Вы находитесь на странице: 1из 35

Contents

Articles
Morphism Homomorphism Automorphism Endomorphism Isomorphism Homeomorphism Holomorphic function Diffeomorphism Monomorphism Epimorphism 1 3 7 9 11 15 18 21 26 28

References
Article Sources and Contributors Image Sources, Licenses and Contributors 32 33

Article Licenses
License 34

Morphism

Morphism
In mathematics, a morphism is an abstraction derived from structure-preserving mappings between two mathematical structures. The notion of morphism recurs in much of contemporary mathematics. In set theory, morphisms are functions; in linear algebra, linear transformations; in group theory, group homomorphisms; in topology, continuous functions, and so on. The study of morphisms and of the structures (called objects) over which they are defined, is central to category theory. Much of the terminology of morphisms, as well as the intuition underlying them, comes from concrete categories, where the objects are simply sets with some additional structure, and morphisms are structure-preserving functions.

Definition
A category C consists of two classes, one of objects and the other of morphisms. There are two operations which are defined on every morphism, the domain (or source) and the codomain (or target). If a morphism f has domain X and codomain Y, we write f : X Y. Thus a morphism is represented by an arrow from its domain to its codomain. The collection of all morphisms from X to Y is denoted homC(X,Y) or simply hom(X, Y) and called the hom-set between X and Y. Some authors write MorC(X,Y) or Mor(X, Y). Note that the term hom-set is a bit of a misnomer as the collection of morphisms is not required to be a set. For every three objects X, Y, and Z, there exists a binary operation hom(X, Y) hom(Y, Z) hom(X, Z) called composition. The composite of f : X Y and g : Y Z is written g f or gf. The composition of morphisms is often represented by a commutative diagram. For example,

Morphisms satisfy two axioms: Identity: for every object X, there exists a morphism idX : X X called the identity morphism on X, such that for every morphism f : A B we have idB f = f = f idA. Associativity: h (g f) = (h g) f whenever the operations are defined. When C is a concrete category, the identity morphism is just the identity function, and composition is just the ordinary composition of functions. Associativity then follows, because the composition of functions is associative. Note that the domain and codomain are in fact part of the information determining a morphism. For example, in the category of sets, where morphisms are functions, two functions may be identical as sets of ordered pairs (may have the same range), while having different codomains. The two functions are distinct from the viewpoint of category theory. Thus many authors require that the hom-classes hom(X, Y) be disjoint. In practice, this is not a problem because if this disjointness does not hold, it can be assured by appending the domain and codomain to the morphisms, (say, as the second and third components of an ordered triple).

Morphism

Some specific morphisms


Monomorphism: f : X Y is called a monomorphism if f g1 = f g2 implies g1 = g2 for all morphisms g1, g2 : Z X. It is also called a mono or a monic.[1] The morphism f has a left inverse if there is a morphism g:Y X such that g f = idX. The left inverse g is also called a retraction of f.[1] Morphisms with left inverses are always monomorphisms, but the converse is not always true in every category; a monomorphism may fail to have a left-inverse. A split monomorphism h : X Y is a monomorphism having a left inverse g : Y X, so that g h = idX. Thus h g : Y Y is idempotent, so that (h g)2 = h g. In concrete categories, a function that has a left inverse is injective. Thus in concrete categories, monomorphisms are often, but not always, injective. The condition of being an injection is stronger than that of being a monomorphism, but weaker than that of being a split monomorphism. Epimorphism: Dually, f : X Y is called an epimorphism if g1 f = g2 f implies g1 = g2 for all morphisms g1, g2 : Y Z. It is also called an epi or an epic.[1] The morphism f has a right-inverse if there is a morphism g : Y X such that f g = idY. The right inverse g is also called a section of f.[1] Morphisms having a right inverse are always epimorphisms, but the converse is not always true in every category, as an epimorphism may fail to have a right inverse. A split epimorphism is an epimorphism having a right inverse. Note that if a monomorphism f splits with left-inverse g, then g is a split epimorphism with right-inverse f. In concrete categories, a function that has a right inverse is surjective. Thus in concrete categories, epimorphisms are often, but not always, surjective. The condition of being a surjection is stronger than that of being an epimorphism, but weaker than that of being a split epimorphism. In the category of sets, every surjection has a section, a result equivalent to the axiom of choice. A bimorphism is a morphism that is both an epimorphism and a monomorphism. Isomorphism: f : X Y is called an isomorphism if there exists a morphism g : Y X such that f g = idY and g f = idX. If a morphism has both left-inverse and right-inverse, then the two inverses are equal, so f is an isomorphism, and g is called simply the inverse of f. Inverse morphisms, if they exist, are unique. The inverse g is also an isomorphism with inverse f. Two objects with an isomorphism between them are said to be isomorphic or equivalent. Note that while every isomorphism is a bimorphism, a bimorphism is not necessarily an isomorphism. For example, in the category of commutative rings the inclusion Z Q is a bimorphism, which is not an isomorphism. However, any morphism that is both an epimorphism and a split monomorphism, or both a monomorphism and a split epimorphism, must be an isomorphism. A category, such as Set, in which every bimorphism is an isomorphism is known as a balanced category. Endomorphism: f : X X is an endomorphism of X. A split endomorphism is an idempotent endomorphism f if f admits a decomposition f = h g with g h = id. In particular, the Karoubi envelope of a category splits every idempotent morphism. An automorphism is a morphism that is both an endomorphism and an isomorphism.

Examples
In the concrete categories studied in universal algebra (groups, rings, modules, etc.), morphisms are usually homomorphisms. Likewise, the notions of automorphism, endomorphism, epimorphism, homeomorphism, isomorphism, and monomorphism all find use in universal algebra. In the category of topological spaces, morphisms are continuous functions and isomorphisms are called homeomorphisms. In the category of smooth manifolds, morphisms are smooth functions and isomorphisms are called diffeomorphisms.

Morphism In the category of small categories, functors can be thought of as morphisms. In a functor category, the morphisms are natural transformations. For more examples, see the entry category theory.

Notes
[1] Jacobson (2009), p. 15.

References
Jacobson, Nathan (2009), Basic algebra, 2 (2nd ed.), Dover, ISBN978-0-486-47187-7.

External links
Category (http://planetmath.org/?op=getobj&from=objects&id=965), PlanetMath.org. TypesOfMorphisms (http://planetmath.org/?op=getobj&from=objects&id=8114), PlanetMath.org.

Homomorphism
In abstract algebra, a homomorphism is a structure-preserving map between two algebraic structures (such as groups, rings, or vector spaces). The word homomorphism comes from the ancient Greek language: (homos) meaning "same" and (morphe) meaning "shape". Isomorphisms, automorphisms, and endomorphisms are all types of homomorphism.

Definition and illustration


Definition
The definition of homomorphism depends on the type of algebraic structure under consideration. Particular definitions of homomorphism include the following: A group homomorphism is a homomorphism between two groups. A ring homomorphism is a homomorphism between two rings. A linear map is a homomorphism between two vector spaces. An algebra homomorphism is a homomorphism between two algebras. A functor is a homomorphism between two categories.

The common theme is that a homomorphism is a function between two algebraic objects that respects the algebraic structure. For example, a group is an algebraic object consisting of a set together with a single binary operation, satisfying certain axioms. If and are groups, a homomorphism from to is a function : such that for any elements g1,g2G. When an algebraic structure includes more than one operation, homomorphisms are required to preserve each operation. For example, a ring possesses both addition and multiplication, and a homomorphism from the ring to the ring is a function such that

for any elements r and s of the domain ring. The notion of a homomorphism can be given a formal definition in the context of universal algebra, a field which studies ideas common to all algebraic structures. In this setting, a homomorphism :AB is a function between two

Homomorphism algebraic structures of the same type such that

for each n-ary operation and for all elements a1,...,anA.

Basic examples
The real numbers are a ring, having both addition and multiplication. The set of all 22matrices is also a ring, under matrix addition and matrix multiplication. If we define a function between these rings as follows:

where r is a real number. Then is a homomorphism of rings, since preserves both addition:

and multiplication:

For another example, the nonzero complex numbers form a group under the operation of multiplication, as do the nonzero real numbers. (Zero must be excluded from both groups since it does not have a multiplicative inverse, which is required for elements of a group.) Define a function from the nonzero complex numbers to the nonzero real numbers by

That is, (z) is the absolute value (or modulus) of the complex number z. Then is a homomorphism of groups, since it preserves multiplication:

Note that cannot be extended to a homomorphism of rings (from the complex numbers to the real numbers), since it does not preserve addition:

Informal discussion
Because abstract algebra studies sets endowed with operations that generate interesting structure or properties on the set, functions which preserve the operations are especially important. These functions are known as homomorphisms. For example, consider the natural numbers with addition as the operation. A function which preserves addition should have this property: f(a + b) = f(a) + f(b). For example, f(x) = 3x is one such homomorphism, since f(a + b) = 3(a + b) = 3a + 3b = f(a) + f(b). Note that this homomorphism maps the natural numbers back into themselves. Homomorphisms do not have to map between sets which have the same operations. For example, operation-preserving functions exist between the set of real numbers with addition and the set of positive real numbers with multiplication. A function which preserves operation should have this property: f(a + b) = f(a) * f(b), since addition is the operation in the first set and multiplication is the operation in the second. Given the laws of exponents, f(x) = ex satisfies this condition : 2 + 3 = 5 translates into e2 * e3 = e5. If we are considering multiple operations on a set, then all operations must be preserved for a function to be considered as a homomorphism. Even though the set may be the same, the same function might be a homomorphism, say, in group theory (sets with a single operation) but not in ring theory (sets with two related operations), because it fails to preserve the additional operation that ring theory considers.

Homomorphism

Relation to category theory


Since homomorphisms are morphisms, the following specific kinds of morphisms defined in any category are defined for homomorphisms as well. However, the definitions in category theory are somewhat technical. In the important special case of module homomorphisms, and for some other classes of homomorphisms, there are much simpler descriptions, as follows: An isomorphism is a bijective homomorphism. An epimorphism (sometimes called a cover) is a surjective homomorphism. A monomorphism (sometimes called an embedding or extension) is an injective homomorphism. An endomorphism is a homomorphism from an object to itself. An automorphism is an endomorphism which is also an isomorphism, i.e., an isomorphism from an object to itself. These descriptions may be used in order to derive several interesting properties. For instance, since a function is bijective if and only if it is both injective and surjective, a module homomorphism is an isomorphism if and only if it is both a monomorphism and an epimorphism. For endomorphisms and automorphisms, the descriptions above coincide with the category theoretic definitions; the first three descriptions do not. For instance, the precise definition for a homomorphism f to be iso is not only that it is bijective, and thus has an inverse f-1, but also that this inverse is a homomorphism, too. This has the important consequence that two objects are completely indistinguishable as far as the structure in question is concerned, if there is an isomorphism between them. Two such objects are said to be isomorphic. Actually, in the algebraic setting (at least within the context of universal algebra) this extra condition on isomorphisms is automatically satisfied. However, the same is not true for epimorphisms; for instance, the inclusion of Z as a (unitary) subring of Q is not surjective, but an epimorphic ring homomorphism.[1] This inclusion thus also is an example of a ring homomorphism which is both mono and epi, but not iso.

