De Wikipedia, la enciclopedia libre
Saltar a navegación Saltar a búsqueda

En los campos matemáticos de la geometría diferencial y el cálculo tensorial , las formas diferenciales son un enfoque del cálculo multivariable que es independiente de las coordenadas . Las formas diferenciales proporcionan un enfoque unificado para definir integrandos sobre curvas, superficies, sólidos y variedades de dimensiones superiores . La noción moderna de formas diferenciales fue pionera en Élie Cartan . Tiene muchas aplicaciones, especialmente en geometría, topología y física.

Por ejemplo, la expresión f ( x ) dx del cálculo de una variable es un ejemplo de una forma de 1 y se puede integrar en un intervalo orientado [ a , b ] en el dominio de f :

De manera similar, la expresión f ( x , y , z ) dxdy + g ( x , y , z ) dzdx + h ( x , y , z ) dydz es una forma 2 que tiene una integral de superficie sobre una superficie orientada S :

El símbolo denota el producto exterior , a veces llamado producto de la cuña , de dos formas diferenciales. Del mismo modo, un 3 -forma f ( x , y , z ) dxdydz representa un elemento de volumen que se puede integrar sobre una región orientada de espacio. En general, una forma k es un objeto que puede integrarse sobre una variedad orientada a una dimensión k , y es homogénea de grado k en los diferenciales de coordenadas.

El álgebra de formas diferenciales está organizado de una manera que refleja naturalmente la orientación del dominio de integración. Existe una operación d sobre formas diferenciales conocida como derivada exterior que, cuando se le da una forma k como entrada, produce una forma ( k + 1) como salida. Esta operación se extiende el diferencial de una función , y está directamente relacionada con la divergencia y el enrollamiento de un campo vectorial de una manera que hace que el teorema fundamental del cálculo , el teorema de la divergencia , el teorema de Green, y casos especiales del teorema de Stokes del mismo resultado general, conocido en este contexto también como el teorema de Stokes generalizado . De manera más profunda, este teorema relaciona la topología del dominio de integración con la estructura de las propias formas diferenciales; la conexión precisa se conoce como teorema de de Rham .

El marco general para el estudio de las formas diferenciales se encuentra en una variedad diferenciable . Las formas diferenciales 1 son naturalmente duales a los campos vectoriales en una variedad, y el emparejamiento entre los campos vectoriales y las formas 1 se extiende a formas diferenciales arbitrarias por el producto interior . El álgebra de formas diferenciales junto con la derivada exterior definida en ella se conserva mediante el retroceso en funciones suaves entre dos variedades. Esta característica permite que la información geométricamente invariante se mueva de un espacio a otro a través del retroceso, siempre que la información se exprese en términos de formas diferenciales. Como ejemplo, la fórmula de cambio de variables para la integración se convierte en una simple declaración de que una integral se conserva bajo retroceso.

Historia [ editar ]

Las formas diferenciales son parte del campo de la geometría diferencial, influenciadas por el álgebra lineal. Aunque la noción de diferencial es bastante antigua, el intento inicial de una organización algebraica de formas diferenciales suele atribuirse a Élie Cartan con referencia a su artículo de 1899. [1] Algunos aspectos del álgebra exterior de formas diferenciales aparecen en la obra de 1844 de Hermann Grassmann , Die Lineale Ausdehnungslehre, ein neuer Zweig der Mathematik (La teoría de la extensión lineal, una nueva rama de las matemáticas) .

Concepto [ editar ]

Las formas diferenciales proporcionan un enfoque del cálculo multivariable que es independiente de las coordenadas .

Integración y orientación [ editar ]

Se puede integrar una forma k diferencial sobre un colector orientado de dimensión k . Se puede pensar que una forma diferencial 1 mide una longitud orientada infinitesimal o una densidad orientada unidimensional. Se puede pensar en una forma diferencial 2 como la medición de un área orientada infinitesimal, o una densidad orientada bidimensional. Y así.

La integración de formas diferenciales está bien definida solo en variedades orientadas . Un ejemplo de una variedad unidimensional es un intervalo [ a , b ] , y a los intervalos se les puede dar una orientación: están orientados positivamente si a < b , y orientados negativamente en caso contrario. Si a < b, entonces la integral de la forma diferencial 1 f ( x ) dx sobre el intervalo [ a , b ] (con su orientación positiva natural) es

que es el negativo de la integral de la misma forma diferencial sobre el mismo intervalo, cuando está equipado con la orientación opuesta. Es decir:

Esto da un contexto geométrico a las convenciones para integrales unidimensionales, que el signo cambia cuando se invierte la orientación del intervalo. Una explicación estándar de esto en la teoría de integración de una variable es que, cuando los límites de integración están en el orden opuesto ( b < a ), el incremento dx es negativo en la dirección de integración.

