Injective function
- العربية
- Беларуская
- Български
- Bosanski
- Català
- Čeština
- Dansk
- Deutsch
- Eesti
- Ελληνικά
- Español
- Esperanto
- Euskara
- فارسی
- Français
- Galego
- 한국어
- हिन्दी
- Hrvatski
- Ido
- Bahasa Indonesia
- Interlingua
- Íslenska
- Italiano
- עברית
- Қазақша
- Latina
- Lietuvių
- Lombard
- Magyar
- Македонски
- Bahasa Melayu
- Nederlands
- 日本語
- Norsk bokmål
- Norsk nynorsk
- Occitan
- Polski
- Português
- Română
- Русский
- Simple English
- Slovenčina
- Slovenščina
- Ślůnski
- کوردی
- Српски / srpski
- Suomi
- Svenska
- தமிழ்
- ไทย
- Türkçe
- Українська
- Tiếng Việt
- 粵語
- 中文
Function |
---|
x ↦ f (x) |
History of the function concept |
Types by domain and codomain |
Classes/properties |
Constructions |
Generalizations |
List of specific functions |
In mathematics, an injective function (also known as injection, or one-to-one function[1] ) is a function f that maps distinct elements of its domain to distinct elements of its codomain; that is, x1 ≠ x2 implies f(x1) ≠ f(x2) (equivalently by contraposition, f(x1) = f(x2) implies x1 = x2). In other words, every element of the function's codomain is the image of at most one element of its domain.[2] The term one-to-one function must not be confused with one-to-one correspondence that refers to bijective functions, which are functions such that each element in the codomain is an image of exactly one element in the domain.
A homomorphism between algebraic structures is a function that is compatible with the operations of the structures. For all common algebraic structures, and, in particular for vector spaces, an injective homomorphism is also called a monomorphism . However, in the more general context of category theory, the definition of a monomorphism differs from that of an injective homomorphism.[3] This is thus a theorem that they are equivalent for algebraic structures; see Homomorphism § Monomorphism for more details.
A function {\displaystyle f} that is not injective is sometimes called many-to-one.[2]
Definition
[edit ]Let {\displaystyle f} be a function whose domain is a set {\displaystyle X.} The function {\displaystyle f} is said to be injective provided that for all {\displaystyle a} and {\displaystyle b} in {\displaystyle X,} if {\displaystyle f(a)=f(b),} then {\displaystyle a=b}; that is, {\displaystyle f(a)=f(b)} implies {\displaystyle a=b.} Equivalently, if {\displaystyle a\neq b,} then {\displaystyle f(a)\neq f(b)} in the contrapositive statement.
Symbolically,{\displaystyle \forall a,b\in X,\;\;f(a)=f(b)\Rightarrow a=b,} which is logically equivalent to the contrapositive,[4] {\displaystyle \forall a,b\in X,\;\;a\neq b\Rightarrow f(a)\neq f(b).}An injective function (or, more generally, a monomorphism) is often denoted by using the specialized arrows ↣ or ↪ (for example, {\displaystyle f:A\rightarrowtail B} or {\displaystyle f:A\hookrightarrow B}), although some authors specifically reserve ↪ for an inclusion map.[5]
Examples
[edit ]For visual examples, readers are directed to the gallery section.
- For any set {\displaystyle X} and any subset {\displaystyle S\subseteq X,} the inclusion map {\displaystyle S\to X} (which sends any element {\displaystyle s\in S} to itself) is injective. In particular, the identity function {\displaystyle X\to X} is always injective (and in fact bijective).
- If the domain of a function is the empty set, then the function is the empty function, which is injective.
- If the domain of a function has one element (that is, it is a singleton set), then the function is always injective.
- The function {\displaystyle f:\mathbb {R} \to \mathbb {R} } defined by {\displaystyle f(x)=2x+1} is injective.
- The function {\displaystyle g:\mathbb {R} \to \mathbb {R} } defined by {\displaystyle g(x)=x^{2}} is not injective, because (for example) {\displaystyle g(1)=1=g(-1).} However, if {\displaystyle g} is redefined so that its domain is the non-negative real numbers [0,+∞), then {\displaystyle g} is injective.
- The exponential function {\displaystyle \exp :\mathbb {R} \to \mathbb {R} } defined by {\displaystyle \exp(x)=e^{x}} is injective (but not surjective, as no real value maps to a negative number).
- The natural logarithm function {\displaystyle \ln :(0,\infty )\to \mathbb {R} } defined by {\displaystyle x\mapsto \ln x} is injective.
