Jump to content

Helmholtz decomposition: Difference between revisions

From Wikipedia, the free encyclopedia
Content deleted Content added
Hvh2023 (talk | contribs)
expanded, based on german wikipedia: added short history, clearly separated 3- from d-dimensional case, add options for fields not decaying at infinity, added applications, general cleanup of references. All references in the text.
Line 1: Line 1:
{{Short description|Certain vector fields are the sum of an irrotational and a solenoidal vector field}}
{{Short description|Certain vector fields are the sum of an irrotational and a solenoidal vector field}}
In [[physics]] and [[mathematics]], in the area of [[vector calculus]], '''Helmholtz's theorem''',<ref>On Helmholtz's Theorem in Finite Regions. By [[Jean Bladel]]. Midwestern Universities Research Association, 1958.</ref><ref>Hermann von Helmholtz. Clarendon Press, 1906. By [[Leo Koenigsberger]]. p357</ref> also known as the '''fundamental theorem of vector calculus''',<ref>An Elementary Course in the Integral Calculus. By [[Daniel Alexander Murray]]. American Book Company, 1898. p8.</ref><ref>[[J. W. Gibbs]] & [[Edwin Bidwell Wilson]] (1901) [https://archive.org/stream/117714283#page/236/mode/2up Vector Analysis], page 237, link from [[Internet Archive]]</ref><ref>Electromagnetic theory, Volume 1. By [[Oliver Heaviside]]. "The Electrician" printing and publishing company, limited, 1893.</ref><ref>Elements of the differential calculus. By [[Wesley Stoker Barker Woolhouse]]. Weale, 1854.</ref><ref>An Elementary Treatise on the Integral Calculus: Founded on the Method of Rates Or Fluxions. By [[William Woolsey Johnson]]. John Wiley & Sons, 1881.<br />See also: [[Method of Fluxions]].</ref><ref>Vector Calculus: With Applications to Physics. By [[James Byrnie Shaw]]. D. Van Nostrand, 1922. p205.<br />See also: [[Green's Theorem]].</ref><ref>A Treatise on the Integral Calculus, Volume 2. By [[Joseph Edwards (Mathematician)|Joseph Edwards]]. Chelsea Publishing Company, 1922.</ref> states that any sufficiently [[smooth function|smooth]], rapidly decaying [[vector field]] in three dimensions can be resolved into the sum of an [[irrotational vector field|irrotational]] ([[Curl (mathematics)|curl]]-free) vector field and a [[solenoidal]] ([[divergence]]-free) vector field; this is known as the '''Helmholtz decomposition''' or '''Helmholtz representation'''. It is named after [[Hermann von Helmholtz]].<ref>See:
In [[physics]] and [[mathematics]], in the area of [[vector calculus]], '''Helmholtz's theorem''',<ref name="bladel1958" /><ref name="koenigsberger1906"/> also known as the '''fundamental theorem of vector calculus''',<ref name="murray1898"/><ref name="gibbs1901"/><ref name="heaviside1893"/><ref name="woolhouse1854"/><ref name="johnson1881"/><ref name="shaw1922"/><ref name="edwards1922"/> states that any sufficiently [[smooth function|smooth]], rapidly decaying [[vector field]] in three dimensions can be resolved into the sum of an [[irrotational vector field|irrotational]] ([[Curl (mathematics)|curl]]-free) vector field and a [[solenoidal]] ([[divergence]]-free) vector field; this is known as the '''Helmholtz decomposition''' or '''Helmholtz representation'''. It is named after [[Hermann von Helmholtz]].
* H. Helmholtz (1858) [https://books.google.com/books?id=6gwPAAAAIAAJ&pg=PA25 "Über Integrale der hydrodynamischen Gleichungen, welcher der Wirbelbewegungen entsprechen"] (On integrals of the hydrodynamic equations which correspond to vortex motions), ''Journal für die reine und angewandte Mathematik'', '''55''': 25–55. On page 38, the components of the fluid's velocity (''u'',&nbsp;''v'',&nbsp;''w'') are expressed in terms of the gradient of a scalar potential P and the curl of a vector potential&nbsp;(''L'',&nbsp;''M'',&nbsp;''N'').
* However, Helmholtz was largely anticipated by George Stokes in his paper: G. G. Stokes (presented: 1849; published: 1856) [https://books.google.com/books?id=L_NYAAAAYAAJ&pg=PA1 "On the dynamical theory of diffraction,"] ''Transactions of the Cambridge Philosophical Society'', vol. 9, part I, pages 1–62; see pages 9–10.</ref>


== Definition ==
As an irrotational vector field has a [[scalar potential]] and a solenoidal vector field has a [[vector potential]], the Helmholtz decomposition states that a vector field (satisfying appropriate smoothness and decay conditions) can be decomposed as the sum of the form <math>-\nabla \phi + \nabla \times \mathbf{A}</math>, where <math>\phi</math> is a scalar field called "scalar potential", and {{math|'''A'''}} is a vector field, called a vector potential.


