Persistent obstruction theory for a model category of measures with applications to data merging

By Abraham D. Smith, Paul Bendich, and John Harer

Abstract

Collections of measures on compact metric spaces form a model category (“data complexes”), whose morphisms are marginalization integrals. The fibrant objects in this category represent collections of measures in which there is a measure on a product space that marginalizes to any measures on pairs of its factors. The homotopy and homology for this category allow measurement of obstructions to finding measures on larger and larger product spaces. The obstruction theory is compatible with a fibrant filtration built from the Wasserstein distance on measures.

Despite the abstract tools, this is motivated by a widespread problem in data science. Data complexes provide a mathematical foundation for semi-automated data-alignment tools that are common in commercial database software. Practically speaking, the theory shows that database JOIN operations are subject to genuine topological obstructions. Those obstructions can be detected by an obstruction cocycle and can be resolved by moving through a filtration. Thus, any collection of databases has a persistence level, which measures the difficulty of JOINing those databases. Because of its general formulation, this persistent obstruction theory also encompasses multi-modal data fusion problems, some forms of Bayesian inference, and probability couplings.

1. Introduction

We begin this paper with an abstraction of a problem familiar to any large enterprise. Imagine that the branch offices within the enterprise have access to many data sources. The data sources exposed to each office are related and overlapping but non-identical. Each office attempts to merge its own data sources into a cohesive whole, and reports its findings to the head office. This is done by humans, often aided by ad-hoc data-merging software solutions. Presumably, each office does a good job of merging the data that they see. Now, the head office receives these cohesive reports, and must combine them into an overall understanding.

This paper provides a mathematical foundation combining methods from measure theory, simplicial homotopy, obstruction theory, and persistent cohomology (Section 1(a) gives an overview) for semi-automated data-table-alignment tools (e.g, HumMer Reference 14) that are common in commercial database software. Data tables are abstracted as measures over value spaces. The problem of merging tables, or indeed merging previously-merged tables, is recast as the search for a measure that marginalizes correctly.

For example, one data table might record the ages and heights and weights of patients in a hospital, abstracted as a formal sum of point-mass atoms. Another data table might be an actuarial table giving ages and heights and weights for an entire population, abstracted as a smooth distribution where the heights and weights form 2-dimensional elliptical Gaussian distributions for each age and height, the means and singular values varying with age. Both examples would be data tables on the same age-by-height-by-weight space. A third data table might be a simulated probabilistic model of injury severity during vehicle collisions based on height and weight of the passenger. This data table on height-by-weight-by-severity space may or may not merge with each of the previous examples over height-by-weight, within some error. One can imagine many other data tables collected from myriad sources (motor-vehicle records, longitudinal studies, clinical trials) related to this example that may be of interest.

Our first fundamental result (Theorem 3.11) uses this measure-theoretic lens to draw a surprising correspondence between the process of JOIN in database engineering and the Kan extension property for simplicial sets.

This abstraction, and the model-theoretic tools that come with it, permits several advances over the current state of the art, which are collected in our second fundamental result (Theorem 4.13). First, inconsistencies in table collections are automatically detected as obstructions in the sense of Steenrod (i.e, a certain co-cycle is not zero). Second, when inconsistencies are detected, the obstruction theoretic tools, combined with persistent cohomology, provide two potential remedies: a) if algebraically permitted (i.e, a certain co-class is zero), the user may retreat back one level of merging, repair, and proceed; b) else, the user may settle for a measure that only marginalizes approximately correctly, with the degree of possible correctness computed automatically by persistent obstruction theory.

More broadly, we are interested in the following three meta-problems:

Problem 1.1 (Testing problem).

Given several sources of partial information, how do we test that a hypothetical explanation is reasonably consistent with that partial information?

Problem 1.2 (Merging problem).

Given several sources of partial information, how do we merge that partial information into a cohesive whole?

Problem 1.3 (Meta-merging problem).

Given many sources of partial information, and several partial attempts to merge some combinations of them, is there a way to merge these partial merges into a cohesive whole?

By “sources of partial information” we mean, roughly, collected data (databases, spreadsheets, pictures), statistical models, established theories, simulations, and general population trends. In this article, we define a formal mathematical structure—a Data Complex (Section 2)—that can encapsulate a wide range of problems like 1.1, 1.2, and 1.3. A data complex is meant to encode each source of information as a finite measure on an appropriate value space. Whether these measures arise from collected data as in Problem 1.2 or some model/theory/simulation/trend/merger derived from previous work as in Problem 1.3, we call them data tables. By using measures, we combine Problems 1.2 and 1.3 into a single problem.

1(a). Overview of technical approach

Often, formal mathematics looks very different than its intuitive purpose, so we want to make sure the reader understands our intent, absent its formal notation.

We want a mathematically rigorous way to encode and solve Problems 1.1, ?GPMerge, and 1.3. When translated to the language of data complexes, a physically reasonable process for “merge [data tables] into a cohesive whole” can be expressed in terms of four key mathematical ingredients: homological algebra for simplicial sets, simplicial homotopy Reference 10Reference 18, obstruction theory Reference 23, and persistent (co)homology across a filtration Reference 4.

The first ingredient (homological algebra) is used because data tables may overlap partially, meaning that we need a formal simplicial complex to encode all possible intersections of all possible data tables. Moreover, simplicial sets allow data tables with repeated columns. The marginalization integral provides a face map and a boundary operator, and an analogue of the diagonal measure within a product provides the degeneracy map. The face and boundary operators tell us whether one “narrow” data table reproduces the data of another “wider” data table via marginalization integrals. Thus, the question of whether several “narrow” data tables can be merged into a few “wider” data tables becomes the question of whether a -chain is the boundary of a -chain. That is, the ability to merge overlapping partial information sources as in Problem 1.2 is encoded as the homology of a data complex.

The second ingredient (simplicial homotopy) arises because Problems 1.1, 1.2, 1.3 suggest that we want “simple” solutions. Even when partial merging is possible in the sense of homology, it may be that the result is too complicated to be merged further. In the study of geometric bundles, the fundamental group and higher homotopy groups of the fiber play a key role, and we use simplicial homotopy in a similar way here. A simple solution to Problem 1.2/1.3 or a simple hypothesis in Problem 1.1 corresponds to a single data table (as opposed to a complicated chain of many data tables), which is indicated by triviality in the simplicial homotopy group.

An important side effect of introducing simplicial homotopy (via model categories) is that we see that the Kan extension condition means “merging operations are possible.” The process we call merging is similar to JOIN’ing in database software, to fusion in multi-modal data analysis, and to coupling in probability theory. This link reinforces the intuition that data complexes are a good way to encode Problems 1.1/1.2/1.3 for modern data mining when using spreadsheets, DataFrames, and SQL databases. Indeed, our first fundamental result (Theorem 3.11) explicitly formalizes this correspondence.

The reader may be wondering why we introduce something as abstract as simplicial homotopy into something so concrete and common as data merging. Consider the typical database operation

SELECT * FROM table1 INNER JOIN table2
ON table1.column1 = table2.column2
WHERE condition;

When issuing such a command, the administrator must designate two tables to be JOINed and choose specific columns from the two tables to be identified via the ON clause. The SELECT * …; command returns a table, whose columns must appear in some order that is determined by the ordering of attributes in table1 and table2, by their placement in the command, and by the columns in the ON clause. Thus, in the language of Section 2, the database software and the working administrator must agree on a total set of attributes, the attributes in each table, and an ordered attribute inclusion to be used for the ON clause.

This command also indicates why we formalize “data tables” as measures over products of attribute value spaces. Replacing SELECT * with a SELECT columnList corresponds to the ability to re-order the attribute list and to marginalize the output to a sublist of attributes; hence, arbitrary finite products are possible. The optional WHERE condition clause allows one to impose additional restrictions on the values to be considered by imposing logical conditions on the entries, such as WHERE (age > 18 AND height > 200). These conditions allow one to restrict the data table to any⁠Footnote1 measurable subset of the value space. The entries of a WHERE-restricted data table constitute the mass of this measurable set, with respect to the data table. (Finally, for those fluent in SQL subtleties, note that the ability to perform LEFT, RIGHT, and OUTER JOIN instead of INNER JOIN will be encompassed by approximate join and face operations in Section 4.)

1

Any measurable subset—in principle and given a sufficiently generous SQL implementation.

The third ingredient (Steenrod’s obstruction theory as in Reference 22) provides guidance on how to combine homological algebra and homotopy theory to detect and describe any obstructions to an iterative merging process. In its original formulation, obstruction theory asks whether a section of a fiber bundle defined over a topological space can be extended to a section defined over a larger topological space ? The most famous example is the smooth category, where one computes characteristic cohomology classes to indicate whether sections of a bundle can be extended globally. Steenrod studied this problem in the case of fibrations over general topological spaces. Typically, assuming one has some sort of CW structure on , one tries to extend first over the -skeleton of and then the -skeleton of , and so forth. Assuming one has already extended to a section over the -skeleton of , Steenrod’s obstruction cocycle is an element of , where is the homotopy group of the fiber of , as twisted by ; loosely, the co-chain is defined on each -cell of by restricting to the boundary of , but there is some nuance coming from the twisting needed to turn this into a homotopy class of the fiber. If this cocyle is trivial in homotopy, then the section can be extended, and otherwise it cannot. However, if this cocycle is a coboundary, , then there is another section , agreeing with on the -skeleton of , with trivial in homotopy. Hence, obstructions are discovered dimension-by-dimension via homotopy-valued cohomology, and the obstruction computation can often permit the “correction” of initial extensions of the section to avoid higher-dimensional obstructions.

This concept of an obstruction cocycle was introduced by Steenrod in Reference 22 and revisited many times, such as Reference 15 and Reference 11. Its importance motivated early work in category theory. The entire raison d’être of defining fibrant objects and model categories in Reference 10 and Reference 18 was to establish the most general context in which these (co)homology and homotopy calculations remain sensible for more general notions of “weak equivalence.” In particular, one does not require actual topological spaces to perform obstruction theory, merely fibrant objects in a model category.

Here, we establish homology and homotopy theory for data tables by relying on these categorical foundations, giving us an obstruction theory directly analogous to Steenrod’s. When sequential merging is impossible, the obstruction cochain can compute specific data tables that obstruct the process. That is, obstruction theory determines when Problem 1.3 is solvable locally but not globally.

The fourth ingredient, persistent (co)homology, provides a mathematically robust way to measure how much the underlying original data tables would have to be altered, in order to overcome an obstruction. This is a key feature of the theory, because from a practical perspective, multiple information sources are never perfectly consistent. Typos and transpositions and omissions and error bars always exist, and must be accounted for. We use a filtration built from the Wasserstein distance on measures to ensure that the desired simplicial homotopy is possible throughout all levels of the filtration. This allows for a well-defined notion of persistent obstruction theory. Our second fundamental result (Theorem 4.13) formalizes the idea that when inconsistencies are identified, one of two remedies may be available⁠Footnote2

2

In fact, the second is always available, but may be less desirable!

the head office should retreat back one merging level, repair (with repair suggested by algebra), then again seek consensus

the head office should settle for only approximate consensus, where the desired measure only approximately marginalizes correctly, with the degree of approximation computed via persistence.

The ultimate result of this article is a mathematically robust framework for data merging that is reasonably applicable to real-world data. In this framework, Problems 1.1 and 1.2/1.3 become Problems 2.38 and 2.39, which are answered by Theorem 4.13 and Definition 4.14.

