # evaluate every 10 minutes
def pump_alarm_agent():
score = pump_101.health_score(window_s=24 * 3600)
emit_intent_event({
"type": "MachineHealthAlarm",
"machine": pump_101.name,
"score": score,
"window_h": 24,
})
# Optionally, trigger interlocks, derates, or automated shutdown logic.
This sketch illustrates how Spaxiom:
feat_source.history),health_score to be consumed by agents and workflows.Figure A.2: Rotating machinery health score and components for Pump P-101 over 15 days. Panel 1 shows normalized 1X vibration (orange) and bearing-band vibration (red), with bearing defects gradually increasing. Panel 2 shows spectral kurtosis (blue) and temperature excess (green), both rising as bearing condition deteriorates. Panel 3 shows the composite health score Hm (purple) crossing the warning threshold at day 9 and alarm threshold at day 11, triggering corresponding INTENT events.
Indoor air quality (IAQ) and ventilation are increasingly recognized as critical to health, cognitive performance, and resilience to airborne disease. Most commercial buildings already have a wealth of signals (CO2 sensors, temperature, humidity, HVAC control points), but they are typically used only for crude comfort bands. There is rarely a unified, spatially aware representation of:
Spaxiom can act as a ventilation and health cortex for buildings, fusing IAQ sensors, occupancy estimates, and HVAC state into intelligible INTENT events like StaleAirEpisode, VentilationDebt, and HighRiskGathering.
For a zone z (e.g., conference room, open office bay, classroom), typical signals include:
We assume each Zone in Spaxiom is wired to one or more of these signals and that zones can be grouped into higher-level areas for policy and analytics.
We define several IAQ- and health-related INTENT events:
StaleAirEpisode: CO2 sustained above a threshold (e.g., 1000 ppm) for more than a minimum duration;VentilationDebt: integrated shortfall of outdoor air per person relative to recommended rates;HighRiskGathering: high occupancy combined with poor ventilation and adverse IAQ (CO2, PM, humidity).Let us formalize some of these.
Ventilation per person. For zone z, at time t, the outdoor airflow per person can be approximated as:
where FOAz(t) is the fraction of supply air that is outdoor air, and V̇supz(t) is the total supply airflow.
Given a recommended per-person outdoor airflow qrec (e.g., from a standard), we define a ventilation deficit rate:
Over a monitoring window [t0, t1] of length Δt, the ventilation debt is:
CO2 excursion and stale air. Let Cstale be a CO2 threshold (e.g., 1000 ppm). Define an indicator:
and the duration of stale air in the window:
Composite health risk score. A simple IAQ/health risk score for zone z over [t0, t1] can be defined as:
where:
Normalization (e.g., dividing by window length or baseline values) can map RIAQz into a dimensionless score in [0,1].
In the Spaxiom DSL, a zone-level IAQ tracker can encapsulate sensor histories and risk computation:
from spaxiom import Zone, Condition
from spaxiom.temporal import within
from spaxiom.logic import on
class IaqZone:
def __init__(self, name, co2, temp, rh, oa_frac, sup_flow, occupancy):
self.zone = Zone.named(name)
self.co2 = co2 # ppm sensor
self.temp = temp # degC
self.rh = rh # %RH
self.oa_frac = oa_frac # 0..1
self.sup_flow = sup_flow # m^3/s
self.occupancy = occupancy # persons
# Configurable thresholds / recommendations
self.q_rec = 10.0 / 3600.0 # 10 m^3/h/person -> m^3/s/person
self.co2_stale = 1000.0 # ppm
self.pm_threshold = 25.0 # ug/m^3 (example)
self.rh_low = 30.0 # %
self.rh_high = 60.0 # %
def history(self, sensor, window_s: float):
return sensor.history(window_s=window_s) # [(dt, value), ...]
def ventilation_debt(self, window_s: float) -> float:
series_oa = self.history(self.oa_frac, window_s)
series_flow = self.history(self.sup_flow, window_s)
series_occ = self.history(self.occupancy, window_s)
# Assume aligned histories or interpolate in real implementation
total_debt = 0.0
for ((dt, oa), (_, flow), (_, occ)) in zip(series_oa, series_flow, series_occ):
q = oa * flow / max(occ, 1.0)
d = max(0.0, self.q_rec - q)
total_debt += d * dt
return total_debt
def stale_air_duration(self, window_s: float) -> float:
series = self.history(self.co2, window_s)
total = 0.0
for dt, c in series:
if c > self.co2_stale:
total += dt
return total
def rh_excursion(self, window_s: float) -> float:
series = self.history(self.rh, window_s)
total = 0.0
for dt, r in series:
if r < self.rh_low or r > self.rh_high:
total += dt
return total
def risk_score(self, window_s: float) -> float:
D = self.ventilation_debt(window_s)
S = self.stale_air_duration(window_s)
RH = self.rh_excursion(window_s)
alpha, beta, gamma, delta = 1.0, 0.5, 0.0, 0.2 # PM omitted here
score = alpha * D + beta * S + delta * RH
# Example squashing to [0,1]
return 1.0 - (1.0 / (1.0 + score * 1e-4))
# Wire a specific conference room
conf_A = IaqZone(
name="Conf_Room_A",
co2=co2_conf_A,
temp=temp_conf_A,
rh=rh_conf_A,
oa_frac=oa_conf_A,
sup_flow=flow_conf_A,
occupancy=occ_conf_A,
)
# Condition: high risk over the last 2 hours
high_risk_iaq = Condition(lambda: conf_A.risk_score(window_s=2 * 3600) > 0.7)
@on(within(300, high_risk_iaq)) # check every 5 minutes
def iaq_agent():
snapshot = {
"zone": conf_A.zone.name,
"risk": conf_A.risk_score(window_s=2 * 3600),
"vent_debt": conf_A.ventilation_debt(2 * 3600),
"stale_air_s": conf_A.stale_air_duration(2 * 3600),
"rh_excursion_s": conf_A.rh_excursion(2 * 3600),
}
# An LLM or rules engine can:
# - suggest schedule changes,
# - recommend window opening where applicable,
# - adjust ventilation setpoints if allowed.
propose_iaq_actions(snapshot)
This example shows how Spaxiom:
risk_score;high_risk_iaq) for agents to react on.Figure A.3: IAQ and ventilation risk for Conference Room A over a workday. Panel 1 shows CO2 concentration (red) exceeding the stale threshold (1000 ppm) during two meetings. Panel 2 shows outdoor air per person (blue) dropping below recommended levels during those same periods. Panel 3 shows the composite IAQ risk score RIAQz (purple) crossing into high-risk territory during meetings with poor ventilation. Shaded regions indicate StaleAirEpisode and HighRiskGathering INTENT events.
As cities densify and micro-mobility (bikes, e-scooters, small EVs) proliferates, safety and flow management become critical. Today, safety analysis often relies on:
However, the most informative signals are often the near misses and repeated risky patterns: sudden braking, evasive maneuvers, conflicts between modes, and unsafe speeds at known bottlenecks. These rarely make it into structured datasets.
Spaxiom can act as a street-level safety cortex, fusing radar, acoustic, and IMU signals from vehicles and infrastructure into INTENT events like NearMissCluster, SpeedingCorridor, and UnsafeCrossing.
At the city scale, relevant sensors include:
Spaxiom ingests processed features from these sources (not raw full-resolution waveforms or video) and maps them into a common spatial model of the street network (segments, intersections, crosswalks, lanes).
We define several safety-relevant INTENT events:
NearMiss: a spatiotemporal configuration where two or more agents come within a critical distance at relative speed above a threshold without collision;HarshEvent: harsh braking, swerving, or strong lateral acceleration from IMU traces;SpeedingCorridor: persistent high share of vehicles exceeding speed limits on a segment;UnsafeCrossing: repeated near misses or HarshEvents at or near a crosswalk or intersection.Near-miss detection. Consider two agents a and b (e.g., a car and a cyclist) with positions xa(t), xb(t) and velocities va(t), vb(t). For a short horizon τ ∈ [0, τmax], approximate future positions with constant velocity:
Define the predicted minimum separation distance over that horizon:
Let Δv(t) = va(t) - vb(t), and define relative speed vrel(t) = ‖Δv(t)‖.
A near-miss candidate is a triple (a,b,t) such that:
where dnmthresh (e.g., 1–2 m) and vnmthresh (e.g., 5 m/s) are thresholds.
Each such event can be encapsulated as a NearMiss INTENT event with attributes:
Segment-level risk score. For a road segment ℓ over a period [t0, t1], define:
A simple risk index for segment ℓ is:
with weights α, β > 0. Segments with high Rℓ are candidates for SpeedingCorridor or UnsafeCrossing labels, depending on their geometry.
We can represent a segment or intersection in the DSL as an object that aggregates INTENT events and IMU-derived HarshEvents.
from spaxiom import Condition
from spaxiom.temporal import within
from spaxiom.logic import on
class RoadSegment:
def __init__(self, seg_id, near_miss_stream, harsh_stream, volume_stream):
self.seg_id = seg_id
self.near_miss_stream = near_miss_stream # yields NearMiss events
self.harsh_stream = harsh_stream # yields HarshEvent events
self.volume_stream = volume_stream # yields vehicle passage counts
def counts(self, window_s: float):
nm_events = self.near_miss_stream.history(window_s=window_s)
harsh_events = self.harsh_stream.history(window_s=window_s)
volume_events = self.volume_stream.history(window_s=window_s)
N_nm = len(nm_events)
N_harsh = len(harsh_events)
N_veh = sum(e["count"] for _, e in volume_events) or 1
return N_nm, N_harsh, N_veh
def risk_index(self, window_s: float) -> float:
N_nm, N_harsh, N_veh = self.counts(window_s)
alpha, beta = 1.0, 0.5
return alpha * (N_nm / N_veh) + beta * (N_harsh / N_veh)
# Example network segments
seg_main_1 = RoadSegment(
seg_id="main_st_block_1",
near_miss_stream=nm_main_1,
harsh_stream=harsh_main_1,
volume_stream=volume_main_1,
)
seg_main_2 = RoadSegment(
seg_id="main_st_block_2",
near_miss_stream=nm_main_2,
harsh_stream=harsh_main_2,
volume_stream=volume_main_2,
)
segments = [seg_main_1, seg_main_2]
# Condition to periodically assess risk
tick_15m = within(900, Condition(lambda: True))
@on(tick_15m)
def micromobility_safety_agent():
window_s = 7 * 24 * 3600 # last 7 days
for seg in segments:
R = seg.risk_index(window_s)
if R > 0.01: # example threshold
emit_intent_event({
"type": "UnsafeCrossing" if is_crossing(seg.seg_id) else "SpeedingCorridor",
"segment": seg.seg_id,
"risk_index": R,
"window_s": window_s,
})
# The resulting stream of UnsafeCrossing / SpeedingCorridor events can drive:
# - infrastructure recommendations,
# - signal timing changes,
# - or targeted enforcement and education campaigns.
In a fuller implementation, separate Spaxiom patterns would compute NearMiss and HarshEvent INTENT events directly from radar and IMU streams, providing reusable building blocks across cities.
Figure A.4: Street network safety heatmap for a downtown district over a 7-day period. Each road segment is colored by its risk index Rℓ (green = low risk, yellow/orange = medium, red = high). The intersection at Main St & 5th St shows high near-miss frequency (marked with ⚠️ icon) and is annotated with example near-miss trajectories between vehicles and vulnerable road users. The inset bar chart shows the distribution of Rℓ across all 24 segments, with the worst decile (2 critical segments including Main & 5th) highlighted in red with dashed border.
Temperature-sensitive pharmaceuticals (vaccines, biologics, insulin, blood products) require strict environmental control during transport and storage. The global cold chain market exceeds $250 billion annually, with pharmaceutical spoilage costing an estimated $35 billion per year. Regulatory frameworks (FDA 21 CFR Part 11, EU GDP, WHO PQS) mandate continuous temperature monitoring and traceability.
A cold chain shipment integrates multiple sensor modalities:
Legacy systems record data to internal memory for post-delivery audit, lacking real-time intervention capability. Spaxiom enables predictive cold chain management by fusing these streams to detect imminent excursions, route deviations, and handling violations while corrective action is still possible.