For a vector field <math>\mathbf{F} \in C^1(V, \mathbb{R}^n)</math> defined on a domain <math>V \subseteq \mathbb{R}^n</math>, a Helmholtz decomposition is a pair of vector fields <math>\mathbf{G} \in C^1(V, \mathbb{R}^n)</math> and <math>\mathbf{R} \in C^1(V, \mathbb{R}^n)</math> such that:
==Statement of the theorem==
<math display="block">
Let <math>\mathbf{F}</math> be a vector field on a bounded domain <math>V\subseteq\mathbb{R}^3</math>, which is twice continuously differentiable inside <math>V</math>, and let <math>S</math> be the surface that encloses the domain <math>V</math>. Then <math>\mathbf{F}</math> can be decomposed into a curl-free component and a divergence-free component:<ref>{{cite web |url=http://www.cems.uvm.edu/~oughstun/LectureNotes141/Topic_03_(Helmholtz'%20Theorem).pdf |title=Helmholtz' Theorem |publisher=University of Vermont| access-date=2011-03-11 | archive-url=https://web.archive.org/web/20120813005804/http://www.cems.uvm.edu/~oughstun/LectureNotes141/Topic_03_(Helmholtz'%20Theorem).pdf| archive-date=2012-08-13| url-status=dead}}</ref>
\begin{align}
\mathbf{F}(\mathbf{r}) &= \mathbf{G}(\mathbf{r}) + \mathbf{R}(\mathbf{r}), \\
\mathbf{G}(\mathbf{r}) &= - \nabla \Phi(\mathbf{r}), \\
\nabla \mathbf{R}(\mathbf{r}) &= 0.
\end{align}
</math>
Here, <math>\Phi \in C^2(\mathbb{R}^{n}, \mathbb{R})</math> is a [[scalar potential]], <math>\nabla \Phi</math> is its [[gradient]], and <math>\nabla \mathbf{R}</math> is the [[divergence]]. The irrotational vector field <math>\mathbf{G}</math> is called a ''gradient field'' and <math>\mathbf{R}</math> is called a ''[[solenoidal]] field'' or ''rotation field''. This decomposition does not exist for all vector fields and is not [[uniqueness|unique]].<ref name="glotzl2023"/>

== History ==

The Helmholtz decomposition in three dimensions was first described in 1849<ref name="stokes1849" /> by [[George Gabriel Stokes]] for a theory of [[diffraction]]. [[Hermann von Helmholtz]] published his paper on some [[hydrodynamics|hydrodynamic]] basic equations in 1858,<ref name="helmholtz1858" /><ref name="kustepeli2016" /> which was part of his research on the [[Helmholtz's theorems]] describing the motion of fluid in the vicinity of vortex lines.<ref name="kustepeli2016"/> Their derivation required the vector fields to decay sufficiently fast at infinity. Later, this condition could be relaxed, and the Helmholtz decomposition could be extended to higher dimensions.<ref name="glotzl2023"/><ref name="trancong1993" /><ref name="petrascheck2017"/>

The decomposition has become an important tool for many problems in [[theoretical physics|theoretical physics]],<ref name="kustepeli2016" /><ref name="sprossig2009" /> but has also found applications in [[animation]], [[computer vision]] as well as [[robotics]].<ref name="bhatia2013" /> In this context, the Helmholtz decomposition has been extended to higher-dimensional spaces and applied to [[Riemannian manifold|riemannian manifolds]] as the Helmholtz-Hodge decomposition using [[differential geometry]] and [[tensor calculus]].<ref name="glotzl2023" /><ref name="kustepeli2016"/><ref name="sprossig2009"/><ref name="bhatia2013" />