Relationships between different kinds of module homomorphisms. H = set of Homomorphisms, M = set of Monomorphisms, P = set of Epimorphisms, S = set of Isomorphisms, N = set of Endomorphism, A = set of Automorphisms. Notice that: M P = S, S N = A,

Homomorphism (M N) \ A and (P N) \ A contain only homomorphisms from some infinite modules to themselves.

Kernel of a homomorphism
Any homomorphism f : X Y defines an equivalence relation ~ on X by a~ b if and only if f(a)= f(b). The relation ~ is called the kernel of f. It is a congruence relation on X. The quotient set X/~ can then be given an object-structure in a natural way, i.e. [x] * [y] = [x * y]. In that case the image of X in Y under the homomorphism f is necessarily isomorphic to X/~; this fact is one of the isomorphism theorems. Note in some cases (e.g. groups or rings), a single equivalence class K suffices to specify the structure of the quotient; so we can write it X/K. (X/K is usually read as "X mod K".) Also in these cases, it is K, rather than ~, that is called the kernel of f (cf. normal subgroup).

Homomorphisms of relational structures


In model theory, the notion of an algebraic structure is generalized to structures involving both operations and relations. Let L be a signature consisting of function and relation symbols, and A, B be two L-structures. Then a homomorphism from A to B is a mapping h from the domain of A to the domain of B such that h(FA(a1,,an)) = FB(h(a1),,h(an)) for each n-ary function symbol F in L, RA(a1,,an) implies RB(h(a1),,h(an)) for each n-ary relation symbol R in L. In the special case with just one binary relation, we obtain the notion of a graph homomorphism.

Homomorphisms and e-free homomorphisms in formal language theory


Homomorphisms are also used in the study of formal languages[2] (although within this context, often they are briefly referred to as morphisms[3]). Given alphabets and , a function h : such that for all u and v in is called a homomorphism (or simply morphism) on and for all in .[4] Let e denote the empty word. If h is a homomorphism on , then h is called an e-free the set

homomorphism. This type of homomorphism can be thought of as (and is equivalent to) a monoid homomorphism where of all words over a finite alphabet and the empty word as the identity. is a monoid (in fact it is the free monoid on

) with operation concatenation

References
[1] Exercise 4 in section I.5, in Saunders Mac Lane, Categories for the Working Mathematician, ISBN 0-387-90036-5 [2] Seymour Ginsburg, Algebraic and automata theoretic properties of formal languages, North-Holland, 1975, ISBN 0-7204-2506-9. [3] T. Harju, J. Karhumki, Morphisms in Handbook of Formal Languages, Volume I, edited by G. Rozenberg, A. Salomaa, Springer, 1997, ISBN 3-540-61486-9. [4] In homomorphisms on formal languages, the * operation is the Kleene star operation. The and are both concatenation, commonly denoted by juxtaposition.

A monograph available free online: Burris, Stanley N., and H.P. Sankappanavar, H. P., 1981. A Course in Universal Algebra. (http://www.thoralf. uwaterloo.ca/htdocs/ualg.html) Springer-Verlag. ISBN 3-540-90578-2.

Automorphism

Automorphism
In mathematics, an automorphism is an isomorphism from a mathematical object to itself. It is, in some sense, a symmetry of the object, and a way of mapping the object to itself while preserving all of its structure. The set of all automorphisms of an object forms a group, called the automorphism group. It is, loosely speaking, the symmetry group of the object.

Definition
The exact definition of an automorphism depends on the type of "mathematical object" in question and what, precisely, constitutes an "isomorphism" of that object. The most general setting in which these words have meaning is an abstract branch of mathematics called category theory. Category theory deals with abstract objects and morphisms between those objects. In category theory, an automorphism is an endomorphism (i.e. a morphism from an object to itself) which is also an isomorphism (in the categorical sense of the word). This is a very abstract definition since, in category theory, morphisms aren't necessarily functions and objects aren't necessarily sets. In most concrete settings, however, the objects will be sets with some additional structure and the morphisms will be functions preserving that structure. In the context of abstract algebra, for example, a mathematical object is an algebraic structure such as a group, ring, or vector space. An isomorphism is simply a bijective homomorphism. (The definition of a homomorphism depends on the type of algebraic structure; see, for example: group homomorphism, ring homomorphism, and linear operator). The identity morphism (identity mapping) is called the trivial automorphism in some contexts. Respectively, other (non-identity) automorphisms are called nontrivial automorphisms.

Automorphism group
If the automorphisms of an object X form a set (instead of a proper class), then they form a group under composition of morphisms. This group is called the automorphism group of X. That this is indeed a group is simple to see: Closure: composition of two endomorphisms is another endomorphism. Associativity: composition of morphisms is always associative. Identity: the identity is the identity morphism from an object to itself which exists by definition. Inverses: by definition every isomorphism has an inverse which is also an isomorphism, and since the inverse is also an endomorphism of the same object it is an automorphism.

The automorphism group of an object X in a category C is denoted AutC(X), or simply Aut(X) if the category is clear from context.

Examples
In set theory, an automorphism of a set X is an arbitrary permutation of the elements of X. The automorphism group of X is also called the symmetric group on X. In elementary arithmetic, the set of integers, Z, considered as a group under addition, has a unique nontrivial automorphism: negation. Considered as a ring, however, it has only the trivial automorphism. Generally speaking, negation is an automorphism of any abelian group, but not of a ring or field. A group automorphism is a group isomorphism from a group to itself. Informally, it is a permutation of the group elements such that the structure remains unchanged. For every group G there is a natural group homomorphism G Aut(G) whose image is the group Inn(G) of inner automorphisms and whose kernel is the center of G. Thus, if

Automorphism G has trivial center it can be embedded into its own automorphism group.[1] In linear algebra, an endomorphism of a vector space V is a linear operator V V. An automorphism is an invertible linear operator on V. When the vector space is finite-dimensional, the automorphism group of V is the same as the general linear group, GL(V). A field automorphism is a bijective ring homomorphism from a field to itself. In the cases of the rational numbers (Q) and the real numbers (R) there are no nontrivial field automorphisms. Some subfields of R have nontrivial field automorphisms, which however do not extend to all of R (because they cannot preserve the property of a number having a square root in R). In the case of the complex numbers, C, there is a unique nontrivial automorphism that sends R into R: complex conjugation, but there are infinitely (uncountably) many "wild" automorphisms (assuming the axiom of choice).[2] Field automorphisms are important to the theory of field extensions, in particular Galois extensions. In the case of a Galois extension L/K the subgroup of all automorphisms of L fixing K pointwise is called the Galois group of the extension. In graph theory an automorphism of a graph is a permutation of the nodes that preserves edges and non-edges. In particular, if two nodes are joined by an edge, so are their images under the permutation. For relations, see relation-preserving automorphism.

In order theory, see order automorphism. In geometry, an automorphism may be called a motion of the space. Specialized terminology is also used: In metric geometry an automorphism is a self-isometry. The automorphism group is also called the isometry group. In the category of Riemann surfaces, an automorphism is a bijective biholomorphic map (also called a conformal map), from a surface to itself. For example, the automorphisms of the Riemann sphere are Mbius transformations. An automorphism of a differentiable manifold M is a diffeomorphism from M to itself. The automorphism group is sometimes denoted Diff(M). In topology, morphisms between topological spaces are called continuous maps, and an automorphism of a topological space is a homeomorphism of the space to itself, or self-homeomorphism (see homeomorphism group). In this example it is not sufficient for a morphism to be bijective to be an isomorphism.

History
One of the earliest group automorphisms (automorphism of a group, not simply a group of automorphisms of points) was given by the Irish mathematician William Rowan Hamilton in 1856, in his Icosian Calculus, where he discovered an order two automorphism,[3] writing: so that is a new fifth root of unity, connected with the former fifth root by relations of perfect reciprocity.

Inner and outer automorphisms


In some categoriesnotably groups, rings, and Lie algebrasit is possible to separate automorphisms into two types, called "inner" and "outer" automorphisms. In the case of groups, the inner automorphisms are the conjugations by the elements of the group itself. For each element a of a group G, conjugation by a is the operation a : GG given by a(g) = aga1 (or a1ga; usage varies). One can easily check that conjugation by a is a group automorphism. The inner automorphisms form a normal subgroup of Aut(G), denoted by Inn(G); this is called Goursat's lemma. The other automorphisms are called outer automorphisms. The quotient group Aut(G)/Inn(G) is usually denoted by Out(G); the non-trivial elements are the cosets that contain the outer automorphisms.

Automorphism The same definition holds in any unital ring or algebra where a is any invertible element. For Lie algebras the definition is slightly different.

References
[1] PJ Pahl, R Damrath (2001). "7.5.5 Automorphisms" (http:/ / books. google. com/ ?id=kvoaoWOfqd8C& pg=PA376). Mathematical foundations of computational engineering (Felix Pahl translation ed.). Springer. p.376. ISBN3-540-67995-2. . [2] Yale, Paul B. (May 1966). "Automorphisms of the Complex Numbers" (http:/ / mathdl. maa. org/ images/ upload_library/ 22/ Ford/ PaulBYale. pdf). Mathematics Magazine 39 (3): 135141. doi:10.2307/2689301. JSTOR2689301. . [3] Sir William Rowan Hamilton (1856). "Memorandum respecting a new System of Roots of Unity" (http:/ / www. maths. tcd. ie/ pub/ HistMath/ People/ Hamilton/ Icosian/ NewSys. pdf). Philosophical Magazine 12: 446. .

External links
Weisstein, Eric W., " Automorphism (http://mathworld.wolfram.com/Automorphism.html)" from MathWorld.

Endomorphism
In mathematics, an endomorphism is a morphism (or homomorphism) from a mathematical object to itself. For example, an endomorphism of a vector space V is a linear map :VV, and an endomorphism of a group G is a group homomorphism :GG. In general, we can talk about endomorphisms in any category. In the category of sets, endomorphisms are simply functions from a set S into itself. In any category, the composition of any two endomorphisms of X is again an endomorphism of X. It follows that the set of all endomorphisms of X forms a monoid, denoted End(X) (or EndC(X) to emphasize the category C). An invertible endomorphism of X is called an automorphism. The set of all automorphisms is a subset of End(X) with a group structure, called the automorphism group of X and denoted Aut(X). In the following diagram, the arrows denote implication:
automorphism isomorphism

endomorphism

(homo)morphism

Any two endomorphisms of an abelian group A can be added together by the rule (+g)(a)=(a)+g(a). Under this addition, the endomorphisms of an abelian group form a ring (the endomorphism ring). For example, the set of endomorphisms of Zn is the ring of all nn matrices with integer entries. The endomorphisms of a vector space or module also form a ring, as do the endomorphisms of any object in a preadditive category. The endomorphisms of a nonabelian group generate an algebraic structure known as a nearring. Every ring with one is the endomorphism ring of its regular module, and so is a subring of an endomorphism ring of an abelian group,[1] however there are rings which are not the endomorphism ring of any abelian group.

Endomorphism

10

Operator theory
In any concrete category, especially for vector spaces, endomorphisms are maps from a set into itself, and may be interpreted as unary operators on that set, acting on the elements, and allowing to define the notion of orbits of elements, etc. Depending on the additional structure defined for the category at hand (topology, metric, ...), such operators can have properties like continuity, boundedness, and so on. More details should be found in the article about operator theory.

