Skip to main content
Top
Published in: Data Mining and Knowledge Discovery 4/2023

Open Access 17-11-2022

FiSH: fair spatial hot spots

Authors: Deepak P., Sowmya S. Sundaram

Published in: Data Mining and Knowledge Discovery | Issue 4/2023

Activate our intelligent search to find suitable subject content or patents.

search-config
loading …

Abstract

Pervasiveness of tracking devices and enhanced availability of spatially located data has deepened interest in using them for various policy interventions, through computational data analysis tasks such as spatial hot spot detection. In this paper, we consider, for the first time to our best knowledge, fairness in detecting spatial hot spots. We motivate the need for ensuring fairness through statistical parity over the collective population covered across chosen hot spots. We then characterize the task of identifying a diverse set of solutions in the noteworthiness-fairness trade-off spectrum, to empower the user to choose a trade-off justified by the policy domain. Being a novel task formulation, we also develop a suite of evaluation metrics for fair hot spots, motivated by the need to evaluate pertinent aspects of the task. We illustrate the computational infeasibility of identifying fair hot spots using naive and/or direct approaches and devise a method, codenamed FiSH, for efficiently identifying high-quality, fair and diverse sets of spatial hot spots. FiSH traverses the tree-structured search space using heuristics that guide it towards identifying noteworthy and fair sets of spatial hot spots. Through an extensive empirical analysis over a real-world dataset from the domain of human development, we illustrate that FiSH generates high-quality solutions at fast response times. Towards assessing the relevance of FiSH in real-world context, we also provide a detailed discussion of how it could fit within the current practice of hot spots policing, as read within the historical context of the evolution of the practice.
Notes
Responsible editor: Toon Calders, Salvatore Ruggieri, Bodo Rosenhahn, Mykola Pechenizkiy and Eirini Ntoutsi.

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

1 Introduction

With sensing and tracking devices such as mobile phones and IoT becoming pervasive in this web-driven era, there is an abundance of spatial data across real-world settings. Within such spatial datasets, it is often of interest to identify geographically localized groups of entities that are of sufficient size and express a distinctive character so strongly that it is unlikely to have occurred by chance. To illustrate an example from our times, COVID-19 contact tracing apps accumulate large amounts of spatial data of people, of which some are known to have a COVID-19 infection. It would be of interest to automatically identify localized regions of high COVID-19 incidence, referred to as hot spots in contemporary reporting,1 so that the information could be channelized to health experts to identify causal reasons, or to public policy experts to develop a mitigation strategy for those regions.
While COVID-19 hot spots are characterized by high disease incidence rates, other web and new age data scenarios may call for different formulations of hot spot character, viz., high crime rates (law enforcement), intense poverty (development studies), high mobile data usage (mobile network optimization) and so on. For example, Fig. 1 illustrates hot spots of educational underachievement in India as identified from a human development dataset. In each case, identifying a set of hot spots would be of use so they could be subjected to an appropriate policy action. The unsupervised learning task of detecting spatial hot spots was pioneered by the spatial scan statistic (SSS) (Kulldorff 1997). The spatial scan statistic and its variants within the SaTScan2 toolkit have remained extremely popular for detecting spatial hot spots over the past two decades. While health and communicable diseases form the most popular application area of SSS (e.g., Pinchoff et al. 2015), they have been used within domains as diverse as archaeology (Wilczek et al. 2015) and urban planning (Helbich 2012).

1.1 Fairness in hot spots

In scenarios where spatial hot spots are to be used to inform government and public sector action, especially in sensitive policy domains (e.g., law enforcement (Mohler et al. 2018), development), it is often important to ensure that the collective population subject to the policy action is diverse in terms of protected/sensitive attributes3 such as ethnicity, caste, religion, nationality or language, among others.
Consider hot spot detection on a crime database to inform policy action that could include sanctioning higher levels of police patrols for those regions. This would likely lead to higher levels of stop and frisk checks in the identified hot spots, and would translate to heightened inconvenience to the population in the region. Against this backdrop, consider a sensitive attribute such as ethnicity. If the distribution of those who land up in crime hot spots is skewed towards particular ethnicities, say minorities as often happens (Meehan and Ponder 2002), it directly entails that they are subject to much more inconvenience than others. These, and analogous scenarios in various other sectors, provide a normatively compelling case to ensure that the inconvenience load stemming from crime hot spot detection (and other downstream processing) be proportionally distributed across ethnicities. The same kind of reasoning holds for groups defined over other sensitive attributes such as religion and nationality. It is also notable that ethnically skewed hot spot detection and patrolling would exacerbate the bias in future data, leading to viscious cycles and runaway feedback loops (Ensign et al. 2018). Minor crimes are recorded in the data only when committed as well as observed. Thus, majority and minority areas with similar real crime prevalance, alongside minority-oriented patrolling, would lead to data that records higher crime prevalance in the latter. Second, even in cases where the intended policy action is positive (e.g., setting up job support centres for unemployment hot spots), the policy being perceived as aligned to particular ethnicities could risk social solidarity and open avenues for populist backlash (Greven 2016), which could ultimately jeopardize the policy action itself.
While considerations as above are most felt in policy domains such as policing and human development, these find expression in hot spot based prioritization in provisioning any common good. Ensuring fair distribution of the impact of any policy action, across sensitive attributes such as ethnicities, is aligned with the theory of luck egalitarianism (Knight 2009), one that suggests distributive shares (of inconvenience or benefits) be not influenced by arbitrary factors, especially those of ‘brute luck’ that manifest as membership in sensitive attribute groups such as ethnicity, religion and gender (since individuals do not choose those memberships are are often just born into one). Such notions have been interpreted as a need for orthogonality between groups in the output and groups defined on sensitive attributes, and has been embedded into machine learning algorithms through the formulation of statistical parity (e.g., Abraham et al. 2020).
In summary, there is an compelling case, as in the case of other machine learning tasks, for hot spot detection to be tailored in a way that the population covered across the chosen hot spots be diverse along protected demographic groups such as ethnicity, gender religion, caste and similar. We discuss some practical scenarios for fair hot spots further in Sect. 7.

1.2 Our contributions

We now outline our contributions in this paper. First, we characterize the novel task of detection of fair spatial hot spots, for the first time. In particular, we outline a task formulation for enumerating a diverse sample of trade-off points in the noteworthiness-fairness spectrum, to suit diverse scenarios that require different trade-off points between noteworthiness and fairness. We note that straightforward solutions for the task would be computationally infeasible for even moderate dataset sizes. Second, we develop a method, FiSH, short for Fair Spatial Hot Spots, for efficiently enumerating sets of hot spots along the quality-fairness trade-off. FiSH works as a layer over any chosen fairness-agnostic spatial hot spot detection method, making it available across diverse scenarios and existing methodologies for those scenarios. Third, we outline a suite of evaluation measures to assess the quality and fairness of results for the novel fair spatial hot spots task. Lastly, we perform an extensive empirical evaluation over a real-world dataset over two separate contexts, which illustrates the effectiveness and efficiency of FiSH in identifying diverse and fair hot spots.
Given that fairness in spatial hot spots is a novel problem, we consider related work across two streams. We start by considering work on identifying outliers and spatial hot spots. These tasks are starkly different in terms of how the results are characterized. Outliers are determined based on neighborhood density over all relevant attributes in the data, whereas hot spots are determined based on hotness on a chosen hotness attribute (e.g., diseased, poor etc.) within an area of pre-specified type defined over spatial attributes. In particular, the notions of hotness attribute and spatial attributes are absent in the formulation for outlier detection, making them fundamentally different tasks. The interested reader may refer to Deepak (2016) (Fig. 2 and associated text) for a systematic analysis of the relationships between several allied tasks within this space. Despite being very different tasks, there are similarities in the overall spirit of outlier detection and hot spots, which makes outlier identification relevant to the interested reader. We start with a discussion on methods for the tasks of outlier detection and spatial hot spots, and then move on to work on fairness in machine learning as applied to tasks allied to ours.
We now discuss related work along three streams viz., outlier detection, spatial hot spots, and fairness in unsupervised learning.

2.1 Outlier identification

There have been a variety of problem settings that seek to identify objects that are distinct from either their surroundings or the broader dataset. The more popular formulations use the former notion, that of measuring contrast from the surroundings of the data object, i.e., making use of neighborhood density. LOF (Breunig et al. 2000) (and improvements (Kriegel et al. 2009)) consider identifying individual objects, aka outliers, which are interesting due to their (relatively sparser) spatial neighborhoods. It is noteworthy that these make object-level decisions informed purely by spatial attributes without reference to hotness attributes like diseased/non-diseased, as required for COVID-19 hot spot determination. SLOM (Chawla and Sun 2006) extends the object-level local neighborhood-based decision making framework to incorporate information from non-spatial attributes (e.g., age and gender of a COVID-19 patient), but do not consider hotness attributes such as diseased/non-diseased. Among outlier detection methods that assess the contrast of individual data objects with the dataset as a whole, the popular paradigm is to build a dataset level statistical model, followed by assessing the conformance of individual objects to the model; those that are less conformant would be regarded as outliers. Such statistical models could be a clustering (Yu et al. 2002), dirichlet mixture (Fan et al. 2011), or more recently, auto-encoders (Chen et al. 2017; Lai et al. 2020).