== Three-dimensional space ==

Many physics textbooks restrict the Helmholtz decomposition to the three-dimensional space and limit its application to vector fields that decay sufficiently fast at infinity or to [[bump function]] that are defined on a [[bounded domain]]. Then, a [[vector potential]] <math>A</math> can be defined, such that the rotation field is given by <math>\mathbf{R} = \nabla \times \mathbf{A}</math>, using the [[Curl (mathematics)|Curl]] of a vector field.<ref name="petrascheck2015" />

Let <math>\mathbf{F}</math> be a vector field on a bounded domain <math>V\subseteq\mathbb{R}^3</math>, which is twice continuously differentiable inside <math>V</math>, and let <math>S</math> be the surface that encloses the domain <math>V</math>. Then <math>\mathbf{F}</math> can be decomposed into a curl-free component and a divergence-free component as follows:<ref name="vermont" />


<math display="block">\mathbf{F}=-\nabla \Phi+\nabla\times\mathbf{A},</math>
<math display="block">\mathbf{F}=-\nabla \Phi+\nabla\times\mathbf{A},</math>
Line 20: Line 37:
and <math>\nabla'</math> is the nabla operator with respect to <math>\mathbf{r'}</math>, not <math> \mathbf{r} </math>.
and <math>\nabla'</math> is the nabla operator with respect to <math>\mathbf{r'}</math>, not <math> \mathbf{r} </math>.


If <math>V = \R^3</math> and is therefore unbounded, and <math>\mathbf{F}</math> vanishes at least as fast as <math>1/r</math> as <math>r \to \infty</math>, then one has<ref name="griffiths">[[David J. Griffiths]], ''Introduction to Electrodynamics'', Prentice-Hall, 1999, p. 556.</ref>
If <math>V = \R^3</math> and is therefore unbounded, and <math>\mathbf{F}</math> vanishes at least as fast as <math>1/r</math> as <math>r \to \infty</math>, then one has<ref name="griffiths1999"/>


<math display="block">\begin{align}
<math display="block">\begin{align}
Line 28: Line 45:
This holds in particular if <math>\mathbf F</math> is twice continuously differentiable in <math>\mathbb R^3</math> and of bounded support.
This holds in particular if <math>\mathbf F</math> is twice continuously differentiable in <math>\mathbb R^3</math> and of bounded support.


==Derivation==
=== Derivation ===

{{math proof| proof =
Suppose we have a vector function <math>\mathbf{F}(\mathbf{r})</math> of which we know the curl, <math>\nabla\times\mathbf{F}</math>, and the divergence, <math>\nabla\cdot\mathbf{F}</math>, in the domain and the fields on the boundary. Writing the function using [[delta function]] in the form
Suppose we have a vector function <math>\mathbf{F}(\mathbf{r})</math> of which we know the curl, <math>\nabla\times\mathbf{F}</math>, and the divergence, <math>\nabla\cdot\mathbf{F}</math>, in the domain and the fields on the boundary. Writing the function using [[delta function]] in the form
<math display="block">\delta^3(\mathbf{r}-\mathbf{r}')=-\frac 1 {4\pi} \nabla^2 \frac{1}{|\mathbf{r}-\mathbf{r}'|}\, ,</math>
<math display="block">\delta^3(\mathbf{r}-\mathbf{r}')=-\frac 1 {4\pi} \nabla^2 \frac{1}{|\mathbf{r}-\mathbf{r}'|}\, ,</math>
Line 151: Line 170:
we finally obtain
we finally obtain
<math display="block">\mathbf{F}=-\nabla\Phi+\nabla\times\mathbf{A}.</math>
<math display="block">\mathbf{F}=-\nabla\Phi+\nabla\times\mathbf{A}.</math>
}}


=== Fields with prescribed divergence and curl ===
===Generalization to higher dimensions===
The term "Helmholtz theorem" can also refer to the following. Let {{math|'''C'''}} be a [[solenoidal vector field]] and ''d'' a scalar field on {{math|'''R'''<sup>3</sup>}} which are sufficiently smooth and which vanish faster than {{math|1/''r''<sup>2</sup>}} at infinity. Then there exists a vector field {{math|'''F'''}} such that

<math display="block">\nabla \cdot \mathbf{F} = d \quad \text{ and } \quad \nabla \times \mathbf{F} = \mathbf{C};</math>

if additionally the vector field {{math|'''F'''}} vanishes as {{math|''r'' → ∞}}, then {{math|'''F'''}} is unique.<ref name="griffiths1999"/>