1(b). Related work

To the best of our knowledge, this is the first work to combine all of the tools above to build a robust obstruction theory for databases. Other authors have used different aspects of these tools to address databases. For example, recent work by Fong and Spivak uses database schemas and type/value relationships as a motivational example to introduce functors and (co)limits Reference 6, Chapter 3. Specifically Example 3.99 and the chapter’s final remark are somewhat in the same spirit as the approach taken here. Our category of data complexes in Section 2 is similar to the categorical presentation in Reference 20Reference 21, but our data tables are built from measures (not sets) in order to flexibly address the errors that are inevitable in applications. Finally, other recent work by Abramsky, Morton, and collaborators uses obstruction theory (in the sheaf-theoretic context) to detect non-contextuality in quantum theory, with an application to the non-existence of a universal data table that contains a set of given tables Reference 1Reference 2Reference 13. We expect that further interweaving of these measure-theoretic, sheaf-theoretic, and simplicial/categorical perspectives will be fruitful in the future.

1(c). Outline

The rest of this paper is organized as follows. Section 2 defines the basic object of study, a data complex, and draws a mapping between its simplicial set structure and the choices that must be made by any database administrator. Categorical language is alluded to in this section, but a full categorical treatment of data complexes is confined to the Appendix. Section 3 connects simplicial homotopy to the notion of JOIN, and shows how obstruction theory detects the impossibility of merging. Section 4 describes our notion of persistent obstruction theory and its application to the idea of fuzziness of consensus. The paper concludes with discussion of practical considerations for applications in Section 5.

2. Attributes and data tables

This section provides a practical developmental discussion of a Data Subcomplex that should be accessible to a fairly wide mathematical audience, with full categorical language found in the Appendix. The basic definitions appear in Sections 2(a) and 2(b), culminating in Theorem 2.14 which shows that we have indeed defined a simplicial set. Operations that are specifically useful to standard database operations (inclusion/merge/join) are defined in Section 2(c). Then Section 2(d) makes plain the analogue of “section of a bundle,” which permits the rephrasing of our fundamental problems in mathematical language, and Section 2(e) defines the (co)homology of data complexes needed for obstruction theory.

2(a). Data subcomplex as a simplicial set

Our definitions are aimed at making precise the following real-life scenario in data administration.

(1)

The administrator chooses a set of all attributes (column names and variable types) of interest.

(2)

For each attribute in the list , the administrator chooses a space of possible values, and a “reasonable” metric that can provide the distance between any two values in that space. Our notion of “reasonable” includes compactness, which is typically guaranteed by boundedness of realistic integer or vector-valued entries.

(3)

The administrator acquires “data” for some lists of attributes, and attempts to reconcile these into a joint view across all attributes in . The reconciliation process involves “join” operations that could be represented by SQL commands such as

SELECT * FROM table1 INNER JOIN table2
ON table1.column1 = table2.column2
WHERE condition;
(4)

When reconciling, the administrator may choose to alter the data, as long as the alterations are “small” with respect to both the individual values via and with respect to the overall information-theoretic content of the data.

The former two items are choices that must be made. The latter two items are a process to be accomplished. The mathematical structure developed here is informed deeply by the example SQL command, as discussed in Section 1(a).

Let us define our objects. It is convenient to use language of category theory; see Appendix A for our conventions.

Consider a finite set . The elements are called attributes. For each attribute , there is a compact metric space , called the value space.⁠Footnote3 These assumed objects (the finite set of attributes and a compact metric space assigned to each attribute) are user-supplied by a data administrator; after these choices are made, everything else proceeds as defined.

3

These assumptions imply that is complete, separable and is a Radon space. In many applications, the space is finite or a closed interval in , so one needn’t imagine esoteric spaces to grasp the theory.

Each is a Radon space (in particular, a measurable space) using the usual Borel algebra from the metric . These metrics will be used in Section 4 to quantify levels of acceptable imprecision when marginalizing measures.

An attribute list is a finite sequence of attributes; that is, an attribute list is a function . The length of an attribute list is . An attribute list is called nondegenerate if it contains no repetitions; that is if the function is one-to-one. The longest nondegenerate attribute lists are permutations of .

For any attribute list , the product space is well-defined. The product space admits the metric and is measurable via the corresponding tensor-product algebra.⁠Footnote4 For any list representing a permutation of the set , then is the correspondingly ordered total product of all the measurable spaces of all attributes. At the other extreme, we equip the empty attribute list , of length , with the trivial value space as , a singleton set.

4

We use the -metric for ease of proof when studying filtrations. Other -metrics or more general product metrics might carry the whole theory, too, but we have not yet verified this.

Definition 2.1 (Set of attribute lists).

Let denote the set of all attribute lists in . For each , let denote the set of all attribute lists of length . is a small category. Using the notation from Appendix A, an object in is a function . The case , giving the empty list , is allowed. A morphism of attribute lists is given by (an order-preserving function, which is a morphism of as in Appendix A) such that , which is natural for the commutative diagram Equation 2.1.

In Section 2(b) it is shown that for , each is equipped with face maps (by omission of the th element as in Definition 2.7) and degeneracy maps (by repetition of the th element as in Definition 2.11). When omitting the trivial -level, is the simplicial set whose elements are generated by the permutations of via the face and degeneracy maps. Including the trivial -level, is the augmented simplicial set generated this way. See Appendix A for a summary of the standard definition of (augmented) simplicial sets.

For any attribute list , let denote the space of finite measures on . A data table is a pair for for any . Note that , as a measure on the singleton set is determined by the mass of . A trivial data table is any data table of the form where and is a measure on the singleton set . We sometimes abbreviate our notation for data tables from to , because any is equipped with a domain (the measurable sets in ), so is understood in context.

In the first example alluded to in the introduction, we could have = [age, height, weight] with in integer years, in centimeters, and in kilograms, each with the standard metric. The space would be the set of measures on the compact set given by the product. An attribute list is also permissible, and might arise for example if heights were compared from two different sources (driver’s license versus medical chart).

For practical purposes, because is a compact metric space, one might use the Radon–Nikodym theorem to write any using a density function with respect to the uniform⁠Footnote5 probability measure on the compact set; however, for simplicity we use the language and notation of measures instead of the language of functions and integrals.

5

That is, the measure depends only on , for metric balls of sufficiently small radius.

Definition 2.2 (Ambient data complex).

Given , the ambient data complex over is the set of all data tables,

For , let . Let denote the forgetful map .

Theorem 2.14 shows that the ambient data complex is a simplicial set (augmented when including ) with faces given by the marginalization integrals (Definition 2.9) and degeneracies given by Dirac diagonalizations or intersections (Definition 2.13). The ambient data complex is a small category, whose morphisms are generated by faces and degeneracies. The forgetful functor is a simplicial map between the small categories and .

Definition 2.3 (Data subcomplex).

Given an ambient data complex , a Data Subcomplex is a subset/subcategory that is closed under the face and degeneracy maps defined in 2.9 and 2.13. Because is a simplicial map, the attribute base

is a simplicial subset of .

Definition 2.4 (Finitely generated).

A data subcomplex of an ambient is said to be finitely generated iff there is a finite set such that every data table in is obtained from this finite set via face and degeneracy maps. We write or just .

Definition 2.5 (Closed under permutation).

A subset of is said to be closed under permutation iff for any with and for any permutation (that is, bijection)⁠Footnote6 , there exists . A data subcomplex of an ambient is said to be closed under permutation iff for any with and for any permutation , there exists and such that the measure is evaluated on the basis sets of the Borel algebra of by

6

Note that a nontrivial permutation is not a morphism in .

Remark 2.6.

Actual database merging problems encountered in real-life situations such as Problems 1.1, 1.2, and 1.3 always present themselves as Finitely Generated Data Subcomplexes, because there is some finite set of database tables or spreadsheets under consideration. The face and degeneracy maps provide the logical relations between these tables that allow or prevent joining. Real-life situations are also closed under permutation; because, the “SELECT * FROM …” clause in SQL allows the database engineer to re-order the columns of any table. In our earlier example, a data table given by listing patients’ age-by-height-by-weight might be permuted to height-by-weight-by-age simply by reordering the columns of the spreadsheet.

Notational note.

We always use to refer to an ambient data complex. We use either or to refer to a data subcomplex of . We tend to use when we imagine that this data subcomplex came from an actual data merging problem (so it is likely to be finitely generated and closed under permutation); however, we state explicitly these conditions when they are required for a result. When the projection and the attribute simplicial sets are not used in a statement, we omit them and write “a data subcomplex of an ambient .”

2(b). Morphisms of data tables

This section establishes notation for common operations and proves that and are simplicial sets, establishing that they are small categories with morphisms that are well-understood in language of measures.

Definition 2.7 (Face of attribute list).

The face map on attribute lists, , is defined as omission of the th entry in , so

Remark 2.8 (Categorical interpretation).

In Definition 2.7, , where is the co-face monomorphism in , as in Appendix A.

Definition 2.9 (Face of data table).

For a data table with , let be the measure evaluated on the basis sets of the Borel algebra on as

This is the measure obtained by marginalization to omit the th factor, which could also be written as . Let , which is well-defined in .

In our earlier example with individual patients as atomic point-masses, the face map from age-by-height-by-weight to height-by-weight represents deleting the age column of the spreadsheet, and allowing new duplicate entries to add (that is, integrate) measure.

Face maps can be applied multiple times, and the following lemma provides the desired re-ordered “commutation” property. For attribute lists the proof is immediate; for data tables it is the Fubini–Tonelli Theorem applied to the measures.

Lemma 2.10 (Fubini–Tonelli Theorem).

For any , .

Definition 2.11 (Degeneracy of attribute list).

The degeneracy map on attribute lists, , is defined as repetition of the th entry in , so .

Remark 2.12 (Categorical interpretation).

In Definition 2.11, , where is the co-degeneracy epimorphism in , as in Appendix A.

Definition 2.13 (Degeneracy of data table).

For a data table , let be the measure evaluated on the basis sets of the Borel algebra on as

Then, is well-defined in .

If the measure is expressed as a density function via the Radon–Nikodym theorem, then this is the Dirac-delta

Theorem 2.14 (Simplicial sets).

Let be the ambient data complex over an attribute set . For any , consider the face maps and degeneracy maps as in the definitions above. Then

(1)

, if ;

(2)

, if ;

(3)

;

(4)

, if ; and

(5)

, if .

Moreover, the forgetful map commutes with and . That is, and are augmented simplicial sets as in Lemma A.4. They are simplicial sets when omitting the trivial and . Reference 7, Definition 3.2, Reference 9, Equation (1.3), Reference 12, Definition 1.1

Proof.

This is direct with no surprises, by working on the Borel basis sets for . The condition was already seen as Fubini–Tonelli.

2(c). Inclusions, merges, and joins

We now establish⁠Footnote7 additional operations (inclusion, sum, merge, join) that are special to and and do not apply to general simplicial sets.

7

We are particularly indebted to Tony Falcone for technical discussions that motivated the formalism in this subsection.

Definition 2.15 (Attribute inclusions).

An attribute inclusion

is given by a map such that

(1)

(2)

(3)

is one-to-one (implying ),

Although itself is a map of index sets, we use the compatibility property to overload notation and write .

Remark 2.16 (Categorical interpretation).

