Jet (mathematics)

Last updated

In mathematics, the jet is an operation that takes a differentiable function f and produces a polynomial, the truncated Taylor polynomial of f, at each point of its domain. Although this is the definition of a jet, the theory of jets regards these polynomials as being abstract polynomials rather than polynomial functions.

Contents

This article first explores the notion of a jet of a real valued function in one real variable, followed by a discussion of generalizations to several real variables. It then gives a rigorous construction of jets and jet spaces between Euclidean spaces. It concludes with a description of jets between manifolds, and how these jets can be constructed intrinsically. In this more general context, it summarizes some of the applications of jets to differential geometry and the theory of differential equations.

Jets of functions between Euclidean spaces

Before giving a rigorous definition of a jet, it is useful to examine some special cases.

One-dimensional case

Suppose that is a real-valued function having at least k + 1 derivatives in a neighborhood U of the point . Then by Taylor's theorem,

where

Then the k-jet of f at the point is defined to be the polynomial

Jets are normally regarded as abstract polynomials in the variable z, not as actual polynomial functions in that variable. In other words, z is an indeterminate variable allowing one to perform various algebraic operations among the jets. It is in fact the base-point from which jets derive their functional dependency. Thus, by varying the base-point, a jet yields a polynomial of order at most k at every point. This marks an important conceptual distinction between jets and truncated Taylor series: ordinarily a Taylor series is regarded as depending functionally on its variable, rather than its base-point. Jets, on the other hand, separate the algebraic properties of Taylor series from their functional properties. We shall deal with the reasons and applications of this separation later in the article.

Mappings from one Euclidean space to another

Suppose that is a function from one Euclidean space to another having at least (k + 1) derivatives. In this case, Taylor's theorem asserts that

The k-jet of f is then defined to be the polynomial

in , where .

Algebraic properties of jets

There are two basic algebraic structures jets can carry. The first is a product structure, although this ultimately turns out to be the least important. The second is the structure of the composition of jets.

If are a pair of real-valued functions, then we can define the product of their jets via

Here we have suppressed the indeterminate z, since it is understood that jets are formal polynomials. This product is just the product of ordinary polynomials in z, modulo . In other words, it is multiplication in the ring , where is the ideal generated by polynomials homogeneous of order  k + 1.

We now move to the composition of jets. To avoid unnecessary technicalities, we consider jets of functions that map the origin to the origin. If and with f(0) = 0 and g(0) = 0, then . The composition of jets is defined by It is readily verified, using the chain rule, that this constitutes an associative noncommutative operation on the space of jets at the origin.

In fact, the composition of k-jets is nothing more than the composition of polynomials modulo the ideal of polynomials homogeneous of order .

Examples:

and

Jets at a point in Euclidean space: rigorous definitions

Analytic definition

The following definition uses ideas from mathematical analysis to define jets and jet spaces. It can be generalized to smooth functions between Banach spaces, analytic functions between real or complex domains, to p-adic analysis, and to other areas of analysis.

Let be the vector space of smooth functions . Let k be a non-negative integer, and let p be a point of . We define an equivalence relation on this space by declaring that two functions f and g are equivalent to order k if f and g have the same value at p, and all of their partial derivatives agree at p up to (and including) their k-th-order derivatives. In short, iff to k-th order.

The k-th-order jet space of at p is defined to be the set of equivalence classes of , and is denoted by .

The k-th-order jet at p of a smooth function is defined to be the equivalence class of f in .

Algebro-geometric definition

The following definition uses ideas from algebraic geometry and commutative algebra to establish the notion of a jet and a jet space. Although this definition is not particularly suited for use in algebraic geometry per se, since it is cast in the smooth category, it can easily be tailored to such uses.

Let be the vector space of germs of smooth functions at a point p in . Let be the ideal consisting of germs of functions that vanish at p. (This is the maximal ideal for the local ring .) Then the ideal consists of all function germs that vanish to order k at p. We may now define the jet space at p by

If is a smooth function, we may define the k-jet of f at p as the element of by setting

This is a more general construction. For an -space , let be the stalk of the structure sheaf at and let be the maximal ideal of the local ring . The kth jet space at is defined to be the ring ( is the product of ideals).

Taylor's theorem

Regardless of the definition, Taylor's theorem establishes a canonical isomorphism of vector spaces between and . So in the Euclidean context, jets are typically identified with their polynomial representatives under this isomorphism.

Jet spaces from a point to a point

We have defined the space of jets at a point . The subspace of this consisting of jets of functions f such that f(p) = q is denoted by

Jets of functions between two manifolds

If M and N are two smooth manifolds, how do we define the jet of a function ? We could perhaps attempt to define such a jet by using local coordinates on M and N. The disadvantage of this is that jets cannot thus be defined in an invariant fashion. Jets do not transform as tensors. Instead, jets of functions between two manifolds belong to a jet bundle.