2.2 Spatial hot spots

Spatial scan statistics (SSS), pioneered by Kulldorff (1997), are methods that identify localized regions that encompass multiple objects (in contrast to making decisions on individual objects, as in LOF) which collectively differ from overall dataset on chosen non-spatial hotness attributes (e.g. diseased, poor etc.). The focus is on characterizing regions which may be interpreted as hot spots due to the divergence of their character from the overall dataset. This makes SSS a markedly different task from outlier identification in specification, input data requirements, internal workings and output format. SSS spatial hot spots are vetted using a statistical likelihood ratio test to ascertain significant divergence from global character. This makes SSS as well as its various variants, as implemented within SaTScan, a statistically principled family of methods to detect spatial hot spots. While Kulldorff’s initial proposal is designed to identify circular hot spots, the framework has been generalized to identify arbitrary shapes in several ways; ULS (Patil and Taillie 2004) is a notable work along that direction. Other methods such as bump hunting (Friedman and Fisher 1999) and LHA (Telang et al. 2014) address related problems and leverage data mining methods. Despite an array of diverse research in identifying spatial hot spots, SSS methods have remained extremely popular. Just since 2020, there have been 2000+ papers4 that make use of SSS and other scan statistics within SaTScan. Our technique, FiSH, can work alongside any method that can provide an ordered output of hot spots, such as SaTScan methodologies.

2.3 Fair unsupervised learning

While most attention within the flourishing field of fairness in machine learning (Chouldechova and Roth 2020) has focused on supervised learning tasks, there has been some recent interest in fairness for unsupervised learning tasks (Jose et al. 2020). Among the two streams of fairness explored in ML, viz., individual and group fairness (refer (Binns 2020) for a critical comparative analysis), most work on fair unsupervised learning has focused on group fairness. Group fairness targets to ensure that the outcomes of the analytics task (e.g., clusters, top-k results etc.) embody a fair distribution of groups defined on protected attributes such as gender, ethnicity, language, religion, nationality or others. As alluded to earlier, the most common instantiation of group fairness has been through the computational notion of statistical parity, initially introduced within the context of classification (Dwork et al. 2012). Group fair unsupervised learning work includes those on fair clustering (e.g., Chierichetti et al. 2017), retrieval (e.g., Zehlike et al. 2017) and representation learning (e.g., Olfat and Aswani 2019). While there has been no work on fair spatial hot spots yet, there has been some recent work on fairness in outlier detection which we discuss below.

2.3.1 Fair outliers

There has been some recent work on fair outlier detection. We start by outlining the core differences between outlier detection and hot spots to illustrate why fairness enhancements targeted at outlier detection would not be applicable for spatial hot spots. First, outlier detection often involves object-level decision making, whereas hot spotness can intrinsically be determined only at the level of object groups. Second, outlier detection methods do not make use of any non-spatial hotness attribute (e.g., diseased, poor etc.) to determine outlierness, whereas a key non-spatial attribute is used to characterize hot spots. The second difference makes algorithms for outlier detection contrast highly from those for identifying spatial hot spots. The first paper on fair outlier detection (Davidson and Ravi 2020) develops a framework for the task which is unique on multiple fronts. First, it is designed to be able to address unfairness over combinations of protected attributes leading to a deeper notions of fairness than methods that treat protected attributes one by one. Secondly, it leverages human know-how through seeking expert inputs on interpreting an explanation of potential unfairness. FairLOF (Deepak and Abraham 2020, 2021) focuses on automated group fair outlier detection, developing a technique that extends LOF (discussed above) for fairness. FairLOF adapts LOF to incorporate adjustments based on protected attribute memberships of the object in question and its neighbors, to ensure that protected groups are fairly represented among outliers. It may be noted that the protected attributes are used exclusively to embed fairness, and not to characterize outlierness. FairOD (Shekhar et al. 2020) makes a proposition of achieving group fairness (on protected attributes) while being expressly unaware of protected attributes at decision time (perhaps to avoid what is known as formal disparate treatment). A recent work (Zhang and Davidson 2021) on deep learning for fair anomalies/outliers proposes the usage of adversarial training and de-correlated representation learning to ensure that protected attributes are not correlated with outputs. To our best knowledge, there has been no prior work on fairness in detecting spatial hot spots or anomalous object groups of other kinds.

3 Problem statement

Consider a finite dataset \({\mathcal {D}} = \{ \ldots , D, \ldots \}\). Each object D is associated with a set of spatial attributes such as (xy) for a 2D space, or (latlong) for locations of people. Further, each D is associated with a non-spatial hotness attribute \(v \in \{0,1\}\) such as diseased (for epidemiology) or poor (for human development), which is used to determine spatial hot spots. D is also associated with protected attributes (e.g., ethnicity, religion) as we will see momentarily.
Consider a method for detecting spatial hot spots, such as spatial scan statistics, that is capable of providing a ranked list of top spatial hot spots, as \({\mathcal {S}} = [ S_1, S_2, \ldots , S_m]\). Each \(S_i\) is associated with a spatial region \(R_i\) (circular/spherical in the case of the basic SSS) such that the data objects (from \({\mathcal {D}}\)) that fall within \(R_i\) have a significantly different hotness profile than the overall dataset. For example, the population within \(R_i\) may have a significant high (or low) incidence rate of poverty as compared to the whole population. Noteworthiness of spatial hot spots, analyzed statistically (as in SSS), is directly related to both the size of the population within the hot spot and the extent of divergence on the hotness attribute. \({\mathcal {S}}\) is the list of spatial hot spots ordered in decreasing statistical noteworthiness; thus \(S_i\) is more noteworthy than \(S_{i+1}\). When k (typically, \(k<<m\)) noteworthy spatial hot spots are to be chosen to action upon without consideration to fairness, the most noteworthy k hot spots, i.e., \({\mathcal {S}}_{topk} = [S_1, \ldots , S_k]\), would be a natural choice.

3.1 Fair spatial hot spots

The task of fair spatial hot spots detection is to ensure that the k hot spots chosen for policy action, in addition to noteworthiness considerations as above, together encompass a diverse population when profiled along protected attributes such as ethnicity, religion, nationality etc, as motivated earlier. In other words, each demographic group is to be accorded a fair share within the collective population across the chosen hot spots. As mentioned earlier, this notion of statistical parity has been widely used as the natural measure of fairness in unsupervised learning (Chierichetti et al. 2017; Deepak and Abraham 2020; Bera et al. 2019). When the protected attributes are chosen as those that an individual has no opportunity to actively decide for herself (observe that this is the case with ethnicity, gender as well as religion and nationality to lesser extents), statistical parity aligns particularly well with the philosophy of luck egalitarianism (Knight 2013), as noted in Sect. 1.1.
We will use \({\mathcal {S}}_{fairk}\) to denote a set of k hot spots (from \({\mathcal {S}}\)) that are selected in a fairness-conscious way. It is desired that \({\mathcal {S}}_{fairk}\) fares well on both the following measures:
$$\begin{aligned}{} & {} N({\mathcal {S}}_{fairk}) = \sum _{S \in {\mathcal {S}}_{fairk}} rank_{{\mathcal {S}}}(S) \end{aligned}$$
(1)
$$\begin{aligned}{} & {} F({\mathcal {S}}_{fairk}) = \sum _{P \in {\mathcal {P}}} Div_P({\mathcal {D}}, Pop({\mathcal {S}}_{fairk})) \end{aligned}$$
(2)
The first, N(.), relates with noteworthiness and is simply the sum of the ranks (ranks within \({\mathcal {S}}\)) of the chosen spatial hot spots (S denotes a spatial hot spot, a set of items); \(rank_{{\mathcal {S}}}(S)\) denotes the rank of S within the list \({\mathcal {S}}\). Lower values of N(.) are desirable, and \({\mathcal {S}}_{topk}\) scores best on N(.), due to comprising the top-k (so, \(N({\mathcal {S}}_{topk}) = \sum _{i=1}^k i = \frac{k \times (k+1)}{2}\)). The second, F(.), is a fairness measure, which requires that the population subset covered across the hot spots within \({\mathcal {S}}_{fairk}\) (denoted \(Pop({\mathcal {S}}_{fairk})\)) be minimally divergent to the overall population, when measured on a pre-specified set of protected attributes \({\mathcal {P}}\) (e.g., ethnicity, gender); \(Div_P(.,.)\) measures divergence on attribute \(P \in {\mathcal {P}}\). In other words, \(Pop({\mathcal {S}}_{fairk})\) denotes the population subset chosen for policy action, and we are interested in measuring how divergent this population subset is, from the overall population, on the sensitive attributes. We use Wasserstein distance (Vallender 1974; Yoon et al. 2020) to compute divergence yielding the following formula:
$$\begin{aligned} Div_P({\mathcal {D}}, Pop({\mathcal {S}}_{fairk})) = Wass(Distrib_P({\mathcal {D}}), Distrib_P(Pop({\mathcal {S}}_{fairk}))) \end{aligned}$$
(3)
where \(Distrib_P(.)\) is the normalized distribution of the population across the values of attribute P5. For example, if \({\mathcal {D}}\) has a 50:50 distribution on males and females and the sub-population in \(Pop({\mathcal {S}}_{fairk})\) has a 55:45 distribution, \(Div_P(.,.)\) for this case will be Wass([0.5, 0.5], [0.55, 0.45]), where Wass(., .) denotes the Wasserstein distance. When the distributions are identical, statistical parity is achieved, and \(Div_P(.,.)\) evaluates to 0.0. The usage of Wasserstein measure as an evaluation measure for various fair ML algorithms (Wang and Davidson 2019; Deepak and Abraham 2020) and other contexts in fair ML (Miroshnikov et al. 2020) motivate its usage here. Intuitively, Wasserstein measures the minimal cost of transporting one distribution to another. \(Div_P(.,.)\) can be easily modified to use another distance measure should such a motivation arise in a particular domain. As in the case for N(.), lower values of F(.) are desirable too. Since F(.) is an aggregate of \(Div_P(.)\)s, achievement of statistical parity over each of the protected attributes would naturally lead to F(.) evaluating to 0.0. Though lower, and not higher, values of N(.) and F(.) indicate deeper noteworthiness and fairness, we refer to these measures as noteworthiness and fairness to avoid introducing new terminology.
Fair hot spots and fair ranking The notion of fair hot spots is quite different, and arguably more complex, than tasks such as fair ranking that have been explored in literature (e.g., Zehlike et al. 2017). We briefly discuss the relationship between these tasks. Without loss of generality, fair ranking may be easily conceptualized within a hiring shortlisting scenario where the fairness need is to ensure a fair representation of gender and race groups within the pool of shortlisted candidates. Here each object—which is an applicant in the hiring scenario—has a membership within the protected group (e.g., male when gender is the chosen protected group), and group fairness is sought to be achieved within the shortlisted pool. Despite the superficial high-level structural similarity that fair hot spots seeks to re-order objects in \({\mathcal {S}}\) in accordance with fairness, the sharp departure is evident when one observes that what is sought to be ranked within the fair hot spots formulation are hot spots, which are not individual objects, but sets of objects. Hot spots relate to one another in set relationships (e.g., disjoint, overlapping, subset etc.), and each hot spot, due to comprising multiple objects, has a distribution of memberships over a protected attribute (e.g., gender). These two factors make the task of fair hot spots much more nuanced and intricate than fair ranking. It is also notable that given the structure of fair hot spots task, there is no direct dependency on the dataset size other than through the hot spots detection method employed.