De manera más general, una forma m es una densidad orientada que se puede integrar sobre una variedad orientada m- dimensional. (Por ejemplo, una forma de 1 se puede integrar sobre una curva orientada, una forma de 2 se puede integrar sobre una superficie orientada, etc.) Si M es una variedad orientada de m -dimensional, y M es la misma variedad con opuesto orientación y ω es una forma m , entonces uno tiene:

Estas convenciones corresponden a interpretar el integrando como una forma diferencial, integrada en una cadena . En la teoría de la medida , por el contrario, se interpreta el integrando como una función f con respecto a una medida μ y se integra sobre un subconjunto A , sin ninguna noción de orientación; uno escribe para indicar integración sobre un subconjunto A . Ésta es una distinción menor en una dimensión, pero se vuelve más sutil en variedades de dimensiones superiores; consulte a continuación para obtener más detalles.

Hacer precisa la noción de densidad orientada y, por tanto, de forma diferencial, implica el álgebra exterior . Los diferenciales de un conjunto de coordenadas, dx 1 , ..., dx n se pueden utilizar como base para todas las formas 1 . Cada uno de estos representa un covector en cada punto del colector que se puede pensar que mide un pequeño desplazamiento en la dirección de coordenadas correspondiente. Una forma 1 general es una combinación lineal de estos diferenciales en cada punto de la variedad:

donde f k = f k ( x 1 , ..., x n ) son funciones de todas las coordenadas. Una forma diferencial 1 se integra a lo largo de una curva orientada como una integral de línea.

Las expresiones dx idx j , donde i < j se pueden usar como base en cada punto de la variedad para todas las dos formas. Esto se puede considerar como un cuadrado infinitesimal orientado paralelo al plano x i - x j . A dos forma general es una combinación lineal de estos en cada punto en el colector de: , y está integrado como una integral de superficie.

Una operación fundamental definida en formas diferenciales es el producto exterior (el símbolo es la cuña ). Esto es similar al producto cruzado del cálculo vectorial, en el sentido de que es un producto alterno. Por ejemplo,

porque el cuadrado cuyo primer lado es dx 1 y el segundo lado es dx 2 debe considerarse que tiene la orientación opuesta al cuadrado cuyo primer lado es dx 2 y cuyo segundo lado es dx 1 . Es por eso que solo necesitamos sumar las expresiones dx idx j , con i < j ; por ejemplo: a ( dx idx j ) + b ( dx jdx i ) = ( a - b) dx idx j . El producto exterior permite construir formas diferenciales de mayor grado a partir de las de menor grado, de la misma manera que el producto cruzado en el cálculo vectorial permite calcular el vector de área de un paralelogramo a partir de vectores que apuntan hacia los dos lados. Alternar también implica que dx idx i = 0 , de la misma manera que el producto cruzado de vectores paralelos, cuya magnitud es el área del paralelogramo generado por esos vectores, es cero. En dimensiones superiores, dx i 1 ∧ ⋅⋅⋅ ∧ dx i m = 0 si dos de los índicesi 1 , ..., i m son iguales, de la misma manera que el "volumen" encerrado por un paralelootopo cuyos vectores de borde son linealmente dependientes es cero.

Notación de índices múltiples [ editar ]

A common notation for the wedge product of elementary k-forms is so called multi-index notation: in an n-dimensional context, for , we define .[2] Another useful notation is obtained by defining the set of all strictly increasing multi-indices of length k, in a space of dimension n, denoted . Then locally (wherever the coordinates apply), spans the space of differential k-forms in a manifold M of dimension n, when viewed as a module over the ring C(M) of smooth functions on M. By calculating the size of combinatorially, the module of k-forms on a n-dimensional manifold, and in general space of k-covectors on an n-dimensional vector space, is n choose k: . This also demonstrates that there are no nonzero differential forms of degree greater than the dimension of the underlying manifold.

The exterior derivative[edit]

In addition to the exterior product, there is also the exterior derivative operator d. The exterior derivative of a differential form is a generalization of the differential of a function, in the sense that the exterior derivative of fC(M) = Ω0(M) is exactly the differential of f. When generalized to higher forms, if ω = f dxI is a simple k-form, then its exterior derivative is a (k + 1)-form defined by taking the differential of the coefficient functions:

with extension to general k-forms through linearity: if , then its exterior derivative is

In R3, with the Hodge star operator, the exterior derivative corresponds to gradient, curl, and divergence, although this correspondence, like the cross product, does not generalize to higher dimensions, and should be treated with some caution.