In other words, a vector field can be constructed with both a specified divergence and a specified curl, and if it also vanishes at infinity, it is uniquely specified by its divergence and curl. This theorem is of great importance in [[electrostatics]], since [[Maxwell's equations]] for the electric and magnetic fields in the static case are of exactly this type.<ref name="griffiths1999"/> The proof is by a construction generalizing the one given above: we set

<math display="block">\mathbf{F} = - \nabla(\mathcal{G} (d)) + \nabla \times (\mathcal{G}(\mathbf{C})),</math>

where <math>\mathcal{G}</math> represents the [[Newtonian potential]] operator. (When acting on a vector field, such as {{math|∇ × '''F'''}}, it is defined to act on each component.)

=== Weak formulation ===
The Helmholtz decomposition can be generalized by reducing the regularity assumptions (the need for the existence of strong derivatives). Suppose {{math|Ω}} is a bounded, simply-connected, [[Lipschitz domain]]. Every [[square-integrable]] vector field {{math|'''u''' ∈ (''L''<sup>2</sup>(Ω))<sup>3</sup>}} has an [[orthogonality|orthogonal]] decomposition:<ref name="amrouche1998" /><ref name="dautray1990" /><ref name="girault1986" />

<math display="block">\mathbf{u}=\nabla\varphi+\nabla \times \mathbf{A}</math>

where {{mvar|φ}} is in the [[Sobolev space]] {{math|''H''<sup>1</sup>(Ω)}} of square-integrable functions on {{math|Ω}} whose partial derivatives defined in the [[distribution (mathematics)|distribution]] sense are square integrable, and {{math|'''A''' ∈ ''H''(curl, Ω)}}, the Sobolev space of vector fields consisting of square integrable vector fields with square integrable curl.

For a slightly smoother vector field {{math|'''u''' ∈ ''H''(curl, Ω)}}, a similar decomposition holds:

<math display="block">\mathbf{u}=\nabla\varphi+\mathbf{v}</math>

where {{math|''φ'' ∈ ''H''<sup>1</sup>(Ω), '''v''' ∈ (''H''<sup>1</sup>(Ω))<sup>''d''</sup>}}.

=== Derivation from the Fourier transform ===
Note that in the theorem stated here, we have imposed the condition that if <math>\mathbf{F}</math> is not defined on a bounded domain, then <math>\mathbf{F}</math> shall decay faster than <math>1/r</math>. Thus, the [[Fourier transform]] of <math>\mathbf{F}</math>, denoted as <math>\mathbf{G}</math>, is guaranteed to exist. We apply the convention
<math display="block">\mathbf{F}(\mathbf{r}) = \iiint \mathbf{G}(\mathbf{k}) e^{i\mathbf{k} \cdot \mathbf{r}} dV_k </math>

The Fourier transform of a scalar field is a scalar field, and the Fourier transform of a vector field is a vector field of same dimension.

Now consider the following scalar and vector fields:
<math display="block">\begin{align}
G_\Phi(\mathbf{k}) &= i \frac{\mathbf{k} \cdot \mathbf{G}(\mathbf{k})}{\|\mathbf{k}\|^2} \\
\mathbf{G}_\mathbf{A}(\mathbf{k}) &= i \frac{\mathbf{k} \times \mathbf{G}(\mathbf{k})}{\|\mathbf{k}\|^2} \\ [8pt]
\Phi(\mathbf{r}) &= \iiint G_\Phi(\mathbf{k}) e^{i \mathbf{k} \cdot \mathbf{r}} dV_k \\
\mathbf{A}(\mathbf{r}) &= \iiint \mathbf{G}_\mathbf{A}(\mathbf{k}) e^{i \mathbf{k} \cdot \mathbf{r}} dV_k
\end{align} </math>

Hence

<math display="block">\begin{align}
\mathbf{G}(\mathbf{k}) &= - i \mathbf{k} G_\Phi(\mathbf{k}) + i \mathbf{k} \times \mathbf{G}_\mathbf{A}(\mathbf{k}) \\ [6pt]
\mathbf{F}(\mathbf{r}) &= -\iiint i \mathbf{k} G_\Phi(\mathbf{k}) e^{i \mathbf{k} \cdot \mathbf{r}} dV_k + \iiint i \mathbf{k} \times \mathbf{G}_\mathbf{A}(\mathbf{k}) e^{i \mathbf{k} \cdot \mathbf{r}} dV_k \\
&= - \nabla \Phi(\mathbf{r}) + \nabla \times \mathbf{A}(\mathbf{r})
\end{align}</math>