3.2 Diverse selection of \({\mathcal {S}}_{fairk}\) candidates

The noteworthiness and fairness considerations are expected to be in tension (an instance of the often discussed fairness-accuracy tension (Menon and Williamson 2018)), since fairness is not expected to come for free (as argued extensively in Kearns and Roth (2019)). One can envision a range of possibilities for \({\mathcal {S}}_{fairk}\), each of which choose a different point in the trade-off between N(.) and F(.). At one end is the \({\mathcal {S}}_{topk}\) (best N(.), likely worst F(.)), and the other end is a maximally fair configuration that may include extremely low-ranked hot spots from \({\mathcal {S}}\). These would form the Pareto frontier6 when all the \(^{m}C_k\) (k sized) subsets of \({\mathcal {S}}\) are visualized as points in the 2D noteworthiness-fairness space, as illustrated in Fig. 2. Each point in the Pareto frontier (often called skyline (Borzsony et al. 2001)) is said to be Pareto efficient or Pareto optimal since there is no realizable point which is strictly better than it on \({\underline{\hbox {both}}}\) N and F measures. Thus, \({\mathcal {S}}_{fairk}\) candidates that are not part of the Pareto frontier can be safely excluded from consideration, since there would be a Pareto frontier candidate that is strictly better than it on both noteworthiness and fairness.
Each policy domain may choose a different point in the trade-off offered across candidates in the Pareto frontier, after due consideration of several available trade-off points. For example, policing may require a high-degree of fairness, whereas epidemiology interventions may be able to justify policy actions on less diverse populations based on the extent of supporting medical evidence. The Pareto frontier may be large (could contain hundreds of candidates, bounded above only by \({\mathcal {O}}(^{m}C_k)\)) for a human user to fully peruse. The extreme case, where each of \(^mC_k\) k-sized subsets of C are in the Pareto frontier, appears where no pairs are involved in a Pareto domination relationship. Thus, an obvious recourse would be to identify \(\tau \) diverse Pareto efficient candidates (henceforth, \(\tau \)-dpe), where \(\tau \) is a pre-specified parameter, so the human user may be able to choose appropriately from a varied set of possibilities. A natural and simple but incredibly inefficient solution would be to (i) enumerate the entire Pareto frontier, (ii) trace the sequence of Pareto efficient points from the top-left to the bottom-right (i.e., the dotted line), (iii) split the sequence into \(\tau - 1\) equally sized segments, and (iv) take the \(\tau \) segment end points as the result.
To summarize, the diverse candidate selection task outlined as \(\tau \)-dpe requires a diverse set of Pareto efficient candidates in the N–F space, each candidate representing a k sized subset of \({\mathcal {S}}\).

3.3 Approximate \(\tau \)-dpe

It may be observed that it is infeasible to enumerate the \(^mC_k\) subsets (e.g., \(^{40}C_{10} = 8.5E\)+8) in the N–F space just due to the profusion of possibilities, making exact \(\tau \)-dpe identification (as outlined in the four-step process in the previous section) infeasible for practical scenarios. This makes the task of identifying a close approximation of \(\tau \)-dpe results efficiently a natural alternative for a policy expert to examine the trade-off points and arrive at a reasonable choice of \({\mathcal {S}}_{fairk}\) to subject to policy action. This brings us to the approximate \(\tau \)-dpe task, which is that of efficiently identifying a close approximation of the exact \(\tau \)-dpe result. The set of circled points in Fig. 2 illustrates a possible solution to the approximate \(\tau \)-dpe task. All pertinent notations are outlined in Table 1 for easy reference.
It is useful to note the nature of likely usage contexts for \(\tau \)-dpe to provide some perspective on scalability. Whether it be the case of crime hot spots to inform police patrolling strategies, the case of poverty hot spots to inform public policy or even mobile data usage hot spots to inform network provisioning decisions, all of these are what we could call as offline tasks. In other words, while it is necessary to ensure that hot spots be identified in reasonable time, real-time responses are neither expected nor necessary. For example, while a response time in days or months (as would be required to traverse an exponential space) would be unacceptable, a response time of a few hours would be quite fine for practical scenarios. This is often common in unsupervised outlier detection as well; for example, the response time of a recently proposed random projection based outlier detection method (Bhattacharya et al. 2021) is of the order of several minutes or hours7. The approximate \(\tau \)-dpe formulation thus intends to bring the \(\tau \)-dpe task from the region of infeasible response times to the region of feasibility. Our method, FiSH, that addresses the approximate \(\tau \)-dpe task, is detailed below.
Table 1
Table of notations for easy reference
Notation
What it stands for
\({\mathcal {S}}\)
The ordered list of spatial hot spots used as
Starting point for \(\tau \)-dpe task
\({\mathcal {S}}_{topk}\)
The subset of k most noteworthy hot spots from \({\mathcal {S}}\)
\({\mathcal {S}}_{fairk}\)
k-sized subset of \({\mathcal {S}}\), a candidate for fair
selection of hot spots
\(N({\mathcal {S}}_{fairk})\)
Sum of ranks of the spatial hot spots within
\({\mathcal {S}}_{fairk}\); lower denotes better noteworthiness
\(F({\mathcal {S}}_{fairk})\)
Deviation of \({\mathcal {S}}_{fairk}\)’s population from dataset on
Protected attributes; Lower denotes better fairness
m
Cardinality of \({\mathcal {S}}\)
k
# hot spots from \({\mathcal {S}}\) Desired in each output candidate
\(\tau \)
Number of candidates desired in output
b
Beam width parameter used by FiSH (Sect. 4)

4 FiSH: fair spatial hot spots

FiSH is an efficient heuristic-driven technique addressing the approximate \(\tau \)-dpe task outlined above. We first describe a systematic organization of the search space, followed by a heuristic method that traverses the space prioritizing the search using three considerations: Pareto efficiency, diversity and efficient search.

4.1 Search space organization

Recall that we start with a noteworthiness-ordered list of spatial hot spots \({\mathcal {S}} = [S_1, \ldots , S_m]\). Our full search space comprises the \(^mC_k\) distinct k-sized subsets of \({\mathcal {S}}\). We use the lexical ordering in \({\mathcal {S}}\) to organize these candidates as leaves of a tree structure, as shown in Fig. 3. Each node in the tree is labelled with an element from \({\mathcal {S}}\), and no node in the FiSH search tree has a child that is lexically prior to itself. Such a hierarchical organization is popular for string matching tasks, where they are called prefix trees (Yazdani and Min 2001). In devising FiSH, we draw inspiration from using prefix structures for skyline search over databases (Deshpande et al. 2009). Each internal node at level l (root level \(=0\)) represents a l-sized subset of \({\mathcal {S}}\) comprising the l nodes indicated in the path from root to itself. The lexical ordering ensures that each subset of \({\mathcal {S}}\) has a unique position in the tree, one arrived at by following branches corresponding to nodes in the subset according to the lexical ordering. The \(^mC_k\) candidates would be the nodes at level k. It is infeasible to enumerate them fully, as observed earlier. Thus, FiSH adopts a heuristic search strategy to traverse the tree selectively to follow paths leading to a good solution (i.e., set of \(\tau \) nodes at level k) for the approximate \(\tau \)-dpe task.

