Anda di halaman 1dari 168

FUNCTIONS

PDF generated using the open source mwlib toolkit. See http://code.pediapress.com/ for more information.
PDF generated at: Tue, 25 Jan 2011 11:35:56 UTC
Contents
Articles
Additive function 1
Algebraic function 3
Analytic function 6
Antiholomorphic function 9
Arithmetic function 10
Bijection 20
Binary function 23
Bochner measurable function 24
Bounded function 25
Cauchy-continuous function 26
Closed convex function 27
Coarse function 27
Completely multiplicative function 28
Concave function 29
Constant function 31
Continuous function 32
Convex function 41
Differentiable function 45
Doubly periodic function 47
Elementary function 48
Elliptic function 49
Empty function 51
Entire function 52
Even and odd functions 54
Flat function 57
Function of a real variable 58
Function composition 58
Functional (mathematics) 61
Harmonic function 63
Hermitian function 68
Holomorphic function 69
Homogeneous function 72
Identity function 75
Implicit and explicit functions 76
Indicator function 79
Injective function 82
Invex function 85
List of types of functions 86
Locally integrable function 87
Measurable function 89
Meromorphic function 91
Monotonic function 93
Multiplicative function 97
Multivalued function 99
Negligible function 102
Nowhere continuous function 103
Periodic function 104
Piecewise linear function 107
Pluriharmonic function 109
Plurisubharmonic function 110
Polyconvex function 112
Positive-definite function 112
Proper convex function 114
Pseudoconvex function 115
Quasi-analytic function 117
Quasiconvex function 118
Quasiperiodic function 121
Radially unbounded function 122
Rational function 122
Real-valued function 126
Ring of symmetric functions 126
Simple function 132
Single-valued function 133
Singular function 133
Smooth function 134
Subharmonic function 140
Sublinear function 143
Surjective function 144
Symmetrically continuous function 147
Quasisymmetric function 147
Transcendental function 150
Unary function 152
Univalent function 152
Vector-valued function 153
Weakly harmonic function 158
Weakly measurable function 158

References
Article Sources and Contributors 159
Image Sources, Licenses and Contributors 163

Article Licenses
License 164
Additive function 1

Additive function
In mathematics the term additive function has two different definitions, depending on the specific field of
application.
In algebra an additive function (or additive map) is a function that preserves the addition operation:
f(x + y) = f(x) + f(y)
for any two elements x and y in the domain. For example, any linear map is additive. When the domain is the real
numbers, this is Cauchy's functional equation.
In number theory, an additive function is an arithmetic function f(n) of the positive integer n such that whenever a
and b are coprime, the function of the product is the sum of the functions:
f(ab) = f(a) + f(b).
The remainder of this article discusses number theoretic additive functions, using the second definition. For a
specific case of the first definition see additive polynomial. Note also that any homomorphism f between Abelian
groups is "additive" by the first definition.

Completely additive
An additive function f(n) is said to be completely additive if f(ab) = f(a) + f(b) holds for all positive integers a and
b, even when they are not co-prime. Totally additive is also used in this sense by analogy with totally multiplicative
functions. If f is a completely additive function then f(1) = 0.
Every completely additive function is additive, but not vice versa.

Examples
Example of arithmetic functions which are completely additive are:
• The restriction of the logarithmic function to N.
• The multiplicity of a prime factor p in n, that is the largest exponent m for which pm divides n.
• a0(n) - the sum of primes dividing n counting multiplicity, sometimes called sopfr(n), the potency of n or the
integer logarithm of n (sequence A001414 [1] in OEIS). For example:
a0(4) = 2 + 2 = 4
a0(20) = a0(22 · 5) = 2 + 2+ 5 = 9
a0(27) = 3 + 3 + 3 = 9
a0(144) = a0(24 · 32) = a0(24) + a0(32) = 8 + 6 = 14
a0(2,000) = a0(24 · 53) = a0(24) + a0(53) = 8 + 15 = 23
a0(2,003) = 2003
a0(54,032,858,972,279) = 1240658
a0(54,032,858,972,302) = 1780417
a0(20,802,650,704,327,415) = 1240681
• The function Ω(n), defined as the total number of prime factors of n, counting multiple factors multiple times,
sometimes called the "Big Omega function" (sequence A001222 [2] in OEIS). For example;
Ω(1) = 0, since 1 has no prime factors
Ω(20) = Ω(2·2·5) = 3
Ω(4) = 2
Additive function 2

Ω(27) = 3
Ω(144) = Ω(24 · 32) = Ω(24) + Ω(32) = 4 + 2 = 6
Ω(2,000) = Ω(24 · 53) = Ω(24) + Ω(53) = 4 + 3 = 7
Ω(2,001) = 3
Ω(2,002) = 4
Ω(2,003) = 1
Ω(54,032,858,972,279) = 3
Ω(54,032,858,972,302) = 6
Ω(20,802,650,704,327,415) = 7
Example of arithmetic functions which are additive but not completely additive are:
• ω(n), defined as the total number of different prime factors of n (sequence A001221 [3] in OEIS). For example:
ω(4) = 1
ω(20) = ω(22·5) = 2
ω(27) = 1
ω(144) = ω(24 · 32) = ω(24) + ω(32) = 1 + 1 = 2
ω(2,000) = ω(24 · 53) = ω(24) + ω(53) = 1 + 1 = 2
ω(2,001) = 3
ω(2,002) = 4
ω(2,003) = 1
ω(54,032,858,972,279) = 3
ω(54,032,858,972,302) = 5
ω(20,802,650,704,327,415) = 5
• a1(n) - the sum of the distinct primes dividing n, sometimes called sopf(n) (sequence A008472 [4] in OEIS). For
example:
a1(1) = 0
a1(4) = 2
a1(20) = 2 + 5 = 7
a1(27) = 3
a1(144) = a1(24 · 32) = a1(24) + a1(32) = 2 + 3 = 5
a1(2,000) = a1(24 · 53) = a1(24) + a1(53) = 2 + 5 = 7
a1(2,001) = 55
a1(2,002) = 33
a1(2,003) = 2003
a1(54,032,858,972,279) = 1238665
a1(54,032,858,972,302) = 1780410
a1(20,802,650,704,327,415) = 1238677
Additive function 3

Multiplicative functions
From any additive function f(n) it is easy to create a related multiplicative function g(n) i.e. with the property that
whenever a and b are coprime we have:
g(ab) = g(a) × g(b).
One such example is g(n) = 2f(n).

See also
• Sigma additivity

Further reading
• Janko Bračič, Kolobar aritmetičnih funkcij (Ring of arithmetical functions), (Obzornik mat, fiz. 49 (2002) 4, pp.
97–108) (MSC (2000) 11A25)

References
[1] http:/ / en. wikipedia. org/ wiki/ Oeis%3Aa001414
[2] http:/ / en. wikipedia. org/ wiki/ Oeis%3Aa001222
[3] http:/ / en. wikipedia. org/ wiki/ Oeis%3Aa001221
[4] http:/ / en. wikipedia. org/ wiki/ Oeis%3Aa008472

Algebraic function
In mathematics, an algebraic function is informally a function that satisfies a polynomial equation whose
coefficients are themselves polynomials. For example, an algebraic function in one variable x is a solution y for an
equation

where the coefficients ai(x) are polynomial functions of x. A function which is not algebraic is called a
transcendental function.
In more precise terms, an algebraic function may not be a function at all, at least not in the conventional sense.
Consider for example the equation of a circle:

This determines y, except only up to an overall sign:

However, both branches are thought of as belonging to the "function" determined by the polynomial equation. Thus
an algebraic function is most naturally considered as a multiple valued function.
An algebraic function in n variables is similarly defined as a function y which solves a polynomial equation in
n + 1 variables:

It is normally assumed that p should be an irreducible polynomial. The existence of an algebraic function is then
guaranteed by the implicit function theorem.
Formally, an algebraic function in n variables over the field K is an element of the algebraic closure of the field of
rational functions K(x1,...,xn). In order to understand algebraic functions as functions, it becomes necessary to
introduce ideas relating to Riemann surfaces or more generally algebraic varieties, and sheaf theory.
Algebraic function 4

Algebraic functions in one variable

Introduction and overview


The informal definition of an algebraic function provides a number of clues about the properties of algebraic
functions. To gain an intuitive understanding, it may be helpful to regard algebraic functions as functions which can
be formed by the usual algebraic operations: addition, multiplication, division, and taking an nth root. Of course, this
is something of an oversimplification; because of casus irreducibilis (and more generally the fundamental theorem of
Galois theory), algebraic functions need not be expressible by radicals.
First, note that any polynomial is an algebraic function, since polynomials are simply the solutions for y of the
equation

More generally, any rational function is algebraic, being the solution of

Moreover, the nth root of any polynomial is an algebraic function, solving the equation

Surprisingly, the inverse function of an algebraic function is an algebraic function. For supposing that y is a solution
of

for each value of x, then x is also a solution of this equation for each value of y. Indeed, interchanging the roles of x
and y and gathering terms,

Writing x as a function of y gives the inverse function, also an algebraic function.


However, not every function has an inverse. For example, y = x2 fails the horizontal line test: it fails to be
one-to-one. The inverse is the algebraic "function" . In this sense, algebraic functions are often not true
functions at all, but instead are multiple valued functions.
Another way to understand this, which will become important later in the article, is that an algebraic function is the
graph of an algebraic curve.

The role of complex numbers


From an algebraic perspective, complex numbers enter quite naturally into the study of algebraic functions. First of
all, by the fundamental theorem of algebra, the complex numbers are an algebraically closed field. Hence any
polynomial relation
p(y, x) = 0
is guaranteed to have at least one solution (and in general a number of solutions not exceeding the degree of p in x)
for y at each point x, provided we allow y to assume complex as well as real values. Thus, problems to do with the
domain of an algebraic function can safely be minimized.
Algebraic function 5

Furthermore, even if one is ultimately interested in real algebraic


functions, there may be no adequate means to express the function in a
simple manner without resorting to complex numbers (see casus
irreducibilis). For example, consider the algebraic function determined
by the equation

Using the cubic formula, one solution is (the red curve in the
A graph of three branches of the algebraic
accompanying image)
function y, where y3 − xy + 1 = 0, over the
domain 3/22/3 < x < 50.

There is no way to express this function in terms of real numbers only, even though the resulting function is
real-valued on the domain of the graph shown.
On a more significant theoretical level, using complex numbers allow one to use the powerful techniques of complex
analysis to discuss algebraic functions. In particular, the argument principle can be used to show that any algebraic
function is in fact an analytic function, at least in the multiple-valued sense.
Formally, let p(x, y) be a complex polynomial in the complex variables x and y. Suppose that x0 ∈ C is such that the
polynomial p(x0,y) of y has n distinct zeros. We shall show that the algebraic function is analytic in a neighborhood
of x0. Choose a system of n non-overlapping discs Δi containing each of these zeros. Then by the argument principle

By continuity, this also holds for all x in a neighborhood of x0. In particular, p(x,y) has only one root in Δi, given by
the residue theorem:

which is an analytic function.

Monodromy
Note that the foregoing proof of analyticity derived an expression for a system of n different function elements fi(x),
provided that x is not a critical point of p(x, y). A critical point is a point where the number of distinct zeros is
smaller than the degree of p, and this occurs only where the highest degree term of p vanishes, and where the
discriminant vanishes. Hence there are only finitely many such points c1, ..., cm.
A close analysis of the properties of the function elements fi near the critical points can be used to show that the
monodromy cover is ramified over the critical points (and possibly the point at infinity). Thus the entire function
associated to the fi has at worst algebraic poles and ordinary algebraic branchings over the critical points.
Note that, away from the critical points, we have

since the fi are by definition the distinct zeros of p. The monodromy group acts by permuting the factors, and thus
forms the monodromy representation of the Galois group of p. (The monodromy action on the universal covering
space is related but different notion in the theory of Riemann surfaces.)
Algebraic function 6

History
The ideas surrounding algebraic functions go back at least as far as René Descartes. The first discussion of algebraic
functions appears to have been in Edward Waring's 1794 An Essay on the Principles of Human Knowledge in which
he writes:
let a quantity denoting the ordinate, be an algebraic function of the abscissa x, by the common methods of
division and extraction of roots, reduce it into an infinite series ascending or descending according to the
dimensions of x, and then find the integral of each of the resulting terms.

References
• Ahlfors, Lars (1979). Complex Analysis. McGraw Hill.
• van der Waerden, B.L. (1931). Modern Algebra, Volume II. Springer.

Analytic function
This article is about both real and complex analytic functions. The article holomorphic function is solely about
analytic functions in complex analysis. An analytic signal is a signal with no negative-frequency components.
In mathematics, an analytic function is a function that is locally given by a convergent power series. There exist
both real analytic functions and complex analytic functions, categories that are similar in some ways, but different
in others. Functions of each type are infinitely differentiable, but complex analytic functions exhibit properties that
do not hold generally for real analytic functions. A function is analytic if and only if it is equal to its Taylor series in
some neighborhood of every point.

Definitions
Formally, a function ƒ is real analytic on an open set D in the real line if for any x0 in D one can write

in which the coefficients a0, a1, ... are real numbers and the series is convergent to ƒ(x) for x in a neighborhood of x0.
Alternatively, an analytic function is an infinitely differentiable function such that the Taylor series at any point x0 in
its domain

converges to ƒ(x) for x in a neighborhood of x0. The set of all real analytic functions on a given set D is often denoted
by Cω(D).
A function ƒ defined on some subset of the real line is said to be real analytic at a point x if there is a neighborhood D
of x on which ƒ is real analytic.
The definition of a complex analytic function is obtained by replacing, in the definitions above, "real" with
"complex" and "real line" with "complex plane."
Analytic function 7

Examples
Most special functions are analytic (at least in some range of the complex plane). Typical examples of analytic
functions are:
• Any polynomial (real or complex) is an analytic function. This is because if a polynomial has degree n, any terms
of degree larger than n in its Taylor series expansion will vanish, and so this series will be trivially convergent.
Furthermore, every polynomial is its own Maclaurin series.
• The exponential function is analytic. Any Taylor series for this function converges not only for x close enough to
x0 (as in the definition) but for all values of x (real or complex).
• The trigonometric functions, logarithm, and the power functions are analytic on any open set of their domain.
Typical examples of functions that are not analytic are:
• The absolute value function when defined on the set of real numbers or complex numbers is not everywhere
analytic because it is not differentiable at 0. Piecewise defined functions (functions given by different formulas in
different regions) are typically not analytic where the pieces meet.
• The complex conjugate function is not complex analytic, although its restriction to the real line is the
identity function and therefore real analytic.

Alternate characterizations
If ƒ is an infinitely differentiable function defined on an open set , then the following conditions are
equivalent.
1) ƒ is real analytic.
2) There is a complex analytic extension of ƒ to an open set which contains .
3) For every compact set there exists a constant such that for every and every
non-negative integer k the following estimate holds:

The real analyticity of a function ƒ at a given point x can be characterized using the FBI transform.
Complex analytic functions are exactly equivalent to holomorphic functions, and are thus much more easily
characterized.

Properties of analytic functions


• The sums, products, and compositions of analytic functions are analytic.
• The reciprocal of an analytic function that is nowhere zero is analytic, as is the inverse of an invertible analytic
function whose derivative is nowhere zero. (See also the Lagrange inversion theorem.)
• Any analytic function is smooth, that is, infinitely differentiable. The converse is not true; in fact, in a certain
sense, the analytic functions are sparse compared to all infinitely differentiable functions.
• For any open set Ω ⊆ C, the set A(Ω) of all analytic functions u : Ω → C is a Fréchet space with respect to the
uniform convergence on compact sets. The fact that uniform limits on compact sets of analytic functions are
analytic is an easy consequence of Morera's theorem. The set of all bounded analytic functions with the
supremum norm is a Banach space.
A polynomial cannot be zero at too many points unless it is the zero polynomial (more precisely, the number of zeros
is at most the degree of the polynomial). A similar but weaker statement holds for analytic functions. If the set of
zeros of an analytic function ƒ has an accumulation point inside its domain, then ƒ is zero everywhere on the
connected component containing the accumulation point. In other words, if (rn) is a sequence of distinct numbers
Analytic function 8

such that ƒ(rn) = 0 for all n and this sequence converges to a point r in the domain of D, then ƒ is identically zero on
the connected component of D containing r.
Also, if all the derivatives of an analytic function at a point are zero, the function is constant on the corresponding
connected component.
These statements imply that while analytic functions do have more degrees of freedom than polynomials, they are
still quite rigid.

Analyticity and differentiability


As noted above, any analytic function (real or complex) is infinitely differentiable (also known as smooth, or C∞).
(Note that this differentiability is in the sense of real variables; compare complex derivatives below.) There exist
smooth real functions which are not analytic: see non-analytic smooth function. In fact there are many such
functions, and the space of real analytic functions is a proper subspace of the space of smooth functions.
The situation is quite different when one considers complex analytic functions and complex derivatives. It can be
proved that any complex function differentiable (in the complex sense) in an open set is analytic. Consequently, in
complex analysis, the term analytic function is synonymous with holomorphic function.

Real versus complex analytic functions


Real and complex analytic functions have important differences (one could notice that even from their different
relationship with differentiability). Analyticity of complex functions is a more restrictive property, as it has more
restrictive necessary conditions and complex analytic functions have more structure than their real-line
counterparts.[1]
According to Liouville's theorem, any bounded complex analytic function defined on the whole complex plane is
constant. The corresponding statement for real analytic functions, with the complex plane replaced by the real line, is
clearly false; this is illustrated by

Also, if a complex analytic function is defined in an open ball around a point x0, its power series expansion at x0 is
convergent in the whole ball. This statement for real analytic functions (with open ball meaning an open interval of
the real line rather than an open disk of the complex plane) is not true in general; the function of the example above
gives an example for x0 = 0 and a ball of radius exceeding 1, since the power series 1 − x2 + x4 − x6... diverges for
|x| > 1.
Any real analytic function on some open set on the real line can be extended to a complex analytic function on some
open set of the complex plane. However, not every real analytic function defined on the whole real line can be
extended to a complex function defined on the whole complex plane. The function ƒ (x) defined in the paragraph
above is a counterexample, as it is not defined for x = ±i.
Analytic function 9

Analytic functions of several variables


One can define analytic functions in several variables by means of power series in those variables (see power series).
Analytic functions of several variables have some of the same properties as analytic functions of one variable.
However, especially for complex analytic functions, new and interesting phenomena show up when working in 2 or
more dimensions. For instance, zero sets of complex analytic functions in more than one variable are never discrete.

Notes
[1] Krantz, Steven; Harold R., Parks (2002), A Primer of Real Analytic Functions (Second ed.), Birkhäuser, ISBN 0817642641,
ISBN 0-8176-4264-1, 3-7643-4264-1

References
• Conway, John B. (1978). Functions of One Complex Variable I (Graduate Texts in Mathematics 11).
Springer-Verlag. ISBN 0-387-90328-3.
• Krantz, Steven; Harold R., Parks (2002), A Primer of Real Analytic Functions (Second ed.), Birkhäuser, ISBN
0817642641, ISBN 0-8176-4264-1, 3-7643-4264-1

External links
• Weisstein, Eric W., " Analytic Function (http://mathworld.wolfram.com/AnalyticFunction.html)" from
MathWorld.
• Analytic Functions Module by John H. Mathews (http://math.fullerton.edu/mathews/c2003/
AnalyticFunctionMod.html)

Antiholomorphic function
In mathematics, antiholomorphic functions (also called antianalytic functions) are a family of functions closely
related to but distinct from holomorphic functions.
A function of the complex variable z defined on an open set in the complex plane is said to be antiholomorphic if
its derivative with respect to z* exists in the neighbourhood of each and every point in that set, where z* is the
complex conjugate.
One can show that if f(z) is a holomorphic function on an open set D, then f(z*) is an antiholomorphic function on
D*, where D* is the reflection against the x-axis of D, or in other words, D* is the set of complex conjugates of
elements of D. Moreover, any antiholomorphic function can be obtained in this manner from a holomorphic
function. This implies that a function is antiholomorphic if and only if it can be expanded in a power series in z* in a
neighborhood of each point in its domain.
If a function is both holomorphic and antiholomorphic, then it is constant on any connected component of its
domain.
Arithmetic function 10

Arithmetic function
In number theory, an arithmetic (or arithmetical) function is a real or complex valued function ƒ(n) defined on the
set of natural numbers (i.e. positive integers) that "expresses some arithmetical property of n."[1]
An example of an arithmetic function is the non-principal character (mod 4) defined by

    where is the Kronecker symbol.

To emphasize that they are being thought of as functions rather than sequences, values of an arithmetic function are
usually denoted by a(n) rather than an.
There is a larger class of number-theoretic functions that do not fit the above definition, e.g. the prime-counting
functions. This article provides links to functions of both classes.

Notation
  and     mean that the sum or product is over all prime numbers:

  

Similarly,     and     mean that the sum or product is over all prime powers with strictly positive

exponent (so 1 is not counted):

  and     mean that the sum or product is over all positive divisors of n, including 1 and n. E.g., if

n = 12,

The notations can be combined:     and     mean that the sum or product is over all prime

divisors of n. E.g., if n = 18,

and similarly     and     mean that the sum or product is over all prime powers dividing n.

E.g., if n = 24,
Arithmetic function 11

Multiplicative and additive functions


An arithmetic function a is
• completely additive if a(mn) = a(m) + a(n) for all natural numbers m and n;
• completely multiplicative if a(mn) = a(m)a(n) for all natural numbers m and n;
Two whole numbers m and n are called coprime if their greatest common divisor is 1; i.e., if there is no prime
number that divides both of them.
Then an arithmetic function a is
• additive if a(mn) = a(m) + a(n) for all coprime natural numbers m and n;
• multiplicative if a(mn) = a(m)a(n) for all coprime natural numbers m and n.

Ω(n), ω(n), νp(n) - prime power decomposition


The fundamental theorem of arithmetic states that any positive integer n can be represented uniquely as a product of
powers of primes:     where p1 < p2 < ... < pk are primes and the aj are positive integers. (1 is given
by the empty product.)
It is often convenient to write this as an infinite product over all the primes, where all but a finite number have a zero
exponent. Define νp(n) as the exponent of the highest power of the prime p that divides n. I.e. if p is one of the pi
then νp(n) = ai, otherwise it is zero. Then

In terms of the above the functions ω and Ω are defined by


ω(n) = k,
Ω(n) = a1 + a2 + ... + ak.
To avoid repetition, whenever possible formulas for the functions listed in this article are given in terms of n and the
corresponding pi, ai, ω, and Ω.

Multiplicative functions

σk(n), τ(n), d(n) - divisor sums


σk(n) is the sum of the kth powers of the positive divisors of n, including 1 and n, where k is a complex number.
σ1(n), the sum of the (positive) divisors of n, is usually denoted by σ(n).
Since a positive number to the zero power is one, σ0(n) is therefore the number of (positive) divisors of n; it is
usually denoted by d(n) or τ(n) (for the German Teiler = divisors).

Setting k = 0 in the second product gives


Arithmetic function 12

φ(n) - Euler totient function


φ(n), the Euler totient function, is the number of positive integers not greater than n that are coprime to n.

μ(n) - Möbius function


μ(n), the Möbius function, is important because of the Möbius inversion formula. See Dirichlet convolution, below.

This implies that μ(1) = 1. (Because Ω(1) = ω(1) = 0.)

τ(n) - Ramanujan tau function


τ(n), the Ramanujan tau function, is defined by its generating function identity:

Although it is hard to say exactly what "arithmetical property of n" it "expresses",[2] (τ(n) is (2π)−12 times the nth
Fourier coefficient in the q-expansion of the modular discriminant function)[3] it is included among the arithmetical
functions because it is multiplicative and it occurs in identities involving certain σk(n) and rk(n) functions (because
these are also coefficients in the expansion of modular forms).

cq(n) - Ramanujan's sum


cq(n), Ramanujan's sum, is the sum of the nth powers of the primitive qth roots of unity:

Even though it is defined as a sum of complex numbers (irrational for most values of q), it is an integer. For a fixed
value of n it is multiplicative in q:
If q and r are coprime,
Many of the functions mentioned in this article have expansions as series involving these sums; see the article
Ramanujan's sum for examples.

Completely multiplicative functions

λ(n) - Liouville function


λ(n), the Liouville function, is defined by

χ(n) - characters
All Dirichlet characters χ(n) are completely multiplicative; e.g. the non-principal character (mod 4) defined in the
introduction, or the principal character (mod n) defined by
Arithmetic function 13

Additive functions

ω(n) - distinct prime divisors


ω(n), defined above as the number of distinct primes dividing n, is additive

Completely additive functions

Ω(n) - prime divisors


Ω(n), defined above as the number of prime factors of n counted with multiplicities, is completely additive.

νp(n) - prime power dividing n


For a fixed prime p, νp(n), defined above as the exponent of the largest power of p dividing n, is completely additive.

Neither multiplicative nor additive

π(x), Π(x), θ(x), ψ(x) - prime count functions


Unlike the other functions listed in this article, these are defined for non-negative real (not just integer) arguments.
They are used in the statement and proof of the prime number theorem.
π(x), the prime counting function, is the number of primes not exceeding x.

A related function counts prime powers with weight 1 for primes, 1/2 for their squares, 1/3 for cubes, ...

θ(x) and ψ(x), the Chebyshev functions are defined as sums of the natural logarithms of the primes not exceeding x:

Λ(n) - von Mangoldt function


Λ(n), the von Mangoldt function, is 0 unless the argument is a prime power, in which case it is the natural log of the
prime:

p(n) - partition function


p(n), the partition function, is the number of ways of representing n as a sum of positive integers, where two
representations with the same summands in a different order are not counted as being different:
Arithmetic function 14

λ(n) - Carmichael function


λ(n), the Carmichael function, is the smallest positive number such that   for all a coprime
to n. Equivalently, it is the least common multiple of the orders of the elements of the multiplicative group of
integers modulo n.
For powers of odd primes and for 2 and 4, λ(n) is equal to the Euler totient function of n; for powers of 2 greater than
4 it is equal to one half of the Euler totient function of n:

and for general n it is the least common multiple of λ of each of the prime power factors of n:

h(n) - Class number


h(n), the class number function, is the order of the ideal class group of an algebraic extension of the rationals with
discriminant n. The notation is ambiguous, as there are in general many extensions with the same discriminant. See
quadratic field and cyclotomic field for classical examples.

rk(n) - Sum of k squares


rk(n) is the number of ways n can be represented as the sum of k squares, where representations that differ only in the
order of the summands or in the signs of the square roots are counted as different.

Summation functions
Given an arithmetic function a(n), its summation function A(x) is defined by

A can be regarded as a function of a real variable. Given a positive integer m, A is constant along open intervals m <
x < m + 1, and has a jump discontinuity at each integer for which a(m) ≠ 0.
Since such functions are often represented by series and integrals, to achieve pointwise convergence it is usual to
define the value at the discontinuities as the average of the values to the left and right:

Individual values of arithmetic functions may fluctuate wildly - as in most of the above examples. Summation
functions "smooth out" these fluctuations. In some cases it may be possible to find asymptotic behaviour for the
summation function for large x.
A classical example of this phenomenon[4] is given by d(n), the number of divisors of n:
Arithmetic function 15

Dirichlet convolution
Given an arithmetic function a(n), let Fa(s), for complex s, be the function defined by the corresponding Dirichlet
series (where it converges):[5]

Fa(s) is called a generating function of a(n). The simplest such series, corresponding to the constant function a(n) = 1
for all n, is ς(s) the Riemann zeta function.
The generating function of the Möbius function is the inverse of the zeta function:

Consider two arithmetic functions a and b and their respective generating functions Fa(s) and Fb(s). The product
Fa(s)Fb(s) can be computed as follows:

It is a straightforward exercise to show that if c(n) is defined by

then

This function c is called the Dirichlet convolution of a and b, and is denoted by .


A particularly important case is convolution with the constant function a(n) = 1 for all n, corresponding to
multiplying the generating function by the zeta function:

Multiplying by the inverse of the zeta function gives the Möbius inversion formula:

If f is multiplicative, then so is g. If f is completely multiplicative, then g is multiplicative, but may or may not be
completely multiplicative. The article multiplicative function has a short proof.

Relations among the functions


There are a great many formulas connecting arithmetical functions with each other and with the other functions of
analysis - in fact, a large part of elementary and analytic number theory is a detailed study of these relations.
Here are a few examples:

Dirichlet convolutions

    where λ is the Liouville function.    [6]

     [7]
Arithmetic function 16

      Möbius inversion

     [8]

     [9]

      Möbius inversion

      Möbius inversion

      Möbius inversion

    where λ is the Liouville function.

     [10]

      Möbius inversion

Sums of squares
    (Lagrange's four-square theorem).

    where χ is the non-principal character (mod 4) defined in the introduction.[11]

    where ν = ν2(n).    [12]

[13] [14]

   [15]

Define the function σk*(n) as[16]

That is, if n is odd, σk*(n) is the sum of the kth powers of the divisors of n, i.e. σk(n), and if n is even it is the sum of
the kth powers of the even divisors of n minus the sum of the kth powers of the odd divisors of n.
   [15] [17]
Adopt the convention that Ramanujan's τ(x) = 0 if x is not an integer.
Arithmetic function 17

   [18]

Divisor sum convolutions


Here "convolution" does not mean "Dirichlet convolution" but instead refers to the formula for the coefficients of the
product of two power series:

The sequence is called the convolution or the Cauchy product of the seqences an and bn. See

Eisenstein series for a discussion of the series and functional identities involved in these formulas.

   [19]

   [20]

[20] [21]

[19] [22]

    where τ(n) is Ramanujan's

function.    [23] [24]


Since σk(n) (for natural number k) and τ(n) are integers, the above formulas can be used to prove congruences[25] for
the functions. See Tau-function for some examples.
Extend the domain of the partition function by setting p(0) = 1.

   [26]   This recurrence can be used to compute p(n).

Prime-count related

Let   be the nth harmonic number.   Then

  is true for every natural number n if and only if the Riemann hypothesis is
true.    [27]

   [28]
Arithmetic function 18

   [29]

   [30]

   [31]

Miscellaneous

    and         where λ(n) is Liouville's function.

   [32] [33]
    where λ(n) is Carmichael's function. Further,

     [34]

   [35]

[36]
    Note that      [37]

   [38]   Compare this with 13 + 23 + 33 + ... + n3 = (1 + 2 + 3 + ... + n)2

   [39]

   [40]

    where τ(n) is Ramanujan's function.    [41]


Arithmetic function 19

Notes
[1] Hardy & Wright, intro. to Ch. XVI
[2] Hardy, Ramanujan, § 10.2
[3] Apostol, Modular Functions ..., § 1.15, Ch. 4, and ch. 6
[4] Hardy & Wright, §§ 18.1–18.2
[5] Hardy & Wright, § 17.6, show how the theory of generating functions can be constructed in a purely formal manner with no attention paid to
convergence.
[6] Hardy & Wright, Thm. 263
[7] Hardy & Wright, Thm. 63
[8] Hardy & Wright, Thm. 288–290
[9] Hardy & Wright, Thm. 264
[10] Hardy & Wright, Thm. 296
[11] Hardy & Wright, Thm. 278
[12] Hardy & Wright, Thm. 386
[13] Hardy, Ramanujan, eqs 9.1.2, 9.1.3
[14] Koblitz, Ex. III.5.2
[15] Hardy & Wright, § 20.13
[16] Hardy, Ramanujan, § 9.7
[17] Hardy, Ramanujan, § 9.13
[18] Hardy, Ramanujan, § 9.17
[19] Ramanujan, On Certain Arithmetical Functions, Table IV; Papers, p. 146
[20] Koblitz, ex. III.2.8
[21] Koblitz, ex. III.2.3
[22] Koblitz, ex. III.2.2
[23] Koblitz, ex. III.2.4
[24] Apostol, Modular Functions ..., Ex. 6.10
[25] Apostol, Modular Functions..., Ch. 6 Ex. 10
[26] G.H. Hardy, S. Ramannujan, Asymptotic Formulæ in Combinatory Analysis, § 1.3; in Ramannujan, Papers p. 279
[27] See Divisor function.
[28] Hardy & Wright, eq. 22.1.2
[29] See prime counting functions.
[30] Hardy & Wright, eq. 22.1.1
[31] Hardy & Wright, eq. 22.1.3
[32] Hardy Ramanujan, eq. 3.10.3
[33] Hardy & Wright, § 22.13
[34] See Multiplicative group of integers modulo n and Primitive root modulo n.
[35] Hardy & Wright, Thm. 329
[36] Hardy & Wright, Thms. 271, 272
[37] Hardy & Wright, eq. 16.3.1
[38] Ramanujan, Some Formulæ in the Analytic Theory of Numbers, eq. (C); Papers p.133
[39] Ramanujan, Some Formulæ in the Analytic Theory of Numbers, eq. (F); Papers p.134
[40] Apostol, Modular Functions ..., ch. 6 eq. 4
[41] Apostol, Modular Functions ..., ch. 6 eq. 3

References
• Tom M. Apostol (1976), Introduction to Analytic Number Theory, Springer Undergraduate Texts in Mathematics,
ISBN 0387901639
• Apostol, Tom M. (1989), Modular Functions and Dirichlet Series in Number Theory (2nd Edition), New York:
Springer, ISBN 0-387-97127
• Hardy, G. H. (1999), Ramanujan: Twelve Lectures on Subjects Suggested by his Life and work, Providence RI:
AMS / Chelsea, ISBN 978-0821820230
• Hardy, G. H.; Wright, E. M. (1980), An Introduction to the Theory of Numbers (Fifth edition), Oxford: Oxford
University Press, ISBN 978-0198531715
• G. J. O. Jameson (2003), The Prime Number Theorem, Cambridge University Press, ISBN 0-521-89110-8
Arithmetic function 20

• Koblitz, Neal (1984), Introduction to Elliptic Curves and Modular Forms, New York: Springer,
ISBN 0-387-97966-2
• William J. LeVeque (1996), Fundamentals of Number Theory, Courier Dover Publications, ISBN 0486689069
• Elliott Mendelson (1987), Introduction to Mathematical Logic, CRC Press, ISBN 0412808307
• Ramanujan, Srinivasa (2000), Collected Papers, Providence RI: AMS / Chelsea, ISBN 978-0821820766

External links
• Elementary Evaluation of Certain Convolution Sums Involving Divisor Functions (http://mathstat.carleton.ca/
~williams/papers/pdf/249.pdf) PDF of a paper by Huard, Ou, Spearman, and Williams. Contains elementary
(i.e. not relying on the theory of modular forms) proofs of divisor sum convolutions, formulas for the number of
ways of representing a number as a sum of triangular numbers, and related results.

Bijection
In mathematics, a bijection, or a bijective function, is a function f
from a set X to a set Y with the property that, for every y in Y, there
is exactly one x in X such that f(x) = y. It follows from this
definition that no unmapped element exists in either X or Y.
Alternatively, f is bijective if it is a one-to-one correspondence
between those sets; i.e., both one-to-one (injective) and onto
(surjective).
For example, consider the function succ, defined from the set of
integers to , that to each integer x associates the integer
succ(x) = x + 1. For another example, consider the function sumdif
that to each pair (x,y) of real numbers associates the pair
sumdif(x,y) = (x + y, x − y).
A bijective function, f:X→Y, where set X is {1,2,3,4}
A bijective function from a set to itself is also called a
and set Y is {A,B,C,D}. For example, f(1)=D.
permutation.
The set of all bijections from X to Y is denoted as X ↔ Y. (Sometimes this notation is reserved for binary relations,
and bijections are denoted by X ⤖ Y instead.) Occasionally, the set of permutations of a single set X may be denoted
X!.
Bijective functions play a fundamental role in many areas of mathematics, for instance in the definition of
isomorphism (and related concepts such as homeomorphism and diffeomorphism), permutation group, projective
map, and many others.
Bijection 21

Composition and inverses


A function f is bijective if and only if its inverse relation f −1 is a function. In that case, f −1 is also a bijection.
The composition of two bijections and is a bijection. The inverse of is
.
Conversely, if the composition of
two functions is bijective, we can only say
that f is injective and g is surjective.
A relation f from X to Y is a bijective
function if and only if there exists another
relation g from Y to X such that is
the identity function on X, and is the
identity function on Y. Consequently, the
sets have the same cardinality.

Bijections and cardinality


If X and Y are finite sets, then there exists a A bijection composed of an injection (left) and a surjection (right).
bijection between the two sets X and Y iff X
and Y have the same number of elements. Indeed, in axiomatic set theory, this is taken as the very definition of "same
number of elements", and generalising this definition to infinite sets leads to the concept of cardinal number, a way
to distinguish the various sizes of infinite sets.

Examples and counterexamples


• For any set X, the identity function idX from X to X, defined by idX(x) = x, is bijective.
• The function f from the real line R to R defined by f(x) = 2x + 1 is bijective, since for each y there is a unique x =
(y − 1)/2 such that f(x) = y.
• The exponential function g : R R, with g(x) = ex, is not bijective: for instance, there is no x in R such that g(x)
= −1, showing that g is not surjective. However if the codomain is restricted to the positive real numbers R+ =
(0,+∞), then g becomes bijective; its inverse is the natural logarithm function ln.
• The function h : R → [0,+∞) with h(x) = x² is not bijective: for instance, h(−1) = h(+1) = 1, showing that h is not
injective. However, if the domain is restricted to [0,+∞), then h becomes bijective; its inverse is the positive
square root function.

Properties
• A function f from the real line R to R is bijective if and only if its plot is intersected by any horizontal or vertical
line at exactly one point.
• If X is a set, then the bijective functions from X to itself, together with the operation of functional composition (∘),
form a group, the symmetric group of X, which is denoted variously by S(X), SX, or X! (the last reads "X
factorial").
• For a subset A of the domain with cardinality |A| and subset B of the codomain with cardinality |B|, one has the
following equalities:
|f(A)| = |A| and |f−1(B)| = |B|.
• If X and Y are finite sets with the same cardinality, and f: X → Y, then the following are equivalent:
1. f is a bijection.
Bijection 22

2. f is a surjection.
3. f is an injection.
• At least for a finite set S, there is a bijection between the set of possible total orderings of the elements and the set
of bijections from S to S. That is to say, the number of permutations of elements of S is the same as the number of
total orderings of that set—namely, n!.

Bijections and category theory


Formally, bijections are precisely the isomorphisms in the category Set of sets and functions. However, the
bijections are not always the isomorphisms. For example, in the category Top of topological spaces and continuous
functions, the isomorphisms must be homeomorphisms in addition to being bijections.

See also
• Category theory
• Injective function
• Symmetric group
• Surjective function
• Bijective numeration
• Bijective proof
• Cardinality

References
Weisstein, Eric W., "Bijection [1]" from MathWorld.

External links
• Earliest Uses of Some of the Words of Mathematics: entry on Injection, Surjection and Bijection has the history
of Injection and related terms. [2]

References
[1] http:/ / mathworld. wolfram. com/ Bijection. html
[2] http:/ / jeff560. tripod. com/ i. html
Binary function 23

Binary function
In mathematics, a binary function, or function of two variables, is a function which takes two inputs.
Precisely stated, a function is binary if there exists sets such that

where is the Cartesian product of and


For example, if Z is the set of integers, N+ is the set of natural numbers (except for zero), and Q is the set of rational
numbers, then division is a binary function from Z and N+ to Q.
Set-theoretically, one may represent a binary function as a subset of the Cartesian product X × Y × Z, where (x,y,z)
belongs to the subset if and only if f(x,y) = z. Conversely, a subset R defines a binary function if and only if, for any x
in X and y in Y, there exists a unique z in Z such that (x,y,z) belongs to R. We then define f (x,y) to be this z.
Alternatively, a binary function may be interpreted as simply a function from X × Y to Z. Even when thought of this
way, however, one generally writes f (x,y) instead of f((x,y)). (That is, the same pair of parentheses is used to indicate
both function application and the formation of an ordered pair.)
In turn, one can also derive ordinary functions of one variable from a binary function. Given any element x of X,
there is a function f x, or f (x,·), from Y to Z, given by f x(y) := f (x,y). Similarly, given any element y of Y, there is a
function f y, or f (·,y), from X to Z, given by f y(x) := f (x,y). (In computer science, this identification between a
function from X × Y to Z and a function from X to ZY is called Currying.)
The various concepts relating to functions can also be generalised to binary functions. For example, the division
example above is surjective (or onto) because every rational number may be expressed as a quotient of an integer
and a natural number. This example is injective in each input separately, because the functions f x and f y are always
injective. However, it's not injective in both variables simultaneously, because (for example) f (2,4) = f (1,2).
One can also consider partial binary functions, which may be defined only for certain values of the inputs. For
example, the division example above may also be interpreted as a partial binary function from Z and N to Q, where
N is the set of all natural numbers, including zero. But this function is undefined when the second input is zero.
A binary operation is a binary function where the sets X, Y, and Z are all equal; binary operations are often used to
define algebraic structures.
In linear algebra, a bilinear transformation is a binary function where the sets X, Y, and Z are all vector spaces and
the derived functions f x and fy are all linear transformations. A bilinear transformation, like any binary function, can
be interpreted as a function from X × Y to Z, but this function in general won't be linear. However, the bilinear
transformation can also be interpreted as a single linear transformation from the tensor product X Y to Z.
The concept of binary function generalises to ternary (or 3-ary) function, quaternary (or 4-ary) function, or more
generally to n-ary function for any natural number n. A 0-ary function to Z is simply given by an element of Z. One
can also define an A-ary function where A is any set; there is one input for each element of A.
In category theory, n-ary functions generalise to n-ary morphisms in a multicategory. The interpretation of an n-ary
morphism as an ordinary morphisms whose domain is some sort of product of the domains of the original n-ary
morphism will work in a monoidal category. The construction of the derived morphisms of one variable will work in
a closed monoidal category. The category of sets is closed monoidal, but so is the category of vector spaces, giving
the notion of bilinear transformation above.
Bochner measurable function 24

Bochner measurable function


In mathematics – specifically, in functional analysis – a Bochner-measurable function taking values in a Banach
space is a function that equals a.e. the limit of a sequence of measurable countably-valued functions, i.e.,

where the functions each have a countable range and for which the pre-image is measurable for
each x. The concept is named after Salomon Bochner.
Bochner-measurable functions are sometimes called strongly measurable, -measurable or just measurable (or
uniformly measurable in case that the Banach space is the space of continuous linear operators between Banach
spaces).

Properties
The relationship between measurability and weak measurability is given by the following result, known as Pettis'
theorem or Pettis measurability theorem.
Function f is almost surely separably valued (or essentially separably valued) if there exists a subset
N ⊆ X with μ(N) = 0 such that f(X \ N) ⊆ B is separable.
A function  : X → B defined on a measure space (X, Σ, μ) and taking values in a Banach space B is
(strongly) measurable (with respect to Σ and the Borel σ-algebra on B) if and only if it is both weakly
measurable and almost surely separably valued.
In the case that B is separable, since any subset of a separable Banach space is itself separable, one can take N above
to be empty, and it follows that the notions of weak and strong measurability agree when B is separable.