An attribute inclusion is a morphism in the category such that where is a monomorphism in . We overload notation (that is, omit the upper-star) and write . The functor is contravariant, so attribute inclusions are actually epimorphisms in ; however, it is reasonable to call them “inclusions” because the -ordered multiset is an ordered subset of the -ordered multiset . One could avoid this overloaded notation by working in the opposite category, but we decline to add another layer of notation since the meaning is always clear in context.

Example 2.17.

Consider . Example attribute lists⁠Footnote8 are and . There are 20 possible inclusions , which are obtained by choosing the ordered image of the ’s and ’s. One possible inclusion is

8

The fact that these lists are in alphabetical order is merely aesthetic, and is not required in the definition of an attribute list.

which can be summarized as . We can abbreviate this by decorating the entries in that are included from ,

Lemma 2.18 (Quotient inclusion).

For any attribute inclusion , there is an attribute list (called the quotient) that enumerates the entries of that are not in the image of . This enumeration equips the quotient with an attribute inclusion , and corresponds to the complimentary monomorphism from Lemma A.1.

Example 2.19.

Consider the earlier example of an attribute inclusion. The quotient list is . The quotient inclusion is

The next lemma and corollary make clear that face maps and attribute inclusions are related tightly.

Lemma 2.20.

Any face map in is equipped with an attribute inclusion defined by the index function that skips , namely the co-face monomorphism in . Its quotient inclusion is the index function by that gives .

Corollary 2.21.

For any attribute inclusion , there is a sequence⁠Footnote9 of face maps such that and such that the attribute inclusion induced by the sequence of face maps is . Moreover, any permutation of this sequence obtained by re-indexing the face-maps according to Lemma 2.10 is equivalent. If , then is the index function for , the quotient inclusion.

9

The backwards ordering here is intentional. Because of the indexing situation and Lemma 2.10, it is simpler to remove attributes from the end. To remove an entire list, one could write or , because is like “pop” on the front of the list.

Remark 2.22.

In light of Theorem 2.14, Corollary 2.21 is a partial version of Lemma A.4, which says face maps and degeneracy maps generate all the morphisms in a simplicial set. This is because the co-face and co-degeneracy maps in generate all order-preserving maps.

Attribute inclusions provide surjections on value spaces and measures, according to the following “contravariant” definition.

Definition 2.23 (Reduction).

Consider an attribute inclusion , where and . Write an element of as where and write an element of as where . Define the surjective function by

Similarly, define the surjective function by sequential application of face maps according to the previous corollary: For any , let

That is, is the measure on obtained by marginalizing to remove the factors specified by .

When the attribute inclusion is understood from context, we abuse notation and write instead of . Note that , so we use this notation as shorthand for “the total integral of a measure.”

Definition 2.24 (Sum of attribute lists).

Given attribute lists and in , define as the attribute list obtained by concatenating and .

Note that and are related by a permutation, which (excepting the identity permutation) does not correspond to a morphism in the categories or . The concatenation process provides specific attribute inclusions and . More generally, for attribute inclusion as in Lemma 2.18, it is true that and are related by a permutation; because, the concatenation provides inclusions and that may not be the original and . On the other hand, for any sum , it is true that is the quotient of by the concatenation-induced inclusion of , and vice-versa.

Definition 2.24 implies and are well-defined. But, beware of multivariable calculus: , as not every measure on a product space is an elementary product of measures!

Example 2.25.

Consider and . Then their sum is . The concatenation is equipped with inclusions

Definition 2.26 (Permutation notation).

Suppose that , , and are attribute lists such that for a permutation . Then and are complimentary attribute inclusions. If the permutation or attribute inclusions are well-known in context, then for any subsets and , let denote the subset for which elements and correspond with respect to the -permuted indices.

Because Lemma A.2 provides an ordered form of the inclusion–exclusion principle, we can define an indexed form of the inclusion–exclusion principle.

Definition 2.27 (Merge of attribute lists).

Suppose , and that and are attribute inclusions. Define as the attribute list obtained by performing the index merge specified by Figure1 as in Lemma A.2; this merge concatenates sublists spliced between the entries aligned by . Writing for , Diagram Equation A.1 becomes a diagram of attribute inclusions.

Note that the choice of ordering in Definition 2.27 and Figure 1 is partially arbitrary. In particular, one may draw an equivalent diagram with any choice of interleaving pattern, as long as the entries remain fixed. However, this choice is irrelevant, as the theory developed in Section 3 will encompass all allowable permutations. Regarding the permutation notation introduced earlier, for any Borel sets , , and , we write for the appropriately permuted copy of the set in , since the definition and algorithm give a well-defined permutation. This notation is required in Theorem 3.11 and elsewhere.

Example 2.28.

Compare this to Example A.3. Consider . Consider attribute lists and , and with the attribute inclusions and . Visually, the merged indexing means

Our choice of ordering in provides that the trivial merge

is the sum from Definition 2.24.

As with Definition 2.24, the attribute list is well-defined regardless of the preference of versus and regardless of the indices specified by and . This list is identical to the list obtained by constructing the sum then applying face maps to remove the image of for each indexing . But, again beware that the partitioned merge-sort construction equips with specific attribute inclusions and such that the composed attribute inclusion is well-defined through both compositions. In general, these inclusions are not the same as the inclusions obtained through the “sum and face” construction.

Lemma 2.29 (Decomposition of merged lists).

Suppose , and that and are attribute inclusions. Let denote . Let and denote the complements of these inclusions, so and .

Then is partitioned by the inclusions , , and . That is, is a permutation of , with ordering of projections determined by the inclusion and quotient maps, , , and .

2(d). Data sections

Because the forgetful map acts like a projection, it allows a notion of section.

Definition 2.30 (Data section).

Consider a data subcomplex of an ambient . A data section is a natural⁠Footnote10 map such that .

10

Natural means that it respects the face and degeneracy maps, as in Equation A.2 and Lemma A.4.

Remark 2.31.

In Section 4, data sections will be specified as , on a single level of the simplicial-set grading, where the other levels are inferred by the face and degeneracy maps. This omits all nondegenerate elements of level , so is interpreted as a section on the -skeleton.

The following definition captures a condition describing data subcomplexes that are “as compatible as possible.”

Definition 2.32 (Well-aligned).

A data subcomplex of an ambient is called well-aligned if: for all and all with attribute inclusions and , there exists with

The next lemma shows that well-aligned data subcomplexes in this theory play the role of “submanifolds transverse to the fiber” from classical bundle theory and of “holonomic submanifolds” in geometric PDE theory. That is, they represent local sections.

Lemma 2.33.

Suppose that is a data subcomplex of an ambient data complex such that contains a nontrivial data table. The following are equivalent:

(1)

is well-aligned.

(2)

There is a data section such that .

(3)

is a simple cover via the isomorphism .

Proof.

(2) implies (1): Note that well-alignedness is implied by the commutation of with the face maps.

(1) implies (2): The case of implies that all data tables in have the same mass, , which is non-zero since contains at least one non-trivial table. The case of implies that each admits exactly one .

It is immediate that (2) and (3) are equivalent.

Remark 2.34.

A database engineer would appreciate a database system that could be described as a well-aligned data subcomplex, because for each list of columns present within any combination of the given tables, there is only one possible table; that is, for each there is exactly one . Compare the well-aligned condition to the space of joins, Definition 3.5. Note too that well-aligned implies finitely generated. (Not every finitely-generated data subcomplex is well-aligned, as it could have multiple data tables over the same attribute lists.) Moreover, if is well-aligned (and contains a nontrivial data table), then all data tables can be re-scaled by their shared mass to yield probability measures.

In our definitions above, the set of attributes is finite, and each level of the simplicial set is finite. Therefore, for any simplicial subset , we can consider the finite graph whose vertices are the 0-cells of (singleton attribute lists) and whose edges are the 1-cells (including loops, as degenerate 1-cells like ).

Definition 2.35 (Connected).

Suppose that is a data subcomplex of an ambient data complex such that contains a nontrivial data table. The simplicial set of attributes is called connected if the finite graph with vertices and edges is connected.

Definition 2.36 (Path-connected).

A data subcomplex of an ambient is called path-connected if for any attributes in , there is a sequence in such that and and for all there is an attribute list equipped with inclusions and such that

Lemma 2.37.

Suppose is connected as a simplicial set. If is well-aligned, then is path-connected.

With the language of simplicial sets, we can now re-state our original motivating questions. The remaining sections of this document construct a precise way to answer these questions, and ensure that the notion of “distance” is well-defined. An appropriate notion of distance appears in Definition 4.1. When all the definitions and lemmas are in place, these problems are answered by the Obstruction Cocycle in Definition 4.8.

Problem 2.38 (Testing problem, bis).

Consider a data subcomplex of an ambient . Given a data section of the form for , is it true that lies entirely within ? If not, what is the distance from to in ?

Problem 2.39 (Merging and meta-merging problems, bis).

Consider a data subcomplex of an ambient . Suppose that there is a simplicial map of the form . Does there exist an extension of , meaning for all such that ? If not, what is the minimal distance that would allow an approximate extension?

2(e). Homology

We use the traditional definition of chains, summarized here to fix notation. Fix⁠Footnote11 a ring . For , a -chain is a formal linear combination

11

For practical reasons we use in applications; however, chains are sensible for any ring.

where negative coefficients indicate formally reversed orientation. We define -graded chains as elements of the 1-dimensional -module, . For , define the usual simplicial boundary operator , as

It is immediate that satisfies , so homology is well-defined.

For , a -chain is a formal linear combination

where negative coefficients indicate formally reversed orientation. We can also define -graded chains as , the -module generated by . Moreover, for any , note that and are formally distinct unless ; hence, the graded module is very large, especially if is infinite for any . For , define the usual simplicial boundary operator , as

The next lemma is easy, but important; it means the usual notions of cycle/closed and boundary/exact apply to chains in .

Lemma 2.40.

satisfies . In particular, the homology is well-defined.

Proof.

Suppose that and that . Recall that is a product of the attributes’ measurable spaces, and by our definitions, the measure is finite, therefore -finite, so the Fubini–Tonelli theorem holds. In particular, Corollary 2.21 shows that the reduction is symmetric, so because the double-sum is alternating, all terms will cancel.

For example, suppose on , where the index shows which variables are still free. Then

Note that it is irrelevant in this proof whether is a degenerate attribute list, as repeated attribute value spaces are treated as distinct factors for the sake of integration.

Define the projection by and extending by linearity.

Lemma 2.41.

Proof.

Suppose that and that . Then

Corollary 2.42.

The induced homomorphism is well-defined.

Similarly, the chain modules and homology are well-defined for any data subcomplex of an ambient .

Corollary 2.43.

If is well-aligned, then there are canonical isomorphisms and induced by .

We are particularly interested in the case , so that a chain (respectively ) is interpreted as a set of attribute lists (respectively, data tables), without any consideration for multiplicity or orientation. It is therefore sensible to apply the condition well-aligned to a chain , so that a well-aligned chain in can be interpreted equivalently to a section , where is the data subcomplex generated by the elements of .

3. Homotopy as joins

In the previous section, we established that a data complex is equipped with simplicial homology, and framed data complexes as simplicial sets. This section contains several payoffs for that effort. First, Section 3(a) builds to Theorem 3.11, our first key result, which shows that the simplicial set language enables a connection between our framework and standard database engineering; later results show that the framework enables further insights into data merging problems that transcend standard database engineering. Then, Section 3(b) explores the simplicial homotopy of data complexes and reframes Problems 2.38 and 2.39 in the language of obstruction theory for simplicial sets, as in Reference 7Reference 9Reference 10Reference 12Reference 18.

