Equivalent definitions of mathematical structures

Last updated

In mathematics, equivalent definitions are used in two somewhat different ways. First, within a particular mathematical theory (for example, Euclidean geometry), a notion (for example, ellipse or minimal surface) may have more than one definition. These definitions are equivalent in the context of a given mathematical structure (Euclidean space, in this case). Second, a mathematical structure may have more than one definition (for example, topological space has at least seven definitions; ordered field has at least two definitions).

Contents

In the former case, equivalence of two definitions means that a mathematical object (for example, geometric body) satisfies one definition if and only if it satisfies the other definition.

In the latter case, the meaning of equivalence (between two definitions of a structure) is more complicated, since a structure is more abstract than an object. Many different objects may implement the same structure.

Isomorphic implementations

Natural numbers may be implemented as 0 = { }, 1 = {0} = {{ }}, 2 = {0, 1} = {{ }, {{ }}}, 3 = {0, 1, 2} = {{ }, {{ }}, {{ }, {{ }}}} and so on; or alternatively as 0 = { }, 1 = {0} ={{ }}, 2 = {1} = {{{ }}} and so on. These are two different but isomorphic implementations of natural numbers in set theory. They are isomorphic as models of Peano axioms, that is, triples (N,0,S) where N is a set, 0 an element of N, and S (called the successor function) a map of N to itself (satisfying appropriate conditions). In the first implementation S(n) = n{n}; in the second implementation S(n) = {n}. As emphasized in Benacerraf's identification problem, the two implementations differ in their answer to the question whether 0 ∈ 2; however, this is not a legitimate question about natural numbers (since the relation ∈ is not stipulated by the relevant signature(s), see the next section). [details 1] Similarly, different but isomorphic implementations are used for complex numbers.

Deduced structures and cryptomorphisms

The successor function S on natural numbers leads to arithmetic operations, addition and multiplication, and the total order, thus endowing N with an ordered semiring structure. This is an example of a deduced structure. The ordered semiring structure (N, +, ·, ≤) is deduced from the Peano structure (N, 0, S) by the following procedure: n + 0 = n,  m + S (n) = S (m + n),  m · 0 = 0,  m · S (n) = m + (m · n), and mn if and only if there exists kN such that m + k = n. And conversely, the Peano structure is deduced from the ordered semiring structure as follows: S (n) = n + 1, and 0 is defined by 0 + 0 = 0. It means that the two structures on N are equivalent by means of the two procedures.

The two isomorphic implementations of natural numbers mentioned in the previous section are isomorphic as triples (N,0,S), that is, structures of the same signature (0,S) consisting of a constant symbol 0 and a unary function S. An ordered semiring structure (N, +, ·, ≤) has another signature (+, ·, ≤) consisting of two binary functions and one binary relation. The notion of isomorphism does not apply to structures of different signatures. In particular, a Peano structure cannot be isomorphic to an ordered semiring. However, an ordered semiring deduced from a Peano structure may be isomorphic to another ordered semiring. Such relation between structures of different signatures is sometimes called a cryptomorphism.

Ambient frameworks

A structure may be implemented within a set theory ZFC, or another set theory such as NBG, NFU, ETCS. [1] Alternatively, a structure may be treated in the framework of first-order logic, second-order logic, higher-order logic, a type theory, homotopy type theory etc. [details 2]

Structures according to Bourbaki

"Mathematics [...] cannot be explained completely by a single concept such as the mathematical structure. Nevertheless, Bourbaki's structuralist approach is the best that we have." (Pudlák 2013 , page 3)
"Evident as the notion of mathematical structure may seem these days, it was at least not made explicit until the middle of the 20th century. Then it was the influence of the Bourbaki-project and then later the development of category theory which made the notion explicit" (nLab).

According to Bourbaki, the scale of sets on a given set X consists of all sets arising from X by taking Cartesian products and power sets, in any combination, a finite number of times. Examples: X; X × X; P(X); P(P(X × X) × X × P(P(X))) × X. (Here A × B is the product of A and B, and P(A) is the powerset of A.) In particular, a pair (0,S) consisting of an element 0 ∈ N and a unary function S : NN belongs to N × P(N × N) (since a function is a subset of the Cartesian product). A triple (+, ·, ≤) consisting of two binary functions N × NN and one binary relation on N belongs to P(N × N × N) × P(N × N × N) × P(N × N). Similarly, every algebraic structure on a set belongs to the corresponding set in the scale of sets on X.

