Canonical form







Algorithmic anagram test using multisets as canonical forms: The strings "madam curie" and "radium came" are given as C arrays. Each one is converted into a canonical form by sorting. Since both sorted strings literally agree, the original strings were anagrams of each other.


In mathematics and computer science, a canonical, normal, or standard form of a mathematical object is a standard way of presenting that object as a mathematical expression. The distinction between "canonical" and "normal" forms varies by subfield. In most fields, a canonical form specifies a unique representation for every object, while a normal form simply specifies its form, without the requirement of uniqueness.


The canonical form of a positive integer in decimal representation is a finite sequence of digits that does not begin with zero.


More generally, for a class of objects on which an equivalence relation is defined, a canonical form consists in the choice of a specific object in each class. For example, Jordan normal form is a canonical form for matrix similarity, and the row echelon form is a canonical form, when one considers as equivalent a matrix and its left product by an invertible matrix.


In computer science, and more specifically in computer algebra, when representing mathematical objects in a computer, there are usually many different ways to represent the same object. In this context, a canonical form is a representation such that every object has a unique representation. Thus, the equality of two objects can easily be tested by testing the equality of their canonical forms. However canonical forms frequently depend on arbitrary choices (like ordering the variables), and this introduces difficulties for testing the equality of two objects resulting on independent computations. Therefore, in computer algebra, normal form is a weaker notion: A normal form is a representation such that zero is uniquely represented. This allows testing for equality by putting the difference of two objects in normal form.


Canonical form can also mean a differential form that is defined in a natural (canonical) way.


In computer science, data that has more than one possible representation can often be canonicalized into a completely unique representation called its canonical form. Putting something into canonical form is canonicalization.[1]




Contents






  • 1 Definition


  • 2 Examples


    • 2.1 Large number notation


    • 2.2 Number theory


    • 2.3 Linear algebra


    • 2.4 Algebra


    • 2.5 Geometry


    • 2.6 Integrable systems


    • 2.7 Dynamical systems


    • 2.8 Three dimensional geometry


    • 2.9 Functional analysis


    • 2.10 Classical logic


    • 2.11 Set theory


    • 2.12 Game theory


    • 2.13 Proof theory


    • 2.14 Rewriting systems


    • 2.15 Lambda calculus


    • 2.16 Graph theory


    • 2.17 Computing




  • 3 See also


  • 4 Notes


  • 5 References





Definition


Suppose we have some set S of objects, with an equivalence relation R. A canonical form is given by designating some objects of S to be "in canonical form", such that every object under consideration is equivalent to exactly one object in canonical form. In other words, the canonical forms in S represent the equivalence classes, once and only once. To test whether two objects are equivalent, it then suffices to test their canonical forms for equality.
A canonical form thus provides a classification theorem and more, in that it not just classifies every class, but gives a distinguished (canonical) representative.


Formally, a canonicalization with respect to an equivalence relation R on a set S is a mapping c:SS such that for all s, s1, s2S:




  1. c(s) = c(c(s))   (idempotence),


  2. s1R s2 if and only if c(s1) = c(s2)   (decisiveness), and


  3. s R c(s)   (representativeness).


Property 3 is redundant, it follows by applying 2 to 1.


In practical terms, one wants to be able to recognize the canonical forms. There is also a practical, algorithmic question to consider: how to pass from a given object s in S to its canonical form s*? Canonical forms are generally used to make operating with equivalence classes more effective. For example, in modular arithmetic, the canonical form for a residue class is usually taken as the least non-negative integer in it. Operations on classes are carried out by combining these representatives and then reducing the result to its least non-negative residue.
The uniqueness requirement is sometimes relaxed, allowing the forms to be unique up to some finer equivalence relation, like allowing reordering of terms (if there is no natural ordering on terms).


A canonical form may simply be a convention, or a deep theorem.


For example, polynomials are conventionally written with the terms in descending powers: it is more usual to write x2 + x + 30 than x + 30 + x2, although the two forms define the same polynomial. By contrast, the existence of Jordan canonical form for a matrix is a deep theorem.



Examples


Note: in this section, "up to" some equivalence relation E means that the canonical form is not unique in general, but that if one object has two different canonical forms, they are E-equivalent.



Large number notation


Standard form is used by many mathematicians and scientists to write extremely large numbers in a more concise and understandable way.