The exterior derivative itself applies in an arbitrary finite number of dimensions, and is a flexible and powerful tool with wide application in differential geometry, differential topology, and many areas in physics. Of note, although the above definition of the exterior derivative was defined with respect to local coordinates, it can be defined in an entirely coordinate-free manner, as an antiderivation of degree 1 on the exterior algebra of differential forms. The benefit of this more general approach is that it allows for a natural coordinate-free approach to integration on manifolds. It also allows for a natural generalization of the fundamental theorem of calculus, called the (generalized) Stokes' theorem, which is a central result in the theory of integration on manifolds.

Differential calculus[edit]

Let U be an open set in Rn. A differential 0-form ("zero-form") is defined to be a smooth function f on U – the set of which is denoted C(U). If v is any vector in Rn, then f has a directional derivative v f, which is another function on U whose value at a point pU is the rate of change (at p) of f in the v direction:

(This notion can be extended point-wise to the case that v is a vector field on U by evaluating v at the point p in the definition.)

In particular, if v = ej is the jth coordinate vector then v f is the partial derivative of f with respect to the jth coordinate function, i.e., f / ∂xj, where x1, x2, ..., xn are the coordinate functions on U. By their very definition, partial derivatives depend upon the choice of coordinates: if new coordinates y1, y2, ..., yn are introduced, then

The first idea leading to differential forms is the observation that v f (p) is a linear function of v:

for any vectors v, w and any real number c. At each point p, this linear map from Rn to R is denoted dfp and called the derivative or differential of f at p. Thus dfp(v) = ∂v f (p). Extended over the whole set, the object df can be viewed as a function that takes a vector field on U, and returns a real-valued function whose value at each point is the derivative along the vector field of the function f. Note that at each p, the differential dfp is not a real number, but a linear functional on tangent vectors, and a prototypical example of a differential 1-form.

Since any vector v is a linear combination vjej of its components, df is uniquely determined by dfp(ej) for each j and each pU, which are just the partial derivatives of f on U. Thus df provides a way of encoding the partial derivatives of f. It can be decoded by noticing that the coordinates x1, x2, ..., xn are themselves functions on U, and so define differential 1-forms dx1, dx2, ..., dxn. Let f = xi. Since xi / ∂xj = δij, the Kronecker delta function, it follows that

The meaning of this expression is given by evaluating both sides at an arbitrary point p: on the right hand side, the sum is defined "pointwise", so that

Applying both sides to ej, the result on each side is the jth partial derivative of f at p. Since p and j were arbitrary, this proves the formula (*).

More generally, for any smooth functions gi and hi on U, we define the differential 1-form α = ∑i gi dhi pointwise by

for each pU. Any differential 1-form arises this way, and by using (*) it follows that any differential 1-form α on U may be expressed in coordinates as

for some smooth functions fi on U.

The second idea leading to differential forms arises from the following question: given a differential 1-form α on U, when does there exist a function f on U such that α = df? The above expansion reduces this question to the search for a function f whose partial derivatives f / ∂xi are equal to n given functions fi. For n > 1, such a function does not always exist: any smooth function f satisfies

so it will be impossible to find such an f unless

for all i and j.

The skew-symmetry of the left hand side in i and j suggests introducing an antisymmetric product on differential 1-forms, the exterior product, so that these equations can be combined into a single condition

where is defined so that:

This is an example of a differential 2-form. This 2-form is called the exterior derivative of α = ∑n
j=1
fj dxj
. It is given by

To summarize: = 0 is a necessary condition for the existence of a function f with α = df.

Differential 0-forms, 1-forms, and 2-forms are special cases of differential forms. For each k, there is a space of differential k-forms, which can be expressed in terms of the coordinates as

for a collection of functions fi1i2⋅⋅⋅ik. Antisymmetry, which was already present for 2-forms, makes it possible to restrict the sum to those sets of indices for which i1 < i2 < ... < ik−1 < ik.

Differential forms can be multiplied together using the exterior product, and for any differential k-form α, there is a differential (k + 1)-form called the exterior derivative of α.

Differential forms, the exterior product and the exterior derivative are independent of a choice of coordinates. Consequently, they may be defined on any smooth manifold M. One way to do this is cover M with coordinate charts and define a differential k-form on M to be a family of differential k-forms on each chart which agree on the overlaps. However, there are more intrinsic definitions which make the independence of coordinates manifest.

Intrinsic definitions[edit]

Let M be a smooth manifold. A smooth differential form of degree k is a smooth section of the kth exterior power of the cotangent bundle of M. The set of all differential k-forms on a manifold M is a vector space, often denoted Ωk(M).