=== Longitudinal and transverse fields ===
A terminology often used in physics refers to the curl-free component of a vector field as the '''longitudinal component''' and the divergence-free component as the '''transverse component'''.<ref name="stewart2011"/> This terminology comes from the following construction: Compute the three-dimensional [[Fourier transform]] <math>\hat\mathbf{F}</math> of the vector field <math>\mathbf{F}</math>. Then decompose this field, at each point '''k''', into two components, one of which points longitudinally, i.e. parallel to '''k''', the other of which points in the transverse direction, i.e. perpendicular to '''k'''. So far, we have

<math display="block">\hat\mathbf{F} (\mathbf{k}) = \hat\mathbf{F}_t (\mathbf{k}) + \hat\mathbf{F}_l (\mathbf{k})</math>
<math display="block">\mathbf{k} \cdot \hat\mathbf{F}_t(\mathbf{k}) = 0.</math>
<math display="block">\mathbf{k} \times \hat\mathbf{F}_l(\mathbf{k}) = \mathbf{0}.</math>

Now we apply an inverse Fourier transform to each of these components. Using properties of Fourier transforms, we derive:

<math display="block">\mathbf{F}(\mathbf{r}) = \mathbf{F}_t(\mathbf{r})+\mathbf{F}_l(\mathbf{r})</math>
<math display="block">\nabla \cdot \mathbf{F}_t (\mathbf{r}) = 0</math>
<math display="block">\nabla \times \mathbf{F}_l (\mathbf{r}) = \mathbf{0}</math>

Since <math>\nabla\times(\nabla\Phi)=0</math> and <math>\nabla\cdot(\nabla\times\mathbf{A})=0</math>,

we can get

<math display="block">\mathbf{F}_t=\nabla\times\mathbf{A}=\frac{1}{4\pi}\nabla\times\int_V\frac{\nabla'\times\mathbf{F}}{\left|\mathbf{r}-\mathbf{r}'\right|}\mathrm{d}V'</math>
<math display="block">\mathbf{F}_l=-\nabla\Phi=-\frac{1}{4\pi}\nabla\int_V\frac{\nabla'\cdot\mathbf{F}}{\left|\mathbf{r}-\mathbf{r}'\right|}\mathrm{d}V'</math>

so this is indeed the Helmholtz decomposition.<ref name="littlejohn"/>

== Generalization to higher dimensions ==

=== Matrix approach ===

The generalization to <math>d</math> dimensions cannot be done with a vector potential, since the rotation operator and the [[cross product]] are defined only in three dimensions.

Let <math>\mathbf{F}</math> be a vector field on a bounded domain <math>V\subseteq\mathbb{R}^d</math> which decays faster then <math>|\mathbf{r}|^{-\delta}</math> for <math>|\mathbf{r}| \to \infty</math> and <math>\delta > 2</math>.

The scalar potential is defined similar to the three dimensional case as:
<math display="block">\Phi(\mathbf{r}) = - \int_{\mathbb{R}^d} \operatorname{div}(\mathbf{F}(\mathbf{r}')) K(\mathbf{r}, \mathbf{r}') \mathrm{d}V' = - \int_{\mathbb{R}^d} \sum_i \frac{\partial F_i}{\partial r_i}(\mathbf{r}') K(\mathbf{r}, \mathbf{r}') \mathrm{d}V',</math>
where as the integration kernel <math>K(\mathbf{r}, \mathbf{r}')</math> is again the [[fundamental solution]] of [[Laplace's equation]], but in d-dimensional space:
<math display="block">K(\mathbf{r}, \mathbf{r}') = \begin{cases} \frac{1}{2\pi} \log{ | \mathbf{r}-\mathbf{r}' | } & d=2, \\ \frac{1}{d(2-d)V_d} | \mathbf{r}-\mathbf{r}' | ^{2-d} & \text{otherwise}, \end{cases}</math>
with <math>V_d = \pi^\frac{d}{2} / \Gamma\big(\tfrac{d}{2}+1\big)</math> the volume of the d-dimensional [[unit ball]]s and <math>\Gamma(\vec{x})</math> the [[gamma function]].

