Differential form

In mathematics, differential forms provide a unified approach to define integrands over curves, surfaces, solids, and higher-dimensional manifolds. The modern notion of differential forms was pioneered by Élie Cartan. It has many applications, especially in geometry, topology and physics.

For instance, the expression f(x) dx is an example of a 1-form, and can be integrated over an interval [a, b] contained in the domain of f:

{\displaystyle \int _{a}^{b}f(x),dx.}\int _{a}^{b}f(x),dx.
Similarly, the expression f(x, y, z) dx ∧ dy + g(x, y, z) dz ∧ dx + h(x, y, z) dy ∧ dz is a 2-form that can be integrated over a surface S:

{\displaystyle \int _{S}(f(x,y,z),dx\wedge dy+g(x,y,z),dz\wedge dx+h(x,y,z),dy\wedge dz).}{\displaystyle \int _{S}(f(x,y,z),dx\wedge dy+g(x,y,z),dz\wedge dx+h(x,y,z),dy\wedge dz).}
The symbol ∧ denotes the exterior product, sometimes called the wedge product, of two differential forms. Likewise, a 3-form f(x, y, z) dx ∧ dy ∧ dz represents a volume element that can be integrated over a region of space. In general, a k-form is an object that may be integrated over a k-dimensional manifold, and is homogeneous of degree k in the coordinate differentials {\displaystyle dx,dy,\ldots .}{\displaystyle dx,dy,\ldots .} On an n-dimensional manifold, the top-dimensional form (n-form) is called a volume form.

The differential forms form an alternating algebra. This implies that {\displaystyle dy\wedge dx=-dx\wedge dy}{\displaystyle dy\wedge dx=-dx\wedge dy} and {\displaystyle dx\wedge dx=0.}{\displaystyle dx\wedge dx=0.} This alternating property reflects the orientation of the domain of integration.

The exterior derivative is an operation on differential forms that, given a k-form {\displaystyle \varphi }\varphi , produces a (k+1)-form {\displaystyle d\varphi .}{\displaystyle d\varphi .} This operation extends the differential of a function (a function can be considered as a 0-form, and its differential is {\displaystyle df(x)=f’(x)dx.}{\displaystyle df(x)=f’(x)dx.}) This allows expressing the fundamental theorem of calculus, the divergence theorem, Green’s theorem, and Stokes’ theorem as special cases of a single general result, the generalized Stokes theorem.

Differential 1-forms are naturally dual to vector fields on a differentiable manifold, and the pairing between vector fields and 1-forms is extended to arbitrary differential forms by the interior product. The algebra of differential forms along with the exterior derivative defined on it is preserved by the pullback under smooth functions between two manifolds. This feature allows geometrically invariant information to be moved from one space to another via the pullback, provided that the information is expressed in terms of differential forms. As an example, the change of variables formula for integration becomes a simple statement that an integral is preserved under pullback.

1 History

Differential forms are part of the field of differential geometry, influenced by linear algebra. Although the notion of a differential is quite old, the initial attempt at an algebraic organization of differential forms is usually credited to Élie Cartan with reference to his 1899 paper.[1] Some aspects of the exterior algebra of differential forms appears in Hermann Grassmann’s 1844 work, Die Lineale Ausdehnungslehre, ein neuer Zweig der Mathematik (The Theory of Linear Extension, a New Branch of Mathematics).

2 Concept

Differential forms provide an approach to multivariable calculus that is independent of coordinates.

2.1 Integration and orientation

A differential k-form can be integrated over an oriented manifold of dimension k. A differential 1-form can be thought of as measuring an infinitesimal oriented length, or 1-dimensional oriented density. A differential 2-form can be thought of as measuring an infinitesimal oriented area, or 2-dimensional oriented density. And so on.

Integration of differential forms is well-defined only on oriented manifolds. An example of a 1-dimensional manifold is an interval [a, b], and intervals can be given an orientation: they are positively oriented if a < b, and negatively oriented otherwise. If a < b then the integral of the differential 1-form f(x) dx over the interval [a, b] (with its natural positive orientation) is

{\displaystyle \int _{a}^{b}f(x),dx}{\displaystyle \int _{a}^{b}f(x),dx}
which is the negative of the integral of the same differential form over the same interval, when equipped with the opposite orientation. That is:

{\displaystyle \int _{b}^{a}f(x),dx=-\int _{a}^{b}f(x),dx.}{\displaystyle \int _{b}^{a}f(x),dx=-\int _{a}^{b}f(x),dx.}
This gives a geometrical context to the conventions for one-dimensional integrals, that the sign changes when the orientation of the interval is reversed. A standard explanation of this in one-variable integration theory is that, when the limits of integration are in the opposite order (b < a), the increment dx is negative in the direction of integration.