4.2 FiSH search strategy

The exact \(\tau \)-dpe result set is characterized by Pareto efficiency and diversity, when applied over the \(^mC_k\) candidates. The FiSH search strategy uses precisely these criteria as heuristics to traverse the search tree efficiently from the root downward. The core idea behind this search strategy is our conjecture that Pareto efficiency and diversity at a given level in the FiSH search tree would be predictive of Pareto efficiency and diversity at the next level. We operationalize this heuristic strategy using beam search, a classical memory-optimized search meta-heuristic (Steinbiss et al. 1994) that has received much recent attention (Wiseman and Rush 2016).
Having outlined all relevant notation and an overview of the search strategy, we now describe it in simple terms. FiSH starts its search from the root node, expanding to the first-level child nodes, each of which represent singleton-sets denoting the choice of a particular spatial hot spot from \({\mathcal {S}}\). This forms the candidate set at level 1 of the FiSH tree, \({\mathcal {C}}_1 = \{ \{S_1\}, \{S_2\}, \ldots \}\). These 1-sized subsets of \({\mathcal {S}}\) are then arranged in an N–F space as in Fig. 2. Note that the N–F space of 1-sized subsets is distinct and different from the N–F space of k-sized subsets (Fig. 2). The Pareto-efficient subset of \({\mathcal {C}}_1\) is then identified as \(P({\mathcal {C}}_1)\). The candidates in \(P({\mathcal {C}}_1)\) are then arranged in a linear sequence tracing the Pareto frontier from the top-left to the bottom-right point (similar to the illustration of Pareto frontier in Fig. 2). This linear sequence is split into \(b-1\) equally spaced segments, and the b points at the segment end-points are chosen as \(D_b(P({\mathcal {C}}_1))\), a b-sized subset of Pareto efficient points from \({\mathcal {C}}_1\). The candidate set at the next level of the tree search process, i.e., \({\mathcal {C}}_2\), is simply the set of all children of nodes in \(D_b(P({\mathcal {C}}_1))\) (actually, the subsets of \({\mathcal {S}}\) that they stand for).
$$\begin{aligned} {\mathcal {C}}_2 = \bigcup _{c \in D_b(P({\mathcal {C}}_1))} children(c) \end{aligned}$$
(4)
It may be noted that \({\mathcal {C}}_2\) is a small subset of the set of all 2-sized subsets of \({\mathcal {S}}\), since only children of the b nodes selected from the previous level are selected for inclusion in \({\mathcal {C}}_2\). Next, \({\mathcal {C}}_2\) is subject to the same processing as \({\mathcal {C}}_1\) comprising:
1.
identifying Pareto efficient candidates \(P({\mathcal {C}}_2)\),
 
2.
identifying a diverse b sized subset \(D_b(P({\mathcal {C}}_2))\), and
 
3.
following the children pointers,
 
to arrive at the candidate set for the next level. This process continues up until \({\mathcal {C}}_k\) whereby the Pareto frontier \(P({\mathcal {C}}_k)\) is identified followed by the choice of \(\tau \) diverse candidates which will eventually form FiSH’s result set for the approximate \(\tau \)-dpe task. This search strategy is illustrated formally in Algorithm 1. The one-to-one correspondence between nodes in the search tree and subsets of \({\mathcal {S}}\) allows us to use them interchangeably in the pseudocode.

4.3 Discussion

FiSH’s search strategy makes use of Pareto efficiency and diversity directly towards identifying a small set of nodes to visit at each level of the tree. Restricting the search to only b nodes at each level before moving to the next enables efficiency. Smaller values of b enable more efficient traversal, but at the cost of risking missing out on nodes that could lead to more worthwhile members of the eventual result set. In other words, a high value of b allows a closer approximation of the \(\tau \)-dpe result, but at a slower response time. It may be suggested that b be set to \(\ge \tau \), since the algorithm can likely afford to visit more options than a human may be able to peruse eventually in the result set. The candidate set size at any point, and thus the memory requirement, is in \({\mathcal {O}}(bm)\). The computational complexity is in \({\mathcal {O}}(kb^2m^2)\), and is dominated by the Pareto frontier identification (which is in \({\mathcal {O}}(b^2m^2)\)) at each level. While b is a controllable hyperparameter (likely in the range of 5-20), m can be constrained by limiting FiSH to work with the top-m result set (as \({\mathcal {S}}\)) from the upstream spatial hot spot technique.

5 Evaluating approx \(\tau \)-dpe results

Given that (approximate) \(\tau \)-dpe is a new task we proposed, we now describe novel evaluation metrics to assess the quality of FiSH’s results. Recall that, given the N–F space comprising all k-sized subsets of \({\mathcal {S}}\), the choice of \(\tau \) equally spaced skyline candidates forms the result set for the exact \(\tau \)-dpe task that we propose in this paper. This result set, which we call Exact, is computationally infeasible for moderate datasets, but forms our natural baseline for measuring FiSH’s effectiveness. In other words, the results of Exact form the ground truth that FiSH seeks to approximate efficiently. Approximate \(\tau \)-dpe results from FiSH may be evaluated either directly based on how well they approximate the expected results of the exact \(\tau \)-dpe task, or based on how well they adhere to the spirit of the \(\tau \)-dpe task of identifying a diverse group of Pareto efficient subsets of \({\mathcal {S}}\). We now devise evaluation measures along the lines above. In what follows, we use \({\mathcal {P}}\) to denote the \(^mC_k\) k-sized subsets of \({\mathcal {S}}\).

5.1 Direct comparison

Let the result of the exact \(\tau \)-dpe task be \({\mathcal {E}} = [E_1, \ldots , E_{\tau }]\), and FiSH’s result be \({\mathcal {F}} = [F_1, \ldots , F_{\tau }]\). We would like the average distance between corresponding elements to be as low as possible.
$$\begin{aligned} DC({\mathcal {E}}, {\mathcal {F}}) = \frac{1}{\tau } \sum _{i=1}^{\tau } Eucl(E_i,F_i) \end{aligned}$$
(5)
where Eucl(., .) is the euclidean distance in the N–F space. Notice that when \({\mathcal {E}} = {\mathcal {F}}\), DC(., .) evaluates to 0.0. Given that N(.) and F(.) would be in different ranges, we will compute the distance after normalizing both of these to [0, 1] across the dataset. As may be obvious, smaller values, i.e., as close to 0.0 as possible, of DC(., .) are desirable.

5.2 Quantifying Pareto-ness: coverage

A diverse and Pareto efficient set may be expected to collectively dominate most objects in the N-F space. Accordingly, we devise a measure, called coverage, that measures the fraction of candidates in \({\mathcal {P}}\) that are Pareto dominated by at least one candidate in \({\mathcal {F}}\):
$$\begin{aligned} Cov({\mathcal {F}}) = \frac{1}{\vert {\mathcal {P}} \vert } \sum _{P \in {\mathcal {P}}} {\mathbb {I}}(\exists F \in {\mathcal {F}} \vert F \succ P) \end{aligned}$$
(6)
where \(F \succ P\) is true when F Pareto dominates P. A point Pareto dominates another if the latter is no better than the former on both attributes, excluding the case where both are identical in terms of their N–F co-ordinates. A candidate being dominated by another indicates that the latter characterizes an absolutely better trade-off point than the former (on both N(.) and F(.)). Thus, we would like the result set to be in a way that most, if not all, candidates are dominated by one or more candidates in the result set. Cov(.) is measured as a fraction of the candidates dominated, hence it is in the range [0, 1]. Full coverage (i.e., \(Cov(.) = 1.0\)) may not be attainable given that only \(\tau \) candidates can be chosen in the result; instead, if we were to choose the entire skyline, we would get \(Cov=1.0\) by design. Thus, the extent to which \(Cov({\mathcal {F}})\) (FiSH’s coverage) approaches \(Cov({\mathcal {E}})\) (coverage attained by the exact result) is a measure of FiSH’s quality. Coverage, being modelled using Pareto domination, may be seen as modelling Pareto-ness of FiSH’s result.

5.3 Diversity of results

Given that our formulation of the approximate \(\tau \)-dpe task hinges on the idea that the candidates should be diverse (so that they may embody a variety of different trade-off points), diversity is a key aspect to measure the adherence of the solution to the spirit of the approximate \(\tau \)-dpe task. We model diversity as the minimum among pairwise distances between candidates in \({\mathcal {F}}\):
$$\begin{aligned} MD({\mathcal {F}}) = min \{ Eucl(F_i,F_j) \vert \{ F_i, F_j \} \subseteq {\mathcal {F}}, F_i \ne F_j \} \end{aligned}$$
(7)
Unlike the average of pairwise distances that allows nearby pairs to be compensated by the existence of far away ones, this is a stricter measure of diversity. On the other hand, this is quite brittle, in the sense just one pair of results being proximal would cause MD(.) to go down significantly; in such cases, the MD(.) would not be that representative of the overall diversity in \({\mathcal {F}}\). Hence, all the evaluation measures must be seen in cognisance of the others. Coming to desirable values of MD(.), we would like \(MD({\mathcal {F}})\), which measures the lower bound of distances among elements in \({\mathcal {F}}\), to be as high as possible, and approach the diversity of \({\mathcal {E}}\), i.e., \(MD({\mathcal {E}})\).

5.4 Discussion

As obvious from the construction, lower values of DC, and higher values on both Cov and MD indicate the quality of FiSH’s approach. It is also to be seen that Cov and MD should be judged together, since it is easy to maximize coverage without being diverse and vice versa. Cov and MD requires all \(^mC_k\) subsets of \({\mathcal {S}}\) to be enumerated, whereas DC requires additionally that the exact \(\tau \)-dpe results be computed. This makes these evaluations feasible only in cases where such enumeration can be done, i.e., reasonably low values of m. In addition to the above quality measures, a key performance metric that FiSH seeks to optimize for, is the response time.