The definition of a differential form may be restated as follows. At any point pM, a k-form β defines an element

where TpM is the tangent space to M at p and Tp*M is its dual space. This space is naturally isomorphic to the fiber at p of the dual bundle of the kth exterior power of the tangent bundle of M. That is, β is also a linear functional , i.e. the dual of the kth exterior power is isomorphic to the kth exterior power of the dual:

By the universal property of exterior powers, this is equivalently an alternating multilinear map:

Consequently, a differential k-form may be evaluated against any k-tuple of tangent vectors to the same point p of M. For example, a differential 1-form α assigns to each point pM a linear functional αp on TpM. In the presence of an inner product on TpM (induced by a Riemannian metric on M), αp may be represented as the inner product with a tangent vector Xp. Differential 1-forms are sometimes called covariant vector fields, covector fields, or "dual vector fields", particularly within physics.

The exterior algebra may be embedded in the tensor algebra by means of the alternation map. The alternation map is defined as a mapping

For a tensor at a point p,

where Sk is the symmetric group on k elements. The alternation map is constant on the cosets of the ideal in the tensor algebra generated by the symmetric 2-forms, and therefore descends to an embedding

This map exhibits β as a totally antisymmetric covariant tensor field of rank k. The differential forms on M are in one-to-one correspondence with such tensor fields.

Operations[edit]

As well as the addition and multiplication by scalar operations which arise from the vector space structure, there are several other standard operations defined on differential forms. The most important operations are the exterior product of two differential forms, the exterior derivative of a single differential form, the interior product of a differential form and a vector field, the Lie derivative of a differential form with respect to a vector field and the covariant derivative of a differential form with respect to a vector field on a manifold with a defined connection.

Exterior product[edit]

The exterior product of a k-form α and an -form β, denoted αβ, is a (k + )-form. At each point p of the manifold M, the forms α and β are elements of an exterior power of the cotangent space at p. When the exterior algebra is viewed as a quotient of the tensor algebra, the exterior product corresponds to the tensor product (modulo the equivalence relation defining the exterior algebra).

The antisymmetry inherent in the exterior algebra means that when αβ is viewed as a multilinear functional, it is alternating. However, when the exterior algebra embedded a subspace of the tensor algebra by means of the alternation map, the tensor product αβ is not alternating. There is an explicit formula which describes the exterior product in this situation. The exterior product is

This description is useful for explicit computations. For example, if k = = 1, then αβ is the 2-form whose value at a point p is the alternating bilinear form defined by

for v, w ∈ TpM.

The exterior product is bilinear: If α, β, and γ are any differential forms, and if f is any smooth function, then

It is skew commutative (also known as graded commutative), meaning that it satisfies a variant of anticommutativity that depends on the degrees of the forms: if α is a k-form and β is an -form, then

Riemannian manifold[edit]

On a Riemannian manifold, or more generally a pseudo-Riemannian manifold, the metric defines a fibre-wise isomorphism of the tangent and cotangent bundles. This makes it possible to convert vector fields to covector fields and vice versa. It also enables the definition of additional operations such as the Hodge star operator and the codifferential , which has degree −1 and is adjoint to the exterior differential d.

Vector field structures[edit]

On a pseudo-Riemannian manifold, 1-forms can be identified with vector fields; vector fields have additional distinct algebraic structures, which are listed here for context and to avoid confusion.

Firstly, each (co)tangent space generates a Clifford algebra, where the product of a (co)vector with itself is given by the value of a quadratic form – in this case, the natural one induced by the metric. This algebra is distinct from the exterior algebra of differential forms, which can be viewed as a Clifford algebra where the quadratic form vanishes (since the exterior product of any vector with itself is zero). Clifford algebras are thus non-anticommutative ("quantum") deformations of the exterior algebra. They are studied in geometric algebra.

Another alternative is to consider vector fields as derivations. The (noncommutative) algebra of differential operators they generate is the Weyl algebra and is a noncommutative ("quantum") deformation of the symmetric algebra in the vector fields.

Exterior differential complex[edit]

One important property of the exterior derivative is that d2 = 0. This means that the exterior derivative defines a cochain complex:

This complex is called the de Rham complex, and its cohomology is by definition the de Rham cohomology of M. By the Poincaré lemma, the de Rham complex is locally exact except at Ω0(M). The kernel at Ω0(M) is the space of locally constant functions on M. Therefore, the complex is a resolution of the constant sheaf R, which in turn implies a form of de Rham's theorem: de Rham cohomology computes the sheaf cohomology of R.

Pullback[edit]

Suppose that f : MN is smooth. The differential of f is a smooth map df : TMTN between the tangent bundles of M and N. This map is also denoted f and called the pushforward. For any point pM and any vTpM, there is a well-defined pushforward vector f(v) in Tf(p)N. However, the same is not true of a vector field. If f is not injective, say because qN has two or more preimages, then the vector field may determine two or more distinct vectors in TqN. If f is not surjective, then will be a point qN at which f does not determine any tangent vector at all. Since a vector field on N determines, by definition, a unique tangent vector at every point of N, the pushforward of a vector field does not always exist.