Non-algebraic structures on a set X often involve sets of subsets of X (that is, subsets of P(X), in other words, elements of P(P(X))). For example, the structure of a topological space, called a topology on X, treated as the set of "open" sets; or the structure of a measurable space, treated as the σ-algebra of "measurable" sets; both are elements of P(P(X)). These are second-order structures. [2]

More complicated non-algebraic structures combine an algebraic component and a non-algebraic component. For example, the structure of a topological group consists of a topology and the structure of a group. Thus it belongs to the product of P(P(X)) and another ("algebraic") set in the scale; this product is again a set in the scale.

Transport of structures; isomorphism

Given two sets X, Y and a bijection f : XY, one constructs the corresponding bijections between scale sets. Namely, the bijection X × XY × Y sends (x1,x2) to (f(x1),f(x2)); the bijection P(X) → P(Y) sends a subset A of X into its image f(A) in Y; and so on, recursively: a scale set being either product of scale sets or power set of a scale set, one of the two constructions applies.

Let (X,U) and (Y,V) be two structures of the same signature. Then U belongs to a scale set SX, and V belongs to the corresponding scale set SY. Using the bijection F : SXSY constructed from a bijection f : XY, one defines:

f is an isomorphism between (X,U) and (Y,V) if F(U) = V.

This general notion of isomorphism generalizes many less general notions listed below.

In fact, Bourbaki stipulates two additional features. First, several sets X1, ..., Xn (so-called principal base sets) may be used, rather than a single set X. However, this feature is of little use. All the items listed above use a single principal base set. Second, so-called auxiliary base sets E1, ..., Em may be used. This feature is widely used. Indeed, the structure of a vector space stipulates not only addition X × XX but also scalar multiplication R × XX (if R is the field of scalars). Thus, R is an auxiliary base set (called also "external" [3] ). The scale of sets consists of all sets arising from all base sets (both principal and auxiliary) by taking Cartesian products and power sets. Still, the map f (possibly an isomorphism) acts on X only; auxiliary sets are endowed by identity maps. (However, the case of n principal sets leads to n maps.)

Functoriality

Several statements formulated by Bourbaki without mentioning categories can be reformulated readily in the language of category theory. First, some terminology.

Proposition. [7] Each echelon construction scheme leads to a functor from Set* to itself.

In particular, the permutation group of a set X acts on every scale set SX.

In order to formulate one more proposition, the notion "species of structures" is needed, since echelon construction scheme gives only preliminary information on a structure. For example, commutative groups and (arbitrary) groups are two different species of the same echelon construction scheme. Another example: topological spaces and measurable spaces. They differ in the so-called axiom of the species. This axiom is the conjunction of all required properties, such as "multiplication is associative" for groups, or "the union of open sets is an open set" for topological spaces.

Proposition. [8] Each species of structures leads to a functor from Set* to itself.

Example. For the species of groups, the functor F maps a set X to the set F(X) of all group structures on X. For the species of topological spaces, the functor F maps a set X to the set F(X) of all topologies on X. The morphism F(f) : F(X) → F(Y) corresponding to a bijection f : XY is the transport of structures. Topologies on Y correspond bijectively to topologies on X. The same holds for group structures, etc.

In particular, the set of all structures of a given species on a given set is invariant under the action of the permutation group on the corresponding scale set SX, and is a fixed point of the action of the group on another scale set P(SX). However, not all fixed points of this action correspond to species of structures. [details 5]

Given two species, Bourbaki defines the notion "procedure of deduction" (of a structure of the second species from a structure of the first species). [9] A pair of mutually inverse procedures of deduction leads to the notion "equivalent species". [10]

Example. The structure of a topological space may be defined as an open set topology or alternatively, a closed set topology. The two corresponding procedures of deduction coincide; each one replaces all given subsets of X with their complements. In this sense, these are two equivalent species.