6 Experimental evaluation

We now describe our empirical study evaluating FiSH. In this section, we describe the dataset used, the experimental setup, our evaluation measures and our experimental results.

6.1 Dataset and experimental setup

We describe the dataset and experimental setup in separate subsections.

6.1.1 Dataset

We used the Indian Human Development Survey (IHDS)8 dataset, a large-scale survey of India’s population conducted in 2011-12. This is one among very rare datasets—the only one we came across—that comprises personal information attributes along with locations. In particular, we used a random sample of 10000 individuals from the data with distinct locations. The location (lat, long) was determined through querying Google Maps based on the district and other location information available in the data. The binary hotness attribute was chosen as either (i) (annual) income < It may be noted that100k,9 or (ii) education < 2 years. For each setting, we use caste and religion as sensitive attributes and low income/education as hot spot criterion. In other words, we would like to identify a set of spatial hot spots such that the population across them fare poorly on income (education) but religion and caste groups are fairly represented. These choices of attributes for hotness and fairness are abundantly informed by social realities in contemporary India; for example, caste discrimination remains rampant across India, including in urban settlements.10

6.1.2 Experimental setup

We used SaTScan Bernoulli model to discover hot spots. SaTScan is among the earliest and most popular hot spots detection methods; most improvements upon them are for more specialized scenarios. This backdrop makes usage of SaTScan most appropriate to showcase generality of FiSH, given that it operates as a layer over hot spots detection. We implemented FiSH as well as the Exact \(\tau \)-dpe computation (i.e., enumerate all \(^mC_k\) subsets, find Pareto efficient frontier, and identify \(\tau \) diverse subsets) on Python 3 on an Intel 64 bit i5-8265 at 1.6 GHz with 8 GB RAM. Unless otherwise mentioned, we use the following parameter settings: \(m=20\), \(k=5\) and \(\tau = b = 5\). The source code for FiSH is available at https://​github.​com/​Sowms/​FiSH-Fair-Spatial-Hotspots.

6.2 Overall comparison

We performed extensive empirical analyses over varying settings. We present representative results and analyses herein. Table 2 illustrates a representative sample of the overall trends in the comparison between FiSH and Exact. The low values of DC indicate that FiSH’s results are quite close to those of Exact, which is further illustrated by the trends on the Cov measure where FiSH follows Exact closely. For MD, we observe a 20% deterioration in the case of Income, and a 50% deterioration in the case of Education. We looked at the case of Education and found that the low value of MD for FiSH was due to one pair being quite similar (distance of 0.041), possibly a chance occurrence that coincided with this setting; the second least distance was more than three times higher, at 0.1349. On an average, the pairwise distances for FiSH was only 20% less than that for Exact. Across varying parameter settings, a 15-20% deterioration of MD was observed for FiSH vis-a-vis Exact. For the record, we note that the choice of first k hot spots from \({\mathcal {S}}\) as the result yielded \(DC \approx 0.8\) and Cov 3 to 10 percentage points lower; this confirms that \(\tau \)-dpe task formulation is significantly different from top-k not just analytically, but empirically too.
Table 2
Comparative results (task setting: \(\tau = 5\), \(k=5\), \(m=20\) and parameter setting: \(b=5\) for FiSH); arrows denote whether low or high values are desirable
Setting
Method
DC \(\downarrow \)
Cov \(\uparrow \)
MD \(\uparrow \)
Time(s) \(\downarrow \)
Income
FiSH
0.112
0.995
0.034
23.11
Exact
N/A
0.998
0.042
6536.54
Education
FiSH
0.045
0.987
0.041
23.87
Exact
N/A
0.997
0.081
4413.78
Table 3
Scalability analysis: running time (in seconds) with varying m; Exact did not complete in reasonable time for \(m>25\)
 
Education
m
FiSH
Exact
15
17.83
840.37
20
23.87
4413.78
25
39.46
33151.91
30
49.28
 
35
61.49
 
40
71.09
 
Apart from being able to approximate the Exact results well, FiSH is also seen to be able to generate results exceptionally faster (as expected, given the design of FiSH), a key point to note given that bringing the \(\tau \)-dpe task into the realm of computational feasibility was our main motivation in devising FiSH, as outlined in Sect. 3.3. In particular, FiSH’s sub-minute response times compare extremely favourably against those of Exact which is seen to take more than an hour; we will illustrate later that Exact scales poorly and rapidly becomes infeasible for usage within most practical real-life scenarios.
The FiSH versus Exact trends, reported in Table 2 is representative of results across variations in parameter settings. FiSH was consistently seen to record 0–10% deteriorations in Cov, around 15–25% deterioration in MD, and multiple orders of magnitude improvements in response time. The trends on the effectiveness measures as well as the response time underline the effectiveness of the design of the FiSH method.

6.3 Scalability analysis

With FiSH being designed for efficient computation of a reasonable approximation of \(\tau \)-dpe results, it is critical to ensure that FiSH scales with larger m; recall that \(m=\vert {\mathcal {S}}\vert \), the size of the initial list of hot spots chosen to work upon. Table 3 illustrates the FiSH and Exact response times with varying m. While Exact failed to complete in reasonable time (we set a timeout to 12 h) for \(m>25\), FiSH was seen to scale well with m, producing results many orders of magnitude faster than Exact. In particular, it was seen to finish its computation in a few minutes even for \(m \approx 100\), which is highly promising in terms of applicability for practical scenarios. Similar trends were obtained with scalability with higher values of k and \(\tau \); Exact quickly becomes infeasible, whereas FiSH’s response time grows gradually.

6.4 FiSH fairness analysis

In addition to being able to generate good approximations of Exact at fast response times, it is also pertinent to analyze the fairness achieved over sensitive attributes by FiSH to get a sense of the levels of fairness achieved by FiSH. One may recollect that FiSH generates a set of results spread across the N–F spectrum. At the Fairness end, we expect high degrees of fairness, where we expect that the distribution over sensitive attributes achieved over \(Pop({\mathcal {S}}_{fairk})\) (Ref. Eq. 2) closely approximates that in the whole dataset; statistical parity is said to be achieved absolutely when these distributions are identical. The extent to which the F-end result’s distribution on the sensitive attributes reflects the dataset distribution is thus a key fairness metric for FiSH. Table 4 tabulates the distributions for this analysis. As may be seen therein, the distribution over the religion and caste for the Income case closely follows the full dataset distributions, with deviations under \(2\%\) on an average. The corresponding deviations are just over \(2\%\) for the Education dataset. These indicate that FiSH is able to provide a result option that achieves very high levels of fairness over the sensitive attributes.

6.5 Analysis over varying settings

We now analyze the performance of FiSH in varying settings. This analysis helps us evaluate the sensitivity of FiSH to specific parameter values; for example, smooth movements along small variations in parameter values will help build confidence in the generalizability of FiSH across varying scenarios. With Exact being unable to complete running within reasonable amounts of time for higher search spaces (e.g., \(m>25\), \(k=7\), \(\tau > 5\) etc.), we restrict our attention to FiSH trends over Cov and MD; this is so since results from Exact are necessary to compute the DC measure. Among Cov and MD, our expectation is that the brittleness of the MD measure, as noted in Sect. 5.3, could lead to more fluctuations in MD when compared to Cov, even when FiSH results change only gradually. We now study the trends with varying parameter settings, changing parameters one at a time, keeping all parameters at their reference settings from Sect. 6.1.2, except the one being varied.

6.5.1 Varying m

We now analyze the effectiveness of FiSH when operating over a larger set of SaTScan results, i.e., with larger values of m (recall \(m=\vert {\mathcal {S}}\vert \)). With the number of points in the N–F space being \(^mC_k\), increases in m lead rapidly to much denser N–F spaces, and correspondingly larger search spaces. We vary m from 15 to 30 in steps of 5; the Cov and MD trends appear in Figs. 4 and 7 respectively. As expected, Cov consistently remains at high values, higher than 0.985, whereas there is higher volatility in the case of MD. The trends indicate that FiSH is not highly sensitive to m and the quality of its results varies gradually with varying values of m.
Table 4
FiSH fairness analysis
Scenario
Income
Education
Religion
Caste
Religion
Caste
Full dataset
\([0.270\ 0.730]\)
\([0.210\ 0.790]\)
\([0.310\ 0.690]\)
\([0.200\ 0.800]\)
F-end FiSH
\([0.267\ 0.733]\)
\([0.217\ 0.783]\)
\([0.303\ 0.697]\)
\([0.195\ 0.805]\)

6.5.2 Varying \(\tau \)

The number of trade-off points that is provided to the user, or \(\tau \), is another important parameter in the \(\tau \)-dpe task. The beam size in FiSH, as observed earlier in Sect. 5.4, is intimately related to \(\tau \), and may be expected to be set such that \(b \ge \tau \). Higher values of b yield better results at the cost of slower responses; we consistently set \(b=\tau \) in our result quality analysis. Higher values of \(\tau \) enable choosing more points from the N–F space in the output, and this provides an opportunity to improve on Cov. However, choosing more points obviously would lead to deterioration in the MD measure that measures the minimum of pairwise distances. We vary \(\tau \) (and thus b) from 3 to 7, and plot the Cov and MD trends in Figs. 5 and 8 respectively, which show gentle and consistent variations. As expected, Cov is seen to improve and saturate close to the upper bound of 1.0. MD on the other hand, is seen to deteriorate but stabilizes soon; the patterns are consistent except for the case of \(\tau =5\) for Education, likely a chance occurrence as analyzed in Sect. 6.2.