- The function {\displaystyle g:\mathbb {R} \to \mathbb {R} } defined by {\displaystyle g(x)=x^{n}-x} is not injective, since, for example, {\displaystyle g(0)=g(1)=0.}
More generally, when {\displaystyle X} and {\displaystyle Y} are both the real line {\displaystyle \mathbb {R} ,} then an injective function {\displaystyle f:\mathbb {R} \to \mathbb {R} } is one whose graph is never intersected by any horizontal line more than once. This principle is referred to as the horizontal line test .[2]
Injections can be undone
[edit ]Functions with left inverses are always injections. That is, given {\displaystyle f:X\to Y,} if there is a function {\displaystyle g:Y\to X} such that for every {\displaystyle x\in X}, {\displaystyle g(f(x))=x}, then {\displaystyle f} is injective. In this case, {\displaystyle g} is called a retraction of {\displaystyle f.} Conversely, {\displaystyle f} is called a section of {\displaystyle g.}
Conversely, every injection {\displaystyle f} with a non-empty domain has a left inverse {\displaystyle g}. It can be defined by choosing an element {\displaystyle a} in the domain of {\displaystyle f} and setting {\displaystyle g(y)} to the unique element of the pre-image {\displaystyle f^{-1}[y]} (if it is non-empty) or to {\displaystyle a} (otherwise).[6]
The left inverse {\displaystyle g} is not necessarily an inverse of {\displaystyle f,} because the composition in the other order, {\displaystyle f\circ g,} may differ from the identity on {\displaystyle Y.} In other words, an injective function can be "reversed" by a left inverse, but is not necessarily invertible, which requires that the function is bijective.
Injections may be made invertible
[edit ]In fact, to turn an injective function {\displaystyle f:X\to Y} into a bijective (hence invertible) function, it suffices to replace its codomain {\displaystyle Y} by its actual image {\displaystyle J=f(X).} That is, let {\displaystyle g:X\to J} such that {\displaystyle g(x)=f(x)} for all {\displaystyle x\in X}; then {\displaystyle g} is bijective. Indeed, {\displaystyle f} can be factored as {\displaystyle \operatorname {In} _{J,Y}\circ g,} where {\displaystyle \operatorname {In} _{J,Y}} is the inclusion function from {\displaystyle J} into {\displaystyle Y.}
More generally, injective partial functions are called partial bijections.
Other properties
[edit ]- If {\displaystyle f} and {\displaystyle g} are both injective then {\displaystyle f\circ g} is injective.
- If {\displaystyle g\circ f} is injective, then {\displaystyle f} is injective (but {\displaystyle g} need not be).
- {\displaystyle f:X\to Y} is injective if and only if, given any functions {\displaystyle g,} {\displaystyle h:W\to X} whenever {\displaystyle f\circ g=f\circ h,} then {\displaystyle g=h.} In other words, injective functions are precisely the monomorphisms in the category Set of sets.
- If {\displaystyle f:X\to Y} is injective and {\displaystyle A} is a subset of {\displaystyle X,} then {\displaystyle f^{-1}(f(A))=A.} Thus, {\displaystyle A} can be recovered from its image {\displaystyle f(A).}
- If {\displaystyle f:X\to Y} is injective and {\displaystyle A} and {\displaystyle B} are both subsets of {\displaystyle X,} then {\displaystyle f(A\cap B)=f(A)\cap f(B).}
- Every function {\displaystyle h:W\to Y} can be decomposed as {\displaystyle h=f\circ g} for a suitable injection {\displaystyle f} and surjection {\displaystyle g.} This decomposition is unique up to isomorphism, and {\displaystyle f} may be thought of as the inclusion function of the range {\displaystyle h(W)} of {\displaystyle h} as a subset of the codomain {\displaystyle Y} of {\displaystyle h.}
- If {\displaystyle f:X\to Y} is an injective function, then {\displaystyle Y} has at least as many elements as {\displaystyle X,} in the sense of cardinal numbers. In particular, if, in addition, there is an injection from {\displaystyle Y} to {\displaystyle X,} then {\displaystyle X} and {\displaystyle Y} have the same cardinal number. (This is known as the Cantor–Bernstein–Schroeder theorem.)
- If both {\displaystyle X} and {\displaystyle Y} are finite with the same number of elements, then {\displaystyle f:X\to Y} is injective if and only if {\displaystyle f} is surjective (in which case {\displaystyle f} is bijective).
- An injective function which is a homomorphism between two algebraic structures is an embedding.
- Unlike surjectivity, which is a relation between the graph of a function and its codomain, injectivity is a property of the graph of the function alone; that is, whether a function {\displaystyle f} is injective can be decided by only considering the graph (and not the codomain) of {\displaystyle f.}
Proving that functions are injective
[edit ]A proof that a function {\displaystyle f} is injective depends on how the function is presented and what properties the function holds. For functions that are given by some formula there is a basic idea. We use the definition of injectivity, namely that if {\displaystyle f(x)=f(y),} then {\displaystyle x=y.}[7]
Here is an example: {\displaystyle f(x)=2x+3}
Proof: Let {\displaystyle f:X\to Y.} Suppose {\displaystyle f(x)=f(y).} So {\displaystyle 2x+3=2y+3} implies {\displaystyle 2x=2y,} which implies {\displaystyle x=y.} Therefore, it follows from the definition that {\displaystyle f} is injective.