References
• Showalter, Ralph E. (1997). "Theorem III.1.1". Monotone operators in Banach space and nonlinear partial
differential equations. Mathematical Surveys and Monographs 49. Providence, RI: American Mathematical
Society. p. 103. MR1422252. ISBN 0-8218-0500-2..
Bounded function 25

Bounded function
In mathematics, a function f defined on some set X with real or
complex values is called bounded, if the set of its values is bounded.
In other words, there exists a real number M < ∞ such that

for all x in X.
Sometimes, if for all x in X, then the function is said to be
bounded above by A. On the other hand, if for all x in X,
then the function is said to be bounded below by B.
The concept should not be confused with that of a bounded operator.
An important special case is a bounded sequence, where X is taken to
A schematic illustration of a bounded function
be the set N of natural numbers. Thus a sequence f = ( a0, a1, a2, ... ) is
(red) and an unbounded one (blue). Intuitively,
bounded if there exists a real number M < ∞ such that the graph of a bounded function stays within a
|an| ≤ M horizontal band, while the graph of an unbounded
function does not.
for every natural number n. The set of all bounded sequences, equipped
with a vector space structure, forms a sequence space.
This definition can be extended to functions taking values in a metric space Y. Such a function f defined on some set
X is called bounded if for some a in Y there exists a real number M < ∞ such that

for all x in X.
If this is the case, there is also such an M for each other a.

Examples
• The function f:R → R defined by f (x)=sin x is bounded. The sine function is no longer bounded if it is defined
over the set of all complex numbers
• The function

defined for all real x which do not equal −1 or 1 is not bounded. As x gets closer to −1 or to 1, the values of this
function get larger and larger in magnitude. This function can be made bounded if one considers its domain to be, for
example, [2, ∞).
• The function

defined for all real x is bounded.


• Every continuous function f:[0,1] → R is bounded. This is really a special case of a more general fact: Every
continuous function from a compact space into a metric space is bounded.
• The function f which takes the value 0 for x rational number and 1 for x irrational number is bounded. Thus, a
function does not need to be "nice" in order to be bounded. The set of all bounded functions defined on [0,1] is
much bigger than the set of continuous functions on that interval.
Cauchy-continuous function 26

Cauchy-continuous function
In mathematics, a Cauchy-continuous, or Cauchy-regular, function is a special kind of continuous function
between metric spaces (or more general spaces). Cauchy-continuous functions have the useful property that they can
always be (uniquely) extended to the Cauchy completion of their domain.

Definition
Let X and Y be metric spaces, and let f be a function from X to Y. Then f is Cauchy-continuous if and only if, given
any Cauchy sequence (x1, x2, …) in X, the sequence (f(x1), f(x2), …) is a Cauchy sequence in Y.

Properties
Every uniformly continuous function is also Cauchy-continuous, and any Cauchy-continuous function is continuous.
Conversely, if X is a complete space, then every continuous function on X is Cauchy-continuous too. More generally,
even if X is not complete, as long as Y is complete, then any Cauchy-continuous function from X to Y can be
extended to a function defined on the Cauchy completion of X; this extension is necessarily unique.

Examples and non-examples


Since the real line ℝ is complete, the Cauchy-continuous functions on ℝ are the same as the continuous ones. On the
subspace ℚ of rational numbers, however, matters are different. For example, define a two-valued function so that
f(x) is 0 when x2 is less than 2 but 1 when x2 is greater than 2. (Note that x2 is never equal to 2 for any rational
number x.) This function is continuous on ℚ but not Cauchy-continuous, since it can't be extended to ℝ. On the other
hand, any uniformly continuous function on ℚ must be Cauchy-continuous. For a non-uniform example on ℚ, let f(x)
be 2x; this is not uniformly continuous (on all of ℚ), but it is Cauchy-continuous.
A Cauchy sequence (y1, y2, …) in Y can be identified with a Cauchy-continuous function from {1, 1/2, 1/3, …} to Y,
defined by f(1/n) = yn. If Y is complete, then this can be extended to {1, 1/2, 1/3, …, 0}; f(0) will be the limit of the
Cauchy sequence.

Generalisations
Cauchy continuity makes sense in situations more general than metric spaces, but then one must move from
sequences to nets (or equivalently filters). The definition above applies, as long as the Cauchy sequence (x1, x2, …)
is replaced with an arbitrary Cauchy net. Equivalently, a function f is Cauchy-continuous if and only if, given any
Cauchy filter F on X, then f(F) is a Cauchy filter on Y. This definition agrees with the above on metric spaces, but it
also works for uniform spaces and, most generally, for Cauchy spaces.
Any directed set A may be made into a Cauchy space. Then given any space Y, the Cauchy nets in Y indexed by A
are the same as the Cauchy-continuous functions from A to Y. If Y is complete, then the extension of the function to
A ∪ {∞} will give the value of the limit of the net. (This generalises the example of sequences above, where 0 is to
be interpreted as 1/∞.)
Cauchy-continuous function 27

References
• Eva Lowen-Colebunders (1989). Function Classes of Cauchy Continuous Maps. Dekker, New York.

Closed convex function


In mathematics, a convex function is called closed if its epigraph is a closed set.

Properties
A closed convex function f is the pointwise supremum of the collection of all affine functions h such that h≤f.

References
• Rockafellar, R. Tyrell, Convex Analysis, Princeton University Press (1996). ISBN 0-691-01586-4

Coarse function
In mathematics, coarse functions are functions that may appear to be continuous at a distance, but in reality are not
necessarily continuous.[1] Although continuous functions are usually observed on a small scale, coarse functions are
usually observed on a large scale.[1] [2]

See also
• Continuous function

References
[1] Chul-Woo Lee and Jared Duke (2007), Coarse Function Value Theorems (http:/ / www. rose-hulman. edu/ mathjournal/ archives/ 2007/
vol8-n2/ paper4/ v8n2-4pd. pdf). Rose-Hulman Undergraduate Mathematics Journal 8 (2)
[2] Dongarra, Jack; Madsen, Kaj; Wasniewski, Jerzy, eds (2006). Applied Parallel Computing: State of the Art in Scientific Computing (http:/ /
books. google. com/ books?id=ZqFkI1MufzMC). Germany: Springer-Verlag Berlin Heidelberg. pp. 316–322. ISBN 978-3-540-29067-4. .
Completely multiplicative function 28

Completely multiplicative function


In number theory, functions of positive integers which respect products are important and are called completely
multiplicative functions or totally multiplicative functions. Especially in number theory, a weaker condition is
also important, respecting only products of coprime numbers, and such functions are called multiplicative functions.
Outside of number theory, the term "multiplicative function" is often taken to be synonymous with "completely
multiplicative function" as defined in this article.

Definition
A completely multiplicative function (or totally multiplicative function) is an arithmetic function (that is, a
function whose domain is the natural numbers), such that f(1) = 1 and f(ab) = f(a) f(b) holds for all positive integers a
and b.
Without the requirement that f(1) = 1, one could still have f(1) = 0, but then f(a) = 0 for all positive integers a, so this
is not a very strong restriction.

Examples
The easiest example of a multiplicative function is a monomial: For any particular positive integer n, define f(a) = an.

Properties
A completely multiplicative function is completely determined by its values at the prime numbers, a consequence of
the fundamental theorem of arithmetic. Thus, if n is a product of powers of distinct primes, say n = pa qb ..., then f(n)
= f(p)a f(q)b ...
Concave function 29

Concave function
In mathematics, a concave function is the negative of a convex function. A concave function is also synonymously
called concave downwards, concave down, convex cap or upper convex.

Definition
A real-valued function f defined on an interval (or on any convex set C of some vector space) is called concave if,
for any two points x and y in its domain C and any t in [0,1], we have

A function is called strictly concave if

for any t in (0,1) and x ≠ y.


For a function f:R→R, this definition merely states that for every z between x and y, the point (z, f(z) ) on the graph of
f is above the straight line joining the points (x, f(x) ) and (y, f(y) ).

A function f(x) is a quasiconcave if the upper contour sets of the function are convex
sets.[1]

Properties
A function f(x) is concave over a convex set if and only if the function −f(x) is a convex function over the set.
A differentiable function f is concave on an interval if its derivative function f ′ is monotonically decreasing on that
interval: a concave function has a decreasing slope. ("Decreasing" here means "non-increasing", rather than "strictly
decreasing", and thus allows zero slopes.)
For a twice-differentiable function f, if the second derivative, f ′′(x), is positive (or, if the acceleration is positive),
then the graph is convex; if f ′′(x) is negative, then the graph is concave. Points where concavity changes are
inflection points.
If a convex (i.e., concave upward) function has a "bottom", any point at the bottom is a minimal extremum. If a
concave (i.e., concave downward) function has an "apex", any point at the apex is a maximal extremum.
Concave function 30

If f(x) is twice-differentiable, then f(x) is concave if and only if f ′′(x) is non-positive. If its second derivative is
negative then it is strictly concave, but the opposite is not true, as shown by f(x) = -x4.
If f is concave and differentiable then
[2]

A continuous function on C is concave if and only if for any x and y in C

If a function f is concave, and f(0) ≥ 0, then f is subadditive. Proof:


• since f is concave, let y = 0,

Examples
• The functions and are concave, as the second derivative is always negative.
• Any linear function is both concave and convex.
• The function is concave on the interval .
• The function , where is the determinant of matrix nonnegative-definite matrix B, is concave[3] .
• Practical application: rays bending in Computation of radiowave attenuation in the atmosphere.

References
[1] Varian, Hal A. (1992) Microeconomic Analysis. Third Edition. W.W. Norton and Company. p. 496
[2] Varian, Hal A. (1992) Microeconomic Analysis. Third Edition. W.W. Norton and Company. p. 489
[3] Thomas M. Cover and J. A. Thomas (1988). "Determinant inequalities via information theory". SIAM journal on matrix analysis and
applications 9 (3): 384–392.

• Rao, Singiresu S. (2009). Engineering Optimization: Theory and Practice. John Wiley and Sons. p. 779.
ISBN 0470183527.
Constant function 31

Constant function
In mathematics, a constant function is a function whose values do not vary and thus are constant. For example, if
we have the function f(x) = 4, then f is constant since f maps any value to 4. More formally, a function f : A → B is a
constant function if f(x) = f(y) for all x and y in A.
Every empty function is constant, vacuously, since there are no x and y in A for which f(x) and f(y) are different when
A is the empty set. Some find it more convenient, however, to define constant function so as to exclude empty
functions.
In the context of polynomial functions, a non-zero constant function is called a polynomial of degree zero.

Properties
Constant functions can be characterized with respect to function composition in two ways.
The following are equivalent:
1. f : A → B is a constant function.
2. For all functions g, h : C → A, f o g = f o h, (where "o" denotes function composition).
3. The composition of f with any other function is also a constant function.
The first characterization of constant functions given above, is taken as the motivating and defining property for the
more general notion of constant morphism in category theory.
In contexts where it is defined, the derivative of a function measures how that function varies with respect to the
variation of some argument. It follows that, since a constant function does not vary, its derivative, where defined,
will be zero. Thus for example:
• If f is a real-valued function of a real variable, defined on some interval, then f is constant if and only if the
derivative of f is everywhere zero.
For functions between preordered sets, constant functions are both order-preserving and order-reversing; conversely,
if f is both order-preserving and order-reversing, and if the domain of f is a lattice, then f must be constant.
Other properties of constant functions include:
• Every constant function whose domain and codomain are the same is idempotent.
• Every constant function between topological spaces is continuous.
A function on a connected set is locally constant if and only if it is constant.

References
• Herrlich, Horst and Strecker, George E., Category Theory, Allen and Bacon, Inc. Boston (1973)
• Constant function [1] on PlanetMath

References
[1] http:/ / planetmath. org/ ?op=getobj& amp;from=objects& amp;id=4727
Continuous function 32

Continuous function
In mathematics, a continuous function is a function for which, intuitively, small changes in the input result in small
changes in the output. Otherwise, a function is said to be "discontinuous". A continuous function with a continuous
inverse function is called "bicontinuous".
Continuity of functions is one of the core concepts of topology, which is treated in full generality below. The
introductory portion of this article focuses on the special case where the inputs and outputs of functions are real
numbers. In addition, this article discusses the definition for the more general case of functions between two metric
spaces. In order theory, especially in domain theory, one considers a notion of continuity known as Scott continuity.
Other forms of continuity do exist but they are not discussed in this article.
As an example, consider the function h(t) which describes the height of a growing flower at time t. This function is
continuous. In fact, there is a dictum of classical physics which states that in nature everything is continuous. By
contrast, if M(t) denotes the amount of money in a bank account at time t, then the function jumps whenever money
is deposited or withdrawn, so the function M(t) is discontinuous. (However, if one assumes a discrete set as the
domain of function M, for instance the set of points of time at 4:00 PM on business days, then M becomes
continuous function, as every function whose domain is a discrete subset of reals is.)

Real-valued continuous functions

Historical infinitesimal definition


Cauchy defined continuity of a function in the following intuitive terms: an infinitesimal change in the independent
variable corresponds to an infinitesimal change of the dependent variable (see Cours d'analyse, page 34).

Definition in terms of limits


Suppose we have a function that maps real numbers to real numbers and whose domain is some interval, like the
functions h and M above. Such a function can be represented by a graph in the Cartesian plane; the function is
continuous if, roughly speaking, the graph is a single unbroken curve with no "holes" or "jumps".
In general, we say that the function f is continuous at some point c of its domain if, and only if, the following holds:
• The limit of f(x) as x approaches c through domain of f does exist and is equal to f(c); in mathematical notation,
. If the point c in the domain of f is not a limit point of the domain, then this condition is
vacuously true, since x cannot approach c through values not equal c. Thus, for example, every function whose
domain is the set of all integers is continuous.
We call a function continuous if and only if it is continuous at every point of its domain. More generally, we say
that a function is continuous on some subset of its domain if it is continuous at every point of that subset.
The notation C(Ω) or C0(Ω) is sometimes used to denote the set of all continuous functions with domain Ω.
Similarly, C1(Ω) is used to denote the set of differentiable functions whose derivative is continuous, C²(Ω) for the
twice-differentiable functions whose second derivative is continuous, and so on (see differentiability class). In the
field of computer graphics, these three levels are sometimes called g0 (continuity of position), g1 (continuity of
tangency), and g2 (continuity of curvature). The notation C(n, α)(Ω) occurs in the definition of a more subtle concept,
that of Hölder continuity.
Continuous function 33

Weierstrass definition (epsilon-delta) of continuous functions


Without resorting to limits, one can define continuity of real functions as follows.
Again consider a function ƒ that maps a set of real numbers to another set of real numbers, and suppose c is an
element of the domain of ƒ. The function ƒ is said to be continuous at the point c if the following holds: For any
number ε > 0, however small, there exists some number δ > 0 such that for all x in the domain of ƒ with
c − δ < x < c + δ, the value of ƒ(x) satisfies

Alternatively written: Given subsets I, D of R, continuity of ƒ : I → D at c ∈ I means that for every ε > 0 there exists
a δ > 0 such that for all x ∈ I,:

A form of this epsilon-delta definition of continuity was first given by Bernard Bolzano in 1817. Preliminary forms
of a related definition of the limit were given by Cauchy,[1] though the formal definition and the distinction between
pointwise continuity and uniform continuity were first given by Karl Weierstrass.
More intuitively, we can say that if we want to get all the ƒ(x) values to stay in some small neighborhood around ƒ(c),
we simply need to choose a small enough neighborhood for the x values around c, and we can do that no matter how
small the ƒ(x) neighborhood is; ƒ is then continuous at c.
In modern terms, this is generalized by the definition of continuity of a function with respect to a basis for the
topology, here the metric topology.

Definition using oscillation

The failure of a function to be continuous at a


point is quantified by its oscillation.

Continuity can also be defined in terms of oscillation: a function ƒ is continuous at a point x0 if and only if the
oscillation is zero;[2] in symbols, A benefit of this definition is that it quantifies discontinuity: the
oscillation gives how much the function is discontinuous at a point.
This definition is useful in descriptive set theory to study the set of discontinuities and continuous points – the
continuous points are the intersection of the sets where the oscillation is less than ε (hence a Gδ set) – and gives a
very quick proof of one direction of the Lebesgue integrability condition.[3]
The oscillation is equivalent to the ε-δ definition by a simple re-arrangement, and by using a limit (lim sup, lim inf)
to define oscillation: if (at a given point) for a given ε0 there is no δ that satisfies the ε-δ definition, then the
oscillation is at least ε0, and conversely if for every ε there is a desired δ, the oscillation is 0. The oscillation
definition can be naturally generalized to maps from a topological space to a metric space.
Continuous function 34

Definition using the hyperreals


Non-standard analysis is a way of making Newton-Leibniz-style infinitesimals mathematically rigorous. The real
line is augmented by the addition of infinite and infinitesimal numbers to form the hyperreal numbers. In
nonstandard analysis, continuity can be defined as follows.
A function ƒ from the reals to the reals is continuous if its natural extension to the hyperreals has the property
that for real x and infinitesimal dx, ƒ(x+dx) − ƒ(x) is infinitesimal.[4]
In other words, an infinitesimal increment of the independent variable corresponds to an infinitesimal change of the
dependent variable, giving a modern expression to Augustin-Louis Cauchy's definition of continuity.

Examples
• All polynomial functions are continuous.
• If a function has a domain which is not an interval, the notion of a continuous function as one whose graph you
can draw without taking your pencil off the paper is not quite correct. Consider the functions f(x) = 1/x and g(x) =
(sin x)/x. Neither function is defined at x = 0, so each has domain R \ {0} of real numbers except 0, and each
function is continuous. The question of continuity at x = 0 does not arise, since x = 0 is neither in the domain of f
nor in the domain of g. The function f cannot be extended to a continuous function whose domain is R, since no
matter what value is assigned at 0, the resulting function will not be continuous. On the other hand, since the limit
of g at 0 is 1, g can be extended continuously to R by defining its value at 0 to be 1.
• The exponential functions, logarithms, square root function, trigonometric functions and absolute value function
are continuous. Rational functions, however, are not necessarily continuous on all of R.
• An example of a rational continuous function is f(x)=1⁄x-2. The question of continuity at x= 2 does not arise, since
x = 2 is not in the domain of f.
• An example of a discontinuous function is the function f defined by f(x) = 1 if x > 0, f(x) = 0 if x ≤ 0. Pick for
instance ε = 1⁄2. There is no δ-neighborhood around x = 0 that will force all the f(x) values to be within ε of f(0).
Intuitively we can think of this type of discontinuity as a sudden jump in function values.
• Another example of a discontinuous function is the signum or sign function.
• A more complicated example of a discontinuous function is Thomae's function.
• Dirichlet's function

is nowhere continuous.

Facts about continuous functions


If two functions f and g are continuous, then f + g, fg, and f/g are continuous. (Note. The only possible points x of
discontinuity of f/g are the solutions of the equation g(x) = 0; but then any such x does not belong to the domain of
the function f/g. Hence f/g is continuous on its entire domain, or - in other words - is continuous.)
The composition f o g of two continuous functions is continuous.
If a function is differentiable at some point c of its domain, then it is also continuous at c. The converse is not true: a
function that is continuous at c need not be differentiable there. Consider for instance the absolute value function at
c = 0.
Continuous function 35

Intermediate value theorem


The intermediate value theorem is an existence theorem, based on the real number property of completeness, and
states:
If the real-valued function f is continuous on the closed interval [a, b] and k is some number between f(a) and
f(b), then there is some number c in [a, b] such that f(c) = k.
For example, if a child grows from 1 m to 1.5 m between the ages of two and six years, then, at some time between
two and six years of age, the child's height must have been 1.25 m.
As a consequence, if f is continuous on [a, b] and f(a) and f(b) differ in sign, then, at some point c in [a, b], f(c) must
equal zero.

Extreme value theorem


The extreme value theorem states that if a function f is defined on a closed interval [a,b] (or any closed and bounded
set) and is continuous there, then the function attains its maximum, i.e. there exists c ∈ [a,b] with f(c) ≥ f(x) for all
x ∈ [a,b]. The same is true of the minimum of f. These statements are not, in general, true if the function is defined
on an open interval (a,b) (or any set that is not both closed and bounded), as, for example, the continuous function
f(x) = 1/x, defined on the open interval (0,1), does not attain a maximum, being unbounded above.

Directional continuity

A right continuous function A left continuous function

A function may happen to be continuous in only one direction, either from the "left" or from the "right". A
right-continuous function is a function which is continuous at all points when approached from the right.
Technically, the formal definition is similar to the definition above for a continuous function but modified as
follows:
The function ƒ is said to be right-continuous at the point c if the following holds: For any number ε > 0 however
small, there exists some number δ > 0 such that for all x in the domain with c < x < c + δ, the value of ƒ(x) will satisfy

Notice that x must be larger than c, that is on the right of c. If x were also allowed to take values less than c, this
would be the definition of continuity. This restriction makes it possible for the function to have a discontinuity at c,
but still be right continuous at c, as pictured.
Likewise a left-continuous function is a function which is continuous at all points when approached from the left,
that is, c − δ < x < c.
Continuous function 36

A function is continuous if and only if it is both right-continuous and left-continuous.

Continuous functions between metric spaces


Now consider a function f from one metric space (X, dX) to another metric space (Y, dY). Then f is continuous at the
point c in X if for any positive real number ε, there exists a positive real number δ such that all x in X satisfying dX(x,
c) < δ will also satisfy dY(f(x), f(c)) < ε.
This can also be formulated in terms of sequences and limits: the function f is continuous at the point c if for every
sequence (xn) in X with limit lim xn = c, we have lim f(xn) = f(c). Continuous functions transform limits into limits.
This latter condition can be weakened as follows: f is continuous at the point c if and only if for every convergent
sequence (xn) in X with limit c, the sequence (f(xn)) is a Cauchy sequence, and c is in the domain of f. Continuous
functions transform convergent sequences into Cauchy sequences.
The set of points at which a function between metric spaces is continuous is a Gδ set – this follows from the ε-δ
definition of continuity.

Continuous functions between topological spaces


The above definitions of continuous
functions can be generalized to functions
from one topological space to another in a
natural way; a function f : X → Y, where X
and Y are topological spaces, is continuous
if and only if for every open set V ⊆ Y, the
inverse image

Continuity of a function at a point

is open.
However, this definition is often difficult to use directly. Instead, suppose we have a function f from X to Y, where X,
Y are topological spaces. We say f is continuous at x for some x ∈ X if for any neighborhood V of f(x), there is a
neighborhood U of x such that f(U) ⊆ V. Although this definition appears complex, the intuition is that no matter how
"small" V becomes, we can always find a U containing x that will map inside it. If f is continuous at every x ∈ X, then
we simply say f is continuous.
In a metric space, it is equivalent to consider the neighbourhood system of open balls centered at x and f(x) instead of
all neighborhoods. This leads to the standard ε-δ definition of a continuous function from real analysis, which says
roughly that a function is continuous if all points close to x map to points close to f(x). This only really makes sense
in a metric space, however, which has a notion of distance.
Note, however, that if the target space is Hausdorff, it is still true that f is continuous at a if and only if the limit of f
as x approaches a is f(a). At an isolated point, every function is continuous.
Continuous function 37

Definitions
Several equivalent definitions for a topological structure exist and thus there are several equivalent ways to define a
continuous function.

Open and closed set definition


The most common notion of continuity in topology defines continuous functions as those functions for which the
preimages (or inverse images) of open sets are open. Similar to the open set formulation is the closed set
formulation, which says that preimages (or inverse images) of closed sets are closed.

Neighborhood definition
Definitions based on preimages are often difficult to use directly. Instead, suppose we have a function f : X → Y,
where X and Y are topological spaces.[5] We say f is continuous at x for some x ∈ X if for any neighborhood V of
f(x), there is a neighborhood U of x such that f(U) ⊆ V. Although this definition appears complicated, the intuition is
that no matter how "small" V becomes, we can always find a U containing x that will map inside it. If f is continuous
at every x ∈ X, then we simply say f is continuous.

In a metric space, it is equivalent to consider the neighbourhood system of open balls centered at x and f(x) instead of
all neighborhoods. This leads to the standard δ-ε definition of a continuous function from real analysis, which says
roughly that a function is continuous if all points close to x map to points close to f(x). This only really makes sense
in a metric space, however, which has a notion of distance.
Note, however, that if the target space is Hausdorff, it is still true that f is continuous at a if and only if the limit of f
as x approaches a is f(a). At an isolated point, every function is continuous.

Sequences and nets


In several contexts, the topology of a space is conveniently specified in terms of limit points. In many instances, this
is accomplished by specifying when a point is the limit of a sequence, but for some spaces that are too large in some
sense, one specifies also when a point is the limit of more general sets of points indexed by a directed set, known as
nets. A function is continuous only if it takes limits of sequences to limits of sequences. In the former case,
preservation of limits is also sufficient; in the latter, a function may preserve all limits of sequences yet still fail to be
continuous, and preservation of nets is a necessary and sufficient condition.
In detail, a function f : X → Y is sequentially continuous if whenever a sequence (xn) in X converges to a limit x, the
sequence (f(xn)) converges to f(x). Thus sequentially continuous functions "preserve sequential limits". Every
continuous function is sequentially continuous. If X is a first-countable space, then the converse also holds: any
function preserving sequential limits is continuous. In particular, if X is a metric space, sequential continuity and
continuity are equivalent. For non first-countable spaces, sequential continuity might be strictly weaker than
continuity. (The spaces for which the two properties are equivalent are called sequential spaces.) This motivates the
Continuous function 38

consideration of nets instead of sequences in general topological spaces. Continuous functions preserve limits of
nets, and in fact this property characterizes continuous functions.

Closure operator definition


Given two topological spaces (X,cl) and (X ' ,cl ') where cl and cl ' are two closure operators then a function

is continuous if for all subsets A of X

One might therefore suspect that given two topological spaces (X,int) and (X ' ,int ') where int and int ' are two
interior operators then a function

is continuous if for all subsets A of X

or perhaps if

however, neither of these conditions is either necessary or sufficient for continuity.


Instead, we must resort to inverse images: given two topological spaces (X,int) and (X ' ,int ') where int and int ' are
two interior operators then a function

is continuous if for all subsets A of X '

We can also write that given two topological spaces (X,cl) and (X ' ,cl ') where cl and cl ' are two closure operators
then a function

is continuous if for all subsets A of X '

Closeness relation definition


Given two topological spaces (X,δ) and (X' ,δ') where δ and δ' are two closeness relations then a function

is continuous if for all points x and of X and all subsets A of X,

This is another way of writing the closure operator definition.


Continuous function 39

Useful properties of continuous maps


Some facts about continuous maps between topological spaces:
• If f : X → Y and g : Y → Z are continuous, then so is the composition g ∘ f : X → Z.
• If f : X → Y is continuous and
• X is compact, then f(X) is compact.
• X is connected, then f(X) is connected.
• X is path-connected, then f(X) is path-connected.
• X is Lindelöf, then f(X) is Lindelöf.
• X is separable, then f(X) is separable.
• The identity map idX : (X, τ2) → (X, τ1) is continuous if and only if τ1 ⊆ τ2 (see also comparison of topologies).

Other notes
If a set is given the discrete topology, all functions with that space as a domain are continuous. If the domain set is
given the indiscrete topology and the range set is at least T0, then the only continuous functions are the constant
functions. Conversely, any function whose range is indiscrete is continuous.
Given a set X, a partial ordering can be defined on the possible topologies on X. A continuous function between two
topological spaces stays continuous if we strengthen the topology of the domain space or weaken the topology of the
codomain space. Thus we can consider the continuity of a given function a topological property, depending only on
the topologies of its domain and codomain spaces.
For a function f from a topological space X to a set S, one defines the final topology on S by letting the open sets of S
be those subsets A of S for which f−1(A) is open in X. If S has an existing topology, f is continuous with respect to
this topology if and only if the existing topology is coarser than the final topology on S. Thus the final topology can
be characterized as the finest topology on S which makes f continuous. If f is surjective, this topology is canonically
identified with the quotient topology under the equivalence relation defined by f. This construction can be
generalized to an arbitrary family of functions X → S.
Dually, for a function f from a set S to a topological space, one defines the initial topology on S by letting the open
sets of S be those subsets A of S for which f(A) is open in X. If S has an existing topology, f is continuous with
respect to this topology if and only if the existing topology is finer than the initial topology on S. Thus the initial
topology can be characterized as the coarsest topology on S which makes f continuous. If f is injective, this topology
is canonically identified with the subspace topology of S, viewed as a subset of X. This construction can be
generalized to an arbitrary family of functions S → X.
Symmetric to the concept of a continuous map is an open map, for which images of open sets are open. In fact, if an
open map f has an inverse, that inverse is continuous, and if a continuous map g has an inverse, that inverse is open.
If a function is a bijection, then it has an inverse function. The inverse of a continuous bijection is open, but need not
be continuous. If it is, this special function is called a homeomorphism. If a continuous bijection has as its domain a
compact space and its codomain is Hausdorff, then it is automatically a homeomorphism.
Continuous function 40

Continuous functions between partially ordered sets


In order theory, continuity of a function between posets is Scott continuity. Let X be a complete lattice, then a
function f : X → X is continuous if, for each subset Y of X, we have sup f(Y) = f(sup Y).

Continuous binary relation


A binary relation R on A is continuous if R(a, b) whenever there are sequences (ak)i and (bk)i in A which converge to
a and b respectively for which R(ak, bk) for all k. Clearly, if one treats R as a characteristic function in two variables,
this definition of continuous is identical to that for continuous functions.

Continuity space
A continuity space[6] [7] is a generalization of metric spaces and posets, which uses the concept of quantales, and
that can be used to unify the notions of metric spaces and domains.[8]

See also
• Absolute continuity
• Bounded linear operator
• Classification of discontinuities
• Coarse function
• Continuous functor
• Continuous stochastic process
• Dini continuity
• Discrete function
• Equicontinuity
• Lipschitz continuity
• Normal function
• Piecewise
• Scott continuity
• Semicontinuity
• Smooth function
• Symmetrically continuous function
• Uniform continuity

Notes
[1] Grabiner, Judith V. (March 1983). "Who Gave You the Epsilon? Cauchy and the Origins of Rigorous Calculus" (http:/ / www. maa. org/
pubs/ Calc_articles/ ma002. pdf). The American Mathematical Monthly 90 (3): 185–194. doi:10.2307/2975545. .
[2] Introduction to Real Analysis (http:/ / ramanujan. math. trinity. edu/ wtrench/ texts/ TRENCH_REAL_ANALYSIS. PDF), updated April
2010, William F. Trench, Theorem 3.5.2, p. 172
[3] Introduction to Real Analysis (http:/ / ramanujan. math. trinity. edu/ wtrench/ texts/ TRENCH_REAL_ANALYSIS. PDF), updated April
2010, William F. Trench, 3.5 "A More Advanced Look at the Existence of the Proper Riemann Integral", pp. 171–177
[4] http:/ / www. math. wisc. edu/ ~keisler/ calc. html
[5] f is a function f : X → Y between two topological spaces (X,TX) and (Y,TY). That is, the function f is defined on the elements of the set X, not
on the elements of the topology TX. However continuity of the function does depend on the topologies used.
[6] Quantales and continuity spaces (http:/ / citeseerx. ist. psu. edu/ viewdoc/ download?doi=10. 1. 1. 48. 851& rep=rep1& type=pdf), RC Flagg -
Algebra Universalis, 1997
[7] All topologies come from generalized metrics, R Kopperman - American Mathematical Monthly, 1988
[8] Continuity spaces: Reconciling domains and metric spaces, B Flagg, R Kopperman - Theoretical Computer Science, 1997
Continuous function 41

References
• Visual Calculus (http://archives.math.utk.edu/visual.calculus/) by Lawrence S. Husch, University of
Tennessee (2001)

Convex function
In mathematics, a real-valued function
defined on an interval (or on any
convex subset of some vector space) is
called convex, concave upwards, concave
up or convex cup, if for any two points
and in its domain X and any ,

Convex function on an interval.

A function (in black) is convex if and only if the region above its graph (in green)
is a convex set.

A function is called strictly convex if

for every , , and .


Note that the function must be defined over a convex set, otherwise the point may not lie in the
function domain.
A function ƒ is said to be (strictly) concave if −ƒ is (strictly) convex.
Pictorially, a function is called 'convex' if the function lies below or on the straight line segment connecting two
points, for any two points in the interval.
Sometimes an alternative definition is used:
Convex function 42

A function is convex if its epigraph (the set of points lying on or above the graph) is a convex set.
These two definitions are equivalent, i.e., one holds if and only if the other one is true.

Properties
Suppose ƒ is a function of one real variable defined on an interval, and let

(note that R(x,y) is the slope of the red line in the above drawing; note also that the function R is symmetric in x,y). ƒ
is convex if and only if R(x,y) is monotonically non-decreasing in x, for y fixed (or viceversa). This characterization
of convexity is quite useful to prove the following results.
A convex function ƒ defined on some open interval C is continuous on C and Lipschitz continuous on any closed
subinterval. ƒ admits left and right derivatives, and these are monotonically non-decreasing. As a consequence, ƒ is
differentiable at all but at most countably many points. If C is closed, then ƒ may fail to be continuous at the
endpoints of C (an example is shown in the examples' section).
A function is midpoint convex on an interval C if

for all x and y in C. This condition is only slightly weaker than convexity. For example, a real valued Lebesgue
measurable function that is midpoint convex will be convex.[1] In particular, a continuous function that is midpoint
convex will be convex.
A differentiable function of one variable is convex on an interval if and only if its derivative is monotonically
non-decreasing on that interval. If a function is differentiable and convex then it is also continuously differentiable.
A continuously differentiable function of one variable is convex on an interval if and only if the function lies above
all of its tangents:
[2]

for all x and y in the interval. In particular, if ƒ '(c) = 0, then c is a global minimum of ƒ(x).
A twice differentiable function of one variable is convex on an interval if and only if its second derivative is
non-negative there; this gives a practical test for convexity. If its second derivative is positive then it is strictly
convex, but the converse does not hold. For example, the second derivative of ƒ(x) = x4 is ƒ "(x) = 12 x2, which is zero
for x = 0, but x4 is strictly convex.
More generally, a continuous, twice differentiable function of several variables is convex on a convex set if and only
if its Hessian matrix is positive semidefinite on the interior of the convex set.
Any local minimum of a convex function is also a global minimum. A strictly convex function will have at most one
global minimum.
For a convex function ƒ, the sublevel sets {x | ƒ(x) < a} and {x | ƒ(x) ≤ a} with a ∈ R are convex sets. However, a
function whose sublevel sets are convex sets may fail to be a convex function. A function whose sublevel sets are
convex is called a quasiconvex function.
Jensen's inequality applies to every convex function ƒ. If X is a random variable taking values in the domain of ƒ, then
(Here denotes the mathematical expectation.)
If a function f is convex, and f(0) ≤ 0, then f is superadditive. Proof:
• since f is convex, let y = 0,


Convex function 43

Convex function calculus


• If and are convex functions, then so are and
• If and are convex functions and if is non-decreasing, then is convex.
• Convexity is invariant under affine maps: that is, if is convex with , then so is
with , where
• If is convex in then is convex in provided for some
• If is convex, then its perspective function (whose domain is
) is convex.

Strongly convex functions


The concept of strong convexity extends the notion of strict convexity. A strongly convex function is also strictly
convex, but not vice-versa. A differentiable function f is called strongly convex with parameter m > 0 if the
following equation holds for all points x,y in its domain:

This is equivalent to the following

It is not necessary for a function to be differentiable in order to be strongly convex. A third definition for a strongly
convex function, with parameter m, is that, for all x,y in the domain and ,

(given that for .)


If the function f is twice continuously differentiable, then f is strongly convex with parameter m if and only if
for all x in the domain, where I is the identity and is the Hessian matrix, and the inequality
means that is positive definite. This is equivalent to requiring that the minimum eigenvalue of
be at least m for all x. If the domain is just the real line, then is just the second derivative
, so the condition becomes . If m = 0, then this means the Hessian is positive semidefinite (or
if the domain is the real line, it means that ), which implies the function is convex, and perhaps strictly
convex, but not strongly convex.
Assuming still that the function is twice continuously differentiable, we show that the lower bound of
implies that it is strongly convex. Start by using Taylor's Theorem:

for some (unknown) . Then by the assumption


about the eigenvalues, and hence we recover the second strong convexity equation above.
The distinction between convex, strictly convex, and strongly convex can be subtle at first glimpse. If is twice
continuously differentiable and the domain is the real line, then we can characterize it as follows:
convex if and only if for all
strictly convex if for all (note: this is sufficient, but not necessary)
strongly convex if and only if for all
For example, consider a function that is strictly convex, and suppose there is a sequence of points such that

. Even though , the function is not strongly convex because will become

arbitrarily small.
Convex function 44

Strongly convex functions are in general easier to work with than convex or strictly convex functions, since they are
a smaller class. Like strictly convex functions, strongly convex functions have unique minima.

Examples
• The function has at all points, so ƒ is a convex function. It is also strongly convex
(and hence strictly convex too), with strong convexity constant 2.
• The function has , so ƒ is a convex function. It is strictly convex, even
though the second derivative is not strictly positive at all points. It is not strongly convex.
• The absolute value function is convex, even though it does not have a derivative at the point x = 0.
It is not strictly convex.
• The function for 1 ≤ p is convex.
• The exponential function is convex. It is also strictly convex, since , but it is not
strongly convex since the second derivative can be arbitrarily close to zero. More generally, the function
is logarithmically convex if ƒ is a convex function.
• The function ƒ with domain [0,1] defined by ƒ(0) = ƒ(1) = 1, ƒ(x) = 0 for 0 < x < 1 is convex; it is continuous on the
open interval (0, 1), but not continuous at 0 and 1.
• The function x3 has second derivative 6x; thus it is convex on the set where x ≥ 0 and concave on the set
where x ≤ 0.
• Every linear transformation taking values in is convex but not strictly convex, since if f is linear, then
This statement also holds if we replace "convex" by "concave".
• Every affine function taking values in , i.e., each function of the form , is simultaneously
convex and concave.
• Every norm is a convex function, by the triangle inequality and positive homogeneity.
• If ƒ is convex, the perspective function is convex for t > 0.
• Examples of functions that are monotonically increasing but not convex include and g(x) = log(x).
• Examples of functions that are convex but not monotonically increasing include and .
• The function ƒ(x) = 1/x2, with f(0) = +∞, is convex on the interval (0, +∞) and convex on the interval (-∞,0), but
not convex on the interval (-∞, +∞), because of the singularity at x = 0.

References
[1] Sierpinski Theorem, Donoghue (1969), p. 12 (http:/ / books. google. com/ books?id=P30Y7daiGvQC& pg=PA12)
[2] Varian, Hal A. (1992) Microeconomic Analysis Third Edition. W.W. Norton and Company. p. 490

• Bertsekas, Dimitri (2003). Convex Analysis and Optimization. Athena Scientific.


• Borwein, Jonathan, and Lewis, Adrian. (2000). Convex Analysis and Nonlinear Optimization. Springer.
• Donoghue, William F. (1969). Distributions and Fourier Transforms. Academic Press.
• Hiriart-Urruty, Jean-Baptiste, and Lemaréchal, Claude. (2004). Fundamentals of Convex analysis. Berlin:
Springer.
• Krasnosel'skii M.A., Rutickii Ya.B. (1961). Convex Functions and Orlicz Spaces. Groningen: P.Noordhoff Ltd.
• Luenberger, David (1984). Linear and Nonlinear Programming. Addison-Wesley.
• Luenberger, David (1969). Optimization by Vector Space Methods. Wiley & Sons.
• Rockafellar, R. T. (1970). Convex analysis. Princeton: Princeton University Press.
• Thomson, Brian (1994). Symmetric Properties of Real Functions. CRC Press.
• Zălinescu, C.. Convex analysis in general vector spaces. World Scientific Publishing  Co., Inc. pp. xx+367.
MR1921556. ISBN 981-238-067-1.
Convex function 45

External links
• Stephen Boyd and Lieven Vandenberghe, Convex Optimization (http://www.stanford.edu/~boyd/cvxbook/)
(PDF)

Differentiable function
In calculus (a branch of mathematics), a differentiable function is a
function whose derivative exists at each point in its domain. The graph
of a differentiable function must have a non-vertical tangent line at
each point in its domain. As a result, the graph of a differentiable
function must be relatively smooth, and cannot contain any breaks,
bends, or cusps, or any points with a vertical tangent.

More generally, if x0 is a point in the domain of a function ƒ, then ƒ is


said to be differentiable at x0 if the derivative ƒ′(x0) is defined. This
means that the graph of ƒ has a non-vertical tangent line at the point
(x0, ƒ(x0)), and therefore cannot have a break, bend, or cusp at this
point.
A differentiable function

The absolute value function is not differentiable


at x = 0.

Differentiability and continuity


If ƒ is differentiable at a point x0, then ƒ must also be continuous at x0.
In particular, any differentiable function must be continuous at every
point in its domain. The converse does not hold: a continuous function
need not be differentiable. For example, a function with a bend, cusp,
or vertical tangent may be continuous, but fails to be differentiable at
the location of the anomaly.

Most functions which occur in practice have derivatives at all points or


at almost every point. However, a result of Stefan Banach states that The Weierstrass function is continuous, but is not
the set of functions which have a derivative at some point is a meager differentiable at any point.
[1]
set in the space of all continuous functions. Informally, this means
that differentiable functions are very atypical among continuous functions. The first known example of a function
that is continuous everywhere but differentiable nowhere is the Weierstrass function.
Differentiable function 46

Differentiability classes
A function ƒ is said to be continuously differentiable if the derivative ƒ′(x) exists, and is itself a continuous function.
Though the derivative of a differentiable function never has a jump discontinuity, it is possible for the derivative to
have an essential discontinuity. For example, the function

is differentiable at 0 (with the derivative being 0), but the derivative is not continuous at this point.
Sometimes continuously differentiable functions are said to be of class C1. A function is of class C2 if the first and
second derivative of the function both exist and are continuous. More generally, a function is said to be of class Ck if
the first k derivatives ƒ′(x), ƒ″(x), ..., ƒ(k)(x) all exist and are continuous.

Differentiability in higher dimensions


A function f: Rm → Rn is said to be differentiable at a point x0 if there exists a linear map J: Rm → Rn such that

If a function is differentiable at x0, then all of the partial derivatives must exist at x0, in which case the linear map J
is given by the Jacobian matrix.
Note that existence of the partial derivatives (or even all of the directional derivatives) does not guarantee that a
function is differentiable at a point. For example, the function ƒ: R2 → R defined by

is not differentiable at (0, 0), but all of the partial derivatives and directional derivatives exist at this point. For a
continuous example, the function

is not differentiable at (0, 0), but again all of the partial derivatives and directional derivatives exist.
It is known that if the partial derivatives of a function all exist and are continuous in a neighborhood of a point, then
the function must be differentiable at that point, and is in fact of class C1.

Differentiability in complex analysis


In complex analysis, any function that is complex-differentiable in a neighborhood of a point is called holomorphic.
Such a function is necessarily infinitely differentiable, and in fact analytic.

Differentiable functions on manifolds


If M is a differentiable manifold, a real or complex-valued function ƒ on M is said to be differentiable at a point p if it
is differentiable with respect to some (or any) coordinate chart defined around p. More generally, if M and N are
differentiable manifolds, a function ƒ: M → N is said to be differentiable at a point p if it is differentiable with respect
to some (or any) coordinate charts defined around p and ƒ(p).
Differentiable function 47

References
[1] Banach, S. (1931). "Uber die Baire'sche Kategorie gewisser Funktionenmengen". Studia. Math. (3): 174–179.. Cited by Hewitt, E and
Stromberg, K (1963). Real and abstract analysis. Springer-Verlag. Theorem 17.8.

Doubly periodic function


In mathematics, a doubly periodic function is a function defined at all points on the complex plane and having two
"periods", which are complex numbers u and v that are linearly independent as vectors over the field of real numbers.
That u and v are periods of a function ƒ means that

for all values of the complex number z.