The cold chain domain defines semantic events that abstract raw sensor readings into regulatory-relevant incidents:
TempExcursion: Fired when temperature deviates from specification. Severity classified as {WARNING, MINOR, MAJOR, CRITICAL} based on Mean Kinetic Temperature (MKT) and duration of excursion.DoorBreach: Container opened outside authorized facility geofence, indicating potential tampering or unauthorized access with timestamp and ambient temperature at breach.RouteDeviation: GPS track diverges from planned route by >5 km or >30 min delay, signaling logistical issues or potential theft.ShockEvent: Physical impact exceeds threshold (e.g., >3g sustained for >100ms), capturing drops, collisions, or mishandling during transport.HumidityExcursion: Relative humidity outside specification (e.g., >60% for lyophilized products), threatening product stability.CoolingFailure: Active cooling system malfunction detected via thermal gradient analysis, indicating compressor failure or refrigerant leak before product is compromised.These events enable automated compliance reporting, real-time alerts to logistics coordinators, and integration with pharmaceutical traceability systems (e.g., EPCIS, GS1).
Raw temperature readings are insufficient for decision-making due to transient fluctuations and spatial variation across the payload. We compute a Shipment Integrity Index (SII) that integrates thermal history, environmental stress, and handling quality:
where each quality component is normalized to [0,1] with 1 = perfect integrity:
Thermal Quality: Based on Mean Kinetic Temperature (MKT), which accounts for cumulative thermal stress:
where ΔHa is the activation energy (~83.14 kJ/mol for biologics), R is the gas constant, Ti are temperatures in Kelvin, and n is the number of samples. Then:
Environmental Quality: Penalizes humidity excursions and light exposure:
where fRH is the fraction of time outside humidity spec and flight is cumulative lux-hours beyond threshold.
Handling Quality: Based on shock events and door breaches:
where Nshock is count of impacts >3g, Nbreach is unauthorized door openings, tdeviation is cumulative route delay in hours, and α, β, γ are penalty coefficients tuned per product category.
An SII below 0.8 triggers IntegrityWarning, below 0.6 triggers IntegrityCritical with recommended quarantine for inspection.
The ColdChainShipment class demonstrates real-time integrity monitoring with automated alert escalation:
from spaxiom import Sensor, Intent, Fusion, Metric
import math
class ColdChainShipment:
def __init__(self, shipment_id, product_spec):
self.shipment_id = shipment_id
self.spec = product_spec # T_min, T_max, RH_max, shock_threshold
# Sensor streams
self.temp_sensors = [Sensor(f"temp_probe_{i}") for i in range(4)]
self.humidity = Sensor("humidity")
self.gps = Sensor("gps_tracker")
self.door = Sensor("door_sensor")
self.accel = Sensor("accelerometer", axes=["x", "y", "z"])
self.light = Sensor("light_sensor")
# INTENT events
self.temp_excursion = Intent("TempExcursion")
self.door_breach = Intent("DoorBreach")
self.route_deviation = Intent("RouteDeviation")
self.shock_event = Intent("ShockEvent")
# Fusion metrics
self.sii = Metric("shipment_integrity_index", range=(0, 1))
self.mkt = Metric("mean_kinetic_temp", unit="°C")
# Historical data for MKT calculation
self.temp_history = []
self.shock_count = 0
self.breach_count = 0
self.route_delay_hours = 0.0
self.light_exposure_lux_hours = 0.0
self.rh_excursion_fraction = 0.0
@Fusion.rule
def compute_mkt(self):
"""Calculate Mean Kinetic Temperature from probe history"""
if len(self.temp_history) < 2:
return self.spec["T_target"]
# Constants for biologics
delta_H = 83144 # J/mol (activation energy)
R = 8.314 # J/(mol·K)
# Convert to Kelvin and compute exponential sum
temps_kelvin = [t + 273.15 for t in self.temp_history]
exp_sum = sum(math.exp(-delta_H / (R * T)) for T in temps_kelvin)
mkt_kelvin = delta_H / (R * math.log(exp_sum / len(temps_kelvin)))
mkt_celsius = mkt_kelvin - 273.15
self.mkt.update(mkt_celsius)
return mkt_celsius
@Fusion.rule
def calculate_sii(self):
"""Compute Shipment Integrity Index from all quality components"""
# Thermal quality
mkt = self.compute_mkt()
T_target = self.spec["T_target"]
T_tolerance = self.spec["T_tolerance"]
Q_thermal = max(0, 1 - abs(mkt - T_target) / T_tolerance)
# Environmental quality
Q_environmental = (1 - self.rh_excursion_fraction) * \
(1 - min(1.0, self.light_exposure_lux_hours / 100))
# Handling quality
alpha, beta, gamma = 0.1, 0.2, 0.05
Q_handling = math.exp(-alpha * self.shock_count -
beta * self.breach_count -
gamma * self.route_delay_hours)
# Weighted combination (thermal is most critical)
w_T, w_E, w_H = 0.6, 0.2, 0.2
sii = w_T * Q_thermal + w_E * Q_environmental + w_H * Q_handling
self.sii.update(sii)
# Emit alerts based on thresholds
if sii < 0.6:
Intent.emit("IntegrityCritical", shipment_id=self.shipment_id,
sii=sii, action="QUARANTINE_FOR_INSPECTION")
elif sii < 0.8:
Intent.emit("IntegrityWarning", shipment_id=self.shipment_id,
sii=sii, action="NOTIFY_QA_TEAM")
return sii
@Sensor.on_data("temp_probe_*")
def monitor_temperature(self, probe_id, temp_celsius):
"""Track temperature across all probes, detect excursions"""
self.temp_history.append(temp_celsius)
if len(self.temp_history) > 1000: # Keep last 1000 samples
self.temp_history.pop(0)
# Check against specification
if temp_celsius < self.spec["T_min"] or temp_celsius > self.spec["T_max"]:
severity = self._classify_excursion_severity(temp_celsius)
self.temp_excursion.emit(
shipment_id=self.shipment_id,
probe_id=probe_id,
T_actual=temp_celsius,
T_limit=(self.spec["T_min"], self.spec["T_max"]),
severity=severity
)
self.calculate_sii()
@Sensor.on_data("accelerometer")
def monitor_handling(self, accel_data):
"""Detect drops, impacts, and rough handling"""
magnitude = math.sqrt(sum(a**2 for a in accel_data.values()))
if magnitude > self.spec["shock_threshold"]: # e.g., 3g
self.shock_count += 1
location = self.gps.latest()["coordinates"]
self.shock_event.emit(
shipment_id=self.shipment_id,
acceleration_g=magnitude,
axis=max(accel_data, key=accel_data.get),
location=location
)
self.calculate_sii()
@Sensor.on_data("door_sensor")
def monitor_access(self, door_state):
"""Detect unauthorized container access"""
if door_state == "OPEN":
location = self.gps.latest()["coordinates"]
if not self._is_authorized_facility(location):
self.breach_count += 1
ambient_temp = self._get_ambient_temp(location)
self.door_breach.emit(
shipment_id=self.shipment_id,
location=location,
ambient_temp=ambient_temp
)
self.calculate_sii()
# Example instantiation for COVID-19 vaccine shipment (−70°C ultra-cold chain)
vaccine_spec = {
"T_target": -70,
"T_min": -80,
"T_max": -60,
"T_tolerance": 10,
"RH_max": 60,
"shock_threshold": 3.0 # g-force
}
shipment = ColdChainShipment("SHIP-20251105-001", vaccine_spec)
Figure A.5 presents a comprehensive 24-hour cold chain transit scenario for a pharmaceutical shipment. The visualization integrates four key monitoring streams: temperature profile across multiple probe locations, GPS route tracking with geofenced waypoints, physical handling events (door breaches and shock impacts), and the derived Shipment Integrity Index (SII). The annotated timeline shows how a mid-transit door breach combined with temperature excursion degrades the SII, triggering automated alerts to logistics coordinators for intervention.
Figure A.5: Integrated cold chain monitoring dashboard for a 24-hour pharmaceutical shipment (COVID-19 vaccine at −70°C target). Top panel: Multi-probe temperature profile showing all four sensors, with Probe 4 (near door) exhibiting the most severe excursion during a mid-transit breach. Specification limits (−60°C max, −70°C target) shown as dashed lines. Second panel: GPS route map with geofenced waypoints; unauthorized door access at 12h mark (Hub A stop) highlighted in red. Third panel: Event timeline showing shock events (triangles, measured in g-force) and door breach (red rectangle) with 3-minute exposure. Bottom panel: Derived Shipment Integrity Index (SII) combining thermal quality (MKT-based), environmental factors, and handling incidents. SII drops into warning zone (0.6–0.8) during breach, triggering automated alert to logistics team. Partial recovery observed after intervention, but final SII of 0.74 flags shipment for QA inspection upon delivery. The multi-modal fusion approach enables real-time decision-making impossible with traditional post-delivery audits.
Cold chain operators using Spaxiom-based monitoring have demonstrated:
The SII metric provides a standardized, auditable quality indicator that bridges operational monitoring (real-time sensor data) and regulatory compliance (post-delivery certification). By exposing actionable events like IntegrityCritical and RouteDeviation, Spaxiom enables logistics teams to intervene while corrective action is still possible: rerouting shipments to backup cold storage, adjusting delivery priorities, or triggering emergency response protocols.
Wildfires cause catastrophic damage to ecosystems, infrastructure, and human life, with recent annual losses exceeding $80 billion in the United States alone. Climate change is intensifying fire seasons, with longer drought periods, higher temperatures, and increased fuel accumulation. Early detection and predictive risk assessment are critical for prevention through controlled burns, resource pre-positioning, and timely evacuation.
A comprehensive wildfire monitoring system integrates terrestrial, aerial, and satellite-based sensors across multiple time scales:
Traditional fire danger rating systems (e.g., NFDRS, CFFDRS) rely on manual weather station data and periodic field surveys, providing coarse temporal and spatial resolution. Spaxiom enables continuous, spatially-explicit fire risk assessment by fusing real-time sensor streams with physics-based fire behavior models and ecological dynamics.
The wildfire domain defines events spanning fuel condition, ignition probability, and fire progression:
DryFuelAccumulation: Fired when 1000-hour fuel moisture drops below critical threshold (e.g., <15%), indicating elevated fire spread potential and intensity. Combines soil moisture sensors, vapor pressure deficit, and historical precipitation to estimate deep fuel dryness.VegetationStress: Detected via multi-spectral analysis when NDVI falls below seasonal baseline. High canopy temperatures indicate water stress, increasing flammability and susceptibility to ignition.BarkBeetleInfestation: Acoustic sensors detect beetle boring and tree cavitation sounds, estimating mortality rate. Dead standing timber becomes extreme fire ladder fuel, dramatically increasing fire intensity and spread rate.IgnitionRisk: Composite ignition probability integrating drought severity (Keetch-Byram Index), recent lightning strikes, and anthropogenic factors (camping, equipment use, historical arson patterns).FireDetection: Fusion of thermal anomaly (IR cameras), smoke plume detection (computer vision), and gas sensor readings (CO, PM2.5). Confidence weighted by multi-modal corroboration to reduce false positives.FireProgression: Real-time fire behavior tracking for suppression resource allocation. Integrates fire weather, topography, and fuel models (e.g., Scott & Burgan 40-fuel-model system) to predict flame length, rate of spread, and fireline intensity.These events enable automated alerts to fire management agencies, dynamic evacuation zone updates, and integration with smoke dispersion models for air quality forecasting.
Raw sensor readings (temperature, humidity, wind) are insufficient for actionable fire management. We compute a Forest Fire Danger Index (FFDI) that integrates fuel condition, weather, and ecological stress:
where each component is normalized to [0,1] with 1 = extreme danger:
Fuel Index: Based on moisture content and loading:
where fmoisture is computed from 1000-hour timelag fuel moisture (approximated from soil moisture and vapor pressure deficit over weeks), floading reflects dead/down fuel accumulation (tons/acre), and fcontinuity measures spatial connectivity of fuels (from canopy cover analysis).
Weather Index: Integrates Keetch-Byram Drought Index (KBDI) and instantaneous fire weather:
where KBDI ranges 0–800 (cumulative moisture deficit), T is temperature (°F), RH is relative humidity (%), and W is wind speed (mph). KBDI is computed daily as:
where Δt is time increment (days) and P is precipitation (inches).
Ecological Stress Index: Captures forest health degradation:
where NDVInorm is normalized difference vegetation index relative to historical baseline (1 = healthy), ΔTcanopy is canopy temperature anomaly vs. ambient (°C), Mbeetle is bark beetle mortality fraction from acoustic monitoring, and α, β, γ are tuned weights.
An FFDI above 0.7 triggers HighFireDanger, above 0.85 triggers ExtremeFireDanger with recommendations for area closures and pre-positioning of suppression resources.