3(a). Database joins and the kan condition

Recall these three standard definitions from the well-established theory of simplicial sets, as in Appendix A and Reference 7Reference 9Reference 10Reference 12Reference 18.

Definition 3.1 (Simplex).

The standard -simplex is the simplicial set generated (via face and degeneracy maps) by the ordered set in the simplex category .

Definition 3.2 (Horn).

The th horn of the -simplex is the simplicial subset generated by the union of all the faces of except the th face.

As is standard in the literature, we abuse notation slightly by referring to both (which is an infinite collection of sets) and (which is the generator of that collection) as “an -simplex in .”

Note! The (categorical) -simplex is not the same as the (topological) -simplex . The former is an infinite set of formal objects in the simplex category; it has no notion of “interior” or “continuity.” The latter is a compact topological space obtained defined via convex linear combinations in . There is a relationship between their respective categories called realization, as discussed in Reference 18, §3 and Reference 9, Chap I.2.

Remark 3.3 (Data tables as simplices).

A data subcomplex in an ambient is an (augmented) simplicial set by Theorem 2.14. Thus, a data table with can be seen as (the generator of) an -simplex, which includes its faces and degeneracies , and so-on. The “vertices” are (generated by) the single-attribute data tables obtained by applying sequences of face maps. For , the -simplices within are (generated by) the data tables obtained by applying sequences of face maps and degeneracy maps until the result has attributes. The picture of “two simplices that share a boundary component” is realized in as a pair of data tables and and attribute inclusions and such that there is a data table with . If and , then this information generates a -horn. A completion of the -horn to a -simplex would be (generated by) a data table that has and as two of its three faces. Depending on the available simplices in , it may or may not be possible to find such .

Definition 3.4 (Kan condition).

A simplicial set is said to satisfy the Kan condition iff any map from a horn extends to a compatible map from the simplex, .

The Kan condition means that the simplicial set is closed under simplicial deformation, so it has a well-defined homotopy group. The Kan condition is not specific to data complexes; it is a definition for general simplicial sets, and gives the appropriate notion of fibrant for many model categories. For our purposes, we require a slight variation on the Kan condition to provide an adequate notion of fibrant data contexts, which we now develop as Definition 3.13.

We define the space of joins for two data tables with designated attribute inclusions.

Definition 3.5 (Space of joins).

Suppose attribute lists are equipped with attribute inclusions and , and let denote the attribute list as in Definition 2.27, which is equipped with inclusions and . For any data tables , let

Note that requires , as must be well-defined.

Similarly to Definition 2.27, we write this set as for notational convenience when the attribute inclusions are understood from context.

Note! This is not the same notion of “join” that one sees in traditional topology, or in categorical references such as Reference 5Reference 19 and https://ncatlab.org/nlab/show/join+of+simplicial+sets. It is not yet clear whether there is a useful relationship to joins in ergodic theory Reference 8. We choose the term “join” to mimic the terminology in database engineering discussed in Section 1. Definition 3.5 reminds one of couplings from statistics, as in Reference 3; however, the generality here allows repetition and distinct measures and overlaps.

Definition 3.6 (Join conditions).

A data subcomplex of an ambient is said to satisfy the weak join condition iff, for any and with attribute inclusions and and , then is nonempty. It satisfies the strong join condition iff .

The weak join condition means that the simplicial set admits some database JOIN operation between any well-aligned pair of data tables. The strong join condition requires that all possible joins exist in .

The trivial join (when ) is of particular interest as it provides a generalization of independent products of measures.

Definition 3.7 (Admission of trivial joins).

A data subcomplex of an ambient is said to admit trivial joins iff, for every with , there exists some such that and .

Definition 3.8 (Closure under independent products).

A data subcomplex of an ambient is said to be closed under independent products iff, for every and , we have .

Remark 3.9.

The independent product is an example of a trivial join. If a data subcomplex is closed under independent products, then it also includes all IID measures built from its various data tables; this property is important for applications to statistics.

Lemma 3.10.

If a data subcomplex of an ambient satisfies the strong join condition, then is closed under permutations.

Proof.

Because is finite, it suffices to prove that is closed under permutations that are swaps (that is, transpositions or 2-cycles). Moreover, it suffices to consider only swaps of adjacent entries, as any swap can be written by migration of past , then to the original position of .

Suppose that with with . Then, consider and , and . By construction, there are well-defined attribute inclusions and that satisfy . Note that , and . Consider the attribute list obtained by swapping the adjacent entries and , and let denote the correspondingly permuted measure obtained from . Note that , and . By the strong join condition, .

Theorem 3.11 (Fundamental theorem of data complexes).

For any data subcomplex of an ambient .

(1)

If satisfies the strong join condition, then admits trivial joins and satisfies the Kan condition as a simplicial set.

(2)

If admits trivial joins and satisfies the Kan condition, then satisfies the weak join condition.

(We would not be surprised if the Kan condition and the strong join condition are equivalent, under some reasonable assumptions, but we have not pursued that claim.)

Proof of .

Suppose that satisfies the Kan condition and admits trivial joins. Admission of trivial joins provides the weak join condition in the case . Suppose that and are elements of . Suppose that there are inclusions and for some , and suppose that . Each of and and provides all faces of all lower dimensions. We prove the existence of by induction on the dimension. For simplicity, we use the language of simplicial sets, instead of the language of measures. Recall that a “vertex” is a data table obtained by marginalizing to a single attribute, and an -simplex is a data table obtained by marginalizing to attributes, as in Remark 3.3. Fix a preferred vertex in . For any vertex in and any vertex in , the 1-simplices and exist a priori (up to notational ordering). This is an example of a horn . Therefore, by the Kan condition, the 2-simplex exists in . Hence, every 2-face including and vertices in or exists in . Assume for induction that every -face containing vertex exists. Any of those -faces form a horn , so their -face exists in . So, every -face containing vertex exists in . Therefore, there is a Data Table that involves all vertices in and .

Proof of .

We prove part (1) under the notable assumption that is a compact metric space for all . Hence, for any attribute list , the space of measures includes a uniform⁠Footnote12 probability measure .

12

That is, depends only on , for metric balls of sufficiently small radius.

Suppose that satisfies the strong join condition. The case implies admission of trivial joins.

In this proof, we assume that is the common vertex in a horn , but that is only for notational simplicity; the proof certainly applies for any other specified vertex , by appropriate re-ordering. Consider data tables giving a horn . These data tables are of the form for , where . Let . As a horn, these data tables are well-aligned; that is, they match on all corresponding faces according to for as noted after Definition A.7. In particular, all these data tables share the same total mass, . To establish the Kan condition, we construct a compatible -simplex; that is, a data table such that for .

A brief outline of the argument: First, we construct a data table such that and . The measure is built from a parameterized family of partial measures on by recursively bifurcating the parameter set into dyadic sets, which allows to be defined via countable disjoint unions. This data table serves as the base case for an inductive argument for a sequence of partial solutions such that has the desired faces through . Finally, is the desired -simplex. Let us proceed.

For , a compact set, consider the measures

Consider any trivial join

that is, and . Note that

Fix any open such⁠Footnote13 that . Note that the measures

13

Of course the value of is not special, but aesthetic. Any will do.

can be joined to provide

That is, and . Note that

Further, by their definitions via trivial joins from , one can choose to guarantee that for all Borel sets . In particular, . Likewise, for the closed set , define , which is also a measure in by construction. Note that both the closure and the complement are closed in , therefore both are compact. Replacing with or in Equation 3.2 means that we can establish measures for a countable bifurcating collection of open sets; any measurable set in can be -almost covered by disjoint sets in the collection. By analogy, we refer to the as a dyadic collection.

Given such a countable collection , define a measure on Borel sets by disjoint -additivity,

By construction, and . Therefore, by the strong join condition, . The data table provides the base case for induction on faces.

Assume for induction that for some satisfying there exists a data table such that for all . Denote the “error” of the face as

The error is a signed measure—not a measure—on , but the face operation of marginalization is still sensible. Then for ,

Also, for application below, consider the pre-measure on Borel sets as defined by

Observe the inequality

which follows because for all Borel sets satisfying , we have

Let be a probability measure satisfying the condition

for all Borel . Such probability measures are guaranteed to exist by Equation 3.10.

Then, define for any⁠Footnote14 Borel ,

14

Not necessarily meeting the condition above.

and extend by additivity. By construction, is additive and zero-null. Non-negativity follows from Equation 3.12 and the definition of in Equation 3.9; therefore, is a measure on . Moreover, satisfies the desired marginalizations, shown here:

And, for , the properties Equation 3.8 apply to give

Therefore, is a data table that has the desired faces through . The inductive step is established. The ultimate data table provides the -simplex completing .

Remark 3.12 (Freedom).

The flexibility in choosing arises from an initial parametric choice of joined measures and a finite set of probability measures , , …, .

3(b). Simplicial homotopy of data complexes

Definition 3.13 (Fibrant).

A data complex satisfying the strong join condition is called fibrant.

See Appendix A for a categorical version of this definition. The entire raison d’être of fibrant objects is that they admit homotopy, as proven by Reference 10 and Reference 18, which allows obstruction theory to be studied in direct analogy to Steenrod. In the category of simplicial sets, the term fibrant refers only to the Kan extension condition. Our practical desire to use joins as a weak-equivalence compels us to require the strong join condition. By Theorem 3.11(1) the traditional definition and all of its consequences are implied.

Corollary 3.14.

Suppose a data subcomplex of an ambient is fibrant, and fix a basepoint data table . The homotopy group is well-defined for all , and satisfies the typical properties of homotopy categories over model categories.

Theorem 3.15.

For any attribute set and value spaces , the ambient data complex is fibrant.

Proof.

The very definition of an ambient is that it includes all finite measures over the relevant metric spaces, so it includes the set in particular.

We now want to explore how a data subcomplex of an ambient interacts with any other attribute list . The following sets are of interest.

Definition 3.16.

Let , the set of data tables in the data subcomplex that are detected by . Let , the set of attribute lists in the subcomplex that are detected by .

A data subcomplex may not be fibrant, so we define a convenient fibrant space that contains it. The notation is meant to be suggestive; in Section 4, a larger filtration of simplicial sets is created (Definition 4.5) by turning the equality in the definition below into an inequality involving Wasserstein distance.

Definition 3.17 (Complex of perfect joins).

For any data subcomplex of an ambient , let denote the subset of defined by

if and only if , such that and .

Note: the quantifier refers to each entry in the attribute list, which means repeated entries must have corresponding measures.

The definition of is a convenient way to say “consider everything that can be generated from using ,” as justified by the following lemma. Similarly, the upcoming Definition 4.5 of gives a convenient way of saying “consider everything that can be approximated to an acceptable level of uncertainty from using .”

Lemma 3.18.

if and only if there is a sequence such that

, and

for some , and

.

Proof.

Suppose . Let denote the first attribute of . By the definition of , there exists with an attribute inclusion such that and such that is in the image of . Let . By reducing if necessary, we may ensure that is contiguous within . If , then the sequence is complete.