Number theory



  • canonical representation of a positive integer

  • canonical form of a continued fraction



Linear algebra



















































Objects

A is equivalent to B if:
Normal form
Notes

Normal matrices over the complex numbers

A=U∗BU{displaystyle A=U^{*}BU}A=U^* B U for some unitary matrix U

Diagonal matrices (up to reordering)
This is the Spectral theorem
Matrices over the complex numbers

A=UBV∗{displaystyle A=UBV^{*}}A=U B V^* for some unitary matrices U and V
Diagonal matrices with real positive entries (in descending order)

Singular value decomposition
Matrices over an algebraically closed field

A=P−1BP{displaystyle A=P^{-1}BP}A=P^{-1} B P for some invertible matrix P

Jordan normal form (up to reordering of blocks)

Matrices over an algebraically closed field

A=P−1BP{displaystyle A=P^{-1}BP}A=P^{-1} B P for some invertible matrix P

Weyr canonical form (up to reordering of blocks)

Matrices over a field

A=P−1BP{displaystyle A=P^{-1}BP}A=P^{-1} B P for some invertible matrix P

Frobenius normal form

Matrices over a principal ideal domain

A=P−1BQ{displaystyle A=P^{-1}BQ}A=P^{-1} B Q for some invertible Matrices P and Q

Smith normal form
The equivalence is the same as allowing invertible elementary row and column transformations
Finite-dimensional vector spaces over a field K

A and B are isomorphic as vector spaces

Kn{displaystyle K^{n}}K^{n}, n a non-negative integer



Algebra













Objects

A is equivalent to B if:
Normal form
Finitely generated R-modules with R a principal ideal domain

A and B are isomorphic as R-modules

Primary decomposition (up to reordering) or invariant factor decomposition


Geometry


In analytic geometry:



  • The equation of a line: Ax + By = C, with A2 + B2 = 1 and C ≥ 0

  • The equation of a circle: (x−h)2+(y−k)2=r2{displaystyle (x-h)^{2}+(y-k)^{2}=r^{2}}{displaystyle (x-h)^{2}+(y-k)^{2}=r^{2}}


By contrast, there are alternative forms for writing equations. For example, the equation of a line may be written as a linear equation in point-slope and slope-intercept form.


Convex polyhedra can be put into canonical form such that:



  • All faces are flat,

  • All edges are tangent to the unit sphere, and

  • The centroid of the polyhedron is at the origin.[2]



Integrable systems


Every differentiable manifold has a cotangent bundle. That bundle can always be endowed with a certain differential form, called the canonical one-form. This form gives the cotangent bundle the structure of a symplectic manifold. This allows vector fields on the manifold to be integrated by means of the Euler-Lagrange equations, or by means of Hamiltonian mechanics. Such systems of integrable differential equations are called integrable systems.



Dynamical systems


The study of dynamical systems overlaps with that of integrable systems; there one has the idea of a normal form (dynamical systems).



Three dimensional geometry


In the study of manifolds in three dimensions, one has the first fundamental form, the second fundamental form and the third fundamental form.



Functional analysis


















Objects

A is equivalent to B if:
Normal form

Hilbert spaces
If A and B are both separable Hilbert spaces of infinite dimension, then A and B are isometrically isomorphic.

2(I){displaystyle ell ^{2}(I)}ell^2(I) sequence spaces (up to exchanging the index set I with another index set of the same cardinality)
Commutative C∗{displaystyle C^{*}}C^*-algebras with unit

A and B are isomorphic as C∗{displaystyle C^{*}}C^*-algebras
The algebra C(X){displaystyle C(X)}C(X) of continuous functions on a compact Hausdorff space, up to homeomorphism of the base space.


Classical logic




  • Negation normal form

  • Conjunctive normal form

  • Disjunctive normal form

  • Algebraic normal form

  • Prenex normal form

  • Skolem normal form


  • Blake canonical form, also known as the complete sum of prime implicants, the complete sum, or the disjunctive prime form



Set theory



  • Cantor normal form of an ordinal number


Game theory


  • Normal form game


Proof theory


  • Normal form (natural deduction)


Rewriting systems



The symbolic manipulation of a formula from one form to another is called a "rewriting" of that formula. One can study the abstract properties of rewriting generic formulas; one need only specify a collection of rules by which formulas can be validly manipulated. These are the "rewriting rules", they form an abstract rewriting system. A common question is whether it is possible to bring some generic expression to a single, common form, the normal form. If different sequences of rewrites still result in the same form, then that form can be termed a normal form; and the rewrite is called confluent. It is not always possible to obtain a normal form.