Endofunctions in mathematics
In mathematics, an endofunction is a function whose codomain is equal to its domain. A homomorphic endofunction is an endomorphism. Let S be an arbitrary set. Among endofunctions on S one finds permutations of S and constant functions associating to each a given . Every permutation of S has the codomain equal to its domain and is bijective and invertible. A constant function on S, if S has more than 1 element, has a codomain that is a proper subset of its domain, is not bijective (and non invertible). The function associating to each natural integer n the floor of n/2 has its codomain equal to its domain and is not invertible. Finite endofunctions are equivalent to monogeneous digraphs, i.e. digraphs having all nodes with outdegree equal to 1, and can be easily described. For sets of size n, there are n^n endofunctions on the set. Particular bijective endofunctions are the involutions, i.e. the functions coinciding with their inverses.

Notes
[1] Jacobson (2009), p. 162, Theorem 3.2.

References
Jacobson, Nathan (2009), Basic algebra, 1 (2nd ed.), Dover, ISBN978-0-486-47189-1

External links
Endomorphism (http://planetmath.org/?op=getobj&from=objects&id=7462), PlanetMath.org.

Isomorphism

11

Isomorphism
In abstract algebra, an isomorphism[1] is a bijective homomorphism.[2] Two mathematical structures are said to be isomorphic if there is an isomorphism between them. In category theory, an isomorphism is a morphism f: X Y in a category for which there exists an "inverse" f 1: Y X, with the property that both f 1f = idX and f f 1 = idY.[3]

Purpose
Isomorphisms are studied in mathematics in order to extend insights from one phenomenon to others: if two objects are isomorphic, then any property that is preserved by an isomorphism and that is true of one of the objects, is also true of the other. If an isomorphism can be found from a relatively unknown part of mathematics into some well studied division of mathematics, where many theorems are already proved, and many methods are already available to find answers, then the function can be used to map whole problems out of unfamiliar territory over to "solid ground" where the problem is easier to understand and work with.

Practical examples
The following are examples of isomorphisms from ordinary algebra. Consider the logarithm function: For any fixed base b, the logarithm function logb maps from the positive real numbers R+ onto the real numbers R; formally:

This mapping is one-to-one and onto, that is, it is a bijection from the domain to the codomain of the logarithm function. In addition to being an isomorphism of sets, the logarithm function also preserves certain operations. Specifically, consider the group (R+,) of positive real numbers under ordinary multiplication. The logarithm function obeys the following identity:

But the real numbers under addition also form a group. So the logarithm function is in fact a group isomorphism from the group (R+,) to the group (R,+). Logarithms can therefore be used to simplify multiplication of positive real numbers. By working with logarithms, multiplication of positive real numbers is replaced by addition of logs. This way it is possible to multiply real numbers using a ruler and a table of logarithms, or using a slide rule with a logarithmic scale. Consider the group (Z6, +), the integers from 0 to 5 with addition modulo 6. Also consider the group (Z2 Z3, +), the ordered pairs where the x coordinates can be 0 or 1, and the y coordinates can be 0, 1, or 2, where addition in the x-coordinate is modulo 2 and addition in the y-coordinate is modulo 3. These structures are isomorphic under addition, if you identify them using the following scheme: (0,0) 0 (1,1) 1 (0,2) 2 (1,0) 3 (0,1) 4 (1,2) 5 or in general (a,b) (3a + 4b) mod 6. For example note that (1,1) + (1,0) = (0,1), which translates in the other system as 1 + 3 = 4. Even though these two groups "look" different in that the sets contain different elements, they are indeed isomorphic: their structures are exactly the same. More generally, the direct product of two cyclic

Isomorphism groups Zm and Zn is isomorphic to Zmn if and only if m and n are coprime.

12

Abstract examples
A relation-preserving isomorphism
If one object consists of a set X with a binary relation R and the other object consists of a set Y with a binary relation S then an isomorphism from X to Y is a bijective function : X Y such that[4]:

S is reflexive, irreflexive, symmetric, antisymmetric, asymmetric, transitive, total, trichotomous, a partial order, total order, strict weak order, total preorder (weak order), an equivalence relation, or a relation with any other special properties, if and only if R is. For example, R is an ordering and S an ordering Y such that , then an isomorphism from X to Y is a bijective function : X

Such an isomorphism is called an order isomorphism or (less commonly) an isotone isomorphism. If X = Y, then this is a relation-preserving automorphism.

An operation-preserving isomorphism
Suppose that on these sets X and Y, there are two binary operations and that happen to constitute the groups (X, ) and (Y, ). Note that the operators operate on elements from the domain and range, respectively, of the "one-to-one" and "onto" function . There is an isomorphism from X to Y if the bijective function : X Y happens to produce results, that sets up a correspondence between the operator and the operator .

for all u, v in X.

Applications
In abstract algebra, two basic isomorphisms are defined: Group isomorphism, an isomorphism between groups Ring isomorphism, an isomorphism between rings. (Note that isomorphisms between fields are actually ring isomorphisms) Just as the automorphisms of an algebraic structure form a group, the isomorphisms between two algebras sharing a common structure form a heap. Letting a particular isomorphism identify the two structures turns this heap into a group. In mathematical analysis, the Laplace transform is an isomorphism mapping hard differential equations into easier algebraic equations. In category theory, Iet the category C consist of two classes, one of objects and the other of morphisms. Then a general definition of isomorphism that covers the previous and many other cases is: an isomorphism is a morphism : a b that has an inverse, i.e. there exists a morphism g : b a with g = 1b and g = 1a. For example, a bijective linear map is an isomorphism between vector spaces, and a bijective continuous function whose inverse is also continuous is an isomorphism between topological spaces, called a homeomorphism. In graph theory, an isomorphism between two graphs G and H is a bijective map f from the vertices of G to the vertices of H that preserves the "edge structure" in the sense that there is an edge from vertex u to vertex v in G if and only if there is an edge from (u) to (v) in H. See graph isomorphism.

Isomorphism In mathematical analysis, an isomorphism between two Hilbert spaces is a bijection preserving addition, scalar multiplication, and inner product. In early theories of logical atomism, the formal relationship between facts and true propositions was theorized by Bertrand Russell and Ludwig Wittgenstein to be isomorphic. An example of this line of thinking can be found in Russell's Introduction to Mathematical Philosophy. In cybernetics, the Good Regulator or Conant-Ashby theorem is stated "Every Good Regulator of a system must be a model of that system". Whether regulated or self-regulating an isomorphism is required between regulator part and the processing part of the system.

13

Relation with equality


In certain areas of mathematics, notably category theory, it is valuable to distinguish between equality on the one hand and isomorphism on the other.[5] Equality is when two objects are exactly the same, and everything that's true about one object is true about the other, while an isomorphism implies everything that's true about one object's structure is true about the other's. For example, the sets and are equal; they are merely different presentationsthe first an intensional one (in set builder notation), and the second extensional (by explicit enumeration)of the same subset of the integers. By contrast, the sets {A,B,C} and {1,2,3} are not equal the first has elements that are letters, while the second has elements that are numbers. These are isomorphic as sets, since finite sets are determined up to isomorphism by their cardinality (number of elements) and these both have three elements, but there are many choices of isomorphism one isomorphism is while another is and no one isomorphism is intrinsically better than any other.[6][7] On this view and in this sense, these two sets are not equal because one cannot consider them identical: one can choose an isomorphism between them, but that is a weaker claim than identityand valid only in the context of the chosen isomorphism. Sometimes the isomorphisms can seem obvious and compelling, but are still not equalities. As a simple example, the genealogical relationships among Joe, John, and Bobby Kennedy are, in a real sense, the same as those among the American football quarterbacks in the Manning family: Archie, Peyton, and Eli. The father-son pairings and the elder-brother-younger-brother pairings correspond perfectly. That similarity between the two family structures illustrates the origin of the word isomorphism (Greek iso-, "same," and -morph, "form" or "shape"). But because the Kennedys are not the same people as the Mannings, the two genealogical structures are merely isomorphic and not equal. Another example is more formal and more directly illustrates the motivation for distinguishing equality from isomorphism: the distinction between a finite-dimensional vector space V and its dual space V* = { : V K} of linear maps from V to its field of scalars K. These spaces have the same dimension, and thus are isomorphic as abstract vector spaces (since algebraically, vector spaces are classified by dimension, just as sets are classified by cardinality), but there is no "natural" choice of isomorphism . If one chooses a basis for V, then this yields an isomorphism: For all u. v V, . This corresponds to transforming a column vector (element of V) to a row vector (element of V*) by transpose, but a different choice of basis gives a different isomorphism: the isomorphism "depends on the choice of basis". More subtly, there is a map from a vector space V to its double dual V** = { x : V* K} that does not depend on the choice of basis: For all v V and V*, .

Isomorphism This leads to a third notion, that of a natural isomorphism: while V and V** are different sets, there is a "natural" choice of isomorphism between them. This intuitive notion of "an isomorphism that does not depend on an arbitrary choice" is formalized in the notion of a natural transformation; briefly, that one may consistently identify, or more generally map from, a vector space to its double dual, , for any vector space in a consistent way. Formalizing this intuition is a motivation for the development of category theory. If one wishes to draw a distinction between an arbitrary isomorphism (one that depends on a choice) and a natural isomorphism (one that can be done consistently), one may write for an unnatural isomorphism and for a natural isomorphism, as in V V* and V V**. This convention is not universally followed, and authors who wish to distinguish between unnatural isomorphisms and natural isomorphisms will generally explicitly state the distinction. Generally, saying that two objects are equal is reserved for when there is a notion of a larger (ambient) space that these objects live in. Most often, one speaks of equality of two subsets of a given set (as in the integer set example above), but not of two objects abstractly presented. For example, the 2-dimensional unit sphere in 3-dimensional space and the Riemann sphere which can be presented as the one-point compactification of the complex plane C {} or as the complex projective line (a quotient space)

14

are three different descriptions for a mathematical object, all of which are isomorphic, but not equal because they are not all subsets of a single space: the first is a subset of R3, the second is C R2[8] plus an additional point, and the third is a subquotient of C2 In the context of category theory, objects are usually at most isomorphic indeed, a motivation for the development of category theory was showing that different constructions in homology theory yielded equivalent (isomorphic) groups. Given maps between two objects X and Y, however, one asks if they are equal or not (they are both elements of the set Hom(X,Y), hence equality is the proper relationship), particularly in commutative diagrams.