Jets of functions from the real line to a manifold

Suppose that M is a smooth manifold containing a point p. We shall define the jets of curves through p, by which we henceforth mean smooth functions such that f(0) = p. Define an equivalence relation as follows. Let f and g be a pair of curves through p. We will then say that f and g are equivalent to order k at p if there is some neighborhood U of p, such that, for every smooth function , . Note that these jets are well-defined since the composite functions and are just mappings from the real line to itself. This equivalence relation is sometimes called that of k-th-order contact between curves at p.

We now define the k-jet of a curve f through p to be the equivalence class of f under , denoted or . The k-th-order jet space is then the set of k-jets at p.

As p varies over M, forms a fibre bundle over M: the k-th-order tangent bundle, often denoted in the literature by TkM (although this notation occasionally can lead to confusion). In the case k=1, then the first-order tangent bundle is the usual tangent bundle: T1M = TM.

To prove that TkM is in fact a fibre bundle, it is instructive to examine the properties of in local coordinates. Let (xi)= (x1,...,xn) be a local coordinate system for M in a neighborhood U of p. Abusing notation slightly, we may regard (xi) as a local diffeomorphism .

Claim. Two curves f and g through p are equivalent modulo if and only if .

Indeed, the only if part is clear, since each of the n functions x1,...,xn is a smooth function from M to . So by the definition of the equivalence relation , two equivalent curves must have .
Conversely, suppose that ; is a smooth real-valued function on M in a neighborhood of p. Since every smooth function has a local coordinate expression, we may express ; as a function in the coordinates. Specifically, if q is a point of M near p, then
for some smooth real-valued function ψ of n real variables. Hence, for two curves f and g through p, we have
The chain rule now establishes the if part of the claim. For instance, if f and g are functions of the real variable t , then
which is equal to the same expression when evaluated against g instead of f, recalling that f(0)=g(0)=p and f and g are in k-th-order contact in the coordinate system (xi).

Hence the ostensible fibre bundle TkM admits a local trivialization in each coordinate neighborhood. At this point, in order to prove that this ostensible fibre bundle is in fact a fibre bundle, it suffices to establish that it has non-singular transition functions under a change of coordinates. Let be a different coordinate system and let be the associated change of coordinates diffeomorphism of Euclidean space to itself. By means of an affine transformation of , we may assume without loss of generality that ρ(0)=0. With this assumption, it suffices to prove that is an invertible transformation under jet composition. (See also jet groups.) But since ρ is a diffeomorphism, is a smooth mapping as well. Hence,

which proves that is non-singular. Furthermore, it is smooth, although we do not prove that fact here.

Intuitively, this means that we can express the jet of a curve through p in terms of its Taylor series in local coordinates on M.

Examples in local coordinates:

Given such a tangent vector v, let f be the curve given in the xi coordinate system by . If φ is a smooth function in a neighborhood of p with φ(p) = 0, then
is a smooth real-valued function of one variable whose 1-jet is given by
which proves that one can naturally identify tangent vectors at a point with the 1-jets of curves through that point.
In a local coordinate system xi centered at a point p, we can express the second-order Taylor polynomial of a curve f(t) through p by
So in the x coordinate system, the 2-jet of a curve through p is identified with a list of real numbers . As with the tangent vectors (1-jets of curves) at a point, 2-jets of curves obey a transformation law upon application of the coordinate transition functions.
Let (yi) be another coordinate system. By the chain rule,
Hence, the transformation law is given by evaluating these two expressions at t = 0.
Note that the transformation law for 2-jets is second-order in the coordinate transition functions.

Jets of functions from a manifold to a manifold

We are now prepared to define the jet of a function from a manifold to a manifold.

Suppose that M and N are two smooth manifolds. Let p be a point of M. Consider the space consisting of smooth maps defined in some neighborhood of p. We define an equivalence relation on as follows. Two maps f and g are said to be equivalent if, for every curve γ through p (recall that by our conventions this is a mapping such that ), we have on some neighborhood of 0.

The jet space is then defined to be the set of equivalence classes of modulo the equivalence relation . Note that because the target space N need not possess any algebraic structure, also need not have such a structure. This is, in fact, a sharp contrast with the case of Euclidean spaces.

If is a smooth function defined near p, then we define the k-jet of f at p, , to be the equivalence class of f modulo .

Multijets

John Mather introduced the notion of multijet. Loosely speaking, a multijet is a finite list of jets over different base-points. Mather proved the multijet transversality theorem, which he used in his study of stable mappings.