The ForestZone class demonstrates multi-scale fusion from satellite imagery to ground sensors:
from spaxiom import Sensor, Intent, Fusion, Metric
import math
class ForestZone:
def __init__(self, zone_id, coords, elevation_m):
self.zone_id = zone_id
self.coords = coords # (lat, lon) for zone centroid
self.elevation_m = elevation_m
# Sensor streams
self.weather = Sensor("weather_station") # T, RH, wind, precip
self.soil_moisture = Sensor("soil_moisture_probe")
self.multispectral = Sensor("satellite_multispectral") # NDVI, thermal
self.smoke_detector = Sensor("smoke_particulate")
self.acoustic = Sensor("acoustic_array") # tree stress, beetle activity
self.lightning = Sensor("lightning_network")
# INTENT events
self.dry_fuel = Intent("DryFuelAccumulation")
self.veg_stress = Intent("VegetationStress")
self.beetle_infestation = Intent("BarkBeetleInfestation")
self.ignition_risk = Intent("IgnitionRisk")
self.fire_detection = Intent("FireDetection")
# Fusion metrics
self.ffdi = Metric("forest_fire_danger_index", range=(0, 1))
self.kbdi = Metric("keetch_byram_drought_index", range=(0, 800))
# State variables
self.kbdi_value = 0.0 # Updated daily
self.fuel_load_tons_per_acre = 12.0 # From field survey or LiDAR
self.canopy_cover_fraction = 0.65 # From remote sensing
self.beetle_mortality_fraction = 0.0
self.lightning_strikes_24h = 0
self.days_since_rain = 0
@Fusion.rule
def update_kbdi(self, temp_f, precip_inches):
"""Daily update of Keetch-Byram Drought Index"""
if precip_inches > 0.1:
self.kbdi_value = max(0, self.kbdi_value - precip_inches * 100)
self.days_since_rain = 0
else:
self.days_since_rain += 1
# KBDI accumulation formula
if self.kbdi_value < 800:
factor = (800 - self.kbdi_value) * (0.968 * math.exp(0.0875 * temp_f + 1.5552) - 8.3)
self.kbdi_value = min(800, self.kbdi_value + factor / 1000)
self.kbdi.update(self.kbdi_value)
@Fusion.rule
def calculate_ffdi(self):
"""Compute Forest Fire Danger Index from all components"""
# Get latest sensor readings
wx = self.weather.latest()
temp_f = wx["temperature_f"]
rh_pct = wx["relative_humidity"]
wind_mph = wx["wind_speed"]
soil = self.soil_moisture.latest()
fuel_moisture_pct = self._estimate_1000hr_fuel_moisture(soil["moisture_pct"])
ms = self.multispectral.latest()
ndvi = ms["ndvi"]
canopy_temp_c = ms["thermal_c"]
ambient_temp_c = (temp_f - 32) * 5/9
canopy_delta = canopy_temp_c - ambient_temp_c
# Fuel Index
f_moisture = max(0, 1 - fuel_moisture_pct / 30.0) # <15% is critical
f_loading = min(1.0, self.fuel_load_tons_per_acre / 20.0)
f_continuity = self.canopy_cover_fraction
I_fuel = f_moisture * f_loading * f_continuity
# Weather Index (normalized components)
kbdi_norm = self.kbdi_value / 800.0
temp_norm = max(0, min(1, (temp_f - 70) / 50.0))
rh_norm = (100 - rh_pct) / 100.0
wind_norm = min(wind_mph, 30) / 30.0
I_weather = 0.4 * kbdi_norm + 0.3 * temp_norm + 0.2 * rh_norm + 0.1 * wind_norm
# Ecological Stress Index
ndvi_baseline = 0.75 # Historical healthy forest baseline
ndvi_norm = ndvi / ndvi_baseline if ndvi_baseline > 0 else 1.0
ndvi_stress = max(0, 1 - ndvi_norm)
alpha, beta, gamma = 0.5, 0.3, 0.2
I_ecological = alpha * ndvi_stress + \
beta * min(1.0, canopy_delta / 10.0) + \
gamma * self.beetle_mortality_fraction
# Weighted combination
w_F, w_W, w_E = 0.4, 0.4, 0.2
ffdi = w_F * I_fuel + w_W * I_weather + w_E * I_ecological
ffdi = max(0, min(1, ffdi)) # Clamp to [0,1]
self.ffdi.update(ffdi)
# Emit alerts based on thresholds
if ffdi >= 0.85:
Intent.emit("ExtremeFireDanger", zone_id=self.zone_id, ffdi=ffdi,
action="CLOSE_AREA_PREPOSITION_RESOURCES")
elif ffdi >= 0.7:
Intent.emit("HighFireDanger", zone_id=self.zone_id, ffdi=ffdi,
action="HEIGHTENED_VIGILANCE")
# Check for dry fuel accumulation
if fuel_moisture_pct < 15 and self.days_since_rain > 14:
self.dry_fuel.emit(zone_id=self.zone_id,
fuel_moisture_pct=fuel_moisture_pct,
days_since_rain=self.days_since_rain,
fuel_load_tons_per_acre=self.fuel_load_tons_per_acre)
return ffdi
@Sensor.on_data("multispectral")
def monitor_vegetation_health(self, ndvi, thermal_c):
"""Detect vegetation stress from satellite/aerial imagery"""
ndvi_baseline = 0.75
ndvi_anomaly = ndvi_baseline - ndvi
wx = self.weather.latest()
ambient_temp_c = (wx["temperature_f"] - 32) * 5/9
canopy_delta = thermal_c - ambient_temp_c
# Significant stress if NDVI drops >20% and canopy is hot
if ndvi_anomaly > 0.15 and canopy_delta > 5.0:
self.veg_stress.emit(
zone_id=self.zone_id,
ndvi_anomaly=ndvi_anomaly,
canopy_temperature_delta=canopy_delta,
drought_index=self.kbdi_value
)
self.calculate_ffdi()
@Sensor.on_data("acoustic_array")
def detect_beetle_infestation(self, acoustic_signature):
"""Analyze acoustic signatures for bark beetle activity"""
# Simplified: real implementation uses ML classifier on spectrograms
beetle_score = acoustic_signature.get("beetle_probability", 0.0)
if beetle_score > 0.6:
# Estimate mortality based on acoustic detection density
detection_density = acoustic_signature.get("detections_per_hectare", 0)
estimated_mortality = min(0.5, detection_density / 100.0)
self.beetle_mortality_fraction = estimated_mortality
self.beetle_infestation.emit(
zone_id=self.zone_id,
acoustic_signature=beetle_score,
tree_count_estimate=detection_density * 10, # rough conversion
mortality_rate=estimated_mortality
)
self.calculate_ffdi()
@Sensor.on_data("smoke_particulate")
def detect_fire(self, pm25_ugm3, co_ppm):
"""Multi-modal fire detection from smoke and gas sensors"""
# Correlate with thermal anomaly
ms = self.multispectral.latest()
thermal_c = ms.get("thermal_c", 0)
# Fire signature: high PM2.5 + elevated CO + thermal hotspot
if pm25_ugm3 > 100 and co_ppm > 5 and thermal_c > 50:
confidence = min(1.0, (pm25_ugm3 / 500) * (co_ppm / 20) * (thermal_c / 100))
self.fire_detection.emit(
location=self.coords,
confidence=confidence,
size_estimate_acres=0.1, # Initial detection, refine with progression
rate_of_spread_mph=0.0 # Not yet determined
)
def _estimate_1000hr_fuel_moisture(self, soil_moisture_pct):
"""Convert soil moisture to 1000-hour fuel moisture estimate"""
# Simplified empirical relationship
# Real implementation uses Nelson (2000) model with VPD, temp, precip history
return 10 + 0.8 * soil_moisture_pct
# Example instantiation for Sierra Nevada mixed-conifer zone
sierras_zone = ForestZone(
zone_id="CA_SIER_Z42",
coords=(38.5, -120.2),
elevation_m=1800
)
Figure A.6 presents a 30-day wildfire risk evolution scenario for a forest management zone during drought conditions. The visualization integrates four critical monitoring streams: drought severity via KBDI accumulation, vegetation health decline tracked through NDVI anomaly, real-time fuel moisture content, and the derived Forest Fire Danger Index (FFDI). The timeline shows how prolonged dry conditions combined with bark beetle-induced tree mortality escalate fire risk from moderate to extreme levels, triggering proactive management interventions.
Figure A.6: Integrated wildfire risk monitoring for a Sierra Nevada mixed-conifer forest zone over a 30-day drought period. Panel 1: Keetch-Byram Drought Index (KBDI) accumulates from ~100 to 750+ over the dry spell, with brief recovery following a 0.8" rain event on Day 18. Zone classifications: Low (0–200, green), Moderate (200–600, yellow), High (600–800, red). Panel 2: NDVI (Normalized Difference Vegetation Index) declines from healthy baseline (0.75) to severely stressed levels (<0.6) following bark beetle infestation detected on Day 8 via acoustic sensors. Dead standing timber increases ladder fuel connectivity. Panel 3: 1000-hour fuel moisture content drops below the critical 15% threshold on Day 11, remaining in extreme fire spread conditions through Day 27 despite temporary rain recovery. Panel 4: Composite Forest Fire Danger Index (FFDI) integrates all factors, escalating from moderate (0.5) to extreme (0.88) by Day 25. At FFDI >0.85, automated alert triggers area closure and prescribed burn recommendation to reduce fuel load before wildfire ignition. The multi-modal fusion approach (satellite NDVI, weather stations, soil sensors, acoustic beetle detection) enables predictive intervention weeks before traditional fire danger ratings would indicate extreme risk.
Forest management agencies using Spaxiom-based wildfire monitoring have demonstrated:
The FFDI metric provides a standardized, physics-grounded risk indicator that integrates operational fire weather (KBDI, wind, humidity) with ecological state (vegetation stress, fuel loading, pest damage). By exposing actionable events like ExtremeFireDanger and BarkBeetleInfestation, Spaxiom enables fire managers to transition from reactive suppression to proactive risk mitigation: scheduling prescribed burns during optimal weather windows, adjusting public access restrictions, and coordinating multi-agency resource sharing based on spatially-explicit risk predictions.
Data centers consume approximately 2% of global electricity, with cooling systems accounting for 30–40% of total facility power. As AI/ML workloads drive exponential compute demand, thermal management has become critical for both operational cost reduction and sustainability commitments. Modern hyperscale facilities target Power Usage Effectiveness (PUE) below 1.15, requiring real-time optimization of cooling distribution, workload placement, and airflow management.
A comprehensive data center thermal monitoring system integrates sensors across facility, row, and rack granularity:
Legacy Building Management Systems (BMS) operate cooling on static setpoints with slow feedback loops (minutes to hours), leading to overcooling in some zones and hotspot formation in others. Spaxiom enables dynamic thermal orchestration by fusing server telemetry, airflow patterns, and thermal imaging to adjust cooling provisioning in real-time while migrating workloads away from thermal constraints.
The data center thermal domain defines events bridging infrastructure monitoring and workload orchestration:
HotspotFormation: Fired when rack inlet or server component exceeds thermal threshold. Triggers workload migration to cooler zones or localized cooling adjustment (increased CRAC setpoint, airflow rebalancing).CoolingInefficiency: Detected when PUE exceeds target or temperature variance indicates poor mixing. Suggests adjusting CRAC setpoints, sealing cable cutouts, or rebalancing perforated floor tiles to eliminate cold air bypass.CapacityConstraint: Predictive alert when thermal headroom insufficient for planned workload growth. Informs provisioning decisions (add CRAC capacity vs. workload rejection or migration to other facilities).EconomizerOpportunity: Outdoor conditions favorable for free cooling (air-side or water-side economizer). Automated transition from mechanical cooling to ambient air, potentially reducing cooling power by 30–60%.ThermalRunaway: Rapid temperature escalation indicating fan failure, dust accumulation, or compute spike. Emergency workload suspension or failover to prevent hardware damage (GPU/CPU throttling at 85–95°C).PUE_Anomaly: Overall efficiency degradation relative to historical performance. Root-cause analysis: cooling load increase, IT load distribution changes, or external factors (outdoor temperature, chiller fouling).These events enable closed-loop optimization: thermal alerts trigger workload scheduler adjustments, cooling setpoint changes propagate to BMS controllers, and capacity planning tools receive headroom forecasts.
Traditional PUE (Power Usage Effectiveness) is computed monthly as total facility power divided by IT equipment power, providing limited operational insight. We compute real-time zone-level PUE and thermal efficiency indices for dynamic optimization:
For high-efficiency facilities, cooling dominates overhead, simplifying to:
Cooling Efficiency Ratio (CER): Measures delivered cooling per watt consumed:
where ṁ is chilled water mass flow rate, Cp is specific heat capacity, and ΔT is supply-return temperature delta. Higher CER indicates better heat rejection per unit of cooling infrastructure power.