By contrast, it is always possible to pull back a differential form. A differential form on N may be viewed as a linear functional on each tangent space. Precomposing this functional with the differential df : TMTN defines a linear functional on each tangent space of M and therefore a differential form on M. The existence of pullbacks is one of the key features of the theory of differential forms. It leads to the existence of pullback maps in other situations, such as pullback homomorphisms in de Rham cohomology.

Formally, let f : MN be smooth, and let ω be a smooth k-form on N. Then there is a differential form fω on M, called the pullback of ω, which captures the behavior of ω as seen relative to f. To define the pullback, fix a point p of M and tangent vectors v1, ..., vk to M at p. The pullback of ω is defined by the formula

There are several more abstract ways to view this definition. If ω is a 1-form on N, then it may be viewed as a section of the cotangent bundle TN of N. Using to denote a dual map, the dual to the differential of f is (df) : TNTM. The pullback of ω may be defined to be the composite

This is a section of the cotangent bundle of M and hence a differential 1-form on M. In full generality, let denote the kth exterior power of the dual map to the differential. Then the pullback of a k-form ω is the composite

Another abstract way to view the pullback comes from viewing a k-form ω as a linear functional on tangent spaces. From this point of view, ω is a morphism of vector bundles

where N × R is the trivial rank one bundle on N. The composite map

defines a linear functional on each tangent space of M, and therefore it factors through the trivial bundle M × R. The vector bundle morphism defined in this way is fω.

Pullback respects all of the basic operations on forms. If ω and η are forms and c is a real number, then

The pullback of a form can also be written in coordinates. Assume that x1, ..., xm are coordinates on M, that y1, ..., yn are coordinates on N, and that these coordinate systems are related by the formulas yi = fi(x1, ..., xm) for all i. Locally on N, ω can be written as

where, for each choice of i1, ..., ik, ωi1⋅⋅⋅ik is a real-valued function of y1, ..., yn. Using the linearity of pullback and its compatibility with exterior product, the pullback of ω has the formula

Each exterior derivative dfi can be expanded in terms of dx1, ..., dxm. The resulting k-form can be written using Jacobian matrices:

Here, denotes the determinant of the matrix whose entries are , .

Integration[edit]

A differential k-form can be integrated over an oriented k-dimensional manifold. When the k-form is defined on an n-dimensional manifold with n > k, then the k-form can be integrated over oriented k-dimensional submanifolds. If k = 0, integration over oriented 0-dimensional submanifolds is just the summation of the integrand evaluated at points, with according to the orientation of those points. Other values of k = 1, 2, 3, ... correspond to line integrals, surface integrals, volume integrals, and so on. There are several equivalent ways to formally define the integral of a differential form, all of which depend on reducing to the case of Euclidean space.

Integration on Euclidean space[edit]

Let U be an open subset of Rn. Give Rn its standard orientation and U the restriction of that orientation. Every smooth n-form ω on U has the form

for some smooth function f : RnR. Such a function has an integral in the usual Riemann or Lebesgue sense. This allows us to define the integral of ω to be the integral of f:

Fixing an orientation is necessary for this to be well-defined. The skew-symmetry of differential forms means that the integral of, say, dx1dx2 must be the negative of the integral of dx2dx1. Riemann and Lebesgue integrals cannot see this dependence on the ordering of the coordinates, so they leave the sign of the integral undetermined. The orientation resolves this ambiguity.

Integration over chains[edit]

Let M be an n-manifold and ω an n-form on M. First, assume that there is a parametrization of M by an open subset of Euclidean space. That is, assume that there exists a diffeomorphism

where DRn. Give M the orientation induced by φ. Then (Rudin 1976) defines the integral of ω over M to be the integral of φω over D. In coordinates, this has the following expression. Fix a chart on M with coordinates x1, ..., xn. Then

Suppose that φ is defined by

Then the integral may be written in coordinates as

where

is the determinant of the Jacobian. The Jacobian exists because φ is differentiable.

In general, an n-manifold cannot be parametrized by an open subset of Rn. But such a parametrization is always possible locally, so it is possible to define integrals over arbitrary manifolds by defining them as sums of integrals over collections of local parametrizations. Moreover, it is also possible to define parametrizations of k-dimensional subsets for k < n, and this makes it possible to define integrals of k-forms. To make this precise, it is convenient to fix a standard domain D in Rk, usually a cube or a simplex. A k-chain is a formal sum of smooth embeddings DM. That is, it is a collection of smooth embeddings, each of which is assigned an integer multiplicity. Each smooth embedding determines a k-dimensional submanifold of M. If the chain is

then the integral of a k-form ω over c is defined to be the sum of the integrals over the terms of c:

This approach to defining integration does not assign a direct meaning to integration over the whole manifold M. However, it is still possible to assign such a meaning indirectly because every smooth manifold may be smoothly triangulated in an essentially unique way, and the integral over M may be defined to be the integral over the chain determined by a triangulation.