Otherwise, there exists some first attribute in . By the definition of , there exists with an attribute inclusion such that and such that is in the image of . By reducing if necessary, we may ensure that is contiguous within , and that is also contiguous. With these reductions, the orderings are consistent such that is equipped with a list inclusion . Because is given, let , which by construction is an element of . Repeat this process until all elements of are in the image of some inclusion .

For the converse, note that each is included in some , which is sufficient.

Corollary 3.19.

includes all independent products formed from data tables in .

Lemma 3.20.

For any data subcomplex of an ambient , the complex of perfect joins is fibrant.

Proof.

The data subcomplex is closed under face maps and degeneracy maps, so application of those maps to all in the definition shows that is closed under the face maps and degeneracy maps as well. To verify that is fibrant, suppose that is any join of and in . Because every appears in or , the existence of is inherited from and .

We conclude this section by tying simplicial homotopy theory to Problem 2.39.

Lemma 3.21.

Suppose is a fibrant data subcomplex of an ambient . A basepoint-preserving simplicial map defines a class in . Moreover, if and only if admits an extension .

Proof.

The first claim reduces to Lemma 9.6 in Reference 7. The second claim reduces to Lemma 7.4 in Reference 9. Our definition of fibrant implies path-connectedness, so a spanning tree can be used for locality such as in Reference 10.

Corollary 3.22.

Suppose that is a data subcomplex of an ambient such that for some . Fix a simplicial section . The following are equivalent (omitting basepoints for brevity).

(1)

For every composition

we have .

(2)

admits an extension of the form .

Proof.

Because , the boundary of every -simplex in appears in . Apply the previous lemma for each as a map for .

This corollary is revisited as Lemma 4.9. The corollary fails when no such extension can be found. Then, the question remains: how to measure the failure of this corollary? That measurement is the purpose of filtered obstruction theory.

4. Filtrations and obstructions

This section concludes the theoretical framework outlined in Section 1(a). Together, obstructions and filtrations allow us to detect when merging is possible; if merging appears obstructed, we can determine whether merging can be achieved by reverting a previous merge or by altering some of the data tables. Section 4(a) introduces a filtration from a data subcomplex to its ambient using the Wasserstein distance. Each level of the filtration is fibrant, which allows one to define an obstruction cocycle (Section 4(b)) at each level of the filtration. Eventually, for a high enough level in the filtration, the obstruction cocycle becomes trivial, so the importance of the obstruction cocycle can be measured using topological persistence. This statement is formalized in Theorem 4.13, which can be seen as the main payoff of our theoretical development in terms of database engineering. As promised in the introduction, the theory of data complexes does not just mathematize the notion of table merging; rather, it provides further powerful operations when traditional merging is impossible.

4(a). Filtrations from data subcomplexes

A general notion of persistence on simplicial sets appears in Reference 16. In summary, a fibrant filtration of simplicial sets is a bi-graded collection of sets for and equipped with maps and such that

(1)

is a simplicial set for each ,

(2)

for all , and

(3)

is fibrant for each .

The fibrant condition implies that is well-defined for all , and the inclusion maps induce maps on homotopy, .

We now define a specific filtration for a data subcomplex that is designed to meet our application regarding joining data tables. Recall that is a Radon space for each attribute .

Definition 4.1 (Wasserstein distance).

For any with , and , let

The reductions and refer to the two copies of the attribute .

For any and , let

The reductions and refer to the two interwoven copies of the attribute list .

Remark 4.2.

Recall that , the -metric obtained from the individual attribute metrics. Also, in the special case that , the infimum argument lies in the space of trivial joins, , so the Wasserstein distance is tied to our notion of fibrant data complexes.

Lemma 4.3.

Suppose that . If , then .

Proof.

Let indicate the th attribute of , and write with inclusion and quotient inclusion . Then write and . For any such that and , let be the reduction of obtained by applying both copies of . Also, we use the notational convention for , , , so the metric gives .

Therefore, the infimum defining cannot be greater than the infimum defining .

Lemma 4.4.

Suppose that . If , then .

Proof.

Let indicate the th attribute of . Let , equipped with the degeneracy inclusion and its quotient . Then write and . Now, for any such that and , let be the degeneracy of obtained by applying both copies of . Consider the integral . Note that the distributional form of the degeneracy is a delta,

Moreover, if , for some , then . Together, these give . Therefore, the infimum defining cannot be greater than the infimum defining .

Now we produce a particular fibrant filtration for a data subcomplex.

Definition 4.5 (The complex of approximate joins).

Let be a data subcomplex of an ambient . For any , let

Note that the case reproduces the complex of perfect joins, . Note also that .

Theorem 4.6.

For each , is a fibrant data subcomplex of .

The proof is identical to the proof of Lemma 3.20, replacing the equality with an inequality.

Proof.

Recall that the data complex is closed under face maps and degeneracy maps. Note the face and degeneracy bounds for the Wasserstein distance given above. Application of those maps to the and in the definition shows that is closed under the face maps and degeneracy maps as well. Therefore, is a data subcomplex.

To verify that is fibrant, apply Theorem 3.15 to obtain all joins from any and in . We must show such lies in . Fix . Because every , it appears in or . For concreteness, assume . There is some such that . By the construction of , we have , so . Hence, .

Because is fibrant, all of the usual consequences apply in homotopical algebra, such as

Corollary 4.7.

Fix a data subcomplex of an ambient . For each , and for each , the pointed homotopy group is well-defined. Moreover, for , the inclusion of data subcomplexes induces a homomorphism of pointed homotopy groups .

4(b). Persistent obstruction theory for data subcomplexes

Because we have established fibrant objects with resulting homotopy and homology, we are equipped to extend obstruction theory to our application. Although our category is not classical, the next several results are modeled on the classical work summarized in Section 6 of Reference 24, Section 34 of Reference 22, Section 4 of Reference 11, and Reference 15. The discussion culminates in Definition 4.8 and Theorem 4.13.

Definition 4.8 (Obstruction cocycle).

Let be the filtration of a path-connected data complex. Fix a dimension such that for all faces of all . Let be a data section. For a fixed basepoint , define

to be the element of that is represented by the loop corresponding⁠Footnote15 to the cycle for any . Extend by linearity for . We typically omit the basepoint and ring for brevity, so .

15

The well-definedness of this loop is implied by our assumption .

Lemma 4.9.

Fix . If , then there exists such that the diagram commutes

Lemma 4.10.

The cochain is a cocycle. So, it defines a cohomology class .

Proof.

For any , we have , but then the trivial cycle represents the trivial class .

Remark 4.11.

Obstructions in dimension detect loops in , which will prevent some data tables from being mutually joinable.

Obstructions in dimension detect spheres in , which will prevent some data tables from being mutually joinable.

Obstructions in dimension detect non-path-connectedness of , which would prevent some data tables from being joinable (but this is impossible with our definitions including trivial joins).

The next theorem is an adaptation of Theorem 34.6 and Corollary 34.7 in Reference 22, which is summarized in Theorem 4.5 of Reference 11. It relies on defining a difference cochain that compares a homology class of sections.

Theorem 4.12.

Fix a data section . Suppose for some . Then there exists a data section such that . The converse holds as well.

Theorem 4.13 restates Theorems 4.9 and 4.12 in practical language.

Theorem 4.13 (Steenrod’s trichotomy).

Fix a data subcomplex of an ambient , with Wasserstein filtration . Exactly one of the following is true.

(1)

as a cocycle. Every -cycle of data tables in over a total of attributes can be approximately joined to a single data table over those attributes, allowing error at-most in any reduction to the original data.

(2)

as a cocycle, but as a cohomology class. There is some -cycle of data tables in such that the combined attribute list does not admit an approximate join with error at-most . However, if one considers all of the faces of these data tables, then there is an approximate join to of error at-most .

(3)

as a cohomology class. There is some -cycle of data tables in such that the combined attribute list does not admit an approximate join with error at-most , even when omitting attributes from the original data tables. The only way to produce a single joined table is to increase the error threshold .

Definition 4.14 (Persistence of obstruction).

Let be the filtration of a path-connected data complex. Fix a dimension such that for all faces of all . Let be a data section. Fix a basepoint . Define

and

Note that .

Remark 4.15.

Consider a data section . A specific value means that admits an extension into , but not for any level of the filtration less than . In other words, there is no obstruction to extension beyond the mere existence of the data section . Similarly, by Theorem 4.13, a specific value means that there is no obstruction to extension beyond the mere existence of the data section .

Remark 4.16.

When obstructions are resolved, there are typically many solutions to Problems 1.2/1.3. That is, if any hypothesis is consistent in 1.1, then there are typically many other hypotheses that are consistent as well. Typical methods for choosing among them often involve posing and then solving some optimization problem. We might propose enriching those optimization problems via inclusion of a measure of global inconsistency. More precisely, the cost of a proposed data section might be some combination of a local cost and some decreasing function of or ; in other words, one might penalize proposed local mergers based on the degree of difficulty they cause in forming global consensus with other local mergers.

5. Discussion

This paper provides a mathematical foundation for semi-automated data-table-alignment tools that are common in commercial database software. Data tables are abstracted as measures over value spaces, and the problem of merging tables, or indeed merging previously-merged tables, is recast as the search for a measure that marginalizes correctly. This abstraction, and the simplicial set structure built with it, permits several advances over the current state of the art in database engineering. Ongoing and future work will focus on developing clear algorithms for application of persistent obstruction theory to real-world database engineering and related problems in data science.

We conclude this paper with several brief remarks about further work and also some practicalities for future use of this theory:

A data sample in any metric space provides a measure, by counting. The measure is or normalized as for any .

For computational purposes, most infinite metric spaces can be considered as compact or finite spaces, using bounds or bins or kernel methods or distributional coordinates that are appropriate to the problem at hand.

On the compact metric spaces , measures of interest can be described as density functions via a Radon–Nikodym comparison to the uniform probability measure .

One attribute can represent models on other attributes, providing an interpretation of Bayesian inference and an opportunity to apply persistent obstruction theory to compact parameterized model spaces. In machine learning, one could use this framework to describe the compatibility of solutions in ensemble methods.

Any list of attributes can be considered as a single attribute, because it still provides measures over some metric space. There is no requirement that attribute value spaces are “minimal” or “1-dimensional” in any sense.

Filtrations other than -Wasserstein might work, too, but someone has to prove that all levels of the filtration are fibrant.

To study a complex of approximate joins, , one must compute Wasserstein distances as in Definition 4.1. This can be done efficiently using the tools of optimal transport as in Reference 17.

To apply Theorem 4.13, one must compute in the simplicial homotopy group . This is definitely the greatest challenge for realizing these mathematical advances as actual software, because homotopy groups are notoriously difficult to compute in general. The task is simplified in our case by several factors. First, we do not necessarily need to know the group structure of to know whether a particular element is trivial in that group. Second, a data subcomplex is always finitely generated with finite, and that finite number is small (several, not several trillion) in most use-cases. Third, because any list of attributes can be considered as a single attribute, problems that are a priori high-dimensional can be studied with a smaller list of formal attributes. Fourth, we expect the homotopy to simplify as increases, so for practical purposes it may be easy to bound even if each is difficult to compute. We expect (or hope) that and are often sufficient for practical problems.

The most important conclusions of this work are: Any manual or automatic data-merging system must analyze homotopy in order to guarantee success; and Obstructions can only be resolved two ways—backing up one step, or allowing additional leeway in the data comparison.