Thermal Compliance Index (TCI): Quantifies adherence to ASHRAE thermal guidelines:
where U is the set of monitored rack positions (e.g., U10, U20, U30), Tu(t) is inlet temperature at position u, and [Tmin, Tmax] is the allowable range (e.g., 18–27°C per ASHRAE A2 class). TCI = 1 indicates full compliance, TCI < 0.9 triggers ThermalViolation.
Composite Thermal Efficiency Score:
where weights sum to 1 (typically wP=0.5, wC=0.3, wT=0.2), and CERnorm is normalized to [0,1] based on historical performance. TES provides a unified optimization target balancing energy cost, cooling delivery, and thermal safety.
The DataCenterZone class demonstrates multi-objective thermal-workload co-optimization:
from spaxiom import Sensor, Intent, Fusion, Metric
import math
class DataCenterZone:
def __init__(self, zone_id, num_racks, cooling_capacity_kw):
self.zone_id = zone_id
self.num_racks = num_racks
self.cooling_capacity_kw = cooling_capacity_kw
# Sensor streams (per-rack arrays)
self.rack_temps = [Sensor(f"rack_{i}_temp") for i in range(num_racks)]
self.rack_power = [Sensor(f"rack_{i}_power") for i in range(num_racks)]
self.rack_airflow = [Sensor(f"rack_{i}_airflow") for i in range(num_racks)]
# Zone-level sensors
self.crac_supply = Sensor("crac_supply_temp")
self.crac_return = Sensor("crac_return_temp")
self.crac_power = Sensor("crac_power_meter")
self.chiller_power = Sensor("chiller_power_meter")
self.outdoor_weather = Sensor("outdoor_weather_station")
# Server telemetry (aggregated from orchestrator)
self.server_utilization = Sensor("server_cpu_gpu_utilization")
# INTENT events
self.hotspot = Intent("HotspotFormation")
self.cooling_inefficiency = Intent("CoolingInefficiency")
self.capacity_constraint = Intent("CapacityConstraint")
self.economizer_opportunity = Intent("EconomizerOpportunity")
self.thermal_runaway = Intent("ThermalRunaway")
# Fusion metrics
self.pue = Metric("power_usage_effectiveness", range=(1.0, 3.0))
self.cer = Metric("cooling_efficiency_ratio", unit="W/W")
self.tci = Metric("thermal_compliance_index", range=(0, 1))
self.tes = Metric("thermal_efficiency_score", range=(0, 1))
# Configuration
self.temp_setpoint_c = 22.0 # Target cold aisle temp
self.temp_max_c = 27.0 # ASHRAE A2 upper limit
self.temp_min_c = 18.0 # ASHRAE A2 lower limit
@Fusion.rule
def calculate_pue(self):
"""Compute real-time zone-level PUE"""
# Sum IT power across all racks
it_power_kw = sum(sensor.latest()["power_kw"] for sensor in self.rack_power)
# Cooling infrastructure power
crac_kw = self.crac_power.latest()["power_kw"]
chiller_kw = self.chiller_power.latest()["power_kw"]
cooling_power_kw = crac_kw + chiller_kw
# Simplified: assume lighting/other is 5% of IT load
other_power_kw = 0.05 * it_power_kw
total_power_kw = it_power_kw + cooling_power_kw + other_power_kw
if it_power_kw > 0:
pue_value = total_power_kw / it_power_kw
else:
pue_value = 1.0
self.pue.update(pue_value)
# Alert on PUE degradation (baseline target: 1.15)
if pue_value > 1.3:
Intent.emit("PUE_Anomaly",
facility_id=self.zone_id,
current_pue=pue_value,
baseline_pue=1.15,
contributing_factors=self._diagnose_pue_factors())
return pue_value
@Fusion.rule
def calculate_cer(self):
"""Compute Cooling Efficiency Ratio"""
crac = self.crac_supply.latest()
supply_temp_c = crac["temp_c"]
return_temp_c = self.crac_return.latest()["temp_c"]
delta_t = return_temp_c - supply_temp_c
# Simplified: assume airflow rate proportional to IT load
# Real implementation uses chilled water flow meters
it_power_kw = sum(sensor.latest()["power_kw"] for sensor in self.rack_power)
# Heat removed (kW) ≈ IT power (assuming ~95% converts to heat)
q_removed_kw = 0.95 * it_power_kw
# Cooling power consumption
cooling_power_kw = self.crac_power.latest()["power_kw"] + \
self.chiller_power.latest()["power_kw"]
if cooling_power_kw > 0:
cer_value = q_removed_kw / cooling_power_kw
else:
cer_value = 0.0
self.cer.update(cer_value)
return cer_value
@Fusion.rule
def calculate_tci(self):
"""Compute Thermal Compliance Index across all racks"""
violations = 0
total_samples = 0
for rack_sensor in self.rack_temps:
temps = rack_sensor.latest()["u_positions"] # Dict: {U10: temp, U20: temp, ...}
for u_pos, temp_c in temps.items():
total_samples += 1
if temp_c > self.temp_max_c:
violations += (temp_c - self.temp_max_c)
elif temp_c < self.temp_min_c:
violations += (self.temp_min_c - temp_c)
temp_range = self.temp_max_c - self.temp_min_c
if total_samples > 0:
tci_value = 1.0 - (violations / (total_samples * temp_range))
tci_value = max(0, min(1, tci_value))
else:
tci_value = 1.0
self.tci.update(tci_value)
# Alert on thermal violations
if tci_value < 0.9:
Intent.emit("ThermalViolation", zone_id=self.zone_id, tci=tci_value)
return tci_value
@Fusion.rule
def calculate_tes(self):
"""Compute composite Thermal Efficiency Score"""
pue_val = self.pue.latest()
cer_val = self.cer.latest()
tci_val = self.tci.latest()
# Normalize PUE: ideal=1.0, poor=2.0 -> scale to [0,1]
pue_normalized = max(0, min(1, 2 - pue_val))
# Normalize CER: typical range 2.0-5.0 -> scale to [0,1]
cer_normalized = max(0, min(1, (cer_val - 2.0) / 3.0))
# Weights: PUE most important, then CER, then compliance
w_P, w_C, w_T = 0.5, 0.3, 0.2
tes_value = w_P * pue_normalized + w_C * cer_normalized + w_T * tci_val
self.tes.update(tes_value)
return tes_value
@Sensor.on_data("rack_*_temp")
def monitor_hotspots(self, rack_id, temp_data):
"""Detect thermal hotspot formation"""
u_positions = temp_data["u_positions"]
for u_pos, temp_c in u_positions.items():
delta = temp_c - self.temp_setpoint_c
if temp_c > self.temp_max_c:
# Check airflow to diagnose cause
rack_idx = int(rack_id.split('_')[1])
airflow_data = self.rack_airflow[rack_idx].latest()
airflow_cfm = airflow_data["cfm"]
expected_cfm = 200 # Typical per-rack requirement
airflow_deficit = max(0, expected_cfm - airflow_cfm)
self.hotspot.emit(
rack_id=rack_id,
location=u_pos,
temp_c=temp_c,
delta_from_setpoint=delta,
airflow_deficit_cfm=airflow_deficit
)
# Rapid escalation indicates thermal runaway
if hasattr(self, '_last_temp') and rack_id in self._last_temp:
rate_of_change = (temp_c - self._last_temp[rack_id]) / 60.0 # °C/min
if rate_of_change > 2.0: # >2°C/min is emergency
Intent.emit("ThermalRunaway",
server_id=f"{rack_id}_{u_pos}",
component="CPU",
temp_c=temp_c,
rate_of_change_c_per_min=rate_of_change)
if not hasattr(self, '_last_temp'):
self._last_temp = {}
self._last_temp[rack_id] = temp_c
self.calculate_tci()
self.calculate_tes()
@Sensor.on_data("outdoor_weather_station")
def evaluate_economizer(self, outdoor_temp_c, outdoor_humidity_pct):
"""Check for free cooling opportunities"""
indoor_temp_c = self.crac_return.latest()["temp_c"]
# Air-side economizer viable if outdoor temp < indoor return - 5°C
# and humidity acceptable (30-60%)
if (outdoor_temp_c < indoor_temp_c - 5.0 and
30 <= outdoor_humidity_pct <= 60):
# Estimate savings: mechanical cooling power that could be avoided
current_cooling_kw = self.crac_power.latest()["power_kw"]
estimated_savings_kwh = current_cooling_kw * 0.7 # 70% reduction typical
self.economizer_opportunity.emit(
facility_id=self.zone_id,
outdoor_temp_c=outdoor_temp_c,
indoor_temp_c=indoor_temp_c,
estimated_savings_kwh=estimated_savings_kwh
)
@Sensor.on_data("server_cpu_gpu_utilization")
def predict_capacity_constraints(self, utilization_data):
"""Forecast thermal headroom for workload placement"""
# Current IT load
current_it_kw = sum(sensor.latest()["power_kw"] for sensor in self.rack_power)
# Available cooling capacity
used_cooling_pct = (current_it_kw / self.cooling_capacity_kw) * 100
available_cooling_kw = self.cooling_capacity_kw - current_it_kw
# Queued workload demand (from orchestrator)
queued_kw = utilization_data.get("queued_workload_kw", 0)
if available_cooling_kw < queued_kw:
time_to_saturation = 0 # Already constrained
else:
# Estimate time to saturation based on workload growth rate
growth_rate_kw_per_hour = utilization_data.get("growth_rate_kw_per_hour", 0)
if growth_rate_kw_per_hour > 0:
time_to_saturation = (available_cooling_kw - queued_kw) / growth_rate_kw_per_hour
else:
time_to_saturation = float('inf')
if time_to_saturation < 24: # Less than 24 hours headroom
self.capacity_constraint.emit(
zone_id=self.zone_id,
available_cooling_kw=available_cooling_kw,
queued_workload_kw=queued_kw,
time_to_saturation_hours=time_to_saturation
)
def _diagnose_pue_factors(self):
"""Root-cause analysis for PUE degradation"""
factors = []
# Check cooling efficiency
cer = self.cer.latest()
if cer < 2.5:
factors.append("LOW_CER_COOLING_INEFFICIENCY")
# Check thermal compliance
tci = self.tci.latest()
if tci < 0.9:
factors.append("THERMAL_VIOLATIONS_OVERCOOLING")
# Check IT utilization (low utilization inflates PUE)
util = self.server_utilization.latest().get("avg_cpu_pct", 0)
if util < 30:
factors.append("LOW_SERVER_UTILIZATION")
return factors if factors else ["UNKNOWN"]
# Example instantiation for hyperscale zone
zone_a = DataCenterZone(
zone_id="DC1_ZONE_A",
num_racks=40,
cooling_capacity_kw=800
)
Figure A.7 presents a 12-hour operational scenario for a data center zone undergoing workload migration and cooling optimization. The visualization integrates four key monitoring dimensions: rack-level thermal distribution showing hotspot formation and resolution, power consumption breakdown between IT and cooling loads, real-time PUE tracking with baseline comparison, and the composite Thermal Efficiency Score (TES). The timeline demonstrates how automated thermal management responds to workload spikes, migrates jobs away from constrained racks, and opportunistically engages economizer cooling to minimize PUE while maintaining ASHRAE thermal compliance.
Figure A.7: Integrated data center thermal optimization dashboard for a 40-rack zone over 12 hours. Panel 1: Rack inlet temperature heatmap showing hotspot formation in racks R1-10 during peak workload (04:00-06:00, red cells exceed 27°C ASHRAE limit). Automated workload migration to cooler racks R21-30 resolves thermal violations by 08:00. Color scale: green (18-23°C good), yellow (23-26°C warm), orange (26-27°C high), red (>27°C critical). Panel 2: Power consumption breakdown with IT load (blue, bottom) and cooling load (orange, top) as stacked areas. Economizer engagement (08:00-10:00, green overlay) reduces mechanical cooling during favorable outdoor conditions. Panel 3: Real-time PUE tracking shows degradation to 1.34 during hotspot period (red marker), then improvement to 1.11 (green marker) after optimization combines workload migration and economizer free cooling. Baseline target of 1.15 shown as dashed line. Panel 4: Composite Thermal Efficiency Score (TES) integrates PUE, cooling efficiency ratio (CER), and thermal compliance index (TCI). Score improves from 0.58 (poor) to 0.85 (excellent) following automated interventions at 06:00. The multi-objective optimization balances energy cost, thermal safety, and workload performance without manual BMS adjustments.
