Hunting for Foxes with Sheaves

Michael Robinson

Introduction Michael Robinson is an Associate Professor in the Department of Mathemat- To a radio amateur (or “ham”), fox hunting has nothing ics and Statistics at American University. His email address is michaelr to do with animals. It is a sport in which individuals race @american.edu. All figures can be reproduced using The Jupyter Notebook at: https://github For permission to reprint this article, please contact: .com/kb1dds/foxsheaf. [email protected]. Communicated by Notices Associate Editor Emilie Purvine. DOI: https://doi.org/10.1090/noti1867

MAY 2019 NOTICES OF THE AMERICAN MATHEMATICAL SOCIETY 661 each other to locate a hidden radio transmitter on a known We are left with the need for a general deterministic frequency. Since hams are encouraged to design and build method for finding the fox from a small number ofmea- their own equipment, the typical fox hunt involves a vari- surements. This article explains how to meet this need us- ety of different receivers and antennas with different ca- ing sheaves, mathematical objects that describe local consis- pabilities. Some of these can display the received signal tency within data. We can perform data fusion for any , strength from the hidden transmitter (loosely measuring though the fox hunting problem will guide our selection distance to the transmitter), while others estimate the com- of the specific sheaf we need and will be the context for pass bearing. Both of these estimates vary in accuracy and its interpretation. The resulting fox hunting sheaf is mod- in precision depending on terrain, environmental condi- ular; different sensors or models of their performance can tions, equipment quality, and the skill of the operator. be substituted easily without changing how their data are Fox hunting also serves the purpose of preparing radio analyzed. amateurs for emergency or disaster operations. Because This article explains how to model a collection of sen- disaster operations require the concerted efforts of mul- sors in the section called “Formalizing the Sensors in the tiple radio operators, it seems fitting to explore how the Fox Hunt” so that they can be combined into a sheaf model sport changes if fox hunting becomes cooperative. When in “Formalizing the Interactions between Sensors.” Once participants combine their estimates of distance and bear- the sheaf is constructed, we show how to locate the fox ing, how much faster can they find the transmitter? Part transmitter in “Consistency Radius: Where is the Fox?” and of the challenge of fox hunting is that measurements are determine if there are actually multiple fox transmitters in taken infrequently, only once every few minutes. To win “Local Consistency Radius: Finding Multiple Foxes.” Fi- the hunt, every minute must count! In the most demand- nally, since applied sheaf theory is still in its infancy, “Fron- ing scenario, each sensor only gets to take one measure- tiers”points the reader to some interesting directions for ment of the fox transmitter. future study. Locating the fox transmitter from a collection of differ- ent sensors is a model-based data fusion problem: combining Formalizing the Sensors in the Fox Hunt disparate local observations into a global inference. With- Each receiver (or sensor) 퐴 used in the fox hunt produces a out a model that describes how signals from the transmit- signal report concerning its observation of the fox transmit- ter arrive at each receiver, the signal reports are not helpful ter. Signal reports may be of different types, depending on for locating the transmitter. Even with such a model, the the sensor. For instance, the strength of the received signal effects of terrain, the transmitter’s antenna system, and the is typically reported as a single real number. In contrast, a environment can cause substantial differences between the compass bearing is reported as an angle, properly an ele- modeled signal and an actual received signal. Therefore, it ment of the metric space 푆1—the unit circle. To handle is important that we remain even-handed about assump- both of these situations (and more), let us suppose that a tions of the quality of the estimates and the quality of the signal report is an element of a pseudometric space 퐷퐴, de- model. pending on the sensor 퐴. The sensor 퐴 produces reports Though there are many techniques for solving data fu- through a continuous measurement function sion problems, they broadly fall into two categories: (1) 2 2 푀퐴 ∶ ℝ × 퐶퐹 × ℝ × 퐶퐴 → 퐷퐴, problem-specific deterministic methods, and (2) general statistical methods. Well-crafted problem-specific determin- depending on fox transmitter location (in the plane), the istic methods are very effective. Because of the physics of fox transmitter equipment settings 퐶퐹 (such as transmitter radio propagation, it is not too difficult to construct a de- power and antenna orientation), the receiver location in terministic method specifically for locating a fox transmit- the plane, and the receiver equipment settings 퐶퐴 (such as ter. However, problem-specific methods often carry hid- antenna orientation). den assumptions that make it hard to transfer useful tech- Our sensor data will be drawn from a parameterized dis- niques to another problem. Worse, the bookkeeping as- tribution 푆퐴, in which the noise level 휎 is taken as a pa- sociated with all combinations of sensors grows exponen- rameter. To ensure consistency between the deterministic tially as more sensors are deployed. Regardless of their at- model and the stochastic one, our stochastic models sat- traction, problem-specific methods for solving data fusion isfy ′ ′ ′ ′ problems are costly and difficult to manage. Statistical 푆퐴(푥, 푦, 푐푓, 푥 , 푦 , 푐퐴; 휎) → 훿푀퐴(푥,푦,푐푓,푥 ,푦 ,푐퐴) as 휎 → 0, methods automate the bookkeeping and tailoring needed in which 훿 is the unit impulse at 훼. for a problem-specific method, but they usually require 훼 There are typically two kinds of sensors that are used in many observations to produce accurate results. Since we radio fox hunting: calibrated signal strength meters and are only using one measurement from each sensor, statis- directional antennas. Given knowledge of the fox trans- tical methods are not the best option. mitter’s power output, a calibrated signal strength meter