In the general definition of Bourbaki, deduction procedure may include a change of the principal base set(s), but this case is not treated here. In the language of category theory one have the following result.

Proposition. [10] Equivalence between two species of structures leads to a natural isomorphism between the corresponding functors.

However, in general, not all natural isomorphisms between these functors correspond to equivalences between the species. [details 6]

Mathematical practice

"We often do not distinguish structures that are isomorphic and often say that 'two structures are the same, up to isomorphism'." [11]
"When studying structures we are interested only in their form, but when we prove their existence we need to construct them." [12]
'Mathematicians are of course used to identifying isomorphic structures in practice, but they generally do so by "abuse of notation", or some other informal device, knowing that the objects involved are not "really" identical.' [13] (A radically better approach is expected; but for now, Summer 2014, the definitive book quoted above does not elaborate on structures.)

In practice, one makes no distinction between equivalent species of structures. [10]

Usually, a text based on natural numbers (for example, the article "prime number") does not specify the used definition of natural numbers. Likewise, a text based on topological spaces (for example, the article "homotopy", or "inductive dimension") does not specify the used definition of a topological space. Thus, it is possible (and rather probable) that the reader and the author interpret the text differently, according to different definitions. Nevertheless, the communication is successful, which means that such different definitions may be thought of as equivalent.

A person acquainted with topological spaces knows basic relations between neighborhoods, convergence, continuity, boundary, closure, interior, open sets, closed sets, and does not need to know that some of these notions are "primary", stipulated in the definition of a topological space, while others are "secondary", characterized in terms of "primary" notions. Moreover, knowing that subsets of a topological space are themselves topological spaces, as well as products of topological spaces, the person is able to construct some new topological spaces irrespective of the definition.

Thus, in practice a topology on a set is treated like an abstract data type that provides all needed notions (and constructors) but hides the distinction between "primary" and "secondary" notions. The same applies to other kinds of mathematical structures. "Interestingly, the formalization of structures in set theory is a similar task as the formalization of structures for computers." [14]

Canonical, not just natural

As was mentioned, equivalence between two species of structures leads to a natural isomorphism between the corresponding functors. However, "natural" does not mean "canonical". A natural transformation is generally non-unique.

Example. Consider again the two equivalent structures for natural numbers. One is the "Peano structure" (0,S), the other is the structure (+, ·, ≤) of ordered semiring. If a set X is endowed by both structures then, on one hand, X = { a0, a1, a2, ... } where S(an) = an+1 for all n and 0 = a0; and on the other hand, X = { b0, b1, b2, ... } where bm+n = bm + bn, bm·n = bm · bn, and bmbn if and only if mn. Requiring that an = bn for all n one gets the canonical equivalence between the two structures. However, one may also require a0 = b1, a1 = b0, and an = bn for all n > 1, thus getting another, non-canonical, natural isomorphism. Moreover, every permutation of the index set { 0, 1, 2, ... } leads to a natural isomorphism; they are uncountably many!

Another example. A structure of a (simple) graph on a set V = { 1, 2, ..., n } of vertices may be described by means of its adjacency matrix, a (0,1)-matrix of size n×n (with zeros on the diagonal). More generally, for arbitrary V an adjacency function on V × V may be used. The canonical equivalence is given by the rule: "1" means "connected" (with an edge), "0" means "not connected". However, another rule, "0" means "connected", "1" means "not", may be used, and leads to another, natural but not canonical, equivalence. In this example, canonicity is rather a matter of convention. But here is a worse case. Instead of "0" and "1" one may use, say, the two possible orientations of the plane R2 ("clockwise" and "counterclockwise"). It is difficult to choose a canonical rule in this case!

"Natural" is a well-defined mathematical notion, but it does not ensure uniqueness. "Canonical" does, but generally is more or less conventional. A consistent choice of canonical equivalences is an inevitable component of equivalent definitions of mathematical structures.