Appendix A. Categorical definitions

This appendix provides a rapid summary of a categorical interpretation of the development in Section 2. For more on these topics, and for the notion of homotopy for fibrant objects in model categories, see Reference 7Reference 9Reference 10Reference 12Reference 18. The reader is warned that each of these references uses a slightly different convention for ordering, opposite categories, and co-/contra-variant functors.

A(a). Simplex

Let denote the set category, whose objects are sets and whose morphisms are functions.

Let denote the simplex category, whose objects are the nonempty sets of natural numbers with the standard ordering , written , and whose morphisms are order-preserving functions. Let denote the augmented simplex category, whose objects are sets of natural numbers with the standard ordering, and whose morphisms are order-preserving functions. The augmented simplicial category includes the empty set, denoted or , which is the initial object in the category. So, . A monomorphism in is a one-to-one order-preserving function. The only bimorphisms/isomorphisms in are the identity maps. Among the morphisms in and are the co-faces and co-degeneracies , defined as follows.

These morphisms satisfy the conditions

(1)

, if ;

(2)

, if ;

(3)

;

(4)

, if ; and

(5)

, if .

Every non-identity morphism in or can be written as a finite composition of co-face and co-degeneracy morphisms, so these five properties essentially characterize and .

For our applications, the following lemmas about monomorphisms in are very useful. They are elementary, but do not appear in the standard references in this form. Merged indexing is merely an ordered formulation of the inclusion–exclusion principle.

Lemma A.1 (Complimentary monomorphism).

For any monomorphism in , write . There is a monomorphism in that enumerates the entries of that are not in the image of .

Lemma A.2 (Merged indexing).

In the category , suppose are equipped with monomorphisms and . Then, for , there are monomorphisms and such that is well-defined. Moreover, the complementary monomorphisms and provide monomorphisms and , as in Diagram Equation A.1. The images of are disjoint.

Proof.

The sets , , have sizes , , respectively. Then, satisfies and defines the object in .

Monomorphisms and can be constructed via the algorithm in Figure 1, which is merely a sequence of concatenations spliced between aligned entries of and . The resulting maps are indeed morphisms, as they are guaranteed to be order-preserving.

Example A.3.

Consider and and . Then . Let be the monomorphism that is written as the sequence . Let be the monomorphism that is written as the sequence . Visually, the merged indexing means

For abbreviation and programming, the constructed monomorphisms can be written as lists.

A(b). Simplicial sets

For any category , the “simplicial category over is . An object in is a contravariant functor . That is, an object in is an assignment of:

for each object in , an object in ;

for each morphism (order-preserving function) in , a morphism in .

The augmented simplicial category, , allows a terminal object in to correspond to the initial object . That is, the trivial map yields a corresponding map , if the category happens to admit a terminal object.

The morphisms in or are the natural transformations as in Equation A.2.

The most important case is , the category of simplicial sets, which is augmented to . The following lemma shows that augmented simplicial sets are given by face and degeneracy maps.

Lemma A.4.

Any object in is a set (called an augmented simplicial set) graded by and equipped with morphisms and for such that

(1)

, if ;

(2)

, if ;

(3)

;

(4)

, if ; and

(5)

, if .

Proof.

The objects are apparent. As for morphisms, each co-face and co-degeneracy morphism in must correspond to face and boundary morphisms in . Because the co-face and co-degeneracy morphisms generate all non-identity morphisms in , it is sufficient to specify the face and degeneracy maps.

Corollary A.5 (Simplicial maps).

The morphisms of or (called simplicial maps) from Equation A.2 are set functions such that and .

A particularly important example of a simplicial set is , the -simplex. (See 3(a).)

Definition A.6 (Simplex).

The standard -simplex is the simplicial set generated (via face and degeneracy maps) by the ordered set in the simplex category .

By the Yoneda Lemma, a simplicial set is characterized by the simplicial maps ; that is, a simplicial set is characterized by its simplices.

Definition A.7 (Horn).

The th horn of the -simplex is the simplicial subset generated by the union of all the faces of except the th face.

By Lemma A.4 and the Yoneda Lemma, if is a simplicial set, then a horn in is a collection of -simplices such that for .

A simplicial map is called a cofibration iff it is a monomorphism. A simplicial map is called a fibration iff for any cofibration , the commutative diagram Equation A.3 can be completed.

Weak-equivalences are defined to be compatible with fibrations and cofibrations according to Reference 18. See also Reference 9. These definitions of cofibration, fibration, and weak equivalence make into a (closed) model category.

A simplicial set is called fibrant or to satisfy the Kan extension condition if is a fibration; that is, a simplicial set satisfies the Kan condition if and only if each horn in can be extended to a simplex in . Let denote the subcategory of fibrant simplicial sets. Then there is a homotopy category , and any admits pointed homotopy groups that characterize the weak equivalence. Moreover, the simplicial homotopy groups of are isomorphic to the continuous homotopy groups of its topological realization, , as discussed in Reference 18, §3 and Reference 9, Chap I.2. See also Reference 10 and Reference 12 for historical explanations that minimize categorical language.

A(c). Data complexes

Let denote the category of data complexes. An object in is a pair of augmented simplicial sets with simplicial map such that for each , the set is a set of data tables over attribute lists from some attribute set , as in Section 2, with and by marginalization and Dirac-delta intersection, respectively.

A morphism in is a simplicial map as in Equation A.4 with some compatibility conditions.

The vertical maps are tuples satisfying the following compatibility conditions.

(1)

is a level of a simplicial map on sets of attribute lists.

(2)

is a level of a simplicial map on sets of measures, with .

(3)

is a continuous function on metric spaces, where . This induces for all .

(4)

If with and , then as measures. That is,

These conditions guarantee simply that the attribute lists , the value spaces , and the measure spaces remain compatible. As with , in Equation A.4, the map can be taken to be or so that the diagram describes naturality with respect to face and degeneracy maps on and . These conditions are sensible for , so they apply to the trivial data table .

For each real number , there is a singleton data complex with , . For each , there is a single attribute list with a singleton value space and one measure, . For brevity, we refer to this singleton data complex as .

Slightly more generally, there is a terminal data complex with , . For each , there is a single attribute list with a singleton value space and measures for each . The terminal data complex is the union of all the singleton data complexes. For brevity, we refer to the terminal data complex as .

Every data complex admits a morphism to the terminal data complex . This terminal morphism maps each data table to the singleton mass . If all data tables in share the same mass (say, ), then the image of the terminal morphism goes to some .

A morphism in is called a cofibration iff it is a monomorphism. A morphism in is called a fibration iff for any cofibration of from a well-aligned pair to a join , the commutative diagram Equation A.6 can be completed.

A data complex is called fibrant if the terminal morphism is a fibration. By Theorem 3.11, if is a fibrant data complex, then is a fibrant simplicial set. Thus, the category is a (closed) model category, and the fibrant subcategory inherits a well-defined homotopy category from , and any admits pointed homotopy groups that characterize the weak equivalence. Moreover, the homotopy groups are isomorphic to the continuous homotopy groups of the topological realization of the underlying simplicial set.

Acknowledgments

We are very grateful to John Paschewitz and Tony Falcone for project guidance and technical direction, and to Greg Friedman, Justin Curry, Jose Perea, and Jonathan Mattingly for helpful discussions at various stages of the theoretical development.

Table of Contents

  1. Abstract
  2. 1. Introduction
    1. Problem 1.1 (Testing problem).
    2. Problem 1.2 (Merging problem).
    3. Problem 1.3 (Meta-merging problem).
    4. 1(a). Overview of technical approach
    5. 1(b). Related work
    6. 1(c). Outline
  3. 2. Attributes and data tables
    1. 2(a). Data subcomplex as a simplicial set
    2. Definition 2.1 (Set of attribute lists).
    3. Definition 2.2 (Ambient data complex).
    4. Definition 2.3 (Data subcomplex).
    5. Definition 2.4 (Finitely generated).
    6. Definition 2.5 (Closed under permutation).
    7. 2(b). Morphisms of data tables
    8. Definition 2.7 (Face of attribute list).
    9. Definition 2.9 (Face of data table).
    10. Lemma 2.10 (Fubini–Tonelli Theorem).
    11. Definition 2.11 (Degeneracy of attribute list).
    12. Definition 2.13 (Degeneracy of data table).
    13. Theorem 2.14 (Simplicial sets).
    14. 2(c). Inclusions, merges, and joins
    15. Definition 2.15 (Attribute inclusions).
    16. Example 2.17.
    17. Lemma 2.18 (Quotient inclusion).
    18. Example 2.19.
    19. Lemma 2.20.
    20. Corollary 2.21.
    21. Definition 2.23 (Reduction).
    22. Definition 2.24 (Sum of attribute lists).
    23. Example 2.25.
    24. Definition 2.26 (Permutation notation).
    25. Definition 2.27 (Merge of attribute lists).
    26. Example 2.28.
    27. Lemma 2.29 (Decomposition of merged lists).
    28. 2(d). Data sections
    29. Definition 2.30 (Data section).
    30. Definition 2.32 (Well-aligned).
    31. Lemma 2.33.
    32. Definition 2.35 (Connected).
    33. Definition 2.36 (Path-connected).
    34. Lemma 2.37.
    35. Problem 2.38 (Testing problem, bis).
    36. Problem 2.39 (Merging and meta-merging problems, bis).
    37. 2(e). Homology
    38. Lemma 2.40.
    39. Lemma 2.41.
    40. Corollary 2.42.
    41. Corollary 2.43.
  4. 3. Homotopy as joins
    1. 3(a). Database joins and the kan condition
    2. Definition 3.1 (Simplex).
    3. Definition 3.2 (Horn).
    4. Definition 3.4 (Kan condition).
    5. Definition 3.5 (Space of joins).
    6. Definition 3.6 (Join conditions).
    7. Definition 3.7 (Admission of trivial joins).
    8. Definition 3.8 (Closure under independent products).
    9. Lemma 3.10.
    10. Theorem 3.11 (Fundamental theorem of data complexes).
    11. 3(b). Simplicial homotopy of data complexes
    12. Definition 3.13 (Fibrant).
    13. Corollary 3.14.
    14. Theorem 3.15.
    15. Definition 3.16.
    16. Definition 3.17 (Complex of perfect joins).
    17. Lemma 3.18.
    18. Corollary 3.19.
    19. Lemma 3.20.
    20. Lemma 3.21.
    21. Corollary 3.22.
  5. 4. Filtrations and obstructions
    1. 4(a). Filtrations from data subcomplexes
    2. Definition 4.1 (Wasserstein distance).
    3. Lemma 4.3.
    4. Lemma 4.4.
    5. Definition 4.5 (The complex of approximate joins).
    6. Theorem 4.6.
    7. Corollary 4.7.
    8. 4(b). Persistent obstruction theory for data subcomplexes
    9. Definition 4.8 (Obstruction cocycle).
    10. Lemma 4.9.
    11. Lemma 4.10.
    12. Theorem 4.12.
    13. Theorem 4.13 (Steenrod’s trichotomy).
    14. Definition 4.14 (Persistence of obstruction).
  6. 5. Discussion
  7. Appendix A. Categorical definitions
    1. A(a). Simplex
    2. Lemma A.1 (Complimentary monomorphism).
    3. Lemma A.2 (Merged indexing).
    4. Example A.3.
    5. A(b). Simplicial sets
    6. Lemma A.4.
    7. Corollary A.5 (Simplicial maps).
    8. Definition A.6 (Simplex).
    9. Definition A.7 (Horn).
    10. A(c). Data complexes
  8. Acknowledgments

Mathematical Fragments

Problem 1.1 (Testing problem).