Jets of sections

Suppose that E is a finite-dimensional smooth vector bundle over a manifold M, with projection . Then sections of E are smooth functions such that is the identity automorphism of M. The jet of a section s over a neighborhood of a point p is just the jet of this smooth function from M to E at p.

The space of jets of sections at p is denoted by . Although this notation can lead to confusion with the more general jet spaces of functions between two manifolds, the context typically eliminates any such ambiguity.

Unlike jets of functions from a manifold to another manifold, the space of jets of sections at p carries the structure of a vector space inherited from the vector space structure on the sections themselves. As p varies over M, the jet spaces form a vector bundle over M, the k-th-order jet bundle of E, denoted by Jk(E).

We work in local coordinates at a point and use the Einstein notation. Consider a vector field
in a neighborhood of p in M. The 1-jet of v is obtained by taking the first-order Taylor polynomial of the coefficients of the vector field:
In the x coordinates, the 1-jet at a point can be identified with a list of real numbers . In the same way that a tangent vector at a point can be identified with the list (vi), subject to a certain transformation law under coordinate transitions, we have to know how the list is affected by a transition.
So consider the transformation law in passing to another coordinate system yi. Let wk be the coefficients of the vector field v in the y coordinates. Then in the y coordinates, the 1-jet of v is a new list of real numbers . Since
it follows that
So
Expanding by a Taylor series, we have
Note that the transformation law is second-order in the coordinate transition functions.

Differential operators between vector bundles

See also

Related Research Articles

In calculus, the chain rule is a formula that expresses the derivative of the composition of two differentiable functions f and g in terms of the derivatives of f and g. More precisely, if is the function such that for every x, then the chain rule is, in Lagrange's notation,

<span class="mw-page-title-main">Gradient</span> Multivariate derivative (mathematics)

In vector calculus, the gradient of a scalar-valued differentiable function of several variables is the vector field whose value at a point gives the direction and the rate of fastest increase. The gradient transforms like a vector under change of basis of the space of variables of . If the gradient of a function is non-zero at a point , the direction of the gradient is the direction in which the function increases most quickly from , and the magnitude of the gradient is the rate of increase in that direction, the greatest absolute directional derivative. Further, a point where the gradient is the zero vector is known as a stationary point. The gradient thus plays a fundamental role in optimization theory, where it is used to maximize a function by gradient ascent. In coordinate-free terms, the gradient of a function may be defined by:

In differential geometry, a subject of mathematics, a symplectic manifold is a smooth manifold, , equipped with a closed nondegenerate differential 2-form , called the symplectic form. The study of symplectic manifolds is called symplectic geometry or symplectic topology. Symplectic manifolds arise naturally in abstract formulations of classical mechanics and analytical mechanics as the cotangent bundles of manifolds. For example, in the Hamiltonian formulation of classical mechanics, which provides one of the major motivations for the field, the set of all possible configurations of a system is modeled as a manifold, and this manifold's cotangent bundle describes the phase space of the system.

<span class="mw-page-title-main">Taylor's theorem</span> Approximation of a function by a truncated power series

In calculus, Taylor's theorem gives an approximation of a -times differentiable function around a given point by a polynomial of degree , called the -th-order Taylor polynomial. For a smooth function, the Taylor polynomial is the truncation at the order of the Taylor series of the function. The first-order Taylor polynomial is the linear approximation of the function, and the second-order Taylor polynomial is often referred to as the quadratic approximation. There are several versions of Taylor's theorem, some giving explicit estimates of the approximation error of the function by its Taylor polynomial.

A mathematical symbol is a figure or a combination of figures that is used to represent a mathematical object, an action on mathematical objects, a relation between mathematical objects, or for structuring the other symbols that occur in a formula. As formulas are entirely constituted with symbols of various types, many symbols are needed for expressing all mathematics.

<span class="mw-page-title-main">Spherical harmonics</span> Special mathematical functions defined on the surface of a sphere

In mathematics and physical science, spherical harmonics are special functions defined on the surface of a sphere. They are often employed in solving partial differential equations in many scientific fields. A list of the spherical harmonics is available in Table of spherical harmonics.

<span class="mw-page-title-main">Algebraic curve</span> Curve defined as zeros of polynomials

In mathematics, an affine algebraic plane curve is the zero set of a polynomial in two variables. A projective algebraic plane curve is the zero set in a projective plane of a homogeneous polynomial in three variables. An affine algebraic plane curve can be completed in a projective algebraic plane curve by homogenizing its defining polynomial. Conversely, a projective algebraic plane curve of homogeneous equation h(x, y, t) = 0 can be restricted to the affine algebraic plane curve of equation h(x, y, 1) = 0. These two operations are each inverse to the other; therefore, the phrase algebraic plane curve is often used without specifying explicitly whether it is the affine or the projective case that is considered.