6.5.3 Varying k

The third parameter of importance for the \(\tau \)-dpe task is k, which denotes the number of hot spots to be chosen within each trade-off point in the result. Increasing values of k (up to m/2) lead to larger number of points in the N–F space. With the number of trade-off points to be output pegged at \(\tau \), achieving the same coverage would become harder with increasing k. This is in contrast with MD where there is no expectation of a consistent deterioration or improvement. From the Cov and MD plots in Figs. 6 and 9, the Cov is quite stable with a deterioration kicking in at \(k=7\) (even there, Cov remains at \(0.90+\)), whereas MD remains consistent.

6.5.4 Setting b

The beam width, b in FiSH, offers a mechanism to trade-off effectiveness for efficiency. We experimented with varying values of b and found that the gains on effectiveness measures (i.e., DC, Cov and MD) taper off beyond \(b > 2 \times \tau \). The response times were seen to increase with b; there are two ways in which b affects the complexity, one is by providing more candidates at each level (which increases linearly with b), and another by increasing the cost of Pareto frontier identification (which is in \({\mathcal {O}}(b^2)\)). From the trends which indicated a linear trend between response time and b, it may be reasonably suspected that the former factor dominates.

6.6 Example results in the N–F space

Having analyzed FiSH quantitatively, we now consider a qualitative evaluation of FiSH vis-a-vis Exact. Fig 10 illustrates the N–F space for our reference setting (Sect. 6.1.2) for Income, with results from FiSH (green points) juxtaposed against Exact results (mustard yellow) and other points in red. This result is representative of FiSH’s strengths and weaknesses. While three of five FiSH results are seen to be almost on the Pareto frontier, the others are only slightly inward. As in the case of any heuristic-driven method, FiSH may miss some good results; here, FiSH’s sampling misses out on the top-left region of the Pareto frontier, which explains the slight deterioration in Cov for FiSH when compared with Exact.

7 Fair hot spots in practice

Against the backdrop of having discussed the technical details of our method, we now discuss the applicability of FiSH in practical scenarios. In particular, we motivate the role of fair hot spots—and methods such as FiSH—within the domain of crime prevention and surveillance, focusing on the specific context of hot spot policing.

7.1 Hot spots policing: the context

Policing and crime prevention is a policy domain where hot spot detection is used in an apparent and visible manner, with much social and political implications. We scan the historical context of policing and crime prevention, and how the role of hot spot policing has evolved within it. Traditional notions of policing and crime prevention have focused on people, given the inarguable role of human agency in crime. With such people-focused policing, geographical considerations have traditionally focused on high-level and long-term decision making such as determining locations where police force needs to be stationed to ensure timely responses. The emergence of place as an important and explicit consideration within crime prevention arguably may be traced across four decades. A 1982 article titled ’broken windows’ (Wilson and Kelling 1982) foregrounded the idea that visible signs of anti-social behavior, such as broken windows, could be read as an evidence of enhanced crime-proneness within a place. This theory found much favor under the mayoralty of Rudy Giuliani11 in New York in the 1990s whose implementation of the theory included harsh crackdowns on minor crimes such as graffiti and turnstile jumping. This, probably among the first widespread place-focused policing in a wide scale, has led to much criticism and accusations of racism, with a 2007 book titled ’Why blacks fear America’s Mayor’ (Noel 2007) using the phrase ’one of the most racially divisive leaders in the nation’ to describe him. On the other hand, there has been much support for the approach, often citing sharp drops in crime in New York city under the Mayoralty of Giuliani (Langan and Durose 2003).
While broken windows is a place-based approach, modern approaches towards crime prevention often use the phrase hot spot policing12 explicitly, to denote a variety of place-based approaches in crime prevention. This relies on the assumption that police can be effective in addressing crime and disorder when they focus on small units of geography with high rates of crime. These hot spots and policing strategies and tactics focused on such areas are usually referred to as hot spots policing. The effectiveness of such strategies in deterring and/or reducing crime have been established through various studies (Sherman and Weisburd 1995). Over the past decades, the logic of hot spots policing has only gained in prominence, and has even been referred to as ’the law of crime concentration in places’ within an article (Braga et al. 2017) that says ’The empirical observation that a small number of micro places generate the bulk of urban crime problems has become a criminological axiom.’ The high impetus on hot spot policing is reflected in an increasing prioritization of government funding towards it13. The widespread emergence of the usage of AI-based tooling to assist place-based predictive policing (e.g., PredPol (Meliani 2018)) has played a significant role in the increased impetus on hot spot policing within contemporary society.

7.2 Fairness and hot spots policing

As often observed in fair machine learning research, effectiveness and fairness are often in tension (Kearns and Roth 2019). The growth of place-based policing—often enacted through increased surveillance of hot spot locations within the city—has sharply coincided with concerns on systemic racism within policing. A recent book (Gordon 2022) uses the phrase remaking of segregation and—through a police shadowing study within a US context—draws a contrast between two sides of the city: ’one where rich, white neighborhoods are protected, and another where poor, black neighborhoods are punished’. Such observations, one might recollect, dominated the public discourse during the George Floyd protests of 2020.14 While most studies have focused on US contexts, similar issues of bias in place-based policing have been brought up in other contexts within the Global South, such as the impact gradient of policing along caste lines within India (Narayan 2021). The developing narrative arguing against new policing, a term often used for technology-based and hot spot policing, and instead proposing a newer policing, one that is focused on rights, fairness and policing legitimacy, has seen growing acceptance. An article in The Hill15 says: ’The epidemic of police brutality—primarily affecting black males—can be linked to the history of a technique called hot spot policing, ...’. The prominence of this narrative was acknowledged by a pioneer of hot spot policing, David Weisburd, whose 2016 article (Weisburd 2016) is titled: Does Hot Spots Policing Inevitably Lead to Unfair and Abusive Police Practices, or Can We Maximize Both Fairness and Effectiveness in the New Proactive Policing?. While Weisburd offers a positive view of hot spot policing in addressing this question through observing that hot spots policing encompasses a wide variety of implementation possibilities, he notes that ’Hot spots policing programs should be developed and implemented by police managers with the ideas of legitimacy and fairness in mind.’
Apart from such extant observations that policing is being unfair and thus antithetical to a modern society, these additionally undermine the legitimacy of the police force, sowing the seeds for greater disharmony. We observe a sharp contrast between such emerging understanding within social science literature on the tensions between fairness and hot spot policing, and the lack of AI literature to provide enabling technological pathways towards bridging this tension. It is this deficit that our paper seeks to foreground and make initial strides towards addressing. Having outlined the context of hot spot policing and the need for fairness, we now consider how aspects of FiSH could be aligned with the challenges of fair hot spot policing.

7.3 FiSH and fair hot spots policing

A key aspect of the fairness conceptualization within FiSH is that fairness is sought to be achieved across the collection of hot spots to be chosen for policy action rather than within each hot spot. This is based on the observation that imposing fairness conditions at the individual hot spot level could lead to an extremely constrained technical formulation. As emphasized in a recent interdisciplinary critical studies work (Webber and Burrows 2018), demographic identities in large urban locations are increasingly correlated with postcodes, making fairness imposition at the hot spot level impractical. This makes the fairness constraint best placed at the across hot spots level rather than any lower level of granularity. The next characterizing aspect of FiSH is the identification of multiple possible result sets comprising trade-offs in the noteworthiness-fairness spectrum. This is a conscious decision choice to distribute the agency in determining precise choice within the noteworthiness-fairness trade-off across the technique and the user, rather than being decided solely by the technology. These aspects, we believe, are highly aligned with the practical realities of policing, making FiSH a potentially important step towards fairness in hot spot policing.

7.4 FiSH’s applicability gradient in hot spots policing

FiSH obviously has sub-contexts within hot spots policing where its applicability varies. While teasing out details of the applicability gradient requires on-field studies and correlation with pertinent literature (such studies are outside the scope of this work), we discuss some important considerations herewith. The most widely discussed contexts of policing fairness within the Western world relate to race, a protected attribute that happens to also be geo-correlated (Webber and Burrows 2018), making the across-hotspots fairness conceptualization within FiSH reasonably appropriate. We also note that there are similar cases within other geographies, such as fairness over the caste/religion dimension in India.16 However, in scenarios where fairness over other dimensions (e.g., affluence, education, and class) are more important and the geo-correlation on them may not be as pervasive within them, there may exist an opportunity to use a deeper notion of fairness viz., fairness constraints at the level of each hot spot. The next point where FiSH’s applicability gradient is apparent is to do with the nature and number of protected attribute to ensure fairness over. The current construction of FiSH addresses the categorical attribute case, but is amenable to be adapted to numeric attributes to accommodate protected attributes such as age and income, observed to be pertinent in the relationship between policing and gentrification.17 Towards using a numeric protected attribute, the entire numeric range can be bucketized, so it may be treated as a categorical attribute to leverage FiSH as it is. However, this may require bespoke bucketing mechanisms which are informed by domain knowledge. FiSH, in its present form, can only admit one protected attribute. FiSH may be extended—with non-trivial technical effort—to consider N–F trade-offs across a multi-dimensional space comprising one noteworthiness dimension and several fairness dimensions, one for each protected attribute.