The doubly periodic function is thus a two-dimensional extension of the simpler singly periodic function, which
repeats itself in a single dimension. Familiar examples of functions with a single period on the real number line
include the trigonometric functions like cosine and sine. In the complex plane the exponential function ez is a singly
periodic function, with period 2πi.
As an arbitrary mapping from pairs of reals (or complex numbers) to reals, a doubly periodic function can be
constructed with little effort. For example, assume that the periods are 1 and i, so that the repeating lattice is the set
of unit squares with vertices at the Gaussian integers. Values in the prototype square (i.e. x + iy where 0 ≤ x < 1 and
0 ≤ y < 1) can be assigned rather arbitrarily and then 'copied' to adjacent squares. This function will then be
necessarily doubly periodic.
If the vectors 1 and i in this example are replaced by linearly independent vectors u and v the prototype square
becomes a prototype parallelogram, which still tiles the plane. And the "origin" of the lattice of parallelograms does
not have to be the point 0; the lattice can start from any point. In other words, we can think of the plane and its
associated functional values as remaining fixed, and mentally translate the lattice to gain insight into the function's
characteristics.
If a doubly periodic function is also a complex function that satisfies the Cauchy–Riemann equations and provides an
analytic function away from some set of isolated poles – in other words, a meromorphic function – then a lot of
information about such a function can be obtained by applying some basic theorems from complex analysis.
• A non-constant meromorphic doubly periodic function cannot be bounded on the prototype parallelogram. For if
it were it would be bounded everywhere, and therefore constant by Liouville's theorem.
• Since the function is meromorphic, it has no essential singularities and its poles are isolated. Therefore a
translated lattice that does not pass through any pole can be constructed. The contour integral around any
parallelogram in the lattice must vanish, because the values assumed by the doubly periodic function along the
two pairs of parallel sides are identical, and the two pairs of sides are traversed in opposite directions as we move
around the contour. Therefore, by the residue theorem, the function cannot have a single simple pole inside each
parallelogram – it must have at least two simple poles within each parallelogram (Jacobian case), or it must have
one or more poles of order greater than one (Weierstrassian case).
• A similar argument can be applied to the function g = 1/ƒ where ƒ is meromorphic and doubly periodic. Under this
inversion the zeroes of ƒ become the poles of g, and vice versa. So the meromorphic doubly periodic function ƒ
cannot have one simple zero lying within each parallelogram on the lattice—it must have at least two simple
zeroes, or it must have at least one zero of multiplicity greater than one. It follows that ƒ cannot attain any value
just once, since ƒ minus that value would itself be a meromorphic doubly periodic function with just one zero.
Doubly periodic function 48

See also
• Elliptic function
• Fundamental pair of periods
• Jacobi's elliptic functions
• Period mapping
• Weierstrass's elliptic functions

Elementary function
In mathematics, an elementary function is a function built from a finite number of exponentials, logarithms,
constants, one variable, and nth roots through composition and combinations using the four elementary operations (+
– × ÷). By allowing these functions (and constants) to be complex numbers, trigonometric functions and their
inverses become included in the elementary functions (see trigonometric functions and complex exponentials).
The roots of equations are the functions implicitly defined as solving a polynomial equation with constant
coefficients. For polynomials of degree four and smaller there are explicit formulas for the roots (the formulas are
elementary functions).
Elementary functions were introduced by Joseph Liouville in a series of papers from 1833 to 1841. An algebraic
treatment of elementary functions was started by Joseph Fels Ritt in the 1930s.

Examples
Examples of elementary functions include:

and

.
This last function is equal to the inverse cosine trigonometric function in the entire complex domain.
Hence, is an elementary function, too. An example of a function that is not elementary is the error
function

a fact that cannot be seen directly from the definition of elementary function but can be proven using the Risch
algorithm.

Differential algebra
The mathematical definition of an elementary function, or a function in elementary form, is considered in the
context of differential algebra. A differential algebra is an algebra with the extra operation of derivation (algebraic
version of differentiation). Using the derivation operation new equations can be written and their solutions used in
extensions of the algebra. By starting with the field of rational functions, two special types of transcendental
extensions (the logarithm and the exponential) can be added to the field building a tower containing elementary
functions.
A differential field F is a field F0 (rational functions over the rationals Q for example) together with a derivation
map u → ∂u. (Here ∂u is a new function. Sometimes the notation u′ is used.) The derivation captures the properties
of differentiation, so that for any two elements of the base field, the derivation is linear
Elementary function 49

and satisfies the Leibniz product rule

An element h is a constant if ∂h = 0. If the base field is over the rationals, care must be taken when extending the
field to add the needed transcendental constants.
A function u of a differential extension F[u] of a differential field F is an elementary function over F if the function
u
• is algebraic over F, or
• is an exponential, that is, ∂u = u ∂a for a ∈ F, or
• is a logarithm, that is, ∂u = ∂a / u for a ∈ F.
(this is Liouville's theorem).

References
• Maxwell Rosenlicht (1972). "Integration in finite terms" [1]. American Mathematical Monthly (The American
Mathematical Monthly, Vol. 79, No. 9) 79 (9): 963–972. doi:10.2307/2318066.
• Joseph Ritt, Differential Algebra [2], AMS, 1950.

References
[1] http:/ / jstor. org/ stable/ 2318066
[2] http:/ / www. ams. org/ online_bks/ coll33/

Elliptic function
In complex analysis, a mathematical discipline, an elliptic function is a function defined on the complex plane that
is periodic in two directions (a doubly periodic function) and at the same time is meromorphic. Historically, elliptic
functions were discovered as inverse functions of elliptic integrals; these in turn were studied in connection with the
problem of the arc length of an ellipse, whence the name derives.

Definition
Formally, an elliptic function is a meromorphic function f defined on C for which there exist two non-zero complex
numbers a and b with a/b not real, such that
f(z + a) = f(z + b) = f(z)   for all z in C
wherever f(z) is defined. From this it follows that
f(z + ma + nb) = f(z)   for all z in C and all integers m and n.
There are two methods of constructing 'canonical' elliptic functions: those of Jacobi and Weierstrass. In the theory,
modern authors mostly follow Karl Weierstrass: the notations of Weierstrass's elliptic functions based on his
-function are convenient, and any elliptic function can be expressed in terms of these. However it is the functions of
Jacobi that appear most commonly in practical problems, especially the need to avoid complex numbers, having a
mapping from real to real, where the imaginary part is unnecessary or physically insignificant. Weierstrass became
interested in these functions as a student of Christoph Gudermann, a student of Carl Friedrich Gauss.
The elliptic functions introduced by Jacobi, and the auxiliary theta functions (not doubly periodic), are more
complicated but important both for the history and for general theory. The primary difference between these two
theories is that the Weierstrass functions have second-order and higher-order poles located at the corners of the
Elliptic function 50

periodic lattice, whereas the Jacobi functions have simple poles. The development of the Weierstrass theory is easier
to present and understand, having fewer complications.
More generally, the study of elliptic functions is closely related to the study of modular functions and modular
forms, a relationship proven by the modularity theorem. Examples of this relationship include the j-invariant, the
Eisenstein series and the Dedekind eta function.

Properties
• Any complex number ω such that f(z + ω) = f(z) for all z in C is called a period of f. If the two periods a and b are
such that any other period ω can be written as ω = ma + nb with integers m and n, then a and b are called
fundamental periods. Every elliptic function has a fundamental pair of periods, but this pair is not unique, as
described below.
• If a and b are fundamental periods describing a lattice, then exactly the same lattice can be obtained by the
fundamental periods a' and b' where a' = p a + q b and b' = r a + s b where p, q, r and s being integers satisfying p
s − q r = 1. That is, the matrix

has determinant one, and thus belongs to the modular group. In other words, if a and b are fundamental
periods of an elliptic function, then so are a' and b' .
• If a and b are fundamental periods, then any parallelogram with vertices z, z + a, z + b, z + a + b is called a
fundamental parallelogram. Shifting such a parallelogram by integral multiples of a and b yields a copy of the
parallelogram, and the function f behaves identically on all these copies, because of the periodicity.
• The number of poles in any fundamental parallelogram is finite (and the same for all fundamental
parallelograms). Unless the elliptic function is constant, any fundamental parallelogram has at least one pole, a
consequence of Liouville's theorem.
• The sum of the orders of the poles in any fundamental parallelogram is called the order of the elliptic function.
The sum of the residues of the poles in any fundamental parallelogram is equal to zero, so in particular no elliptic
function can have order one.
• The number of zeros (counted with multiplicity) in any fundamental parallelogram is equal to the order of the
elliptic function.
• The set of all elliptic functions which share some two periods form a field.
• The derivative of an elliptic function is again an elliptic function, with the same periods.
• The Weierstrass elliptic function ℘ is the prototypical elliptic function, and in fact, the field of elliptic functions
with respect to a given lattice is generated by ℘ and its derivative ℘′.

References
• Abramowitz, Milton; Stegun, Irene A., eds. (1965), "Chapter 16" [1], Handbook of Mathematical Functions with
Formulas, Graphs, and Mathematical Tables, New York: Dover, pp. 567, MR0167642, ISBN 978-0486612720
See also chapter 18 [2]. (only considers the case of real invariants).
• Naum Illyich Akhiezer, Elements of the Theory of Elliptic Functions, (1970) Moscow, translated into English as
AMS Translations of Mathematical Monographs Volume 79 (1990) AMS, Rhode Island ISBN 0-8218-4532-2
• Tom M. Apostol, Modular Functions and Dirichlet Series in Number Theory, Springer-Verlag, New York, 1976.
ISBN 0-387-97127-0 (See Chapter 1.)
• E. T. Whittaker and G. N. Watson. A course of modern analysis, Cambridge University Press, 1952
Elliptic function 51

External links
• Translation of Niels Abel's Research on Elliptic Functions [3] at Convergence [4]

References
[1] http:/ / www. math. sfu. ca/ ~cbm/ aands/ page_567. htm
[2] http:/ / www. math. sfu. ca/ ~cbm/ aands/ page_627. htm
[3] http:/ / mathdl. maa. org/ convergence/ 1/ ?pa=content& sa=viewDocument& nodeId=1557
[4] http:/ / mathdl. maa. org/ convergence/ 1/

Empty function
In mathematics, an empty function is a function whose domain is the empty set. For each set A, there is exactly one
such empty function

The graph of an empty function is a subset of the Cartesian product ∅×A. Since the product is empty the only such
subset is the empty set ∅. The empty subset is a valid graph since for every x in the domain ∅ there is a unique y in
the codomain A such that (x,y) ∈ ∅. This is an example of a vacuous truth since there is no x in the domain.
Most authors will not care, when defining the term “constant function” precisely, whether or not the empty function
qualifies, and will use whatever definition is most convenient. Sometimes, however, it is best not to consider the
empty function to be constant, and a definition that makes reference to the range is preferable in those situations.
This is much along the same lines of not considering 1 to be a prime number, an empty topological space to be
connected, or the trivial group to be simple.
The existence of an empty function from ∅ to ∅ is required to make the category of sets a category. (In a category,
each objects need to have an "identity morphism", and only the empty function is the identity on the object ∅.) The
existence of a unique empty function from ∅ into each set A means that the empty set is an initial object in the
category of sets. In terms of cardinal arithmetic, it means that k0 = 1 for every cardinal number k - particularly
profound when k=0 to illustrate the strong statement of indices pertaining to 0.

References
• Herrlich, Horst and Strecker, George E.; Category Theory, Allen and Bacon, Inc. Boston (1973).
Entire function 52

Entire function
In complex analysis, an entire function, also called an integral function, is a complex-valued function that is
holomorphic over the whole complex plane. Typical examples of entire functions are the polynomials and the
exponential function, and any sums, products and compositions of these, including the error function and the
trigonometric functions sine and cosine and their hyperbolic counterparts the hyperbolic sine and hyperbolic cosine
functions. Neither the natural logarithm nor the square root functions can be continued analytically to an entire
function.
A transcendental entire function is an entire function that is not a polynomial (see transcendental function).

Properties
Every entire function can be represented as a power series which converges uniformly on compact sets. The
Weierstrass factorization theorem asserts that any entire function can be represented by a product involving its
zeroes.
The entire functions on the complex plane form a commutative ring (in fact a Prüfer domain).
Any entire function f satisfying the inequality for all z with , with n a natural number
[1]
and M and R positive constants, is necessarily a polynomial, of degree at most n.
The special case n = 0 is called Liouville's theorem: any bounded entire function must be constant. Liouville's
theorem may be used to elegantly prove the fundamental theorem of algebra.
As a consequence of Liouville's theorem, any function which is entire on the whole Riemann sphere (complex plane
and the point at infinity) is constant. Thus any non-constant entire function must have a singularity at the complex
point at infinity, either a pole for a polynomial or an essential singularity for a transcendental entire function.
Specifically, by the Casorati–Weierstrass theorem, for any transcendental entire function f and any complex w there
is a sequence (zm)m∈N with and .

Picard's little theorem is a much stronger result: any non-constant entire function takes on every complex number as
value, except possibly one. The latter exception is illustrated by the exponential function, which never takes on the
value 0.
Any entire function f satisfying the inequality for all z with , with n a natural number
and M and R positive constants, is necessarily a polynomial, of degree at least n.

Order and growth


The order (at infinity) of an entire function is defined using the limit superior as:

where is the disk of radius and denotes the supremum norm of on If


one can also define the type:

In other words, the order of is the infimum of all m such that as . The
order need not be finite.
Entire functions may grow as fast as any increasing function: for any increasing function there
exists an entire function such that for all real . Such a function may be easily found of
the form:
Entire function 53

for a conveniently chosen strictly increasing sequence of positive integers . Any such sequence defines an entire
series ; and if it is conveniently chosen, the inequality also holds, for all real .

Other examples
J. E. Littlewood chose the Weierstrass sigma function as a 'typical' entire function in one of his books. Other
examples include the Fresnel integrals, the Jacobi theta function, and the reciprocal Gamma function. The
exponential function and the error function are special cases of the Mittag-Leffler function.

See also
• Jensen's formula
• Carlson's theorem
• Exponential type
• Paley–Wiener theorem

Notes
[1] The converse is also true as for any polynomial of degree n the inequality
holds for any |z| ≥ 1.

References
• Ralph P. Boas (1954). Entire Functions. Academic Press. OCLC 847696 (http://worldcat.org/oclc/847696).
Even and odd functions 54

Even and odd functions


In mathematics, even functions and odd functions are functions which satisfy particular symmetry relations, with
respect to taking additive inverses. They are important in many areas of mathematical analysis, especially the theory
of power series and Fourier series. They are named for the parity of the powers of the power functions which satisfy
each condition: the function f(x) = xn is an even function if n is an even integer, and it is an odd function if n is an
odd integer.

Even functions
Let f(x) be a real-valued function of a real variable. Then f is even if
the following equation holds for all x in the domain of f:

ƒ(x) = x2 is an example of an even function.

Geometrically, the graph of an even function is symmetric with respect to the y-axis, meaning that its graph remains
unchanged after reflection about the y-axis.
Examples of even functions are |x|, x2, x4, cos(x), and cosh(x).

Odd functions
Again, let f(x) be a real-valued function of a real variable. Then f is
odd if the following equation holds for all x in the domain of f:

ƒ(x) = x3 is an example of an odd function.


Even and odd functions 55

or

Geometrically, the graph of an odd function has rotational symmetry with respect to the origin, meaning that its
graph remains unchanged after rotation of 180 degrees about the origin.
Examples of odd functions are x, x3, sin(x), sinh(x), and erf(x).

Some facts
A function's being odd or even does not imply differentiability, or even
continuity. For example, the Dirichlet function is even, but is nowhere
continuous. Properties involving Fourier series, Taylor series,
derivatives and so on may only be used when they can be assumed to
exist.

Basic properties
• The only function which is both even and odd is the constant
function which is identically zero (i.e., f(x) = 0 for all x).
• The sum of an even and odd function is neither even nor odd, unless
one of the functions is identically zero.
• The sum of two even functions is even, and any constant multiple of
ƒ(x) = x3 + 1 is neither even nor odd.
an even function is even.
• The sum of two odd functions is odd, and any constant multiple of
an odd function is odd.
• The product of two even functions is an even function.
• The product of two odd functions is an even function.
• The product of an even function and an odd function is an odd function.
• The quotient of two even functions is an even function.
• The quotient of two odd functions is an even function.
• The quotient of an even function and an odd function is an odd function.
• The derivative of an even function is odd.
• The derivative of an odd function is even.
• The composition of two even functions is even, and the composition of two odd functions is odd.
• The composition of an even function and an odd function is even.
• The composition of any function with an even function is even (but not vice versa).
• The integral of an odd function from −A to +A is zero (where A is finite, and the function has no vertical
asymptotes between −A and A).
• The integral of an even function from −A to +A is twice the integral from 0 to +A (where A is finite, and the
function has no vertical asymptotes between −A and A).
Even and odd functions 56

Series
• The Maclaurin series of an even function includes only even powers.
• The Maclaurin series of an odd function includes only odd powers.
• The Fourier series of a periodic even function includes only cosine terms.
• The Fourier series of a periodic odd function includes only sine terms.

Algebraic structure
• Any linear combination of even functions is even, and the even functions form a vector space over the reals.
Similarly, any linear combination of odd functions is odd, and the odd functions also form a vector space over the
reals. In fact, the vector space of all real-valued functions is the direct sum of the subspaces of even and odd
functions. In other words, every function f(x) can be written uniquely as the sum of an even function and an odd
function:

where

is even and

is odd. For example, if f is exp, then fe is cosh and fo is sinh.


• The even functions form a commutative algebra over the reals. However, the odd functions do not form an
algebra over the reals.

Harmonics
In signal processing, harmonic distortion occurs when a sine wave signal is multiplied by a non-linear transfer
function. The type of harmonics produced depend on the transfer function:[1]
• When the transfer function is even, the resulting signal will consist of only even harmonics of the input sine wave;

• The fundamental is also an odd harmonic, so will not be present.


• A simple example is a full-wave rectifier.
• When it is odd, the resulting signal will consist of only odd harmonics of the input sine wave;
• The output signal will be half-wave symmetric.
• A simple example is clipping in a symmetric push-pull amplifier.
• When it is asymmetric, the resulting signal may contain either even or odd harmonics;
• A simple example is clipping in an asymmetrical class A amplifier.

Notes
[1] Ask the Doctors: Tube vs. Solid-State Harmonics (http:/ / www. uaudio. com/ webzine/ 2005/ october/ content/ content2. html)
Flat function 57

Flat function
In mathematics, especially real analysis, a flat function is a smooth
function ƒ : ℝ → ℝ all of whose derivatives vanish at a given point
x0 ∈ ℝ. The flat functions are, in some sense, the antitheses of the
analytic functions. An analytic function ƒ : ℝ → ℝ is given by a
convergent power series close to some point x0 ∈ ℝ:

The function y = e−1/x2 is flat at x = 0.

In the case of a flat function we see that all derivatives vanish at x0 ∈ ℝ, i.e. ƒ(k)(x0) = 0 for all k ∈ ℕ. This means that
a Taylor series expansion is impossible; i.e. there is no convergent infinite power series. In the language of Taylor's
theorem, the non-constant part of the function always lies in the remainder Rn(x) for all n ∈ ℕ.
Notice that the function need not be flat everywhere. The constant functions on ℝ are flat functions at all of their
points. But there are other, non-trivial, examples.

Example
The function defined by

is flat at x = 0.

References
• Glaister, P. (December 1991), A Flat Function with Some Interesting Properties and an Application [1], The
Mathematical Gazette, Vol. 75, No. 474, pp. 438–440
• Thulin, Fred (2010 November 9), Some Mathematics for Essay 3 [2]

References
[1] http:/ / www. jstor. org/ stable/ 3618627
[2] http:/ / www. math. uic. edu/ ~fthulin/ essay3math. pdf
Function of a real variable 58

Function of a real variable


In mathematics, a function of a real variable is a mathematical function whose domain is the real line. More
loosely, a function of a real variable is sometimes taken to mean any function whose domain is a subset of the real
line.
Functions of a real variable were the classical object of study in mathematical analysis, specifically real analysis. In
that context, a function of a real variable was usually meant a real-valued function of a real variable, that is, a
function whose domain and codomain were the real numbers. However, because of their convenience in fields such
as Fourier analysis, it was also common to consider complex functions of a real variable, that is, a function whose
domain was the real numbers and whose range was the complex numbers.
Until the introduction of functional analysis in the 1920s, the study of functions of a real variable was one of the two
major subdivisions of mathematical analysis. The other was the study of functions of a complex variable, otherwise
known as complex analysis.

Function composition
In mathematics, function composition is the
application of one function to the results of another. For
instance, the functions f: X → Y and g: Y → Z can be
composed by computing the output of g when it has an
argument of f(x) instead of x. Intuitively, if z is a
function g of y and y is a function f of x, then z is a
function of x.

Thus one obtains a composite function g ∘ f: X → Z


g ∘ f, the composition of f and g. For example, (g ∘ f)(c) = #.
defined by (g ∘ f)(x) = g(f(x)) for all x in X. The
notation g ∘ f is read as "g circle f", or "g composed
with f", "g after f", "g following f", or just "g of f".
The composition of functions is always associative. That is, if f, g, and h are three functions with suitably chosen
domains and codomains, then f ∘ (g ∘ h) = (f ∘ g) ∘ h, where the parentheses serve to indicate that composition is to
be performed first for the parenthesized functions. Since there is no distinction between the choices of placement of
parentheses, they may be safely left off.
The functions g and f are said to commute with each other if g ∘ f = f ∘ g. In general, composition of functions will
not be commutative. Commutativity is a special property, attained only by particular functions, and often in special
circumstances. For example, only when . But a function always commutes with its
inverse to produce the identity mapping.
Considering functions as special cases of relations (namely functional relations), one can analogously define
composition of relations, which gives the formula for in terms of and
.
Derivatives of compositions involving differentiable functions can be found using the chain rule. Higher derivatives
of such functions are given by Faà di Bruno's formula.
The structures given by composition are axiomatized and generalized in category theory.
Function composition 59

Example
As an example, suppose that an airplane's elevation at time t is given by the function h(t) and that the oxygen
concentration at elevation x is given by the function c(x). Then (c ∘ h)(t) describes the oxygen concentration around
the plane at time t.

Functional powers
If then may compose with itself; this is sometimes denoted . Thus:

Repeated composition of a function with itself is called function iteration.


The functional powers for natural follow immediately.
• By convention, the identity map on the domain of .
• If admits an inverse function, negative functional powers are defined as the
opposite power of the inverse function, .
Note: If f takes its values in a ring (in particular for real or complex-valued f ), there is a risk of confusion, as f n
could also stand for the n-fold product of f, e.g. f 2(x) = f(x) · f(x).
(For trigonometric functions, usually the latter is meant, at least for positive exponents. For example, in
trigonometry, this superscript notation represents standard exponentiation when used with trigonometric functions:
sin2(x) = sin(x) · sin(x). However, for negative exponents (especially −1), it nevertheless usually refers to the inverse
function, e.g., tan−1 = arctan (≠ 1/tan).
In some cases, an expression for f in g(x) = f r(x) can be derived from the rule for g given non-integer values of r.
This is called fractional iteration. For instance, a half iterate of a function f is a function g satisfying g(g(x)) = f(x).
Another example would be that where f is the successor function, f r(x) = x + r. This idea can be generalized so that
the iteration count becomes a continuous parameter; in this case, such a system is called a flow.
Iterated functions and flows occur naturally in the study of fractals and dynamical systems.

Composition monoids
Suppose one has two (or more) functions f: X → X, g: X → X having the same domain and codomain. Then one can
form long, potentially complicated chains of these functions composed together, such as f ∘ f ∘ g ∘ f. Such long
chains have the algebraic structure of a monoid, called transformation monoid or composition monoid. In general,
composition monoids can have remarkably complicated structure. One particular notable example is the de Rham
curve. The set of all functions f: X → X is called the full transformation semigroup on X.
If the functions are bijective, then the set of all possible combinations of these functions forms a transformation
group; and one says that the group is generated by these functions.
The set of all bijective functions f: X → X form a group with respect to the composition operator. This is the
symmetric group, also sometimes called the composition group.
Function composition 60

Alternative notations
• Many mathematicians omit the composition symbol, writing gf for g ∘ f.
• In the mid-20th century, some mathematicians decided that writing "g ∘ f" to mean "first apply f, then apply g"
was too confusing and decided to change notations. They write "xf" for "f(x)" and "(xf)g" for "g(f(x))". This can be
more natural and seem simpler than writing functions on the left in some areas – in linear algebra, for instance,
where x is a row vector and f and g denote matrices and the composition is by matrix multiplication. This
alternative notation is called postfix notation. The order is important because matrix multiplication is
non-commutative. Successive transformations applying and composing to the right agrees with the left-to-right
reading sequence.
• Mathematicians who use postfix notation may write "fg", meaning first do f then do g, in keeping with the order
the symbols occur in postfix notation, thus making the notation "fg" ambiguous. Computer scientists may write
"f;g" for this, thereby disambiguating the order of composition. To distinguish the left composition operator from
a text semicolon, in the Z notation a fat semicolon ⨟ (U+2A1F) is used for left relation composition. Since all
functions are binary relations, it is correct to use the fat semicolon for function composition as well (see the article
on Composition of relations for further details on this notation).

Composition operator
Given a function g, the composition operator is defined as that operator which maps functions to functions as

Composition operators are studied in the field of operator theory.

External links
• "Composition of Functions [1]" by Bruce Atwood, the Wolfram Demonstrations Project, 2007.

References
[1] http:/ / demonstrations. wolfram. com/ CompositionOfFunctions/
Functional (mathematics) 61

Functional (mathematics)
In mathematics, and particularly in functional analysis, a functional is traditionally a map from a vector space to the
field underlying the vector space, which is usually the real numbers. In other words, it is a function which takes for
its input-argument a vector and returns a scalar. Commonly the vector space is a space of functions, thus the
functional takes a function for its input-argument, then it is sometimes considered a function of a function. Its use
originates in the calculus of variations where one searches for a function which minimizes a certain functional. A
particularly important application in physics is searching for a state of a system which minimizes the energy
functional.
Transformations of functions is a rather more general concept, see Operator (mathematics).

Examples

Duality
Observe that the mapping

is a function, here is an argument of a function . At the same time, the mapping of a function to the value of
the function at a point

is a functional, here is a parameter.


Provided that f is a linear function from a linear vector space to the underlying scalar field, the above linear maps are
dual to each other, and in functional analysis both are called linear functionals.

Integral
Integrals such as

form a special class of functionals. They map a function f into a real number, provided that H is real-valued.
Examples include
• the area underneath the graph of a positive function f

• Lp norm of functions

• the arclength of a curve in 2-dimensional Euclidean space


Functional (mathematics) 62

Vector scalar product


Given any vector in a vector space , the scalar product with another vector , denoted or , is
a scalar. The set of vectors such that this product is zero is a vector subspace of , called the null space or kernel
of .

Functional equation
The traditional usage also applies when one talks about a functional equation, meaning an equation between
functionals: an equation between functionals can be read as an 'equation to solve', with solutions being
themselves functions. In such equations there may be several sets of variable unknowns, like when it is said that an
additive function is one satisfying the functional equation
.

Functional derivative and functional integration


Functional derivatives are used in Lagrangian mechanics. They are derivatives of functionals: i.e. they carry
information on how a functional changes, when the function changes by a small amount. See also calculus of
variations.
Richard Feynman used functional integrals as the central idea in his sum over the histories formulation of quantum
mechanics. This usage implies an integral taken over some function space.

See also
• Linear functional
• Optimization (mathematics)
• Tensor

References
• Rowland, Todd, "Functional [1]" from MathWorld.
• Lang, Serge (2002), "III. Modules, §6. The dual space and dual module", Algebra, Graduate Texts in
Mathematics, 211 (Revised third ed.), New York: Springer-Verlag, pp. 142–146, MR1878556,
ISBN 978-0-387-95385-4

References
[1] http:/ / mathworld. wolfram. com/ Functional. html
Harmonic function 63

Harmonic function
In mathematics, mathematical physics and
the theory of stochastic processes, a
harmonic function is a twice continuously
differentiable function f : U → R (where U
is an open subset of Rn) which satisfies
Laplace's equation, i.e.

A harmonic function defined on an annulus.

everywhere on U. This is usually written as

Examples
Examples of harmonic functions of two variables are:
• The real and imaginary part of any holomorphic function
• The function

defined on (e.g. the electric potential due to a line charge, and the gravity potential due to a long
cylindrical mass)
• The function
Examples of harmonic functions of n variables are:
• The constant, linear and affine functions on all of (for example, the electric potential between the plates of a
capacitor, and the gravity potential of a slab)
• The function on for .
Examples of harmonic functions of three variables are given in the table below with .
Harmonic functions are determined by their singularities. The singular points of the harmonic functions below are
expressed as "charges" and "charge densities" using the terminology of electrostatics, and so the corresponding
harmonic function will be proportional to the electrostatic potential due to these charge distributions. Each function
below will yield another harmonic function when multiplied by a constant, rotated, and/or has a constant added. The
inversion of each function will yield another harmonic function which has singularities which are the images of the
original singularities in a spherical "mirror". Also, the sum of any two harmonic functions will yield another
harmonic function.
Harmonic function 64

Function Singularity

Unit point charge at origin

x-directed dipole at origin

Line of unit charge density on entire z-axis

Line of unit charge density on negative z-axis

Line of x-directed dipoles on entire z axis

Line of x-directed dipoles on negative z axis

Remarks
The set of harmonic functions on a given open set U can be seen as the kernel of the Laplace operator Δ and is
therefore a vector space over R: sums, differences and scalar multiples of harmonic functions are again harmonic.
If f is a harmonic function on U, then all partial derivatives of f are also harmonic functions on U. The Laplace
operator Δ and the partial derivative operator will commute on this class of functions.
In several ways, the harmonic functions are real analogues to holomorphic functions. All harmonic functions are
analytic, i.e. they can be locally expressed as power series. This is a general fact about elliptic operators, of which
the Laplacian is a major example.
The uniform limit of a convergent sequence of harmonic functions is still harmonic. This is true because any
continuous function satisfying the mean value property is harmonic. Consider the sequence on ( , 0)× R
defined by . This sequence is harmonic and converges uniformly to the zero function;
however note that the partial derivatives are not uniformly convergent to the zero function (the derivative of the zero
function). This example shows the importance of relying on the mean value property and continuity to argue that the
limit is harmonic.

Connections with complex function theory


The real and imaginary part of any holomorphic function yield harmonic functions on R2 (these are said to be a pair
of harmonic conjugate functions). Conversely, any harmonic function on an open set is locally the real
part of a holomorphic function. This is immediately seen observing that, writing the complex
function is holomorphic in because it satisfies the Cauchy-Riemann equations. Therefore, g
has locally a primitive , and is the real part of up to a constant, as is the real part of .
Although the above correspondence with holomorphic functions only holds for functions of two real variables, still
harmonic functions in n variables enjoy a number of properties typical of holomorphic functions. They are (real)
analytic; they have a maximum principle and a mean-value principle; a theorem of removal of singularities as well as
a Liouville theorem one holds for them in analogy to the corresponding theorems in complex functions theory.
Harmonic function 65

Properties of harmonic functions


Some important properties of harmonic functions can be deduced from Laplace's equation.

Regularity theorem for harmonic functions


Harmonic functions are infinitely differentiable. In fact, harmonic functions are real analytic.

Maximum principle
Harmonic functions satisfy the following maximum principle: if K is any compact subset of U, then f, restricted to K,
attains its maximum and minimum on the boundary of K. If U is connected, this means that f cannot have local
maxima or minima, other than the exceptional case where f is constant. Similar properties can be shown for
subharmonic functions.

Mean value property


If B(x,r) is a ball with center x and radius r which is completely contained in the open set , then the value
of a harmonic function at the center of the ball is given by the average value of on the
surface of the ball; this average value is also equal to the average value of in the interior of the ball. In other
words

where is the volume of the unit ball in n dimensions and is the n−1 dimensional surface measure. The mean
value theorem follows by verifying that the spherical mean of u is constant:

which in turn follows by making a change of variable and then applying Green's theorem.
As a consequence of the mean value theorem, u is preserved by the convolution of a harmonic function u with any
radial function η with total integral one. More precisely, if η is an integrable radial function supported in B(0,ε) and
∫η = 1, then

provided that B(x,ε) ⊂ Ω. In particular, by taking η to be a C∞ function, the convolution η∗u is also smooth, and
therefore harmonic functions are smooth throughout their domains (in fact, real analytic, by the Poisson integral
representation). Similar arguments also show that harmonic distributions are, in fact, (smooth) harmonic functions
(Weyl's lemma).
The converse to the mean value theorem also holds: all locally integrable functions satisfying the (volume)
mean-value property are infinitely differentiable and harmonic functions as well. This follows for C2 functions again
by the method of spherical means. For locally integrable functions, it follows since the mean value property implies
that u is unchanged when convolved with any radial mollifier of total integral one, but convolutions with mollifiers
are smooth and so the C2 result can still be applied.
Harmonic function 66

Harnack's inequality
Let u be a non-negative harmonic function in a bounded domain Ω. Then for every connected set

Harnack's inequality

holds for some constant C that depends only on V and Ω.

Removal of singularities
The following principle of removal of singularities holds for harmonic functions. If f is a harmonic function defined
on a dotted open subset of Rn, which is less singular at than the fundamental solution, that is

then f extends to a harmonic function on (compare Riemann's theorem for functions of a complex variable).

Liouville's theorem
If f is a harmonic function defined on all of Rn which is bounded above or bounded below, then f is constant
(compare Liouville's theorem for functions of a complex variable).

Generalizations

Weakly harmonic function


A function (or, more generally, a distribution) is weakly harmonic if it satisfies Laplace's equation

in a weak sense (or, equivalently, in the sense of distributions). A weakly harmonic function coincides almost
everywhere with a strongly harmonic function, and is in particular smooth. A weakly harmonic distribution is
precisely the distribution associated to a strongly harmonic function, and so also is smooth. This is Weyl's lemma.
There are other weak formulations of Laplace's equation that are often useful. One of which is Dirichlet's principle,
representing harmonic functions in the Sobolev space H1(Ω) as the minimizers of the Dirichlet energy integral

with respect to local variations, that is, all functions such that holds for all
or equivalently, for all

Harmonic functions on manifolds


Harmonic functions can be defined on an arbitrary Riemannian manifold, using the Laplace-Beltrami operator Δ. In
this context, a function is called harmonic if

Many of the properties of harmonic functions on domains in Euclidean space carry over to this more general setting,
including the mean value theorem (over geodesic balls), the maximum principle, and the Harnack inequality. With
the exception of the mean value theorem, these are easy consequences of the corresponding results for general linear
elliptic partial differential equations of the second order.
Harmonic function 67

Subharmonic functions
A C2 function that satisfies is called subharmonic. This condition guarantees that the maximum principle
will hold, although other properties of harmonic functions may fail. More generally, a function is subharmonic if and
only if, in the interior of any ball in its domain, its graph lies below that of the harmonic function interpolating its
boundary values on the ball.

Harmonic forms
One generalization of the study of harmonic functions is the study of harmonic forms on Riemannian manifolds, and
it is related to the study of cohomology. Also, it is possible to define harmonic vector-valued functions, or harmonic
maps of two Riemannian manifolds, which are critical points of a generalized Dirichlet energy functional (this
includes harmonic functions as a special case, a result known as Dirichlet principle). These kind of harmonic maps
appear in the theory of minimal surfaces. For example, a curve, that is, a map from an interval in R to a Riemannian
manifold, is a harmonic map if and only if it is a geodesic.

Harmonic maps between manifolds


If M and N are two Riemannian manifolds, then a harmonic map u : M → N is defined to be a stationary point of the
Dirichlet energy

in which du : TM → TN is the differential of u, and the norm is that induced by the metric on M and that on N on the
tensor product bundle T∗M⊗u−1TN.
Important special cases of harmonic maps between manifolds include minimal surfaces, which are precisely the
harmonic immersions of a surface into three-dimensional Euclidean space. More generally, minimal submanifolds
are harmonic immersions of one manifold in another. Harmonic coordinates are a harmonic diffeomorphism from a
manifold to an open subset of a Euclidean space of the same dimension.

References
• Evans, Lawrence C. (1998), Partial Differential Equations, American Mathematical Society.
• Gilbarg, David; Trudinger, Neil, Elliptic Partial Differential Equations of Second Order, ISBN 3-540-41160-7.
• Jost, Jürgen (2005), Riemannian Geometry and Geometric Analysis (4th ed.), Berlin, New York: Springer-Verlag,
ISBN 978-3-540-25907-7.
• Han, Q.; Lin, F. (2000), Elliptic Partial Differential Equations, American Mathematical Society.

External links
• Weisstein, Eric W., "Harmonic Function [1]" from MathWorld.
• Harmonic Functions Module by John H. Mathews [2]
• Harmonic Function Theory by S.Axler, Paul Bourdon, and Wade Ramey [3]

References
[1] http:/ / mathworld. wolfram. com/ HarmonicFunction. html
[2] http:/ / math. fullerton. edu/ mathews/ c2003/ HarmonicFunctionMod. html
[3] http:/ / www. axler. net/ HFT. html
Hermitian function 68

Hermitian function
In mathematical analysis, a Hermitian function is a complex function with the property that its complex conjugate
is equal to the original function with the variable changed in sign:

for all in the domain of .


This definition extends also to functions of two or more variables, e.g., in the case that is a function of two
variables it is Hermitian if

for all pairs in the domain of .


From this definition it follows immediately that, if is a Hermitian function, then
• the real part of is an even function
• the imaginary part of is an odd function

Motivation
Hermitian functions appear frequently in mathematics, physics, and signal processing. For example, the following
two statements follows from basic properties of the Fourier transform:
• The function is real-valued if and only if the Fourier transform of is Hermitian.
• The function is Hermitian if and only if the Fourier transform of is real-valued.
Since the Fourier transform of a real signal is guaranteed to be Hermitian, it can be compressed using the Hermitian
even/odd symmetry. This, for example, allows the discrete Fourier transform of a signal (which is in general
complex) to be stored in the same space as the original real signal.
• If either f or g is Hermitian, then
Where the is correlation, and is convolution. Because convolution is commutative we can infer also that:
• If either f or g is Hermitian, then , which in general is not true.

See also
• Even and odd functions
Holomorphic function 69

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 neighbourhood of z0 in the
A rectangular grid (top) and its image under a
complex plane.
holomorphic function f (bottom).

Definition
Given a complex-valued function ƒ of a single complex variable, the derivative of ƒ 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 ƒ is 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 ƒ is complex differentiable at every point z0 in U, we say that ƒ is holomorphic on U. We say that ƒ is holomorphic
at the point z0 if it is holomorphic on some neighborhood of z0. We say that ƒ is holomorphic on some non-open set A
if it is holomorphic in an open set containing A.
The relationship between real differentiability and complex differentiability is the following. If a complex function
ƒ(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 Cauchy–Riemann equations:

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 Cauchy–Riemann equations, then ƒ is holomorphic. A more satisfying
converse, which is much harder to prove, is the Looman–Menchoff theorem: if ƒ is continuous, u and v have first
Holomorphic function 70

partial derivatives, and they satisfy the Cauchy–Riemann equations, then ƒ is holomorphic.

Terminology
The word "holomorphic" was introduced by two of Cauchy's students, Briot (1817–1882) and Bouquet (1819–1895),
and derives from the Greek ὅλος (holos) meaning "entire", and μορφή (morphē) 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 around .


2
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.
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 geometrical 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

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 γ from z0 to z lying entirely in U, define

;
Holomorphic function 71

in light of the Jordan curve theorem and the generalized Stokes' theorem, Fγ(z) is independent of the particular
choice of path γ, and thus F(z) is a well-defined function on U having F(z0) = F0 and dF = f dz.)

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 Cauchy–Riemann 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 Cauchy–Riemann equations; in fact it can be stated as
follows:
A function of several complex variables is holomorphic if and only if it satisfies the Cauchy–Riemann 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 Fréchet or Gâteaux 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. ISBN 0-8218-3780-X. .

See also
• Antiderivative (complex analysis)
• Antiholomorphic function
• Biholomorphy
• Meromorphic function
• Quadrature domains
Homogeneous function 72

Homogeneous function
In mathematics, an homogeneous function is a function with multiplicative scaling behaviour: if the argument is
multiplied by a factor, then the result is multiplied by some power of this factor. More precisely, if ƒ : V → W is a
function between two vector spaces over a field F, then ƒ is said to be homogeneous of degree k ∈ F if {{{}}}

(1)

{{{}}} for all nonzero α ∈ F and v ∈ V. When the vector spaces involved are over the real numbers, a slightly more
general form of homogeneity is often used, requiring only that (1) hold for all α > 0.
Homogeneous functions can also be defined for vector spaces with the origin deleted, a fact that is used in the
definition of sheaves on projective space in algebraic geometry. More generally, if S ⊂ V is any subset that is
invariant under scalar multiplication by elements of the field (a "cone"), then an homogeneous function from S to W
can still be defined by (1).

Examples

Linear functions
Any linear function ƒ : V → W is homogeneous of degree 1, since by the definition of linearity

for all α ∈ F and v ∈ V. Similarly, any multilinear function ƒ : V1 × V2 × ... Vn → W is homogeneous of degree n,
since by the definition of multilinearity

for all α ∈ F and v1 ∈ V1, v2 ∈ V2, ..., vn ∈ Vn. It follows that the n-th differential of a function ƒ : X → Y between two
Banach spaces X and Y is homogeneous of degree n.

Homogeneous polynomials
Monomials in n variables define homogeneous functions ƒ : Fn → F. For example,

is homogeneous of degree 10 since


.
The degree is the sum of the exponents on the variables; in this example, 10=5+2+3.
A homogeneous polynomial is a polynomial made up of a sum of monomials of the same degree. For example,

is a homogeneous polynomial of degree 5. Homogeneous polynomials also define homogeneous functions.


Homogeneous function 73

Polarization
A multilinear function g : V × V × ... V → F from the n-th Cartesian product of V with itself to the groundfield F
gives rise to an homogeneous function ƒ : V → F by evaluating on the diagonal:

The resulting function ƒ is a polynomial on the vector space V.


Conversely, if F has characteristic zero, then given an homogeneous polynomial ƒ of degree n on V, the polarization
of ƒ is a multilinear function g : V × V × ... V → F on the n-th Cartesian product of V. The polarization is defined by

These two constructions, one of an homogeneous polynomial from a multilinear form and the other of a multilinear
form from an homogeneous polynomial, are mutually inverse to one another. In finite dimensions, they establish an
isomorphism of graded vector spaces from the symmetric algebra of V∗ to the algebra of homogeneous polynomials
on V.

Rational functions
Rational functions formed as the ratio of two homogeneous polynomials are homogeneous functions off of the affine
cone cut out by the zero locus of the denominator. Thus, if f is homogeneous of degree m and g is homogeneous of
degree n, then f/g is homogeneous of degree m − n away from the zeros of g.

Non-Examples

Logarithms
The natural logarithm scales additively and so is not homogeneous.
This can be proved by noting that , , and
. Therefore such that .

Affine functions
The function does not scale multiplicatively.

Positive homogeneity
In the special case of vector spaces over the real numbers, the notation of positive homogeneity often plays a more
important role than homogeneity in the above sense. A function ƒ : V \ {0} → R is positive homogeneous of degree k
if

for all α > 0. Here k can be any complex number. A (nonzero) continuous function homogeneous of degree k on
Rn \ {0} extends continuously to Rn if and only if Re{k} > 0.
Positive homogeneous functions are characterized by Euler's homogeneous function theorem. Suppose that the
function ƒ : Rn \ {0} → R is continuously differentiable. Then ƒ is positive homogeneous of degree k if and only if

This result follows at once by differentiating both sides of the equation ƒ(αy) = αkƒ(y) with respect to α and applying
the chain rule. The converse holds by integrating.
As a consequence, suppose that ƒ : Rn → R is differentiable and homogeneous of degree k. Then its first-order partial
derivatives are homogeneous of degree k − 1. The result follows from Euler's theorem by commuting the
Homogeneous function 74

operator with the partial derivative.