662 NOTICES OF THE AMERICAN MATHEMATICAL SOCIETY VOLUME 66, NUMBER 5 Figure 1. Simulated received power for a transmitter located at the top of a mountain using (a) the Longley-Rice model [8] and (b) the simplified model in (1). Colors are on a Figure 2. The author’s son holding a direction-sensitive logarithmic scale, in decibels referenced to the power of 1 antenna (a Yagi-Uda array) built by the author for 144 MHz. milliwatt (dBm). can help estimate the distance from the receiver to the fox. where 퐼0 is the modified Bessel function of the first kind. A directional antenna tells the operator the direction from The Rice distribution models the distance between the ori- which the fox’s signal appears to be the strongest. gin and a point in the plane drawn from a bivariate Gauss- Calibrated signal strength (RSSI) sensors. A received sig- ian with mean 푣 and standard deviation 휎. Assuming the nal strength indication (RSSI) sensor measures the amount received signal is a combination of the inverse square law of power absorbed by its antenna from the fox transmit- and Rician fading, the RSSI signal reports 푃 will follow the ter. To model this accurately requires careful specification distribution of the terrain and any obstacles between the transmitter ′ ′ and receiver. The Longley-Rice model is popular among 푆푅푆푆퐼(푃; 푥, 푦, 푝, 푥 , 푦 ) (2) engineers because it incorporates the effect of terrain and 1 2 = (푅푖푐푒 (푃; 2푀 (푥, 푦, 푝, 푥′, 푦′)푅, 휎)) , atmospheric losses on the received power, and its predic- 2푅 √ 푅푆푆퐼 tions are realistic [8]. Figure 1(a) shows the received signal power predicted by the Longley-Rice model for a fox trans- in which antenna characteristic impedance 푅 (typically 50Ω mitter placed on a mountain. for amateur equipment) and root mean square noise volt- While the realism of the Longley-Rice model is a bene- age 휎 are assumed to be known in advance and constant. fit, its precise specification is quite complicated. Noneof Bearing sensors. A bearing sensor measures the angle be- this complexity is necessary to demonstrate our approach, tween true north (or some other convenient, global direc- because all our analyses are modular. A different measure- tion) and the apparent direction of arrival of signals from ment function can be substituted later if desired without the fox transmitter. Most amateur bearing sensors con- changing the analysis techniques. Therefore, we will sim- sist of an antenna (like the one shown in Figure 2) that ply model the power transferred from transmitter to re- is preferentially sensitive to signals arriving from a specific ceiver by an inverse-square law (Figure 1(b)). This corre- direction. The operator rotates the antenna until the sig- sponds to a measurement function in which there is only nal strength is greatest, and then records its direction. For fox transmitter power 푝 ∈ 퐶퐹 = [0, ∞) and no receiver strong signals, the operator can also block incoming sig- configuration, nals by holding a hand held radio against the chest. When 푝 (1) the operator slowly turns around until the signal strength 푀 (푥, 푦, 푝, 푥′, 푦′) = , 푅푆푆퐼 4휋 ((푥 − 푥′)2 + (푦 − 푦′)2) is minimized, the transmitter is then directly behind the ′ ′ operator! where (푥, 푦) is the transmitter location and (푥 , 푦 ) is the A useful antenna for fox hunting produces a response receiver location. like one of the two shown in Figure 3 (p.664). It is a con- From a statistical perspective, after the inverse square 푆1 → ℝ law, the next largest effect on the received signal is caused tinuous, unimodal function on the circle , since by self-interference as the signal travels along nearby paths such a function has a single maximum (or minimum) in with slightly different lengths. This effect is called Rician the direction of the fox transmitter (or in the exact oppo- fading and is governed by the Rice distribution for signal site direction). A key performance criterion for an antenna amplitude (usually voltage) 푉 ∈ [0, ∞) is its beamwidth, the length of the interval for which the re- ceived signal exceeds half of the maximum value. In the 2 2 푉 −(푉 + 푣 ) 푣푉 presence of noise, an antenna with a small beamwidth will 푅푖푐푒(푉; 푣, 휎) = exp ( ) 퐼0 ( ) 휎2 2휎2 휎2 give more accurate bearing readings.

MAY 2019 NOTICES OF THE AMERICAN MATHEMATICAL SOCIETY 663 Table 1. Independent sensors for the fox hunting examples Sensor ID Location Type Measurement Noise model 1 (1.0, 0.0) Position ℝ2 (None) Bearing 푆1 von Mises 2 (1.0, 1.0) Position ℝ2 (None) Bearing 푆1 von Mises 3 (0.0, 0.0) Position ℝ2 (None) Signal strength ℝ Rician 4 (0.0, 0.5) Position ℝ2 (None) Signal strength ℝ Rician We will model the observations recorded by a bearing Given a fox transmitter located at (푥, 푦) and sensor lo- sensor stochastically by way of the von Mises distribution, cated at (푥′, 푦′), the true bearing (in degrees) can be ob- the probability distribution that is the closest analogue of tained by the measurement function the Gaussian distribution for the circle [9]. The von Mises ′ ′ distribution is given by 푀푏푒푎푟푖푛푔(푥, 푦, 푥 , 푦 ) (3) ′ 휅 cos(휃−휇) 180 −1 푥−푥 ′ ⎧ tan ′ if 푦 − 푦 > 0 푒 1 ⎪ 휋 푦−푦 푉표푛푀푖푠푒푠(휃; 휇, 휅) = 휃 ∈ 푆 ⎪ ′ , for ⎪ 180 −1 푥−푥 ′ ′ 2휋퐼0(휅) ⎪ 휋 tan 푦−푦′ +180 if 푦−푦 <0 and 푥−푥 ≥0 180 −1 푥−푥′ ′ ′ where again 퐼0 is the modified Bessel function of the first = tan ′ −180 if 푦−푦 <0 and 푥−푥 <0 ⎪⎨ 휋 푦−푦 kind. By analogy with the Gaussian distribution, 휇 in the ⎪90 if 푦−푦′ =0 and 푥−푥′ >0 ⎪ von Mises distribution plays the role of the mean, while ⎪ ′ ′ ⎩−90 if 푦−푦 =0 and 푥−푥 <0 1/휅 is analogous to the variance. In order to model the measured bearing 휃 for a fox transmitter located at the true where true north is oriented along the positive 푦-axis and bearing 휇 from an antenna with beamwidth 퐵, we set 휅 = angles are measured clockwise. This simple model ignores 2 4/퐵 . some important effects, such as the fact that the apparent 90 bearing can be distorted by reflections, but it is a good 1 120 60 model for relatively flat terrain. Then the appropriate sto- chastic model is

′ ′ 푆푏푒푎푟푖푛푔(휃; 푥, 푦, 푥 , 푦 ) (4) 150 30 ′ ′ 2 0.5 = 푉표푛푀푖푠푒푠(휃; 푀푏푒푎푟푖푛푔(푥, 푦, 푥 , 푦 ), 4/퐵 )

for an antenna with a known, constant beamwidth 퐵. Coordinating multiple sensors. Consider a team of four radio operators each equipped with a position sensor (a 180 0 GPS receiver) that records their own location and a sen- sor that measures either the received signal strength from or the bearing to the fox transmitter according to Table 1. Since GPS errors are small compared to those reported from amateur radio equipment, we will assume that GPS posi- 210 330 tions are known exactly. In all of the examples, we use the unit-less positions shown in the table rather than the latitude and longitude that would be reported by a GPS 240 300 receiver. Using this configuration of sensors, we will study five 270 Cases (Table 2) that address two questions: (1) “Where Figure 3. Simulated received signal strength as a function of is the fox?” and (2) “Are there multiple foxes?” We will angle for typical antennas used in fox hunting. The solid red start our analysis of the first question using Cases 1 and 2, curve is typical for a directional antenna like the one shown in Figure 2, while the dashed blue curve is typical for a handheld before we demonstrate it with different realizations of sto- receiver held against the body. chastic noise in Case 3. Cases 4 and 5 will be used to ad- dress the second question.