Notes
[1] From the Greek: isos "equal", and morphe "shape" [2] Buchmann, Johannes (2004). Introduction to cryptography (http:/ / books. google. com/ books?id=JEpVP9FC4o4C& pg=PA54). Springer. p.54. ISBN9780387207568. . [3] Awodey, Steve (2006). "Isomorphisms" (http:/ / books. google. com/ books?id=IK_sIDI2TCwC& pg=PA11). Category theory. Oxford University Press. p.11. ISBN9780198568612. . [4] Vinberg, rnest Borisovich (2003). A Course in Algebra (http:/ / books. google. com/ books?id=kd24d3mwaecC& pg=PA3). American Mathematical Society. p.3. ISBN9780821834138. . [5] (Mazur 2007) [6] The careful reader may note that A, B, C have a conventional order, namely alphabetical order, and similarly 1, 2, 3 have the order from the integers, and thus one particular isomorphism is "natural", namely

.
More formally, as sets these are isomorphic, but not naturally isomorphic (there are multiple choices of isomorphism), while as ordered sets they are naturally isomorphic (there is a unique isomorphism, given above), since finite total orders are uniquely determined up to unique isomorphism by cardinality.</math>

This intuition can be formalized by saying that any two finite totally ordered sets of the same cardinality have a natural isomorphism, the one that sends the least element of the first to the least element of the second, the least element of what remains in the first to the least element of what remains in the second, and so forth, but in general, pairs of sets of a given finite cardinality are not naturally isomorphic because there is more than one choice of map except if the cardinality is 0 or 1, where there is a unique choice.
[7] In fact, there are precisely different isomorphisms between two sets with three elements. This is equal to the number of automorphisms of a given three-element set (which in turn is equal to the order of the symmetric group on three letters), and more generally one has that the set of isomorphisms between two objects, denoted is a torsor for the automorphism group of A,

Isomorphism
and also a torsor for the automorphism group of B. In fact, automorphisms of an object are a key reason to be concerned with the distinction between isomorphism and equality, as demonstrated in the effect of change of basis on the identification of a vector space with its dual or with its double dual, as elaborated in the sequel. [8] Being precise, the identification of the complex numbers with the real plane,

15

depends on a choice of

one can just as easily choose

, which yields a different identification formally,

complex conjugation is an automorphism but in practice one often assumes that one has made such an identification.

References Further reading


Mazur, Barry (12 June 2007), When is one thing equal to some other thing? (http://www.math.harvard.edu/ ~mazur/preprints/when_is_one.pdf)

External links
Isomorphism (http://planetmath.org/?op=getobj&amp;from=objects&amp;id=1936), PlanetMath.org. Weisstein, Eric W., " Isomorphism (http://mathworld.wolfram.com/Isomorphism.html)" from MathWorld.

Homeomorphism
In the mathematical field of topology, a homeomorphism or 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 spacesthat 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. Roughly speaking, a topological space is a geometric object, and the homeomorphism is a continuous stretching and bending of the object into a new shape. Thus, a square and a circle are homeomorphic to each other, but a sphere and a donut are not. An often-repeated mathematical joke is that topologists can't tell their coffee cup from their donut,[1] since a sufficiently pliable donut could be reshaped to the form of a coffee cup by creating a dimple and progressively enlarging it, while shrinking the hole into a handle.

A continuous deformation between a coffee mug and a donut illustrating that they are homeomorphic. But there need not be a continuous deformation for two spaces to be homeomorphic only a continuous mapping with a continuous inverse.

Topology is the study of those properties of objects that do not change when homeomorphisms are applied. As Henri Poincar famously said, mathematics is not the study of objects, but instead, the relations (isomorphisms for instance) between them.[2]

Homeomorphism

16

Definition
A function f: X Y between two topological spaces (X, TX) and (Y, TY) is called a homeomorphism if it has the following properties: f is a bijection (one-to-one and onto), f is continuous, the inverse function f 1 is continuous (f is an open mapping). A function with these three properties is sometimes called bicontinuous. If such a function exists, we say X and Y are homeomorphic. A self-homeomorphism is a homeomorphism of a topological space and itself. The homeomorphisms form an equivalence relation on the class of all topological spaces. The resulting equivalence classes are called homeomorphism classes.

Examples
The unit 2-disc D2 and the unit square in R2 are homeomorphic. The open interval (a, b) is homeomorphic to the real numbers R for any a < b. The product space S1 S1 and the two-dimensional torus are homeomorphic. Every uniform isomorphism and isometric isomorphism is a homeomorphism. The 2-sphere with a single point removed is homeomorphic to the set of all points in R2 (a 2-dimensional plane). Let A be a commutative ring with unity and let S be a multiplicative subset of A. Then Spec(AS) is homeomorphic to {p Spec(A) : p S = }. Rm and Rn are not homeomorphic for m n. The Euclidean real line is not homeomorphic to the unit circle as a subspace of R2 as the unit circle is compact as a subspace of Euclidean R2 but the real line is not compact.

A trefoil knot is homeomorphic to a circle. Continuous mappings are not always realizable as deformations. Here the knot has been thickened to make the image understandable.

Notes
The third requirement, that f 1 be continuous, is essential. Consider for instance the function f: [0, 2) S1 defined by f() = (cos(), sin()). This function is bijective and continuous, but not a homeomorphism (S1 is compact but [0, 2) is not). Homeomorphisms are the isomorphisms in the category of topological spaces. As such, the composition of two homeomorphisms is again a homeomorphism, and the set of all self-homeomorphisms X X forms a group, called the homeomorphism group of X, often denoted Homeo(X); this group can be given a topology, such as the compact-open topology, making it a topological group. For some purposes, the homeomorphism group happens to be too big, but by means of the isotopy relation, one can reduce this group to the mapping class group. Similarly, as usual in category theory, given two spaces that are homeomorphic, the space of homeomorphisms between them, Homeo(X, Y), is a torsor for the homeomorphism groups Homeo(X) and Homeo(Y), and given a specific homeomorphism between X and Y, all three sets are identified.

Homeomorphism

17

Properties
Two homeomorphic spaces share the same topological properties. For example, if one of them is compact, then the other is as well; if one of them is connected, then the other is as well; if one of them is Hausdorff, then the other is as well; their homotopy & homology groups will coincide. Note however that this does not extend to properties defined via a metric; there are metric spaces that are homeomorphic even though one of them is complete and the other is not. A homeomorphism is simultaneously an open mapping and a closed mapping; that is, it maps open sets to open sets and closed sets to closed sets. Every self-homeomorphism in trick). can be extended to a self-homeomorphism of the whole disk (Alexander's

Informal discussion
The intuitive criterion of stretching, bending, cutting and gluing back together takes a certain amount of practice to apply correctlyit may not be obvious from the description above that deforming a line segment to a point is impermissible, for instance. It is thus important to realize that it is the formal definition given above that counts. This characterization of a homeomorphism often leads to confusion with the concept of homotopy, which is actually defined as a continuous deformation, but from one function to another, rather than one space to another. In the case of a homeomorphism, envisioning a continuous deformation is a mental tool for keeping track of which points on space X correspond to which points on Yone just follows them as X deforms. In the case of homotopy, the continuous deformation from one map to the other is of the essence, and it is also less restrictive, since none of the maps involved need to be one-to-one or onto. Homotopy does lead to a relation on spaces: homotopy equivalence. There is a name for the kind of deformation involved in visualizing a homeomorphism. It is (except when cutting and regluing are required) an isotopy between the identity map on X and the homeomorphism from X to Y.

References
[1] Hubbard, John H.; West, Beverly H. (1995). Differential Equations: A Dynamical Systems Approach. Part II: Higher-Dimensional Systems (http:/ / books. google. com/ books?id=SHBj2oaSALoC& pg=PA204& dq="coffee+ cup"+ topologist+ joke#v=onepage& q="coffee cup" topologist joke& f=false). Texts in Applied Mathematics. 18. Springer. p.204. ISBN978-0-387-94377-0. . [2] Poincar, Henri. "Chapter II: Mathematical Magnitude and Experiment" (https:/ / en. wikisource. org/ wiki/ Science_and_Hypothesis/ PART_I#b). Science and Hypothesis. .

External links
Homeomorphism (http://planetmath.org/?op=getobj&amp;from=objects&amp;id=912), PlanetMath.org.

Holomorphic function

18

Holomorphic function
In mathematics, holomorphic functions are the central objects of study in complex analysis. A holomorphic function is a complex-valued function of one or more complex variables that is complex differentiable in a neighborhood of every point in its domain. The existence of a complex derivative is a very strong condition, for it implies that any holomorphic function is actually infinitely differentiable and equal to its own Taylor series. The term analytic function is often used interchangeably with holomorphic function, although the word analytic is also used in a broader sense to describe any function (real, complex, or of more general type) that is equal to its Taylor series in a neighborhood of each point in its domain. The fact that the class of complex analytic functions coincides with the class of holomorphic functions is a major theorem in complex analysis. Holomorphic functions are also sometimes referred to as regular functions[1] or as conformal maps. A holomorphic function whose domain is the whole complex plane is called an entire function. The phrase "holomorphic at a point z0" means not just differentiable at z0, but differentiable everywhere within some neighborhood of z0 in the complex plane.

A rectangular grid (top) and its image under a conformal map f (bottom).

Definition
Given a complex-valued function of a single complex variable, the derivative of at a point z0 in its domain is defined by the limit

This is the same as the definition of the derivative for real functions, except that all of the quantities are complex. In particular, the limit is taken as the complex number z approaches z0, and must have the same value for any sequence of complex values for z that approach z0 on the complex plane. If the limit exists, we say that is complex-differentiable at the point z0. This concept of complex differentiability shares several properties with real differentiability: it is linear and obeys the product rule, quotient rule, and chain rule. If is complex differentiable at every point z0 in an open set U, we say that is holomorphic on U. We say that is holomorphic at the point z0 if it is holomorphic on some neighborhood ofz0. We say that is holomorphic on some non-open set A if it is holomorphic in an open set containingA. The relationship between real differentiability and complex differentiability is the following. If a complex function (x + i y)= u(x, y) + i v(x, y) is holomorphic, then u and v have first partial derivatives with respect to x and y, and satisfy the CauchyRiemann equations:

or, equivalently, the Wirtinger derivative of with respect to the complex conjugate of z is zero: to say that, roughly, is functionally independendent from the complex conjugate of z.

which is

Holomorphic function If continuity is not a given, the converse is not necessarily true. A simple converse is that if u and v have continuous first partial derivatives and satisfy the CauchyRiemann equations, then is holomorphic. A more satisfying converse, which is much harder to prove, is the LoomanMenchoff theorem: if is continuous, u and v have first partial derivatives (but not necessarily continuous), and they satisfy the CauchyRiemann equations, then is holomorphic.

19

Terminology
The word "holomorphic" was introduced by two of Cauchy's students, Briot (18171882) and Bouquet (18191895), and derives from the Greek (holos) meaning "entire", and (morph) meaning "form" or "appearance".[2] Today, the term "holomorphic function" is sometimes preferred to "analytic function", as the latter is a more general concept. This is also because an important result in complex analysis is that every holomorphic function is complex analytic, a fact that does not follow directly from the definitions. The term "analytic" is however also in wide use.

Properties
Because complex differentiation is linear and obeys the product, quotient, and chain rules, the sums, products and compositions of holomorphic functions are holomorphic, and the quotient of two holomorphic functions is holomorphic wherever the denominator is not zero. The derivative can be written as a contour integral using Cauchy's differentiation formula:

for any simple loop positively winding once around

, and

for infinitesimal positive loops


2

around

If one identifies C with R , then the holomorphic functions coincide with those functions of two real variables with continuous first derivatives which solve the Cauchy-Riemann equations, a set of two partial differential equations. Every holomorphic function can be separated into its real and imaginary parts, and each of these is a solution of Laplace's equation on R2. In other words, if we express a holomorphic function f(z) as u(x,y)+i v(x,y) both u and v are harmonic functions, where v is the harmonic conjugate of u and vice-versa. In regions where the first derivative is not zero, holomorphic functions are conformal in the sense that they preserve angles and the shape (but not size) of small figures. Cauchy's integral formula states that every function holomorphic inside a disk is completely determined by its values on the disk's boundary. Every holomorphic function is analytic. That is, a holomorphic function f has derivatives of every order at each point a in its domain, and it coincides with its own Taylor series at a in a neighborhood of a. In fact, f coincides with its Taylor series at a in any disk centered at that point and lying within the domain of the function. From an algebraic point of view, the set of holomorphic functions on an open set is a commutative ring and a complex vector space. In fact, it is a locally convex topological vector space, with the seminorms being the suprema on compact subsets. From a geometric perspective, a function f is holomorphic at z0 if and only if its exterior derivative df in a neighborhood U of z0 is equal to f(z) dz for some continuous function f. It follows from

Holomorphic function that df is also proportional to dz, implying that the derivative f is itself holomorphic and thus that f is infinitely differentiable. Similarly, the fact that d(f dz) = f dz dz = 0 implies that any function f that is holomorphic on the simply connected region U is also integrable on U. (For a path from z0 to z lying entirely in U, define ; in light of the Jordan curve theorem and the generalized Stokes' theorem, F(z) is independent of the particular choice of path , and thus F(z) is a well-defined function on U having F(z0) = F0 and dF = f dz.)