Homogeneous distributions
A compactly supported continuous function ƒ on Rn is homogeneous of degree k if and only if

for all compactly supported test functions φ and nonzero real t. Equivalently, making a change of variable y = tx, ƒ is
homogeneous of degree k if and only if

for all t and all test functions φ. The last display makes it possible to define homogeneity of distributions. A
distribution S is homogeneous of degree k if

for all nonzero real t and all test functions φ. Here the angle brackets denote the pairing between distributions and
test functions, and μt : Rn → Rn is the mapping of scalar multiplication by the real number t.

Application to differential equations


The substitution v = y/x converts the ordinary differential equation

where I and J are homogeneous functions of the same degree, into the separable differential equation

See also
• Weierstrass elliptic function
• Triangle center function
• Production function

References
• Blatter, Christian (1979). "20. Mehrdimensionale Differentialrechnung, Aufgaben, 1." (in German). Analysis II
(2nd ed.). Springer Verlag. pp. 188. ISBN 3-540-09484-9.

External links
• Homogeneous function [1] on PlanetMath

References
[1] http:/ / planetmath. org/ ?op=getobj& amp;from=objects& amp;id=6381
Identity function 75

Identity function
In mathematics, an identity function, also called identity map or identity transformation, is a function that always
returns the same value that was used as its argument. In terms of equations, the function is given by f(x) = x.

Definition
Formally, if M is a set, the identity function f on M is defined to be that function with domain and codomain M which
satisfies
f(x) = x    for all elements x in M.
In other words, the function assigns to each element x of M the element x of M.
The identity function f on M is often denoted by idM or 1M.
In terms of set theory, where a function is defined as a particular kind of binary relation, the identity function is
given by the identity relation, or diagonal of M.

Algebraic property
If f : M → N is any function, then we have f o idM = f = idN o f (where "o" denotes function composition). In
particular, idM is the identity element of the monoid of all functions from M to M.
Since the identity element of a monoid is unique, one can alternately define the identity function on M to be this
identity element. Such a definition generalizes to the concept of an identity morphism in category theory, where the
endomorphisms of M need not be functions.

Examples
• The identity function is a linear operator, when applied to vector spaces.
• The identity function on the positive integers is a completely multiplicative function (essentially multiplication by
1), considered in number theory.
• In an n-dimensional vector space the identity function is represented by the identity matrix In, regardless of the
basis.
• In a metric space the identity is trivially an isometry. An object without any symmetry has as symmetry group the
trivial group only containing this isometry (symmetry type C1).
Implicit and explicit functions 76

Implicit and explicit functions


In mathematics, an implicit function is a function in which the dependent variable has not been given "explicitly" in
terms of the independent variable. To give a function f explicitly is to provide a prescription for determining the
output value of the function y in terms of the input value x:
y = f(x).
By contrast, the function is implicit if the value of y is obtained from x by solving an equation of the form:
R(x,y) = 0.
That is, it is defined as the level set of a function in two variables: one variable or the other may determine the other,
but one is not given an explicit formula for one in terms of the other.
Implicit functions can often be useful in situations where it is inconvenient to solve explicitly an equation of the form
R(x,y) = 0 for y in terms of x. Even if it is possible to rearrange this equation to obtain y as an explicit function f(x), it
may not be desirable to do so since the expression of f may be much more complicated than the expression of R. In
other situations, the equation R(x,y) = 0 may fail to define a function at all, and rather defines a kind of
multiple-valued function. Nevertheless, in many situations, it is still possible to work with implicit functions. Some
techniques from calculus, such as differentiation, can be performed with relative ease using implicit differentiation.
The implicit function theorem provides a link between implicit and explicit functions. It states that if the equation
R(x, y) = 0 satisfies some mild conditions on its partial derivatives, then one can in principle solve this equation for y,
at least over some small interval. Geometrically, the graph defined by R(x,y) = 0 will overlap locally with the graph
of a function y = f(x).
Various numerical methods exist for solving the equation R(x,y)=0 to find an approximation to the implicit function
y. Many of these methods are iterative in that they produce successively better approximations, so that a prescribed
accuracy can be achieved. Many of these iterative methods are based on some form of Newton's method.

Examples

Inverse functions
Implicit functions commonly arise as one way of describing the notion of an inverse function. If f is a function, then
the inverse function of f is a solution of the equation

for y in terms of x. Intuitively, an inverse function is obtained from f by interchanging the roles of the dependent and
independent variables. Stated another way, the inverse function is the solution y of the equation

Examples.
1. The natural logarithm y = ln(x) is the solution of the equation x − ey = 0.
2. The product log is an implicit function given by x − y ey = 0.
Implicit and explicit functions 77

Algebraic functions
An algebraic function is a solution y for an equation R(x,y) = 0 where R is a polynomial of two variables. Algebraic
functions play an important role in mathematical analysis and algebraic geometry. A simple example of an algebraic
function is given by the unit circle:

Solving for y gives

Note that there are two "branches" to the implicit function: one where the sign is positive and the other where it is
negative.

Caveats
Not every equation R(x, y) = 0 has a graph that is the graph of a function, the circle equation being one prominent
example. Another example is an implicit function given by x − C(y) = 0 where C is a cubic polynomial having a
"hump" in its graph. Thus, for an implicit function to be a true function it might be necessary to use just part of the
graph. An implicit function can sometimes be successfully defined as a true function only after "zooming in" on
some part of the x-axis and "cutting away" some unwanted function branches. A resulting formula may only then
qualify as a legitimate explicit function.
The defining equation R = 0 can also have other pathologies. For example, the implicit equation x = 0 does not define
a function at all; it is a vertical line. In order to avoid a problem like this, various constraints are frequently imposed
on the allowable sorts of equations or on the domain. The implicit function theorem provides a uniform way of
handling these sorts of pathologies.

Implicit differentiation
In calculus, a method called implicit differentiation makes use of the chain rule to differentiate implicitly defined
functions.
As explained in the introduction, y can be given as a function of x implicitly rather than explicitly. When we have an
equation R(x, y) = 0, we may be able to solve it for y and then differentiate. However, sometimes it is simpler to
differentiate R(x, y) with respect to x and then solve for dy/dx.

Examples
1. Consider for example

This function normally can be manipulated by using algebra to change this equation to an explicit function:

Differentiation then gives . Alternatively, one can differentiate the equation:

Solving for :
Implicit and explicit functions 78

2. An example of an implicit function, for which implicit differentiation might be easier than attempting to use
explicit differentiation, is

In order to differentiate this explicitly with respect to x, one would have to obtain (via algebra)

and then differentiate this function. This creates two derivatives: one for y > 0 and another for y < 0.
One might find it substantially easier to implicitly differentiate the implicit function;

thus,

3. Sometimes standard explicit differentiation cannot be used and, in order to obtain the derivative, another method
such as implicit differentiation must be employed. An example of such a case is the implicit function y5 − y = x. It is
impossible to express y explicitly as a function of x and dy/dx therefore this cannot be found by explicit
differentiation. Using the implicit method, dy/dx can be expressed:

factoring out shows that

which yields the final answer

where

Formula for two variables


"The Implicit Function Theorem states that if is defined on an open disk containing , where
, , and and are continuous on the disk, then the equation defines as a
[1] :§ 11.5
function of near the point and the derivative of this function is given by..."

, indicates the derivative of with respect to x and y


The above formula comes from using the generalized chain rule to obtain the total derivative—with respect to x—of
both sides of F(x, y) = 0:
Implicit and explicit functions 79

Marginal rate of substitution


In economics, when the level set is an indifference curve, the implicit derivative (or rather, −1 times the implicit
derivative) is interpreted as the marginal rate of substitution of the two variables: how much more of y one must
receive in order to be indifferent to a loss of 1 unit of x.

Implicit function theorem


It can be shown that if is given by a smooth submanifold in , and is a point of this

submanifold such that the tangent space there is not vertical (that is ), then in some small enough

neighbourhood of is given by a parametrization where is a smooth function. In less technical


language, implicit functions exist and can be differentiated, unless the tangent to the supposed graph would be
vertical. In the standard case where we are given an equation

the condition on can be checked by means of partial derivatives.[1] :§ 11.5

References
[1] Stewart, James (1998). Calculus Concepts And Contexts. Brooks/Cole Publishing Company. ISBN 0-534-34330-9.

• Rudin, Walter (1976). Principles of Mathematical Analysis. McGraw-Hill. ISBN 0-07-054235-X.


• Spivak, Michael (1965). Calculus on Manifolds. HarperCollins. ISBN 0-8053-9021-9.
• Warner, Frank (1983). Foundations of Differentiable Manifolds and Lie Groups. Springer. ISBN 0-387-90894-3.

Indicator function
In mathematics, an indicator function or a characteristic function is
a function defined on a set that indicates membership of an
element in a subset of , having the value 1 for all elements of A
and the value 0 for all elements of X not in A.

Definition
The indicator function of a subset of a set is a function
The graph of the indicator function of a
two-dimensional subset of a square.

defined as

The Iverson bracket allows the equivalent notation, , to be used instead of


The indicator function of is sometimes denoted
or or even .
(The Greek letter χ because it is the initial letter of the Greek etymon of the word characteristic.)
Indicator function 80

Remark on notation and terminology


• The notation may signify the identity function.
• The notation may signify the characteristic function in convex analysis.
A related concept in statistics is that of a dummy variable (this must not be confused with "dummy variables" as that
term is usually used in mathematics, also called a bound variable).
The term "characteristic function" has an unrelated meaning in probability theory. For this reason, probabilists use
the term indicator function for the function defined here almost exclusively, while mathematicians in other fields
are more likely to use the term characteristic function to describe the function which indicates membership in a set.

Basic properties
The indicator or characteristic function of a subset of some set , maps elements of to the range
.
This mapping is surjective only when is a proper subset of . If , then . By a similar
argument, if then .
In the following, the dot represents multiplication, 1·1 = 1, 1·0 = 0 etc. "+" and "−" represent addition and
subtraction. " " and " " is intersection and union, respectively.
If and are two subsets of , then

and the "complement" of the indicator function of A i.e. AC is:

More generally, suppose is a collection of subsets of . For any ,

is clearly a product of s and s. This product has the value 1 at precisely those which belong to none of
the sets and is otherwise. That is

Expanding the product on the left hand side,

where is the cardinality of . This is one form of the principle of inclusion-exclusion.


As suggested by the previous example, the indicator function is a useful notational device in combinatorics. The
notation is used in other places as well, for instance in probability theory: if is a probability space with
probability measure and is a measurable set, then becomes a random variable whose expected value is
equal to the probability of

This identity is used in a simple proof of Markov's inequality.


In many cases, such as order theory, the inverse of the indicator function may be defined. This is commonly called
the generalized Möbius function, as a generalization of the inverse of the indicator function in elementary number
theory, the Möbius function. (See paragraph below about the use of the inverse in classical recursion theory.)
Indicator function 81

Mean, Variance and covariance


Given a probability space: with A , the indicator random variable , is defined by
if otherwise (i.e basically is an indicator random variable)
(mean)
(Variance)
(Covariance)

Characteristic function in recursion theory, Gödel's and Kleene's representing


function
Kurt Gödel described the representing function in his 1934 paper "On Undecidable Propositions of Formal
Mathematical Systems". (The paper appears on pp. 41-74 in Martin Davis ed. The Undecidable):
"There shall correspond to each class or relation R a representing function φ(x1, . . ., xn) = 0 if R(x1, . . ., xn)
and φ(x1, . . ., xn)=1 if ~R(x1, . . ., xn)." (p. 42; the "~" indicates logical inversion i.e. "NOT")
Stephen Kleene (1952) (p. 227) offers up the same definition in the context of the primitive recursive functions as a
function φ of a predicate P, takes on values 0 if the predicate is true and 1 if the predicate is false.
For example, because the product of characteristic functions φ1*φ2* . . . *φn = 0 whenever any one of the functions
equals 0, it plays the role of logical OR: IF φ1=0 OR φ2=0 OR . . . OR φn=0 THEN their product is 0. What appears
to the modern reader as the representing function's logical-inversion, i.e. the representing function is 0 when the
function R is "true" or satisfied", plays a useful role in Kleene's definition of the logical functions OR, AND, and
IMPLY (p. 228), the bounded- (p. 228) and unbounded- (p. 279ff) mu operators (Kleene (1952)) and the CASE
function (p. 229).

Characteristic function in fuzzy set theory


In classical mathematics, characteristic functions of sets only take values 1 (members) or 0 (non-members). In fuzzy
set theory, characteristic functions are generalized to take value in the real unit interval [0, 1], or more generally, in
some algebra or structure (usually required to be at least a poset or lattice). Such generalized characteristic functions
are more usually called membership functions, and the corresponding "sets" are called fuzzy sets. Fuzzy sets model
the gradual change in the membership degree seen in many real-world predicates like "tall", "warm", etc.

References
• Folland, G.B. (1999). Real Analysis: Modern Techniques and Their Applications (Second ed.). John Wiley &
Sons, Inc..
• Cormen, Thomas H.; Leiserson, Charles E.; Rivest, Ronald L.; Stein, Clifford (2001). "Section 5.2: Indicator
random variables". Introduction to Algorithms (Second Edition ed.). MIT Press and McGraw-Hill. pp. 94–99.
ISBN 0-262-03293-7.
• Davis, Martin, ed (1965). The Undecidable. New York: Raven Press Books, Ltd..
• Kleene, Stephen (1971) [1952]. Introduction to Metamathematics (Sixth Reprint with corrections). Netherlands:
Wolters-Noordhoff Publishing and North Holland Publishing Company.
• Boolos, George; Burgess, John P.; Jeffrey, Richard C. (2002). Computability and Logic. Cambridge UK:
Cambridge University Press. ISBN 0-521-00758-5.
• Zadeh, Lotfi A. (June 1965). "Fuzzy sets" [1] (PDF). Information and Control 8 (3): 338–353.
doi:10.1016/S0019-9958(65)90241-X.
• Goguen, Joseph (1967). "L-fuzzy sets". Journal of Mathematical Analysis and Applications 18 (1): 145–174.
doi:10.1016/0022-247X(67)90189-8.
Indicator function 82

References
[1] http:/ / www-bisc. cs. berkeley. edu/ zadeh/ papers/ Fuzzy%20Sets-1965. pdf

Injective function
In mathematics, an injective function is a function that preserves
distinctness: it never maps distinct elements of its domain to the
same element of its codomain. In other words, every element of
the function's codomain is mapped to by at most one element of its
domain. If in addition all of the elements in the codomain are in
fact mapped to by some element of the domain, then the function
is said to be bijective (see figures).

An injective function is called an injection, and is also said to be a


one-to-one function (not to be confused with one-to-one
correspondence, i.e. a bijective function). Occasionally, an
injective function from X to Y is denoted f: X ↣ Y, using an arrow
with a barbed tail. Alternately, it may be denoted YX using a
notation derived from that used for falling factorial powers, since An injective function (not a bijection)
if X and Y are finite sets with respectively x and y elements, the
number of injections X ↣ Y is yx (see the twelvefold way).

A function f that is not injective is sometimes called many-to-one.


(However, this terminology is also sometimes used to mean
"single-valued", i.e., each argument is mapped to at most one
value; this is the case for any function, but is used to stress the
opposition with multi-valued functions, which are not true
functions.)

A monomorphism is a generalization of an injective function in


category theory.

Definition
Let f be a function whose domain is a set A. The function f is
Another injective function (is a bijection)
injective if for all a and b in A, if f(a) = f(b), then a = b; that is,
f(a) = f(b) implies a = b.  Equivalently, if a ≠ b, then f(a) ≠ f(b).

Examples
• For any set X and any subset S of X the inclusion map S → X (which sends any element s of S to itself) is
injective. In particular the identity function X → X is always injective (and in fact bijective).
• The function f : R → R defined by f(x) = 2x + 1 is injective.
Injective function 83

• The function g : R → R defined by g(x) = x2 is not injective,


because (for example) g(1) = 1 = g(−1). However, if g is
redefined so that its domain is the non-negative real numbers
[0,+∞), then g is injective.
• The exponential function exp : R → R defined by exp(x) = ex is
injective (but not surjective as no value maps to a negative
number).
• The natural logarithm function ln : (0, ∞) → R defined by x ↦
ln x is injective.
• The function g : R → R defined by g(x) = xn − x is not
injective, since, for example, g(0) = g(1).
More generally, when X and Y are both the real line R, then an
injective function f : R → R is one whose graph is never A non-injective function (this one happens to be a
intersected by any horizontal line more than once. This principle is surjection)

referred to as the horizontal line test.

Injections can be undone


Functions with left inverses are always injections. That is, given f : X → Y, if there is a function g : Y → X such that,
for every x ∈ X
g(f(x)) = x (f can be undone by g)
then f is injective. In this case, f is called a section of g and g is called a retraction of f.
Conversely, every injection f with non-empty domain has a left inverse g (in conventional mathematics[1] ). Note that
g may not be a complete inverse of f because the composition in the other order, f ∘ g, may not be the identity on Y.
In other words, a function that can be undone or "reversed", such as f, is not necessarily invertible (bijective).
Injections are "reversible" but not always invertible.
Although it is impossible to reverse a non-injective (and therefore information-losing) function, one can at least
obtain a "quasi-inverse" of it, that is a multiple-valued function.
Injective function 84

Injections may be made invertible


In fact, to turn an injective function f : X → Y into a bijective (hence invertible) function, it suffices to replace its
codomain Y by its actual range J = f(X). That is, let g : X → J such that g(x) = f(x) for all x in X; then g is bijective.
Indeed, f can be factored as inclJ,Y ∘ g, where inclJ,Y is the inclusion function from J into Y.

Other properties
• If f and g are both injective, then f ∘ g is injective.
• If g ∘ f is injective, then f is injective (but
g need not be).
• f : X → Y is injective if and only if, given
any functions g, h : W → X, whenever
f ∘ g = f ∘ h, then g = h. In other words,
injective functions are precisely the
monomorphisms in the category Set of
sets.
• If f : X → Y is injective and A is a subset
of X, then f −1(f(A)) = A. Thus, A can be
recovered from its image f(A).
• If f : X → Y is injective and A and B are
both subsets of X, then f(A ∩ B) = The composition of two injective functions is injective.
f(A) ∩ f(B).
• Every function h : W → Y can be decomposed as h = f ∘ g for a suitable injection f and surjection g. This
decomposition is unique up to isomorphism, and f may be thought of as the inclusion function of the range h(W)
of h as a subset of the codomain Y of h.
• If f : X → Y is an injective function, then Y has at least as many elements as X, in the sense of cardinal numbers. In
particular, if, in addition, there is an injection from to , then and has the same cardinal number.
(This is known as the Cantor–Bernstein–Schroeder theorem.)
• If both X and Y are finite with the same number of elements, then f : X → Y is injective if and only if f is surjective
(in which case they are bijective).
• An injective function which is a homomorphism between two algebraic structures is an embedding.

See also
• Surjective function
• Injective module
• Horizontal line test
• Injective metric space

Notes
[1] This principle is valid in conventional mathematics, but may fail in constructive mathematics. For instance, a left inverse of the inclusion
{0,1} → R of the two-element set in the reals violates indecomposability by giving a retraction of the real line to the set {0,1}.
Injective function 85

References
• 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.

External links
• Earliest Uses of Some of the Words of Mathematics: entry on Injection, Surjection and Bijection has the history
of Injection and related terms. (http://jeff560.tripod.com/i.html)

Invex function
In vector calculus, an invex function is a differentiable function ƒ from Rn to R for which there exists a vector valued
function g such that

for all x and u.


Invex functions were introduced by Hanson [1] as a generalization of convex functions. Ben-Israel and Mond [2]

provided a simple proof that a function is invex if and only if every stationary point is a global minimum.
Hanson also showed that if the objective and the constraints of an optimization problem are invex with respect to the
same function g(x, u), then the Karush–Kuhn–Tucker conditions are sufficient for a global minimum.
A slight generalization of invex functions called Type 1 invex functions are the most general class of functions for
which the Karush–Kuhn–Tucker conditions are necessary and sufficient for a global minimum [3] .

References
[1] M.A. Hanson, On sufficiency of the Kuhn–Tucker conditions, J. Math. Anal. Appl. 80, pp. 545–550 (1981)
[2] Ben-Israel, A. and Mond, B., What is invexity?, The ANZIAM Journal 28, pp. 1–9 (1986)
[3] M.A. Hanson, Invexity and the Kuhn-Tucker Theorem, J. Math. Anal. Appl. vol. 236, pp. 594–604 (1999)

Further reading
S. K. Mishra and G. Giorgi, Invexity and optimization, Nonconvex optimization and Its Applications, Vol. 88,
Springer-Verlag, Berlin, 2008.
List of types of functions 86

List of types of functions


Functions can be classified according to the properties they have. These properties describe the functions behaviour
under certain conditions.

Relative to set theory


These properties concern the domain, the codomain and the range of functions.
• Bijective function: is both an injective and a surjection, and thus invertible.
• Composite function: is formed by the composition of two functions f and g, by mapping x to f(g(x)).
• Constant function: has a fixed value regardless of arguments.
• Empty function: whose domain equals the empty set.
• Inverse function: is declared by "doing the reverse" of a given function (e.g. arcsine is the inverse of sine).
• Injective function: has a distinct value for each distinct argument. Also called an injection or, sometimes,
one-to-one function.
• Surjective function: has a preimage for every element of the codomain, i.e. the codomain equals the range. Also
called a surjection or onto function.
• Identity function: maps any given element to itself.
• Piecewise function: is defined by different expressions at different intervals.

Relative to an operator (c.q. a group)


These properties concern how the function is affected by arithmetic operations on its operand.
• Additive function: preserves the addition operation: f(x+y) = f(x)+f(y).
• Multiplicative function: preserves the multiplication operation: f(xy) = f(x)f(y).
• Even function: is symmetric with respect to the Y-axis. Formally, for each x: f(x) = f(−x).
• Odd function: is symmetric with respect to the origin. Formally, for each x: f(−x) = −f(x).
• Subadditive function: for which the value of f(x+y) is less than or equal to f(x)+f(y).
• Superadditive function: for which the value of f(x+y) is greater than or equal to f(x)+f(y).

Relative to a topology
• Continuous function: in which preimages of open sets are open.
• Nowhere continuous function: is not continuous at any point of its domain (e.g. Dirichlet function).
• Homeomorphism: is an injective function that is also continuous, whose inverse is continuous.

Relative to an ordering
• Monotonic function: does not reverse ordering of any pair.
• Strict Monotonic function: preserves the given order.
List of types of functions 87

Relative to the real/complex numbers


• Analytic function: Can be defined locally by a convergent power series.
• Arithmetic function: A function from the positive integers into the complex numbers.
• Differentiable function: Has a derivative.
• Smooth function: Has derivatives of all orders.
• Holomorphic function: Complex valued function of a complex variable which is differentiable at every point in its
domain.
• Meromorphic function: Complex valued function that is holomorphic everywhere, apart from at isolated points
where there are poles.
• Entire function: A holomorphic function whose domain is the entire complex plane.

Locally integrable function


In mathematics, a locally integrable function is a function which is integrable on any compact set of its domain of
definition. Their importance lies on the fact that we do not care about their behavior at infinity.

Formal definition
Formally, let be an open set in the Euclidean space ℝn and be a Lebesgue measurable function. If the
Lebesgue integral of is such that

i.e. it is finite for all compact subsets in , then is called locally integrable. The set of all such functions is
denoted by :

where is the set of all compact subsets of the set .

Properties
Theorem. Every function belonging to , , where is an open subset of ℝn is locally
integrable. To see this, consider the characteristic function of a compact subset of : then, for

where
• is the positive number such that for a given
• is the Lebesgue measure of the compact set
Then by Hölder's inequality, the product is integrable i.e. belongs to and

therefore

Note that since the following inequality is true


Locally integrable function 88

the theorem is true also for functions belonging only to for each compact subset of .

Examples
• The constant function defined on the real line is locally integrable but not globally integrable. More generally,
continuous functions and constants are locally integrable.
• The function

is not locally integrable near .

Applications
Locally integrable functions play a prominent role in distribution theory. Also they occur in the definition of various
classes of functions and function spaces, like functions of bounded variation.

See also
• Compact set
• Distribution (mathematics)
• Lebesgue integral
• Lebesgue measure
• Lp(Ω) space

References
• Saks, Stanisław (1937), Theory of the Integral [1], Monografie Matematyczne [2], 7 (2nd ed.), Warszawa-Lwów:
G.E. Stechert & Co., pp. VI+347, Zbl: 63.0183.05, Zbl: 0017.30004. English translation by Laurence Chisholm
Young, with two additional notes by Stefan Banach.
• Strichartz, Robert S. (2003), A Guide to Distribution Theory and Fourier Transforms [3] (2nd ed.), River Edge, NJ:
World Scientific Publishers, Zbl: 1029.46039, ISBN 981-238-430-8

External links
• Rowland, Todd, "Locally integrable [4]" from MathWorld.
• Vinogradova, I.A. (2001), "Locally integrable function" [5], in Hazewinkel, Michiel, Encyclopaedia of
Mathematics, Springer, ISBN 978-1556080104
This article incorporates material from Locally integrable function on PlanetMath, which is licensed under the
Creative Commons Attribution/Share-Alike License.
Locally integrable function 89

References
[1] http:/ / matwbn. icm. edu. pl/ kstresc. php?tom=7& wyd=10& jez=pl
[2] http:/ / matwbn. icm. edu. pl/ ksspis. php?wyd=10& jez=pl
[3] http:/ / books. google. it/ books?id=T7vEOGGDCh4C& printsec=frontcover& dq=A+ Guide+ to+ Distribution+ Theory+ and+ Fourier+
Transforms#v=onepage& q=& f=false
[4] http:/ / mathworld. wolfram. com/ LocallyIntegrable. html
[5] http:/ / eom. springer. de/ L/ l060460. htm

Measurable function
In mathematics, particularly in measure theory, measurable functions are structure-preserving functions between
measurable spaces; as such, they form a natural context for the theory of integration. Specifically, a function between
measurable spaces is said to be measurable if the preimage of each measurable set is measurable, analogous to the
situation of continuous functions between topological spaces.
This definition can be deceptively simple, however, as special care must be taken regarding the -algebras
involved. In particular, when a function is said to be Lebesgue measurable what is actually meant is
that is a measurable function -- that is, the domain and range represent different
-algebras on the same underlying set (here is the sigma algebra of Lebesgue measurable sets, and is the Borel
algebra on ). As a result, the composition of Lebesgue-measurable functions need not be Lebesgue-measurable.
By convention a topological space is assumed to be equipped with the Borel algebra generated by its open subsets
unless otherwise specified. Most commonly this space will be the real or complex numbers. For instance, a
real-valued measurable function is a function for which the preimage of each Borel set is measurable. A
complex-valued measurable function is defined analogously. In practice, some authors use measurable functions
to refer only to real-valued measurable functions with respect to the Borel algebra.[1] If the values of the function lie
in an infinite-dimensional vector space instead of R or C, usually other definitions of measurability are used, such as
weak measurability and Bochner measurability.
In probability theory, the sigma algebra often represents the set of available information, and a function (in this
context a random variable) is measurable if and only if it represents an outcome that is knowable based on the
available information. In contrast, functions that are not Lebesgue measurable are generally considered pathological,
at least in the field of analysis.

Formal definition
Let and be measurable spaces, meaning that and are sets equipped with respective sigma
algebras and . A function

is said to be measurable if for every . The notion of measurability depends on the sigma
algebras and . To emphasize this dependency, if is a measurable function, we will write
Measurable function 90

Special measurable functions


• If and are Borel spaces, a measurable function is also called a Borel
function. Continuous functions are Borel functions but not all Borel functions are continuous. However, a
measurable function is nearly a continuous function; see Luzin's theorem. If a Borel function happens to be a
section of some map , it is called a Borel section.
• A Lebesgue measurable function is a measurable function , where is the sigma
algebra of Lebesgue measurable sets, and is the Borel algebra on the complex numbers . Lebesgue
measurable functions are of interest in mathematical analysis because they can be integrated.
• Random variables are by definition measurable functions defined on sample spaces.

Properties of measurable functions


• The sum and product of two complex-valued measurable functions are measurable.[2] . So is the quotient, so long
as there is no division by zero.[1]
• The composition of measurable functions is measurable; i.e, if and
are measurable functions, then so is .[1] But see the
caveat regarding Lebesgue-measurable functions in the introduction.
• The (pointwise) supremum, infimum, limit superior, and limit inferior of a sequence (viz., countably many) of
real-valued measurable functions are all measurable as well.[1] [3]
• The pointwise limit of a sequence of measurable functions is measurable; note that the corresponding statement
for continuous functions requires stronger conditions than pointwise convergence, such as uniform convergence.
(This is correct when the counter domain of the elements of the sequence is a metric space. It is false in general;
see pages 125 and 126 of [4] .)

Non-measurable functions
Real-valued functions encountered in applications tend to be measurable; however, it is not difficult to find
non-measurable functions.
• So long as there are non-measurable sets in a measure space, there are non-measurable functions from that space.
If is some measurable space and is a non-measurable set, i.e. if , then the indicator
function is non-measurable (where is equipped with the Borel algebra as usual), since
the preimage of the measurable set is the non-measurable set . Here is given by

• Any non-constant function can be made non-measurable by equipping the domain and range with appropriate
-algebras. If is an arbitrary non-constant, real-valued function, then is non-measurable if is
equipped with the indiscrete algebra , since the preimage of any point in the range is some proper,
nonempty subset of , and therefore does not lie in .
Measurable function 91

Notes
[1] Strichartz, Robert (2000). The Way of Analysis. Jones and Bartlett. ISBN 0-7637-1497-6.
[2] Folland, Gerald B. (1999). Real Analysis: Modern Techniques and their Applications. Wiley. ISBN 0471317160.
[3] Royden, H. L. (1988). Real Analysis. Prentice Hall. ISBN 0-02-404151-3.
[4] Dudley, R. M. (2002). Real Analysis and Probability (2 ed.). Cambridge University Press. ISBN 0521007542.

Meromorphic function
In complex analysis, a meromorphic function on an open subset D of the complex plane is a function that is
holomorphic on all D except a set of isolated points, which are poles for the function. (The terminology comes from
the Ancient Greek meros (μέρος), meaning part, as opposed to holos (ὅλος), meaning whole.)
Every meromorphic function on D can be expressed as the ratio between two holomorphic functions (with the
denominator not constant 0) defined on D: any pole must coincide with a zero of the denominator.
Intuitively then, a meromorphic function is a ratio of two well-behaved
(holomorphic) functions. Such a function will still be well-behaved,
except possibly at the points where the denominator of the fraction is
zero. (If the denominator has a zero at z and the numerator does not,
then the value of the function will be infinite; if both parts have a zero
at z, then one must compare the multiplicities of these zeros.)

From an algebraic point of view, if D is connected, then the set of


meromorphic functions is the field of fractions of the integral domain
of the set of holomorphic functions. This is analogous to the
The Gamma function is meromorphic in the
relationship between , the rational numbers, and , the integers.
whole complex plane.
Additionally, in group theory of the 1930s, a meromorphic function (or
simply a meromorph) was a function from a group G into itself which preserves the product on the group. The image
of this function was called an automorphism of G [1] . (Similarly, a homomorphic function (or homomorph) was a
function between groups which preserved the product while a homomorphism was the image of a homomorph.) This
terminology has been replaced with use of endomorphism for the function itself with no special name given to the
image of the function and thus meromorph no longer has an implied meaning within group theory.

Examples
• All rational functions such as

are meromorphic on the whole complex plane.


• The functions

as well as the gamma function and the Riemann zeta function are meromorphic on the whole complex plane.
• The function

is defined in the whole complex plane except for the origin, 0. However, 0 is not a pole of this function, rather
an essential singularity. Thus, this function is not meromorphic in the whole complex plane. However, it is
Meromorphic function 92

meromorphic (even holomorphic) on .


• The complex logarithm function

is not meromorphic on the whole complex plane, as it cannot be defined on the whole complex plane less an
isolated set of points.
• The function

is not meromorphic in the whole plane, since the point is an accumulation point of poles and is thus
not an isolated singularity. The function

is not meromorphic either, as it has an essential singularity at 0.

Properties
Since the poles of a meromorphic function are isolated, there are at most countably many. The set of poles can be
infinite, as exemplified by the function

By using analytic continuation to eliminate removable singularities, meromorphic functions can be added,
subtracted, multiplied, and the quotient can be formed unless on a connected component of D.
Thus, if D is connected, the meromorphic functions form a field, in fact a field extension of the complex numbers.

Meromorphic functions on Riemann surfaces


On a Riemann surface every point admits an open neighborhood which is homeomorphic to an open subset of the
complex plane. Thereby the notion of a meromorphic function can be defined for every Riemann surface.
When D is the entire Riemann sphere, the field of meromorphic functions is simply the field of rational functions in
one variable over the complex field, since one can prove that any meromorphic function on the sphere is rational.
(This is a special case of the so-called GAGA principle.)
For every Riemann surface, a meromorphic function is the same as a holomorphic function that maps to the Riemann
sphere and which is not constant ∞. The poles correspond to those complex numbers which are mapped to ∞.
On a non-compact Riemann surface every meromorphic function can be realized as a quotient of two (globally
defined) holomorphic functions. In contrast, on a compact Riemann surface every holomorphic function is constant,
while there always exist non-constant meromorphic functions.
Meromorphic functions on an elliptic curve are also known as elliptic functions.

Higher dimensions
In several complex variables, a meromorphic function is defined to be locally a quotient of two holomorphic
functions. For example, is a meromorphic function on the two-dimensional complex affine
space. Here it is no longer true that every meromorphic function can be regarded as holomorphic function with
values in the Riemann sphere: There is a set of "indeterminacy" of codimension two (in the given example this set
consists of the origin ).
Meromorphic function 93

Unlike in dimension one, in higher dimensions there do exist complex manifolds on which there are no non-constant
meromorphic functions, for example, most complex tori.

References
[1] Zassenhaus pp. 29, 41

• Lang, Serge (1999), Complex analysis (4th ed.), Berlin, New York: Springer-Verlag, ISBN 978-0-387-98592-3
• Zassenhaus, Hans (1937), Lehrbuch der Gruppentheorie (1st ed.), Leipzig, Berlin: Verlag und Druck von
B.G.Teubner

Monotonic function
In mathematics, a monotonic function (or monotone function) is a
function which preserves the given order. This concept first arose in
calculus, and was later generalized to the more abstract setting of order
theory.

Monotonicity in calculus and analysis


In calculus, a function f defined on a subset of the real numbers with
real values is called monotonic (also monotonically increasing,
increasing or non-decreasing), if for all x and y such that x ≤ y one
has f(x) ≤ f(y), so f preserves the order (see Figure 1). Likewise, a
function is called monotonically decreasing (also decreasing, or Figure 1. A monotonically increasing function (it
non-increasing) if, whenever x ≤ y, then f(x) ≥ f(y), so it reverses the is strictly increasing on the left and right while
just non-decreasing in the middle.)
order (see Figure 2).

If the order ≤ in the definition of monotonicity is replaced by the strict


order <, then one obtains a stronger requirement. A function with this
property is called strictly increasing. Again, by inverting the order
symbol, one finds a corresponding concept called strictly decreasing.
Functions that are strictly increasing or decreasing are one-to-one
(because for x not equal to y, either x < y or x > y and so, by
monotonicity, either f(x) < f(y) or f(x) > f(y), thus f(x) is not equal to
f(y)).

The terms "non-decreasing" and "non-increasing" are meant to avoid


confusion with "strictly increasing" respectively "strictly decreasing",
but should not be confused with the (much weaker) negative
qualifications "not decreasing" and "not increasing"; see also strict. Figure 2. A monotonically decreasing function.
When functions between discrete sets are considered in combinatorics,
it is not always obvious that "increasing" and "decreasing" are taken to include the possibility of repeating the same
value at successive arguments, so one finds the terms weakly increasing and weakly decreasing to stress this
possibility.
Monotonic function 94

The term monotonic transformation can also possibly cause some


confusion because it refers to a transformation by a strictly increasing
function. Notably, this is the case in Economics with respect to the
ordinal properties of a utility function being preserved across a
monotonic transform (see also monotone preferences).[1]
A function f(x) is said to be absolutely monotonic over an interval (a,
b) if the derivatives of all orders of f are nonnegative at all points on
the interval.

Some basic applications and results


The following properties are true for a monotonic function f : R → R: Figure 3. A function that is not monotonic.

• f has limits from the right and from the left at every point of its
domain;
• f has a limit at infinity (either ∞ or −∞) of either a real number, ∞, or −∞.
• f can only have jump discontinuities;
• f can only have countably many discontinuities in its domain.
These properties are the reason why monotonic functions are useful in technical work in analysis. Two facts about
these functions are:
• if f is a monotonic function defined on an interval I, then f is differentiable almost everywhere on I, i.e. the set of
numbers x in I such that f is not differentiable in x has Lebesgue measure zero.
• if f is a monotonic function defined on an interval [a, b], then f is Riemann integrable.
An important application of monotonic functions is in probability theory. If X is a random variable, its cumulative
distribution function
FX(x) = Prob(X ≤ x)
is a monotonically increasing function.
A function is unimodal if it is monotonically increasing up to some point (the mode) and then monotonically
decreasing.

Monotonicity in functional analysis


In functional analysis on a topological vector space X, a (possibly non-linear) operator T : X → X∗ is said to be a
monotone operator if

Kachurovskii's theorem shows that convex functions on Banach spaces have monotonic operators as their
derivatives.
A subset G of X × X∗ is said to be a monotone set if for every pair [u1,w1] and [u2,w2] in G,

G is said to be maximal monotone if it is maximal among all monotone sets in the sense of set inclusion. The graph
of a monotone operator G(T) is a monotone set. A monotone operator is said to be maximal monotone if its graph is
a maximal monotone set.
Monotonic function 95

Monotonicity in order theory


In order theory, one does not restrict to real numbers, but one is concerned with arbitrary partially ordered sets or
even with preordered sets. In these cases, the above definition of monotonicity is relevant as well. However, the
terms "increasing" and "decreasing" are avoided, since they lose their appealing pictorial motivation as soon as one
deals with orders that are not total. Furthermore, the strict relations < and > are of little use in many non-total orders
and hence no additional terminology is introduced for them.
A monotone function is also called isotone, or order-preserving. The dual notion is often called antitone,
anti-monotone, or order-reversing. Hence, an antitone function f satisfies the property
x ≤ y implies f(x) ≥ f(y),
for all x and y in its domain. It is easy to see that the composite of two monotone mappings is also monotone.
A constant function is both monotone and antitone; conversely, if f is both monotone and antitone, and if the domain
of f is a lattice, then f must be constant.
Monotone functions are central in order theory. They appear in most articles on the subject and examples from
special applications are to be found in these places. Some notable special monotone functions are order embeddings
(functions for which x ≤ y if and only if f(x) ≤ f(y)) and order isomorphisms (surjective order embeddings).

Boolean functions
In Boolean algebra, a monotonic
function is one such that for all ai and
bi in {0,1} such that a1 ≤ b1, a2 ≤ b2, ...
, an ≤ bn
it is true that
f(a1, ... , an) ≤ f(b1, ... , bn).
More human-readable that means, a
Boolean function is monotonic if, for
every combination of inputs, switching
one of the inputs from false to true can
only cause the output to switch from
false to true and not from true to false.
Graphically that means, a Boolean
function is monotonic, when in its
Hasse diagram (dual of its Venn
diagram) there is no 1 (red vertex)
The free distributive lattices of monotonic Boolean functions on 0, 1, 2, and 3 arguments
connected to a higher 0 (white vertex).
(move mouse over right diagram to see description)
The monotonic Boolean functions are
precisely those which can be defined as a composition of ands and ors, but no nots.
The number of such functions on n variables is known as the Dedekind number of n.
Monotonic function 96

Monotonic logic
Monotonicity of entailment is a property of many logic systems that states that the hypotheses of any derived fact
may be freely extended with additional assumptions. Any true statement in a logic with this property continues to be
true, even after adding new axioms. Logics with this property may be called monotonic, to differentiate them from
non-monotonic logic.

See also
• Monotone cubic interpolation
• Pseudo-monotone operator
• Total monotonicity

Notes
[1] See the section on Cardinal Versus Ordinal Utility in (Simon and Blume, 1994).

Bibliography
• Bartle, Robert G. (1976). The elements of real analysis (second edition ed.).
• Grätzer, George (1971). Lattice theory: first concepts and distributive lattices. ISBN 0716704420.
• Pemberton, Malcolm; Rau, Nicholas (2001). Mathematics for economists: an introductory textbook. Manchester
University Press. ISBN 0719033411.
• Renardy, Michael and Rogers, Robert C. (2004). An introduction to partial differential equations. Texts in
Applied Mathematics 13 (Second edition ed.). New York: Springer-Verlag. pp. 356. ISBN 0-387-00444-0.
• Riesz, Frigyes and Béla Szőkefalvi-Nagy (1990). Functional Analysis. Courier Dover Publications.
ISBN 9780486662893.
• Simon, Carl P. and Lawrence Blume (April 1994). Mathematics for Economists (first edition ed.).
ISBN 978-0393957334. (Definition 9.31)

External links
• Convergence of a Monotonic Sequence (http://demonstrations.wolfram.com/
ConvergenceOfAMonotonicSequence/) by Anik Debnath and Thomas Roxlo (The Harker School), Wolfram
Demonstrations Project.
Multiplicative function 97

Multiplicative function
Outside number theory, the term multiplicative function is usually used for completely multiplicative
functions. This article discusses number theoretic multiplicative functions.
In number theory, a multiplicative function is an arithmetic function f(n) of the positive integer n with the property
that f(1) = 1 and whenever a and b are coprime, then
f(ab) = f(a) f(b).
An arithmetic function f(n) is said to be completely multiplicative (or totally multiplicative) if f(1) = 1 and f(ab) =
f(a) f(b) holds for all positive integers a and b, even when they are not coprime.

Examples
Examples of multiplicative functions include many functions of importance in number theory, such as:
• (n): Euler's totient function , counting the positive integers coprime to (but not bigger than) n
• (n): the Möbius function, related to the number of prime factors of square-free numbers
• gcd(n,k): the greatest common divisor of n and k, where k is a fixed integer.
• d(n): the number of positive divisors of n,
• (n): the sum of all the positive divisors of n,
• (n): the divisor function, which is the sum of the k-th powers of all the positive divisors of n (where k may be
k
any complex number). In special cases we have
• (n) = d(n) and
0
• (n) = (n),
1
• : the number of non-isomorphic abelian groups of order n.
• 1(n): the constant function, defined by 1(n) = 1 (completely multiplicative)
• the indicator function of the set of squares (or cubes, or fourth powers, etc.)
• Id(n): identity function, defined by Id(n) = n (completely multiplicative)
• Idk(n): the power functions, defined by Idk(n) = nk for any natural (or even complex) number k (completely
multiplicative). As special cases we have
• Id0(n) = 1(n) and
• Id1(n) = Id(n),
• (n): the function defined by (n) = 1 if n = 1 and = 0 otherwise, sometimes called multiplication unit for
Dirichlet convolution or simply the unit function; sometimes written as u(n), not to be confused with (n)
(completely multiplicative).
• (n/p), the Legendre symbol, where p is a fixed prime number (completely multiplicative).
• (n): the Liouville function, related to the number of prime factors dividing n (completely multiplicative).
• (n), defined by (n)=(-1) (n), where the additive function (n) is the number of distinct primes dividing
n.
• All Dirichlet characters are completely multiplicative functions.
An example of a non-multiplicative function is the arithmetic function r2(n) - the number of representations of n as a
sum of squares of two integers, positive, negative, or zero, where in counting the number of ways, reversal of order
is allowed. For example:
1 = 12 + 02 = (-1)2 + 02 = 02 + 12 = 02 + (-1)2
and therefore r2(1) = 4 ≠ 1. This shows that the function is not multiplicative. However, r2(n)/4 is multiplicative.
[1]
In the On-Line Encyclopedia of Integer Sequences, sequences of values of a multiplicative function have the
keyword "mult".
Multiplicative function 98

See arithmetic function for some other examples of non-multiplicative functions.