More generally, an m-form is an oriented density that can be integrated over an m-dimensional oriented manifold. (For example, a 1-form can be integrated over an oriented curve, a 2-form can be integrated over an oriented surface, etc.) If M is an oriented m-dimensional manifold, and M′ is the same manifold with opposite orientation and ω is an m-form, then one has:

{\displaystyle \int _{M}\omega =-\int _{M’}\omega ,.}\int _{M}\omega =-\int _{M’}\omega ,.
These conventions correspond to interpreting the integrand as a differential form, integrated over a chain. In measure theory, by contrast, one interprets the integrand as a function f with respect to a measure μ and integrates over a subset A, without any notion of orientation; one writes {\textstyle \int _{A}f,d\mu =\int _{[a,b]}f,d\mu }{\textstyle \int _{A}f,d\mu =\int _{[a,b]}f,d\mu } to indicate integration over a subset A. This is a minor distinction in one dimension, but becomes subtler on higher-dimensional manifolds; see below for details.

Making the notion of an oriented density precise, and thus of a differential form, involves the exterior algebra. The differentials of a set of coordinates, dx1, …, dxn can be used as a basis for all 1-forms. Each of these represents a covector at each point on the manifold that may be thought of as measuring a small displacement in the corresponding coordinate direction. A general 1-form is a linear combination of these differentials at every point on the manifold:

{\displaystyle f_{1},dx^{1}+\cdots +f_{n},dx^{n},}{\displaystyle f_{1},dx^{1}+\cdots +f_{n},dx^{n},}
where the fk = fk(x1, … , xn) are functions of all the coordinates. A differential 1-form is integrated along an oriented curve as a line integral.

The expressions dxi ∧ dxj, where i < j can be used as a basis at every point on the manifold for all 2-forms. This may be thought of as an infinitesimal oriented square parallel to the xi–xj-plane. A general 2-form is a linear combination of these at every point on the manifold: {\textstyle \sum {1\leq i<j\leq n}f{i,j},dx^{i}\wedge dx^{j}}{\textstyle \sum {1\leq i<j\leq n}f{i,j},dx^{i}\wedge dx^{j}}, and it is integrated just like a surface integral.

A fundamental operation defined on differential forms is the exterior product (the symbol is the wedge ∧). This is similar to the cross product from vector calculus, in that it is an alternating product. For instance,

{\displaystyle dx^{1}\wedge dx{2}=-dx{2}\wedge dx{1}}dx{1}\wedge dx{2}=-dx{2}\wedge dx^{1}
because the square whose first side is dx1 and second side is dx2 is to be regarded as having the opposite orientation as the square whose first side is dx2 and whose second side is dx1. This is why we only need to sum over expressions dxi ∧ dxj, with i < j; for example: a(dxi ∧ dxj) + b(dxj ∧ dxi) = (a − b) dxi ∧ dxj. The exterior product allows higher-degree differential forms to be built out of lower-degree ones, in much the same way that the cross product in vector calculus allows one to compute the area vector of a parallelogram from vectors pointing up the two sides. Alternating also implies that dxi ∧ dxi = 0, in the same way that the cross product of parallel vectors, whose magnitude is the area of the parallelogram spanned by those vectors, is zero. In higher dimensions, dxi1 ∧ ⋅⋅⋅ ∧ dxim = 0 if any two of the indices i1, …, im are equal, in the same way that the “volume” enclosed by a parallelotope whose edge vectors are linearly dependent is zero.