Integration using partitions of unity[edit]

There is another approach, expounded in (Dieudonne 1972), which does directly assign a meaning to integration over M, but this approach requires fixing an orientation of M. The integral of an n-form ω on an n-dimensional manifold is defined by working in charts. Suppose first that ω is supported on a single positively oriented chart. On this chart, it may be pulled back to an n-form on an open subset of Rn. Here, the form has a well-defined Riemann or Lebesgue integral as before. The change of variables formula and the assumption that the chart is positively oriented together ensure that the integral of ω is independent of the chosen chart. In the general case, use a partition of unity to write ω as a sum of n-forms, each of which is supported in a single positively oriented chart, and define the integral of ω to be the sum of the integrals of each term in the partition of unity.

It is also possible to integrate k-forms on oriented k-dimensional submanifolds using this more intrinsic approach. The form is pulled back to the submanifold, where the integral is defined using charts as before. For example, given a path γ(t) : [0, 1] → R2, integrating a 1-form on the path is simply pulling back the form to a form f(t) dt on [0, 1], and this integral is the integral of the function f(t) on the interval.

Integration along fibers[edit]

Fubini's theorem states that the integral over a set that is a product may be computed as an iterated integral over the two factors in the product. This suggests that the integral of a differential form over a product ought to be computable as an iterated integral as well. The geometric flexibility of differential forms ensures that this is possible not just for products, but in more general situations as well. Under some hypotheses, it is possible to integrate along the fibers of a smooth map, and the analog of Fubini's theorem is the case where this map is the projection from a product to one of its factors.

Because integrating a differential form over a submanifold requires fixing an orientation, a prerequisite to integration along fibers is the existence of a well-defined orientation on those fibers. Let M and N be two orientable manifolds of pure dimensions m and n, respectively. Suppose that f : MN is a surjective submersion. This implies that each fiber f−1(y) is (mn)-dimensional and that, around each point of M, there is a chart on which f looks like the projection from a product onto one of its factors. Fix xM and set y = f(x). Suppose that

and that ηy does not vanish. Following (Dieudonne 1972), there is a unique

which may be thought of as the fibral part of ωx with respect to ηy. More precisely, define j : f−1(y) → M to be the inclusion. Then σx is defined by the property that

where

is any (mn)-covector for which

The form σx may also be notated ωx / ηy.

Moreover, for fixed y, σx varies smoothly with respect to x. That is, suppose that

is a smooth section of the projection map; we say that ω is a smooth differential m-form on M along f−1(y). Then there is a smooth differential (mn)-form σ on f−1(y) such that, at each xf−1(y),

This form is denoted ω / ηy. The same construction works if ω is an m-form in a neighborhood of the fiber, and the same notation is used. A consequence is that each fiber f−1(y) is orientable. In particular, a choice of orientation forms on M and N defines an orientation of every fiber of f.

The analog of Fubini's theorem is as follows. As before, M and N are two orientable manifolds of pure dimensions m and n, and f : MN is a surjective submersion. Fix orientations of M and N, and give each fiber of f the induced orientation. Let θ be an m-form on M, and let ζ be an n-form on N that is almost everywhere positive with respect to the orientation of N. Then, for almost every yN, the form θ / ζy is a well-defined integrable mn form on f−1(y). Moreover, there is an integrable n-form on N defined by

Denote this form by

Then (Dieudonne 1972) proves the generalized Fubini formula

It is also possible to integrate forms of other degrees along the fibers of a submersion. Assume the same hypotheses as before, and let α be a compactly supported (mn + k)-form on M. Then there is a k-form γ on N which is the result of integrating α along the fibers of f. The form α is defined by specifying, at each yN, how α pairs against each k-vector v at y, and the value of that pairing is an integral over f−1(y) that depends only on α, v, and the orientations of M and N. More precisely, at each yN, there is an isomorphism

defined by the interior product

If xf−1(y), then a k-vector v at y determines an (mk)-covector at x by pullback:

Each of these covectors has an exterior product against α, so there is an (mn)-form βv on M along f−1(y) defined by

This form depends on the orientation of N but not the choice of ζ. Then the k-form γ is uniquely defined by the property

and γ is smooth (Dieudonne 1972). This form also denoted α and called the integral of α along the fibers of f. Integration along fibers is important for the construction of Gysin maps in de Rham cohomology.