20

Examples
All polynomial functions in z with complex coefficients are holomorphic on C, and so are sine, cosine and the exponential function. (The trigonometric functions are in fact closely related to and can be defined via the exponential function using Euler's formula). The principal branch of the complex logarithm function is holomorphic on the set C \ {z R : z 0}. The square root function can be defined as

and is therefore holomorphic wherever the logarithm log(z) is. The function 1/z is holomorphic on {z : z 0}. As a consequence of the CauchyRiemann equations, a real-valued holomorphic function must be constant. Therefore, the absolute value of z, the argument of z, the real part of z and the imaginary part of z are not holomorphic. Another typical example of a continuous function which is not holomorphic is complex conjugation.

Several variables
A complex analytic function of several complex variables is defined to be analytic and holomorphic at a point if it is locally expandable (within a polydisk, a Cartesian product of disks, centered at that point) as a convergent power series in the variables. This condition is stronger than the CauchyRiemann equations; in fact it can be stated as follows: A function of several complex variables is holomorphic if and only if it satisfies the CauchyRiemann equations and is locally square-integrable.

Extension to functional analysis


The concept of a holomorphic function can be extended to the infinite-dimensional spaces of functional analysis. For instance, the Frchet or Gteaux derivative can be used to define a notion of a holomorphic function on a Banach space over the field of complex numbers.

References
[1] Springer Online Reference Books (http:/ / eom. springer. de/ a/ a012240. htm), Wolfram MathWorld (http:/ / mathworld. wolfram. com/ RegularFunction. html) [2] Markushevich, A.I.; Silverman, Richard A. (ed.) (2005) [1977]. Theory of functions of a Complex Variable (http:/ / books. google. com/ books?id=H8xfPRhTOcEC& dq) (2nd ed. ed.). New York: American Mathematical Society. p.112. ISBN0-8218-3780-X. .

Diffeomorphism

21

Diffeomorphism
In mathematics, a diffeomorphism is an isomorphism in the category of smooth manifolds. It is an invertible function that maps one differentiable manifold to another, such that both the function and its inverse are smooth.

Definition
Given two manifolds M and N, a bijective map f from M to N is called a diffeomorphism if both

and its inverse

are differentiable (if these functions are r times continuously differentiable, f is called a -diffeomorphism). Two manifolds M and N are diffeomorphic (symbol usually being ) if there is a smooth bijective map f from M to N with a smooth inverse. They are diffeomorphic if there is an r times continuously differentiable bijective map between them whose inverse is also r times continuously differentiable.

The image of a rectangular grid on a square under a diffeomorphism from the square onto itself.

Diffeomorphisms of subsets of manifolds


Given a subset X of a manifold M and a subset Y of a manifold N, a function f : X Y is said to be smooth if for all p in X there is a neighborhood of p and a smooth function g: U N such that the restrictions agree (note that g is an extension of f). We say that f is a diffeomorphism if it is bijective, smooth, and if its inverse is smooth.

Local description
Model example: if U and V are two connected open subsets of Rn such that V is simply connected, a differentiable map f: U V is a diffeomorphism if it is proper and if the differential Dfx: Rn Rn is bijective at each point x in U. Remarks It is essential for U to be simply connected for the function f to be globally invertible (under the sole condition that its derivative is a bijective map at each point). For example, consider the map (which is the "realification" of the complex square function) where U = V = R2 \ {(0,0)}. Then the map f is surjective and its satisfies (thus Dfx is bijective at each point) yet f is not invertible, because it fails to be injective, e.g., f(1,0) = (1,0) = f(-1,0). Since the differential at a point (for a differentiable function) is a linear map it has a

well defined inverse if, and only if, Dfx is a bijection. The matrix representation of Dfx is the n n matrix of first order partial derivatives whose entry in the i-th row and j-th colomn is . We often use this so-called Jacobian matrix for explicit computations. Diffeomorphisms are necessarily between manifolds of the same dimension. Imagine that f were going from dimension n to dimension k. If n < k then Dfx could never be surjective, and if n > k then Dfx could never be injective. So in both cases Dfx fails to be a bijection.

Diffeomorphism If Dfx is a bijection at x then we say that f is a local diffeomorphism (since by continuity Dfy will also be bijective for all y sufficiently close to x). Given a smooth map from dimension n to dimension k, if Df (resp. Dfx) is surjective then we say that f is a submersion (resp. local submersion), and if Df (resp. Dfx) is injective we say that f is an immersion (resp. local immersion). A differentiable bijection is not necessarily a diffeomorphism, e.g. f(x) = x3 is not a diffeomorphism from R to itself because its derivative vanishes at 0 (and hence its inverse is not differentiable at 0). This is an example of a homeomorphism that is not a diffeomorphism. f being a diffeomorphism is a stronger condition than f being a homeomorphism (when f is a map between differentiable manifolds). For a diffeomorphism we need f and its inverse to be differentiable. For a homeomorphism we only require that f and its inverse be continuous. Thus every diffeomorphism is a homeomorphism, but the converse is false: not every homeomorphism is a diffeomorphism. Now, f: M N is called a diffeomorphism if in coordinates charts it satisfies the definition above. More precisely, pick any cover of M by compatible coordinate charts, and do the same for N. Let and be charts on M and N respectively, with U being the image of and V the image of . Then the conditions says that the map f 1: U V is a diffeomorphism as in the definition above (whenever it makes sense). One has to check that for every couple of charts , of two given atlases, but once checked, it will be true for any other compatible chart. Again we see that dimensions have to agree.

22

Examples
Since any manifold can be locally parametrised, we can consider some explicit maps from two-space into two-space. Let . We can calculate the Jacobian matrix:

The Jacobian matrix has zero determinant if, and only if xy = 0. We see that f is a diffeomorphism away from the x-axis and the y-axis. Let where the and are

arbitrary real numbers, and the omitted terms are of degree at least two in x and y. We can calculate the Jacobian matrix at 0:

We see that g is a local diffeomorphism at 0 if, and only if, components of g are linearly independent as polynomials. Now let

, i.e. the linear terms in the

. We can calculate the Jacobian matrix:

The Jacobian matrix has zero determinant everywhere! In fact we see that the image of h is the unit circle.

Diffeomorphism

23

Diffeomorphism group
Let M be a differentiable manifold that is second-countable and Hausdorff. The diffeomorphism group of M is the group of all Cr diffeomorphisms of M to itself, and is denoted by Diffr(M) or Diff(M) when r is understood. This is a 'large' group, in the sense that it is not locally compact (provided M is not zero-dimensional).

Topology
The diffeomorphism group has two natural topologies, called the weak and strong topology (Hirsch 1997). When the manifold is compact, these two topologies agree. The weak topology is always metrizable. When the manifold is not compact, the strong topology captures the behavior of functions "at infinity", and is not metrizable. It is, however, still Baire. Fixing a Riemannian metric on M, the weak topology is the topology induced by the family of metrics

as K varies over compact subsets of M. Indeed, since M is -compact, there is a sequence of compact subsets Kn whose union is M. Then, define

The diffeomorphism group equipped with its weak topology is locally homeomorphic to the space of Cr vector fields (Leslie 1967). Over a compact subset of M, this follows by fixing a Riemannian metric on M and using the exponential map for that metric. If r is finite and the manifold is compact, the space of vector fields is a Banach space. Moreover, the transition maps from one chart of this atlas to another are smooth, making the diffeomorphism group into a Banach manifold. If r= or if the manifold is -compact, the space of vector fields is a Frchet space. Moreover, the transition maps are smooth, making the diffeomorphism group into a Frchet manifold.

Lie algebra
In particular, the Lie algebra of the diffeomorphism group of M consists of all vector fields on M, equipped with the Lie bracket of vector fields. Somewhat formally, this is seen by making a small change to the coordinate x at each point in space:

so the infinitesimal generators are the vector fields

Examples
When M=G is a Lie group, there is a natural inclusion of G in its own diffeomorphism group via left-translation. Let Diff(G) denote the diffeomorphism group of G, then there is a splitting Diff(G) G Diff(G,e) where Diff(G,e) is the subgroup of Diff(G) that fixes the identity element of the group. The diffeomorphism group of Euclidean space Rn consists of two components, consisting of the orientation preserving and orientation reversing diffeomorphisms. In fact, the general linear group is a deformation retract of subgroup Diff(Rn,0) of diffeomorphisms fixing the origin under the map (x) (tx)/t, t(0,1]. Hence, in particular, the general linear group is also a deformation retract of the full diffeomorphism group as well. For a finite set of points, the diffeomorphism group is simply the symmetric group. Similarly, if M is any manifold there is a group extension 0 Diff0(M) Diff(M) (0M). Here Diff0(M)is the subgroup of Diff(M) that preserves all the components of M, and (0M) is the permutation group of the set 0M (the

Diffeomorphism components of M). Moreover, the image of the map Diff(M) (0M) is the bijections of 0M that preserve diffeomorphism classes.

24

Transitivity
For a connected manifold M the diffeomorphism group acts transitively on M. More generally, the diffeomorphism group acts transitively on the configuration space CkM. If the dimension of M is at least two the diffeomorphism group acts transitively on the configuration space FkM: the action on M is multiply transitive (Banyaga 1997, p.29).

Extensions of diffeomorphisms
In 1926, Tibor Rad asked whether the harmonic extension of any homeomorphism (or diffeomorphism) of the unit circle to the unit disc yields a diffeomorphism on the open disc. An elegant proof was provided shortly afterwards by Hellmuth Kneser and a completely different proof was discovered in 1945 by Gustave Choquet, apparently unaware that the theorem was already known. The (orientation-preserving) diffeomorphism group of the circle is pathwise connected. This can be seen by noting that any such diffeomorphism can be lifted to a diffeomorphism f of the reals satisfying f(x+1) = f(x) +1; this space is convex and hence path connected. A smooth eventually constant path to the identity gives a second more elementary way of extending a diffeomorphism from the circle to the open unit disc (this is a special case of the Alexander trick). Moreover, the diffeomorphism group of the circle has the homotopy-type of the orthogonal group O(2). The corresponding extension problem for diffeomorphisms of higher dimensional spheres Sn1 was much studied in the 1950s and 1960s, with notable contributions from Ren Thom, John Milnor and Stephen Smale. An obstruction to such extensions is given by the finite Abelian group n, the "group of twisted spheres", defined as the quotient of the Abelian component group of the diffeomorphism group by the subgroup of classes extending to diffeomorphisms of the ball Bn.