See also

Notes

  1. Technically, "0 ∈ 2" is an example of a non-transportable relation, see Bourbaki 1968 , Sect.IV.1.3, Marshall & Chuaqui 1991.
  2. A reasonable choice of an ambient framework should not alter basic properties of a structure, but can alter provability of finer properties. For example, some theorems about the natural numbers are provable in set theory (and some other strong systems) but not provable in first-order logic; see Paris–Harrington theorem and Goodstein's theorem. The same applies to definability; see for example Tarski's undefinability theorem.
  3. In order to be more formal, Bourbaki encodes such formulas with sequences of ordered pairs of natural numbers.
  4. On one hand, it is possible to exclude the Cartesian products, treating a pair (x,y) as just the set {{x},{x,y}}. On the other hand, it is possible to include the set operation X,Y->YX (all functions from X to Y). "It is possible to simplify the matter by considering operations and functions as a special kind of relations (for example, a binary operation is a ternary relation). However, quite often, it is an advantage to have operations as a primitive concept." Pudlák 2013 , page 17
  5. The set of all possible axioms of species is countable, while the set of all fixed points of the considered action may be uncountable. Tarski's "logical notions of higher order" are closer to the fixed points than to species of structures, see Feferman 2010 and references therefrom.
  6. The set of all possible deduction procedures is countable, while the set of all natural isomorphisms between the considered functors may be uncountable (see an example in Section #Canonical, not just natural).

Footnotes

  1. About ETCS see Type theory#Mathematical foundations
  2. Pudlák 2013 , pages 10–11
  3. Pudlák 2013 , page 12
  4. Bourbaki 1968 , Sect.IV.1.1
  5. Pudlák 2013 , page 10
  6. Marshall & Chuaqui 1991 , §2
  7. Bourbaki 1968 , Sect.IV.1.2
  8. Bourbaki 1968 , Sect.IV.1.5
  9. Bourbaki 1968 , Sect.IV.1.6
  10. 1 2 3 Bourbaki 1968 , Sect.IV.1.7
  11. Pudlák 2013 , page 13
  12. Pudlák 2013 , page 22
  13. The Univalent Foundations Program 2013 , Subsection "Univalent foundations" of Introduction
  14. Pudlák 2013 , page 34

Related Research Articles

In category theory, a branch of mathematics, a Grothendieck topology is a structure on a category C that makes the objects of C act like the open sets of a topological space. A category together with a choice of Grothendieck topology is called a site.

Homeomorphism Isomorphism in topology (mathematics)

In the mathematical field of topology, a homeomorphism, topological isomorphism, or bicontinuous function is a continuous function between topological spaces that has a continuous inverse function. Homeomorphisms are the isomorphisms in the category of topological spaces—that is, they are the mappings that preserve all the topological properties of a given space. Two spaces with a homeomorphism between them are called homeomorphic, and from a topological viewpoint they are the same. The word homeomorphism comes from the Greek words ὅμοιος (homoios) = similar or same and μορφή (morphē) = shape or form, introduced to mathematics by Henri Poincaré in 1895.

Isomorphism In mathematics, invertible homomorphism

In mathematics, an isomorphism is a structure-preserving mapping between two structures of the same type that can be reversed by an inverse mapping. Two mathematical structures are isomorphic if an isomorphism exists between them. The word isomorphism is derived from the Ancient Greek: ἴσοςisos "equal", and μορφήmorphe "form" or "shape".

Topological group Group that is a topological space with continuous group action

In mathematics, topological groups are logically the combination of groups and topological spaces, i.e. they are groups and topological spaces at the same time, such that the continuity condition for the group operations connects these two structures together and consequently they are not independent from each other.

In mathematics, specifically category theory, adjunction is a relationship that two functors may have, intuitively corresponding to a weak form of equivalence between two related categories. Two functors that stand in this relationship are known as adjoint functors, one being the left adjoint and the other the right adjoint. Pairs of adjoint functors are ubiquitous in mathematics and often arise from constructions of "optimal solutions" to certain problems, such as the construction of a free group on a set in algebra, or the construction of the Stone–Čech compactification of a topological space in topology.