664 NOTICES OF THE AMERICAN MATHEMATICAL SOCIETY VOLUME 66, NUMBER 5 Table 2. Fox hunting cases Table 3. Observations used in the fox hunting examples Case Sensors Number Fox location Noise Observation Virtual Description of foxes known present 퐹′ Yes Fox position, transmit power 1 All 1 Yes No 퐹 Yes Fox position ′ 2 Sensors 1 & 2 1 Yes No 퐵1 Yes Fox position, Sensor 1 position (bearings only) 퐵1 No Sensor 1 position and bearing ′ 3 All 1 No Yes 퐵2 Yes Fox position, Sensor 2 position 4 Sensors 1 & 2 2 No No 퐵2 No Sensor 2 position and bearing ′ (bearings only) 푅3 Yes Fox position, Sensor 3 position 5 All 2 No No 푅3 No Sensor 3 position and RSSI ′ 푅4 Yes Fox position, Sensor 4 position Formalizing the Interactions between Sensors 푅4 No Sensor 4 position and RSSI Neither of the measurement functions (푀푅푆푆퐼 in (1) and 푀푏푒푎푟푖푛푔 in (3)) alone are sufficient to determine the loca- tion of the fox transmitter. It is only when multiple mea- This is a good strategy because there may be a statistical de- surements from different locations or when different sen- pendence between the position errors and the signal mea- sors’ measurements are taken together that the location of surements. There are three observations that can be made: the fox may be determined. This joint interaction between (1) the joint fox transmitter location and sensor location, sensors can be encoded as a sheaf of pseudometric spaces on (2) the fox transmitter’s location, and (3) the joint bearing a partial order, by taking account of deterministic, func- sensor’s location and reported bearing. The first of these tional dependencies between sensors’ observations. Ob- completely and functionally determines the other two, a servations are of two types: (1) true observations: those ac- situation that can be expressed as a partial order on the tually made by sensors, and (2) virtual observations: those observations, which is shown in Figure 5(a) (p. 666). The that could have been observed, but were not actually re- joint fox transmitter location and sensor location (the min- ported. The location and the transmitter power of the fox imal element in the partial order) is a virtual observation transmitter are in the latter class: the operator who hides because we cannot make that joint observation without ob- the fox transmitter is legally required to observe its loca- serving both simultaneously. The arrows in the diagram tion and output power, but keeps these observations secret point from smaller to larger elements in the partial order, during the fox hunt! and express functional dependence. The spaces of obser- It is reasonable to record position and a radio measure- vations and the actual functions themselves are shown in ment simultaneously, so the typical measurement reported Figure 5(b), where pr푘(푥1, 푥2, … ) = 푥푘 is the projection by Sensor 1 (a bearing sensor) will be of type ℝ2 × 푆1. from a product onto its 푘-th factor. In much the same way, the relationship for an RSSI sensor is between a vir- tual observation (the fox transmitter location, power level, and sensor location, jointly) and two true observations is shown in Figure 6 (p. 666). Measurement functions play the role of transforming virtual observations into true ob- servations; if a different measurement function is desired, it can be easily substituted without disrupting the partial order. Combining all four sensors under the hypothesis that all of the fox transmitters are actually the same yields a somewhat larger collection of observations, shown in Ta- ble 3 along with their interpretations. They are named so that 퐵 is for bearing sensors, 푅 is for RSSI sensors, and 퐹 is for the fox. These observations form a partial order, shown in Fig- ure 7 (p. 666), obtained by “gluing together” several di- agrams like those shown in Figures 5(a) and 6(a) along the common observations about the fox transmitter. The Figure 4. Spatial layout of the fox hunting examples order relation ≤ is the transitive closure of the relation in- ′ ′ duced by the graph, so 푅3 ≤ 퐹 ≤ 퐹 for instance. This

MAY 2019 NOTICES OF THE AMERICAN MATHEMATICAL SOCIETY 665 Figure 5. Measurement of a bearing to the fox: (a) partial order of the dependencies between observations, (b) the spaces of ′ observations and the measurement functions determining them. (This is a subgraph corresponding to both {퐹, 퐵1, 퐵1} and ′ {퐹, 퐵2, 퐵2} in Figure 7.)

Figure 6. Measurement of an RSSI report of the fox: (a) partial order of the dependencies between observations, (b) the spaces ′ ′ of observations and the measurement functions determining them. (This is a subgraph corresponding to both {퐹 , 푅3, 푅3} and ′ ′ {퐹 , 푅4, 푅4} in Figure 7.)

′ implies that the observation 푅3 functionally determines the observation 퐹. Figure 8 shows each observation’s space of values and specifies each of the measurement functions, which makes all of the functional dependencies between observations explicit. Figure 8 defines a bit more than a partial order; it determines a sheaf on the partial order.

Figure 8. The sheaf diagram for all observations shown in Figure 7.

Definition 1. [1] A sheaf 풮 on a partial order (푋, ≤) consists of the specification of: 1. A set 풮(푥) for each 푥 ∈ 푋, called the on 푥, and 2. A function 풮(푥 ≤ 푦) ∶ 풮(푥) → 풮(푦), called the restriction along 푥 ≤ 푦, for each 푥 ≤ 푦 ∈ 푋, such that 3. Whenever 푥 ≤ 푦 ≤ 푧 ∈ 푋, it follows that 풮(푥 ≤ 푧) = 풮(푦 ≤ 푧) ∘ 풮(푥 ≤ 푦). Notice that we need not specify the restriction along ′ Figure 7. Partial order of all observations shown in Table 3 푅3 ≤ 퐹 in Figure 7, for instance, since this is already completely determined by the composition of restrictions