Connectedness
For manifolds the diffeomorphism group is usually not connected. Its component group is called the mapping class group. In dimension 2, i.e. for surfaces, the mapping class group is a finitely presented group, generated by Dehn twists (Dehn, Lickorish, Hatcher). Max Dehn and Jakob Nielsen showed that it can be identified with the outer automorphism group of the fundamental group of the surface. William Thurston refined this analysis by classifying elements of the mapping class group into three types: those equivalent to a periodic diffeomorphism; those equivalent to a diffeomorphism leaving a simple closed curve invariant; and those equivalent to pseudo-Anosov diffeomorphisms. In the case of the torus S1 x S1 = R2/Z2, the mapping class group is just the modular group SL(2,Z) and the classification reduces to the classical one in terms of elliptic, parabolic and hyperbolic matrices. Thurston accomplished his classification by observing that the mapping class group acted naturally on a compactification of Teichmller space; since this enlarged space was homeomorphic to a closed ball, the Brouwer fixed-point theorem became applicable. If M is an oriented smooth closed manifold, it was conjectured by Smale that the identity component of the group of orientation-preserving diffeomorphisms is simple. This had first been proved for a product of circles by Michel Herman; it was proved in full generality by Thurston.

Diffeomorphism

25

Homotopy types
The diffeomorphism group of S2 has the homotopy-type of the subgroup O(3). This was proven by Steve Smale.[1] The diffeomorphism group of the torus has the homotopy-type of its linear automorphisms: (S1)2 GL(2, Z). The diffeomorphism groups of orientable surfaces of genus g > 1 have the homotopy-type of their mapping class groupsi.e.: the components are contractible. The homotopy-type of the diffeomorphism groups of 3-manifolds are fairly well-understood via the work of Ivanov, Hatcher, Gabai and Rubinstein although there are a few outstanding open cases, primarily 3-manifolds with finite fundamental groups. The homotopy-type of diffeomorphism groups of n-manifolds for n > 3 are poorly undersood. For example, it is an open problem whether or not Diff(S4) has more than two components. But via the work of Milnor, Kahn and Antonelli it's known that Diff(Sn) does not have the homotopy-type of a finite CW-complex provided n > 6.

Homeomorphism and diffeomorphism


It is easy to find a homeomorphism that is not a diffeomorphism, but it is more difficult to find a pair of homeomorphic manifolds that are not diffeomorphic. In dimensions 1, 2, 3, any pair of homeomorphic smooth manifolds are diffeomorphic. In dimension 4 or greater, examples of homeomorphic but not diffeomorphic pairs have been found. The first such example was constructed by John Milnor in dimension 7. He constructed a smooth 7-dimensional manifold (called now Milnor's sphere) that is homeomorphic to the standard 7-sphere but not diffeomorphic to it. There are in fact 28 oriented diffeomorphism classes of manifolds homeomorphic to the 7-sphere (each of them is a total space of the fiber bundle over the 4-sphere with the 3-sphere as the fiber). Much more extreme phenomena occur for 4-manifolds: in the early 1980s, a combination of results due to Simon Donaldson and Michael Freedman led to the discovery of exotic R4s: there are uncountably many pairwise non-diffeomorphic open subsets of R4 each of which is homeomorphic to R4, and also there are uncountably many pairwise non-diffeomorphic differentiable manifolds homeomorphic to R4 that do not embed smoothly in R4.

Notes
[1] Smale, Diffeomorphisms of the 2-sphere, Proc. Amer. Math. Soc. 10 (1959) 621626.

References
Chaudhuri, Shyamoli, Hakuru Kawai and S.-H Henry Tye. "Path-integral formulation of closed strings," Phys. Rev. D, 36: 1148, 1987. Banyaga, Augustin (1997), The structure of classical diffeomorphism groups, Mathematics and its Applications, 400, Kluwer Academic, ISBN0-7923-4475-8 Duren, Peter L. (2004), Harmonic Mappings in the Plane, Cambridge Mathematical Tracts, 156, Cambridge University Press, ISBN0-521-64121-7 Hirsch, Morris (1997), Differential Topology, Berlin, New York: Springer-Verlag, ISBN978-0-387-90148-0 Kriegl, Andreas; Michor, Peter (1997), The convenient setting of global analysis, Mathematical Surveys and Monographs, 53, American Mathematical Society, ISBN0-8218-0780-3 Leslie, J. A. (1967), "On a differential structure for the group of diffeomorphisms", Topology. an International Journal of Mathematics 6 (2): 263271, doi:10.1016/0040-9383(67)90038-9, ISSN0040-9383, MR0210147 Milnor, John W. (2007), Collected Works Vol. III, Differential Topology, American Mathematical Society, ISBN0-8218-4230-7

Diffeomorphism Omori, Hideki (1997), Infinite-dimensional Lie groups, Translations of Mathematical Monographs, 158, American Mathematical Society, ISBN0-8218-4575-6 Kneser, Hellmuth (1926), "Lsung der Aufgabe 41." (in German), Jahresbericht der Deutschen Mathematiker-Vereinigung 35 (2): 123f.

26

Monomorphism
In the context of abstract algebra or universal algebra, a monomorphism is an injective homomorphism. A monomorphism from X to Y is often denoted with the notation . In the more general setting of category theory, a monomorphism (also called a monic morphism or a mono) is a left-cancellative morphism, that is, an arrow f : X Y such that, for all morphisms g1, g2 : Z X, Monomorphisms are a categorical generalization of injective functions; in some categories the notions coincide, but monomorphisms are more general, as in the examples below. The categorical dual of a monomorphism is an epimorphism, i.e. a monomorphism in a category C is an epimorphism in the dual category Cop. Every section is a monomorphism, and every retraction is an epimorphism.

Terminology
The companion terms monomorphism and epimorphism were originally introduced by Nicolas Bourbaki; Bourbaki uses monomorphism as shorthand for an injective function. Early category theorists believed that the correct generalization of injectivity to the context of categories was the cancellation property given above. While this is not exactly true for monic maps, it is very close, so this has caused little trouble, unlike the case of epimorphisms. Saunders Mac Lane attempted to make a distinction between what he called monomorphisms, which were maps in a concrete category whose underlying maps of sets were injective, and monic maps, which are monomorphisms in the categorical sense of the word. This distinction never came into general use. Another name for monomorphism is extension, although this has other uses too.

Relation to invertibility
Left invertible maps are necessarily monic: if l is a left inverse for f (meaning ), then f is monic, as

A left invertible map is called a split mono. A map f : X Y is monic if and only if the induced map f : Hom(Z, X) Hom(Z, Y), defined by for all morphisms h : Z X , is injective for all Z.

Monomorphism

27

Examples
Every morphism in a concrete category whose underlying function is injective is a monomorphism. In the category of sets, the converse also holds so the monomorphisms are exactly the injective morphisms. The converse also holds in most naturally occurring categories of algebras because of the existence of a free object on one generator. In particular, it is true in the categories of groups and rings, and in any abelian category. It is not true in general, however, that all monomorphisms must be injective in other categories. For example, in the category Div of divisible (abelian) groups and group homomorphisms between them there are monomorphisms that are not injective: consider, for example, the quotient map q:QQ/Z. This is not an injective map; nevertheless, it is a monomorphism in this category. This follows from (and in fact, is equivalent to) the implication qh = 0 h = 0, which we now prove. (NB: The converse of this last implication, namely, h = 0 qh = 0, is trivially true, but it is not needed here). If h:G Q, where G is some divisible group, and qh = 0, then h(x) Z, x G. Now fix some x G. Without loss of generality, we may assume that h(x) 0 (otherwise, choose x instead). Then, since G is a divisible group, for some y G, x = (h(x) + 1) y, so h(x) = (h(x) + 1) h(y). From this, and 0 h(x) < h(x) + 1, it follows that

Since h(y) Z, it follows that h(y) = 0, and thus h(x) = 0 = h(x), x G. This says that h = 0, as desired. Now, if qf = qg for some morphisms f, g:G Q, where G is some divisible group then q(f g) = 0, where (f g):x f(x) g(x). (Since (f g)(0) = 0, and (f - g)(x + y) = (f - g)(x) + (f - g)(y), it follows that (f - g) Hom(G, Q)). From the result just proved, q(f g) = 0 f g = 0 x G, f(x) = g(x) f = g. Hence q is a monomorphism, as claimed.

Properties
In a topos, every monic is an equalizer, and any map that is both monic and epic is an isomorphism. Every isomorphism is monic.

Related concepts
There are also useful concepts of regular monomorphism, strong monomorphism, and extremal monomorphism. A regular monomorphism equalizes some parallel pair of morphisms. An extremal monomorphism is a monomorphism that cannot be nontrivially factored through an epimorphism: Precisely, if m=ge with e an epimorphism, then e is an isomorphism. A strong monomorphism satisfies a certain lifting property with respect to commutative squares involving an epimorphism.

References
Francis Borceux (1994), Handbook of Categorical Algebra 1, Cambridge University Press. ISBN 0-521-44178-1. George Bergman (1998), An Invitation to General Algebra and Universal Constructions [1], Henry Helson Publisher, Berkeley. ISBN 0-9655211-4-1. Jaap van Oosten, Basic Category Theory [2]

References
[1] http:/ / math. berkeley. edu/ ~gbergman/ 245/ index. html [2] http:/ / www. math. uu. nl/ people/ jvoosten/ syllabi/ catsmoeder. pdf

Epimorphism

28

Epimorphism
In category theory, an epimorphism (also called an epic morphism or, colloquially, an epi) is a morphism f : X Y which is right-cancellative in the sense that, for all morphisms g1, g2 : Y Z, Epimorphisms are analogues of surjective functions, but they are not exactly the same. The dual of an epimorphism is a monomorphism (i.e. an epimorphism in a category C is a monomorphism in the dual category Cop). Many authors in abstract algebra and universal algebra define an epimorphism simply as an onto or surjective homomorphism. Every epimorphism in this algebraic sense is an epimorphism in the sense of category theory, but the converse is not true in all categories. In this article, the term "epimorphism" will be used in the sense of category theory given above. For more on this, see the section on Terminology below.