In category theory, an epimorphism is a morphism f : XY that is right-cancellative in the sense that, for all objects Z and all morphisms g1, g2: YZ,

In mathematics, specifically in homology theory and algebraic topology, cohomology is a general term for a sequence of abelian groups, usually one associated with a topological space, often defined from a cochain complex. Cohomology can be viewed as a method of assigning richer algebraic invariants to a space than homology. Some versions of cohomology arise by dualizing the construction of homology. In other words, cochains are functions on the group of chains in homology theory.

In category theory, a branch of abstract mathematics, an equivalence of categories is a relation between two categories that establishes that these categories are "essentially the same". There are numerous examples of categorical equivalences from many areas of mathematics. Establishing an equivalence involves demonstrating strong similarities between the mathematical structures concerned. In some cases, these structures may appear to be unrelated at a superficial or intuitive level, making the notion fairly powerful: it creates the opportunity to "translate" theorems between different kinds of mathematical structures, knowing that the essential meaning of those theorems is preserved under the translation.

In mathematics, localization of a category consists of adding to a category inverse morphisms for some collection of morphisms, constraining them to become isomorphisms. This is formally similar to the process of localization of a ring; it in general makes objects isomorphic that were not so before. In homotopy theory, for example, there are many examples of mappings that are invertible up to homotopy; and so large classes of homotopy equivalent spaces. Calculus of fractions is another name for working in a localized category.

In mathematics, particularly category theory, a representable functor is a certain functor from an arbitrary category into the category of sets. Such functors give representations of an abstract category in terms of known structures allowing one to utilize, as much as possible, knowledge about the category of sets in other settings.

In mathematics, the derived categoryD(A) of an abelian category A is a construction of homological algebra introduced to refine and in a certain sense to simplify the theory of derived functors defined on A. The construction proceeds on the basis that the objects of D(A) should be chain complexes in A, with two such chain complexes considered isomorphic when there is a chain map that induces an isomorphism on the level of homology of the chain complexes. Derived functors can then be defined for chain complexes, refining the concept of hypercohomology. The definitions lead to a significant simplification of formulas otherwise described by complicated spectral sequences.

In mathematics, the homotopy category is a category built from the category of topological spaces which in a sense identifies two spaces that have the same shape. The phrase is in fact used for two different categories, as discussed below.

Fibred categories are abstract entities in mathematics used to provide a general framework for descent theory. They formalise the various situations in geometry and algebra in which inverse images of objects such as vector bundles can be defined. As an example, for each topological space there is the category of vector bundles on the space, and for every continuous map from a topological space X to another topological space Y is associated the pullback functor taking bundles on Y to bundles on X. Fibred categories formalise the system consisting of these categories and inverse image functors. Similar setups appear in various guises in mathematics, in particular in algebraic geometry, which is the context in which fibred categories originally appeared. Fibered categories are used to define stacks, which are fibered categories with "descent". Fibrations also play an important role in categorical semantics of type theory, and in particular that of dependent type theories.

This is a glossary of properties and concepts in category theory in mathematics.

In mathematics, a quotient category is a category obtained from another one by identifying sets of morphisms. Formally, it is a quotient object in the category of categories, analogous to a quotient group or quotient space, but in the categorical setting.

Space (mathematics) Mathematical set with some added structure

In mathematics, a space is a set with some added structure.

In mathematics, especially in algebraic topology, an induced homomorphism is a homomorphism derived in a canonical way from another map. For example, a continuous map from a topological space X to a topological space Y induces a group homomorphism from the fundamental group of X to the fundamental group of Y.

In mathematics, a topos is a category that behaves like the category of sheaves of sets on a topological space. Topoi behave much like the category of sets and possess a notion of localization; they are a direct generalization of point-set topology. The Grothendieck topoi find applications in algebraic geometry; the more general elementary topoi are used in logic.

In the mathematical field of order theory, an order isomorphism is a special kind of monotone function that constitutes a suitable notion of isomorphism for partially ordered sets (posets). Whenever two posets are order isomorphic, they can be considered to be "essentially the same" in the sense that either of the orders can be obtained from the other just by renaming of elements. Two strictly weaker notions that relate to order isomorphisms are order embeddings and Galois connections.

In mathematics, a weak equivalence is a notion from homotopy theory that in some sense identifies objects that have the same "shape". This notion is formalized in the axiomatic definition of a model category.

References

Further reading