Integration along fibers satisfies the projection formula (Dieudonne 1972). If λ is any -form on N, then

Stokes's theorem[edit]

The fundamental relationship between the exterior derivative and integration is given by the Stokes' theorem: If ω is an (n − 1)-form with compact support on M and ∂M denotes the boundary of M with its induced orientation, then

A key consequence of this is that "the integral of a closed form over homologous chains is equal": If ω is a closed k-form and M and N are k-chains that are homologous (such that MN is the boundary of a (k + 1)-chain W), then , since the difference is the integral .

For example, if ω = df is the derivative of a potential function on the plane or Rn, then the integral of ω over a path from a to b does not depend on the choice of path (the integral is f(b) − f(a)), since different paths with given endpoints are homotopic, hence homologous (a weaker condition). This case is called the gradient theorem, and generalizes the fundamental theorem of calculus. This path independence is very useful in contour integration.

This theorem also underlies the duality between de Rham cohomology and the homology of chains.

Relation with measures[edit]

On a general differentiable manifold (without additional structure), differential forms cannot be integrated over subsets of the manifold; this distinction is key to the distinction between differential forms, which are integrated over chains or oriented submanifolds, and measures, which are integrated over subsets. The simplest example is attempting to integrate the 1-form dx over the interval [0, 1]. Assuming the usual distance (and thus measure) on the real line, this integral is either 1 or −1, depending on orientation: , while . By contrast, the integral of the measure |dx| on the interval is unambiguously 1 (i.e. the integral of the constant function 1 with respect to this measure is 1). Similarly, under a change of coordinates a differential n-form changes by the Jacobian determinant J, while a measure changes by the absolute value of the Jacobian determinant, |J|, which further reflects the issue of orientation. For example, under the map x ↦ −x on the line, the differential form dx pulls back to dx; orientation has reversed; while the Lebesgue measure, which here we denote |dx|, pulls back to |dx|; it does not change.

In the presence of the additional data of an orientation, it is possible to integrate n-forms (top-dimensional forms) over the entire manifold or over compact subsets; integration over the entire manifold corresponds to integrating the form over the fundamental class of the manifold, [M]. Formally, in the presence of an orientation, one may identify n-forms with densities on a manifold; densities in turn define a measure, and thus can be integrated (Folland 1999, Section 11.4, pp. 361–362).

On an orientable but not oriented manifold, there are two choices of orientation; either choice allows one to integrate n-forms over compact subsets, with the two choices differing by a sign. On non-orientable manifold, n-forms and densities cannot be identified —notably, any top-dimensional form must vanish somewhere (there are no volume forms on non-orientable manifolds), but there are nowhere-vanishing densities— thus while one can integrate densities over compact subsets, one cannot integrate n-forms. One can instead identify densities with top-dimensional pseudoforms.

Even in the presence of an orientation, there is in general no meaningful way to integrate k-forms over subsets for k < n because there is no consistent way to use the ambient orientation to orient k-dimensional subsets. Geometrically, a k-dimensional subset can be turned around in place, yielding the same subset with the opposite orientation; for example, the horizontal axis in a plane can be rotated by 180 degrees. Compare the Gram determinant of a set of k vectors in an n-dimensional space, which, unlike the determinant of n vectors, is always positive, corresponding to a squared number. An orientation of a k-submanifold is therefore extra data not derivable from the ambient manifold.

On a Riemannian manifold, one may define a k-dimensional Hausdorff measure for any k (integer or real), which may be integrated over k-dimensional subsets of the manifold. A function times this Hausdorff measure can then be integrated over k-dimensional subsets, providing a measure-theoretic analog to integration of k-forms. The n-dimensional Hausdorff measure yields a density, as above.

Currents[edit]

The differential form analog of a distribution or generalized function is called a current. The space of k-currents on M is the dual space to an appropriate space of differential k-forms. Currents play the role of generalized domains of integration, similar to but even more flexible than chains.

Applications in physics[edit]

Differential forms arise in some important physical contexts. For example, in Maxwell's theory of electromagnetism, the Faraday 2-form, or electromagnetic field strength, is

where the fab are formed from the electromagnetic fields and ; e.g., f12 = Ez/c, f23 = −Bz, or equivalent definitions.

This form is a special case of the curvature form on the U(1) principal bundle on which both electromagnetism and general gauge theories may be described. The connection form for the principal bundle is the vector potential, typically denoted by A, when represented in some gauge. One then has

The current 3-form is

where ja are the four components of the current density. (Here it is a matter of convention to write Fab instead of fab, i.e. to use capital letters, and to write Ja instead of ja. However, the vector rsp. tensor components and the above-mentioned forms have different physical dimensions. Moreover, by decision of an international commission of the International Union of Pure and Applied Physics, the magnetic polarization vector is called since several decades, and by some publishers J, i.e. the same name is used for different quantities.)