666 NOTICES OF THE AMERICAN MATHEMATICAL SOCIETY VOLUME 66, NUMBER 5 Table 4. Case 1: Known fox at location (0.5, 0.5) Observation Description Value 퐹′ Fox position, transmit power (0.5, 0.5), 1.0 퐹 Fox position (0.5, 0.5) ′ 퐵1 Fox position, Sensor 1 position (0.5, 0.5), (1.0, 0.0) 퐵1 Sensor 1 position and bearing (1.0, 0.0), −45 ′ 퐵2 Fox position, Sensor 2 position (0.5, 0.5), (1.0, 1.0) 퐵2 Sensor 2 position and bearing (1.0, 1.0), −135 ′ 푅3 Fox position and power, Sensor 3 position (0.5, 0.5), 1.0, (0.0, 0.0) 푅3 Sensor 3 position and RSSI (0.0, 0.0), 0.16 ′ 푅4 Fox position and power, Sensor 4 position (0.5, 0.5), 1.0, (0.5, 1.0) 푅4 Sensor 4 position and RSSI (0.5, 1.0), 0.32 ′ ′ ′ ′ along 푅3 ≤ 퐹 and 퐹 ≤ 퐹. If we have obtained the fox 퐹 of ((0.5, 0.5), 1.0). Table 4 shows what all of the obser- ′ position and Sensor 3 position (observation 푅3), then we vations (both true and virtual) would be for this situation automatically know the fox position (observation 퐹). (positions of the sensors are from Figure 4). These obser- An important feature of a sheaf on a partial order is that vations form a global section 푔 of the sheaf 풮 shown in the upward set 푈푥 for an element 푥 ∈ 푋, given by Figure 8. Although it is tedious to verify that 푔 really is a global section, it is enlightening to check a few restrictions. 푈푥 = {푦 ∈ 푋 ∶ 푥 ≤ 푦} The virtual observation for the bearing Sensor 1 (observa- is the set of observations 푦 ∈ 푈푥 that are functionally de- ′ tion 퐵1) is consistent with the fox location (observation 퐹) termined by the observation 푥. because The collection of upward sets forms the for a topol- ′ ′ ogy, called the Alexandrov topology 퐀퐥퐞퐱(푋, ≤). This sug- (풮 (푈퐵1 ⊂ 푈퐹)) 푔(퐵1) gests that we ought to assign measured quantities to open = (풮 (푈 ′ ⊂ 푈퐹)) ((0.5, 0.5), (1.0, 0.0)) sets of observations, not just individual observations. 퐵1 = pr ((0.5, 0.5), (1.0, 0.0)) Remark 1. The reader who is familiar with the usual def- 1 inition of a sheaf on a may notice that our = (0.5, 0.5) = 푔(퐹). definition of a sheaf 풮 on a partial order (푋, ≤) is merely ′ The observation at 퐵 is also consistent with the bearing a from the category generated by the partial order, 1 Sensor’s true observation 퐵1 because and that the gluing axiom is apparently missing. The gluing ′ axiom ensures that the value of 풮 on a typical 푈 ′ (풮 (푈퐵1 ⊂ 푈퐵1 )) 푔(퐵1) is the space of sections for any open cover of 푈. Since we = (풮 (푈 ′ ⊂ 푈 )) ((0.5, 0.5), (1.0, 0.0)) have only defined 풮 on the upward sets, not all open sets, 퐵1 퐵1 we may use the gluing axiom to define sections and the rest = (pr2 ((0.5, 0.5), (1.0, 0.0)) , of the sheaf accordingly. 푀푏푒푎푟푖푛푔(0.5, 0.5, 1.0, 0.0)) Definition 2. Suppose that 푈 is an open set in 퐀퐥퐞퐱(푋, ≤). = ((1.0, 0.0), −45) = 푔(퐵 ). The set of sections on 푈 of a sheaf 풮 on (푋, ≤) is 1 The sheaf structure imposes constraints on sections over 풮(푈) = {푠 ∈ 풮(푥) ∶ 푠(푦) = 풮(푥 ≤ 푦) (푠(푥)) , ′ ∏ unions of overlapping upward sets. Observations 퐵1 and 푥∈푈 ′ 퐵2 both stipulate a position for the fox transmitter in their for all 푥 ∈ 푈 and 푦 ∈ 푈푥}, first factor, but 퐹 also stipulates a position for the fox trans- ′ ′ namely the set of values from each stalk in 푈 that are con- mitter. The model posited by a section over 푈퐵1 ∪ 푈퐵2 sistent with all restrictions. A section on 푋 is called a global argues that there is only one transmitter, so each of these section. positions ought to be the same. ′ ′ 4 So while 풮(퐵1) = 풮(푈퐵1 ) = ℝ represents the fox po- It is immediate that 풮(푥) is in one-to-one correspon- ′ 4 sition and Sensor 1 position, and 풮(퐵2) = 풮(푈퐵′ ) = ℝ dence with 풮(푈푥). 2 represents the fox position and Sensor 2 position, their in- Case 1: All sensors, known fox, no noise. Suppose that 2 tersection 풮(퐹) = 풮(푈퐹) = ℝ represents the fox posi- the four sensors receive (with no noise) reports from a tion only. A section on the union fox located at (0.5, 0.5) transmitting with power level 1.0. ′ ′ ′ ′ This corresponds to the situation of a virtual observation 푈퐵1 ∪ 푈퐵2 = {퐵1, 퐵1, 퐵2, 퐵2, 퐹}

MAY 2019 NOTICES OF THE AMERICAN MATHEMATICAL SOCIETY 667 asserts that the fox positions between the two sensors are Implicit in both the assignment pseudometric and consis- ′ ′ 6 held in common, so that 풮(푈퐵1 ∪ 푈퐵2 ) = ℝ by the inclu- tency radius is a choice of norm that aggregates multiple sion-exclusion principle. pseudometrics. Since different norms respond differently Moreover, any global section is determined by its value to outliers, what guides the selection of that norm? on each of the minimal elements of the partial order in Fig- The central relationship between global sections of 풮 ure 7. For this particular sheaf, specifying the fox position, and assignments is captured by the following bound, which fox transmitter power, and the positions of each of the four interprets consistency radius as an obstruction to an assign- sensors completely determines a global section. Thus the ment being a section. set of global sections is ℝ11 = ℝ2 × ℝ × (ℝ2)4. Proposition 1. [13, Prop. 23] For an assignment 푎 to a sheaf Consistency Radius: Where is the Fox? 풮 of pseudometric spaces on (푋, 풯) in which each restriction Sections on large open sets are rarely observed due to a va- map of 풮 is Lipschitz with constant 퐾, then for every global riety of uncertainties about the model and the presence section 푠 of 풮(푋), of stochastic noise. One cannot detect these uncertain- 푐 (푎, 풯) ≤ (1 + 퐾)퐷(푎, 푠). ties from a single observation; they are only apparent after 풮 comparing with other observations. This leads to the no- Strictly speaking, none of the restriction maps in this ar- tion of an assignment, where the consistency described by ticle are Lipschitz since their derivatives become unbound- restriction functions is not enforced. ed when the fox and receiver approach each other. It is Definition 3. [13]. For a sheaf 풮 on a partial order (푋, ≤) practical, however, to assume that the fox and the receivers with the Alexandrov topology 풯 = 퐀퐥퐞퐱(푋, ≤), and a are always separated by some minimum distance. This 퐾 collection of open sets 풰 ⊆ 풯, an assignment supported on minimum distance establishes the Lipschitz constant . 풰 is an element of ∏푈∈풰 풮(푈). Case 2: Bearings only, unknown fox, no noise. To ex- plore how consistency radius can help locate a fox trans- If the space of sections 풮(푈) is a pseudometric space mitter, consider Sensor 1 and 2 in Figure 9(a) (p. 669), with pseudometric 푑푈 for each open 푈 ∈ 풰, the set of as- both of which are bearing sensors. Sensor 1 reports a bear- signments supported on 풰 has an assignment pseudometric ∘ ing of −45 (corresponding to 퐵1 in Figure 9(b)) and Sen- for two assignments 푎, 푏 in ∏ 풮(푈) given by ∘ 푈∈풰 sor 2 reports −135 (corresponding to 퐵2 in Figure 9(b)). 2 Merely by intersecting sight lines from these sensors, we 퐷(푎, 푏) = ∑ 푑푈 (푎(푈), 푏(푈)) . can infer that the fox is located at (0.5, 0.5). The sheaf √푈∈풰 of observations also recovers this information and a little Suppose that 푉 ⊆ 푊 are two open sets in 풯 = more besides. The two sensors correspond to five obser- 퐀퐥퐞퐱(푋, ≤). This means that ∏푦∈푉 풮(푦) ⊆ ∏푥∈푊 풮(푥), vations shown in Figure 9(b), and this collection of obser- so that there is a natural projection 풮(푊) → 풮(푉) that vations forms the sheaf shown in Figure 9(c) when asso- merely discards the stalks that are in 푊 but not in 푉. In the ciated to their measurement functions. As before, the fox case of upward sets 푈푦 ⊆ 푈푥, this projection is precisely ′ ′ position 퐹 and both joint observations 퐵1, 퐵2 are virtual the restriction 풮(푥 ≤ 푦). Without complicating our no- observations, because we do not know their values. tation, we may unambiguously refer to the natural projec- We can encode the bearing reports as an assignment 푎 tion of sections 풮(푊) → 풮(푉) as a restriction 풮(푉 ⊆ 푊). shown in Figure 10(a) (p. 669), where we note that the fox Definition 4. For an assignment supported on all open sets, location can vary the consistency radius 푎(푈퐹) = (푥, 푦), 2 푐풮(푎, 풯)= ∑ ∑ 푑푈 ((풮(푈⊆푉)) 푎(푉), 푎(푈)) because we do not yet know its location. The true observa- √푉∈풯, 푈⊆푉∈풯 tions from the two sensors are given by quantifies how far a given assignment is from being aglobal 푎(푈퐵1 ) = ((1.0, 0.0), −45), section. 푎(푈퐵2 ) = ((1.0, 1.0), −135), Effective usage of sheaves in practice can hinge on care- since we know both the sensor locations and the bearings. ful weighting among the pseudometrics in Definition 4, ′ ′ The two virtual observations at 퐵1 and 퐵2 are typically over but equal weights among the different pseudometrics in constrained by these three facts because most points (푥, 푦) their natural units works well enough in our fox hunting in the plane do not satisfy the simultaneous system examples. 푀푏푒푎푟푖푛푔(푥, 푦, 1.0, 0.0)=−45, Open question 1. How should one choose the weights for 푀 (푥, 푦, 1.0, 1.0)=−135. the assignment pseudometric and the consistency radius? 푏푒푎푟푖푛푔