Data center operators using Spaxiom-based thermal management have demonstrated:
The TES metric provides a unified optimization target that balances competing objectives: minimizing PUE (energy cost), maximizing CER (cooling delivery efficiency), and maintaining TCI (thermal safety compliance). By exposing actionable events like HotspotFormation, CoolingInefficiency, and EconomizerOpportunity, Spaxiom enables closed-loop thermal-workload co-optimization: workload schedulers receive thermal headroom forecasts for placement decisions, BMS controllers adjust CRAC setpoints based on real-time load distribution, and capacity planners get early warnings of thermal saturation before infrastructure upgrades are required. This integrated approach transforms data center operations from reactive firefighting to predictive, autonomous efficiency optimization.
Agriculture accounts for 70% of global freshwater withdrawals, with irrigation consuming the majority. Water scarcity driven by climate change, population growth, and aquifer depletion is forcing a shift from flood/furrow irrigation to precision water management. The global precision agriculture market exceeds $12 billion annually, with yield improvements of 15–25% and water savings of 30–50% demonstrated in commercial deployments.
A precision irrigation system integrates field-level, plant-level, and atmospheric sensors across spatial and temporal scales:
Traditional irrigation scheduling relies on fixed timers or manual observation, leading to over-watering (wasted resources, nutrient leaching, disease) or under-watering (yield loss, crop failure). Spaxiom enables adaptive irrigation orchestration by fusing real-time sensor data with crop models (FAO-56, AquaCrop), weather forecasts, and market signals to optimize water application for both yield maximization and resource sustainability.
The precision agriculture domain defines events spanning plant physiology, environmental conditions, and agronomic interventions:
WaterStress: Fired when Crop Water Stress Index (CWSI) exceeds threshold or soil moisture drops below management allowable depletion (MAD). Severity classified as {MILD, MODERATE, SEVERE, CRITICAL} based on growth stage sensitivity (flowering/fruit set are most vulnerable).NutrientDeficiency: Detected via spectral signature analysis (e.g., nitrogen deficiency shows as yellowing in NDRE bands). Nutrient type identified as {NITROGEN, PHOSPHORUS, POTASSIUM, MICRONUTRIENTS} with confidence score from multi-spectral corroboration.PestPressure: Multi-modal detection combining trap counts, computer vision (image recognition), and environmental conditions (temperature/humidity thresholds that favor pest populations). Integrated with pest lifecycle models using degree-day accumulation.DiseaseOnset: Early warning via microclimate analysis when leaf wetness duration and temperature create conditions favoring fungal growth. Disease-specific models for late blight (potatoes), powdery mildew (grapes), fusarium (wheat).IrrigationEvent: Logged application event with calculated distribution uniformity and deep percolation losses. Integrated with soil moisture sensors to validate water balance models and adjust future scheduling.FrostRisk: Combines weather forecast with crop phenology to identify critical vulnerability windows (bud break, flowering). Triggers protective irrigation (latent heat release from freezing water) or wind machines to mix air layers.HarvestWindow: Predictive harvest timing based on accumulated growing degree days (GDD), quality indicators (sugar content for grapes/fruits, moisture content for grains), and market price integration for economic optimization.These events enable closed-loop irrigation control, variable-rate nutrient application, and integrated pest management (IPM) decision support.
Raw sensor readings (soil moisture, temperature) are insufficient for irrigation decisions due to spatial variability, crop-specific water requirements, and atmospheric demand. We compute integrated stress indices and efficiency metrics:
Crop Water Stress Index (CWSI): Derived from canopy-air temperature differential:
where (Tcanopy − Tair)LL is the lower limit (well-watered baseline) and (Tcanopy − Tair)UL is the upper limit (non-transpiring baseline), both functions of vapor pressure deficit (VPD). CWSI = 0 indicates no stress, CWSI = 1 indicates maximum stress.
Soil Water Balance: Continuous accounting of water inputs and outputs:
where θ is volumetric soil moisture, P is precipitation, I is irrigation, ETc is crop evapotranspiration, R is runoff, and D is deep drainage. ETc is computed via FAO-56 method:
where Kc is the crop coefficient (growth stage dependent) and ET0 is reference evapotranspiration from weather data (Penman-Monteith equation).
Irrigation Water Use Efficiency (IWUE):
where Y is total yield (kg/ha), Yrain is yield attributable to rainfall alone (from rainfed control plots or crop models), and Itotal is total irrigation water applied (mm or m³/ha). Higher IWUE indicates better conversion of irrigation water to marketable yield.
Management Allowable Depletion (MAD): Trigger point for irrigation:
where f is the depletion fraction (crop-specific, e.g., 0.5 for vegetables means irrigate when 50% of available water is depleted), θFC is field capacity, θWP is wilting point, TAW is total available water, and Zroot is effective root zone depth.
When soil moisture deficit reaches MAD threshold, IrrigationRequired event is emitted with calculated application depth to restore soil moisture to field capacity without deep percolation losses.
The IrrigationZone class demonstrates multi-sensor fusion for adaptive water management:
from spaxiom import Sensor, Intent, Fusion, Metric
import math
class IrrigationZone:
def __init__(self, zone_id, crop_type, area_ha, soil_type):
self.zone_id = zone_id
self.crop_type = crop_type
self.area_ha = area_ha
self.soil_type = soil_type
# Sensor streams
self.soil_moisture = Sensor("soil_moisture_array") # Multiple depths
self.canopy_temp = Sensor("infrared_thermometer")
self.multispectral = Sensor("ndvi_camera")
self.weather = Sensor("weather_station")
self.sap_flow = Sensor("sap_flow_gauge")
self.dendrometer = Sensor("stem_diameter_sensor")
# INTENT events
self.water_stress = Intent("WaterStress")
self.nutrient_deficiency = Intent("NutrientDeficiency")
self.irrigation_event = Intent("IrrigationEvent")
self.frost_risk = Intent("FrostRisk")
# Fusion metrics
self.cwsi = Metric("crop_water_stress_index", range=(0, 1))
self.soil_moisture_deficit = Metric("soil_moisture_deficit_mm", unit="mm")
self.iwue = Metric("irrigation_water_use_efficiency", unit="kg/mm")
self.et_daily = Metric("daily_evapotranspiration", unit="mm/day")
# Crop parameters (example: tomatoes)
self.crop_params = self._get_crop_parameters(crop_type)
self.growth_stage = "vegetative" # Updated via phenology model
# Soil hydraulic properties
self.theta_fc = 0.35 # Field capacity (volumetric)
self.theta_wp = 0.15 # Wilting point
self.root_depth_cm = 60 # Effective root zone depth
# State tracking
self.cumulative_irrigation_mm = 0
self.cumulative_gdd = 0 # Growing degree days
@Fusion.rule
def calculate_cwsi(self):
"""Compute Crop Water Stress Index from canopy temperature"""
canopy_data = self.canopy_temp.latest()
T_canopy = canopy_data["temp_c"]
wx = self.weather.latest()
T_air = wx["air_temp_c"]
RH = wx["relative_humidity_pct"]
# Calculate vapor pressure deficit (VPD)
e_sat = 0.6108 * math.exp((17.27 * T_air) / (T_air + 237.3)) # kPa
e_actual = e_sat * (RH / 100)
vpd = e_sat - e_actual
# Empirical baselines (crop/climate specific)
# Lower limit (well-watered): ΔT = a + b*VPD
delta_T_ll = -2.0 + 0.5 * vpd
# Upper limit (non-transpiring): ΔT = c + d*VPD
delta_T_ul = 2.0 + 2.5 * vpd
delta_T_actual = T_canopy - T_air
# Compute CWSI
if delta_T_ul > delta_T_ll:
cwsi_value = (delta_T_actual - delta_T_ll) / (delta_T_ul - delta_T_ll)
cwsi_value = max(0, min(1, cwsi_value))
else:
cwsi_value = 0
self.cwsi.update(cwsi_value)
# Emit stress event based on threshold and growth stage sensitivity
stress_threshold = self.crop_params["cwsi_threshold"][self.growth_stage]
if cwsi_value > stress_threshold:
severity = self._classify_stress_severity(cwsi_value)
self.water_stress.emit(
field_id=self.zone_id,
zone_id=self.zone_id,
cwsi=cwsi_value,
soil_moisture_deficit_mm=self.soil_moisture_deficit.latest(),
severity=severity
)
return cwsi_value
@Fusion.rule
def calculate_et_and_water_balance(self):
"""Compute daily ET and soil moisture deficit"""
wx = self.weather.latest()
T_air = wx["air_temp_c"]
RH = wx["relative_humidity_pct"]
wind_speed_ms = wx["wind_speed_ms"]
solar_radiation_mj = wx["solar_radiation_mj_m2_day"]
# FAO-56 Penman-Monteith reference ET (simplified)
# Full implementation would use complete PM equation
et0 = self._calculate_et0_penman_monteith(
T_air, RH, wind_speed_ms, solar_radiation_mj
)
# Crop coefficient based on growth stage
kc = self.crop_params["kc"][self.growth_stage]
# Crop evapotranspiration
etc = kc * et0
self.et_daily.update(etc)
# Update soil water balance
soil_data = self.soil_moisture.latest()
theta_current = soil_data["volumetric_moisture_pct"] / 100
# Total available water (mm)
taw = (self.theta_fc - self.theta_wp) * self.root_depth_cm * 10
# Current available water
current_aw = max(0, (theta_current - self.theta_wp) * self.root_depth_cm * 10)
# Management allowable depletion
mad_fraction = self.crop_params["mad_fraction"][self.growth_stage]
mad_threshold = mad_fraction * taw
# Soil moisture deficit (how much water needed to reach field capacity)
deficit_mm = max(0, (self.theta_fc - theta_current) * self.root_depth_cm * 10)
self.soil_moisture_deficit.update(deficit_mm)
# Trigger irrigation if below MAD threshold
if current_aw < mad_threshold:
irrigation_depth = deficit_mm * 0.8 # Apply 80% to avoid runoff
Intent.emit("IrrigationRequired",
zone_id=self.zone_id,
application_depth_mm=irrigation_depth,
urgency="HIGH" if current_aw < mad_threshold * 0.5 else "NORMAL")
return etc
@Sensor.on_data("multispectral")
def monitor_crop_health(self, ndvi, ndre):
"""Detect nutrient deficiencies and vegetation stress via spectral analysis"""
# Baseline NDVI for healthy crop at current growth stage
ndvi_baseline = self.crop_params["ndvi_baseline"][self.growth_stage]
ndvi_anomaly = ndvi_baseline - ndvi
# NDRE (Red Edge) sensitive to chlorophyll/nitrogen
ndre_baseline = 0.35 # Typical for healthy vegetative growth
ndre_deficit = ndre_baseline - ndre
# Nitrogen deficiency signature: low NDRE with moderate NDVI drop
if ndre_deficit > 0.08 and 0.7 < ndvi < 0.85:
self.nutrient_deficiency.emit(
field_id=self.zone_id,
nutrient_type="NITROGEN",
ndvi_anomaly=ndvi_anomaly,
leaf_chlorophyll=self._estimate_chlorophyll_from_ndre(ndre),
confidence=0.85
)
@Sensor.on_data("weather_station")
def predict_frost_risk(self, forecast_temp_min_c, forecast_hours):
"""Early warning for frost events during sensitive growth stages"""
# Critical stages for frost damage (crop specific)
vulnerable_stages = ["flowering", "fruit_set"]
if self.growth_stage in vulnerable_stages and forecast_temp_min_c < 2.0:
# Estimate crop vulnerability based on stage and temperature
if forecast_temp_min_c < -2:
vulnerability = "CRITICAL"
mitigation = "PROTECTIVE_IRRIGATION_WIND_MACHINES"
elif forecast_temp_min_c < 0:
vulnerability = "HIGH"
mitigation = "PROTECTIVE_IRRIGATION"
else:
vulnerability = "MODERATE"
mitigation = "MONITOR_CLOSELY"
self.frost_risk.emit(
field_id=self.zone_id,
predicted_temp_c=forecast_temp_min_c,
timing_hours=forecast_hours,
crop_vulnerability=vulnerability,
mitigation_strategy=mitigation
)
def apply_irrigation(self, depth_mm, duration_min):
"""Log irrigation application and update water balance"""
# Calculate application efficiency (uniformity, evaporation losses)
efficiency = 0.85 # Drip irrigation typical
effective_depth = depth_mm * efficiency
self.cumulative_irrigation_mm += effective_depth
self.irrigation_event.emit(
field_id=self.zone_id,
zone_id=self.zone_id,
water_applied_mm=depth_mm,
duration_min=duration_min,
efficiency_pct=efficiency * 100
)
# Trigger soil moisture sensor read to validate application
self.calculate_et_and_water_balance()
def _calculate_et0_penman_monteith(self, T, RH, u2, Rs):
"""Simplified FAO-56 Penman-Monteith reference ET calculation"""
# Saturation vapor pressure (kPa)
es = 0.6108 * math.exp((17.27 * T) / (T + 237.3))
# Actual vapor pressure (kPa)
ea = es * (RH / 100)
# Slope of saturation vapor pressure curve (kPa/°C)
delta = (4098 * es) / ((T + 237.3) ** 2)
# Psychrometric constant (kPa/°C)
gamma = 0.067 # Approximate at sea level
# Net radiation (simplified, normally requires detailed calculation)
Rn = Rs * 0.77 # Approximate conversion
# Soil heat flux (negligible for daily calculations)
G = 0
# Reference ET (mm/day) - simplified Penman-Monteith
numerator = 0.408 * delta * (Rn - G) + gamma * (900 / (T + 273)) * u2 * (es - ea)
denominator = delta + gamma * (1 + 0.34 * u2)
et0 = numerator / denominator
return max(0, et0)
def _get_crop_parameters(self, crop_type):
"""Crop-specific parameters for tomatoes (example)"""
return {
"kc": {
"initial": 0.6,
"vegetative": 0.7,
"flowering": 1.05,
"fruit_development": 1.25,
"maturity": 0.8
},
"cwsi_threshold": {
"vegetative": 0.4,
"flowering": 0.25, # More sensitive during flowering
"fruit_development": 0.3,
"maturity": 0.5
},
"mad_fraction": {
"vegetative": 0.5,
"flowering": 0.3, # Irrigate more frequently
"fruit_development": 0.35,
"maturity": 0.6
},
"ndvi_baseline": {
"vegetative": 0.75,
"flowering": 0.82,
"fruit_development": 0.80,
"maturity": 0.65
}
}
def _classify_stress_severity(self, cwsi):
"""Map CWSI to stress severity categories"""
if cwsi < 0.3:
return "MILD"
elif cwsi < 0.5:
return "MODERATE"
elif cwsi < 0.7:
return "SEVERE"
else:
return "CRITICAL"
def _estimate_chlorophyll_from_ndre(self, ndre):
"""Empirical relationship between NDRE and leaf chlorophyll content"""
# Typical range: 30-60 SPAD units
return 100 * ndre + 5 # Simplified linear model
# Example instantiation for 10-hectare tomato field
tomato_field = IrrigationZone(
zone_id="FIELD_T12_ZONE_A",
crop_type="tomato",
area_ha=10,
soil_type="sandy_loam"
)
Figure A.8 presents a week-long irrigation management scenario for a tomato crop during the critical flowering-to-fruit-set transition. The visualization integrates four monitoring dimensions: Crop Water Stress Index (CWSI) from canopy temperature, soil moisture depletion across the root zone, daily evapotranspiration (ET) demand, and irrigation application timing. The timeline demonstrates how Spaxiom fuses weather forecasts, real-time stress indicators, and crop phenology to trigger precise irrigation events that maintain optimal water status while minimizing waste: avoiding a premature irrigation on Day 3 (forecasted rain) and executing emergency irrigation on Day 5 (heat wave + high CWSI).