Properties
A multiplicative function is completely determined by its values at the powers of prime numbers, a consequence of
the fundamental theorem of arithmetic. Thus, if n is a product of powers of distinct primes, say n = pa qb ..., then f(n)
= f(pa) f(qb) ...
This property of multiplicative functions significantly reduces the need for computation, as in the following
examples for n = 144 = 24 · 32:
d(144) = (144) = (24) (32) = (10 + 20 + 40 + 80 + 160)(10 + 30 + 90) = 5 · 3 = 15,
0 0 0
(144) = (144) = (24) (32) = (11 + 21 + 41 + 81 + 161)(11 + 31 + 91) = 31 · 13 = 403,
1 1 1
* *
(24) * 2
(3 ) = (1 + 161)(11 + 91) = 17 · 10 = 170.
1
(144) =
Similarly, we have:
(144)= (24) (32) = 8 · 6 = 48
In general, if f(n) is a multiplicative function and a, b are any two positive integers, then
f(a) · f(b) = f(gcd(a,b)) · f(lcm(a,b)).
Every completely multiplicative function is a homomorphism of monoids and is completely determined by its
restriction to the prime numbers.

Convolution
If f and g are two multiplicative functions, one defines a new multiplicative function f * g, the Dirichlet convolution
of f and g, by

where the sum extends over all positive divisors d of n. With this operation, the set of all multiplicative functions
turns into an abelian group; the identity element is .
Relations among the multiplicative functions discussed above include:
• * 1 = (the Möbius inversion formula)
• ( Idk) * Idk = (generalized Möbius inversion)
• * 1 = Id
• d=1*1
• = Id * 1 = *d
• = Idk * 1
k
• Id = *1= *
• Idk = *
k
The Dirichlet convolution can be defined for general arithmetic functions, and yields a ring structure, the Dirichlet
ring.
Multiplicative function 99

Dirichlet series for some multiplicative functions

References
• See chapter 2 of Apostol, Tom M. (1976), Introduction to analytic number theory, Undergraduate Texts in
Mathematics, New York-Heidelberg: Springer-Verlag, MR0434929, ISBN 978-0-387-90163-3

External links
• Planet Math [2]

References
[1] http:/ / www. research. att. com/ ~njas/ sequences/ ?q=keyword:mult
[2] http:/ / planetmath. org/ encyclopedia/ MultiplicativeFunction. html

Multivalued function
In mathematics, a multivalued function (shortly: multifunction, other
names: set-valued function, set-valued map, multi-valued map,
multimap, correspondence, carrier) is a left-total relation; i.e. every
input is associated with one or more outputs. Strictly speaking, a
"well-defined" function associates one, and only one, output to any
particular input. The term "multivalued function" is, therefore, a
misnomer since functions are single-valued. Multivalued functions
often arise from functions which are not injective. Such functions do
not have an inverse function, but they do have an inverse relation. The
multivalued function corresponds to this inverse relation.

Examples This diagram does not represent a "true" function,


because the element 3 in X is associated with two
• Every real number greater than zero or every complex number elements, b and c, in Y.

except 0 has two square roots. The square roots of 4 are in the set
{+2,−2}. The square roots of 0 are described by the multiset {0,0}, because 0 is a root of multiplicity 2 of the
polynomial x².
• Each complex number has three cube roots.
• The complex logarithm function is multiple-valued. The values assumed by log(1) are for all integers .
• Inverse trigonometric functions are multiple-valued because trigonometric functions are periodic. We have
Multivalued function 100

Consequently arctan(1) is intuitively related to several values: π/4, 5π/4, −3π/4, and so on. We can treat arctan
as a single-valued function by restricting the domain of tan x to -π/2 < x < π/2 – a domain over which tan x is
monotonically increasing. Thus, the range of arctan(x) becomes -π/2 < x < π/2. These values from a restricted
domain are called principal values.
• The indefinite integral is a multivalued function of real-valued functions. The indefinite integral of a function is
the set of functions whose derivative is that function. The constant of integration comes follows from the fact that
the difference between any two indefinite integrals is a constant,
These are all examples of multivalued functions which come about from non-injective functions. Since the original
functions do not preserve all the information of their inputs, they are not reversible. Often, the restriction of a
multivalued function is a partial inverse of the original function.
Multivalued functions of a complex variable have branch points. For example the nth root and logarithm functions, 0
is a branch point; for the arctangent function, the imaginary units i and −i are branch points. Using the branch points
these functions may be redefined to be single valued functions, by restricting the range. A suitable interval may be
found through use of a branch cut, a kind of curve which connects pairs of branch points, thus reducing the
multilayered Riemann surface of the function to a single layer. As in the case with real functions the restricted range
may be called principal branch of the function.

Riemann surfaces
A more sophisticated viewpoint replaces "multivalued functions" with functions whose domain is a Riemann surface
(so named in honor of Bernhard Riemann).

Set-valued analysis
Set-valued analysis is the generalization, to set-valued functions, of ideas from mathematical analysis and topology
such as continuity, differentiation, integration, implicit function theorem, contraction mappings, measure theory,
fixed-point theorems, optimization, topological degree theory. Equations and inequalities can be generalized to
intervals and then to inclusions.

Types of multivalued functions


One can differentiate many continuity concepts, primarily closed graph property and upper and lower
hemicontinuity. (One should be warned that often the terms upper and lower semicontinuous are used instead of
upper and lower hemicontinuous reserved for the case of weak topology in domain; yet we arrive at the collision
with the reserved names for upper and lower semicontinuous real-valued function). There exist also various
definitions for measurability of multifunction.
Multivalued function 101

History
The practice of allowing function in mathematics to mean also multivalued function dropped out of usage at some
point in the first half of the twentieth century. Some evolution can be seen in different editions of A Course of Pure
Mathematics by G. H. Hardy, for example. It probably persisted longest in the theory of special functions, for its
occasional convenience.
The theory of multivalued functions was fairly systematically developed for the first time in C. Berge "Topological
spaces" 1963 [1].

Applications
Multifunctions arise in optimal control theory, especially differential inclusions and related subjects as game theory,
where the Kakutani fixed point theorem for multifunctions has been applied to prove existence of Nash equilibria.
This amongst many other properties loosely associated with approximability of upper hemicontinuous multifunctions
via continuous functions explains why upper hemicontinuity is more preferred than lower hemicontinuity.
Nevertheless, lower hemicontinuous multifunctions usually possess continuous selections as stated in the Michael
selection theorem which provides another characterisation of paracompact spaces (see: E. Michael, Continuous
selections I" Ann. of Math. (2) 63 (1956), and D. Repovs, P.V. Semenov, Ernest Michael and theory of continuous
selections" arXiv:0803.4473v1). Other selection theorems, like Bressan-Colombo directional continuous selection,
Kuratowski—Ryll-Nardzewski measurable selection, Aumann measurable selection, Fryszkowski selection for
decomposable maps are important in optimal control and the theory of differential inclusions.
In physics, multivalued functions play an increasingly important role. They form the mathematical basis for Dirac's
magnetic monopoles, for the theory of defects in crystal and the resulting plasticity of materials, for vortices in
superfluids and superconductors, and for phase transitions in these systems, for instance melting and quark
confinement. They are the origin of gauge field structures in many branches of physics.

References
• Jean-Pierre Aubin, Arrigo Cellina Differential Inclusions, Set-Valued Maps And Viability Theory, Grundl. der
Math. Wiss., vol. 264, Springer - Verlag, Berlin, 1984
• J.-P. Aubin and H. Frankowska Set-Valued Analysis, Birkhäuser, Basel, 1990
• Klaus Deimling Multivalued Differential Equations, Walter de Gruyter, 1992
• Kleinert, Hagen, Multivalued Fields in in Condensed Matter, Electrodynamics, and Gravitation, World Scientific
(Singapore, 2008) [2] (also available online [3])
• Kleinert, Hagen, Gauge Fields in Condensed Matter, Vol. I, "SUPERFLOW AND VORTEX LINES", pp.
1—742, Vol. II, "STRESSES AND DEFECTS", pp. 743-1456, World Scientific (Singapore, 1989) [4]; Paperback
ISBN 9971-5-0210-0 (also available online: Vol. I [5] and Vol. II [6])
• Aliprantis, Kim C. Border Infinite dimensional analysis. Hitchhiker's guide Springer
• J. Andres, L. Górniewicz Topological Fixed Point Principles for Boundary Value Problems, Kluwer Academic
Publishers, 2003
• Topological methods for set-valued nonlinear analysis [7], Enayet U. Tarafdar, Mohammad Showkat Rahim
Chowdhury, World Scientific, 2008, ISBN 9789812704672
Multivalued function 102

References
[1] http:/ / www. gbv. de/ dms/ goettingen/ 229406319. pdf
[2] http:/ / www. worldscibooks. com/ physics/ 6742. html
[3] http:/ / www. physik. fu-berlin. de/ ~kleinert/ re. html#B9
[4] http:/ / www. worldscibooks. com/ physics/ 0356. htm
[5] http:/ / www. physik. fu-berlin. de/ ~kleinert/ kleiner_reb1/ contents1. html
[6] http:/ / www. physik. fu-berlin. de/ ~kleinert/ kleiner_reb1/ contents2. html
[7] http:/ / books. google. co. uk/ books?id=Cir88lF64xIC

Negligible function
In mathematics, a negligible function is a function such that for every positive integer c there exists
an integer Nc such that for all x > Nc,

Equivalently, we may also use the following definition. A function is negligible, if for every
positive polynomial poly(·) there exists an integer Npoly > 0 such that for all x > Npoly

History
The concept of negligibility can find its trace back to sound models of analysis. Though the concepts of "continuity"
and "infinitesimal" became important in mathematics during Newton and Leibniz's time (1680s), they were not
well-defined until late 1810s. The first reasonably rigorous definition of continuity in mathematical analysis was due
to Bernard Bolzano, who wrote in 1817 the modern definition of continuity. Lately Cauchy, Weierstrass and Heine
also defined as follows (with all numbers in the real number domain ):
(Continuous function) A function is continuous at if for every , there exists
a positive number such that implies
This classic definition of continuity can be transformed into the definition of negligibility in a few steps by changing
a parameter used in the definition per step. First, in case with , we must define the concept
of "infinitesimal function":
(Infinitesimal) A continuous function is infinitesimal (as goes to infinity) if for every
there exists such that for all

Next, we replace by the functions where or by where is a positive


polynomial. This leads to the definitions of negligible functions given at the top of this article. Since the constants
can be expressed as with a constant polynomial this shows that negligible functions are
a subset of the infinitesimal functions.
In complexity-based modern cryptography, a security scheme is provably secure if the probability of security failure
(e.g., inverting a one-way function, distinguishing cryptographically strong pseudorandom bits from truly random
bits) is negligible in terms of the input = cryptographic key length . Hence comes the definition at the top of
the page because key length must be a natural number.
Nevertheless, the general notion of negligibility has never said that the system input parameter must be the key
length . Indeed, can be any predetermined system metric and corresponding mathematic analysis would
illustrate some hidden analytical behaviors of the system.
Negligible function 103

The reciprocal-of-polynomial formulation is used for the same reason that computational boundedness is defined as
polynomial running time: it has mathematical closure properties that make it tractable in the asymptotic setting. For
example, if an attack succeeds in violating a security condition only with negligible probability, and the attack is
repeated a polynomial number of times, the success probability of the overall attack still remains negligible. In
practice one might want to have more concrete functions bounding the adversary's success probability and to choose
the security parameter large enough that this probability is smaller than some threshold, say 2−128.

References
• Goldreich, Oded (2001). Foundations of Cryptography: Volume 1, Basic Tools. Cambridge University Press.
ISBN 0-521-79172-3. Fragments available at the author's web site [1].
• Michael Sipser (1997). Introduction to the Theory of Computation. PWS Publishing. ISBN 0-534-94728-X.
Section 10.6.3: One-way functions, pp.374–376.
• Christos Papadimitriou (1993). Computational Complexity (1st edition ed.). Addison Wesley.
ISBN 0-201-53082-1. Section 12.1: One-way functions, pp.279–298.
• Jean François Colombeau (1984). New Generalized Functions and Multiplication of Distributions. Mathematics
Studies 84, North Holland. ISBN 0-444-86830-5.

References
[1] http:/ / www. wisdom. weizmann. ac. il/ ~oded/ frag. html

Nowhere continuous function


In mathematics, a nowhere continuous function, also called an everywhere discontinuous function, is a function
that is not continuous at any point of its domain. If f is a function from real numbers to real numbers, then f(x) is
nowhere continuous if for each point x there is an ε > 0 such that for each δ > 0 we can find a point y such that
|x − y| < δ and |f(x) − f(y)| ≥ ε. Therefore, no matter how close we get to any fixed point, there are even closer points
at which the function takes not-nearby values.
More general definitions of this kind of function can be obtained, by replacing the absolute value by the distance
function in a metric space, or by using the definition of continuity in a topological space.
One example of such a function is the indicator function of the rational numbers, also known as the Dirichlet
function, named after German mathematician Peter Gustav Lejeune Dirichlet. This function is written IQ and has
domain and codomain both equal to the real numbers. IQ(x) equals 1 if x is a rational number and 0 if x is not
rational. If we look at this function in the vicinity of some number y, there are two cases:
• If y is rational, then f(y) = 1. To show the function is not continuous at y, we need to find an ε such that no matter
how small we choose δ, there will be points z within δ of y such that f(z) is not within ε of f(y) = 1. In fact, 1/2 is
such an ε. Because the irrational numbers are dense in the reals, no matter what δ we choose we can always find
an irrational z within δ of y, and f(z) = 0 is at least 1/2 away from 1.
• If y is irrational, then f(y) = 0. Again, we can take ε = 1/2, and this time, because the rational numbers are dense in
the reals, we can pick z to be a rational number as close to y as is required. Again, f(z) = 1 is more than 1/2 away
from f(y) = 0.
In plainer terms, between any two irrationals, there is a rational, and vice versa.
The Dirichlet function can be constructed as the double pointwise limit of a sequence of continuous functions, as
follows:
Nowhere continuous function 104

for integer j and k.


This shows that the Dirichlet function is a Baire class 2 function. It cannot be a Baire class 1 function because a
Baire class 1 function can only be discontinuous on a meagre set.[1]
In general, if E is any subset of a topological space X such that both E and the complement of E are dense in X, then
the real-valued function which takes the value 1 on E and 0 on the complement of E will be nowhere continuous.
Functions of this type were originally investigated by Dirichlet.

References
[1] Dunham, William (2005). The Calculus Gallery. Princeton University Press. pp. 197. ISBN 0-691-09565-5.

External links
• Dirichlet Function -- from MathWorld (http://mathworld.wolfram.com/DirichletFunction.html)
• The Modified Dirichlet Function (http://demonstrations.wolfram.com/TheModifiedDirichletFunction/) by
George Beck, The Wolfram Demonstrations Project.

Periodic function
In mathematics, a periodic function is a function that repeats its values in regular intervals or periods. The most
important examples are the trigonometric functions, which repeat over intervals of length 2π. Periodic functions are
used throughout science to describe oscillations, waves, and other phenomena that exhibit periodicity. Any function
which is not periodic is called aperiodic.

Definition
A function f is said to be periodic if (for
some nonzero constant P) we have

An illustration of a periodic function with period

for all values of x. The least positive[1] constant P with this property is called the period. A function with period P
will repeat on intervals of length P, and these intervals are sometimes also referred to as periods.
Geometrically, a periodic function can be defined as a function whose graph exhibits translational symmetry.
Specifically, a function f is periodic with period P if the graph of f is invariant under translation in the x-direction by
a distance of P. This definition of periodic can be extended to other geometric shapes and patterns, such as periodic
tessellations of the plane.
A function that is not periodic is called aperiodic.
Periodic function 105

Examples
For example, the sine function is
periodic with period 2π, since

A graph of the sine function, showing two complete periods.

for all values of x. This function repeats on intervals of length 2π (see the graph to the right).
Everyday examples are seen when the variable is time; for instance the hands of a clock or the phases of the moon
show periodic behaviour. Periodic motion is motion in which the position(s) of the system are expressible as
periodic functions, all with the same period.
For a function on the real numbers or on the integers, that means that the entire graph can be formed from copies of
one particular portion, repeated at regular intervals.
A simple example of a periodic function is the function f that gives the "fractional part" of its argument. Its period is
1. In particular,
f( 0.5 ) = f( 1.5 ) = f( 2.5 ) = ... = 0.5.
The graph of the function f is the sawtooth wave.
The trigonometric functions sine and cosine
are common periodic functions, with period
2π (see the figure on the right). The subject
of Fourier series investigates the idea that an
'arbitrary' periodic function is a sum of
trigonometric functions with matching
periods.

According to the definition above, some


exotic functions, for example the Dirichlet
function, are also periodic; in the case of
Dirichlet function, any nonzero rational
number is a period.
A plot of f(x) = sin(x) and g(x) = cos(x); both functions are periodic with period 2π.

Properties
If a function f is periodic with period P, then for all x in the domain of f and all integers n,
f(x + nP) = f(x).
If f(x) is a function with period P, then f(ax+b), where a is a positive constant, is periodic with period P/a. For
example, f(x)=sinx has period 2π, therefore sin(5x) will have period 2π/5.
Periodic function 106

Double-periodic functions
A function whose domain is the complex numbers can have two incommensurate periods without being constant.
The elliptic functions are such functions. ("Incommensurate" in this context means not real multiples of each other.)

Complex example
Using complex variables we have the common period function:

As you can see, since the cosine and sine functions are periodic, and the complex exponential above is made up of
cosine/sine waves, then the above (actually Euler's formula) has the following property. If L is the period of the
function then:

Generalizations

Antiperiodic functions
One common generalization of periodic functions is that of antiperiodic functions. This is a function f such that
f(x + P) = −f(x) for all x. (Thus, a P-antiperiodic function is a 2P-periodic function.)

Bloch-periodic functions
A further generalization appears in the context of Bloch waves and Floquet theory, which govern the solution of
various periodic differential equations. In this context, the solution (in one dimension) is typically a function of the
form:

where k is a real or complex number (the Bloch wavevector or Floquet exponent). Functions of this form are
sometimes called Bloch-periodic in this context. A periodic function is the special case k = 0, and an antiperiodic
function is the special case k = π/P.

Quotient spaces as domain


In signal processing you encounter the problem, that Fourier series represent periodic functions and that Fourier
series satisfy convolution theorems (i.e. convolution of Fourier series corresponds to multiplication of represented
periodic function and vice versa), but periodic functions cannot be convolved with the usual definition, since the
involved integrals diverge. A possible way out is to define a periodic function on a bounded but periodic domain. To
this end you can use the notion of a quotient space:
.
That is, each element in is an equivalence class of real numbers that share the same fractional part. Thus a
function like is a representation of a 1-periodic function.
Periodic function 107

References
[1] For some functions, like a constant function or the indicator function of the rational numbers, a least positive "period" may not exist (the
infimum of possible positve P being zero).

External links
• Periodic functions at MathWorld (http://mathworld.wolfram.com/PeriodicFunction.html)

Piecewise linear function


In mathematics, a piecewise linear function is a piecewise-defined
function whose pieces are linear.

Examples
The function defined by:

A piecewise linear function

A function (blue) and a piecewise linear


approximation to it (red).
Piecewise linear function 108

A piecewise linear function in two dimensions


(top) and the convex polytopes on which it is
linear (bottom).

is piecewise linear with four pieces. (The graph of this function is shown to the right.) Since the graph of a linear
function is a line, the graph of a piecewise linear function consists of line segments and rays. If the function is
continuous, the graph will be a polygonal curve.
Other examples of piecewise linear functions include the absolute value function, the square wave, the sawtooth
function, and the floor function.

Notation
The notion of a piecewise linear function makes sense in several different contexts. Piecewise linear functions may
be defined on n-dimensional Euclidean space, or more generally any vector space or affine space, as well as on
piecewise linear manifolds, simplicial complexes, and so forth. In each case, the function may be real-valued, or it
may take values from a vector space, an affine space, a piecewise-linear manifold, or a simplicial complex. (In these
contexts, the term “linear” does not refer solely to linear transformations, but to more general affine linear functions.)
In dimensions higher than one, it is common to require the domain of each piece to be a polygon or polytope. This
guarantees that the graph of the function will be composed of polygonal or polytopal pieces.
Important sub-classes of piecewise linear functions include the continuous piecewise linear functions and the convex
piecewise linear functions. Splines generalize piecewise linear functions to higher-order polynomials, and more one
can speak of piecewise-differentiable functions, as in PDIFF.

See also
• Linear interpolation
• Spline interpolation
Pluriharmonic function 109

Pluriharmonic function
Let

be a (twice continuously differentiable) function. is called pluriharmonic if for every complex line

the function

is a harmonic function on the set


.

Notes
Every pluriharmonic function is a harmonic function, but not the other way around. Further, it can be shown that for
holomorphic functions of several complex variables the real (and the imaginary) parts are locally pluriharmonic
functions. However a function being harmonic in each variable separately does not imply that it is pluriharmonic.

Bibliography
• Steven G. Krantz. Function Theory of Several Complex Variables, AMS Chelsea Publishing, Providence, Rhode
Island, 1992.

This article incorporates material from pluriharmonic function on PlanetMath, which is licensed under the Creative
Commons Attribution/Share-Alike License.
Plurisubharmonic function 110

Plurisubharmonic function
In mathematics, plurisubharmonic functions (sometimes abbreviated as psh, plsh, or plush functions) form an
important class of functions used in complex analysis. On a Kahler manifold, plurisubharmonic functions form a
subset of the subharmonic functions. However, unlike subharmonic functions (which are defined on a Riemannian
manifold) plurisubharmonic functions can be defined in full generality on complex spaces.

Formal definition
A function

with domain is called plurisubharmonic if it is upper semi-continuous, and for every complex line
with
the function is a subharmonic function on the set

In full generality, the notion can be defined on an arbitrary complex manifold or even a complex space as
follows. An upper semi-continuous function

is said to be plurisubharmonic if and only if for any holomorphic map the function

is subharmonic, where denotes the unit disk.

Differentiable plurisubharmonic functions


If is of (differentiability) class , then is plurisubharmonic, if and only if the hermitian matrix
, called Levi matrix, with entries

is positive semidefinite.
Equivalently, a -function f is plurisubharmonic if and only if is a positive (1,1)-form.

History
Plurisubharmonic functions were defined in 1942 by Kiyoshi Oka [1] and Pierre Lelong. [2]

Properties
• The set of plurisubharmonic functions form a convex cone in the vector space of semicontinuous functions, i.e.
• if is a plurisubharmonic function and a positive real number, then the function is
plurisubharmonic,
• if and are plurisubharmonic functions, then the sum is a plurisubharmonic function.
• Plurisubharmonicity is a local property, i.e. a function is plurisubharmonic if and only if it is plurisubharmonic in
a neighborhood of each point.
• If is plurisubharmonic and a monotonically increasing, convex function then is
plurisubharmonic.
Plurisubharmonic function 111

• If and are plurisubharmonic functions, then the function is


plurisubharmonic.
• If is a monotonically decreasing sequence of plurisubharmonic functions

then so is .
• Every continuous plurisubharmonic function can be obtained as a limit of monotonically decreasing sequence of
smooth plurisubharmonic functions. Moreover, this sequence can be chosen uniformly convergent.[3]
• The inequality in the usual semi-continuity condition holds as equality, i.e. if is plurisubharmonic then

(see limit superior and limit inferior for the definition of lim sup).
• Plurisubharmonic functions are subharmonic, for any Kähler metric.
• Therefore, plurisubharmonic functions satisfy the maximum principle, i.e. if is plurisubharmonic on the
connected domain and

for some point then is constant.

Applications
In complex analysis, plurisubharmonic functions are used to describe pseudoconvex domains, domains of
holomorphy and Stein manifolds.

Oka theorem
The main geometric application of the theory of plurisubharmonic functions is the famous theorem proven by
Kiyoshi Oka in 1942. [1]
A continuous function is called exhaustive if the preimage is compact for all
. A plurisubharmonic function f is called strongly plurisubharmonic if the form is
positive, for some Kähler form on M.
Theorem of Oka: Let M be a complex manifold, admitting a smooth, exhaustive, strongly plurisubharmonic
function. Then M is Stein. Conversely, any Stein manifold admits such a function.

References
• Steven G. Krantz. Function Theory of Several Complex Variables, AMS Chelsea Publishing, Providence, Rhode
Island, 1992.
• Robert C. Gunning. Introduction to Holomorphic Functions in Several Variables, Wadsworth & Brooks/Cole.

Notes
[1] K. Oka, Domaines pseudoconvexes, Tohoku Math. J. 49 (1942), 15–52.
[2] P. Lelong, Definition des fonctions plurisousharmoniques, C. R. Acd. Sci. Paris 215 (1942), 398–400.
[3] R. E. Greene and H. Wu, -approximations of convex, subharmonic, and plurisubharmonic functions, Ann. Scient. Ec. Norm. Sup. 12
(1979), 47–84.
Polyconvex function 112

Polyconvex function
In mathematics, the notion of polyconvexity is a generalization of the notion of convexity for functions defined on
spaces of matrices. Let Mm×n(K) denote the space of all m × n matrices over the field K, which may be either the real
numbers R, or the complex numbers C. A function f : Mm×n(K) → R ∪ {±∞} is said to be polyconvex if

can be written as a convex function of the p × p subdeterminants of A, for 1 ≤ p ≤ min{m, n}.


Polyconvexity is a weaker property than convexity. For example, the function f given by

is polyconvex but not convex.

References
• Renardy, Michael and Rogers, Robert C. (2004). An introduction to partial differential equations. Texts in
Applied Mathematics 13 (Second edition ed.). New York: Springer-Verlag. pp. 353. ISBN 0-387-00444-0.
(Definition 9.25)

Positive-definite function
In mathematics, the term positive-definite function may refer to a couple of different concepts.

In dynamical systems
A real-valued, continuously differentiable function f is positive definite on a neighborhood of the origin, D, if
and for every non-zero .[1] [2]
A function is negative definite if the inequality is reversed. A function is semidefinite if the strong inequality is
replaced with a weak ( or ) one.

In complex analysis and statistics


A positive-definite function of a real variable x is a complex-valued function
f:R → C
such that for any real numbers
x1, ..., xn
the n×n matrix A with entries
aij = f(xi − xj)
is a positive semi-definite matrix. It is usual to restrict to the case in which f(−x) is the complex conjugate of f(x),
making the matrix A Hermitian.
If a function f is positive semidefinite, we find by taking n = 1 that
f(0) ≥ 0.
By taking n=2 and recognising that a positive semi-definite matrix has a nonnegative determinant we get
f(x − y)f(y − x) ≤ f(0)2
Positive-definite function 113

which implies
|f(x)| ≤ f(0).
Positive-definiteness arises naturally in the theory of the Fourier transform; it is easy to see directly that to be
positive-definite is a necessary condition on f, for it to be the Fourier transform of a function g on the real line with
g(y) ≥ 0.
The converse result is Bochner's theorem, stating that a continuous positive-definite function on the real line is the
Fourier transform of a (positive) measure.[3]
This result generalizes to the context of Pontryagin duality, with positive-definite functions defined on any locally
compact abelian topological group. Positive-definite functions on groups occur naturally in the representation theory
of groups on Hilbert spaces (i.e. the theory of unitary representations).
In statistics, and especially Bayesian statistics, the theorem is usually applied to real functions. Typically, one takes n
scalar measurements of some scalar value at points in and one requires that points that are closely separated
have measurements that are highly correlated. In practice, one must be careful to ensure that the resulting covariance
matrix (an n-by-n matrix) is always positive definite. One strategy is to define a correlation matrix A which is then
multiplied by a scalar to give a covariance matrix: this must be positive definite. Bochner's theorem states that if the
correlation between two points is dependent only upon the distance between them (via function f()), then function f()
must be positive definite to ensure the covariance matrix A is positive definite. See Kriging.
In a this context, one does not usually use Fourier terminology and instead one states that f(x) is the characteristic
function of a symmetric PDF.

References
• Z. Sasvári, Positive Definite and Definitizable Functions, Akademie Verlag, 1994
[1] Verhulst, Ferdinand (1996). Nonlinear Differential Equations and Dynamical Systems (2nd ed. ed.). Springer. ISBN 3-540-60934-2.
[2] Hahn, Wolfgang (1967). Stability of Motion. Springer.
[3] Bochner, Salomon (1959). Lectures on Fourier integrals. Princeton University Press.
Proper convex function 114

Proper convex function


In mathematics, a proper convex function is a convex function f taking values in the extended real number line such
that

for at least one x and

for every x. This definition takes account of the fact that the extended real number line does not constitute a field
because, for example, the value of the expression ∞ − ∞ is left undefined.
It is always possible to consider the restriction of a proper convex function f to its effective domain

instead of f itself, thereby avoiding some minor technicalities that may otherwise arise. The effective domain of a
convex function is always a convex set.

Properties
For every proper convex function f on Rn there exist some b in Rn and β in R such that

for every x.
The sum of two proper convex functions is convex but not necessarily proper convex. The infimal convolute of two
proper convex functions is convex but not necessarily proper convex.

References
• Rockafellar, R. Tyrrell, Convex analysis, Princeton University Press (1996). ISBN 0-691-01586-4
Pseudoconvex function 115

Pseudoconvex function
In convex analysis and the calculus of variations, branches of mathematics, a pseudoconvex function is a function
that behaves like a convex function with respect to finding its local minima, but need not actually be convex.
Informally, a differentiable function is pseudoconvex if it is increasing in any direction where it has a positive
directional derivative.

Formal definition
Formally, a real-valued differentiable function ƒ defined on a (nonempty) convex open set X in the finite-dimensional
Euclidean space Rn is said to be pseudoconvex if, for all x, y ∈ X such that , we have
.[1] Here ∇ƒ is the gradient of ƒ, defined by

Properties
Every convex function is pseudoconvex, but the converse is not true. For example, the function ƒ(x) = x + x3 is
pseudoconvex but not convex. Any pseudoconvex function is quasiconvex, but the converse is not true since the
function ƒ(x) = x3 is quasiconvex but not pseudoconvex. Pseudoconvexity is primarily of interest because a point x*
is a local minimum of a pseudoconvex function ƒ if and only if it is a stationary point of ƒ, which is to say that the
gradient of ƒ vanishes at x*:
[2]

Generalization to nondifferentiable functions


The notion of pseudoconvexity can be generalized to nondifferentiable functions as follows.[3] Given any function ƒ :
X → R we can define the upper Dini derivative of ƒ by

where u is any unit vector. The function is said to be pseudoconvex if it is increasing in any direction where the
upper Dini derivative is positive. More precisely, this is characterized in terms of the subdifferential ∂ƒ as follows:
• For all x, y ∈ X, if there exists an x* ∈ ∂ƒ(x) such that then ƒ(x) ≤ ƒ(z) for all z on the line
segment adjoining x and y.
Pseudoconvex function 116

Related notions
A pseudoconcave function is a function whose negative is pseudoconvex. A pseudolinear function is a function
that is both pseudoconvex and pseudoconcave.[4] For example, linear–fractional programs have pseudolinear
objective functions and linear–inequality constraints: These properties allow fractional–linear problems to be solved
by a variant of the simplex algorithm (of George B. Dantzig).[5] [6] [7]

Notes
[1] Mangasarian 1965
[2] Mangasarian 1965
[3] Floudas & Pardalos 2001
[4] Rapcsak 1991
[5] Chapter five: Craven, B. D. (1988). Fractional programming. Sigma Series in Applied Mathematics. 4. Berlin: Heldermann Verlag. pp. 145.
MR949209. ISBN 3-88538-404-3.
[6] Kruk, Serge; Wolkowicz, Henry (1999). "Pseudolinear programming" (http:/ / www. jstor. org/ stable/ 2653207). SIAM Review 41 (4):
pp. 795-805. MR1723002.JSTOR 2653207.doi:DOI:10.1137/S0036144598335259. .
[7] Mathis, Frank H.; Mathis, Lenora Jane (1995). "A nonlinear programming algorithm for hospital management" (http:/ / www. jstor. org/
stable/ 2132826). SIAM Review 37 (2): pp. 230-234. MR1343214.JSTOR 2132826.doi:DOI:10.1137/1037046. .

References
• Floudas, Christodoulos A.; Pardalos, Panos M. (2001), "Generalized monotone multivalued maps", Encyclopedia
of Optimization, Springer, p. 227, ISBN 9780792369325.
• Mangasarian, O. L. (January 1965). "Pseudo-Convex Functions". Journal of the Society for Industrial and
Applied Mathematics Series A 3 (2): 281–290. doi:10.1137/0303020. ISSN 0363-0129..
• Rapcsak, T. (15 February 1991). "On pseudolinear functions". European Journal of Operational Research 50 (3):
353–360. doi:10.1016/0377-2217(91)90267-Y. ISSN 0377-2217.
Quasi-analytic function 117

Quasi-analytic function
In mathematics, a quasi-analytic class of functions is a generalization of the class of real analytic functions based
upon the following fact. If f is an analytic function on an interval , and at some point f and all of its
derivatives are zero, then f is identically zero on all of . Quasi-analytic classes are broader classes of functions
for which this statement still holds true.

Definitions
Let be a sequence of positive real numbers with . Then we define the class of functions
to be those which satisfy

for all , some constant C, and all non-negative integers k. If this is exactly the class of real
analytic functions on . The class is said to be quasi-analytic if whenever and

for some point and all k, f is identically equal to zero.


A function f is called a quasi-analytic function if f is in some quasi-analytic class.

The Denjoy–Carleman theorem


The Denjoy–Carleman theorem, proved by Carleman (1926) after Denjoy (1921) gave some partial results, gives
criteria on the sequence M under which is a quasi-analytic class. It states that the following conditions
are equivalent:
• is quasi-analytic
• where

• , where Mj* is the largest log convex sequence bounded above by Mj.

The proof that the last two conditions are equivalent to the second uses Carleman's inequality.
Example: Denjoy (1921) pointed out that if Mn is given by one of the sequences
, , , ,…
then the corresponding class is quasi-analytic. The first sequence gives analytic functions.
Quasi-analytic function 118

References
• Carleman, T. (1926), Les fonctions quasi-analytiques, Gauthier-Villars
• Cohen, Paul J. (1968), "A simple proof of the Denjoy-Carleman theorem" [1], The American Mathematical
Monthly (Mathematical Association of America) 75 (1): 26–31, doi:10.2307/2315100, MR0225957,
ISSN 0002-9890
• Denjoy, A. (1921), "Sur les fonctions quasi-analytiques de variable réelle", C.R. Acad. Sci. Paris 173: 1329–1331
• Hörmander, Lars (1990), The Analysis of Linear Partial Differential Operators I, Springer-Verlag,
ISBN 3-540-00662
• Leont'ev, A.F. (2001), "Quasi-analytic class" [2], in Hazewinkel, Michiel, Encyclopaedia of Mathematics,
Springer, ISBN 978-1556080104
• Solomentsev, E.D. (2001), "Carleman theorem" [3], in Hazewinkel, Michiel, Encyclopaedia of Mathematics,
Springer, ISBN 978-1556080104

References
[1] http:/ / www. jstor. org/ stable/ 2315100
[2] http:/ / eom. springer. de/ Q/ q076370. htm
[3] http:/ / eom. springer. de/ C/ c020430. htm

Quasiconvex function
In mathematics, a quasiconvex function is a real-valued function
defined on an interval or on a convex subset of a real vector space such
that the inverse image of any set of the form is a convex
set. In lay language, if you cut off the top of the function and look
down at the remaining part of the domain, you always see something
that is convex.
All convex functions are also quasiconvex, but not all quasiconvex
functions are convex, so quasiconvexity is a weak form of convexity.
Quasiconvexity extends the notion of unimodality for functions with a A quasiconvex function which is not convex.

single real argument.

A function which is not quasiconvex: the set of


points in the domain of the function for which the
function values are below the dashed red line is
the union of the two red intervals, which is not a
convex set.
Quasiconvex function 119

Definition and properties


A function defined on a convex subset S of a real vector space is quasiconvex if whenever
and then

If furthermore

for any and , then is strictly quasiconvex.


A quasiconcave function is a function whose negative is quasiconvex, and a strictly quasiconcave function is a
function whose negative is strictly quasiconvex. Equivalently a function is quasiconcave if

and strictly quasiconcave if

A (strictly) quasiconvex function has (strictly) convex lower contour


sets, while a (strictly) quasiconcave function has (strictly) convex
upper contour sets.
A function that is both quasiconvex and quasiconcave is quasilinear.
Optimization methods that work for quasiconvex functions come under
the heading of quasiconvex programming. This comes under the broad
heading of mathematical programming and generalizes both linear
programming and convex programming.
A graph of a quasiconcave function.
There are also minimax theorems on quasiconvex functions, such as
Sion's minimax theorem, which is a far-reaching generalization of the result of von Neumann and Oskar
Morgenstern.
In economics quasiconcave utility functions are desirable as they imply a strictly convex set of preferences.

Operations preserving quasiconvexity


• non-negative weighted maximum of quasiconvex functions (i.e. with
non-negative)
• composition with a non-decreasing function (i.e. quasiconvex, non-decreasing,
then is quasiconvex)
• minimization (i.e. quasiconvex, convex set, then is quasiconvex)
Quasiconvex function 120

Operations not preserving quasiconvexity


• sum of quasiconvex functions defined on the same domain (i.e. if are quasiconvex,
need not be quasiconvex)
• sum of quasiconvex functions defined on different domains (i.e. if are quasiconvex,
need not be quasiconvex)

Examples
• Every convex function is quasiconvex.
• A concave function can be quasiconvex function. For example log(x) is concave, and it is quasiconvex.
• Any monotonic function is both quasiconvex and quasiconcave. More generally, a function which decreases up to
a point and increases from that point on is quasiconvex.
• The floor function is an example of a quasiconvex function that is neither convex nor continuous.
• If f(x) and g(y) are positive convex decreasing functions, then is quasiconvex.

References
• Avriel, M., Diewert, W.E., Schaible, S. and Zang, I., Generalized Concavity, Plenum Press, 1988.
• Singer, Ivan Abstract convex analysis. Canadian Mathematical Society Series of Monographs and Advanced
Texts. A Wiley-Interscience Publication. John Wiley & Sons, Inc., New York, 1997. xxii+491 pp. ISBN:
0-471-16015-6

External links
• SION, M., "On general minimax theorems", Pacific J. Math. 8 (1958), 171-176. [1]
• Mathematical programming glossary [2]
• Concave and Quasi-Concave Functions [3] - by Charles Wilson, NYU Department of Economics
• Quasiconcave [4] - From Econterms, for About.com
• Quasiconcavity and quasiconvexity [5] - by Martin J. Osborne, University of Toronto Department of Economics
• Anatomy of Cobb-Douglas Type Utility Functions in 3D [6] - several examples of quasiconcave utility functions

References
[1] http:/ / projecteuclid. org/ euclid. pjm/ 1103040253
[2] http:/ / glossary. computing. society. informs. org/ second. php
[3] http:/ / homepages. nyu. edu/ ~caw1/ UMath/ Handouts/ ums06h23convexsetsandfunctions. pdf
[4] http:/ / economics. about. com/ od/ economicsglossary/ g/ quasiconc. htm
[5] http:/ / www. economics. utoronto. ca/ osborne/ MathTutorial/ QCC. HTM
[6] http:/ / students. washington. edu/ fuleky/ anatomy/ anatomy. html
Quasiperiodic function 121

Quasiperiodic function
In mathematics, a function f is said to be quasiperiodic with quasiperiod (sometimes simply called the period) ω if
for certain constants a and b, f satisfies the functional equation

An example of this is the Jacobi theta function, where

shows that for fixed τ it has quasiperiod τ; it also is periodic with period one. Another example is provided by the
Weierstrass sigma function, which is quasiperiodic in two independent quasiperiods, the periods of the
corresponding Weierstrass ℘ function.
Functions with an additive functional equation

are also called quasiperiodic. An example of this is the Weierstrass zeta function, where

for a fixed constant η when ω is a period of the corresponding Weierstrass ℘ function.


In the special case where we say f is periodic with period ω.

Quasiperiodic signals
Quasiperiodic signals in the sense of audio processing are not quasiperiodic functions; instead they have the nature
of almost periodic functions and that article should be consulted.

See also
• Quasiperiodicity
• Quasiperiodic motion
• Almost periodic function

External links
• Quasiperiodic function [1] at PlanetMath

References
[1] http:/ / planetmath. org/ encyclopedia/ QuasiperiodicFunction. html
Radially unbounded function 122

Radially unbounded function


In mathematics, a radially unbounded function is a function for which

Such functions are applied in control theory.

References
• Terrell, William J. (2009), Stability and stabilization, Princeton University Press, MR2482799,
ISBN 978-0-691-13444-4

Rational function
In mathematics, a rational function is any function which can be written as the ratio of two polynomial functions.

Definitions
In the case of one variable, , a function is called a
rational function if and only if it can be written in the
form

Rational function of degree 2 :

where and are polynomial functions in and is not the zero polynomial. The domain of is the set of
all points for which the denominator is not zero, where one assumes that the fraction is written in its lower
degree terms, that is, and have several factors of the positive degree.
Every polynomial function is a rational function with . A function that cannot be written in this form
(for example, ) is not a rational function (but the adjective "irrational" is not generally used for
functions, but only for real numbers).

An expression of the form is called a rational expression. The need not be a variable. In abstract algebra

the is called an indeterminate.


Rational function 123

A rational equation is an equation in which two rational expressions are set equal to each other. These expressions
obey the same rules as fractions. The equations can be solved by cross-multiplying. Division by zero is undefined, so
that a solution causing formal division by zero is rejected.

Examples

Rational function of degree 3 :

The rational function is not defined at .

The rational function is defined for all real numbers, but not for all complex numbers, since if x

were the square root of (i.e. the imaginary unit) or its negative, then formal evaluation would lead to division by

zero: , which is undefined.

The rational function , as x approaches infinity, is asymptotic to .

A constant function such as f(x) = π is a rational function since constants are polynomials. Note that the function
itself is rational, even though f(x) is irrational for all x.

Taylor series
The coefficients of a Taylor series of any rational function satisfy a linear recurrence relation, which can be found by
setting the rational function equal to its Taylor series and collecting like terms.
For example,

Multiplying through by the denominator and distributing,


Rational function 124

After adjusting the indices of the sums to get the same powers of x, we get

Combining like terms gives

Since this holds true for all x in the radius of convergence of the original Taylor series, we can compute as follows.
Since the constant term on the left must equal the constant term on the right it follows that

Then, since there are no powers of x on the left, all of the coefficients on the right must be zero, from which it
follows that

Conversely, any sequence that satisfies a linear recurrence determines a rational function when used as the
coefficients of a Taylor series. This is useful in solving such recurrences, since by using partial fraction
decomposition we can write any rational function as a sum of factors of the form 1 / (ax + b) and expand these as
geometric series, giving an explicit formula for the Taylor coefficients; this is the method of generating functions.