668 NOTICES OF THE AMERICAN MATHEMATICAL SOCIETY VOLUME 66, NUMBER 5 Figure 9. The setup for Case 2: (a) the spatial layout of the two bearing sensors and the fox whose location is unknown, (b) the partial order of the observations, and (c) the sheaf diagram.

Figure 10. The results of Case 2: (a) an assignment to the sheaf in Figure 9(c) with Sensor 1 receiving a bearing of −45∘ and Sensor 2 receiving a bearing of −135∘, (b) consistency radius as a function of fox transmitter position. Notice the prominent minimum where the bearing sight lines (marked with arrows) coincide.

Therefore, the consistency radius of the assignment 푎 is values at the true observations. For instance, if we did not ′ ′ typically not zero. Figure 10(b) shows the consistency ra- specify values at 퐵1, 퐵2, and 퐹 in Figure 9(b), we could dius of 푎 as a function of the fox position (푥, 푦). Two con- infer them from the values at 퐵1 and 퐵2. This warrants a clusions can be drawn immediately from this plot: (1) the more general definition of consistency radius. point of intersection (0.5, 0.5) between the two sight lines 푎 minimizes the consistency radius (zero), (2) no other fox Definition 5. The consistency radius of an assignment sup- 풰 transmitter position will yield an assignment with consis- ported on —rather than all open sets—is the infimum 푏 푎 tency radius of zero. This strongly points to the conclusion of all consistency radii of assignments that restrict to , that the fox transmitter is located at (0.5, 0.5). namely

Optimally extending assignments. Case 2 above indicates 푐풮(푎, 풰) that minimizing the consistency radius is a useful infer- ⎧ ence procedure. Instead of encoding the signal reports = inf 푐 (푏, 풯) ∶ 푏 ∈ 풮(푉) ⎨ 풮 ∏ from the sensors as an assignment with values at both true ⎩ 푉∈풯 and virtual observations, we really only need to consider such that 푏(푈) = 푎(푈) whenever 푈 ∈ 풰} .

MAY 2019 NOTICES OF THE AMERICAN MATHEMATICAL SOCIETY 669 We say that each such assignment 푏 extends 푎. The fox location appears in several different places in ′ an assignment supported on all observations, namely 퐵1, A nonzero consistency radius for an assignment that is ′ ′ ′ ′ 퐵2, 푅3, 푅4, 퐹 , and 퐹. Since supported on 풰, rather than on all open sets, is still the ob- 푈 = {퐹} = 푈 ′ ∩ 푈 ′ ∩ 푈 ′ ∩ 푈 ′ ∩ 푈 ′ , struction to extending that assignment to a global section, 퐹 퐵1 퐵2 푅3 푅4 퐹 as the next Proposition states. we deem that the predicted fox location is specified by the (virtual) observation at 퐹. Proposition 2. [14, Prop. 11] If 푎 is an assignment to a Even though a given extension of an assignment may sheaf 풮 on a topological space (푋, 풯) supported on 풯 (every minimize consistency radius, it may not predict the cor- open set) and 푈 ∈ 풯, then rect location of the fox. Figure 13 (p. 672) shows the er- rors for the inferred location of the fox transmitter. The 푐풮(푎, 풯 ∩ 푈) ≥ 푑푉1 (풮(푉1 ⊆ 푉2)푎(푉2), 푎(푉1)) figure shows results from two ways to infer virtual observa- for every 푉1 ⊆ 푉2 ⊆ 푈. tions when presented with a set of true observations as an assignment 푎 supported on 풰: (1) extension: find the as- Case 3: All sensors, unknown fox, noise present. Con- signment extending 푎 to all open sets that minimizes con- sider the situation of trying to infer the location of a single sistency radius (green triangles), as given by Definition 5 unknown fox transmitter using all four sensors, in which or (2) fusion: find the nearest global section 푠 to 푎 in the as- the reports are contaminated with noise. Several simula- signment pseudometric (red dots) [13]. In both cases, the tions were run with different amounts of noise and differ- overall transmitter location error is largely independent of ent random draws to produce signal reports for each sen- noise level, with angle errors being more detrimental. The sor. Each simulation produced four values, (휃 , 휃 , 푟 , 푟 ) 1 2 3 4 green triangles in Figure 13 correspond to the same simu- for the bearing from sensor 1, the bearing from sensor 2, lations shown in Figure 12, but the vertical scales are dif- RSSI from sensor 3, and RSSI from sensor 4, respectively. ferent: consistency radius is not the same as location error. These four sensors’ reports correspond to the open set {퐵1, 퐵2, 푅3, 푅4} in the partial order of the observations Open question 2. Solving the extension problem gener- shown in Figure 11(a) (p. 671). The simulated reports ally results in substantially better estimates of the fox trans- were encoded as an assignment to the sheaf shown in Fig- mitter location than solving the fusion problem, which is ure 11(b) supported on the open set {퐵1, 퐵2, 푅3, 푅4}. No- why we discuss it here. While this intriguing phenome- tice that the true observations—within the support of the non has been observed in other settings [15], what are the assignment—are shown by an element of the stalk specify- precise conditions under which extension outperforms fu- ing the signal report, while the figure shows the entire stalk sion? at each virtual observation. Each simulation therefore corresponds to an assignment, Local Consistency Radius: Finding Multiple Foxes for which we would like to compute consistency radius The traditional amateur radio fox hunt involves only one using Definition 5. This requires finding an extension of fox transmitter. If there are actually multiple fox transmit- each assignment with minimal consistency radius. This ters and the sensors are in general position, the minimum extension was found using the “SLSQP” method of the consistency radius of an assignment constructed from the popular SciPy optimizer scipy.optimize.minimize, signal reports will not be zero. Sheaves allow us to segment which can solve general optimization problems. The de- the observations into those of different foxes. This deduc- fault options for stopping conditions were sufficient in all tion can be fully justified if we can identify open sets of our simulations, but more sophisticated optimizers or care- sensors whose reports are consistent, even when the collec- ful selection of parameters is sometimes required for other tion of all sensors is not. The extent to which consistency sheaves [15]. is obtained on an open set is formalized by the following Figure 12 (p. 671) shows how consistency radius de- definition. pends on stochastic noise present in simulated signal re- Definition 6. Let 푎 be an assignment to every open set in ports. In Figure 12(a), no noise is present in the reports a sheaf 풮 of pseudometric spaces on (푋, 풯). For an open from Sensor 3 and 4 (RSSI reports 푟 and 푟 ) but is present 3 4 푈 ∈ 풯, the local consistency radius on 푈 is in the reports from Sensors 1 and 2 (bearing reports 휃1 and 휃2). The consistency radius generally increases steadily as 푐풮(푎, 푈) the amount of noise in the bearing reports increases. Fig- ∶= 푑 ((풮(푉 ⊆푉 ))푎(푉 ), 푎(푉 ))2. ure 12(b) shows the same situation but the bearings are ∑ ∑ 푉1 1 2 2 1 푉 ⊆푈∈풰, 푉 ⊆푉 ∈풰 noiseless, while noise is applied to the RSSI reports. The √ 2 1 2 increase in consistency radius is somewhat more abrupt, Using this, the 휖-consistent collection consists of every con- but stabilizes for larger noise values. nected open set 푈 ∈ 풯, such that (1) 푐풮(푎, 푈) < 휖 and