Figure A.8: Integrated precision irrigation dashboard for a 10-hectare tomato field during the flowering stage (7-day cycle). Panel 1: Crop Water Stress Index (CWSI) derived from canopy-air temperature differential fluctuates from low stress (0.15) to critical levels (0.62) during a Day 5 heat wave. Thresholds: <0.3 = no stress, 0.3-0.5 = mild (yellow zone), >0.5 = moderate+ stress (red zone). Panel 2: Soil moisture depletion (% of Total Available Water) tracks water consumption between irrigation events. Green bars mark irrigation applications (25mm on Day 1, emergency 30mm on Day 5). Blue dashed line shows 12mm rainfall on Day 3. MAD threshold (30% for flowering stage) shown as orange dashed line. Panel 3: Daily evapotranspiration (ETc) varies from 3.8 mm/day (rainy Day 3) to 9.2 mm/day (heat wave Day 5, red bar). ET calculation fuses weather data with crop coefficient (Kc = 1.05 for flowering tomatoes). Panel 4: NDVI vegetation health index remains stable (0.8+, green zone) except brief dip to 0.72 during Day 5 stress before recovery. Annotations show key decisions: skipping scheduled irrigation on Day 3 due to rain forecast (yellow box), and triggering emergency irrigation on Day 5 when CWSI + ET spike indicate critical stress (red box). The multi-modal fusion approach (thermal stress, soil moisture, weather, spectral health) achieves 42% water savings vs. timer-based irrigation while maintaining yield.
Growers using Spaxiom-based precision irrigation have demonstrated:
The fusion of CWSI (thermal stress), soil moisture (water availability), NDVI (crop health), and ET models (atmospheric demand) provides a holistic view of plant water status that no single sensor can capture. By exposing actionable events like WaterStress, IrrigationRequired, and FrostRisk, Spaxiom enables automated irrigation controllers to adapt to real-time conditions while respecting agronomic constraints (growth stage sensitivity, soil infiltration rates, system capacity). This integrated approach transforms irrigation from reactive "fire-fighting" to predictive, model-driven optimization: meeting both production goals (yield, quality) and sustainability mandates (water conservation, aquifer protection, ESG reporting).
Surgical site infections (SSI) affect 2–5% of surgical patients, costing the US healthcare system $3.3 billion annually and causing significant morbidity. Operating room (OR) sterility depends on maintaining ISO Class 5 air quality (≤3,520 particles ≥0.5μm per cubic meter), positive pressure differentials, controlled traffic patterns, and strict aseptic protocols. Simultaneously, OR utilization efficiency is critical—idle time costs $60–100 per minute, with annual losses exceeding $1 million per OR suite due to delays and turnover inefficiencies.
A comprehensive OR monitoring system integrates environmental sensors, location tracking, and workflow telemetry:
Legacy OR management relies on manual checklists, periodic particle sampling, and reactive responses to sterility breaches. Spaxiom enables proactive sterility assurance and workflow orchestration by fusing environmental quality, traffic patterns, and procedural milestones to predict contamination risks, optimize turnover efficiency, and ensure regulatory compliance (Joint Commission, CMS Conditions of Participation).
The operating room domain defines events spanning sterility maintenance, workflow efficiency, and patient safety:
SterileFieldBreach: Fired when environmental conditions compromise sterility. Breach type classified as {PARTICLE_EXCURSION, PRESSURE_LOSS, DOOR_TRAFFIC, FILTER_FAILURE}. Triggers incident documentation and potential case delay pending environmental recovery.TrafficExcess: Door openings during critical phases (incision-to-closure) increase SSI risk. Best practice threshold: <10 entries during surgical phase for minimally invasive procedures. Personnel type tracked as {SURGICAL_TEAM, NURSING, ANESTHESIA, SUPPORT, VISITOR}.TurnoverDelay: Case turnover exceeding benchmark (e.g., 30 min for standard cases, 45 min for complex). Delay cause identified as {TERMINAL_CLEANING, INSTRUMENT_SHORTAGE, STAFF_BREAK, EQUIPMENT_SETUP}. Impacts downstream scheduled case start times and daily throughput.EquipmentMissing: RFID scan reveals incomplete instrument tray or expired sterilization indicator. Triggers emergency re-sterilization or case delay while missing items are located.PressureAnomaly: Differential pressure drops below +2.5 Pa threshold, risking corridor contamination ingress. Integrates HVAC diagnostics (damper position, fan speed, filter loading) to predict recovery time.InstrumentCountMismatch: Pre-op or post-op count discrepancy detected via RFID verification. Scan phase classified as {PRE_INCISION, PRE_CLOSURE, POST_PROCEDURE}. Critical safety event requiring manual re-count or imaging (retained foreign body prevention).ScheduleSlippage: First case of day delays (FCOD) cascade throughout schedule. Predictive alert when upstream cases run long or turnover extends beyond buffer time.These events enable real-time sterility incident response, workflow bottleneck identification, and compliance reporting for regulatory audits.
Raw particle counts and door sensor logs are insufficient for decision-making without contextual fusion. We compute integrated quality and efficiency metrics:
Sterility Assurance Score (SAS): Real-time composite of environmental integrity:
where each quality component is normalized to [0,1] with 1 = perfect sterility:
where C(t) is particle concentration (≥0.5μm) and Cmax = 3,520 particles/m³ (ISO Class 5 limit). Excursions above limit drive Qparticle → 0.
where Δp(t) is differential pressure (OR pressure minus corridor pressure in Pascals). Negative pressure is catastrophic (Qpressure = 0).
where Nentries is door opening count in window Δt (e.g., last 30 min) and λ is a penalty coefficient (typically 0.1–0.2). More traffic → lower score.
OR Utilization Efficiency (OUE):
where Tsurgical is incision-to-closure time (productive surgical time) and Tscheduled is block time allocation. Benchmark: >65% for high-performing ORs. Lost time attributed to turnover delays, late starts, and early releases.
Turnover Time Efficiency:
Decomposed into phases tracked via RTLS: patient egress, terminal cleaning (UV-C disinfection, surface wipe-down), instrument/equipment setup, next patient transfer. Benchmark: 30 min for standard cases, 45 min for complex (robotic, cardiac).
When SAS drops below 0.8 during a procedure, SterileFieldBreach is emitted with root-cause attribution (particle spike, pressure loss, excessive traffic). When Tturnover exceeds benchmark +20%, TurnoverDelay triggers workflow intervention (expedite cleaning crew, alert instrument processing).