Complex analysis
In complex analysis, a rational function

is the ratio of two polynomials with complex coefficients, where Q is not the zero polynomial and P and Q have no
common factor (this avoids f taking the indeterminate value 0/0). The domain and range of f are usually taken to be
the Riemann sphere, which avoids any need for special treatment at the poles of the function (where Q(z) is 0).
The degree of a rational function is the maximum of the degrees of its constituent polynomials P and Q. If the degree
of f is d then the equation

has d distinct solutions in z except for certain values of w, called critical values, where two or more solutions
coincide. f can therefore be thought of as a d-fold covering of the w-sphere by the z-sphere.
Rational functions with degree 1 are called Möbius transformations and are automorphisms of the Riemann sphere.
Rational functions are representative examples of meromorphic functions.

Abstract algebra
In abstract algebra the concept of a polynomial is extended to include formal expressions in which the coefficients of
the polynomial can be taken from any field. In this setting given a field F and some indeterminate X, a rational
expression is any element of the field of fractions of the polynomial ring F[X]. Any rational expression can be
written as the quotient of two polynomials P/Q with Q ≠ 0, although this representation isn't unique. P/Q is
equivalent to R/S, for polynomials P, Q, R, and S, when PS = QR. However since F[X] is a unique factorization
domain, there is a unique representation for any rational expression P/Q with P and Q polynomials of lowest degree
and Q chosen to be monic. This is similar to how a fraction of integers can always be written uniquely in lowest
Rational function 125

terms by canceling out common factors.


The field of rational expressions is denoted F(X). This field is said to be generated (as a field) over F by (a
transcendental element) X, because F(X) does not contain any proper subfield containing both F and the element X.
Like polynomials, rational expressions can also be generalized to n indeterminates X1,..., Xn, by taking the field of
fractions of F[X1,..., Xn], which is denoted by F(X1,..., Xn).
An extended version of the abstract idea of rational function is used in algebraic geometry. There the function field
of an algebraic variety V is formed as the field of fractions of the coordinate ring of V (more accurately said, of a
Zariski-dense affine open set in V). Its elements f are considered as regular functions in the sense of algebraic
geometry on non-empty open sets U, and also may be seen as morphisms to the projective line.

Applications
These objects are first encountered in school algebra. In more advanced mathematics they play an important role in
ring theory, especially in the construction of field extensions. They also provide an example of a nonarchimedean
field (see Archimedean property).
Rational functions are used in numerical analysis for interpolation and approximation of functions, for example the
Padé approximations introduced by Henri Padé. Approximations in terms of rational functions are well suited for
computer algebra systems and other numerical software. Like polynomials, they can be evaluated straightforwardly,
and at the same time they express more diverse behavior than polynomials.
Rational functions are used to approximate or model more complex equations in science and engineering including
(i) fields and forces in physics, (ii) spectroscopy in analytical chemistry, (iii) enzyme kinetics in biochemistry, (iv)
electronic circuitry, (v) aerodynamics, (vi) medicine concentrations in vivo, (vii) wave functions for atoms and
molecules, (viii) optics and photography to improve image resolution, and (ix) acoustics and sound.

References
• Hazewinkel, Michiel, ed. (2001), "Rational function" [1], Encyclopaedia of Mathematics, Springer,
ISBN 978-1556080104

External links
• Dynamic visualization of rational functions with JSXGraph [2]

References
[1] http:/ / eom. springer. de/ r/ r077590. htm
[2] http:/ / jsxgraph. uni-bayreuth. de/ wiki/ index. php/ Rational_functions
Real-valued function 126

Real-valued function
In mathematics, a real-valued function is a function that associates to every element of the domain a real number in
the image.

See also
• Function of a real variable

Ring of symmetric functions


In algebra and in particular in algebraic combinatorics, the ring of symmetric functions, is a specific limit of the
rings of symmetric polynomials in n indeterminates, as n goes to infinity. This ring serves as universal structure in
which relations between symmetric polynomials can be expressed in a way independent of the number n of
indeterminates (but its elements are neither polynomials nor functions). Among other things, this ring plays an
important role in the representation theory of the symmetric groups.

Symmetric polynomials
The study of symmetric functions is based on that of symmetric polynomials. In a polynomial ring in some finite set
of indeterminates, there is an action by ring automorphisms of the symmetric group on (the indices of) the
indeterminates (simultaneaously substituting each of them for another according to the permutation used). The
invariants for this action form the subring of symmetric polynomials. If the indeterminates are X1,…,Xn, then
examples of such symmetric polynomials are

and

A somewhat more complicated example is X13X2X3 +X1X23X3 +X1X2X33 +X13X2X4 +X1X23X4 +X1X2X43 +… where
the summation goes on to include all products of the third power of some variable and two other variables. There are
many specific kinds of symmetric polynomials, such as elementary symmetric polynomials, power sum symmetric
polynomials, monomial symmetric polynomials, complete homogeneous symmetric polynomials, and Schur
polynomials.

The ring of symmetric functions


Most relations between symmetric polynomials do not depend on the number n of indeterminates, other than that
some polynomials in the relation might require n to be large enough in order to be defined. For instance the Newton's
identity for the third power sum polynomial leads to

where the denote elementary symmetric polynomials; this formula is valid for all natural numbers n, and the only
notable dependency on it is that ek(X1,…,Xn) = 0 whenever n < k. One would like to write this as an identity
p3 = e13 − 3e2e1 + 3e3 that does not depend on n at all, and this can be done in the ring of symmetric polynomials. In
that ring there are elements ek for all integers k ≥ 1, and an arbitrary element can be given by a polynomial
expression in them.
Ring of symmetric functions 127

Definitions
A ring of symmetric polynomials can be defined over any commutative ring R, and will be denoted ΛR; the basic
case is for R = Z. The ring ΛR is in fact a graded R-algebra. There are two main constructions for it; the first one
given below can be found in (Stanley, 1999), and the second is essentially the one given in (Macdonald, 1979).

As a ring of formal power series


The easiest (though somewhat heavy) construction starts with the ring of formal power series
R''X''<sub>1</sub>,''X''<sub>2</sub>,… over R in infinitely many indeterminates; one defines ΛR as its subring
consisting of power series S that satisfy
1. S is invariant under any permutation of the indeterminates, and
2. the degrees of the monomials occurring in S are bounded.
Note that because of the second condition, power series are used here only to allow infinitely many terms of a fixed
degree, rather than to sum terms of all possible degrees. Allowing this is necessary because an element that contains
for instance a term X1 should also contain a term Xi for every i > 1 in order to be symmetric. Unlike the whole power
series ring, the subring ΛR is graded by the total degree of monomials: due to condition 2, every element of ΛR is a
finite sum of homogeneous elements of ΛR (which are themselves infinite sums of terms of equal degree). For every
k ≥ 0, the element ek ∈ ΛR is defined as the formal sum of all products of k distinct indeterminates, which is clearly
homogeneous of degree k.

As an algebraic limit
Another construction of ΛR takes somewhat longer to describe, but better indicates the relationship with the rings
R[X1,…,Xn]Sn of symmetric polynomials in n indeterminates. For every n there is a surjective ring homomorphism
ρn from the analoguous ring R[X1,…,Xn+1]Sn+1 with one more indeterminate onto R[X1,…,Xn]Sn, defined by setting
the last indeterminate Xn+1 to 0. Although ρn has a non-trivial kernel, the nonzero elements of that kernel have degree
at least (they are multiples of X1X2…Xn+1). This means that the restriction of ρn to elements of degree at
most n is a bijective linear map, and ρn(ek(X1,…,Xn+1)) = ek(X1,…,Xn) for all k ≤ n. The inverse of this restriction
can be extended uniquely to a ring homomorphism φn from R[X1,…,Xn]Sn to R[X1,…,Xn+1]Sn+1, as follows for
instance from the fundamental theorem of symmetric polynomials. Since the images
φn(ek(X1,…,Xn)) = ek(X1,…,Xn+1) for k = 1,…,n are still algebraically independent over R, the homomorphism φn is
injective and can be viewed as a (somewhat unusual) inclusion of rings. The ring ΛR is then the "union" (direct limit)
of all these rings subject to these inclusions. Since all φn are compatible with the grading by total degree of the rings
involved, ΛR obtains the structure of a graded ring.
This construction differs slightly from the one in (Macdonald, 1979). That construction only uses the surjective
morphisms ρn without mentioning the injective morphisms φn: it constructs the homogeneous components of ΛR
separately, and equips their direct sum with a ring structure using the ρn. It is also observed that the result can be
described as an inverse limit in the category of graded rings. That description however somewhat obscures an
important property typical for a direct limit of injective morphisms, namely that every individual element
(symmetric function) is already faithfully represented in some object used in the limit construction, here a ring
R[X1,…,Xd]Sd. It suffices to take for d the degree of the symmetric function, since the part in degree d is of that ring
is mapped isomorphically to rings with more indeterminates by φn for all n ≥ d. This implies that for studying
relations between individual elements, there is no fundamental difference between symmetric polynomials and
symmetric functions.
Ring of symmetric functions 128

Defining individual symmetric functions


It should be noted that the name "symmetric function" for elements of ΛR is a misnomer: in neither construction the
elements are functions, and in fact, unlike symmetric polynomials, no function of independent variables can be
associated to such elements (for instance e1 would be the sum of all infinitely many variables, which is not defined
unless restrictions are imposed on the variables). However the name is traditional and well established; it can be
found both in (Macdonald, 1979), which says (footnote on p.12)
The elements of Λ (unlike those of Λn) are no longer polynomials: they are formal infinite sums of
monomials. We have therefore reverted to the older terminology of symmetric functions.
(here Λn denotes the ring of symmetric polynomials in n indeterminates), and also in (Stanley, 1999)
To define a symmetric function one must either indicate directly a power series as in the first construction, or give a
symmetric polynomial in n indeterminates for every natural number n in a way compatible with the second
construction. An expression in an unspecified number of indeterminates may do both, for instance

can be taken as the definition of an elementary symmetric function if the number of indeterminates is infinite, or as
the definition of an elementary symmetric polynomial in any finite number of indeterminates. Symmetric
polynomials for the same symmetric function should be compatible with the morphisms ρn (decreasing the number
of indeterminates is obtained by setting some of them to zero, so that the coefficients of any monomial in the
remaining indeterminates is unchanged), and their degree should remain bounded. (An example of a family of
symmetric polynomials that fails both conditions is ; the family fails only the second
condition.) Any symmetric polynomial in n indeterminates can be used to construct a compatible family of
symmetric polynomials, using the morphisms ρi for i < n to decrease the number of indeterminates, and φi for i ≥ n
to increase the number of indeterminates (which amounts to adding all monomials in new indeterminates obtained by
symmetry from monomials already present).
The following are fundamental examples of symmetric functions.
• The monomial symmetric functions mα, determined by monomial Xα (where α = (α1,α2,…) is a sequence of
natural numbers); mα is the sum of all monomials obtained by symmetry from Xα. For a formal definition,
consider such sequences to be infinite by appending zeroes (which does not alter the monomial), and define the
relation "~" between such sequences that expresses that one is a permutation of the other; then

This symmetric function corresponds to the monomial symmetric polynomial mα(X1,…,Xn) for any n large
enough to have the monomial Xα. The distinct monomial symmetric functions are parametrized by the integer
partitions (each mα has a unique representative monomial Xλ with the parts λi in weakly decreasing order).
Since any symmetric function containing any of the monomials of some mα must contain all of them with the
same coefficient, each symmetric function can be written as an R-linear combination of monomial symmetric
functions, and the distinct monomial symmetric functions form a basis of ΛR as R-module.
• The elementary symmetric functions ek, for any natural number k; one has ek = mα where . As
a power series, this is the sum of all distinct products of k distinct indeterminates. This symmetric function
corresponds to the elementary symmetric polynomial ek(X1,…,Xn) for any n ≥ k.
• The power sum symmetric functions pk, for any positive integer k; one has pk = m(k), the monomial symmetric
function for the monomial X1k. This symmetric function corresponds to the power sum symmetric polynomial
pk(X1,…,Xn) = X1k+…+Xnk for any n ≥ 1.
• The complete homogeneous symmetric functions hk, for any natural number k; hk is the sum of all monomial
symmetric functions mα where α is a partition of k. As a power series, this is the sum of all monomials of degree
k, which is what motivates its name. This symmetric function corresponds to the complete homogeneous
Ring of symmetric functions 129

symmetric polynomial hk(X1,…,Xn) for any n ≥ k.


• The Schur functions sλ for any partition λ, which corresponds to the Schur polynomial sλ(X1,…,Xn) for any n
large enough to have the monomial Xλ.
There is no power sum symmetric function p0: although it is possible (and in some contexts natural) to define
as a symmetric polynomial in n variables, these values are not compatible
with the morphisms ρn. The "discriminant" is another example of an expression giving a
symmetric polynomial for all n, but not defining any symmetric function. The expressions defining Schur
polynomials as a quotient of alternating polynomials are somewhat similar to that for the discriminant, but the
polynomials sλ(X1,…,Xn) turn out to be compatible for varying n, and therefore do define a symmetric function.

A principle relating symmetric polynomials and symmetric functions


For any symmetric function P, the corresponding symmetric polynomials in n indeterminates for any natural number
n may be designated by P(X1,…,Xn). The second definition of the ring of symmetric functions implies the following
fundamental principle:
If P and Q are symmetric functions of degree d, then one has the identity of symmetric functions if
and only one has the identity P(X1,…,Xd) = Q(X1,…,Xd) of symmetric polynomials in d indeterminates. In this
case one has in fact P(X1,…,Xn) = Q(X1,…,Xn) for any number n of indeterminates.
This is because one can always reduce the number of variables by substituting zero for some variables, and one can
increase the number of variables by applying the homomorphisms φn; the definition of those homomorphisms
assures that φn(P(X1,…,Xn)) = P(X1,…,Xn+1) (and similarly for Q) whenever n ≥ d. See a proof of Newton's
identities for an effective application of this principle.

Properties of the ring of symmetric functions

Identities
The ring of symmetric functions is a convenient tool for writing identities between symmetric polynomials that are
independent of the number of indeterminates: in ΛR there is no such number, yet by the above principle any identity
in ΛR automatically gives identities the rings of symmetric polynomials over R in any number of indeterminates.
Some fundamental identities are

which shows a symmetry between elementary and complete homogeneous symmetric functions; these relations are
explained under complete homogeneous symmetric polynomial.

the Newton identities, which also have a variant for complete homogeneous symmetric functions:
Ring of symmetric functions 130

Structural properties of ΛR
Important properties of ΛR include the following.
1. The set of monomial symmetric functions parametrized by partitions form a basis of ΛR as graded R-module,
those parametrized by partitions of d being homogeneous of degree d; the same is true for the set of Schur
functions (also parametrized by partitions).
2. ΛR is isomorphic as a graded R-algebra to a polynomial ring R[Y1,Y2,…] in infinitely many variables, where Yi is
given degree i for all i > 0, one isomorphism being the one that sends Yi to ei ∈ ΛR for every i.
3. There is a involutary automorphism ω of ΛR that interchanges the elementary symmetric functions ei and the
complete homogeneous symmetric function hi for all i. It also sends each power sum symmetric function pi to
(−1)i−1 pi, and it permutes the Schur functions among each other, interchanging sλ and sλt where λt is the
transpose partition of λ.
Property 2 is the essence of the fundamental theorem of symmetric polynomials. It immediately implies some other
properties:
• The subring of ΛR generated by its elements of degree at most n is isomorphic to the ring of symmetric
polynomials over R in n variables;
• The Hilbert–Poincaré series of ΛR is , the generating function of the integer partitions (this also
follows from property 1);
• For every n > 0, the R-module formed by the homogeneous part of ΛR of degree n, modulo its intersection with
the subring generated by its elements of degree strictly less than n, is free of rank 1, and (the image of) en is a
generator of this R-module;
• For every family of symmetric functions (fi)i>0 in which fi is homogeneous of degree i and gives a generator of the
free R-module of the previous point (for all i), there is an alternative isomorphism of graded R-algebras from
R[Y1,Y2,…] as above to ΛR that sends Yi to fi; in other words, the family (fi)i>0 forms a set of free polynomial
generators of ΛR.
This final point applies in particular to the family (hi)i>0 of complete homogeneous symmetric functions. If R
contains the field Q of rational numbers, it applies also to the family (pi)i>0 of power sum symmetric functions. This
explains why the first n elements of each of these families define sets of symmetric polynomials in n variables that
are free polynomial generators of that ring of symmetric polynomials.
The fact that the complete homogeneous symmetric functions form a set of free polynomial generators of ΛR already
shows the existence of an automorphism ω sending the elementary symmetric functions to the complete
homogeneous ones, as mentioned in property 3. The fact that ω is an involution of ΛR follows from the symmetry
between elementary and complete homogeneous symmetric functions expressed by the first set of relations given
above.

Generating functions
The first definition of ΛR as a subring of R''X''<sub>1</sub>,''X''<sub>2</sub>,… allows expression the generating
functions of several sequences of symmetric functions to be elegantly expressed. Contrary to the relations mentioned
earlier, which are internal to ΛR, these expressions involve operations taking place in R[[X1,X2,…;t]] but outside its
subring ΛR[[t]], so they are meaningful only if symmetric functions are viewed as formal power series in
indeterminates Xi. We shall write "(X)" after the symmetric functions to stress this interpretation.
The generating function for the elementary symmetric functions is

Similarly one has for complete homogeneous symmetric functions


Ring of symmetric functions 131

The obvious fact that explains the symmetry between elementary and
complete homogeneous symmetric functions. The generating function for the power sum symmetric functions can be
expressed as

((Macdonald, 1979) defines P(t) as Σk>0 pk(X)tk−1, and its expressions therefore lack a factor t with respect to those
given here). The two final expressions, involving the formal derivatives of the generating functions E(t) and H(t),
imply Newton's identities and their variants for the complete homogeneous symmetric functions. These expressions
are sometimes written as

which amounts to the same, but requires that R contain the rational numbers, so that the logarithm of power series
with constant term 1 is defined (by ).

References
• Macdonald, I. G. Symmetric functions and Hall polynomials. Oxford Mathematical Monographs. The Clarendon
Press, Oxford University Press, Oxford, 1979. viii+180 pp. ISBN 0-19-853530-9 MR84g:05003
• Macdonald, I. G. Symmetric functions and Hall polynomials. Second edition. Oxford Mathematical Monographs.
Oxford Science Publications. The Clarendon Press, Oxford University Press, New York, 1995. x+475 pp. ISBN
0-19-853489-2 MR96h:05207
• Stanley, Richard P. Enumerative Combinatorics, Vol. 2, Cambridge University Press, 1999. ISBN 0-521-56069-1
(hardback) ISBN 0-521-78987-7 (paperback).
Simple function 132

Simple function
In the mathematical field of real analysis, a simple function is a (sufficiently 'nice' - see below for the formal
definition) real-valued function over a subset of the real line which attains only a finite number of values. Some
authors also require simple functions to be measurable; as used in practice, they invariably are.
A basic example of a simple function is the floor function over the half-open interval [1,9), whose only values are
{1,2,3,4,5,6,7,8}. A more advanced example is the Dirichlet function over the real line, which takes the value 1 if x
is rational and 0 otherwise. (Thus the "simple" of "simple function" has a technical meaning somewhat at odds with
common language.) Note also that all step functions are simple.
Simple functions are used as a first stage in the development of theories of integration, such as the Lebesgue integral,
because it is very easy to create a definition of an integral for a simple function, and also, it is straightforward to
approximate more general functions by sequences of simple functions.

Definition
Formally, a simple function is a finite linear combination of indicator functions of measurable sets. More precisely,
let (X, Σ) be a measurable space. Let A1, ..., An ∈ Σ be a sequence of measurable sets, and let a1, ..., an be a sequence
of real or complex numbers. A simple function is a function of the form

where is the indicator function of the set A.

Properties of simple functions


By definition, the sum, difference, and product of two simple functions are again simple functions, and
multiplication by constant keeps a simple function simple; hence it follows that the collection of all simple functions
on a given measurable space forms a commutative algebra over .

Integration of simple functions


If a measure μ is defined on the space (X,Σ), the integral of f with respect to μ is

if all summands are finite.

Relation to Lebesgue integration


Any non-negative measurable function is the pointwise limit of a monotonic increasing sequence of
non-negative simple functions. Indeed, let be a non-negative measurable function defined over the measure space
as before. For each , subdivide the range of into intervals, of which have

length . For each , set for , and . (Note

that, for fixed , the sets are disjoint and cover the non-negative real line.)
Now define the measurable sets for . Then the increasing sequence

of simple functions converges pointwise to as . Note that, when is


Simple function 133

bounded, the convergence is uniform. This approximation of by simple functions (which are easily integrable) allows
us to define an integral itself; see the article on Lebesgue integration for more details.

References
• J. F. C. Kingman, S. J. Taylor. Introduction to Measure and Probability, 1966, Cambridge.
• S. Lang. Real and Functional Analysis, 1993, Springer-Verlag.
• W. Rudin. Real and Complex Analysis, 1987, McGraw-Hill.
• H. L. Royden. Real Analysis, 1968, Collier Macmillan.

Single-valued function
A single-valued function is an emphatic term for a mathematical function in the usual sense. That is, each element
of the function's domain maps to a single, well-defined element of its range. Single-valued functions are also referred
to as One-to-One. This contrasts with a general binary relation, which can be viewed as being a multi-valued
function. For example : f(x) = x+3 (each element in domain has not more than one image in range set).

Singular function
In mathematics, a singular function is
any function ƒ defined on the interval
[a, b] that has the following properties:
• ƒ is continuous on [a, b]. (**)
• there exists a set N of measure 0
such that for all x outside of N the
derivative ƒ ′(x) exists and is zero,
that is, the derivative of f vanishes
almost everywhere.
• ƒ is nondecreasing on [a, b].
• ƒ(a) < ƒ(b).
A standard example of a singular
function is the Cantor function, which
is sometimes called the devil's
staircase (a term also used for singular
functions in general). There are,
however, other functions that have
been given that name. One is defined
in terms of the circle map.

If ƒ(x) = 0 for all x ≤ a and ƒ(x) = 1 for The graph of the winding number of the circle map is an example of a singular function.

all x ≥ b, then the function can be


taken to represent a cumulative distribution function for a random variable which is neither a discrete random
variable (since the probability is zero for each point) nor an absolutely continuous random variable (since the
probability density is zero everywhere it exists).
Singular function 134

Singular functions occur, for instance, as sequences of spatially modulated phases or structures in solids and
magnets, described in a prototypical fashion by the model of Frenkel and Kontorova and by the ANNNI model, as
well as in some dynamical systems. Most famously, perhaps, they lie at the center of the fractional quantum Hall
effect.

References
(**) This condition depends on the references [1]
[1] singular function -- Springer Online Reference Works (http:/ / eom. springer. de/ s/ s085550. htm),

• Lebesgue, H. (1955-1961), Theory of functions of a real variable, F. Ungar


• Halmos, P.R. (1950), Measure theory, v. Nostrand
• Royden, H.L (1988), Real Analysis, Prentice-Hall, Englewood Cliffs, New Jersey
• Lebesgue, H. (1928), Leçons sur l'intégration et la récherche des fonctions primitives, Gauthier-Villars

Smooth function
In mathematical analysis, a differentiability class is a
classification of functions according to the properties of
their derivatives. Higher order differentiability classes
correspond to the existence of more derivatives.
Functions that have derivatives of all orders are called
smooth.

Most of this article is about real-valued functions of


one real variable. A discussion of the multivariable
case is presented towards the end.

A bump function is a smooth function with compact support.


Differentiability classes
Consider an open set on the real line and a function f defined on that set with real values. Let k be a non-negative
integer. The function f is said to be of class Ck if the derivatives f', f'', ..., f(k) exist and are continuous (the continuity
is automatic for all the derivatives except for f(k)). The function f is said to be of class C∞, or smooth, if it has
derivatives of all orders.[1] f is said to be of class Cω, or analytic, if f is smooth and if it equals its Taylor series
expansion around any point in its domain.
To put it differently, the class C0 consists of all continuous functions. The class C1 consists of all differentiable
functions whose derivative is continuous; such functions are called continuously differentiable. Thus, a C1 function
is exactly a function whose derivative exists and is of class C0. In general, the classes Ck can be defined recursively
by declaring C0 to be the set of all continuous functions and declaring Ck for any positive integer k to be the set of all
differentiable functions whose derivative is in Ck−1. In particular, Ck is contained in Ck−1 for every k, and there are
examples to show that this containment is strict. C∞ is the intersection of the sets Ck as k varies over the non-negative
integers. Cω is strictly contained in C∞; for an example of this, see bump function or also below.
Smooth function 135

Examples
The function

The C0 function f(x)=x for x≥0 and 0 otherwise.

The function f(x)=x2 sin(1/x) for x>0.

A smooth function that is not analytic.

is continuous, but not differentiable at , so it is of class C0 but not of class C1.


The function

is differentiable, with derivative

Because cos(1/x) oscillates as x approaches zero, f ’(x) is not continuous at zero. Therefore, this function is
differentiable but not of class C1. Moreover, if one takes f(x) = x3/2 sin(1/x) (x ≠ 0) in this example, it can be used to
show that the derivative function of a differentiable function can be unbounded on a compact set and, therefore, that
Smooth function 136

a differentiable function on a compact set may not be locally Lipschitz continuous.


The functions

where k is even, are continuous and k times differentiable at all x. But at they are not (k+1) times differentiable,
so they are of class C k but not of class C j where j>k.
The exponential function is analytic, so, of class Cω. The trigonometric functions are also analytic wherever they are
defined.
The function

is smooth, so of class C∞, but it is not analytic at , so it is not of class Cω. (Piecewise defined functions are
typically not analytic where the pieces meet.) The function f is an example of a smooth function with compact
support.

Multivariate differentiability classes


Let n and m be some positive integers. If f is a function from an open subset of Rn with values in Rm, then f has
component functions f1, ..., fm. Each of these may or may not have partial derivatives. We say that f is of class Ck if

all of the partial derivatives exist and are continuous, where each of is an

integer between 1 and n, each of is an integer between 0 and k, .[1] The


classes C∞ and Cω are defined as before.[1]
These criteria of differentiability can be applied to the transition functions of a differential structure. The resulting
space is called a Ck manifold.
If one wishes to start with a coordinate-independent definition of the class Ck, one may start by considering maps
between Banach spaces. A map from one Banach space to another is differentiable at a point if there is an affine map
which approximates it at that point. The derivative of the map assigns to the point x the linear part of the affine
approximation to the map at x. Since the space of linear maps from one Banach space to another is again a Banach
space, we may continue this procedure to define higher order derivatives. A map f is of class Ck if it has continuous
derivatives up to order k, as before.
Note that Rn is a Banach space for any value of n, so the coordinate-free approach is applicable in this instance. It
can be shown that the definition in terms of partial derivatives and the coordinate-free approach are equivalent; that
is, a function f is of class Ck by one definition iff it is so by the other definition.

The space of Ck functions


Let D be an open subset of the real line. The set of all Ck functions defined on and taking real values is a Fréchet
space with the countable family of seminorms

where K varies over an increasing sequence of compact sets whose union is D, and m = 0, 1, …, k.
The set of C∞ functions over also forms a Fréchet space. One uses the same seminorms as above, except that is
allowed to range over all non-negative integer values.
The above spaces occur naturally in applications where functions having derivatives of certain orders are necessary;
however, particularly in the study of partial differential equations, it can sometimes be more fruitful to work instead
with the Sobolev spaces.
Smooth function 137

Parametric continuity
Parametric continuity is a concept applied to parametric curves describing the smoothness of the parameter's value
with distance along the curve.

Definition
A curve can be said to have Cn continuity if

is continuous of value throughout the curve.


As an example of a practical application of this concept, a curve describing the motion of an object with a parameter
of time, must have C1 continuity for the object to have finite acceleration. For smoother motion, such as that of a
camera's path while making a film, higher levels of parametric continuity are required.

Order of continuity
The various order of parametric continuity can be described as
follows:[2]
• C−1: curves include discontinuities
• C0: curves are joined Two Bézier curve segments attached that is only
1
• C : first derivatives are equal C0 continuous.

• C2: first and second derivatives are equal


• Cn: first through nth derivatives are equal
The term parametric continuity was introduced to distinguish it from
geometric continuity (Gn) which removes restrictions on the speed with
which the parameter traces out the curve.[3]
Two Bézier curve segments attached in such a
way that they are C1 continuous.

Geometric continuity
The concept of geometrical or geometric continuity was primarily applied to the conic sections and related shapes
by mathematicians such as Leibniz, Kepler, and Poncelet. The concept was an early attempt at describing, through
geometry rather than algebra, the concept of continuity as expressed through a parametric function.
The basic idea behind geometric continuity was that the five conic sections were really five different versions of the
same shape. An ellipse tends to a circle as the eccentricity approaches zero, or to a parabola as it approaches one; and
a hyperbola tends to a parabola as the eccentricity drops toward one; it can also tend to intersecting lines. Thus, there
was continuity between the conic sections. These ideas led to other concepts of continuity. For instance, if a circle
and a straight line were two expressions of the same shape, perhaps a line could be thought of as a circle of infinite
radius. For such to be the case, one would have to make the line closed by allowing the point x = ∞ to be a point on
the circle, and for x = +∞ and x = −∞ to be identical. Such ideas were useful in crafting the modern, algebraically
defined, idea of the continuity of a function and of ∞.
Smooth function 138

Smoothness of curves and surfaces


A curve or surface can be described as having Gn continuity, n being the increasing measure of smoothness.
Consider the segments either side of a point on a curve:
• G0: The curves touch at the join point.
• G1: The curves also share a common tangent direction at the join point.
• G2: The curves also share a common center of curvature at the join point.
In general, Gn continuity exists if the curves can be reparameterized to have Cn (parametric) continuity.[4] A
reparametrization of the curve is geometrically identical to the original; only the parameter is affected.
Equivalently, two vector functions and have Gn continuity if and , for a scalar
(i.e., if the direction, but not necessarily the magnitude, of the two vectors is equal).
While it may be obvious that a curve would require G1 continuity to appear smooth, for good aesthetics, such as
those aspired to in architecture and sports car design, higher levels of geometric continuity are required. For
example, reflections in a car body will not appear smooth unless the body has G2 continuity.
A rounded rectangle (with ninety degree circular arcs at the four corners) has G1 continuity, but does not have G2
continuity. The same is true for a rounded cube, with octants of a sphere at its corners and quarter-cylinders along its
edges. If an editable curve with G2 continuity is required, then cubic splines are typically chosen; these curves are
frequently used in industrial design.

Smoothness

Relation to analyticity
While all analytic functions are smooth on the set on which they are analytic, the above example shows that the
converse is not true for functions on the reals: there exist smooth real functions which are not analytic. For example,
the Fabius function is smooth but not analytic at any point. Although it might seem that such functions are the
exception rather than the rule, it turns out that the analytic functions are scattered very thinly among the smooth
ones; more rigorously, the analytic functions form a meagre subset of the smooth functions. Furthermore, for every
open subset A of the real line, there exist smooth functions which are analytic on A and nowhere else.
It is useful to compare the situation to that of the ubiquity of transcendental numbers on the real line. Both on the real
line and the set of smooth functions, the examples we come up with at first thought (algebraic/rational numbers and
analytic functions) are far better behaved than the majority of cases: the transcendental numbers and nowhere
analytic functions have full measure (their complements are meagre).
The situation thus described is in marked contrast to complex differentiable functions. If a complex function is
differentiable just once on an open set it is both infinitely differentiable and analytic on that set.

Smooth partitions of unity


Smooth functions with given closed support are used in the construction of smooth partitions of unity (see partition
of unity and topology glossary); these are essential in the study of smooth manifolds, for example to show that
Riemannian metrics can be defined globally starting from their local existence. A simple case is that of a bump
function on the real line, that is, a smooth function f that takes the value 0 outside an interval [a,b] and such that

Given a number of overlapping intervals on the line, bump functions can be constructed on each of them, and on
semi-infinite intervals (-∞, c] and [d,+∞) to cover the whole line, such that the sum of the functions is always 1.
From what has just been said, partitions of unity don't apply to holomorphic functions; their different behavior
relative to existence and analytic continuation is one of the roots of sheaf theory. In contrast, sheaves of smooth
Smooth function 139

functions tend not to carry much topological information.

Smooth functions between manifolds


Smooth maps between smooth manifolds may be defined by means of charts, since the idea of smoothness of
function is independent of the particular chart used. If F is a map from an m-manifold M to an n-manifold N, then F
is smooth if, for every , there is a chart in M containing p and a chart in N containing F(p) with
, such that is a smooth from to as a function from to .
Such a map has a first derivative defined on tangent vectors; it gives a fibre-wise linear mapping on the level of
tangent bundles.

Smooth functions between subsets of manifolds


There is a corresponding notion of smooth map for arbitrary subsets of manifolds. If is a function whose
domain and range are subsets of manifolds and respectively. is said to be smooth if for all there
is an open set with and a smooth function such that for all .

References
[1] Warner (1883), p. 5, Definition 1.2 (http:/ / books. google. com/ books?id=t6PNrjnfhuIC& pg=PA5& dq="f+ is+ differentiable+ of+ class+
Ck").
[2] Parametric Curves (http:/ / www. cs. helsinki. fi/ group/ goa/ mallinnus/ curves/ curves. html)
[3] (Bartels, Beatty & Barsky 1987, Ch. 13)
[4] Brian A. Barsky and Tony D. DeRose, "Geometric Continuity of Parametric Curves: Three Equivalent Characterizations," IEEE Computer
Graphics and Applications, 9(6), Nov. 1989, pp. 60–68.

• Kim, Sung S.; Kwon, Kil H. (March 2000). "Smooth (C-inf) but Nowhere Analytic Functions" (http://links.
jstor.org/sici?sici=0002-9890(200003)107:3<264:S(BNAF>2.0.CO;2-5). The American Mathematical Monthly
(Mathematical Association of America) 107 (3): 264–266. doi:10.2307/2589322. Retrieved 2008-04-04.
• Guillemin, Pollack. Differential Topology. Prentice-Hall (1974).
•  This article incorporates text from a publication now in the public domain: Chisholm, Hugh, ed (1911).
Encyclopædia Britannica (Eleventh ed.). Cambridge University Press.

• Warner, Frank Wilson (1983). Foundations of differentiable manifolds and Lie groups. Springer.
ISBN 9780387908946.
Subharmonic function 140

Subharmonic function
In mathematics, subharmonic and superharmonic functions are important classes of functions used extensively in
partial differential equations, complex analysis and potential theory.
Intuitively, subharmonic functions are related to convex functions of one variable as follows. If the graph of a
convex function and a line intersect at two points, then the graph of the convex function is below the line between
those points. In the same way, if the values of a subharmonic function are no larger than the values of a harmonic
function on the boundary of a ball, then the values of the subharmonic function are no larger than the values of the
harmonic function also inside the ball.
Superharmonic functions can be defined by the same description, only replacing "no larger" with "no smaller".
Alternatively, a superharmonic function is just the negative of a subharmonic function, and for this reason any
property of subharmonic functions can be easily transferred to superharmonic functions.

Formal definition
Formally, the definition can be stated as follows. Let be a subset of the Euclidean space and let

be an upper semi-continuous function. Then, is called subharmonic if for any closed ball of centre
and radius contained in and every real-valued continuous function on that is harmonic in
and satisfies for all on the boundary of we have
for all
Note that by the above, the function which is identically −∞ is subharmonic, but some authors exclude this function
by definition.

Properties
• A function is harmonic if and only if it is both subharmonic and superharmonic.
• If is C2 (twice continuously differentiable) on an open set in , then is subharmonic if and only if
one has
on
where is the Laplacian.
• The maximum of a subharmonic function cannot be achieved in the interior of its domain unless the function is
constant, this is the so-called maximum principle.
• Subharmonic functions are upper semicontinuous, while superharmonic functions are lower semicontinuous.
Subharmonic function 141

Subharmonic functions in the complex plane


Subharmonic functions are of a particular importance in complex analysis, where they are intimately connected to
holomorphic functions.
One can show that a real-valued, continuous function of a complex variable (that is, of two real variables) defined
on a set is subharmonic if and only if for any closed disc of center and radius one
has

Intuitively, this means that a subharmonic function is at any point no greater than the average of the values in a circle
around that point, a fact which can be used to derive the maximum principle.
If is a holomorphic function, then

is a subharmonic function if we define the value of at the zeros of to be −∞. It follows that

is subharmonic for every α > 0. This observation plays a role in the theory of Hardy spaces, especially for the study
of Hp when 0 < p < 1.
In the context of the complex plane, the connection to the convex functions can be realized as well by the fact that a
subharmonic function on a domain that is constant in the imaginary direction is convex in the real
direction and vice versa.

Harmonic Majorants of Subharmonic Functions


If is subharmonic in a region of the complex plane, and is harmonic on , then is a harmonic
majorant of in if ≤ in . Such an inequality can be viewed as a growth condition on .[1]

Subharmonic functions in the unit disc. Radial maximal function


Let φ be subharmonic, continuous and non-negative in an open subset Ω of the complex plane containing the closed
unit disc D(0, 1). The radial maximal function for the function φ (restricted to the unit disc) is defined on the unit
circle by

If Pr denotes the Poisson kernel, it follows from the subharmonicity that

It can be shown that the last integral is less than the value at e iθ of the Hardy–Littlewood maximal function φ∗ of the
restriction of φ to the unit circle T,

so that 0 ≤ M φ ≤ φ∗. It is known that the Hardy–Littlewood operator is bounded on Lp(T) when 1 < p < ∞. It
follows that for some universal constant C,

If f is a function holomorphic in Ω and 0 < p < ∞, then the preceding inequality applies to φ = |f | p/2. It can be
deduced from these facts that any function F in the classical Hardy space Hp satisfies
Subharmonic function 142

With more work, it can be shown that F has radial limits F(e iθ) almost everywhere on the unit circle, and (by the
dominated convergence theorem) that Fr, defined by Fr(e iθ) = F(r e iθ) tends to F in Lp(T).

Subharmonic functions on Riemannian manifolds


Subharmonic functions can be defined on an arbitrary Riemannian manifold.
Definition: Let M be a Riemannian manifold, and an upper semicontinuous function. Assume that
for any open subset , and any harmonic function f1 on U, such that on the boundary of U, the
inequality holds on all U. Then f is called subharmonic.
This definition is equivalent to one given above. Also, for twice differentiable functions, subharmonicity is
equivalent to the inequality , where is the usual Laplacian.[2]

Notes
[1] Rosenblum, Marvin; Rovnyak, James (1994), p.35 (see References)
[2] Greene, R. E.; Wu, H. (1974). "Integrals of subharmonic functions on manifolds of nonnegative curvature". Inventiones mathematicae 27:
265–298. doi:10.1007/BF01425500, MR0382723

References
• Conway, John B. (1978). Functions of one complex variable. New York: Springer-Verlag. ISBN 0-387-90328-3.
• Krantz, Steven G. (1992). Function Theory of Several Complex Variables. Providence, Rhode Island: AMS
Chelsea Publishing. ISBN 0-8218-2724-3.
• Doob, Joseph Leo (1984). Classical Potential Theory and Its Probabilistic Counterpart. Berlin Heidelberg New
York: Springer-Verlag. ISBN 3-540-41206-9.
• Rosenblum, Marvin; Rovnyak, James (1994). Topics in Hardy classes and univalent functions. Birkhauser
Advanced Texts: Basel Textbooks. Birkhauser Verlag.

This article incorporates material from Subharmonic and superharmonic functions on PlanetMath, which is licensed
under the Creative Commons Attribution/Share-Alike License.
Sublinear function 143

Sublinear function
A sublinear function, in linear algebra and related areas of mathematics, is a function on a vector
space V over F, an ordered field (e.g. the real numbers ), which satisfies
  for any and any x ∈ V (positive homogenity),
  for any x, y ∈ V (subadditivity).
In functional analysis the name Banach functional is used for sublinear function, especially when formulating
Hahn–Banach theorem.
In computer science, a function is called sublinear if in asymptotic notation (Notice
the small ). Formally, if and only if, for any given , there exists an such that[1]

This means that for any linear function , for sufficiently large input grows slower than .

Examples
• Every (semi-)norm is a sublinear function. Opposite is not true, because (semi-)norms can have their domain
vector space over any field (not necessarily ordered) and must have as their codomain.

Properties
• Every sublinear function is a convex functional.

Operators
The concept can be extended to operators that are homogeneous and subadditive. This requires only that the
codomain be, say, an ordered vector space to make sense of the conditions.

References
[1] Thomas H. Cormen, Charles E. Leiserson, Ronald L. Rivest, and Clifford Stein (2001) [1990]. "3.1". Introduction to Algorithms (2nd edition
ed.). MIT Press and McGraw-Hill. pp. 47–48. ISBN 0-262-03293-7.
Surjective function 144

Surjective function
In mathematics, a function is said to be surjective or
onto if its image is equal to its codomain. A function f:
X → Y is surjective if and only if for every y in the
codomain Y there is at least one x in the domain X such
that f(x) = y. A surjective function is called a
surjection. Surjections are sometimes denoted by a
two-headed rightwards arrow, as in f: X ↠ Y.

The term surjective and the related terms injective and


bijective were introduced by Nicolas Bourbaki,[1] a
group of mainly French 20th-century mathematicians
who wrote a series of books presenting an exposition of
modern advanced mathematics, beginning in 1935. The
A non-surjective function from domain X to codomain Y. The
French prefix sur means over or above and relates to smaller oval inside Y is the image (also called range) of f. This
the fact that the image of the domain of a surjective function is not surjective, because the image does not fill the whole
function completely covers the function's codomain. codomain. In other words, Y is colored in a two-step process: First,
for every x in X, the point f(x) is colored green; Second, all the rest
of the points in Y, that are not green, are colored blue. The function f

Examples is surjective only if there are no blue points.

For any set X, the identity function idX on X is surjective.


The function f: Z → {0,1} defined by f(n) = n mod 2 and mapping even integers to 0 and odd integers to 1 is
surjective.
The function f: R → R defined by f(x) = 2x + 1 is surjective (and even bijective), because for every real number y we
have an x such that f(x) = y: an appropriate x is (y − 1)/2.
The function g: R → R defined by g(x) = x2 is not surjective, because there is no real number x such that x2 = −1.
However, the function g: R → R+ defined by g(x) = x2 (with restricted codomain) is surjective because for every y in
the positive real codomain Y there is at least one x in the real domain X such that x2 = y.
The natural logarithm function ln: (0,+∞) → R is a surjective and even bijective mapping from the set of positive
real numbers to the set of all real numbers. Its inverse, the exponential function, is not surjective as its range is the
set of positive real numbers and its domain is usually defined to be the set of all real numbers. The matrix
exponential is not surjective when seen as a map from the space of all n×n matrices to itself. It is, however, usually
defined as a map from the space of all n×n matrices to the general linear group of degree n, i.e. the group of all n×n
invertible matrices. Under this definition the matrix exponential is surjective.
The projection from a cartesian product A × B to one of its factors is surjective.

Properties
A function is bijective if and only if it is both surjective and injective.

Surjections as right invertible functions


The function g : Y → X is said to be a right inverse of the function f : X → Y if f(g(y)) = y for every y in Y (g can be
undone by f). In other words, g is a right inverse of f if the composition f o g of g and f in that order is the identity
function on the domain Y of g. The function g need not be a complete inverse of f because the composition in the
other order, g o f, may not be the identity function on the domain X of f. In other words, f can undo or "reverse" g, but
Surjective function 145

not necessarily can be reversed by it.