Examples
Every morphism in a concrete category whose underlying function is surjective is an epimorphism. In many concrete categories of interest the converse is also true. For example, in the following categories, the epimorphisms are exactly those morphisms which are surjective on the underlying sets: Set, sets and functions. To prove that every epimorphism f: X Y in Set is surjective, we compose it with both the characteristic function g1: Y {0,1} of the image f(X) and the map g2: Y {0,1} that is constant 1. Rel, sets with binary relations and relation preserving functions. Here we can use the same proof as for Set, equipping {0,1} with the full relation {0,1}{0,1}. Pos, partially ordered sets and monotone functions. If f : (X,) (Y,) is not surjective, pick y0 in Y \ f(X) and let g1 : Y {0,1} be the characteristic function of {y | y0 y} and g2 : Y {0,1} the characteristic function of {y | y0 < y}. These maps are monotone if {0,1} is given the standard ordering 0 < 1. Grp, groups and group homomorphisms. The result that every epimorphism in Grp is surjective is due to Otto Schreier (he actually proved more, showing that every subgroup is an equalizer using the free product with one amalgamated subgroup); an elementary proof can be found in (Linderholm 1970). FinGrp, finite groups and group homomorphisms. Also due to Schreier; the proof given in (Linderholm 1970) establishes this case as well. Ab, abelian groups and group homomorphisms. K-Vect, vector spaces over a field K and K-linear transformations. Mod-R, right modules over a ring R and module homomorphisms. This generalizes the two previous examples; to prove that every epimorphism f: X Y in Mod-R is surjective, we compose it with both the canonical quotient map g 1: Y Y/f(X) and the zero map g2: Y Y/f(X). Top, topological spaces and continuous functions. To prove that every epimorphism in Top is surjective, we proceed exactly as in Set, giving {0,1} the indiscrete topology which ensures that all considered maps are continuous. HComp, compact Hausdorff spaces and continuous functions. If f: X Y is not surjective, let y in Y-fX. Since fX is closed, by Urysohn's Lemma there is a continuous function g1:Y [0,1] such that g1 is 0 on fX and 1 on y. We compose f with both g1 and the zero function g2: Y [0,1]. However there are also many concrete categories of interest where epimorphisms fail to be surjective. A few examples are: In the category of monoids, Mon, the inclusion map N Z is a non-surjective epimorphism. To see this, suppose that g1 and g2 are two distinct maps from Z to some monoid M. Then for some n in Z, g1(n) g2(n), so g1(-n)

Epimorphism g2(-n). Either n or -n is in N, so the restrictions of g1 and g2 to N are unequal. In the category of algebras over commutative ring R, take R[N] R[Z], where R[G] is the group ring of the group G and the morphism is induced by the inclusion N Z as in the previous example. This follows from the observation that 1 generates the algebra R[Z] (note that the unit in R[Z] is given by 0 of Z), and the inverse of the element represented by n in Z is just the element represented by -n. Thus any homomorphism from R[Z] is uniquely determined by its value on the element represented by 1 of Z. In the category of rings, Ring, the inclusion map Z Q is a non-surjective epimorphism; to see this, note that any ring homomorphism on Q is determined entirely by its action on Z, similar to the previous example. A similar argument shows that the natural ring homomorphism from any commutative ring R to any one of its localizations is an epimorphism. In the category of commutative rings, a finitely generated homomorphism of rings f : R S is an epimorphism if and only if for all prime ideals P of R, the ideal Q generated by f(P) is either S or is prime, and if Q is not S, the induced map Frac(R/P) Frac(S/Q) is an isomorphism (EGA IV 17.2.6). In the category of Hausdorff spaces, Haus, the epimorphisms are precisely the continuous functions with dense images. For example, the inclusion map Q R, is a non-surjective epimorphism.

29

The above differs from the case of monomorphisms where it is more frequently true that monomorphisms are precisely those whose underlying functions are injective. As to examples of epimorphisms in non-concrete categories: If a monoid or ring is considered as a category with a single object (composition of morphisms given by multiplication), then the epimorphisms are precisely the right-cancellable elements. If a directed graph is considered as a category (objects are the vertices, morphisms are the paths, composition of morphisms is the concatenation of paths), then every morphism is an epimorphism.

Properties
Every isomorphism is an epimorphism; indeed only a right-sided inverse is needed: if there exists a morphism j : Y X such that fj = idY, then f is easily seen to be an epimorphism. A map with such a right-sided inverse is called a split epi. In a topos, a map that is both a monic morphism and an epimorphism is an isomorphism. The composition of two epimorphisms is again an epimorphism. If the composition fg of two morphisms is an epimorphism, then f must be an epimorphism. As some of the above examples show, the property of being an epimorphism is not determined by the morphism alone, but also by the category of context. If D is a subcategory of C, then every morphism in D which is an epimorphism when considered as a morphism in C is also an epimorphism in D; the converse, however, need not hold; the smaller category can (and often will) have more epimorphisms. As for most concepts in category theory, epimorphisms are preserved under equivalences of categories: given an equivalence F : C D, then a morphism f is an epimorphism in the category C if and only if F(f) is an epimorphism in D. A duality between two categories turns epimorphisms into monomorphisms, and vice versa. The definition of epimorphism may be reformulated to state that f : X Y is an epimorphism if and only if the induced maps

are injective for every choice of Z. This in turn is equivalent to the induced natural transformation being a monomorphism in the functor category SetC.

Epimorphism Every coequalizer is an epimorphism, a consequence of the uniqueness requirement in the definition of coequalizers. It follows in particular that every cokernel is an epimorphism. The converse, namely that every epimorphism be a coequalizer, is not true in all categories. In many categories it is possible to write every morphism as the composition of a monomorphism followed by an epimorphism. For instance, given a group homomorphism f : G H, we can define the group K = im(f) = f(G) and then write f as the composition of the surjective homomorphism G K which is defined like f, followed by the injective homomorphism K H which sends each element to itself. Such a factorization of an arbitrary morphism into an epimorphism followed by a monomorphism can be carried out in all abelian categories and also in all the concrete categories mentioned above in the Examples section (though not in all concrete categories).

30

Related concepts
Among other useful concepts are regular epimorphism, extremal epimorphism, strong epimorphism, and split epimorphism. A regular epimorphism coequalizes some parallel pair of morphisms. An extremal epimorphism is an epimorphism that has no monomorphism as a second factor, unless that monomorphism is an isomorphism. A strong epimorphism satisfies a certain lifting property with respect to commutative squares involving a monomorphism. A split epimorphism is a morphism which has a right-sided inverse. A morphism that is both a monomorphism and an epimorphism is called a bimorphism. Every isomorphism is a bimorphism but the converse is not true in general. For example, the map from the half-open interval [0,1) to the unit circle S1 (thought of as a subspace of the complex plane) which sends x to exp(2ix) (see Euler's formula) is continuous and bijective but not a homeomorphism since the inverse map is not continuous at 1, so it is an instance of a bimorphism that is not an isomorphism in the category Top. Another example is the embedding Q R in the category Haus; as noted above, it is a bimorphism, but it is not bijective and therefore not an isomorphism. Similarly, in the category of rings, the maps Z Q and Q R are bimorphisms but not isomorphisms. Epimorphisms are used to define abstract quotient objects in general categories: two epimorphisms f1 : X Y1 and f2 : X Y2 are said to be equivalent if there exists an isomorphism j : Y1 Y2 with jf1 = f2. This is an equivalence relation, and the equivalence classes are defined to be the quotient objects of X.

Terminology
The companion terms epimorphism and monomorphism were first introduced by Bourbaki. Bourbaki uses epimorphism as shorthand for a surjective function. Early category theorists believed that epimorphisms were the correct analogue of surjections in an arbitrary category, similar to how monomorphisms are very nearly an exact analogue of injections. Unfortunately this is incorrect; strong or regular epimorphisms behave much more closely to surjections than ordinary epimorphisms. Saunders Mac Lane attempted to create a distinction between epimorphisms, which were maps in a concrete category whose underlying set maps were surjective, and epic morphisms, which are epimorphisms in the modern sense. However, this distinction never caught on. It is a common mistake to believe that epimorphisms are either identical to surjections or that they are a better concept. Unfortunately this is rarely the case; epimorphisms can be very mysterious and have unexpected behavior. It is very difficult, for example, to classify all the epimorphisms of rings. In general, epimorphisms are their own unique concept, related to surjections but fundamentally different.

Epimorphism

31

References
Admek, Ji, Herrlich, Horst, & Strecker, George E. (1990). Abstract and Concrete Categories [1] (4.2MB PDF). Originally publ. John Wiley & Sons. ISBN 0-471-60922-6. (now free on-line edition) Bergman, George M. (1998), An Invitation to General Algebra and Universal Constructions, Harry Helson Publisher, Berkeley. ISBN 0-9655211-4-1. Linderholm, Carl (1970). A Group Epimorphism is Surjective. American Mathematical Monthly 77, pp.176177. Proof summarized by Arturo Magidin in [2]. Lawvere & Rosebrugh: Sets for Mathematics, Cambridge university press, 2003. ISBN 0-521-80444-2.

References
[1] http:/ / katmat. math. uni-bremen. de/ acc/ acc. pdf [2] http:/ / groups. google. com/ group/ sci. math/ msg/ 6d4023d93a2b4300