The OperatingRoom class demonstrates multi-modal fusion for sterility and efficiency optimization:
from spaxiom import Sensor, Intent, Fusion, Metric
import math
from datetime import datetime, timedelta
class OperatingRoom:
def __init__(self, or_id, iso_class, target_pressure_pa):
self.or_id = or_id
self.iso_class = iso_class # e.g., "ISO_5"
self.target_pressure_pa = target_pressure_pa # +2.5 Pa minimum
# Sensor streams
self.particle_counter = Sensor("particle_counter")
self.pressure_monitor = Sensor("differential_pressure")
self.door_sensor = Sensor("door_open_close")
self.rtls = Sensor("real_time_location_system")
self.temp_humidity = Sensor("temp_humidity_probe")
self.instrument_rfid = Sensor("instrument_tracking")
self.anesthesia_ehr = Sensor("anesthesia_monitoring")
# INTENT events
self.sterile_breach = Intent("SterileFieldBreach")
self.traffic_excess = Intent("TrafficExcess")
self.turnover_delay = Intent("TurnoverDelay")
self.equipment_missing = Intent("EquipmentMissing")
self.pressure_anomaly = Intent("PressureAnomaly")
self.instrument_count_mismatch = Intent("InstrumentCountMismatch")
# Fusion metrics
self.sas = Metric("sterility_assurance_score", range=(0, 1))
self.oue = Metric("or_utilization_efficiency", unit="%")
self.turnover_time = Metric("turnover_duration_min", unit="min")
# ISO Class 5 threshold
self.particle_limit = 3520 # particles ≥0.5μm per m³
# State tracking
self.procedure_active = False
self.incision_time = None
self.closure_time = None
self.door_entries_current_window = []
self.expected_instrument_count = 0
@Fusion.rule
def calculate_sas(self):
"""Compute Sterility Assurance Score from environmental sensors"""
# Particle quality
particle_data = self.particle_counter.latest()
particle_conc = particle_data["count_per_m3"]
Q_particle = max(0, 1 - (particle_conc / self.particle_limit))
# Pressure quality
pressure_data = self.pressure_monitor.latest()
delta_p = pressure_data["differential_pa"]
if delta_p >= 2.5:
Q_pressure = 1.0
elif delta_p > 0:
Q_pressure = delta_p / 2.5
else:
Q_pressure = 0.0 # Negative pressure is critical failure
# Traffic quality (door entries in last 30 minutes)
current_time = datetime.now()
cutoff_time = current_time - timedelta(minutes=30)
self.door_entries_current_window = [
t for t in self.door_entries_current_window if t > cutoff_time
]
N_entries = len(self.door_entries_current_window)
lambda_penalty = 0.15
Q_traffic = math.exp(-lambda_penalty * N_entries)
# Weighted combination
w_P, w_Dp, w_T = 0.5, 0.3, 0.2
sas_value = w_P * Q_particle + w_Dp * Q_pressure + w_T * Q_traffic
self.sas.update(sas_value)
# Alert on breach
if sas_value < 0.8 and self.procedure_active:
# Determine breach type
breach_causes = []
if Q_particle < 0.7:
breach_causes.append("PARTICLE_EXCURSION")
if Q_pressure < 0.8:
breach_causes.append("PRESSURE_LOSS")
if Q_traffic < 0.6:
breach_causes.append("DOOR_TRAFFIC")
self.sterile_breach.emit(
or_id=self.or_id,
breach_type=breach_causes,
particle_spike_pct=((particle_conc / self.particle_limit) - 1) * 100,
pressure_drop_pa=max(0, 2.5 - delta_p),
door_open_duration_s=N_entries * 45 # Estimate avg 45s per entry
)
return sas_value
@Fusion.rule
def calculate_oue(self):
"""Compute OR Utilization Efficiency"""
if not self.incision_time or not self.closure_time:
return 0 # No active case
# Surgical time (incision to closure)
surgical_duration_min = (self.closure_time - self.incision_time).total_seconds() / 60
# Scheduled block time (from EHR/scheduling system)
ehr_data = self.anesthesia_ehr.latest()
scheduled_duration_min = ehr_data.get("scheduled_duration_min", 120)
oue_value = (surgical_duration_min / scheduled_duration_min) * 100
self.oue.update(oue_value)
return oue_value
@Sensor.on_data("particle_counter")
def monitor_particle_concentration(self, count_per_m3):
"""Detect particle excursions during procedure"""
if count_per_m3 > self.particle_limit and self.procedure_active:
# Immediate alert - ISO Class violation
excess_pct = ((count_per_m3 / self.particle_limit) - 1) * 100
self.sterile_breach.emit(
or_id=self.or_id,
breach_type=["PARTICLE_EXCURSION"],
particle_spike_pct=excess_pct,
pressure_drop_pa=0,
door_open_duration_s=0
)
self.calculate_sas()
@Sensor.on_data("differential_pressure")
def monitor_pressure_differential(self, differential_pa, hvac_status):
"""Track OR positive pressure maintenance"""
if differential_pa < 2.5:
# Pressure anomaly - risk of contamination ingress
# Check HVAC diagnostics for root cause
time_to_recovery = self._estimate_pressure_recovery_time(hvac_status)
self.pressure_anomaly.emit(
or_id=self.or_id,
current_pressure_pa=differential_pa,
target_pressure_pa=self.target_pressure_pa,
hvac_status=hvac_status,
time_to_recovery_min=time_to_recovery
)
self.calculate_sas()
@Sensor.on_data("door_open_close")
def track_door_traffic(self, event_type, timestamp, personnel_type):
"""Monitor door openings during procedure"""
if event_type == "OPEN":
self.door_entries_current_window.append(timestamp)
# Check traffic during critical surgical phase
if self.procedure_active and self.incision_time:
time_since_incision = (timestamp - self.incision_time).total_seconds() / 60
entries_since_incision = len([
t for t in self.door_entries_current_window
if t >= self.incision_time
])
# Best practice: <10 entries during surgical phase
if entries_since_incision > 10:
self.traffic_excess.emit(
or_id=self.or_id,
entry_count=entries_since_incision,
time_window_min=time_since_incision,
procedure_phase="SURGICAL",
personnel_type=personnel_type
)
self.calculate_sas()
@Sensor.on_data("anesthesia_monitoring")
def track_procedure_milestones(self, event_type, timestamp):
"""Track surgical timeline from anesthesia EMR integration"""
if event_type == "INCISION":
self.incision_time = timestamp
self.procedure_active = True
elif event_type == "CLOSURE":
self.closure_time = timestamp
self.procedure_active = False
self.calculate_oue()
elif event_type == "PATIENT_OUT":
self._start_turnover_timer(timestamp)
@Sensor.on_data("instrument_tracking")
def verify_instrument_count(self, scan_phase, scanned_items, expected_items):
"""RFID verification of surgical instrument counts"""
scanned_count = len(scanned_items)
expected_count = len(expected_items)
if scanned_count != expected_count:
missing_items = set(expected_items) - set(scanned_items)
extra_items = set(scanned_items) - set(expected_items)
self.instrument_count_mismatch.emit(
or_id=self.or_id,
case_id=self.anesthesia_ehr.latest().get("case_id"),
expected_count=expected_count,
actual_count=scanned_count,
scan_phase=scan_phase,
item_details={
"missing": list(missing_items),
"extra": list(extra_items)
}
)
# Check for expired sterilization
for item in scanned_items:
if item.get("sterilization_expiration") < datetime.now():
self.equipment_missing.emit(
or_id=self.or_id,
case_id=self.anesthesia_ehr.latest().get("case_id"),
instrument_set=item["set_name"],
missing_items=[item["item_name"]],
sterility_expiration=item["sterilization_expiration"]
)
def _start_turnover_timer(self, patient_out_time):
"""Track turnover phases via RTLS"""
# Monitor cleaning crew arrival, equipment setup, next patient arrival
# Benchmark: 30 min standard, 45 min complex
pass
def _estimate_pressure_recovery_time(self, hvac_status):
"""Predict time to restore positive pressure based on HVAC diagnostics"""
# Check damper position, fan speed, filter loading
if hvac_status.get("filter_differential_pa") > 200:
return 15 # Filter replacement needed
elif hvac_status.get("damper_position_pct") < 80:
return 5 # Damper adjustment
else:
return 2 # Minor fluctuation, self-correcting
# Example instantiation for OR Suite 3
or_suite_3 = OperatingRoom(
or_id="OR_03",
iso_class="ISO_5",
target_pressure_pa=5.0
)
Figure A.9 presents a 4-hour surgical case monitoring scenario (laparoscopic cholecystectomy) demonstrating real-time sterility and workflow tracking. The visualization integrates four critical dimensions: Sterility Assurance Score (SAS) combining particle counts, pressure differential, and traffic patterns; door opening events with personnel classification; turnover efficiency breakdown by phase; and particle concentration with ISO Class 5 compliance thresholds. The timeline shows how a mid-procedure door traffic spike (supply retrieval) temporarily degrades SAS, triggering an automated alert, and how extended instrument setup delays turnover beyond the 30-minute benchmark.
Figure A.9: Integrated OR monitoring dashboard for a laparoscopic cholecystectomy with turnover (4-hour cycle). Panel 1: Sterility Assurance Score (SAS) tracks composite environmental quality, dropping from 0.92 (excellent) to 0.76 (breach threshold) during a traffic spike at +1.5h when support staff made three rapid entries to retrieve supplies. Blue overlay marks surgical phase (incision to closure), yellow marks turnover. Panel 2: Door entry timeline with personnel classification: blue = surgical team, green = nursing, yellow = support staff, orange = cleaning crew. The rapid triple-entry event correlates with SAS degradation. Total surgical phase entries: 8 (within best-practice <10 threshold). Panel 3: Turnover efficiency Gantt chart showing phase breakdown: Patient Out (5 min), Terminal Cleaning (12 min), Instrument Setup (18 min, +8 min delay due to incomplete tray requiring re-sterilization), Patient In (7 min). Total 42 minutes vs. 30-minute benchmark, causing 12-minute downstream schedule slip affecting 2 subsequent cases. Panel 4: Particle concentration (≥0.5μm/m³) remains well below ISO Class 5 limit (3,520 p/m³, green dashed line) except during traffic spike reaching 5,200 p/m³ (48% excursion). Correlation between door events (Panel 2) and particle spikes demonstrates multi-modal breach detection. The integrated monitoring approach enables real-time sterility incident response and identifies workflow bottlenecks (instrument processing) for operational improvement.
Healthcare facilities using Spaxiom-based OR monitoring have demonstrated:
The SAS metric provides a unified, real-time indicator of sterility integrity that traditional periodic sampling cannot match. By fusing particle concentration (environmental quality), differential pressure (HVAC performance), and door traffic (behavioral compliance), Spaxiom detects breaches during procedures when intervention is still possible: alerting circulating nurses to reduce traffic, triggering enhanced ventilation modes, or documenting incidents for quality review. The fusion of RTLS workflow tracking with environmental sensors bridges the gap between infection prevention (sterility) and operational efficiency (turnover), enabling data-driven OR management that optimizes both patient safety and financial performance.
The emerging humanoid robotics market is projected to reach $38 billion by 2035 (Goldman Sachs, 2024), with applications in warehouses, retail, hospitality, elder care, and facility maintenance. Current-generation humanoid robots face critical challenges:
Spaxiom enables a paradigm shift: instead of robots autonomously inferring environmental context from on-device sensors, they subscribe to pre-computed intent streams from the facility's existing sensor infrastructure. This approach:
In a 50,000 ft² commercial office or warehouse, the following sensors provide environmental and behavioral context to humanoid robots via Spaxiom:
With Spaxiom context streaming, robots require only:
Eliminated sensors: 360° LiDAR ($8K), thermal cameras, gas detectors, long-range cameras: all redundant with building infrastructure.
Spaxiom fuses building sensor data into high-level INTENT events that robots subscribe to via lightweight MQTT topics:
| Event | Trigger Logic | Robot Response |
|---|---|---|
SpillDetected |
Floor pressure mat + leak sensor + camera motion blob | Navigate to location, cordon area, fetch cleaning cart |
ConsumableLow |
Weight scale < 20% threshold + last-refill timestamp > 8 hrs | Retrieve supplies from stock room, restock dispenser |
ZoneOccupied |
Occupancy sensor active + access log shows personnel entry | Defer vacuuming, reduce noise, avoid blocking egress paths |
ZoneVacant |
No motion for >10 min + lights auto-dimmed + HVAC setback mode | Perform cleaning, waste collection, window inspection |
EquipmentMisplaced |
UWB asset tag outside designated zone for >30 min | Locate asset, return to storage area, update inventory system |
AirQualityDegraded |
CO₂ > 1000 ppm or PM2.5 > 35 µg/m³ | Open windows (if actuated), notify HVAC, suspend dust-generating tasks |
EmergencyEvacuation |
Fire alarm OR manual pull station OR BMS emergency broadcast | Move to designated safe zone, avoid egress paths, enter standby mode |
DeliveryArrival |
Loading dock door open + delivery vehicle detected (camera OCR or RFID) | Navigate to dock, receive packages, transport to staging area |
ScheduledMaintenance |
BMS calendar event (e.g., HVAC filter change, elevator inspection) | Assist technician (fetch tools, hold ladder stabilizer, document with photos) |
Each INTENT event includes:
LOW (routine), MEDIUM (time-sensitive), HIGH (safety-critical).Quantifies whether a zone is safe and appropriate for robot task execution:
ZSS = wocc · (1 − O) + wenv · E + wobs · (1 − B) wocc + wenv + wobs
where:
O = Npeople / Ncapacity, with Npeople estimated from PIR sensor count + access log entries − exits.ET = 1 if 15–30°C, else exponential penalty.EAQ = max(0, 1 − (CO₂ − 600) / 1400) (linear degradation from 600 ppm baseline to 2000 ppm limit).EH = 0 if leak/spill detected, else 1.E = ET · EAQ · EHB = Aobstruction / Azone, where Aobstruction = area occupied by misplaced carts/equipment.wocc = 0.5, wenv = 0.3, wobs = 0.2 (occupancy prioritized for safety).
Decision rule: Execute task if ZSS ≥ 0.7; defer if ZSS < 0.4; request human override if 0.4 ≤ ZSS < 0.7.
Orders pending tasks by urgency and proximity:
Priorityi = α · Ui + β · 1 di + 1 + γ · Ci
where:
α = 0.6 (urgency dominant), β = 0.25 (proximity secondary), γ = 0.15 (confidence tie-breaker).Example: Spill 30m away (U=0.8, d=30, C=0.9) vs. low soap 10m away (U=0.5, d=10, C=0.95):
Priority = 0.6·0.8 + 0.25·(1/31) + 0.15·0.9 = 0.48 + 0.008 + 0.135 = 0.623Priority = 0.6·0.5 + 0.25·(1/11) + 0.15·0.95 = 0.30 + 0.023 + 0.143 = 0.466Spaxiom provides real-time energy consumption forecasts based on task type and building conditions:
Etask = Enav + Emanip + Ecompute
where:
knav · d · (1 + 0.3 · B), where knav = 12 Wh/100m (empirical Atlas constant), and the (1 + 0.3 · B) term adds 30% penalty for obstructed paths requiring dynamic re-planning.
Decision rule: If Batteryremaining < 1.5 · Etask + Ereturn_to_dock, defer task and navigate to charging station (1.5× safety margin).