670 NOTICES OF THE AMERICAN MATHEMATICAL SOCIETY VOLUME 66, NUMBER 5 Figure 11. Case 3: Inferring fox location from reports from all sensors. (a) The partial order of observations (true observations are shaded), and (b) the resulting assignment supported on the true observations. Virtual assignments are marked with the stalk from the sheaf in Figure 8 for easy reference.

Figure 12. Consistency radius as a function of input noise level for Case 3. Each point is a separate simulation consisting of a report from each Sensor. (a) Noise is present only in Sensors 1 and 2 (bearing) according to (4). (b) Noise is present only in Sensors 3 and 4 (RSSI) according to (2).

(2) there is no other connected open set 푉 ∈ 풯 with space, then 푐풮(푎, 푉) < 휖 and 푈 ⊂ 푉. 푐풮(푎, 푈) ≤ 푐풮(푎, 푉). Usually, 휖-consistent collections only cover part of the The Lemma follows from the fact that the sum in the space 푋. All observations 푥 ∈ 푋 that are maximal in the expression for 푐풮(푎, 푉) is over a strictly larger set than the partial order of observations will necessarily have 푐풮(푎, 푈푥) sum in the expression for 푐풮(푎, 푈). = 0 for any assignment, so their upward sets are always in- A collection of open sets 풱 is said to refine another col- cluded in any 휖-consistent collection for 휖 > 0. Connect- lection of open sets 풰 if every 푉 ∈ 풱 is contained in edness in Definition 6 is for ease of interpretation: two some 푈 ∈ 풰, which can be thought of—not uniquely—as observations lying in different connected components of a function 풱 → 풰. Therefore, if 휏 < 휖, the 휏-consistent an open set 푈 are never tested for consistency when com- collection for an assignment refines its 휖-consistent collec- puting the local consistency radius on 푈. tion. Lemma 1. Local consistency radius is a monotonic function of Definition 7. For an assignment 푎 to every open set in a the open set. Specifically, if 푎 is an assignment to a sheaf 풮 of sheaf 풮 of pseudometric spaces on (푋, 풯), the consistency pseudometric spaces and 푈 ⊆ 푉 are open of the base filtration 퐂퐅(풮,푎) is the set of all 휖-consistent collections for

MAY 2019 NOTICES OF THE AMERICAN MATHEMATICAL SOCIETY 671 Figure 13. Fox transmitter location error after applying sheaf-based processing described in the text to Case 3. Each point is a separate simulation consisting of a report from each Sensor. Green triangles correspond to the estimates obtained by solving the extension problem, while the red dots correspond to solving the fusion problem. (a) Noise is present only in Sensors 1 and 2 (bearing) according to (4). (b) Noise is present only in Sensors 3 and 4 (RSSI) according to (2).

Figure 14. Case 4: Two bearing sensors receiving conflicting reports. (a) The partial order of observations, (b) the sheaf diagram, (c) the assignment representing the bearing reports shown by arrows in (d), and (d) consistency radius of the assignment as a function of fox transmitter position. The star in (d) marks the fox transmitter for the purpose of computing the consistency filtration in Figure 15 (p. 673). all 휖 ∈ [0, ∞). We will use the notation 퐂퐅(풮,푎)(휖) to setup is thus far the same as in Case 2, but instead sup- refer to a particular value of 휖. pose that Sensor 1 reports a bearing of −180∘ and Sensor 2 reports a bearing of 0∘. This can happen if there are two

Because of Lemma 1, the consistency filtration 퐂퐅(풮,푎) foxes, but cannot happen if there is only one fox. As be- is actually a sheaf on the partial order (ℝ, ≤), in which fore, we can encode the signal reports as an assignment 푎 the restrictions 퐂퐅(풮,푎)(휏 < 휖) are refinement functions. shown in Figure 14(c), with As 휖 decreases, the open sets in 퐂퐅 (휖) become more (풮,푎) 푎(푈 ) = ((1, 0), −180), 푎(푈 ) = ((1, 1), 0). refined. 퐵1 퐵2 Case 4: Bearings only, two unknown foxes, no noise. Figure 14(d) shows the consistency radius of the assign- Let us consider the setting in which there are two bear- ment as a function of the fox location. The infimum consis- ing sensors, shown in Figure 14. The sheaf shown in Fig- tency radius is strictly positive and is never attained, both ure 14(b) assumes that there is one fox transmitter. The of which indicate that there is a problem with the sheaf