Lambda calculus


  • A lambda term is in beta normal form if no beta reduction is possible; lambda calculus is a particular case of an abstract rewriting system. In the untyped lambda calculus, e.g., the term x.(xx)λx.(xx)){displaystyle (lambda x.(xx);lambda x.(xx))}{displaystyle (lambda x.(xx);lambda x.(xx))} doesn't have a normal form. In the typed lambda calculus, every well-formed term can be rewritten to its normal form.


Graph theory



In graph theory, a branch of mathematics, graph canonization is the problem finding a canonical form of a given graph G. A canonical form is a labeled graph Canon(G) that is isomorphic to G, such that every graph that is isomorphic to G has the same canonical form as G. Thus, from a solution to the graph canonization problem, one could also solve the problem of graph isomorphism: to test whether two graphs G and H are isomorphic, compute their canonical forms Canon(G) and Canon(H), and test whether these two canonical forms are identical.



Computing


In computing, the reduction of data to any kind of canonical form is commonly called data normalization.


For instance, database normalization is the process of organizing the fields and tables of a relational database to minimize redundancy and dependency.


In the field of software security, a common vulnerability is unchecked malicious input. The mitigation for this problem is proper input validation. Before input validation may be performed, the input must be normalized, i.e., eliminating encoding (for instance HTML encoding) and reducing the input data to a single common character set.


Other forms of data, typically associated with signal processing (including audio and imaging) or machine learning, can be normalized in order to provide a limited range of values.



See also



  • Canonicalization

  • Canonical basis

  • Canonical class

  • Normalization (disambiguation)

  • Standardization



Notes




  1. ^ The term 'canonization' is sometimes incorrectly used for this.


  2. ^ Ziegler, Günter M. (1995), Lectures on Polytopes, Graduate Texts in Mathematics, 152, Springer-Verlag, pp. 117–118, ISBN 0-387-94365-X.mw-parser-output cite.citation{font-style:inherit}.mw-parser-output .citation q{quotes:"""""""'""'"}.mw-parser-output .citation .cs1-lock-free a{background:url("//upload.wikimedia.org/wikipedia/commons/thumb/6/65/Lock-green.svg/9px-Lock-green.svg.png")no-repeat;background-position:right .1em center}.mw-parser-output .citation .cs1-lock-limited a,.mw-parser-output .citation .cs1-lock-registration a{background:url("//upload.wikimedia.org/wikipedia/commons/thumb/d/d6/Lock-gray-alt-2.svg/9px-Lock-gray-alt-2.svg.png")no-repeat;background-position:right .1em center}.mw-parser-output .citation .cs1-lock-subscription a{background:url("//upload.wikimedia.org/wikipedia/commons/thumb/a/aa/Lock-red-alt-2.svg/9px-Lock-red-alt-2.svg.png")no-repeat;background-position:right .1em center}.mw-parser-output .cs1-subscription,.mw-parser-output .cs1-registration{color:#555}.mw-parser-output .cs1-subscription span,.mw-parser-output .cs1-registration span{border-bottom:1px dotted;cursor:help}.mw-parser-output .cs1-ws-icon a{background:url("//upload.wikimedia.org/wikipedia/commons/thumb/4/4c/Wikisource-logo.svg/12px-Wikisource-logo.svg.png")no-repeat;background-position:right .1em center}.mw-parser-output code.cs1-code{color:inherit;background:inherit;border:inherit;padding:inherit}.mw-parser-output .cs1-hidden-error{display:none;font-size:100%}.mw-parser-output .cs1-visible-error{font-size:100%}.mw-parser-output .cs1-maint{display:none;color:#33aa33;margin-left:0.3em}.mw-parser-output .cs1-subscription,.mw-parser-output .cs1-registration,.mw-parser-output .cs1-format{font-size:95%}.mw-parser-output .cs1-kern-left,.mw-parser-output .cs1-kern-wl-left{padding-left:0.2em}.mw-parser-output .cs1-kern-right,.mw-parser-output .cs1-kern-wl-right{padding-right:0.2em}



References




  • Shilov, Georgi E. (1977), Silverman, Richard A., ed., Linear Algebra, Dover, ISBN 0-486-63518-X.


  • Hansen, Vagn Lundsgaard (2006), Functional Analysis: Entering Hilbert Space, World Scientific Publishing, ISBN 981-256-563-9.




Popular posts from this blog

Lambaréné

Chris Pine

Kashihara Line