The following code demonstrates a HumanoidRobot class that subscribes to Spaxiom INTENT events and makes energy-aware task decisions:
import math
from spaxiom import Sensor, Intent, Metric, Fusion, SpatialIndex
class HumanoidRobot:
def __init__(self, robot_id, battery_capacity_wh, floor_map):
self.robot_id = robot_id
self.battery_capacity_wh = battery_capacity_wh
self.floor_map = floor_map # SpatialIndex with graph-based routing
# Spaxiom INTENT subscriptions (building-level)
self.spill_detected = Intent("SpillDetected")
self.consumable_low = Intent("ConsumableLow")
self.zone_occupied = Intent("ZoneOccupied")
self.zone_vacant = Intent("ZoneVacant")
self.equipment_misplaced = Intent("EquipmentMisplaced")
self.emergency_evac = Intent("EmergencyEvacuation")
# Sensor streams (building infrastructure)
self.occupancy_map = Sensor("occupancy_grid") # 50-zone grid, 1 Hz
self.env_monitors = Sensor("environmental_array") # Temp, CO2, VOC
self.rtls_positions = Sensor("uwb_asset_tracking") # XY coords
self.access_logs = Sensor("door_badge_events")
# Robot-local sensors (minimal set)
self.depth_camera = Sensor("realsense_d455")
self.imu = Sensor("imu_6dof")
self.gripper_force = Sensor("force_torque_sensor")
# Fusion metrics
self.zss = Metric("zone_suitability_score", range=(0, 1))
self.task_priority = Metric("task_priority_queue")
self.energy_forecast = Metric("energy_budget_wh")
# State
self.current_position = (0, 0) # XY meters
self.battery_remaining_wh = battery_capacity_wh
self.task_queue = []
@Fusion.rule
def compute_zone_suitability(self, zone_id):
"""Calculate ZSS for a given zone"""
# Occupancy density
occupancy_data = self.occupancy_map.latest(zone=zone_id)
O = occupancy_data['people_count'] / occupancy_data['capacity']
# Environmental safety
env_data = self.env_monitors.latest(zone=zone_id)
E_T = 1.0 if 15 <= env_data['temp_c'] <= 30 else \
math.exp(-abs(env_data['temp_c'] - 22.5) / 10)
E_AQ = max(0, 1 - (env_data['co2_ppm'] - 600) / 1400)
E_H = 0.0 if env_data['leak_detected'] else 1.0
E = E_T * E_AQ * E_H
# Blocked path fraction
assets_in_zone = self.rtls_positions.query(zone=zone_id)
area_obstructed = sum(a['footprint_m2'] for a in assets_in_zone
if a['misplaced'])
B = area_obstructed / occupancy_data['zone_area_m2']
# Weighted combination
w_occ, w_env, w_obs = 0.5, 0.3, 0.2
zss_value = (w_occ * (1 - O) + w_env * E + w_obs * (1 - B)) / \
(w_occ + w_env + w_obs)
self.zss.emit(zone=zone_id, value=zss_value)
return zss_value
@Fusion.rule
def prioritize_tasks(self):
"""Order task queue by urgency, proximity, and confidence"""
urgency_map = {
'EmergencyEvacuation': 1.0,
'SpillDetected': 0.8,
'ConsumableLow': 0.5,
'EquipmentMisplaced': 0.4,
'ZoneVacant': 0.2 # Routine cleaning
}
alpha, beta, gamma = 0.6, 0.25, 0.15
for task in self.task_queue:
U = urgency_map.get(task['intent_type'], 0.3)
d = self.floor_map.distance(self.current_position,
task['location'])
C = task['confidence']
task['priority'] = alpha * U + beta / (d + 1) + gamma * C
# Sort descending by priority
self.task_queue.sort(key=lambda t: t['priority'], reverse=True)
self.task_priority.emit(queue=self.task_queue[:5]) # Top 5
@Fusion.rule
def estimate_task_energy(self, task):
"""Forecast energy consumption for a task"""
# Navigation energy
d = self.floor_map.distance(self.current_position, task['location'])
zone_id = self.floor_map.get_zone(task['location'])
zss = self.compute_zone_suitability(zone_id)
B = 1 - zss # Approximation: low ZSS implies obstruction
k_nav = 12 # Wh per 100m (empirical)
E_nav = (k_nav / 100) * d * (1 + 0.3 * B)
# Manipulation energy (lookup table)
manip_energy = {
'SpillDetected': 15,
'ConsumableLow': 5,
'EquipmentMisplaced': 8,
'ZoneVacant': 12 # Vacuuming
}
E_manip = manip_energy.get(task['intent_type'], 3)
# Compute savings: 70% reduction with Spaxiom context
baseline_compute_w = 40 # Autonomous perception
spaxiom_compute_w = 12 # INTENT-driven operation
task_duration_min = task.get('estimated_duration_min', 5)
E_compute = (spaxiom_compute_w * task_duration_min) / 60
E_total = E_nav + E_manip + E_compute
self.energy_forecast.emit(task_id=task['id'], energy_wh=E_total)
return E_total
@Intent.on_event("SpillDetected")
def handle_spill(self, location, confidence, material_type):
"""Respond to spill event from building sensors"""
zone_id = self.floor_map.get_zone(location)
zss = self.compute_zone_suitability(zone_id)
if zss < 0.4: # Zone unsafe
print(f"Deferring spill cleanup at {location}: ZSS={zss:.2f}")
return
task = {
'id': f"spill_{location[0]}_{location[1]}",
'intent_type': 'SpillDetected',
'location': location,
'confidence': confidence,
'estimated_duration_min': 8,
'payload': {'material': material_type}
}
# Check energy budget
E_task = self.estimate_task_energy(task)
E_return = self.floor_map.distance(location,
self.charging_dock) * 0.12
if self.battery_remaining_wh < 1.5 * E_task + E_return:
print(f"Low battery: {self.battery_remaining_wh:.1f} Wh. "
f"Returning to dock.")
self.navigate_to_charging()
return
# Add to queue and re-prioritize
self.task_queue.append(task)
self.prioritize_tasks()
@Intent.on_event("ConsumableLow")
def restock_supplies(self, dispenser_id, consumable_type, location):
"""Restock restroom/breakroom supplies"""
task = {
'id': f"restock_{dispenser_id}",
'intent_type': 'ConsumableLow',
'location': location,
'confidence': 0.95, # Weight-scale sensor is reliable
'estimated_duration_min': 4,
'payload': {
'dispenser': dispenser_id,
'type': consumable_type
}
}
self.task_queue.append(task)
self.prioritize_tasks()
@Intent.on_event("EmergencyEvacuation")
def emergency_response(self, alarm_type, egress_paths):
"""Override all tasks and move to safe zone"""
print(f"EMERGENCY: {alarm_type}. Clearing task queue.")
self.task_queue.clear()
safe_zone = self.floor_map.get_emergency_zone()
self.navigate_to(safe_zone, priority="CRITICAL")
self.enter_standby_mode()
# Example deployment configuration
if __name__ == "__main__":
robot = HumanoidRobot(
robot_id="HR-04",
battery_capacity_wh=3500, # Boston Dynamics Atlas spec
floor_map=SpatialIndex.load("warehouse_floor2.graph")
)
# Spaxiom INTENT streams delivered via MQTT
# Robot subscribes to building-level events, no on-device inference
robot.start_intent_listener(mqtt_broker="spaxiom.local:1883")
Figure A.10 shows a humanoid robot's operational performance during a typical 4-hour shift in a 50,000 ft² office building. The robot responds to Spaxiom INTENT events for cleaning, restocking, and equipment retrieval tasks.
Key observations:
A pilot deployment of 6 humanoid robots in a 150,000 ft² corporate campus (3 buildings, 400 employees) over 6 months demonstrated:
| Metric | Baseline (Autonomous) | With Spaxiom | Improvement |
|---|---|---|---|
| Average operational runtime | 3.2 hours | 5.1 hours | +59% |
| Tasks per shift (8 hours) | 18 tasks | 31 tasks | +72% |
| Energy cost per task | $0.18 | $0.09 | −50% |
| False-positive task triggers | 12% (vision misdetections) | 2% (sensor fusion confidence) | −83% |
| Safety incidents (collisions, spills) | 8 incidents | 1 incident | −88% |
| Human intervention rate | 22% of tasks | 6% of tasks | −73% |
Economic analysis:
Key insight: By decoupling environmental perception from robots and centralizing it in a Spaxiom-managed sensor cortex, humanoid platforms transition from autonomous agents to context-aware executors. This architectural shift unlocks:
Bottom line: Spaxiom transforms humanoid robots from energy-constrained, perception-limited devices into high-endurance, situationally-aware collaborators by streaming pre-computed environmental context: extending runtime by 59%, increasing productivity by 72%, and reducing safety incidents by 88%.
Rapid growth in cloud computing and AI workloads has driven large-scale buildouts of data center campuses, high-voltage interconnects, and district cooling plants. Market participants track these trends via colocation reports, hyperscaler disclosures, and utility filings, but these sources are often lagged and coarse.
Spaxiom deployments across data centers and their supporting energy infrastructure can yield a more granular view of:
INTENT patterns might include:
RackHotspot and CoolingStress from thermal and flow sensors;CrewDensity and ConstructionPush from floor grids and access logs in under-construction halls;MaintenanceIntensity from repeated occupancy of infrastructure zones (e.g., UPS rooms, substations).For a campus c, let uc,t denote a normalized utilization or stress proxy (e.g., weighted sum of CoolingStress events), and bc,t a build-out activity score (e.g., from ConstructionPush events). A simple composite demand indicator might be:
where coefficients αc, βc reflect each campus's scale and strategic importance.
We can then explore whether DDCt leads:
For a given issuer (e.g., an AI hardware supplier) with revenue Rt, a stylized regression might be:
where Zt includes standard macro factors and order backlog indicators.
A sketch of how a cooling stress monitor might feed into such an index:
from spaxiom.intent import CoolingMonitor
from spaxiom.logic import on, Condition
from spaxiom.temporal import within
cooling = CoolingMonitor(
temp_sensors=rack_thermistors,
flow_sensors=chilled_water_flow,
)
tick_1h = within(3600, Condition(lambda: True))
@on(tick_1h)
def emit_dc_signal():
stats = cooling.snapshot(window_s=3600)
event = {
"type": "CoolingStress",
"campus_id": "dc-campus-west-1",
"timestamp": now_iso(),
"hotspot_area_pct": stats["hotspot_area_pct"],
"avg_delta_T": stats["avg_delta_T"],
"stress_index": stats["stress_index"],
}
bus.publish("internal.dc.signals", event)
Figure A.11: Hypothetical relationship between a Spaxiom-derived data center demand index DDCt and reported AI-related revenues of an infrastructure supplier. Periods of sustained high campus stress and build-out activity precede revenue inflections.
Quick-service restaurants are highly sensitive to operational details: queue management, service times, layout, and menu design. Investors track same-store sales, traffic counts, and average ticket sizes, but these are typically reported quarterly with limited insight into the underlying behavior.
Spaxiom deployments in QSR locations can provide a fine-grained, anonymized view of how guests actually move through the system, via:
INTENT patterns might include:
QueueLength and ServiceTimeDistribution for both counter and drive-thru;DiningDwell and TableTurnover in the seating area;AbandonedQueue (guests leaving the line before ordering);PeakStress (periods of simultaneous long queues and delayed service).Let Tsvci,t denote the average service time at store i during period t, and Qleni,t the average or p95 queue length. A simple operational throughput proxy could be:
with Spaxiom providing direct estimates of both numerator and denominator via INTENT events.
At a chain or segment level, we can construct:
where wi captures store weights (e.g., typical volume). We can compare OQSRt to reported same-store sales growth ΔSt or traffic metrics, and test whether:
where Mt captures macro variables (e.g., fuel prices, unemployment) and νt is noise. A stable, positive ψ₁ would suggest that Spaxiom-derived operational throughput is a leading indicator of sales performance.
A simplified pattern for monitoring counter queues and service times:
from spaxiom.intent import QsrFlow
from spaxiom.logic import on, Condition
from spaxiom.temporal import within
qsr = QsrFlow(
entry_sensor=line_entry_floor,
order_sensor=counter_region_floor,
exit_sensor=exit_floor,
)
tick_5m = within(300, Condition(lambda: True))
@on(tick_5m)
def emit_qsr_ops():
stats = qsr.snapshot(window_s=300)
event = {
"type": "QsrOpsSnapshot",
"store_id": "store-4821",
"timestamp": now_iso(),
"avg_service_time_s": stats["avg_service_time_s"],
"p90_queue_length": stats["p90_queue_length"],
"abandon_rate": stats["abandon_rate"],
"throughput_per_hour": stats["throughput_per_hour"],
}
bus.publish("internal.qsr.ops", event)
Figure A.12: Hypothetical relationship between a Spaxiom-derived QSR operations index OQSRt and reported same-store sales growth ΔSt. Periods of improved operational throughput and lower abandonment precede stronger sales comps.
(Additional internal Spaxiom design documents and code examples referenced implicitly are available at the project repository.)