672 NOTICES OF THE AMERICAN MATHEMATICAL SOCIETY VOLUME 66, NUMBER 5 Both the maximality and connectedness requirements are ′ important: {퐹, 퐵1, 퐵1, 퐵2} is consistent to a threshold of ∘ 100 but is not connected, and while {퐵1} is sufficiently consistent and is connected, it is contained in a larger such set. In Figure 15, {퐵2} is farther from being absorbed into a larger consistent connected open set than any of {퐹}, ′ {퐵1}, or {퐹, 퐵1, 퐵1}, which suggests that its observation is incompatible with the others. The interpretation is that Sensor 2’s report is harder to reconcile with the proposed fox transmitter location than Sensor 1’s report, because it is receiving a fox transmitter located somewhere else. The consistency filtration is especially useful because it is robust to noise and other variations. If we weight the edges in the consistency filtration graph shown in Figure 15 with the difference in local consistency radius between Figure 15. The consistency filtration of Case 4. Recall that their endpoints, we can interpret the differences between Definition 6 requires only connected and maximal open sets to be shown. length of edges in the consistency filtration graph as the amount of noise (or other uncertainty) that we can accept and yet still detect the inconsistency between the signal model. The problem is that there is no clearly “correct” reports. For instance, in order to reconcile 퐵1 with 퐵2 ∘ location to be inferred for the single fox transmitter. we must accept 135 of total error: the length from {퐵2} ′ ′ Without computing anything, observe that the local con- to {퐵1, 퐵1, 퐵2, 퐵2, 퐹}. The topology of the graph is un- ∘ sistency radius of 푎 on the open set {퐵1, 퐵2, 퐹} is always changed when there is less than 45 of total error: the ′ ′ ′ zero, because no comparisons between these observations length from {퐵1, 퐵1, 퐹} to {퐵1, 퐵1, 퐵2, 퐵2, 퐹}. These are are required. Each of the observations 퐵1, 퐵2, and 퐹 are guaranteed by the following theoretical result. perfectly self-consistent if one does not check for consis- Theorem 1. [14, Thm. 3] For a sheaf of pseudometric spaces tency between them! By choosing the fox location cor- on a finite partial order, the consistency filtration is a continuous rectly at 퐹, one can easily ensure that 푐(푎, 푈 ′ ) = 0 or 퐵1 function of the assignment when the space of consistency filtra- 푐(푎, 푈 ′ ) = 0, but one cannot make both zero at the 퐵2 tions is given the generalized interleaving distance [4, Def. same time. Given this information, Lemma 1 ensures that 2.13] and the space of assignments is given the product topol- the global consistency radius, 푐(푎, 푈 ′ ∪ 푈 ′ ) must be 퐵1 퐵2 ogy. nonzero. Definition 7 of the consistency filtration requires theas- The definition of generalized interleaving distance is signment to be supported on every open set—since every rather technical, but the idea of the Theorem is straightfor- open set is being tested for local consistency!—so let us ward. If we consider two different assignments 푎 and 푏 to suppose that the transmitter is located at the origin (marked a sheaf 풮 such that their consistency filtrations 퐂퐅(풮,푎)(휖) with a star in Figure 14(d)). This can be encoded as the and 퐂퐅(풮,푏)(휙(휖)) are identical under an order preserv- assignment shown in Figure 14(c) with (푥, 푦) = (0, 0). ing bijection 휙, then the local consistency radius of any Figure 15 shows the consistency filtration of this assign- open set cannot change more than a constant factor times ment rendered as a directed graph, in which the vertices are the distance between the two assignments. The size of 휙 labeled with open sets and their local consistency radius. accounts for the difference in local consistency radii of the The edges of this graph denote relations between consistent collections, and also constrains the interleaving open sets. distance after some calculation. 휖 = 0 The consistency filtration begins at with the three Case 5: All sensors, two unknown foxes, no noise. As a {퐵 } {퐵 } {퐹} connected components 1 , 2 , and , since each of final case, consider the situation in Table 5 where allfour these are the largest connected open sets with consistency sensors are receiving noiseless signal reports, but where 퐵 퐹 radius zero. 1 and become consistent above a threshold Sensor 3 (an RSSI sensor) receives signals from a fox lo- 90∘ of as both of these elements are subsumed into the cated at (0.5, 0.5) while the other Sensors receive signals {퐵 , 퐵′ , 퐹} 휖 = 100 open set 1 1 . The -consistent collection, from a fox located at (0.5, 1.0). Both foxes use a trans- 100∘ in which bearings are required to be closer than , is mit power of 1.0. If Sensor 3 had received the same fox as the others, it would report an RSSI observation of 0.063. ′ 퐂퐅(풮,푎)(100) = {{퐵1, 퐵1, 퐹}, {퐵2}}. Let us use the same sheaf model as in Case 1 and Case 3

MAY 2019 NOTICES OF THE AMERICAN MATHEMATICAL SOCIETY 673 Figure 16. Case 5: all sensors and two foxes (a) The data encoded as an assignment, constructed by extending an assignment on the shaded open set to the others by way of minimizing the overall consistency radius. (b) The resulting consistency filtration of this assignment.

(the reader may refer back to Figures 8 or 11, but this is not of 0.063 instead of 0.16. These two values have a differ- necessary), which assumes that there is a single fox trans- ence of 0.097, which is greater than the consistency ra- mitter. We therefore do not expect the assignment that dius. The discrepancy between the consistency radius and encodes these reports to be a global section. this difference is an effect of the unknown fox transmitter The reports from Table 5 are encoded as an assignment power. Even though both of the fox transmitters used a supported on the open set {퐵1, 퐵2, 푅3, 푅4} shown shaded power of 1.0, the consistency radius is minimized with a in Figure 16(a). transmit power of 1.3 (the last component of 퐹′ in Figure To obtain the assignment to all open sets required for 16(a)), which “splits the difference” between the signal re- the consistency filtration (Definition 7), this assignment ports from Sensor 3 and Sensor 4. was extended (again using “SLSQP” in scipy.optimize Figure 16(b) shows the consistency filtration for the as- .minimize) to the one that minimizes overall consistency signment, represented as a directed graph as was done pre- radius. The resulting estimates of virtual observations are viously in Case 4. Most of the open sets are not labeled shown in Figure 16(a). (Note that all open sets are not since they can be inferred by the maximality and connect- shown as this would clutter the figure considerably.) edness conditions. The portion of the diagram correspond- ′ ′ Although there are two foxes present, the assignment ing to Sensors 1, 2, and 4 is the open set {퐵1, 퐵1, 퐵2, 퐵2, 푅4, ′ predicts the location of one fox transmitter at (0.51, 1.0) 푅4, 퐹}, which has a consistency radius of about 0.05. It (the observation at 퐹), which is close to the true value of can be seen that 퐵1 became consistent with the predicted (0.5, 1.0) observed by Sensors 1, 2, and 4. The consistency fox location at a higher local consistency radius than 퐵2 radius of this minimizer is 0.085—evidently not zero— did, because of a difference in predicted fox locations at the ′ ′ which is a result of the conflict between Sensor 3 and the corresponding virtual observations, 퐵1 and 퐵2. Although ′ ′ others. This is not the whole story, though. If Sensor 3 re- the predicted fox location for 푅4 is the same as 퐵2, the dif- ′ ceived the same fox as the others, it would report an RSSI ference in predicted power level at 푅4 (1.4) from the true power level (1.0) results in a local consistency radius for ′ 푈푅4 that is higher than expected from location alone. It is also clear that one particular open set {푅3} is far- thest from the rest in Figure 16(b). At a local consistency Table 5. Case 5 signal reports radius of about 0.08, it becomes consistent with the bear- Sensor ID Type Reported observation ing sensors, but not with the other RSSI sensor. Notice in 1 Position (1.0, 0.0) particular that its predicted location of the fox transmitter Bearing −26.5∘ (0.51, 0.99) is not too far from the others, but that this lo- 2 Position (1.0, 1.0) cation does not correspond to the fox that it is receiving at Bearing −90∘ (0.5, 0.5). The other sensors become consistent with Sen- 3 Position (0.0, 0.0) sor 3 at the global consistency radius of 0.085. The large ′ RSSI 0.16 difference in consistency between open sets containing 푅3 4 Position (0.0, 0.5) and the others indicates that Sensor 3 is the cause of the RSSI 0.16 comparatively high global consistency radius.