2.2 Multi-index notation

A common notation for the wedge product of elementary k-forms is so called multi-index notation: in an n-dimensional context, for {\displaystyle I=(i_{1},i_{2},\ldots ,i_{k}),1\leq i_{1}<i_{2}<\cdots <i_{k}\leq n}{\displaystyle I=(i_{1},i_{2},\ldots ,i_{k}),1\leq i_{1}<i_{2}<\cdots <i_{k}\leq n}, we define {\textstyle dx{I}:=dx{i_{1}}\wedge \cdots \wedge dx^{i_{k}}=\bigwedge {i\in I}dx^{i}}{\textstyle dx{I}:=dx{i{1}}\wedge \cdots \wedge dx^{i_{k}}=\bigwedge {i\in I}dx^{i}}.[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 {\displaystyle {\mathcal {J}}{k,n}:={I=(i_{1},\ldots ,i_{k}):1\leq i_{1}<i_{2}<\cdots <i_{k}\leq n}}{\displaystyle {\mathcal {J}}{k,n}:={I=(i{1},\ldots ,i_{k}):1\leq i_{1}<i_{2}<\cdots <i_{k}\leq n}}. Then locally (wherever the coordinates apply), {\displaystyle {dx^{I}}{I\in {\mathcal {J}}{k,n}}}{\displaystyle {dx^{I}}{I\in {\mathcal {J}}{k,n}}} 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 {\displaystyle {\mathcal {J}}{k,n}}{\displaystyle {\mathcal {J}}{k,n}} combinatorially, the module of k-forms on an n-dimensional manifold, and in general space of k-covectors on an n-dimensional vector space, is n choose k: {\textstyle |{\mathcal {J}}{k,n}|={\binom {n}{k}}}{\textstyle |{\mathcal {J}}{k,n}|={\binom {n}{k}}}. This also demonstrates that there are no nonzero differential forms of degree greater than the dimension of the underlying manifold.

2.3 The exterior derivative

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 f ∈ C∞(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 dω is a (k + 1)-form defined by taking the differential of the coefficient functions:

{\displaystyle d\omega =\sum {i=1}^{n}{\frac {\partial f}{\partial x{i}}},dx{i}\wedge dx^{I}.}{\displaystyle d\omega =\sum {i=1}^{n}{\frac {\partial f}{\partial x{i}}},dx{i}\wedge dx^{I}.}
with extension to general k-forms through linearity: if {\displaystyle \tau =\sum {I\in {\mathcal {J}}{k,n}}a
{I},dx^{I}\in \Omega ^{k}(M)}{\displaystyle \tau =\sum {I\in {\mathcal {J}}{k,n}}a
{I},dx^{I}\in \Omega ^{k}(M)}, then its exterior derivative is

{\displaystyle d\tau =\sum {I\in {\mathcal {J}}{k,n}}\left(\sum {j=1}^{n}{\frac {\partial a{I}}{\partial x{j}}},dx{j}\right)\wedge dx^{I}\in \Omega ^{k+1}(M)}{\displaystyle d\tau =\sum {I\in {\mathcal {J}}{k,n}}\left(\sum {j=1}^{n}{\frac {\partial a{I}}{\partial x{j}}},dx{j}\right)\wedge dx^{I}\in \Omega ^{k+1}(M)}
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 integrate 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.

2.4 Differential calculus

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 p ∈ U is the rate of change (at p) of f in the v direction:

{\displaystyle (\partial {v}f)§=\left.{\frac {d}{dt}}f(p+t\mathbf {v} )\right|{t=0}.}{\displaystyle (\partial {v}f)§=\left.{\frac {d}{dt}}f(p+t\mathbf {v} )\right|{t=0}.}
(This notion can be extended pointwise 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 vector, i.e., ∂f / ∂xj, where x1, x2, …, xn are the coordinate vectors in U. By their very definition, partial derivatives depend upon the choice of coordinates: if new coordinates y1, y2, …, yn are introduced, then

{\displaystyle {\frac {\partial f}{\partial x^{j}}}=\sum _{i=1}^{n}{\frac {\partial y^{i}}{\partial x^{j}}}{\frac {\partial f}{\partial y^{i}}}.}{\displaystyle {\frac {\partial f}{\partial x^{j}}}=\sum _{i=1}^{n}{\frac {\partial y^{i}}{\partial x^{j}}}{\frac {\partial f}{\partial y^{i}}}.}
The first idea leading to differential forms is the observation that ∂v f § is a linear function of v:

{\displaystyle {\begin{aligned}(\partial _{v+w}f)§&=(\partial _{v}f)§+(\partial _{w}f)§\(\partial _{cv}f)§&=c(\partial _{v}f)§\end{aligned}}}{\displaystyle {\begin{aligned}(\partial _{v+w}f)§&=(\partial _{v}f)§+(\partial _{w}f)§\(\partial _{cv}f)§&=c(\partial _{v}f)§\end{aligned}}}
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 §. 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 Σ vjej of its components, df is uniquely determined by dfp(ej) for each j and each p ∈ U, 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

{\displaystyle df=\sum _{i=1}^{n}{\frac {\partial f}{\partial x{i}}},dx{i}.}{\displaystyle df=\sum _{i=1}^{n}{\frac {\partial f}{\partial x{i}}},dx{i}.}