Given several sources of partial information, how do we test that a hypothetical explanation is reasonably consistent with that partial information?

Problem 1.2 (Merging problem).

Given several sources of partial information, how do we merge that partial information into a cohesive whole?

Problem 1.3 (Meta-merging problem).

Given many sources of partial information, and several partial attempts to merge some combinations of them, is there a way to merge these partial merges into a cohesive whole?

Definition 2.1 (Set of attribute lists).

Let denote the set of all attribute lists in . For each , let denote the set of all attribute lists of length . is a small category. Using the notation from Appendix A, an object in is a function . The case , giving the empty list , is allowed. A morphism of attribute lists is given by (an order-preserving function, which is a morphism of as in Appendix A) such that , which is natural for the commutative diagram 2.1.

Definition 2.7 (Face of attribute list).

The face map on attribute lists, , is defined as omission of the th entry in , so

Definition 2.9 (Face of data table).

For a data table with , let be the measure evaluated on the basis sets of the Borel algebra on as

This is the measure obtained by marginalization to omit the th factor, which could also be written as . Let , which is well-defined in .

Lemma 2.10 (Fubini–Tonelli Theorem).

For any , .

Definition 2.11 (Degeneracy of attribute list).

The degeneracy map on attribute lists, , is defined as repetition of the th entry in , so .

Definition 2.13 (Degeneracy of data table).

For a data table , let be the measure evaluated on the basis sets of the Borel algebra on as

Then, is well-defined in .

Theorem 2.14 (Simplicial sets).

Let be the ambient data complex over an attribute set . For any , consider the face maps and degeneracy maps as in the definitions above. Then

(1)

, if ;

(2)

, if ;

(3)

;

(4)

, if ; and

(5)

, if .

Moreover, the forgetful map commutes with and . That is, and are augmented simplicial sets as in Lemma A.4. They are simplicial sets when omitting the trivial and . Reference 7, Definition 3.2, Reference 9, Equation (1.3), Reference 12, Definition 1.1

Lemma 2.18 (Quotient inclusion).

For any attribute inclusion , there is an attribute list (called the quotient) that enumerates the entries of that are not in the image of . This enumeration equips the quotient with an attribute inclusion , and corresponds to the complimentary monomorphism from Lemma A.1.

Corollary 2.21.

For any attribute inclusion , there is a sequence⁠Footnote9 of face maps such that and such that the attribute inclusion induced by the sequence of face maps is . Moreover, any permutation of this sequence obtained by re-indexing the face-maps according to Lemma 2.10 is equivalent. If , then is the index function for , the quotient inclusion.

9

The backwards ordering here is intentional. Because of the indexing situation and Lemma 2.10, it is simpler to remove attributes from the end. To remove an entire list, one could write or , because is like “pop” on the front of the list.

Definition 2.24 (Sum of attribute lists).

Given attribute lists and in , define as the attribute list obtained by concatenating and .

Definition 2.27 (Merge of attribute lists).

Suppose , and that and are attribute inclusions. Define as the attribute list obtained by performing the index merge specified by Figure1 as in Lemma A.2; this merge concatenates sublists spliced between the entries aligned by . Writing for , Diagram A.1 becomes a diagram of attribute inclusions.

Problem 2.38 (Testing problem, bis).

Consider a data subcomplex of an ambient . Given a data section of the form for , is it true that lies entirely within ? If not, what is the distance from to in ?

Problem 2.39 (Merging and meta-merging problems, bis).

Consider a data subcomplex of an ambient . Suppose that there is a simplicial map of the form . Does there exist an extension of , meaning for all such that ? If not, what is the minimal distance that would allow an approximate extension?

Remark 3.3 (Data tables as simplices).

A data subcomplex in an ambient is an (augmented) simplicial set by Theorem 2.14. Thus, a data table with can be seen as (the generator of) an -simplex, which includes its faces and degeneracies , and so-on. The “vertices” are (generated by) the single-attribute data tables obtained by applying sequences of face maps. For , the -simplices within are (generated by) the data tables obtained by applying sequences of face maps and degeneracy maps until the result has attributes. The picture of “two simplices that share a boundary component” is realized in as a pair of data tables and and attribute inclusions and such that there is a data table with . If and , then this information generates a -horn. A completion of the -horn to a -simplex would be (generated by) a data table that has and as two of its three faces. Depending on the available simplices in , it may or may not be possible to find such .

Definition 3.5 (Space of joins).

Suppose attribute lists are equipped with attribute inclusions and , and let denote the attribute list as in Definition 2.27, which is equipped with inclusions and . For any data tables , let

Note that requires , as must be well-defined.

Theorem 3.11 (Fundamental theorem of data complexes).

For any data subcomplex of an ambient .

(1)

If satisfies the strong join condition, then admits trivial joins and satisfies the Kan condition as a simplicial set.

(2)

If admits trivial joins and satisfies the Kan condition, then satisfies the weak join condition.

Equation (3.2)
Equation (3.8)
Equation (3.9)
Equation (3.10)
Equation (3.12)
Definition 3.13 (Fibrant).

A data complex satisfying the strong join condition is called fibrant.

Theorem 3.15.

For any attribute set and value spaces , the ambient data complex is fibrant.

Lemma 3.20.

For any data subcomplex of an ambient , the complex of perfect joins is fibrant.

Definition 4.1 (Wasserstein distance).

For any with , and , let

The reductions and refer to the two copies of the attribute .

For any and , let

The reductions and refer to the two interwoven copies of the attribute list .

Definition 4.5 (The complex of approximate joins).

Let be a data subcomplex of an ambient . For any , let

Definition 4.8 (Obstruction cocycle).

Let be the filtration of a path-connected data complex. Fix a dimension such that for all faces of all . Let be a data section. For a fixed basepoint , define

to be the element of that is represented by the loop corresponding⁠Footnote15 to the cycle for any . Extend by linearity for . We typically omit the basepoint and ring for brevity, so .

15

The well-definedness of this loop is implied by our assumption .

Lemma 4.9.

Fix . If , then there exists such that the diagram commutes

Theorem 4.12.

Fix a data section . Suppose for some . Then there exists a data section such that . The converse holds as well.

Theorem 4.13 (Steenrod’s trichotomy).

Fix a data subcomplex of an ambient , with Wasserstein filtration . Exactly one of the following is true.

(1)

as a cocycle. Every -cycle of data tables in over a total of attributes can be approximately joined to a single data table over those attributes, allowing error at-most in any reduction to the original data.

(2)

as a cocycle, but as a cohomology class. There is some -cycle of data tables in such that the combined attribute list does not admit an approximate join with error at-most . However, if one considers all of the faces of these data tables, then there is an approximate join to of error at-most .

(3)

as a cohomology class. There is some -cycle of data tables in such that the combined attribute list does not admit an approximate join with error at-most , even when omitting attributes from the original data tables. The only way to produce a single joined table is to increase the error threshold .

Definition 4.14 (Persistence of obstruction).

Let be the filtration of a path-connected data complex. Fix a dimension such that for all faces of all . Let be a data section. Fix a basepoint . Define

and

Note that .

Lemma A.1 (Complimentary monomorphism).

For any monomorphism in , write . There is a monomorphism in that enumerates the entries of that are not in the image of .

Lemma A.2 (Merged indexing).

In the category , suppose are equipped with monomorphisms and . Then, for , there are monomorphisms and such that is well-defined. Moreover, the complementary monomorphisms and provide monomorphisms and , as in Diagram A.1. The images of are disjoint.

Example A.3.

Consider and and . Then . Let be the monomorphism that is written as the sequence . Let be the monomorphism that is written as the sequence . Visually, the merged indexing means

For abbreviation and programming, the constructed monomorphisms can be written as lists.

Equation (A.2)
Lemma A.4.

Any object in is a set (called an augmented simplicial set) graded by and equipped with morphisms and for such that

(1)

, if ;

(2)

, if ;

(3)

;

(4)

, if ; and

(5)

, if .

Definition A.7 (Horn).

The th horn of the -simplex is the simplicial subset generated by the union of all the faces of except the th face.

Equation (A.3)
Equation (A.4)
Equation (A.6)