8 Conclusions and future work

In this paper, for the first time to our best knowledge, we considered the task of fair detection of spatial hot spots. In this web era where spatially-anchored digital data is collected extensively, spatial hot spot detection is used extensively to inform substantive policy interventions across a variety of domains, making fairness an important consideration within them. We characterized fairness using the popular notion of statistical parity when computed collectively over k chosen hot spots, and outlined the task of identifying a diverse set of solution candidates along the fairness-noteworthiness Pareto frontier. Observing the computational infeasibility of identifying exact solutions, we developed a method, FiSH, that performs a highly efficient heuristic-driven search to identify good quality approximate solutions for the task. We then formulated a suite of evaluation metrics for the novel task of fair hot spots. We perform an extensive empirical evaluation over a real-world dataset from the human development domain where fairness may be considered indispensable, and illustrated that FiSH delivers high-quality results, and offers good scalability, consistently returning results orders of magnitude faster than what is required to compute exact results. This illustrates the effectiveness of FiSH in achieving fairness in detection of spatial hot spots, and that it offers fast response times, making it appropriate for real-world scenarios. Through a detailed discussion on the context of hot spot policing, we illustrated how FiSH could provide significant strides in deepening fairness within place-based policing.

8.1 Future work

While we have considered enhancing fairness by working upon a ranked list of spatial hot spots, FiSH extends easily to work over techniques that are capable of providing scores (in addition to ranks, which is basically an ordering over the scores) for each hot spot as well; we are considering evaluating FiSH’s effectiveness in working over such scored lists. Our formulation of diverse candidates assumes that the user may be interested equally in all parts of the noteworthiness-fairness trade-off space. However, in several cases, users may have a preference to exclude some parts of the space. For example, the maximum relaxation of noteworthiness may be bounded above in some scenarios. We are considering how user’s trade-off preferences can be factored into the FiSH search process to deliver diverse results within the sub-spectrum of interest.

Declarations

Conflicts of interest

The authors confirm that they do not have any conflicts of interest to declare.
Open AccessThis article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article’s Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://​creativecommons.​org/​licenses/​by/​4.​0/​.

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.
Footnotes
3
We use sensitive and protected interchangeably, in the context of attribute adjectives, throughout this paper.
 
5
While this design is motivated by categorical protected attributes which form the most popular type of protected attributes, this could be extended to other attributes as well.
 
7
They refer only to speedup rates in the paper; however, the KDD presentation has absolute response times.
 
9
100k INR is approximately 1.35k$; India’s per capita income is \(\approx 2k \$\).
 