(*)

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

{\displaystyle df_{p}=\sum {i=1}^{n}{\frac {\partial f}{\partial x{i}}}§(dx{i}){p}.}{\displaystyle df_{p}=\sum {i=1}^{n}{\frac {\partial f}{\partial x{i}}}§(dx{i}){p}.}
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 = Σi gi dhi pointwise by

{\displaystyle \alpha {p}=\sum {i}g{i}§(dh{i}){p}}{\displaystyle \alpha {p}=\sum {i}g{i}§(dh{i}){p}}
for each p ∈ U. 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

{\displaystyle \alpha =\sum _{i=1}{n}f_{i},dx{i}}\alpha =\sum _{i=1}{n}f_{i},dx{i}
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

{\displaystyle {\frac {\partial ^{2}f}{\partial x^{i},\partial x^{j}}}={\frac {\partial ^{2}f}{\partial x^{j},\partial x^{i}}},}{\frac {\partial ^{2}f}{\partial x^{i},\partial x^{j}}}={\frac {\partial ^{2}f}{\partial x^{j},\partial x^{i}}},
so it will be impossible to find such an f unless

{\displaystyle {\frac {\partial f_{j}}{\partial x^{i}}}-{\frac {\partial f_{i}}{\partial x^{j}}}=0}{\displaystyle {\frac {\partial f_{j}}{\partial x^{i}}}-{\frac {\partial f_{i}}{\partial x^{j}}}=0}
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

{\displaystyle \sum {i,j=1}^{n}{\frac {\partial f{j}}{\partial x{i}}},dx{i}\wedge dx^{j}=0,}{\displaystyle \sum {i,j=1}^{n}{\frac {\partial f{j}}{\partial x{i}}},dx{i}\wedge dx^{j}=0,}
where ∧ is defined so that:

{\displaystyle dx^{i}\wedge dx{j}=-dx{j}\wedge dx^{i}.}{\displaystyle dx^{i}\wedge dx{j}=-dx{j}\wedge dx^{i}.}
This is an example of a differential 2-form. This 2-form is called the exterior derivative dα of α = Σn
j=1 fj dxj. It is given by

{\displaystyle d\alpha =\sum {j=1}^{n}df{j}\wedge dx^{j}=\sum {i,j=1}^{n}{\frac {\partial f{j}}{\partial x{i}}},dx{i}\wedge dx^{j}.}{\displaystyle d\alpha =\sum {j=1}^{n}df{j}\wedge dx^{j}=\sum {i,j=1}^{n}{\frac {\partial f{j}}{\partial x{i}}},dx{i}\wedge dx^{j}.}
To summarize: dα = 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

{\displaystyle \sum {i{1},i_{2}\ldots i_{k}=1}^{n}f_{i_{1}i_{2}\ldots i_{k}},dx^{i_{1}}\wedge dx^{i_{2}}\wedge \cdots \wedge dx^{i_{k}}}{\displaystyle \sum {i{1},i_{2}\ldots i_{k}=1}^{n}f_{i_{1}i_{2}\ldots i_{k}},dx^{i_{1}}\wedge dx^{i_{2}}\wedge \cdots \wedge dx^{i_{k}}}
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 dα 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.

3 Intrinsic definitions

4 Operations

4.1 Exterior product

4.2 Riemannian manifold

4.2.1 Vector field structures

4.3 Exterior differential complex

5 Pullback

6 Integration

6.1 Integration on Euclidean space

6.2 Integration over chains

6.3 Integration using partitions of unity

6.4 Integration along fibers

6.5 Stokes’s theorem

6.6 Relation with measures

6.7 Currents

7 Applications in physics

8 Applications in geometric measure theory

9 See also

  • 0
    点赞
  • 0
    收藏
    觉得还不错? 一键收藏
  • 0
    评论
评论
添加红包

请填写红包祝福语或标题

红包个数最小为10个

红包金额最低5元

当前余额3.43前往充值 >
需支付:10.00
成就一亿技术人!
领取后你会自动成为博主和红包主的粉丝 规则
hope_wisdom
发出的红包
实付
使用余额支付
点击重新获取
扫码支付
钱包余额 0

抵扣说明:

1.余额是钱包充值的虚拟货币,按照1:1的比例进行支付金额的抵扣。
2.余额无法直接购买下载,可以购买VIP、付费专栏及课程。

余额充值