References

Reference [1]
Samson Abramsky, Relational databases and Bell’s theorem, In search of elegance in the theory and practice of computation, Lecture Notes in Comput. Sci., vol. 8000, Springer, Heidelberg, 2013, pp. 13–35, DOI 10.1007/978-3-642-41660-6_2. MR3124039,
Show rawAMSref \bib{Abramsky2013}{article}{ author={Abramsky, Samson}, title={Relational databases and Bell's theorem}, conference={ title={In search of elegance in the theory and practice of computation}, }, book={ series={Lecture Notes in Comput. Sci.}, volume={8000}, publisher={Springer, Heidelberg}, }, date={2013}, pages={13--35}, review={\MR {3124039}}, doi={10.1007/978-3-642-41660-6\_2}, }
Reference [2]
Samson Abramsky, Rui Soares Barbosa, Kohei Kishida, Raymond Lal, and Shane Mansfield, Contextuality, cohomology and paradox, 24th EACSL Annual Conference on Computer Science Logic, LIPIcs. Leibniz Int. Proc. Inform., vol. 41, Schloss Dagstuhl. Leibniz-Zent. Inform., Wadern, 2015, pp. 211–228. MR3441764,
Show rawAMSref \bib{Abramsky2015}{article}{ author={Abramsky, Samson}, author={Barbosa, Rui Soares}, author={Kishida, Kohei}, author={Lal, Raymond}, author={Mansfield, Shane}, title={Contextuality, cohomology and paradox}, conference={ title={24th EACSL Annual Conference on Computer Science Logic}, }, book={ series={LIPIcs. Leibniz Int. Proc. Inform.}, volume={41}, publisher={Schloss Dagstuhl. Leibniz-Zent. Inform., Wadern}, }, date={2015}, pages={211--228}, review={\MR {3441764}}, }
Reference [3]
Torgny Lindvall, Lectures on the coupling method, Wiley Series in Probability and Mathematical Statistics: Probability and Mathematical Statistics, John Wiley & Sons, Inc., New York, 1992. A Wiley-Interscience Publication. MR1180522,
Show rawAMSref \bib{Baxter1994}{book}{ author={Lindvall, Torgny}, title={Lectures on the coupling method}, series={Wiley Series in Probability and Mathematical Statistics: Probability and Mathematical Statistics}, note={A Wiley-Interscience Publication}, publisher={John Wiley \& Sons, Inc., New York}, date={1992}, pages={xiv+257}, isbn={0-471-54025-0}, review={\MR {1180522}}, }
Reference [4]
Herbert Edelsbrunner and John L. Harer, Computational topology, American Mathematical Society, Providence, RI, 2010. An introduction, DOI 10.1090/mbk/069. MR2572029,
Show rawAMSref \bib{Edelsbrunner2010}{book}{ author={Edelsbrunner, Herbert}, author={Harer, John L.}, title={Computational topology}, note={An introduction}, publisher={American Mathematical Society, Providence, RI}, date={2010}, pages={xii+241}, isbn={978-0-8218-4925-5}, review={\MR {2572029}}, doi={10.1090/mbk/069}, }
Reference [5]
P. J. Ehlers and T. Porter, Joins for (augmented) simplicial sets, J. Pure Appl. Algebra 145 (2000), no. 1, 37–44, DOI 10.1016/S0022-4049(98)00065-6. MR1732286,
Show rawAMSref \bib{Ehlers1999}{article}{ author={Ehlers, P. J.}, author={Porter, T.}, title={Joins for (augmented) simplicial sets}, journal={J. Pure Appl. Algebra}, volume={145}, date={2000}, number={1}, pages={37--44}, issn={0022-4049}, review={\MR {1732286}}, doi={10.1016/S0022-4049(98)00065-6}, }
Reference [6]
Brendan Fong and David I. Spivak, An invitation to applied category theory: Seven sketches in compositionality, Cambridge University Press, Cambridge, 2019, DOI 10.1017/9781108668804. MR3966447,
Show rawAMSref \bib{Fong2018}{book}{ author={Fong, Brendan}, author={Spivak, David I.}, title={An invitation to applied category theory}, subtitle={Seven sketches in compositionality}, publisher={Cambridge University Press, Cambridge}, date={2019}, pages={xii+338}, isbn={978-1-108-71182-1}, isbn={978-1-108-48229-5}, review={\MR {3966447}}, doi={10.1017/9781108668804}, }
Reference [7]
Greg Friedman, Survey article: An elementary illustrated introduction to simplicial sets, Rocky Mountain J. Math. 42 (2012), no. 2, 353–423, DOI 10.1216/RMJ-2012-42-2-353. MR2915498,
Show rawAMSref \bib{Friedman2008}{article}{ author={Friedman, Greg}, title={Survey article: An elementary illustrated introduction to simplicial sets}, journal={Rocky Mountain J. Math.}, volume={42}, date={2012}, number={2}, pages={353--423}, issn={0035-7596}, review={\MR {2915498}}, doi={10.1216/RMJ-2012-42-2-353}, }
Reference [8]
Eli Glasner, Ergodic theory via joinings, Mathematical Surveys and Monographs, vol. 101, American Mathematical Society, Providence, RI, 2003, DOI 10.1090/surv/101. MR1958753,
Show rawAMSref \bib{Glasner2003}{book}{ author={Glasner, Eli}, title={Ergodic theory via joinings}, series={Mathematical Surveys and Monographs}, volume={101}, publisher={American Mathematical Society, Providence, RI}, date={2003}, pages={xii+384}, isbn={0-8218-3372-3}, review={\MR {1958753}}, doi={10.1090/surv/101}, }
Reference [9]
Paul G. Goerss and John F. Jardine, Simplicial homotopy theory, Modern Birkhäuser Classics, Birkhäuser Verlag, Basel, 2009. Reprint of the 1999 edition [MR1711612], DOI 10.1007/978-3-0346-0189-4. MR2840650,
Show rawAMSref \bib{Seriesa}{book}{ author={Goerss, Paul G.}, author={Jardine, John F.}, title={Simplicial homotopy theory}, series={Modern Birkh\"{a}user Classics}, note={Reprint of the 1999 edition [MR1711612]}, publisher={Birkh\"{a}user Verlag, Basel}, date={2009}, pages={xvi+510}, isbn={978-3-0346-0188-7}, review={\MR {2840650}}, doi={10.1007/978-3-0346-0189-4}, }
Reference [10]
Daniel M. Kan, A combinatorial definition of homotopy groups, Ann. of Math. (2) 67 (1958), 282–312, DOI 10.2307/1970006. MR111032,
Show rawAMSref \bib{Kan1958}{article}{ author={Kan, Daniel M.}, title={A combinatorial definition of homotopy groups}, journal={Ann. of Math. (2)}, volume={67}, date={1958}, pages={282--312}, issn={0003-486X}, review={\MR {111032}}, doi={10.2307/1970006}, }
Reference [11]
Albert T. Lundell, Obstruction theory of principal fibre bundles, Trans. Amer. Math. Soc. 97 (1960), 161–192, DOI 10.2307/1993368. MR130694,
Show rawAMSref \bib{Lundell1960}{article}{ author={Lundell, Albert T.}, title={Obstruction theory of principal fibre bundles}, journal={Trans. Amer. Math. Soc.}, volume={97}, date={1960}, pages={161--192}, issn={0002-9947}, review={\MR {130694}}, doi={10.2307/1993368}, }
Reference [12]
J. Peter May, Simplicial objects in algebraic topology, Chicago Lectures in Mathematics, University of Chicago Press, Chicago, IL, 1992. Reprint of the 1967 original. MR1206474,
Show rawAMSref \bib{May1967}{book}{ author={May, J. Peter}, title={Simplicial objects in algebraic topology}, series={Chicago Lectures in Mathematics}, note={Reprint of the 1967 original}, publisher={University of Chicago Press, Chicago, IL}, date={1992}, pages={viii+161}, isbn={0-226-51181-2}, review={\MR {1206474}}, }
Reference [13]
Jason Morton, Contextuality from missing and versioned data, 1–21, arXiv:1708.03264 2017.
Reference [14]
Felix Naumann, Alexander Bilke, Jens Bleiholder, and Melanie Weis, Data fusion in three steps: resolving inconsistencies at schema-, tuple-, and value-level, Bulletin of the IEEE Computer Society Technical Committee on Data Engineering 29 (2006), no. 2, 21–31.
Reference [15]
Paul Olum, Obstructions to extensions and homotopies, Ann. of Math. (2) 52 (1950), 1–50, DOI 10.2307/1969510. MR36507,
Show rawAMSref \bib{Olum1950}{article}{ author={Olum, Paul}, title={Obstructions to extensions and homotopies}, journal={Ann. of Math. (2)}, volume={52}, date={1950}, pages={1--50}, issn={0003-486X}, review={\MR {36507}}, doi={10.2307/1969510}, }
Reference [16]
Nina Otter, Magnitude meets persistence. homology theories for filtered simplicial sets, 1–21, arXiv:1807.01540 2018.
Reference [17]
Gabriel Peyré and Marco Cuturi, Computational optimal transport, arXiv:1803.00567 2018.
Reference [18]
Daniel G. Quillen, Homotopical algebra, Lecture Notes in Mathematics, vol. 43, Springer Berlin Heidelberg, Berlin, Heidelberg, 1967, doi:10.1007/BFb0097438.
Reference [19]
Egbert Rijke, The join construction, (2017), arXiv:1701.07538.
Reference [20]
Patrick Schultz, David I. Spivak, Christina Vasilakopoulou, and Ryan Wisnesky, Algebraic databases, Theory Appl. Categ. 32 (2017), Paper No. 16, 547–619. MR3641249,
Show rawAMSref \bib{Schultz2017}{article}{ author={Schultz, Patrick}, author={Spivak, David I.}, author={Vasilakopoulou, Christina}, author={Wisnesky, Ryan}, title={Algebraic databases}, journal={Theory Appl. Categ.}, volume={32}, date={2017}, pages={Paper No. 16, 547--619}, review={\MR {3641249}}, }
Reference [21]
Patrick Schultz and Ryan Wisnesky, Algebraic data integration, J. Funct. Programming 27 (2017), e24, 51, DOI 10.1017/S0956796817000168. MR3720789,
Show rawAMSref \bib{Schultz2015}{article}{ author={Schultz, Patrick}, author={Wisnesky, Ryan}, title={Algebraic data integration}, journal={J. Funct. Programming}, volume={27}, date={2017}, pages={e24, 51}, issn={0956-7968}, review={\MR {3720789}}, doi={10.1017/S0956796817000168}, }
Reference [22]
N. E. Steenrod, Homology with local coefficients, Ann. of Math. (2) 44 (1943), 610–627, DOI 10.2307/1969099. MR9114,
Show rawAMSref \bib{Steenrod1943}{article}{ author={Steenrod, N. E.}, title={Homology with local coefficients}, journal={Ann. of Math. (2)}, volume={44}, date={1943}, pages={610--627}, issn={0003-486X}, review={\MR {9114}}, doi={10.2307/1969099}, }
Reference [23]
Norman Steenrod, The Topology of Fibre Bundles, Princeton Mathematical Series, vol. 14, Princeton University Press, Princeton, N. J., 1951. MR0039258,
Show rawAMSref \bib{Steenrod}{book}{ author={Steenrod, Norman}, title={The Topology of Fibre Bundles}, series={Princeton Mathematical Series, vol. 14}, publisher={Princeton University Press, Princeton, N. J.}, date={1951}, pages={viii+224}, review={\MR {0039258}}, }
Reference [24]
Phillip B. Thurber, Semi-localization of a one pointed Kan complex, Pacific J. Math. 178 (1997), no. 1, 147–184, DOI 10.2140/pjm.1997.178.147. MR1447409,
Show rawAMSref \bib{Thurber1997}{article}{ author={Thurber, Phillip B.}, title={Semi-localization of a one pointed Kan complex}, journal={Pacific J. Math.}, volume={178}, date={1997}, number={1}, pages={147--184}, issn={0030-8730}, review={\MR {1447409}}, doi={10.2140/pjm.1997.178.147}, }

Article Information

MSC 2020
Primary: 55U10 (Simplicial sets and complexes in algebraic topology)
Secondary: 55S35 (Obstruction theory in algebraic topology)
Author Information
Abraham D. Smith
Department of Mathematics, Statistics, and Computer Science, University of Wisconsin-Stout, Menomonie, Wisconsin 54751; and Geometric Data Analytics, Inc., Durham, North Carolina 27707
smithabr@uwstout.edu
ORCID
MathSciNet
Paul Bendich
Department of Mathematics, Duke University, Durham, North Carolina 27708; and Geometric Data Analytics, Inc., Durham, North Carolina 27707
bendich@math.duke.edu
MathSciNet
John Harer
Department of Mathematics, Duke University, Durham, North Carolina 27707; and Geometric Data Analytics, Inc., Durham, North Carolina 27707
harer@math.duke.edu
MathSciNet
Additional Notes

Work by all three authors was partially supported by the DARPA Simplex Program, under contract # HR001118C0070. The last two authors were also partially supported by the Air Force Office of Scientific Research under grant AFOSR FA9550-18-1-0266.

Journal Information
Transactions of the American Mathematical Society, Series B, Volume 8, Issue 1, ISSN 2330-0000, published by the American Mathematical Society, Providence, Rhode Island.
Publication History
This article was received on , revised on , and published on .
Copyright Information
Copyright 2021 by the authors under Creative Commons Attribution-Noncommercial 3.0 License (CC BY NC 3.0)
Article References
  • Permalink
  • Permalink (PDF)
  • DOI 10.1090/btran/56
  • MathSciNet Review: 4207891
  • Show rawAMSref \bib{4207891}{article}{ author={Smith, Abraham}, author={Bendich, Paul}, author={Harer, John}, title={Persistent obstruction theory for a model category of measures with applications to data merging}, journal={Trans. Amer. Math. Soc. Ser. B}, volume={8}, number={1}, date={2021}, pages={1-38}, issn={2330-0000}, review={4207891}, doi={10.1090/btran/56}, }

Settings

Change font size
Resize article panel
Enable equation enrichment

Note. To explore an equation, focus it (e.g., by clicking on it) and use the arrow keys to navigate its structure. Screenreader users should be advised that enabling speech synthesis will lead to duplicate aural rendering.

For more information please visit the AMS MathViewer documentation.