<span class="mw-page-title-main">Differential operator</span> Typically linear operator defined in terms of differentiation of functions

In mathematics, a differential operator is an operator defined as a function of the differentiation operator. It is helpful, as a matter of notation first, to consider differentiation as an abstract operation that accepts a function and returns another function.

In differential geometry, the Ricci curvature tensor, named after Gregorio Ricci-Curbastro, is a geometric object which is determined by a choice of Riemannian or pseudo-Riemannian metric on a manifold. It can be considered, broadly, as a measure of the degree to which the geometry of a given metric tensor differs locally from that of ordinary Euclidean space or pseudo-Euclidean space.

In mathematics, specifically differential calculus, the inverse function theorem gives a sufficient condition for a function to be invertible in a neighborhood of a point in its domain: namely, that its derivative is continuous and non-zero at the point. The theorem also gives a formula for the derivative of the inverse function. In multivariable calculus, this theorem can be generalized to any continuously differentiable, vector-valued function whose Jacobian determinant is nonzero at a point in its domain, giving a formula for the Jacobian matrix of the inverse. There are also versions of the inverse function theorem for complex holomorphic functions, for differentiable maps between manifolds, for differentiable functions between Banach spaces, and so forth.

In mathematics, the Schwarzian derivative is an operator similar to the derivative which is invariant under Möbius transformations. Thus, it occurs in the theory of the complex projective line, and in particular, in the theory of modular forms and hypergeometric functions. It plays an important role in the theory of univalent functions, conformal mapping and Teichmüller spaces. It is named after the German mathematician Hermann Schwarz.

In mathematics, a homogeneous function is a function of several variables such that the following holds: If each of the function's arguments is multiplied by the same scalar, then the function's value is multiplied by some power of this scalar; the power is called the degree of homogeneity, or simply the degree. That is, if k is an integer, a function f of n variables is homogeneous of degree k if

<span class="mw-page-title-main">Bump function</span> Smooth and compactly supported function

In mathematics, a bump function is a function on a Euclidean space which is both smooth and compactly supported. The set of all bump functions with domain forms a vector space, denoted or The dual space of this space endowed with a suitable topology is the space of distributions.

In mathematics, differential algebra is, broadly speaking, the area of mathematics consisting in the study of differential equations and differential operators as algebraic objects in view of deriving properties of differential equations and operators without computing the solutions, similarly as polynomial algebras are used for the study of algebraic varieties, which are solution sets of systems of polynomial equations. Weyl algebras and Lie algebras may be considered as belonging to differential algebra.

In mathematics, the (linear) Peetre theorem, named after Jaak Peetre, is a result of functional analysis that gives a characterisation of differential operators in terms of their effect on generalized function spaces, and without mentioning differentiation in explicit terms. The Peetre theorem is an example of a finite order theorem in which a function or a functor, defined in a very general way, can in fact be shown to be a polynomial because of some extraneous condition or symmetry imposed upon it.

In the mathematical field of differential topology, the Lie bracket of vector fields, also known as the Jacobi–Lie bracket or the commutator of vector fields, is an operator that assigns to any two vector fields X and Y on a smooth manifold M a third vector field denoted [X, Y].

In mathematics, the Möbius energy of a knot is a particular knot energy, i.e., a functional on the space of knots. It was discovered by Jun O'Hara, who demonstrated that the energy blows up as the knot's strands get close to one another. This is a useful property because it prevents self-intersection and ensures the result under gradient descent is of the same knot type.

In mathematics, quaternionic analysis is the study of functions with quaternions as the domain and/or range. Such functions can be called functions of a quaternion variable just as functions of a real variable or a complex variable are called.

In complex analysis of one and several complex variables, Wirtinger derivatives, named after Wilhelm Wirtinger who introduced them in 1927 in the course of his studies on the theory of functions of several complex variables, are partial differential operators of the first order which behave in a very similar manner to the ordinary derivatives with respect to one real variable, when applied to holomorphic functions, antiholomorphic functions or simply differentiable functions on complex domains. These operators permit the construction of a differential calculus for such functions that is entirely analogous to the ordinary differential calculus for functions of real variables.

In mathematics, calculus on Euclidean space is a generalization of calculus of functions in one or several variables to calculus of functions on Euclidean space as well as a finite-dimensional real vector space. This calculus is also known as advanced calculus, especially in the United States. It is similar to multivariable calculus but is somewhat more sophisticated in that it uses linear algebra more extensively and covers some concepts from differential geometry such as differential forms and Stokes' formula in terms of differential forms. This extensive use of linear algebra also allows a natural generalization of multivariable calculus to calculus on Banach spaces or topological vector spaces.

References