674 NOTICES OF THE AMERICAN MATHEMATICAL SOCIETY VOLUME 66, NUMBER 5 Theorem 1 is useful because it asserts that the consis- References tency filtration changes continuously as the assignment [1] Bacławski K. Whitney numbers of geometric lattices. Adv. changes, which itself varies continuously with the signal in Math., 16:125–138, 1975. MR0387086 reports. The next largest open set to contain 푅3 has a con- [2] Ghrist R, Hiraoka Y. Applications of sheaf cohomology sistency radius of 0.08. All of the other true observations and exact sequences to network coding. preprint, 2011. [3] Goguen JA. Sheaf semantics for concurrent interact- 퐵1, 퐵2, and 푅4 are all contained in an open set with con- ing objects. Mathematical Structures in Computer Science, sistency radius 0.05. This means that Sensor 3’s report can 2(02):159–191, 1992. MR1171509 be distorted by 0.03 before the topology of the consistency [4] Harker S, Kramar M, Levanger R, Mischaikow K. A com- filtration graph changes. Tolerance to noise is typically ex- parison framework for interleaved persistence modules pressed as a signal-to-noise ratio, with the typical fox hunt- arxiv:1801.06725, 2018. ing system providing a signal-to-noise ratio in excess of 10. [5] Kokar M, Tomasik J, Weyman J. Formalizing classes of in- Sensor 3’s report will be deemed inconsistent with the oth- formation fusion systems. Information Fusion, 5:189–202, ers provided the system maintains a signal-to-noise ratio of 2004. at least 0.16/0.03 = 5.3, which will definitely be met by [6] Kokar MM, Baclawski K, Gao H. Category theory-based a typical fox hunting system. At such a signal-to-noise ra- synthesis of a higher-level fusion algorithm: An example. In Information Fusion, 2006 9th International Conference on, tio, the value of 퐵 will remain closest to the predicted fox 2 pages 1–8. IEEE, 2006. 퐹 location since the noise level would be not more than [7] Lilius J. Sheaf semantics for Petri nets. Technical report, 0.02/5.3 = 0.004, which is smaller than the difference Helsinki University of Technology, Digital Systems Labo- in local consistency radius for any of the other open sets. ratory, 1993. [8] Longley AG, Rice PL. Prediction of tropospheric radio trans- Open question 3. How should consistency filtrations be mission loss over irregular terrain. a computer method-1968. analyzed in a systematic way? Certainly they can be used Technical report, Institute for Telecommunication Sciences for segmentation of the observations, as is done above, but Boulder, Colorado, 1968. how can this be done most effectively? [9] Mardia KV, Jupp PE. Directional statistics, volume 494. John Wiley & Sons, 2009. MR1828667 Frontiers [10] Robinson M. Topological Signal Processing. Springer, Janu- Although sheaf theory got its start as an abstract tool, pur- ary 2014. MR3157249 posefully beyond any application, sheaves have a number [11] Robinson M. Imaging geometric graphs using inter- nal measurements. J. Differential Equations, 260:872–896, of practical applications beyond fox hunting. The inter- 2016. MR3411693 ested reader may enjoy reading how sheaves are useful in [12] Robinson M. Sheaf and duality methods for analyzing the theory of computation [3,7], network coding [2], quan- multi-model systems. In Novel Methods in Harmonic Analy- tum graphs [11], systems of differential equations, graphi- sis. Birkhäuser, 2017. MR3701273 cal models [12], and signal processors [10]. One can also [13] Robinson M. Sheaves are the canonical datastructure for look to more general tools from category theory to com- information integration. Information Fusion, 36:208–224, bine observations, sensors, and decisions under very gen- 2017. eral conditions [5, 6]. [14] Robinson M. Assignments to sheaves of pseudometric The connections between statistics, topology, and sheaves spaces arxiv:1805.08927, 2018. [15] Robinson M, Henrich J, Capraro C, Zulch P. Dynamic have steadfastly resisted our attempts at complete under- sensor fusion using local topology. In 2018 IEEE Aerospace standing. While the simulations performed in this article Conference, pages 1–7. IEEE, March 2018. used realistic noise models to produce signal reports, the sheaves that processed these reports were naïvely determin- istic. Theoretical guarantees are available for small pertur- ACKNOWLEDGMENT. The author would like to thank bations of the data, which provide a measure of robust- the anonymous referees for their detailed critiques and ness to statistical noise, but these appear to be of limited thoughtful suggestions for improving this article. He explanatory value. also wishes to express gratitude to the amateur radio The extension of an assignment supported on some community for continuing inspiration since his child- open sets to the rest of the topology has the appearance hood. This article is based upon work supported by the of a statistical imputation problem. Should the consistency Defense Advanced Research Projects Agency (DARPA) radius play the role of a decision statistic? If so, it seems and SPAWAR Systems Center Pacific (SSC Pacific) un- that given the stochastic models, the sheaf should aid the der Contract No. N66001-15-C-4040 and the Office of specification of a maximum likelihood estimate of thefox Naval Research (ONR) under Contract Nos. N00014- transmitter. These connections are only just starting to be 15-1-2090 and N00014-18-1-2541. Any opinions, find- explored! ings and conclusions or recommendations expressed

MAY 2019 NOTICES OF THE AMERICAN MATHEMATICAL SOCIETY 675 in this article are those of the author and do not nec- essarily reflect the views of the Defense Advanced Re- search Projects Agency (DARPA), SPAWAR Systems Cen- ter Pacific (SSC Pacific), or the Office of Naval Research (ONR).

mathscinet.ams.org

MathSciNet® is the authoritative gateway to the scholarly literature of mathemat- ics. Containing information on more than 3 million articles and books, with direct Credits links to over 2 million articles in more than All figures are courtesy of the author. 1,800 journals. MathSciNet includes expert reviews, customizable author profiles, and citation information on articles, books, journals, and authors. MathSciNet’s extensive resources can help you throughout your entire math career. Use it to: • Quickly get up to speed on a new topic • Look up a researcher’s body of work • Find an article or book and track its reference list • Research a math department to prepare for a job interview or when applying to graduate school

How to Subscribe

Go to www.ams.org/mathsciprice to learn more about MathSciNet, including information about subscription rates, joining a consortium, and a 30-day free trial.

© Marijn Heule

676 NOTICES OF THE AMERICAN MATHEMATICAL SOCIETY VOLUME 66, NUMBER 5