Every function with a right inverse is necessarily a surjection. The proposition that every surjective function has a
right inverse is equivalent to the axiom of choice.
If f: X → Y is surjective and B is a subset of Y, then f(f −1(B)) = B. Thus, B can be recovered from its preimage
f −1(B).
For example, in the first illustration, there is some function g such that g(C) = 4. There is also some function f such
that f(4) = C. It doesn't matter that g(C) can also equal 3; it only matters that f "reverses" g.

A surjective function. Another surjective A non-surjective Surjective composition: the first


(However, this one is function. (This one function. (This one function need not be surjective.
not an injection) happens to be a happens to be an
bijection) injection)

Surjections as epimorphisms
A function f: X → Y is surjective if and only if it is right-cancellative:[2] given any functions g,h:Y → Z, whenever
g o f = h o f, then g = h. This property is formulated in terms of functions and their composition and can be
generalized to the more general notion of the morphisms of a category and their composition. Right-cancellative
morphisms are called epimorphisms. Specifically, surjective functions are precisely the epimorphisms in the
category of sets. The prefix epi is derived from the greek preposition ἐπί meaning over, above, on.
Any morphism with a right inverse is an epimorphism, but the converse is not true in general. A right inverse g of a
morphism f is called a section of f. A morphism with a right inverse is called a split epimorphism.

Surjections as binary relations


Any function with domain X and codomain Y can be seen as a left-total and right-unique binary relation between X
and Y by identifying it with its function graph. A surjective function with domain X and codomain Y is then a binary
relation between X and Y that is right-unique and both left-total and right-total.

Cardinality of the domain of a surjection


The cardinality of the domain of a surjective function is greater than or equal to the cardinality of its codomain: If
f: X → Y is a surjective function, then X has at least as many elements as Y, in the sense of cardinal numbers.
Specifically, if both X and Y are finite with the same number of elements, then f : X → Y is surjective if and only if f
is injective.
Surjective function 146

Composition and decomposition


The composite of surjective functions is always surjective: If f and g are both surjective, then g o f is surjective.
Conversely, if f o g is surjective, then f is surjective (but g need not be). These properties generalize from surjections
in the category of sets to any epimorphisms in any category.
Any function can be decomposed into a surjection and an injection: For any function h: X → Z there exist a
surjection f:X → Y and an injection g:Y → Z such that h = g o f. To see this, define Y to be the sets h −1(z) where z is
in Z. These sets are disjoint and partition X. Then f carries each x to the element of Y which contains it, and g carries
each element of Y to the point in Z to which h sends its points. Then f is surjective since it is a projection map, and g
is injective by definition.

Induced surjection and induced bijection


Any function induces a surjection by restricting its codomain to its range. Any surjective function induces a bijection
defined on a quotient of its domain by collapsing all arguments mapping to a given fixed image. More precisely,
every surjection f : A → B can be factored as a projection followed by a bijection as follows. Let A/~ be the
equivalence classes of A under the following equivalence relation: x ~ y if and only if f(x) = f(y). Equivalently, A/~ is
the set of all preimages under f. Let P(~) : A → A/~ be the projection map which sends each x in A to its equivalence
class [x]~, and let fP : A/~ → B be the well-defined function given by fP([x]~) = f(x). Then f = fP o P(~).

Notes
[1] Earliest Uses of Some of the Words of Mathematics: entry on Injection, Surjection and Bijection has the history of surjection and related
terms. (http:/ / jeff560. tripod. com/ i. html)
[2] Goldblatt, Robert (2006) [1984]. Topoi, the Categorial Analysis of Logic (http:/ / historical. library. cornell. edu/ cgi-bin/ cul. math/
docviewer?did=Gold010& id=3) (Revised ed.). Dover Publications. ISBN 978-0486450261. . Retrieved 2009-11-25.

References
• Bourbaki, Nicolas (2004) [1968]. Theory of Sets. Springer. ISBN 978-3540225256.
Symmetrically continuous function 147

Symmetrically continuous function


In mathematics, a function is symmetrically continuous at a point x if

The usual definition of continuity implies symmetric continuity, but the converse is not true. For example, the
function is symmetrically continuous at , but not continuous.
Also, symmetric differentiability implies symmetric continuity, but the converse is not true just like usual continuity
does not imply differentiability.

References
• Thomson, Brian S. (1994). Symmetric Properties of Real Functions. Marcel Dekker. ISBN 0-8247-9230-0.

Quasisymmetric function
In algebra and in particular in algebraic combinatorics, a quasisymmetric function is any element in the ring of
quasisymmetric functions which is in turn a subring of the formal power series ring with a countable number of
variables. This ring generalizes the ring of symmetric functions. This ring can be realized as a specific limit of the
rings of quasisymmetric polynomials in n variables, as n goes to infinity. This ring serves as universal structure in
which relations between quasisymmetric polynomials can be expressed in a way independent of the number n of
variables (but its elements are neither polynomials nor functions).

Definitions
The ring of quasisymmetric functions, denoted QSym, can be defined over any commutative ring R such as the
integers. Quasisymmetric functions are power series of bounded degree in variables with
coefficients in R, which are shift invariant in the sense that the coefficient of the monomial is equal
to the coefficient of the monomial for any strictly increasing sequence of positive integers
indexing the variables and any positive integer sequence of exponents.[1]
Much of the study of quasisymmetric functions is based on that of symmetric functions.
A quasisymmetric function in finitely many variables is a quasisymmetric polynomial. Both symmetric and
quasisymmetric polynomials may be characterized in terms of actions of the symmetric group on a polynomial
ring in variables . One such action of permutes variables, changing a polynomial
by iteratively swapping pairs of variables having consecutive indices. Those
polynomials unchanged by all such swaps form the subring of symmetric polynomials. A second action of
conditionally permutes variables, changing a polynomial by swapping pairs of
variables except in monomials containing both variables. Those polynomials unchanged by all such conditional
swaps form the subring of quasisymmetric polynomials. One quasisymmetric function in four variables is the
polynomial

The simplest symmetric function containing all of these monomials is


Quasisymmetric function 148

Important bases
QSym is a graded R-algebra, decomposing as

where is the -span of all quasisymmetric functions that are homogeneous of degree . Two natural
bases for are the monomial basis and the fundamental basis indexed by compositions
of , denoted . The monomial basis consists of and all formal power
series

The fundamental basis consists and all formal power series

where means we can obtain by adding together adjacent parts of , for example, (3,2,4,2)
(3,1,1,1,2,1,2). Thus, when the ring is the ring of rational numbers, one has

Then one can define the algebra of symmetric functions as the subalgebra of QSym spanned
by the monomial symmetric functions and all formal power series where the sum is over
all compositions which rearrange to the partition . Moreover, we have . For example,
and
Other important bases for quasisymmetric functions include the basis of quasisymmetric Schur functions[2] , and
bases related to enumeration in matroids[3] [4] .

Applications
Quasisymmetric functions have been applied in enumerative combinatorics, symmetric function theory,
representation theory, and number theory. Applications of quasisymmetric functions include enumeration of
P-partitions [5] [6] , permutations [7] [8] [9] [10] , tableaux [11] , chains of posets [11] [12] , reduced decompositions in
finite Coxeter groups [11] , and parking functions[13] . In symmetric function theory and representation theory,
applications include the study of Schubert polynomials [14] [15] , Macdonald polynomials [16] , Hecke algebras [17] ,
and Kazhdan-Lusztig polynomials [18] . Often quasisymmetric functions provide a powerful bridge between
combinatorial structures and symmetric functions.

Related algebras
As a graded Hopf algebra, the dual of the ring of quasisymmetric functions is the ring of noncommutative symmetric
functions. Every symmetric function is also a quasisymmetric function, and hence the ring of symmetric functions is
a subalgebra of the ring of quasisymmetric functions.
The ring of quasisymmetric functions is the terminal object in category of graded Hopf algebras with a single
character.[19] Hence any such Hopf algebra has a embedding as a subalgebra of the ring of quasisymmetric functions.
One very important example of this is the peak algebra (make page for peak algebra)[20] .
Other Related Algebras: The Malvenuto-Reutenauer algebra[21] is a Hopf algebra based on permutations that relates
the rings of symmetric functions, quasisymmetric functions, and noncommutative symmetric functions, (denoted
Sym, QSym, and NSym respectively), as depicted the following commutative diagram. The duality between QSym
and NSym mentioned above is reflected in the main diagonal of this diagram.
Quasisymmetric function 149

Many related Hopf algebras were constructed from Hopf monoids in the category of species by Aguiar and Majahan
[22]
.
One can also construct the ring of quasisymmetric functions in noncommuting variables.[23] [24]

External links
• BIRS Workshop on Quasisymmetric Functions [25]

References
[1] Stanley, Richard P. Enumerative Combinatorics, Vol. 2, Cambridge University Press, 1999. ISBN 0-521-56069-1 (hardback) ISBN
0-521-78987-7 (paperback).
[2] J. Haglund, K. Luoto, S. Mason ans S. van Willigenburg, Quasisymmetric Schur functions, J. Combin. Theory Ser. A 118 (2011) 463–490
[3] K. Luoto, A matroid-friendly basis for the quasisymmetric functions, J. Combin. Theory Ser. A 115 (2008), 777–798
[4] L. Billera, N. Jia and V. Reiner, A quasisymmetric function for matroids, European J. Combin. 30 (2009), 1727–1757
[5] Stanley, Richard P. Ordered structures and partitions, Memoirs of the American Mathematical Society, No. 119, American Mathematical
Society, 1972.
[6] Gessel, Ira. Multipartite P-partitions and inner products of skew Schur functions, Combinatorics and algebra (Boulder, Colo., 1983),
289–317, Contemp. Math., 34, Amer. Math. Soc., Providence, RI, 1984.
[7] Gessel, Ira M.; Reutenauer, Christophe, Counting permutations with given

cycle structure and descent set. J. Combin. Theory Ser. A 64 (1993), no. 2, 189–215
[8] Shareshian, John; Wachs, Michelle L.; -Eulerian polynomials:
excedance number and major index. Electron. Res. Announc. Amer. Math. Soc. 13 (2007), 33–45
[9] Shareshian, John; Wachs, Michelle L., Eulerian quasisymmetric functions,

Advances in Mathematics, Volume 225, Issue 6, 20 December 2010, Pages 2921-2966


[10] Hyatt, Matthew; Eulerian quasisymmetric functions for the type B Coxeter group and other wreath product groups,arXiv:1007.0459
[11] Stanley, Richard P., On the number of reduced decompositions of elements of Coxeter groups. European J. Combin. 5 (1984), no. 4,
359–372.
[12] Ehrenborg, Richard, On posets and Hopf algebras. Adv. Math. 119

(1996), no. 1, 1–25.


[13] Haglund, James; The q,t-Catalan numbers and the space of diagonal harmonics. University Lecture Series, 41. American Mathematical
Society, Providence, RI, 2008. viii+167 pp. ISBN 978-0-8218-4411-3; 0-8218-4411-3
[14] Billey, Sara C.; Jockusch, William; Stanley, Richard P. Some combinatorial properties of Schubert polynomials. (English summary)

J. Algebraic Combin. 2 (1993), no. 4, 345–374.


[15] Fomin, Sergey; Stanley, Richard P. Schubert polynomials and the nil-Coxeter algebra. (English summary) Adv. Math. 103 (1994), no. 2,
196–207.
[16] Assaf, Sami, Dual Equivalence Graphs I: A combinatorial proof of LLT and Macdonald positivity,arXiv:1005.3759.
[17] Duchamp, Gérard; Krob, Daniel; Leclerc, Bernard; Thibon, Jean-Yves; Fonctions quasi-symétriques, fonctions symétriques non
commutatives et algèbres de Hecke à C. R. Acad. Sci. Paris Sér. I Math. 322 (1996), no. 2, 107–112
[18] Billera, Louis J.; Brenti, Francesco; Quasisymmetric functions and Kazhdan-Lusztig polynomials, arXiv:0710.3965.
Quasisymmetric function 150

[19] Aguiar, Marcelo; Bergeron, Nantel; and Sottile, Frank. Combinatorial Hopf algebras and generalized Dehn-Sommerville relations. Compos.
Math. 142 (2006), no. 1, 1–30.
[20] Stembridge, John R. Enriched P-partitions. Trans. Amer. Math. Soc. 349 (1997), no. 2, 763–788.
[21] Malvenuto, Clauda; Reutenauer, Christophe Duality between quasi-symmetric functions and the Solomon descent algebra. J. Algebra 177
(1995), no. 3, 967–982.
[22] Aguiar, Marcelo; Mahajan, Swapneel Monoidal Functors, Species and Hopf Algebras CRM Monograph Series, no. 29. American
Mathematical Society, Providence, RI, 2010.
[23] Hiver, Florent, Ph.D. Thesis, Marne-la-Valee
[24] Bergeron, Nantel; Zabrocki, Mike The Hopf algebras of symmetric functions and quasi-symmetric functions in non-commutative variables
are free and co-free. J. Algebra Appl. 8 (2009), no. 4, 581–600.
[25] http:/ / www. birs. ca/ events/ 2010/ 5-day-workshops/ 10w5031

Transcendental function
A transcendental function is a function that does not satisfy a polynomial equation whose coefficients are
themselves polynomials, in contrast to an algebraic function, which does satisfy such an equation.[1] In other words,
a transcendental function is a function that "transcends" algebra in the sense that it cannot be expressed in terms of
a finite sequence of the algebraic operations of addition, multiplication, and root extraction.
Examples of transcendental functions include the exponential function, the logarithm, and the trigonometric
functions.
Formally, an analytic function ƒ(z) of the real or complex variables z1,…,zn is transcendental if the n + 1 functions
z1,…,zn, ƒ(z) are algebraically independent.[2] That is, ƒ is transcendental over the field C(z1,…,zn).

Algebraic and transcendental functions


The logarithm and the exponential function are examples of transcendental functions. Transcendental function is a
term often used to describe the trigonometric functions (sine, cosine, tangent, cotangent, secant, cosecant, versine,
haversine, coversine, and so forth).
A function that is not transcendental is said to be algebraic. Examples of algebraic functions are rational functions
and the square root function.
The operation of taking the indefinite integral of an algebraic function is a source of transcendental functions. For
example, the logarithm function arose from the reciprocal function in an effort to find the area of a hyperbolic sector.
Thus the hyperbolic angle and the hyperbolic functions sinh, cosh, and tanh are all transcendental.
Differential algebra examines how integration frequently creates functions that are algebraically independent of
some class, such as when one takes polynomials with trigonometric functions as variables.

Dimensional analysis
In dimensional analysis, transcendental functions are notable because they make sense only when their argument is
dimensionless (possibly after algebraic reduction). Because of this, transcendental functions can be an easy-to-spot
source of dimensional errors. For example, log(10 m) is a nonsensical expression, unlike  log(5 meters / 3 meters)  or
 log(3) meters . One could attempt to apply a logarithmic identity to get log(10) + log(m), which highlights the
problem: applying a non-algebraic operation to a dimension creates meaningless results.
Transcendental function 151

Some examples
All of the following functions are transcendental.

Note that in particular for if we set c equal to , the base of the natural logarithm, then we get that is a
transcendental function. Similarly, if we set c equal to in , then we get that , the natural logarithm, is a
transcendental function.

Exceptional set
If ƒ(z) is an algebraic function and α is an algebraic number then ƒ(α) will also be an algebraic number. The converse
is not true: there are entire transcendental functions ƒ(z) such that ƒ(α) is an algebraic number for any algebraic α. In
many instances, however, the set of algebraic numbers α where ƒ(α) is algebraic is fairly small. For example, if ƒ is
the exponential function, ƒ(z) = ez, then the only algebraic number α where ƒ(α) is also algebraic is α = 0, where
ƒ(α) = 1. For a given transcendental function this set of algebraic numbers giving algebraic results is called the
exceptional set of the function,[3] [4] that is the set

If this set can be calculated then it can often lead to results in transcendence theory. For example, Lindemann proved
in 1882 that the exceptional set of the exponential function is just {0}. In particular exp(1) = e is transcendental.
Also, since exp(iπ) = −1 is algebraic we know that iπ cannot be algebraic. Since i is algebraic this implies that π is a
transcendental number.
In general, finding the exceptional set of a function is a difficult problem, but it has been calculated for some
functions:
• ,
• ,
• Here j is Klein's j-invariant, H is the upper half-plane, and [Q(α):Q] is the degree of the number field Q(α).
This result is due to Theodor Schneider.[5]
• ,
• This result is a corollary of the Gelfond–Schneider theorem which says that if α is algebraic and not 0 or 1, and
if β is algebraic and irrational then αβ is transcendental. Thus the function 2x could be replaced by cx for any
algebraic c not equal to 0 or 1. Indeed, we have:

• A consequence of Schanuel's conjecture in transcendental number theory would be that
• A function with empty exceptional set that doesn't require one to assume this conjecture is the function
ƒ(x) = exp(1 + πx).
While calculating the exceptional set for a given function is not easy, it is known that given any subset of the
algebraic numbers, say A, there is a transcendental function ƒ whose exceptional set is A.[6] Since, as mentioned
above, this includes taking A to be the whole set of algebraic numbers, there is no way to determine if a function is
transcendental just by looking at its values at algebraic numbers. In fact, Alex Wilkie showed that the situation is
even worse: he constructed a transcendental function ƒ from R to R that is analytic everywhere but whose
transcendence cannot be detected by any first-order method.[7]
Transcendental function 152

References
[1] E. J. Townsend, Functions of a Complex Variable, BiblioLife, LLC, (2009).
[2] M. Waldschmidt, Diophantine approximation on linear algebraic groups, Springer (2000).
[3] D. Marques, F. M. S. Lima, Some transcendental functions that yield transcendental values for every algebraic entry, (2010)
arXiv:1004.1668v1.
[4] N. Archinard, Exceptional sets of hypergeometric series, Journal of Number Theory 101 Issue 2 (2003), pp.244–269.
[5] T. Schneider, Arithmetische Untersuchungen elliptischer Integrale, Math. Annalen 113 (1937), pp.1–13.
[6] M. Waldschmidt, Auxiliary functions in transcendental number theory, The Ramanujan journal 20 no3, (2009), pp.341–373.
[7] A. Wilkie, An algebraically conservative, transcendental function, Paris VII preprints, number 66, 1998.

Unary function
A unary function is a function that takes one argument. In computer science, a unary operator is a subset of unary
function.
Many of the elementary functions are unary functions, in particular the trigonometric functions and hyperbolic
function are unary.

References
• Foundations of Genetic Programming [1]

References
[1] http:/ / www. cs. ucl. ac. uk/ staff/ W. Langdon/ FOGP

Univalent function
In mathematics, in the branch of complex analysis, a holomorphic function on an open subset of the complex plane is
called univalent if it is one-to-one.

Examples
Any mapping of the open unit disc to itself, : where is univalent.

Basic properties
One can prove that if and are two open connected sets in the complex plane, and

is a univalent function such that (that is, is onto), then the derivative of is never zero, is
invertible, and its inverse is also holomorphic. More, one has by the chain rule

for all in
Univalent function 153

Comparison with real functions


For real analytic functions, unlike for complex analytic (that is, holomorphic) functions, these statements fail to hold.
For example, consider the function

given by ƒ(x) = x3. This function is clearly one-to-one, however, its derivative is 0 at x = 0, and its inverse is not
analytic, or even differentiable, on the whole interval (−1, 1).

References
• John B. Conway. Functions of One Complex Variable I. Springer-Verlag, New York, 1978. ISBN 0-387-90328-3.
• John B. Conway. Functions of One Complex Variable II. Springer-Verlag, New York, 1996. ISBN
0-387-94460-5.

This article incorporates material from univalent analytic function on PlanetMath, which is licensed under the
Creative Commons Attribution/Share-Alike License.

Vector-valued function
A vector-valued function also referred to
as a vector function is a mathematical
function of one or more variables whose
range is a set of multidimensional vectors or
infinite-dimensional vectors. Often the input
of a vector-valued function is a scalar, but in
general the input can be a vector of both
complex or real variables.

A graph of the vector-valued function r(t) = <2 cos t, 4 sin t, t> indicating a range
of solutions and the vector when evaluated near t = 19.5
Vector-valued function 154

Example
A common example of a vector valued function is one that depends on a single real number parameter t, often
representing time, producing a vector v(t) as the result. In terms of the standard unit vectors i, j, k of Cartesian
3-space, these specific type of vector-valued functions are given by expressions such as
• or

where f(t), g(t) and h(t) are the coordinate functions of the parameter t. The vector r(t) has its tail at the origin and
its head at the coordinates evaluated by the function.
The vector shown in the graph to the right is the evaluation of the function near t=19.5 (between 6π and 6.5π; i.e.,
somewhat more than 3 rotations). The spiral is the path traced by the tip of the vector as t increases from zero
through 8π.
Vector functions can also be referred to in a different notation:
• or

Properties
The domain of a vector-valued function is the intersection of the domain of the functions f, g, and h.

Derivative of a three-dimensional vector function


Many vector-valued functions, like scalar-valued functions, can be differentiated by simply differentiating the
components in the Cartesian coordinate system. Thus, if

is a vector-valued function, then

The vector derivative admits the following physical interpretation: if r(t) represents the position of a particle, then
the derivative is the velocity of the particle

Likewise, the derivative of the velocity is the acceleration

Partial derivative
The partial derivative of a vector function a with respect to a scalar variable q is defined as[1]

where ai is the scalar component of a in the direction of ei. It is also called the direction cosine of a and ei or their dot
product. The vectors e1,e2,e3 form an orthonormal basis fixed in the reference frame in which the derivative is being
taken.
Vector-valued function 155

Ordinary derivative
If a is regarded as a vector function of a single scalar variable, such as time t, then the equation above reduces to the
first ordinary time derivative of a with respect to t,[1]

Total derivative
If the vector a is a function of a number n of scalar variables qr (r = 1,...,n), and each qr is only a function of time t,
then the ordinary derivative of a with respect to t can be expressed, in a form known as the total derivative, as[1]

Some authors prefer to use capital D to indicate the total derivative operator, as in D/Dt. The total derivative differs
from the partial time derivative in that the total derivative accounts for changes in a due to the time variance of the
variables qr.

Reference frames
Whereas for scalar-valued functions there is only a single possible reference frame, to take the derivative of a
vector-valued function requires the choice of a reference frame (at least when a fixed Cartesian coordinate system is
not implied as such). Once a reference frame has been chosen, the derivative of a vector-valued function can be
computed using techniques similar to those for computing derivatives of scalar-valued functions. A different choice
of reference frame will, in general, produce a different derivative function. The derivative functions in different
reference frames have a specific kinematical relationship.

Derivative of a vector function with nonfixed bases


The above formulas for the derivative of a vector function rely on the assumption that the basis vectors e1,e2,e3 are
constant, that is, fixed in the reference frame in which the derivative of a is being taken, and therefore the e1,e2,e3
each has a derivative of identically zero. This often holds true for problems dealing with vector fields in a fixed
coordinate system, or for simple problems in physics. However, many complex problems involve the derivative of a
vector function in multiple moving reference frames, which means that the basis vectors will not necessarily be
constant. In such a case where the basis vectors e1,e2,e3 are fixed in reference frame E, but not in reference frame N,
the more general formula for the ordinary time derivative of a vector in reference frame N is[1]

where the superscript N to the left of the derivative operator indicates the reference frame in which the derivative is
taken. As shown previously, the first term on the right hand side is equal to the derivative of a in the reference frame
where e1,e2,e3 are constant, reference frame E. It also can be shown that the second term on the right hand side is
equal to the relative angular velocity of the two reference frames cross multiplied with the vector a itself.[1] Thus,
after substitution, the formula relating the derivative of a vector function in two reference frames is[1]

where NωE is the angular velocity of the reference frame E relative to the reference frame N.
One common example where this formula is used is to find the velocity of a space-borne object, such as a rocket, in
the inertial reference frame using measurements of the rocket's velocity relative to the ground. The velocity NvR in
inertial reference frame N of a rocket R located at position rR can be found using the formula
Vector-valued function 156

where NωE is the angular velocity of the Earth relative to the inertial frame N. Since velocity is the derivative of
position, NvR and EvR are the derivatives of rR in reference frames N and E, respectively. By substitution,

where EvR is the velocity vector of the rocket as measured from a reference frame E that is fixed to the Earth.

Derivative and vector multiplication


The derivative of the products of vector functions behaves similarly to the derivative of the products of scalar
functions.[2] Specifically, in the case of scalar multiplication of a vector, if p is a scalar variable function of q,[1]

In the case of dot multiplication, for two vectors a and b that are both functions of q,[1]

Similarly, the derivative of the cross product of two vector functions is[1]

Derivative of an n-dimensional vector function


A function f of a real number t with values in the space can be written as .
Its derivative equals
.
If f is a function of several variables, say of , then the partial derivatives of the components of f form a
matrix called the Jacobian matrix of f.

Infinite-dimensional vector functions


If the values of a function f lie in an infinite-dimensional vector space X, such as a Hilbert space, then f may be called
an infinite-dimensional vector function.

Functions with values in a Hilbert space


If the argument of f is a real number and X is a a Hilbert space, then the derivative of f at a point t can be defined as
in the finite-dimensional case:

Most results of the finite-dimensional case also hold in the infinite-dimensional case too, mutatis mutandis.
Differentiation can also be defined to functions of several variables (e.g., or even , where Y is an
infinite-dimensional vector space).
N.B. If X is a Hilbert space, then one can easily show that any derivative (and any other limit) can be computed
componentwise: if

(i.e., , where is an orthonormal basis of the space X), and


exists, then
Vector-valued function 157

.
However, the existence of a componentwise derivative does not guarantee the existence of a derivative, as
componentwise convergence in a Hilbert space does not guarantee convergence with respect to the actual topology
of the Hilbert space.

Other infinite-dimensional vector spaces


Most of the above hold for other topological vector spaces X too. However, not as many classical results hold in the
Banach space setting, e.g., an absolutely continuous function with values in a suitable Banach space need not have a
derivative anywhere. Moreover, in most Banach spaces setting there are no orthonormal bases.

Notes
[1] Kane & Levinson 1996, p. 29–37
[2] In fact, these relations are derived applying the product rule componentwise.

References
• Kane, Thomas R.; Levinson, David A. (1996), "1-9 Differentiation of Vector Functions", Dynamics Online,
Sunnyvale, California: OnLine Dynamics, Inc., pp. 29–37

External links
• Vector-valued functions and their properties (from Lake Tahoe Community College) (http://ltcconline.net/
greenl/courses/202/vectorFunctions/vectorFunctions.htm)
• Weisstein, Eric W., " Vector Function (http://mathworld.wolfram.com/VectorFunction.html)" from
MathWorld.
• Everything2 article (http://www.everything2.com/index.pl?node_id=1525585)
• 3 Dimensional vector-valued functions (from East Tennessee State University) (http://math.etsu.edu/
MultiCalc/Chap1/Chap1-6/part1.htm)
Weakly harmonic function 158

Weakly harmonic function


In mathematics, a function is weakly harmonic in a domain if

for all with compact support in and continuous second derivatives, where Δ is the Laplacian. This is the same
notion as a weak derivative, however, a function can have a weak derivative and not be differentiable. In this case,
we have the somewhat surprising result that a function is weakly harmonic if and only if it is harmonic. Thus weakly
harmonic is actually equivalent to the seemingly stronger harmonic condition.

Weakly measurable function


In mathematics — specifically, in functional analysis — a weakly measurable function taking values in a Banach
space is a function whose composition with any element of the dual space is a measurable function in the usual
(strong) sense. For separable spaces, the notions of weak and strong measurability agree.

Definition
If (X, Σ) is a measurable space and B is a Banach space over a field K (usually the real numbers R or complex
numbers C), then f : X → B is said to be weakly measurable if, for every continuous linear functional g : B → K,
the function

is a measurable function with respect to Σ and the usual Borel σ-algebra on K.

Properties
The relationship between measurability and weak measurability is given by the following result, known as Pettis'
theorem or Pettis measurability theorem.
A function f is said to be almost surely separably valued (or essentially separably valued) if there
exists a subset N ⊆ X with μ(N) = 0 such that f(X \ N) ⊆ B is separable.
Theorem (Pettis). A function f : X → B defined on a measure space (X, Σ, μ) and taking values in a
Banach space B is (strongly) measurable (with respect to Σ and the Borel σ-algebra on B) if and only if it
is both weakly measurable and almost surely separably valued.
In the case that B is separable, since any subset of a separable Banach space is itself separable, one can take N above
to be empty, and it follows that the notions of weak and strong measurability agree when B is separable.

References
• Showalter, Ralph E. (1997). "Theorem III.1.1". Monotone operators in Banach space and nonlinear partial
differential equations. Mathematical Surveys and Monographs 49. Providence, RI: American Mathematical
Society. p. 103. MR1422252. ISBN 0-8218-0500-2..
Article Sources and Contributors 159

Article Sources and Contributors


Additive function  Source: http://en.wikipedia.org/w/index.php?oldid=401216309  Contributors: AlainD, Algebraist, Andres, AxelBoldt, Burn, CRGreathouse, Camembert, DYLAN LENNON,
David Eppstein, Drbreznjev, Dysprosia, Fudo, Gandalf61, Gfis, Giftlite, LOL, Laurentius, Linas, Maksim-e, Memmis, Michael Hardy, Ozob, PV=nRT, Paul August, Petrb, Schneelocke,
Soliloquial, Tango, Tannkrem, TedPavlic, Tonigonenstein, Woodstone, XJamRastafire, 23 anonymous edits

Algebraic function  Source: http://en.wikipedia.org/w/index.php?oldid=375669198  Contributors: Almit39, BenWhitey, Bethling, CBM, Charles Matthews, Ciphers, Cronholm144, Dekimasu,
Dugwiki, Giftlite, GrAfFiT, Haukurth, JackSchmidt, Johnbibby, Kai Su?, Krashlandon, Ksnow, MarSch, Mattsem, Mets501, Michael Hardy, MuthuKutty, PV=nRT, Pomte, Silly rabbit, Sławomir
Biała, VictorAnyakin, Wisems, Wmahan, 33 anonymous edits

Analytic function  Source: http://en.wikipedia.org/w/index.php?oldid=407991013  Contributors: Albmont, Algebraist, Aliotra, BigJohnHenry, Blackcloak, CRGreathouse, Centrx, Charles
Matthews, Cheeser1, Cronholm144, CryptoDerk, Dan Gluck, Domitori, Dratman, Drilnoth, Erast, Fakhredinblog, Gauge, Gene Ward Smith, Geropod, Gesslein, Giftlite, Holmansf, IronGargoyle,
Jitse Niesen, Joriki, Kiefer.Wolfowitz, King Bee, Linas, Madmath789, Marc van Leeuwen, McLaurin, Michael Hardy, Mikewax, Mnemo, Musiphil, Ntmatter, Oleg Alexandrov, Oliphaunt,
Omegatron, PMajer, PV=nRT, Petri Krohn, Phys, Point-set topologist, Por.pl, Quietbritishjim, Rigadoun, Ryang316, Saccade, Shadowjams, Shotwell, Spireguy, Stephen Bain, Sullivan.t.j, T
boyd, TakuyaMurata, Tbsmith, The Diagonal Prince, Thenub314, TimBentley, Tobias Bergemann, Tosha, Vanished User 0001, Vivacissamamente, Weialawaga, WriterHound, Yamamoto Ichiro,
Zygmuund, 74 anonymous edits

Antiholomorphic function  Source: http://en.wikipedia.org/w/index.php?oldid=387764502  Contributors: Altenmann, Charles Matthews, Dysprosia, Frencheigh, Gaius Cornelius, Konradek,
Lethe, Oleg Alexandrov, PV=nRT, Rvollmert, 3 anonymous edits

Arithmetic function  Source: http://en.wikipedia.org/w/index.php?oldid=393996418  Contributors: Anita5192, AxelBoldt, Burn, CBM, CRGreathouse, Colonies Chris, Conversion script, David
Shay, Dr Dec, Drilnoth, Fr3aki, Fredrik, Gandalf61, Georg Muntingh, Giftlite, Gubbubu, Haham hanuka, Jim.belk, Jkelleyy, Linas, Madmath789, Marek69, Mikhail Dvorkin, Mscalculus, Olaf,
Oleg Alexandrov, PV=nRT, Phys, RedWolf, Richard L. Peterson, RobHar, Tomaxer, Uncia, Virginia-American, WillowW, XJamRastafire, 22 anonymous edits

Bijection  Source: http://en.wikipedia.org/w/index.php?oldid=400273630  Contributors: ABCD, AbcXyz, Adrianwn, Alberto da Calvairate, Ash4Math, AxelBoldt, Baudway, Bevo, Biagioli, Bill
Malloy, Bkell, Bongoman666, Bwoodacre, Charles Matthews, Classicalecon, Cobaltcigs, Conversion script, Dallashan, Damian Yerrick, David Shay, Dcoetzee, Domitori, Dreadstar, Dreftymac,
Dysprosia, Ed g2s, FactChecker1199, Fredrik, GTBacchus, GaborLajos, Giftlite, Glenn, Gregbard, Hans Adler, Hawthorn, Hilverd, I Spel Good, Ignignot, JRSpriggs, JackSchmidt, Jan Hidders,
Johnfuhrmann, Jorge Stolfi, Karada, Kevin Lamoreau, Kilva, MarSch, MarcelB612, MarkSweep, Martynas Patasius, Meisam, Mets501, Mhym, MiNombreDeGuerra, Michael Hardy, Michael
Slone, MisterSheik, Mmernex, Nickj, Obradovic Goran, Octahedron80, PV=nRT, Panzer raccoon!, Paolo.dL, Paul August, Pearle, Peiresc, Pizza Puzzle, Poor Yorick, Quistnix, RDBury, RexNL,
Rob Hooft, Salgueiro, Salix alba, Salty-horse, Sam Staton, Schapel, Smaug123, Subversive.sound, Tarquin, Tobias Bergemann, Toby Bartels, Tsemii, UnicornTapestry, VKokielov,
Vivacissamamente, Wshun, XJamRastafire, Yomcat, Александър, 77 anonymous edits

Binary function  Source: http://en.wikipedia.org/w/index.php?oldid=394617373  Contributors: ABCD, Abdull, Aleph4, Andres, Avaragado, Charles Matthews, CommonsDelinker, Conversion
script, Damian Yerrick, Fresheneesz, Herbee, Jan Hidders, Josh Parris, Oleg Alexandrov, PV=nRT, Rpresser, Tarquin, Toby Bartels, WhisperToMe, Wile E. Heresiarch, 13 anonymous edits

Bochner measurable function  Source: http://en.wikipedia.org/w/index.php?oldid=371451842  Contributors: Michael Hardy, Parodi

Bounded function  Source: http://en.wikipedia.org/w/index.php?oldid=354910395  Contributors: Aleph4, Amire80, Calle, Docu, DoubleBlue, Giftlite, Haihe, Jtkiefer, Kompik, Lechatjaune,
Maksim-e, MathMartin, Minvogt, Mormegil, Mpd1989, NoirNoir, Oleg Alexandrov, PV=nRT, Patrick, Phys, Salix alba, Tabletop, Toby Bartels, 14 anonymous edits

Cauchy-continuous function  Source: http://en.wikipedia.org/w/index.php?oldid=345497130  Contributors: CBM, PV=nRT, Toby Bartels

Closed convex function  Source: http://en.wikipedia.org/w/index.php?oldid=384851892  Contributors: Charles Matthews, Kiefer.Wolfowitz, Oleg Alexandrov, PV=nRT, Silverfish, Tobias
Bergemann, 3 anonymous edits

Coarse function  Source: http://en.wikipedia.org/w/index.php?oldid=376524842  Contributors: Hqb, Justin W Smith, MC10, Michael Hardy

Completely multiplicative function  Source: http://en.wikipedia.org/w/index.php?oldid=254308572  Contributors: Andy.melnikov, Dcoetzee, Giftlite, JackSchmidt, Michael Hardy, PV=nRT,
Rich Farmbrough, Vanish2, Virginia-American

Concave function  Source: http://en.wikipedia.org/w/index.php?oldid=408646081  Contributors: Aholtman, Aisaac, Aleph4, Altenmann, Andre Engels, Andykoo1990, Audacity, Charles
Matthews, Dreadstar, Dyaka, Econn, Edemaine, Flavio Guitian, Fredrik, Gamesou, Gene Nygaard, Giftlite, Greg Kuperberg, J heisenberg, Jim.belk, Johngcarlsson, Malin84, MapsMan, Mets501,
Michael Hardy, Nandhp, NeoUrfahraner, Noe, O18, Oleg Alexandrov, Orientalhope, PV=nRT, Pgan002, Pizza Puzzle, RDBury, Robinh, Rridlon, Salgueiro, SirJective, SirPeebles, Stebulus,
StudierMalMarburg, TakuyaMurata, The Gnome, The Scarlet Letter, Thuytnguyen48, Tmy1018, Tobacman, Tosha, ZeroOne, Zoicon5, 35 anonymous edits

Constant function  Source: http://en.wikipedia.org/w/index.php?oldid=387709763  Contributors: Abdull, AndrewHowse, CRGreathouse, David Shay, Dpv, Evil saltine, Giftlite, Glenn, JaGa,
Khakbaz, Kompik, Mhaitham.shammaa, Mormegil, Octahedron80, Oleg Alexandrov, PV=nRT, Paul August, Pred, Prophile, TakuyaMurata, Xario, Zero Thrust, ‫ןריל‬, 25 anonymous edits

Continuous function  Source: http://en.wikipedia.org/w/index.php?oldid=407367738  Contributors: 213.253.39.xxx, ABCD, AdamSmithee, Aetheling, Ams80, Andywall, Ap, Army1987,
Arthena, Arthur Rubin, Ashted, AxelBoldt, Bdmy, BenKovitz, Bethnim, Bloodshedder, CRGreathouse, Charles Matthews, Cheeser1, Cic, Conversion script, Ctmt, D.M. from Ukraine, Dallashan,
Darth Panda, Dcoetzee, DomenicDenicola, Domitori, Dr.K., Dysprosia, EWikist, EdC, Edemaine, Error792, Evilchicken1234, Fabartus, Felix Wiemann, Fgnievinski, Fiedorow, Fresheneesz,
Giftlite, Glenn, Gombang, Graham87, Grinevitski, Gthb, Harriv, Henry Delforn, Hqb, HyDeckar, Hyacinth, Iameukarya, Ian Pitchford, Igiffin, Igrant, Intangir, Isomorphic, Iulianu, Jacj, JahJah,
Jim.belk, Jimp, Jitse Niesen, Joseaperez, Jrtayloriv, Jshadias, K-UNIT, Katzmik, Klutzy, Kompik, LachlanA, Lambiam, Larryisgood, Lee Larson, Leoremy, Linas, Lupin, MC10, MSGJ, Markus
Krötzsch, MathMartin, Mdd, Michael Hardy, Mikez, Monkey 32606, Mormegil, Mplourde, Msh210, Musicpvm, NawlinWiki, Nbarth, Newone, Oleg Alexandrov, PV=nRT, Paul August, Pdn,
Penumbra2000, Pillcrow, Pizza Puzzle, QYV, Qz, RDBury, Ramzzhakim, Rbb l181, Reach Out to the Truth, Rhetth, Rick Norwood, Rinconsoleao, Roman3, Sabbut, Salgueiro, Sapphic, Sbacle,
Schneelocke, Seb35, Sligocki, Smmurphy, Splarka, Stan Lioubomoudrov, Stca74, Stevenj, StradivariusTV, Sullivan.t.j, Svick, T00h00, TedPavlic, Template namespace initialisation script,
Thehotelambush, Thenub314, Thierry Caro, Tiagofassoni, Timhoooey, Tkuvho, Tlevine, Tobias Bergemann, Toby, Tosha, Tuxedo junction, Ulipaul, Ultramarine, Wolfrock, Wshun, Xantharius,
Yacht, Youandme, Zoicon5, Zundark, ZyMOS, 139 anonymous edits

Convex function  Source: http://en.wikipedia.org/w/index.php?oldid=407844558  Contributors: Abhimanyulad, Aleph4, Altenmann, Anonymous Dissident, Arthur Rubin, Bender2k14, Bh3u4m,
Brianjd, Buettcher, CSTAR, Chicocvenancio, Closedmouth, Comfortably Paranoid, Connelly, Dattorro, Dchudz, Derbeth, Dpbert, Dwmalone, Easwaran, Eli Osherovich, Flavio Guitian, Fred
Bauder, Gaius Cornelius, Ged.R, Giftlite, Greg Kuperberg, Hrafeiro, Hu12, Ian Pitchford, Isheden, JF Manning, Jheiv, Jim.belk, Johngcarlsson, Kiefer.Wolfowitz, Kummi, LachlanA, Lavaka,
Lechatjaune, Madmath789, Magister Mathematicae, Maldavir, Mennucc, Michael Hardy, MrOllie, Mrsaad31, NeoUrfahraner, Noe, Oleg Alexandrov, Oli Filth, Oysterofamerica, PV=nRT,
Pgan002, Populus, Rar, Rjwilmsi, Rphb, Salvatore Ingala, Sgorg10, Shreevatsa, Small potato, Stan Lioubomoudrov, Sullivan.t.j, TakuyaMurata, TedPavlic, The Anome, Tobacman, Tobias
Bergemann, Val001, Van helsing, Willking1979, Wlod, 91 anonymous edits

Differentiable function  Source: http://en.wikipedia.org/w/index.php?oldid=406525113  Contributors: 1l2, Jim.belk, Kusluj, Mormegil, StAnselm, Toby Bartels, 5 anonymous edits

Doubly periodic function  Source: http://en.wikipedia.org/w/index.php?oldid=367796267  Contributors: Charles Matthews, DavidCBryant, Gandalf61, Kupirijo, Michael Hardy,
MichaelShoemaker, Ozob, PV=nRT, R.e.b., Singularity, Wavelength, 5 anonymous edits

Elementary function  Source: http://en.wikipedia.org/w/index.php?oldid=406249107  Contributors: 2help, Akriasas, Albmont, Aldaron, Andres, Anonymous Dissident, Appzter, Conversion
script, DMacks, Domitori, Dratman, Eivindgh, Falcor84, Fredrik, Giftlite, Haham hanuka, Indeed123, JJL, Jason Quinn, Kostmo, Kusunose, Kuszi, LokiClock, Magnus Manske, MarSch,
MathFacts, Mercenario97, Mhym, Michael Hardy, Michael Slone, Mormegil, Oleg Alexandrov, PV=nRT, Phil Boswell, R.e.b., Rbk, Rogerbrent, Sligocki, Stefano.gogioso, Stever Augustus,
StradivariusTV, Timwi, Udalov, Wvbailey, XJamRastafire, XaosBits, 23 anonymous edits

Elliptic function  Source: http://en.wikipedia.org/w/index.php?oldid=404610044  Contributors: A. Pichler, Alansohn, Almit39, AxelBoldt, CRGreathouse, Cenarium, Charles Matthews, CiaPan,
Cptheorist, DavidCBryant, Doetoe, Dysprosia, Eric Kvaalen, Fenice, Fredrik, Gene Ward Smith, Giftlite, Jemebius, La goutte de pluie, Linas, Looxix, Michael Hardy, Netrapt, Nic bor, PV=nRT,
PierreAbbat, Puuropyssy, R.e.b., Robinh, Salgueiro, Stepp-Wulf, Tbr00, TomyDuby, Twanvl, Wile E. Heresiarch, William Ackerman, XJamRastafire, 30 anonymous edits