There are multiple other methods of proving that a function is injective. For example, in calculus if {\displaystyle f} is a differentiable function defined on some interval, then it is sufficient to show that the derivative is always positive or always negative on that interval. In linear algebra, if {\displaystyle f} is a linear transformation it is sufficient to show that the kernel of {\displaystyle f} contains only the zero vector. If {\displaystyle f} is a function with finite domain it is sufficient to look through the list of images of each domain element and check that no image occurs twice on the list.
A graphical approach for a real-valued function {\displaystyle f} of a real variable {\displaystyle x} is the horizontal line test. If every horizontal line intersects the curve of {\displaystyle f(x)} in at most one point, then {\displaystyle f} is injective or one-to-one.
Gallery
[edit ]-
An injective non-surjective function (injection, not a bijection)
-
An injective surjective function (bijection)
-
A non-injective surjective function (surjection, not a bijection)
-
A non-injective non-surjective function (also not a bijection)
-
Not an injective function. Here {\displaystyle X_{1}} and {\displaystyle X_{2}} are subsets of {\displaystyle X,Y_{1}} and {\displaystyle Y_{2}} are subsets of {\displaystyle Y}: for two regions where the function is not injective because more than one domain element can map to a single range element. That is, it is possible for more than one {\displaystyle x} in {\displaystyle X} to map to the same {\displaystyle y} in {\displaystyle Y.}
-
Making functions injective. The previous function {\displaystyle f:X\to Y} can be reduced to one or more injective functions (say) {\displaystyle f:X_{1}\to Y_{1}} and {\displaystyle f:X_{2}\to Y_{2},} shown by solid curves (long-dash parts of initial curve are not mapped to anymore). Notice how the rule {\displaystyle f} has not changed – only the domain and range. {\displaystyle X_{1}} and {\displaystyle X_{2}} are subsets of {\displaystyle X,Y_{1}} and {\displaystyle Y_{2}} are subsets of {\displaystyle Y}: for two regions where the initial function can be made injective so that one domain element can map to a single range element. That is, only one {\displaystyle x} in {\displaystyle X} maps to one {\displaystyle y} in {\displaystyle Y.}
-
Injective functions. Diagramatic interpretation in the Cartesian plane, defined by the mapping {\displaystyle f:X\to Y,} where {\displaystyle y=f(x),} {\displaystyle X=} domain of function, {\displaystyle Y=} range of function , and {\displaystyle \operatorname {im} (f)} denotes image of {\displaystyle f.} Every one {\displaystyle x} in {\displaystyle X} maps to exactly one unique {\displaystyle y} in {\displaystyle Y.} The circled parts of the axes represent domain and range sets— in accordance with the standard diagrams above
See also
[edit ]- Bijection, injection and surjection – Properties of mathematical functions
- Injective metric space – Type of metric space
- Monotonic function – Order-preserving mathematical function
- Univalent function – Mathematical concept
Notes
[edit ]- ^ Sometimes one-one function, in Indian mathematical education. "Chapter 1:Relations and functions" (PDF). Archived (PDF) from the original on Dec 26, 2023 – via NCERT.
- ^ a b c "Injective, Surjective and Bijective". Math is Fun. Retrieved 2019年12月07日.
- ^ "Section 7.3 (00V5): Injective and surjective maps of presheaves". The Stacks project. Retrieved 2019年12月07日.
- ^ Farlow, S. J. "Section 4.2 Injections, Surjections, and Bijections" (PDF). Mathematics & Statistics - University of Maine. Archived from the original (PDF) on Dec 7, 2019. Retrieved 2019年12月06日.
- ^ "What are usual notations for surjective, injective and bijective functions?". Mathematics Stack Exchange. Retrieved 2024年11月24日.
- ^ Unlike the corresponding statement that every surjective function has a right inverse, this does not require the axiom of choice, as the existence of {\displaystyle a} is implied by the non-emptiness of the domain. However, this statement may fail in less conventional mathematics such as constructive mathematics. In constructive mathematics, the inclusion {\displaystyle \{0,1\}\to \mathbb {R} } of the two-element set in the reals cannot have a left inverse, as it would violate indecomposability, by giving a retraction of the real line to the set {0,1}.
- ^ Williams, Peter (Aug 21, 1996). "Proving Functions One-to-One". Department of Mathematics at CSU San Bernardino Reference Notes Page. Archived from the original on 4 June 2017.
References
[edit ]- Bartle, Robert G. (1976), The Elements of Real Analysis (2nd ed.), New York: John Wiley & Sons, ISBN 978-0-471-05464-1 , p. 17 ff.
- Halmos, Paul R. (1974), Naive Set Theory , New York: Springer, ISBN 978-0-387-90092-6 , p. 38 ff.