Article Sources and Contributors

32

Article Sources and Contributors


Morphism Source: http://en.wikipedia.org/w/index.php?oldid=502672907 Contributors: 213.253.39.xxx, Alberto da Calvairate, Albmont, Aragorn2, Archelon, AxelBoldt, Bart v M, Blotwell, Bonotake, Brad7777, Bryan Derksen, Cenarium, Ciphergoth, Conversion script, Dysprosia, Edcolins, Elwikipedista, Fropuff, Giftlite, Glenn, Gorobay, Graham87, Greenrd, Helder.wiki, Hotfeba, IstvanWolf, JamesBWatson, Jon Awbrey, Jose Ramos, Jsnx, Kku, Kmarinas86, Larry_Sanger, Lethe, Linas, LokiClock, Looxix, Michael Hardy, Michael K. Edwards, Mikez, Msh210, Ntmatter, Odoncaoa, PV=nRT, Palnot, Phys, Porges, Pthag, Revolver, Salix alba, Sam Hocevar, Semorrison, Simon J Kissane, Sjef, Slawekb, Smimram, TakuyaMurata, The Anome, Tkeu, Weppens, Xantharius, 27 anonymous edits Homomorphism Source: http://en.wikipedia.org/w/index.php?oldid=493862815 Contributors: 213.253.39.xxx, Adriaan Joubert, Allanhalme, Andres, AugPi, AxelBoldt, Bloodshedder, Bryan Derksen, CRGreathouse, Calle, Classicalecon, Clment Pillias, Conversion script, Dirac1933, Dysprosia, EagleFan, Ed g2s, El C, Electiontechnology, EmilJ, Ex13, FractalFusion, Fropuff, Gandalf61, Giftlite, Glenn, Goochelaar, Graham87, Hadal, Hukkinen, Ishboyfay, JAn Dudk, Jim.belk, JoergenB, Johntyree, Jyoshimi, Kanags, Kevinbsmith, KnightRider, Korako, Kostisl, Larry_Sanger, Ligulem, LokiClock, Martynas Patasius, Mehdi.manshadi, Mesoderm, Mgreenbe, Michael Hardy, Michael Slone, Mlm42, Msh210, Netsnipe, Nishantjr, Oblivious, Oleg Alexandrov, PV=nRT, Pcap, Pramcom, Qwfp, R'n'B, Rat144, Rgamble, Riteshsood, Robert K S, Sahar Tomer, Salix alba, Sebastian Goll, Staffwaterboy, The Anome, TheSeven, Toby Bartels, Tong, Vantelimus, Wowulu, XJamRastafire, Youandme, Zero0000, Zundark, 78 anonymous edits Automorphism Source: http://en.wikipedia.org/w/index.php?oldid=502140961 Contributors: Algebraist, Altenmann, Amire80, Archelon, AxelBoldt, Bdesham, Brews ohare, Caiodnh, Cbogart2, Chas zzz brown, Conversion script, Crisfilax, David Eppstein, Dogaroon, Dreftymac, Dysprosia, Elwikipedista, Fropuff, Gauge, Giftlite, Gogo Dodo, Goochelaar, Gregbard, Hannes Eder, Happy-melon, Henning Makholm, Hephaestos, Huppybanny, JRB-Europe, JackSchmidt, Jan Hidders, Keenan Pepper, Ksanyi, LC, MSGJ, Marc van Leeuwen, MatrixFrog, Mcmillin24, Mets501, Michael Hardy, Mild Bill Hiccup, MishaMisha, Mskalak13, Nbarth, Nishantjr, Noisy, Obradovic Goran, Orenburg1, PV=nRT, Patrick, Peruvianllama, Phys, Pixeltoo, Pred, Qwertyus, Rdsmith4, Reyk, Rgdboer, Rich Farmbrough, Salgueiro, Sam Hocevar, Shadowjams, SirJective, Tarquin, TechnoGuyRob, Thehotelambush, TimothyRias, Tommy Jantarek, Topology Expert, Tosha, VKokielov, William M. Connolley, Xezbeth, Youssefsan, Zaslav, 23 anonymous edits Endomorphism Source: http://en.wikipedia.org/w/index.php?oldid=504334772 Contributors: Altenmann, Andres, Arskwad, AugPi, AxelBoldt, CitizenB, Conversion script, David Eppstein, Dbenbenn, Doctorfree, Dysprosia, Elwikipedista, Ewlyahoocom, Fropuff, Giftlite, Graham87, Helder.wiki, Ilmari Karonen, JackSchmidt, Javy tahu, Jim.belk, Joshua Davis, JoshuaGrosse, Kmarinas86, MFH, Magmalex, Melchoir, Oleg Alexandrov, PV=nRT, Point-set topologist, Quiddity, Salix alba, Smimram, Tarquin, Thomas T Howard, Tosha, Vonkje, Zundark, Zyqqh, 21 anonymous edits Isomorphism Source: http://en.wikipedia.org/w/index.php?oldid=502956121 Contributors: 16@r, Albmont, Andre Engels, Anonymous Dissident, Army1987, Avaya1, AxelBoldt, Bahar101, Banus, Bender2k14, Bigdumbdinosaur, Bkell, Borhan0, Brad7777, Brews ohare, Bryanmcdonald, CRGreathouse, Charles Matthews, Conversion script, Coolhandscot, Cpiral, Cronholm144, Dan Granahan, Debresser, Dysprosia, Edemaine, Elwikipedista, EmilJ, Ezequiels.90, Fantomdrives, Fly by Night, Ghakko, Giftlite, Glenn, Grubber, Gryllida, Hannes Eder, Iamthedeus, Ioannis.Demetriou, Isomorphic, Ivan tambuk, Jim.belk, Jlittlet, KSmrq, Kmarinas86, Koberozendaal, LOL, Lars Washington, LokiClock, Lubos, M hariprasad, Marc Venot, MarkSweep, MathKnight, Mathwiz777, Mattmacf, Mcole13, Melongrower, Mesoderm, Mets501, Michael Angelkovich, Michael Hardy, Michael Slone, Mike4ty4, Msh210, Mxn, Nabla, Nbarth, Netsnipe, Nick Green, Oleg Alexandrov, Omnipedian, PV=nRT, Patrick, Paul August, PaulTanenbaum, Peruvianllama, Philip Cross, PhotoBox, Phys, Pokus9999, QuiteUnusual, Quondum, Revolver, Richard Molnr-Szipai, Rljacobson, Rlupsa, Rschwieb, Ryguasu, Sam Staton, Slawekb, Smite-Meister, Spacepotato, Spinality, Spinningspark, Ssavelan, Stevertigo, Stuhacking, Subversive.sound, Superbatfish, Sawomir Biay, TK-925, Taemyr, Thaddeus Slamp, The Rhymesmith, Thehotelambush, Tosha, Trumpet marietta 45750, Uncle Dick, Wrp103, YnnusOiramo, Youandme, Youssefsan, Yuide, Yurik, Zell08v, Zero0000, Zundark, , 98 , anonymous edits Homeomorphism Source: http://en.wikipedia.org/w/index.php?oldid=498554773 Contributors: 9258fahsflkh917fas, Adeliine, Ahoerstemeier, Andrew Delong, Arjun r acharya, Asztal, AxelBoldt, BenFrantzDale, Brians, Charles Matthews, ClamDip, Conversion script, Crasshopper, DekuDekuplex, Dominus, Dr.K., Dysprosia, Elwikipedista, EmilJ, Explorall, Frencheigh, Fropuff, Giftlite, Glenn, Grondemar, Hadal, Hakeem.gadi, Henry Delforn (old), Joule36e5, Juan Marquez, Kieff, Linas, LokiClock, Lorem Ip, Mark Foskey, MathKnight, MathMartin, Mchipofya, Michael Angelkovich, Michael Hardy, Miguel, MisfitToys, Ms2ger, Msh210, Myasuda, Nbarth, Ojigiri, Oleg Alexandrov, Orionus, Orthografer, Pandaritrl, Parudox, Paul August, Pengo, PhilKnight, Prijutme4ty, Renamed user 1, RyanEberhart, Salix alba, Seqsea, Set theorist, SingingDragon, Steelpillow, The cattr, TheObtuseAngleOfDoom, ThomasWinwood, Tobias Bergemann, Toby Bartels, Topology Expert, Tosha, Vonkje, Weialawaga, Wikomidia, Wjmallard, XJamRastafire, Xmlizer, Youandme, Zero sharp, Zhaoway, Zundark, 96 ,anonymous edits Holomorphic function Source: http://en.wikipedia.org/w/index.php?oldid=505855894 Contributors: 212.242.115.xxx, Aaronbrick, Abiola Lapite, Acepectif, Adoniscik, Almit39, Altenmann, AxelBoldt, Bdmy, Ben pcc, Brad7777, CRGreathouse, Charles Matthews, Chinju, Christian.Mercat, Conversion script, Crasshopper, CrniBombarder!!!, Crust, Damian Yerrick, DavidCBryant, DomenicDenicola, Dougher, DragonflySixtyseven, Dratman, Duckbill, Dysprosia, Dzordzm, Eequor, Email4mobile, EmilJ, Fakhredinblog, FilipeS, Frencheigh, FuriousScribble, Giftlite, Graham87, HannsEwald, Hesam7, Howard McCay, Irigi, Isocliff, JamesBWatson, Jao, Jim.belk, Jmath666, Jobh, Jusjih, KnightRider, Laurent MAYER, Lethe, Lhf, Linas, Lunch, Maxim Razin, Michael Hardy, Michael K. Edwards, Mike40033, Mokhtari34, Naddy, NickBush24, ObsessiveMathsFreak, Oleg Alexandrov, PV=nRT, Patrick, PierreAbbat, Prim Ethics, Reaverdrop, Riwnodennyk, Robert Illes, Rvollmert, Saleemsan, Salix alba, Silly rabbit, Slawekb, Sligocki, Some jerk on the Internet, Stephen Bain, TakuyaMurata, Tarquin, Tcnuk, Tetracube, The Diagonal Prince, Thorfinn, Tobias Bergemann, Tobias Hoevekamp, Unco, Vanished User 0001, Weierstra, XJamRastafire, 60 anonymous edits Diffeomorphism Source: http://en.wikipedia.org/w/index.php?oldid=504516901 Contributors: 0, AugPi, Charles Matthews, Chester Markel, Connelly, CryptoDerk, Dharma6662000, Dreadstar, Dysprosia, Ensign beedrill, Fly by Night, Fropuff, Gaius Cornelius, Geometry guy, Giftlite, Headbomb, Helder.wiki, Herve.lombaert, JeLuF, JerroldPease-Atlanta, Kuszi, LachlanA, Lethe, Lmatt, MathMartin, Mathsci, Maury Markowitz, Med, Mhwu, Michael Hardy, Msh210, Muses' house, Nakon, Nosirrom, Oleg Alexandrov, PV=nRT, Pascalromon, Paul August, Physicistjedi, Point-set topologist, Poor Yorick, R.e.b., Redrose64, Rybu, Silly rabbit, Slawekb, Sawomir Biay, TakuyaMurata, Topology Expert, Tosha, Uni.Liu, Whitepaw, Woseph, kebrke, 42 anonymous edits Monomorphism Source: http://en.wikipedia.org/w/index.php?oldid=503687785 Contributors: Adler F, Albmont, Altenmann, Archelon, Blotwell, CRGreathouse, Ckhenderson, Classicalecon, Crasshopper, E-boy, Eric Kvaalen, Fropuff, Gauge, Giftlite, Hairy Dude, Hans Adler, Le schtroumpf, Lethe, Magidin, MarSch, Nbarth, Nick Number, Obradovic Goran, PV=nRT, Perturbationist, Rich Farmbrough, Sam Hocevar, Tesseran, Tobias Bergemann, Toby Bartels, Tosha, Xantharius, YouRang?, 14 anonymous edits Epimorphism Source: http://en.wikipedia.org/w/index.php?oldid=487833533 Contributors: Albmont, Altenmann, Anonymous Dissident, AxelBoldt, Beroal, Ckhenderson, Daniel5Ko, Esap, Fropuff, Gauge, Giftlite, Icey, Itai, Kmarinas86, Kompik, Magidin, Mszamotulski, Nbarth, PV=nRT, Paul August, Pit-trout, Salix alba, Sam Hocevar, Silvonen, Slawekb, Thehotelambush, Toby Bartels, Tosha, VeryVerily, Zoicon5, 28 anonymous edits

Image Sources, Licenses and Contributors

33

Image Sources, Licenses and Contributors


Image:Commutative diagram for morphism.svg Source: http://en.wikipedia.org/w/index.php?title=File:Commutative_diagram_for_morphism.svg License: Public Domain Contributors: User:Cepheus Image:Morphisms.svg Source: http://en.wikipedia.org/w/index.php?title=File:Morphisms.svg License: Creative Commons Attribution-ShareAlike 3.0 Unported Contributors: User:EmilJ, based on a png image by en:User:AugPi Image:Mug and Torus morph.gif Source: http://en.wikipedia.org/w/index.php?title=File:Mug_and_Torus_morph.gif License: Public Domain Contributors: Abnormaal, Durova, Howcheng, Kieff, Kri, Manco Capac, Maximaximax, Rovnet, SharkD, Takabeg, 16 anonymous edits Image:Trefoil knot arb.png Source: http://en.wikipedia.org/w/index.php?title=File:Trefoil_knot_arb.png License: GNU Free Documentation License Contributors: Editor at Large, Ylebru, 2 anonymous edits Image:Conformal map.svg Source: http://en.wikipedia.org/w/index.php?title=File:Conformal_map.svg License: Public Domain Contributors: Oleg Alexandrov Image:Diffeomorphism of a square.svg Source: http://en.wikipedia.org/w/index.php?title=File:Diffeomorphism_of_a_square.svg License: Public Domain Contributors: Oleg Alexandrov Image:Monomorphism-01.png Source: http://en.wikipedia.org/w/index.php?title=File:Monomorphism-01.png License: Public Domain Contributors: Derlay, Ico83, Samulili, 1 anonymous edits Image:Epimorphism-01.png Source: http://en.wikipedia.org/w/index.php?title=File:Epimorphism-01.png License: Public Domain Contributors: Darapti, Derlay, Maksim

License

34

License
Creative Commons Attribution-Share Alike 3.0 Unported //creativecommons.org/licenses/by-sa/3.0/

Вам также может понравиться