Empty function  Source: http://en.wikipedia.org/w/index.php?oldid=368572104  Contributors: Aleph4, Altenmann, CBM, Charles Matthews, Devoutb3nji, EdC, FactSpewer, Fredrik, Fropuff,
Lildoodle, Melchoir, Mets501, Ms2ger, Oleg Alexandrov, PV=nRT, Patrick, Paul August, Ruud Koot, Salix alba, Sionus, SlamDiego, The Anome, WojciechSwiderski, Zundark, 6 anonymous
Article Sources and Contributors 160

edits

Entire function  Source: http://en.wikipedia.org/w/index.php?oldid=398031592  Contributors: Alberto da Calvairate, Arjarj, AxelBoldt, Conversion script, David Newton, Dojarca, ERcheck,
EdJohnston, El C, EmilJ, EoGuy, Eric Kvaalen, Giftlite, Gogobera, J.delanoy, Kevinatilusa, Lasserempe, Lethe, Madmath789, Mattbuck, Michael Hardy, Mon4, Oleg Alexandrov, Oliphaunt,
PMajer, PV=nRT, Randomblue, RedWolf, Rhythm, Robin S, Robinh, Tarquin, Tobias Bergemann, 25 anonymous edits

Even and odd functions  Source: http://en.wikipedia.org/w/index.php?oldid=392025820  Contributors: 16@r, Albedo, Anskas, CALR, CecilBlade, Charles Matthews, Cremepuff222, Dbfirs,
Doctormatt, Dogah, Download, Dysprosia, El aprendelenguas, Fayenatic london, Gandalf61, GeordieMcBain, Giftlite, Ih8evilstuff, Johnuniq, KPH2293, KYN, KYPark, Kilva, Laurentius, Lixy,
MagneticFlux, Marek69, Marino-slo, Melchoir, Mets501, Michael Devore, Michael Hardy, Mochi, Mormegil, Oleg Alexandrov, Omegatron, PV=nRT, Paul August, Pgan002, Pizza1512,
Plastikspork, Rbj, Revolver, RobHar, Salgueiro, Six.oh.six, Spoon!, Staka, Taejo, TeleComNasSprVen, Tobias Bergemann, Tomruen, WOSlinker, Yoshigev, Zvika, გიგა, 121 anonymous edits

Flat function  Source: http://en.wikipedia.org/w/index.php?oldid=401351533  Contributors: Fly by Night, Gandalf61, Michael Hardy, Rich Farmbrough, 2 anonymous edits

Function of a real variable  Source: http://en.wikipedia.org/w/index.php?oldid=332225156  Contributors: Aisaac, Charles Matthews, PV=nRT, 1 anonymous edits

Function composition  Source: http://en.wikipedia.org/w/index.php?oldid=405641133  Contributors: Adam majewski, Anonymous Dissident, CBM, Charles Matthews, Cherry Cotton,
Classicalecon, Constructive editor, DA3N, Danakil, David Eppstein, Dcljr, Dfass, Dmh, EdC, EmilJ, Fallenness, Fuzzybyte, Gcm, Georgia guy, Giftlite, Glenn, Googl, Greenrd, GregorB,
Grubber, Jason Quinn, Jon Awbrey, Jonnyappleseed24, Juansempere, J•A•K, Karl Dickman, Kku, Kuteni, Lethe, LilHelpa, Linas, MFH, Maksim-e, Marino-slo, MattGiuca, Melchoir, Michael
Hardy, Nbarth, Netrapt, Oleg Alexandrov, PV=nRT, Patrick, Paul August, Pcap, Phil Boswell, Phys, Pit-trout, Plastikspork, Pleasantville, Qwertyus, Rasmus Faber, Rheun, SixWingedSeraph,
Slac, Strangelv, TakuyaMurata, Tarquin, Tobias Bergemann, VKokielov, Vonkje, Woohookitty, Wshun, XudongGuan, Zenkat, Zundark, 64 anonymous edits

Functional (mathematics)  Source: http://en.wikipedia.org/w/index.php?oldid=405548639  Contributors: 3mta3, Ae-a, Baterista, Bedivere, Belizefan, BenKovitz, Cícero, Dr. Universe,
Eaglizard, Faramarz.M, Giftlite, Hammertime, Helohe, Igny, JJ Harrison, JSpudeman, JamesBWatson, Jpkotta, Linas, Lzur, Michael Hardy, Mkirsten, Mpatel, Oleg Alexandrov, Oliphaunt,
OoberMick, Outofmine, Poopface001, Qwertyus, R'n'B, Rgrg, RobHar, Salgueiro, Saravask, TheObtuseAngleOfDoom, Thurth, Zenohockey, 29 anonymous edits

Harmonic function  Source: http://en.wikipedia.org/w/index.php?oldid=409111218  Contributors: 4C, Almit39, Aminrahimian, AvicAWB, AxelBoldt, Bowiki, Brian Tvedt, Bryan Derksen,
Camw, Charles Matthews, Ciphergoth, Crowsnest, DavidCBryant, Dysprosia, El C, Elipongo, Giftlite, Greenpickle, Grubber, Guardian of Light, Hanleywashington, Hyacinth, Jerzy, Jim.belk,
Linas, Mathanor, Michael Hardy, Morning277, Nuwewsco, Oleg Alexandrov, PAR, PMajer, PV=nRT, Patrick, Perturbationist, Pjacobi, RayAYang, Red Act, Saihtam, Shoofle, Sławomir Biały,
Tbsmith, The Anome, Tim Starling, Tiphareth, Tparameter, Willking1979, Xpi6, 45 anonymous edits

Hermitian function  Source: http://en.wikipedia.org/w/index.php?oldid=406821504  Contributors: Auntof6, BenFrantzDale, BiH, CBM, Dmmaus, Dzordzm, JCSantos, Jim.belk, KYN, Michael
Hardy, Rbj, 16 anonymous edits

Holomorphic function  Source: http://en.wikipedia.org/w/index.php?oldid=401488926  Contributors: 212.242.115.xxx, Aaronbrick, Abiola Lapite, Acepectif, Adoniscik, Almit39, Altenmann,
AxelBoldt, Bdmy, Ben pcc, CRGreathouse, Charles Matthews, Christian.Mercat, Conversion script, CrniBombarder!!!, Crust, Damian Yerrick, DavidCBryant, DomenicDenicola, Dougher,
DragonflySixtyseven, Dratman, Duckbill, Dysprosia, Dzordzm, Eequor, Email4mobile, EmilJ, Fakhredinblog, FilipeS, Frencheigh, FuriousScribble, Giftlite, Graham87, HannsEwald, Hesam7,
Irigi, JamesBWatson, Jao, Jim.belk, Jmath666, Jobh, Jusjih, KnightRider, Laurent MAYER, Lethe, Lhf, Linas, Lunch, Maxim Razin, Michael Hardy, Michael K. Edwards, Mike40033, Naddy,
NickBush24, ObsessiveMathsFreak, Oleg Alexandrov, PV=nRT, Patrick, PierreAbbat, Prim Ethics, Reaverdrop, Robert Illes, Rvollmert, Saleemsan, Salix alba, Silly rabbit, Sligocki, Stephen
Bain, TakuyaMurata, Tarquin, Tcnuk, Tetracube, The Diagonal Prince, Thorfinn, Tobias Hoevekamp, Unco, Vanished User 0001, Weierstraß, XJamRastafire, 44 anonymous edits

Homogeneous function  Source: http://en.wikipedia.org/w/index.php?oldid=408061009  Contributors: Adiel lo, AlekseyP, Aleph4, Anonymous Dissident, CBM, Charles Matthews,
Crasshopper, Dangla, Duoduoduo, Giftlite, Jitse Niesen, K madhukar, Kauczuk, Lantonov, Linas, Linus500, MFH, Maksim-e, Mathgroves, Mazi, Michael Hardy, PV=nRT, Sbyrnes321,
Sławomir Biały, TakuyaMurata, The Anome, Tobias Bergemann, Yahel Guhan, 31 anonymous edits

Identity function  Source: http://en.wikipedia.org/w/index.php?oldid=383985348  Contributors: Andy Dingley, AxelBoldt, Bo Jacoby, Cenarium, Charles Matthews, Conversion script,
Dreftymac, Fredrik, Gcm, Giftlite, Glenn, Kilva, Marcos, Marino-slo, Melchoir, Numbo3, Octahedron80, Oleg Alexandrov, PV=nRT, Paine Ellsworth, Patrick, Paul August, Pred, Rick Norwood,
Rubicon, Ruud Koot, Salix alba, Sam Staton, Ssd, TakuyaMurata, Terry Bollinger, XJamRastafire, Zundark, 16 anonymous edits

Implicit and explicit functions  Source: http://en.wikipedia.org/w/index.php?oldid=392163603  Contributors: ABCD, Albmont, Amazins490, Anonymous Dissident, Arthena, Baccyak4H,
BenFrantzDale, BiT, Brianjd, Charles Matthews, Chinju, ChopinStudent, ChrisChiasson, DARTH SIDIOUS 2, Deliou, Dfeuer, Download, Dysprosia, ENIAC, ESkog, Ed Cormany,
Email4mobile, Espressobongo, Estudiarme, Evil saltine, Faramir1138, Feinstein, Fgnievinski, Francisco Quiumento, Frpcad, Giftlite, Gvozdet, Igor Yalovecky, J04n, JohnOwens, Kareeser,
Kenneth M Burke, KurtRaschke, LVC, LestatdeLioncourt, Michael Hardy, Mikiemike, Morten, Nbarth, Nonagonal Spider, Odie5533, Oleg Alexandrov, PV=nRT, Pcap, Pizza Puzzle, Point-set
topologist, PseudoSudo, RDBury, Radagast83, Rigadoun, Rjwilmsi, RobHar, Ryan Reich, Saibod, Salix alba, Schissel, Silly rabbit, Simonloach, Sir Dagon, Spireguy, Spoon!, Stewartadcock,
Stillnotelf, Sławomir Biały, TStein, Template namespace initialisation script, Tobias Bergemann, Tosha, Ulner, Waltpohl, Yacht, Yuliyag, 101 anonymous edits

Indicator function  Source: http://en.wikipedia.org/w/index.php?oldid=409783741  Contributors: Albmont, ArnoldReinhold, AxelBoldt, Banus, BiT, Bo Jacoby, CSTAR, Centrx, Charles
Matthews, Colinvella, David Eppstein, DavidHouse, Dcoetzee, Digby Tantrum, Falcor84, Gauss, Giftlite, Helgus, Ht686rg90, Icairns, Jon Awbrey, LBehounek, Linas, Martpol, MathMartin,
Melcombe, Mets501, Michael Hardy, Mormegil, Mpd1989, NickBush24, Obradovic Goran, Octahedron80, Oleg Alexandrov, PV=nRT, Paintman, Paul August, Pcb21, Phil Boswell, Piotrus,
Quietbritishjim, Raystorm, Rjwilmsi, Rosh3000, Salix alba, Sherbrooke, Small potato, Sullivan.t.j, The Anome, Trovatore, Wvbailey, Xetrov, 22 anonymous edits

Injective function  Source: http://en.wikipedia.org/w/index.php?oldid=408026705  Contributors: 16@r, ABCD, AHM, Adrianwn, Atlantia, Austinflorida, AxelBoldt, BioTube, Bob.v.R,
Bookandcoffee, CBM, Cassivs, Cfailde, Charles Matthews, Chinju, Classicalecon, Cyp, Da Joe, Daniel Brockman, David Eppstein, Dbfirs, Dcljr, Dominus, Donotresus, Download, Dreadstar,
Dubhe.sk, Dysprosia, Ed g2s, Edcolins, Giftlite, Glenn, Haham hanuka, Hawk777, Hawthorn, IPonomarev, Jj137, Jorge Stolfi, Karada, Karl Dickman, Keeyu, LarryLACa, LucPereira, MIT
Trekkie, MarSch, Marc van Leeuwen, MarkSweep, Michael Hardy, Mike Fikes, MisterSheik, Naku, Numbo3, Obradovic Goran, Octahedron80, Oleg Alexandrov, PV=nRT, Paolo.dL, Paul
August, Pbroks13, Peruvianllama, Pred, Quistnix, Redclock2, Rohit math, SMP, Saippuakauppias, Sam Staton, Schapel, Skeptical scientist, TakuyaMurata, Tarquin, TechnoFaye, Tendays, Toby
Bartels, Tomhab, Tsemii, Vivacissamamente, Wshun, XJamRastafire, Xantharius, Александър, 87 anonymous edits

Invex function  Source: http://en.wikipedia.org/w/index.php?oldid=409132389  Contributors: Clausen, DRLB, Isheden, Michael Hardy, 2 anonymous edits

List of types of functions  Source: http://en.wikipedia.org/w/index.php?oldid=306021578  Contributors: PV=nRT, Tompw, 1 anonymous edits

Locally integrable function  Source: http://en.wikipedia.org/w/index.php?oldid=408447126  Contributors: Amorette, Daniele.tampieri, Hesam7, Lechatjaune, Oleg Alexandrov, PV=nRT, Salix
alba, WISo, Zvika, 5 anonymous edits

Measurable function  Source: http://en.wikipedia.org/w/index.php?oldid=401198144  Contributors: AxelBoldt, Babcockd, Bdmy, CSTAR, Calle, Charles Matthews, Chungc, CàlculIntegral,
Dan131m, Dingenis, Dysprosia, Fibonacci, Fropuff, Gala.martin, GeeJo, GiM, Giftlite, Ht686rg90, Imran, Jahredtobin, Jka02, Linas, Loewepeter, Maneesh, MarSch, Miguel, Mike Segal,
Msh210, OdedSchramm, Oleg Alexandrov, PV=nRT, Parodi, Pascal.Tesson, Paul August, Richard L. Peterson, Rinconsoleao, Ron asquith, Rs2, Salgueiro, Salix alba, Silverfish, Sławomir Biały,
Takwan, Tosha, Vivacissamamente, Walterfm, Weialawaga, Zundark, Zvika, 37 anonymous edits

Meromorphic function  Source: http://en.wikipedia.org/w/index.php?oldid=407211411  Contributors: Acepectif, Alberto da Calvairate, Altenmann, Ancheta Wis, ArmadilloFromHell,
AxelBoldt, Ben Standeven, Bertik, Ciphergoth, Clamengh, Cptheorist, Dysprosia, Eequor, El C, EmilJ, Fakhredinblog, Giftlite, H1voltage, Herbee, HorsePunchKid, Jakob.scholbach,
JamesBWatson, Joerg Winkelmann, KSmrq, LOL, Linas, Luqui, Michael Hardy, Mordacil, Oleg Alexandrov, PV=nRT, Pafinocious, Paul August, Point-set topologist, Prumpf, Robert Illes,
Romanm, Rylann, Sullivan.t.j, Tarquin, Thorfinn, Toby Bartels, Tong, Vanished User 0001, Wdvorak, 29 anonymous edits

Monotonic function  Source: http://en.wikipedia.org/w/index.php?oldid=409387749  Contributors: 478jjjz, ANDROBETA, Addshore, Albmont, Andre Engels, AppleJuggler, AxelBoldt,
BenFrantzDale, Bender2k14, Berland, CBM, Caesura, Calle, Charles Matthews, Craig t moore, DafadGoch, David Eppstein, Dino, Dugwiki, Econotechie, Edemaine, Epheterson, Fibonacci,
FreplySpang, Gavin.collins, Giftlite, Gregbard, Haham hanuka, Henrygb, Hisoka-san, Intgr, Isheden, Ixfd64, Jackzhp, Justin W Smith, Kausikghatak, Kri, Lipedia, Luqui, MFH, MSGJ, Macrakis,
Marc van Leeuwen, Markus Krötzsch, Mcld, Mhss, Michael Hardy, Miguel, Oleg Alexandrov, PV=nRT, Papppfaffe, Patrick, Paul August, Pleasantville, Qwfp, ResearchRave, Roberto.zanasi,
Saxobob, Scott Ritchie, Simeon, Smmurphy, SteinbDJ, Sullivan.t.j, Supertigerman, TakuyaMurata, Tobias Bergemann, Topology Expert, Tosha, Totalcynic, TrogdorPolitiks, Trovatore, Yecril,
58 anonymous edits

Multiplicative function  Source: http://en.wikipedia.org/w/index.php?oldid=393995369  Contributors: .mau., 212.134.20.xxx, Anton Mravcek, Arwest, Auclairde, AxelBoldt,
Baccala@freesoft.org, Bubba73, Burn, CRGreathouse, Charles Matthews, ChickenMerengo, Conversion script, Cícero, David Eppstein, Dcoetzee, Dogah, Dominus, Dysprosia, Giftlite,
Graham87, HairyFotr, Henrygb, Hoot, Jim.belk, Jshadias, Kirontanvir11, Linas, Maksim-e, MathNerd, Mhym, Miguel, Ozob, PV=nRT, PierreAbbat, Rajsekar, Rich Farmbrough, RobHar,
Article Sources and Contributors 161

Sabbut, Smithpith, Timwi, Virginia-American, XJamRastafire, Zahlentheorie, Zundark, 20 anonymous edits

Multivalued function  Source: http://en.wikipedia.org/w/index.php?oldid=403865118  Contributors: Abdull, Altenmann, Amillar, Anonymous Dissident, Ben pcc, Bggoldie, BlueGuy213, Bo
Jacoby, Bryan Derksen, CBM, Charles Matthews, Charvest, DavidCBryant, Deego, Dmcq, DrBob, EagleFan, Frazzydee, Futurebird, Gaius Cornelius, Giftlite, Gilgamesh he, Henrygb,
Isomorphic, JamesLee, Jujutacular, Kri, Megaloxantha, Michael Hardy, Mike4ty4, Nbarth, Oleg Alexandrov, PV=nRT, Padaneis, Paolo.dL, Pizza Puzzle, Pownuk, Rholton, Ruakh, Schapel,
Sverdrup, WardenWalk, Woohookitty, 23 anonymous edits

Negligible function  Source: http://en.wikipedia.org/w/index.php?oldid=396472671  Contributors: B. Wolterding, Duja, Inwind, Jiejunkong, MarioS, Michael Hardy, Nigelj, Salix alba, Samsara,
5 anonymous edits

Nowhere continuous function  Source: http://en.wikipedia.org/w/index.php?oldid=400362900  Contributors: Bkell, Charles Matthews, CiaPan, Dysprosia, Gandalf61, Gnat79, H.ehsaan,
Henrygb, Hiiiiiiiiiiiiiiiiiiiii, JJL, Kevinatilusa, MathMartin, Michael Hardy, Moberg, Netrapt, Oleg Alexandrov, PV=nRT, Patrick, Pleasantville, Ryan Reich, Salgueiro, Simetrical, Tmonzenet,
Tobias Bergemann, 21 anonymous edits

Periodic function  Source: http://en.wikipedia.org/w/index.php?oldid=409825226  Contributors: Alberto da Calvairate, Bdmy, Bob.v.R, BrokenSegue, Buzz-tardis, COMPATT, Charles
Matthews, Curtdbz, Cutler, Davepape, Dethme0w, Domitori, EagleFan, El C, FalseAxiom, Flewis, Furrykef, Giftlite, HenningThielemann, Heron, Hyacinth, Imkelleyo, J.delanoy,
JamesBWatson, Jim.belk, LachlanA, Linas, Mani1, Mathsci, Mhaitham.shammaa, Michael Hardy, Michael Tiemann, Nedim Ardoğa, Nuwewsco, Octahedron80, Oleg Alexandrov, PV=nRT,
Patrick, Paul August, PierreAbbat, Pjvpjv, Quuxplusone, Res2216firestar, Robo37, Sam Derbyshire, Sarraf.ankit, Sonett72, Sterrys, Stevenj, Stevertigo, Tarquin, Tobias Bergemann, Tomchiukc,
Trevor MacInnis, VKokielov, Wavelength, WikiDao, Xdenizen, Zaidpjd, Περίεργος, 85 anonymous edits

Piecewise linear function  Source: http://en.wikipedia.org/w/index.php?oldid=405675150  Contributors: Abeliavsky, Adam majewski, Charles Matthews, Chasingsol, CiaPan, Fgnievinski,
Fropuff, Gauge, Hulten, Jim.belk, MaxPower, Michael Hardy, Momotaro, Msh210, Nbarth, Oleg Alexandrov, PV=nRT, Pinethicket, Stebulus, Subash.chandran007, Teorth, 21 anonymous edits

Pluriharmonic function  Source: http://en.wikipedia.org/w/index.php?oldid=309414869  Contributors: Charles Matthews, Dolyn, Oleg Alexandrov, PV=nRT, Rich Farmbrough

Plurisubharmonic function  Source: http://en.wikipedia.org/w/index.php?oldid=409082648  Contributors: Charles Matthews, Hottiger, Jbaber, Michael Hardy, Oleg Alexandrov, PV=nRT,
R.e.b., Rich Farmbrough, Tiphareth, Zundark, 16 anonymous edits

Polyconvex function  Source: http://en.wikipedia.org/w/index.php?oldid=280435575  Contributors: Bigweeboy, Oleg Alexandrov, PV=nRT, Sullivan.t.j

Positive-definite function  Source: http://en.wikipedia.org/w/index.php?oldid=313690749  Contributors: Charles Matthews, Giftlite, Haseldon, J04n, Konradek, Linas, Lunch, Maksim-e, Mct
mht, Michael Hardy, Oleg Alexandrov, PV=nRT, Robinh, TedPavlic, 12 anonymous edits

Proper convex function  Source: http://en.wikipedia.org/w/index.php?oldid=389717525  Contributors: Charles Matthews, Kiefer.Wolfowitz, PV=nRT, Tobias Bergemann, Zundark

Pseudoconvex function  Source: http://en.wikipedia.org/w/index.php?oldid=409058946  Contributors: Hrafn, Isheden, JRSpriggs, Kiefer.Wolfowitz, Michael Hardy, Oleg Alexandrov, Sławomir
Biały

Quasi-analytic function  Source: http://en.wikipedia.org/w/index.php?oldid=375965573  Contributors: CBM, CRGreathouse, Giftlite, Holmansf, Michael Hardy, R.e.b., TakuyaMurata

Quasiconvex function  Source: http://en.wikipedia.org/w/index.php?oldid=407844375  Contributors: Aleph4, CBM, CLST, Damiano.varagnolo, David Eppstein, Entropeneur, Isheden, Jackzhp,
Johngcarlsson, Josteinaj, Kiefer.Wolfowitz, Landroni, Mr. Lefty, Oleg Alexandrov, PV=nRT, Robinh, Sabamo, SlamDiego, Sławomir Biały, Tobacman, Tobias Bergemann, Vipul, 22 anonymous
edits

Quasiperiodic function  Source: http://en.wikipedia.org/w/index.php?oldid=371774947  Contributors: Arthena, Charles Matthews, Father Goose, Gene Ward Smith, Googl, Oleg Alexandrov,
PV=nRT, Rbj, Ryan Reich, Uppland, 2 anonymous edits

Radially unbounded function  Source: http://en.wikipedia.org/w/index.php?oldid=390752509  Contributors: Charles Matthews, Fi012345, Michael Hardy, R.e.b., SpecMode, TYelliot, 1
anonymous edits

Rational function  Source: http://en.wikipedia.org/w/index.php?oldid=409762894  Contributors: Adam majewski, Ahoerstemeier, Alan R. Fisher, Alphachimp, Backslash Forwardslash,
Bcherkas, Benzi455, BiT, Bilboq, Brian the Editor, Bsbllstr512, Burn, CRGreathouse, CSTAR, Calle, Cchilas, Charles Matthews, Courcelles, Davin, Dbaird3191, Doctormatt, Everyguy,
Gandalf61, Giftlite, Hippie Metalhead, Joeldl, Kasadkad, Kirbytime, Lambiam, Logan, Lzur, Marc van Leeuwen, MarkSweep, Mathiastck, Merope, Mets501, Michael Hardy, Miquonranger03,
Oleg Alexandrov, Ondenc, Ornil, PAR, PMajer, PV=nRT, Paul August, PawelekO, Petermmurphy, Pomte, Rabidchipmunk666, Rckrone, RedWolf, Rick Norwood, Sam Derbyshire, Tobias
Bergemann, TomyDuby, Wricardoh, YGingras, Yamamoto Ichiro, 68 anonymous edits

Real-valued function  Source: http://en.wikipedia.org/w/index.php?oldid=382443282  Contributors: Alpha 4615, Bagsc, Drilnoth, Michael Hardy, PV=nRT, Schmloof, 4 anonymous edits

Ring of symmetric functions  Source: http://en.wikipedia.org/w/index.php?oldid=398385154  Contributors: Ahoerstemeier, Arthur Rubin, Basemaze, BenFrantzDale, CBM, Charles Matthews,
Giftlite, Haonhien, Hillman, Marc van Leeuwen, Mhym, Michael Hardy, Nbarth, Oleg Alexandrov, PV=nRT, Patrick, Sbilley, Tabletop, Woohookitty, Zaslav, Петър Петров, 9 anonymous edits

Simple function  Source: http://en.wikipedia.org/w/index.php?oldid=377495030  Contributors: 4C, Algebraist, Ashsearle, B.Wind, Calle, El C, Fibonacci, Generalebriety, Helgus, Ht686rg90,
KSmrq, Kelly Martin, Madmath789, Michael Hardy, Oleg Alexandrov, PV=nRT, Salgueiro, Setokaiba, Silverfish, TakuyaMurata, Timhoooey, Tobias Bergemann, WISo, Weyes, 23 anonymous
edits

Single-valued function  Source: http://en.wikipedia.org/w/index.php?oldid=393457644  Contributors: CBM, Charles Matthews, Giftlite, Kingdon, LilHelpa, Oleg Alexandrov, PV=nRT,
Pseinstein, TheParanoidOne, 18 anonymous edits

Singular function  Source: http://en.wikipedia.org/w/index.php?oldid=397687014  Contributors: Aominux, AxelBoldt, Bart133, Borat fan, Charles Matthews, DavidCBryant, Dugwiki, Gauge,
Henrygb, Inside19, Jitse Niesen, JocK, Josh Cherry, LC, Linas, MathHisSci, Meni Rosenfeld, Michael Hardy, Octahedron80, PV=nRT, PierreAbbat, Rumping, Sherbrooke, Stepa, Typhoonuiuc,
WaWe, 41 anonymous edits

Smooth function  Source: http://en.wikipedia.org/w/index.php?oldid=406426745  Contributors: AK Auto, Algebraist, Anonymous Dissident, AugPi, Ben Spinozoan, Berland, Billlion,
CRGreathouse, Celebere, Charles Matthews, Chinju, CiaPan, Compsonheir, Daniele.tampieri, Dmn, Doctormatt, Dysprosia, Eebster the Great, El C, Emily Jensen, Factsofphotos, Filos96,
Fullmetal2887, Gauge, Ghazer, Haseldon, Helder.wiki, Henning Makholm, Idm196884, Indrian, Jondaman21, Linas, Mad2Physicist, Madmath789, Marcowongshuinam, Mat cross, Michael
Hardy, Nickj, Ojs, Oleg Alexandrov, PV=nRT, Paul August, Pomte, Rholton, Rich Farmbrough, Rybu, Salgueiro, Sapphic, Saric, SimpsonDG, Sławomir Biały, Tamfang, TedPavlic, The Stickler,
The way, the truth, and the light, Tosha, Uncle Dick, UtherSRG, Vonkje, Wikomidia, X42bn6, Zundark, Zvika, 67 anonymous edits

Subharmonic function  Source: http://en.wikipedia.org/w/index.php?oldid=406513176  Contributors: Bdmy, Charles Matthews, Crowsnest, Gala.martin, Ghirlandajo, Giftlite, Hottiger, Jbaber,
Madmath789, Michael Hardy, Mkelly86, Oleg Alexandrov, PV=nRT, RayAYang, Szilard.revesz, Tiphareth, 6 anonymous edits

Sublinear function  Source: http://en.wikipedia.org/w/index.php?oldid=407713784  Contributors: Aidsfrag, Andreasvc, Charles Matthews, French Tourist, J04n, LachlanA, MathKnight,
MathMartin, Nils Grimsmo, PV=nRT, Silverfish, William Ackerman, 15 anonymous edits

Surjective function  Source: http://en.wikipedia.org/w/index.php?oldid=406015463  Contributors: 16@r, 216.60.221.xxx, ABCD, Aleph4, Algebraist, AmarChandra, Amillar, Angus Lepper,
Anonymous Dissident, Applebringer, AxelBoldt, CBM, Charles Matthews, Classicalecon, Conversion script, Dallashan, David Shay, Dreadstar, Dubhe.sk, Dysprosia, Ed g2s, Eliuha gmail.com,
Fredrik, Giftlite, Glenn, Gregbard, Hawthorn, II MusLiM HyBRiD II, Inquisitus, Jeandré du Toit, Jorge Stolfi, Karada, Keeyu, Kevin Lamoreau, Ksnortum, LOL, Larry V, Lethe, Malerin,
Manscher, MarSch, MarkSweep, Marqueed, Martynas Patasius, Matt Crypto, MatthewMain, Michael Hardy, MickPurcell, Mindmatrix, Nandhp, Obradovic Goran, Oleg Alexandrov, PV=nRT,
Paolo.dL, Paul August, Peiresc, Phil Boswell, Pit, Pjvpjv, Prolog, Quistnix, Rheun, Rich Farmbrough, Rjwilmsi, Rotemliss, SLMarcus, Saippuakauppias, Salgueiro, Sam Staton, Sasuketiimer,
Sbyrnes321, Schapel, Soapergem, Tarquin, Tbsmith, TechnoFaye, TedE, TheObtuseAngleOfDoom, Theabsurd, Tobias Bergemann, Toby Bartels, Tsemii, UnicornTapestry, Vivacissamamente,
Wshun, XJamRastafire, Александър, ‫ينام‬, 76 anonymous edits

Symmetrically continuous function  Source: http://en.wikipedia.org/w/index.php?oldid=388290090  Contributors: Algebraist, Charles Matthews, Clément Pillias, Eric119, Incnis Mrsi,
Maksim-e, PV=nRT, Rasmus Faber, That Guy, From That Show!
Article Sources and Contributors 162

Quasisymmetric function  Source: http://en.wikipedia.org/w/index.php?oldid=404791030  Contributors: Aaronlauve, Acather96, Cbhrrhwasshs, Chronulator, Drmies, Edison, Frank Sottile,
Michael Hardy, Mskandera, Poset123, Sbilley, Victorreiner, 23 anonymous edits

Transcendental function  Source: http://en.wikipedia.org/w/index.php?oldid=409370318  Contributors: Asmeurer, Bcherkas, BenFrantzDale, Busy1mind, CBM, Charles Matthews, Chenxlee,
Chrisminter, Dcljr, Deineka, Delaszk, Doctormatt, Dysprosia, Etaoin, Exteray, Fgnievinski, Fredrik, Galoubet, Giftlite, Kickflipthecat, Michael Hardy, Michael Larsen, Michael Slone, Miterdale,
Ninly, Ourai, PV=nRT, Passive, Paxinum, Phaedriel, Rgdboer, Rholton, SQFreak, Silly rabbit, Sławomir Biała, The Thing That Should Not Be, Thegeneralguy, Thumperward, Usien6, 51
anonymous edits

Unary function  Source: http://en.wikipedia.org/w/index.php?oldid=254311922  Contributors: Cyrius, Flambergius, Grutness, J3ff, Jacob grace, Markhobley, PV=nRT, Pomte, Salix alba,
Someguy1221, 4 anonymous edits

Univalent function  Source: http://en.wikipedia.org/w/index.php?oldid=399153583  Contributors: CYCC, Madmath789, Michael Hardy, Oleg Alexandrov, PV=nRT

Vector-valued function  Source: http://en.wikipedia.org/w/index.php?oldid=409509461  Contributors: Ac44ck, BrokenSegue, CBM, Charles Matthews, EconoPhysicist, FilipeS, Giftlite, Gurch,
Hannes Eder, Ht686rg90, JackSchmidt, Jecowa, MATThematical, MarcusMaximus, Michael Hardy, Neparis, Nillerdk, PV=nRT, Paolo.dL, Parodi, Plastikspork, Richie, Rror, Salix alba, Spoon!,
StradivariusTV, Sławomir Biały, TexasAndroid, User A1, 10 anonymous edits

Weakly harmonic function  Source: http://en.wikipedia.org/w/index.php?oldid=306752760  Contributors: Charles Matthews, David Eppstein, Dysprosia, Giftlite, Martpol, Mietchen,
NatusRoma, Oleg Alexandrov, PV=nRT, Paul Laroque, Peruvianllama, Rettetast, RxS, 12 anonymous edits

Weakly measurable function  Source: http://en.wikipedia.org/w/index.php?oldid=388963224  Contributors: David Eppstein, Jmath666, Oleg Alexandrov, PV=nRT, Parodi, Sullivan.t.j, 1
anonymous edits
Image Sources, Licenses and Contributors 163

Image Sources, Licenses and Contributors


Image:y^3-xy+1=0.png  Source: http://en.wikipedia.org/w/index.php?title=File:Y^3-xy+1=0.png  License: GNU Free Documentation License  Contributors: Silly rabbit
Image:Bijection.svg  Source: http://en.wikipedia.org/w/index.php?title=File:Bijection.svg  License: Public Domain  Contributors: Darapti, Manscher, Ramac
Image:Bijective composition.svg  Source: http://en.wikipedia.org/w/index.php?title=File:Bijective_composition.svg  License: Public Domain  Contributors: Darapti, Sl
Image:ProduitTensoriel.png  Source: http://en.wikipedia.org/w/index.php?title=File:ProduitTensoriel.png  License: GNU Free Documentation License  Contributors: Loveless, 1 anonymous
edits
Image:Bounded and unbounded functions.svg  Source: http://en.wikipedia.org/w/index.php?title=File:Bounded_and_unbounded_functions.svg  License: Public Domain  Contributors:
User:Oleg Alexandrov
Image:ConcaveDef.png  Source: http://en.wikipedia.org/w/index.php?title=File:ConcaveDef.png  License: Public Domain  Contributors: Flavio Guitian
File:Rapid Oscillation.svg  Source: http://en.wikipedia.org/w/index.php?title=File:Rapid_Oscillation.svg  License: Creative Commons Attribution 3.0  Contributors: --pbroks13talk? Original
uploader was Pbroks13 at en.wikipedia
Image:Right-continuous.svg  Source: http://en.wikipedia.org/w/index.php?title=File:Right-continuous.svg  License: Public Domain  Contributors: w:User:JacjJacj
Image:Left-continuous.svg  Source: http://en.wikipedia.org/w/index.php?title=File:Left-continuous.svg  License: Public Domain  Contributors: Jacj, Plasticspork
Image:continuity topology.svg  Source: http://en.wikipedia.org/w/index.php?title=File:Continuity_topology.svg  License: Public Domain  Contributors: User:Dcoetzee
Image:ConvexFunction.svg  Source: http://en.wikipedia.org/w/index.php?title=File:ConvexFunction.svg  License: Creative Commons Attribution-Sharealike 3.0  Contributors: User:Eli
Osherovich
Image:Epigraph convex.svg  Source: http://en.wikipedia.org/w/index.php?title=File:Epigraph_convex.svg  License: Creative Commons Attribution-Sharealike 3.0  Contributors: User:Eli
Osherovich
Image:Polynomialdeg3.svg  Source: http://en.wikipedia.org/w/index.php?title=File:Polynomialdeg3.svg  License: Public Domain  Contributors: User:N.Mori
Image:Absolute value.svg  Source: http://en.wikipedia.org/w/index.php?title=File:Absolute_value.svg  License: GNU Free Documentation License  Contributors: User:Qef, User:Ævar Arnfjörð
Bjarmason
Image:WeierstrassFunction.svg  Source: http://en.wikipedia.org/w/index.php?title=File:WeierstrassFunction.svg  License: Public Domain  Contributors: User:Eeyore22
Image:Function x^2.svg  Source: http://en.wikipedia.org/w/index.php?title=File:Function_x^2.svg  License: GNU Free Documentation License  Contributors: User:Qualc1
Image:Function-x3.svg  Source: http://en.wikipedia.org/w/index.php?title=File:Function-x3.svg  License: unknown  Contributors: -
Image:Function-x3plus1.svg  Source: http://en.wikipedia.org/w/index.php?title=File:Function-x3plus1.svg  License: unknown  Contributors: -
File:FBN exp(-1x2).jpeg  Source: http://en.wikipedia.org/w/index.php?title=File:FBN_exp(-1x2).jpeg  License: Creative Commons Attribution-Sharealike 3.0  Contributors: Fly_by_Night
Image:Compfun.svg  Source: http://en.wikipedia.org/w/index.php?title=File:Compfun.svg  License: Public Domain  Contributors: User:Tlep
Image:Laplace's equation on an annulus.jpg  Source: http://en.wikipedia.org/w/index.php?title=File:Laplace's_equation_on_an_annulus.jpg  License: Creative Commons
Attribution-Sharealike 3.0  Contributors: User:DavidianSkitzou
Image:Conformal map.svg  Source: http://en.wikipedia.org/w/index.php?title=File:Conformal_map.svg  License: Public Domain  Contributors: User:Oleg Alexandrov
Image:Indicator function illustration.png  Source: http://en.wikipedia.org/w/index.php?title=File:Indicator_function_illustration.png  License: Public Domain  Contributors: User:Oleg
Alexandrov
Image:Injection.svg  Source: http://en.wikipedia.org/w/index.php?title=File:Injection.svg  License: Public Domain  Contributors: Sl, 2 anonymous edits
Image:Surjection.svg  Source: http://en.wikipedia.org/w/index.php?title=File:Surjection.svg  License: Public Domain  Contributors: Darapti, Sl
Image:Injective_composition2.svg  Source: http://en.wikipedia.org/w/index.php?title=File:Injective_composition2.svg  License: Public Domain  Contributors: User:Oleg Alexandrov
File:Gamma abs 3D.png  Source: http://en.wikipedia.org/w/index.php?title=File:Gamma_abs_3D.png  License: GNU Free Documentation License  Contributors: User:Geek3
Image:Monotonicity example1.png  Source: http://en.wikipedia.org/w/index.php?title=File:Monotonicity_example1.png  License: Public Domain  Contributors: User:Oleg Alexandrov
Image:Monotonicity example2.png  Source: http://en.wikipedia.org/w/index.php?title=File:Monotonicity_example2.png  License: Public Domain  Contributors: User:Oleg Alexandrov
Image:Monotonicity example3.png  Source: http://en.wikipedia.org/w/index.php?title=File:Monotonicity_example3.png  License: Public Domain  Contributors: User:Oleg Alexandrov
Image:Monotone Boolean functions 0,1,2,3.svg  Source: http://en.wikipedia.org/w/index.php?title=File:Monotone_Boolean_functions_0,1,2,3.svg  License: Public Domain  Contributors:
User:Lipedia
Image:Multivalued function.svg  Source: http://en.wikipedia.org/w/index.php?title=File:Multivalued_function.svg  License: Public Domain  Contributors: Sl, 1 anonymous edits
Image:Periodic function illustration.svg  Source: http://en.wikipedia.org/w/index.php?title=File:Periodic_function_illustration.svg  License: Public Domain  Contributors: User:Oleg
Alexandrov
Image:Sine.svg  Source: http://en.wikipedia.org/w/index.php?title=File:Sine.svg  License: GNU Free Documentation License  Contributors: User:Geek3
Image:Sine cosine plot.svg  Source: http://en.wikipedia.org/w/index.php?title=File:Sine_cosine_plot.svg  License: Creative Commons Attribution-Sharealike 2.5  Contributors: User:Qualc1
Image:PiecewiseLinear.png  Source: http://en.wikipedia.org/w/index.php?title=File:PiecewiseLinear.png  License: Public Domain  Contributors: User:Jim.belk
Image:Finite element method 1D illustration1.png  Source: http://en.wikipedia.org/w/index.php?title=File:Finite_element_method_1D_illustration1.png  License: Public Domain  Contributors:
Joelholdsworth, Maksim, WikipediaMaster
Image:Piecewise linear function2D.svg  Source: http://en.wikipedia.org/w/index.php?title=File:Piecewise_linear_function2D.svg  License: Public Domain  Contributors: User:Oleg Alexandrov
Image:Quasiconvex function.png  Source: http://en.wikipedia.org/w/index.php?title=File:Quasiconvex_function.png  License: Public Domain  Contributors: BenFrantzDale, Darapti, Oleg
Alexandrov
Image:Nonquasiconvex function.png  Source: http://en.wikipedia.org/w/index.php?title=File:Nonquasiconvex_function.png  License: Public Domain  Contributors: BenFrantzDale, Darapti,
Oleg Alexandrov
Image:Quasi-concave-function-graph.png  Source: http://en.wikipedia.org/w/index.php?title=File:Quasi-concave-function-graph.png  License: Creative Commons Attribution-Sharealike 2.5
 Contributors: Darapti, He1ix
Image:RationalDegree2byXedi.gif  Source: http://en.wikipedia.org/w/index.php?title=File:RationalDegree2byXedi.gif  License: GNU Free Documentation License  Contributors: Sam
Derbyshire
Image:RationalDegree3byXedi.gif  Source: http://en.wikipedia.org/w/index.php?title=File:RationalDegree3byXedi.gif  License: GNU Free Documentation License  Contributors: Sam
Derbyshire
Image:Devils-staircase.svg  Source: http://en.wikipedia.org/w/index.php?title=File:Devils-staircase.svg  License: GNU Free Documentation License  Contributors: User:Linas
Image:Bump2D illustration.png  Source: http://en.wikipedia.org/w/index.php?title=File:Bump2D_illustration.png  License: Public Domain  Contributors: User:Oleg Alexandrov
Image:C0 function.png  Source: http://en.wikipedia.org/w/index.php?title=File:C0_function.png  License: Public Domain  Contributors: User:Oleg Alexandrov
Image:TV pic3.png  Source: http://en.wikipedia.org/w/index.php?title=File:TV_pic3.png  License: Public Domain  Contributors: User:Oleg Alexandrov
Image:Mollifier illustration.png  Source: http://en.wikipedia.org/w/index.php?title=File:Mollifier_illustration.png  License: Public Domain  Contributors: User:Oleg Alexandrov
Image:parametric_continuity_c0.gif  Source: http://en.wikipedia.org/w/index.php?title=File:Parametric_continuity_c0.gif  License: Public Domain  Contributors: BenFrantzDale, Stimpy
File:Parametric_continuity_vector.svg  Source: http://en.wikipedia.org/w/index.php?title=File:Parametric_continuity_vector.svg  License: Public Domain  Contributors: User:Factsofphotos
File:PD-icon.svg  Source: http://en.wikipedia.org/w/index.php?title=File:PD-icon.svg  License: Public Domain  Contributors: User:Duesentrieb, User:Rfl
Image:Codomain2.SVG  Source: http://en.wikipedia.org/w/index.php?title=File:Codomain2.SVG  License: Public Domain  Contributors: Damien Karras (talk). Original uploader was Damien
Karras at en.wikipedia
Image:Surjective_composition.svg  Source: http://en.wikipedia.org/w/index.php?title=File:Surjective_composition.svg  License: Public Domain  Contributors: Darapti, Sl
Image:QSymDiagram.png  Source: http://en.wikipedia.org/w/index.php?title=File:QSymDiagram.png  License: Free Art License  Contributors: Crites
Image:Vector-valued function-2.png  Source: http://en.wikipedia.org/w/index.php?title=File:Vector-valued_function-2.png  License: GNU Free Documentation License  Contributors: Nillerdk
License 164

License
Creative Commons Attribution-Share Alike 3.0 Unported
http:/ / creativecommons. org/ licenses/ by-sa/ 3. 0/

Anda mungkin juga menyukai