Literature
go back to reference Abraham S.S, P D, Sundaram S.S (2020) Fairness in clustering with multiple sensitive attributes. In: EDBT, pp 287–298 Abraham S.S, P D, Sundaram S.S (2020) Fairness in clustering with multiple sensitive attributes. In: EDBT, pp 287–298
go back to reference Bera S.K, Chakrabarty D, Flores N, Negahbani M (2019) Fair algorithms for clustering. In: NeurIPS, pp. 4955–4966 Bera S.K, Chakrabarty D, Flores N, Negahbani M (2019) Fair algorithms for clustering. In: NeurIPS, pp. 4955–4966
go back to reference Bhattacharya A, Varambally S, Bedathur A.B.S (2021) Frocc: fast random projection-based one-class classification. SIGKDD Bhattacharya A, Varambally S, Bedathur A.B.S (2021) Frocc: fast random projection-based one-class classification. SIGKDD
go back to reference Binns R (2020) On the apparent conflict between individual and group fairness. In: FAT* Binns R (2020) On the apparent conflict between individual and group fairness. In: FAT*
go back to reference Borzsony S, Kossmann D, Stocker K (2001) The skyline operator. In: ICDE Borzsony S, Kossmann D, Stocker K (2001) The skyline operator. In: ICDE
go back to reference Braga AA, Andresen MA, Lawton B (2017) The law of crime concentration at places: Editors’s introduction. Springer, Berlin Braga AA, Andresen MA, Lawton B (2017) The law of crime concentration at places: Editors’s introduction. Springer, Berlin
go back to reference Breunig M.M, Kriegel H.-P, Ng R.T, Sander J (2000) Lof: identifying density-based local outliers. In: SIGMOD, pp. 93–104 Breunig M.M, Kriegel H.-P, Ng R.T, Sander J (2000) Lof: identifying density-based local outliers. In: SIGMOD, pp. 93–104
go back to reference Chawla S, Sun P (2006) Slom: a new measure for local spatial outliers. Knowl Inf Syst 9(4):412–429CrossRef Chawla S, Sun P (2006) Slom: a new measure for local spatial outliers. Knowl Inf Syst 9(4):412–429CrossRef
go back to reference Chen J, Sathe S, Aggarwal C, Turaga D (2017) Outlier detection with autoencoder ensembles. In: Proceedings of the 2017 SIAM international conference on data mining, pp 90–98. SIAM Chen J, Sathe S, Aggarwal C, Turaga D (2017) Outlier detection with autoencoder ensembles. In: Proceedings of the 2017 SIAM international conference on data mining, pp 90–98. SIAM
go back to reference Chierichetti F, Kumar R, Lattanzi S, Vassilvitskii S (2017) Fair clustering through fairlets. In: NIPS Chierichetti F, Kumar R, Lattanzi S, Vassilvitskii S (2017) Fair clustering through fairlets. In: NIPS
go back to reference Chouldechova A, Roth A (2020) A snapshot of the frontiers of fairness in machine learning. Commun ACM 63(5):82–89CrossRef Chouldechova A, Roth A (2020) A snapshot of the frontiers of fairness in machine learning. Commun ACM 63(5):82–89CrossRef
go back to reference Davidson I, Ravi S (2020) A framework for determining the fairness of outlier detection. In: ECAI Davidson I, Ravi S (2020) A framework for determining the fairness of outlier detection. In: ECAI
go back to reference Deepak P (2016) Anomaly detection for data with spatial attributes. Unsupervised learning algorithms. Springer, Switzerland, pp 1–32 Deepak P (2016) Anomaly detection for data with spatial attributes. Unsupervised learning algorithms. Springer, Switzerland, pp 1–32
go back to reference Deepak P, Abraham S.S (2020) Fair outlier detection. In: WISE Deepak P, Abraham S.S (2020) Fair outlier detection. In: WISE
go back to reference Deepak P, Abraham S.S (2021) Fairlof: fairness in outlier detection. Data Sci Eng J Deepak P, Abraham S.S (2021) Fairlof: fairness in outlier detection. Data Sci Eng J
go back to reference Dwork C, Hardt M, Pitassi T, Reingold O, Zemel R (2012) Fairness through awareness. In: Proceedings of the 3rd innovations in theoretical computer science conference. ITCS ’12, pp 214–226, New York, NY, USA Dwork C, Hardt M, Pitassi T, Reingold O, Zemel R (2012) Fairness through awareness. In: Proceedings of the 3rd innovations in theoretical computer science conference. ITCS ’12, pp 214–226, New York, NY, USA
go back to reference Ensign D, Friedler S.A, Neville S, Scheidegger C, Venkatasubramanian S (2018) Runaway feedback loops in predictive policing. In: Conference on fairness, accountability and transparency, pp 160–171. PMLR Ensign D, Friedler S.A, Neville S, Scheidegger C, Venkatasubramanian S (2018) Runaway feedback loops in predictive policing. In: Conference on fairness, accountability and transparency, pp 160–171. PMLR
go back to reference Fan W, Bouguila N, Ziou D (2011) Unsupervised anomaly intrusion detection via localized Bayesian feature selection. In: ICDM Fan W, Bouguila N, Ziou D (2011) Unsupervised anomaly intrusion detection via localized Bayesian feature selection. In: ICDM
go back to reference Friedman JH, Fisher NI (1999) Bump hunting in high-dimensional data. Stat Comput 9(2):123–143 Friedman JH, Fisher NI (1999) Bump hunting in high-dimensional data. Stat Comput 9(2):123–143
go back to reference Gordon D (2022) Policing the racial divide: urban growth politics and the remaking of segregation. NYU Press, New York Gordon D (2022) Policing the racial divide: urban growth politics and the remaking of segregation. NYU Press, New York
go back to reference Friedman JH, Fisher NI (1999) Bump hunting in high-dimensional data. Stat Comput 9(2):123–143CrossRef Friedman JH, Fisher NI (1999) Bump hunting in high-dimensional data. Stat Comput 9(2):123–143CrossRef
go back to reference Gordon D (2022) Policing the racial divide: urban growth politics and the remaking of segregation. NYU Press, New YorkCrossRef Gordon D (2022) Policing the racial divide: urban growth politics and the remaking of segregation. NYU Press, New YorkCrossRef
go back to reference Greven T (2016) The rise of right-wing populism in Europe and the United States. A comparative perspective, Friedrich Ebert Foundation, Washington DC Greven T (2016) The rise of right-wing populism in Europe and the United States. A comparative perspective, Friedrich Ebert Foundation, Washington DC
go back to reference Knight C (2009) Luck egalitarianism: equality, responsibility, and justice. Edinburgh University Press, Edinburgh Knight C (2009) Luck egalitarianism: equality, responsibility, and justice. Edinburgh University Press, Edinburgh
go back to reference Kearns M, Roth A (2019) The ethical algorithm: the science of socially aware algorithm design. Oxford University Press, Oxford Kearns M, Roth A (2019) The ethical algorithm: the science of socially aware algorithm design. Oxford University Press, Oxford
go back to reference Knight C (2009) Luck egalitarianism: equality, responsibility, and justice. Edinburgh University Press, EdinburghCrossRef Knight C (2009) Luck egalitarianism: equality, responsibility, and justice. Edinburgh University Press, EdinburghCrossRef
go back to reference Knight C (2013) Luck egalitarianism. Philosophy. Compass 8(10):924–934 Knight C (2013) Luck egalitarianism. Philosophy. Compass 8(10):924–934
go back to reference Lai C.-H, Zou D, Lerman G (2020) Robust subspace recovery layer for unsupervised anomaly detection. In: ICLR Lai C.-H, Zou D, Lerman G (2020) Robust subspace recovery layer for unsupervised anomaly detection. In: ICLR
go back to reference Meehan AJ, Ponder MC (2002) Race and place: the ecology of racial profiling African American motorists. Justice Q 19(3):399–430 Meehan AJ, Ponder MC (2002) Race and place: the ecology of racial profiling African American motorists. Justice Q 19(3):399–430
go back to reference Meliani L (2018) Machine learning at predpol: risks, biases, and opportunities for predictive policing. RC TOM Challenge Meliani L (2018) Machine learning at predpol: risks, biases, and opportunities for predictive policing. RC TOM Challenge
go back to reference Meehan AJ, Ponder MC (2002) Race and place: the ecology of racial profiling African American motorists. Justice Q 19(3):399–430CrossRef Meehan AJ, Ponder MC (2002) Race and place: the ecology of racial profiling African American motorists. Justice Q 19(3):399–430CrossRef
go back to reference Miroshnikov A, Kotsiopoulos K, Franks R, Kannan A.R (2020) Wasserstein-based fairness interpretability framework for machine learning models. arXiv preprint arXiv:2011.03156 Miroshnikov A, Kotsiopoulos K, Franks R, Kannan A.R (2020) Wasserstein-based fairness interpretability framework for machine learning models. arXiv preprint arXiv:​2011.​03156
go back to reference Mohler G, Raje R, Carter J, Valasik M, Brantingham J (2018) A penalized likelihood method for balancing accuracy and fairness in predictive policing. In: 2018 IEEE international conference on systems, man, and cybernetics (SMC), pp 2454–2459 . IEEE Mohler G, Raje R, Carter J, Valasik M, Brantingham J (2018) A penalized likelihood method for balancing accuracy and fairness in predictive policing. In: 2018 IEEE international conference on systems, man, and cybernetics (SMC), pp 2454–2459 . IEEE
go back to reference Narayan S (2021) Guilty until proven guilty: policing caste through preventive policing registers in India. J. Extreme Anthropol. 5(1) Narayan S (2021) Guilty until proven guilty: policing caste through preventive policing registers in India. J. Extreme Anthropol. 5(1)
go back to reference Noel P (2007) Why Blacks Fear’America’s Mayor’: reporting police brutality and black activist politics under Rudy Giuliani. iUniverse, Lincoln Noel P (2007) Why Blacks Fear’America’s Mayor’: reporting police brutality and black activist politics under Rudy Giuliani. iUniverse, Lincoln
go back to reference Olfat M, Aswani A (2019) Convex formulations for fair principal component analysis. In: AAAI, vol 33, pp 663–670 Olfat M, Aswani A (2019) Convex formulations for fair principal component analysis. In: AAAI, vol 33, pp 663–670
go back to reference Olfat M, Aswani A (2019) Convex formulations for fair principal component analysis. AAAI 33:663–670CrossRef Olfat M, Aswani A (2019) Convex formulations for fair principal component analysis. AAAI 33:663–670CrossRef
go back to reference Patil GP, Taillie C (2004) Upper level set scan statistic for detecting arbitrarily shaped hotspots. Environ Ecol Stat 11(2):183–197 Patil GP, Taillie C (2004) Upper level set scan statistic for detecting arbitrarily shaped hotspots. Environ Ecol Stat 11(2):183–197
go back to reference Patil GP, Taillie C (2004) Upper level set scan statistic for detecting arbitrarily shaped hotspots. Environ Ecol Stat 11(2):183–197MathSciNetCrossRef Patil GP, Taillie C (2004) Upper level set scan statistic for detecting arbitrarily shaped hotspots. Environ Ecol Stat 11(2):183–197MathSciNetCrossRef
go back to reference Pinchoff J, Chipeta J, Banda GC, Miti S, Shields T, Curriero F, Moss WJ (2015) Spatial clustering of measles cases during endemic (1998–2002) and epidemic (2010) periods in Lusaka. Zambia. BMC Infect Dis 15(1):121CrossRef Pinchoff J, Chipeta J, Banda GC, Miti S, Shields T, Curriero F, Moss WJ (2015) Spatial clustering of measles cases during endemic (1998–2002) and epidemic (2010) periods in Lusaka. Zambia. BMC Infect Dis 15(1):121CrossRef
go back to reference Sherman LW, Weisburd D (1995) General deterrent effects of police patrol in crime “hot spots’’: A randomized, controlled trial. Justice Q 12(4):625–648CrossRef Sherman LW, Weisburd D (1995) General deterrent effects of police patrol in crime “hot spots’’: A randomized, controlled trial. Justice Q 12(4):625–648CrossRef
go back to reference Steinbiss V, Tran B.-H, Ney H (1994) Improvements in beam search. In: Third international conference on spoken language processing Steinbiss V, Tran B.-H, Ney H (1994) Improvements in beam search. In: Third international conference on spoken language processing
go back to reference Telang A, Deepak P, Joshi S, Deshpande P, Rajendran R (2014) Detecting localized homogeneous anomalies over spatio-temporal data. DMKD 28(5-6) Telang A, Deepak P, Joshi S, Deshpande P, Rajendran R (2014) Detecting localized homogeneous anomalies over spatio-temporal data. DMKD 28(5-6)
go back to reference Vallender S (1974) Calculation of the Wasserstein distance between probability distributions on the line. Theory Probab Appl 18(4):784–786CrossRef Vallender S (1974) Calculation of the Wasserstein distance between probability distributions on the line. Theory Probab Appl 18(4):784–786CrossRef
go back to reference Webber R, Burrows R (2018) The predictive postcode: the geodemographic classification of British society. Sage, LondonCrossRef Webber R, Burrows R (2018) The predictive postcode: the geodemographic classification of British society. Sage, LondonCrossRef
go back to reference Weisburd D (2016) Does hot spots policing inevitably lead to unfair and abusive police practices, or can we maximize both fairness and effectiveness in the new proactive policing. U. Chi. Legal F., 661 Weisburd D (2016) Does hot spots policing inevitably lead to unfair and abusive police practices, or can we maximize both fairness and effectiveness in the new proactive policing. U. Chi. Legal F., 661
go back to reference Wilczek J, Monna F, Gabillot M, Navarro N, Rusch L, Chateau C (2015) Unsupervised model-based clustering for typological classification of middle bronze age flanged axes. J Archaeol Sci Rep 3:381–391 Wilczek J, Monna F, Gabillot M, Navarro N, Rusch L, Chateau C (2015) Unsupervised model-based clustering for typological classification of middle bronze age flanged axes. J Archaeol Sci Rep 3:381–391
go back to reference Wilson JQ, Kelling GL (1982) Broken windows. Atl Mon 249(3):29–38 Wilson JQ, Kelling GL (1982) Broken windows. Atl Mon 249(3):29–38
go back to reference Yazdani N, Min P.S (2001) Prefix trees: new efficient data structures for matching strings of different lengths. In: IDEAS Yazdani N, Min P.S (2001) Prefix trees: new efficient data structures for matching strings of different lengths. In: IDEAS
go back to reference Yoon T, Lee J, Lee W (2020) Joint transfer of model knowledge and fairness over domains using Wasserstein distance. IEEE Access 8:123783–123798CrossRef Yoon T, Lee J, Lee W (2020) Joint transfer of model knowledge and fairness over domains using Wasserstein distance. IEEE Access 8:123783–123798CrossRef
go back to reference Yu D, Sheikholeslami G, Zhang A (2002) Findout: finding outliers in very large datasets. Knowl Inf Syst 4(4):387–412CrossRef Yu D, Sheikholeslami G, Zhang A (2002) Findout: finding outliers in very large datasets. Knowl Inf Syst 4(4):387–412CrossRef
go back to reference Zehlike M, Bonchi F, Castillo C, Hajian S, Megahed M, Baeza-Yates R (2017) Fa* ir: A fair top-k ranking algorithm. In: Proceedings of the 2017 ACM on conference on information and knowledge management, pp 1569–1578 Zehlike M, Bonchi F, Castillo C, Hajian S, Megahed M, Baeza-Yates R (2017) Fa* ir: A fair top-k ranking algorithm. In: Proceedings of the 2017 ACM on conference on information and knowledge management, pp 1569–1578
go back to reference Zhang H, Davidson I (2021) Towards fair deep anomaly detection. In: Proceedings of the 2021 ACM conference on fairness, accountability, and transparency, pp. 138–148 Zhang H, Davidson I (2021) Towards fair deep anomaly detection. In: Proceedings of the 2021 ACM conference on fairness, accountability, and transparency, pp. 138–148
Metadata
Title
FiSH: fair spatial hot spots
Authors
Deepak P.
Sowmya S. Sundaram
Publication date
17-11-2022
Publisher
Springer US
Published in
Data Mining and Knowledge Discovery / Issue 4/2023
Print ISSN: 1384-5810
Electronic ISSN: 1573-756X
DOI
https://doi.org/10.1007/s10618-022-00887-4

Other articles of this Issue 4/2023

Data Mining and Knowledge Discovery 4/2023 Go to the issue

Premium Partner