Using the above-mentioned definitions, Maxwell's equations can be written very compactly in geometrized units as

where denotes the Hodge star operator. Similar considerations describe the geometry of gauge theories in general.

The 2-form , which is dual to the Faraday form, is also called Maxwell 2-form.

Electromagnetism is an example of a U(1) gauge theory. Here the Lie group is U(1), the one-dimensional unitary group, which is in particular abelian. There are gauge theories, such as Yang–Mills theory, in which the Lie group is not abelian. In that case, one gets relations which are similar to those described here. The analog of the field F in such theories is the curvature form of the connection, which is represented in a gauge by a Lie algebra-valued one-form A. The Yang–Mills field F is then defined by

In the abelian case, such as electromagnetism, AA = 0, but this does not hold in general. Likewise the field equations are modified by additional terms involving exterior products of A and F, owing to the structure equations of the gauge group.

Applications in geometric measure theory[edit]

Numerous minimality results for complex analytic manifolds are based on the Wirtinger inequality for 2-forms. A succinct proof may be found in Herbert Federer's classic text Geometric Measure Theory. The Wirtinger inequality is also a key ingredient in Gromov's inequality for complex projective space in systolic geometry.

See also[edit]

  • Closed and exact differential forms
  • Complex differential form
  • Vector-valued differential form
  • Equivariant differential form
  • Calculus on Manifolds
  • Multilinear form
  • Polynomial differential form

Notes[edit]

  1. ^ Cartan, Élie (1899), "Sur certaines expressions différentielles et le problème de Pfaff", Annales Scientifiques de l'École Normale Supérieure: 239–332
  2. ^ Tu, Loring W. (2011). An introduction to manifolds (2nd ed.). New York: Springer. ISBN 9781441974006. OCLC 682907530.

References[edit]

  • Bachman, David (2006), A Geometric Approach to Differential Forms, Birkhäuser, ISBN 978-0-8176-4499-4
  • Bachman, David (2003), A Geometric Approach to Differential Forms, arXiv:math/0306194v1, Bibcode:2003math......6194B
  • Cartan, Henri (2006), Differential Forms, Dover, ISBN 0-486-45010-4—Translation of Formes différentielles (1967)
  • Dieudonné, Jean (1972), Treatise on Analysis, 3, New York-London: Academic Press, Inc., MR 0350769
  • Edwards, Harold M. (1994), Advanced Calculus; A Differential Forms Approach, Boston, Basel, Berlin: Birkhäuser, doi:10.1007/978-0-8176-8412-9, ISBN 978-0-8176-8411-2
  • Folland, Gerald B. (1999), Real Analysis: Modern Techniques and Their Applications (Second ed.), ISBN 978-0-471-31716-6, provides a brief discussion of integration on manifolds from the point of view of measure theory in the last section.CS1 maint: postscript (link)
  • Flanders, Harley (1989) [1964], Differential forms with applications to the physical sciences, Mineola, New York: Dover Publications, ISBN 0-486-66169-5
  • Fleming, Wendell H. (1965), "Chapter 6: Exterior algebra and differential calculus", Functions of Several Variables, Addison-Wesley, pp. 205–238. This textbook in multivariate calculus introduces the exterior algebra of differential forms at the college calculus level.CS1 maint: postscript (link)
  • Morita, Shigeyuki (2001), Geometry of Differential Forms, AMS, ISBN 0-8218-1045-6
  • Rudin, Walter (1976), Principles of Mathematical Analysis, New York: McGraw-Hill, ISBN 0-07-054235-X
  • Spivak, Michael (1965), Calculus on Manifolds, Menlo Park, California: W. A. Benjamin, ISBN 0-8053-9021-9, standard introductory textCS1 maint: postscript (link)
  • Tu, Loring W. (2008), An Introduction to Manifolds, Universitext, Springer, doi:10.1007/978-1-4419-7400-6, ISBN 978-0-387-48098-5
  • Zorich, Vladimir A. (2004), Mathematical Analysis II, Springer, ISBN 3-540-40633-6

External links[edit]

  • Weisstein, Eric W. "Differential form". MathWorld.
  • Sjamaar, Reyer (2006), Manifolds and differential forms lecture notes, a course taught at Cornell University.
  • Bachman, David (2003), A Geometric Approach to Differential Forms, arXiv:math/0306194, Bibcode:2003math......6194B, an undergraduate text.
  • Jones, Frank, Integration on manifolds (PDF)