For <math>d = 3</math>, <math>V_d</math> is just equal to <math>\frac{4 \pi}{3}</math>, yielding the same prefactor as above.
The rotational potential is an [[antisymmetric matrix]] with the elements:
<math display="block">A_{ij}(\mathbf{r}) = \int_{\mathbb{R}^d} \left( \frac{\partial F_i}{\partial x_j}(\mathbf{r}') - \frac{\partial F_j}{\partial x_i}(\mathbf{r}') \right) K(\mathbf{r}, \mathbf{r}') \mathrm{d}V'. </math>
Above the diagonal are <math>\textstyle\binom{d}{2}</math> entries which occur again mirrored at the diagonal, but with a negative sign.
In the three-dimensional case, the matrix elements just correspond to the components of the vector potential <math>\mathbf{A} = [A_1, A_2, A_3] = [A_{23}, A_{31}, A_{12}]</math>.
However, such a matrix potential can be written as a vector only in the three-dimensional case, because <math>\textstyle\binom{d}{2} = d</math> is valid only for <math>d = 3</math>.

As in the three-dimensional case, the gradient field is defined as
<math display="block">
\mathbf{G}(\mathbf{r}) = - \nabla \Phi(\mathbf{r}).
</math>
The rotational field, on the other hand, is defined in the general case as the row divergence of the matrix:
<math display="block">\mathbf{R}(\mathbf{r}) = \left[ \sum\nolimits_k \partial_{r_k} A_{ik}(\mathbf{r}); {1 \leq i \leq d} \right].</math>
In three-dimensional space, this is equivalent to the rotation of the vector potential.<ref name="glotzl2023" /><ref name="glotzl2020" />

=== Tensor approach ===


In a <math>d</math>-dimensional vector space with <math>d\neq 3</math>, <math display="inline">-\frac{1}{4\pi\left|\mathbf{r}-\mathbf{r}'\right|}</math> should be replaced by the appropriate [[Green's function#Green's functions for the Laplacian|Green's function for the Laplacian]], defined by
In a <math>d</math>-dimensional vector space with <math>d\neq 3</math>, <math display="inline">-\frac{1}{4\pi\left|\mathbf{r}-\mathbf{r}'\right|}</math> can be replaced by the appropriate [[Green's function#Green's functions for the Laplacian|Green's function for the Laplacian]], defined by
<math display="block">
<math display="block">
\nabla^2 G(\mathbf{r},\mathbf{r}') = \frac{\partial}{\partial r_\mu}\frac{\partial}{\partial r_\mu}G(\mathbf{r},\mathbf{r}') = \delta^d(\mathbf{r}-\mathbf{r}')
\nabla^2 G(\mathbf{r},\mathbf{r}') = \frac{\partial}{\partial r_\mu}\frac{\partial}{\partial r_\mu}G(\mathbf{r},\mathbf{r}') = \delta^d(\mathbf{r}-\mathbf{r}')
Line 190: Line 309:
For a further generalization to manifolds, see the discussion of [[Hodge decomposition]] [[Helmholtz decomposition#Differential forms|below]].
For a further generalization to manifolds, see the discussion of [[Hodge decomposition]] [[Helmholtz decomposition#Differential forms|below]].


== Differential forms ==
===Another derivation from the Fourier transform===
The [[Hodge decomposition#Hodge decomposition|Hodge decomposition]] is closely related to the Helmholtz decomposition,<ref name="warner1983"/> generalizing from vector fields on '''R'''<sup>3</sup> to [[differential forms]] on a [[Riemannian manifold]] ''M''. Most formulations of the Hodge decomposition require ''M'' to be [[compact space|compact]].<ref name="cantarella2002" /> Since this is not true of '''R'''<sup>3</sup>, the Hodge decomposition theorem is not strictly a generalization of the Helmholtz theorem. However, the compactness restriction in the usual formulation of the Hodge decomposition can be replaced by suitable decay assumptions at infinity on the differential forms involved, giving a proper generalization of the Helmholtz theorem.
Note that in the theorem stated here, we have imposed the condition that if <math>\mathbf{F}</math> is not defined on a bounded domain, then <math>\mathbf{F}</math> shall decay faster than <math>1/r</math>. Thus, the Fourier Transform of <math>\mathbf{F}</math>, denoted as <math>\mathbf{G}</math>, is guaranteed to exist. We apply the convention
<math display="block">\mathbf{F}(\mathbf{r}) = \iiint \mathbf{G}(\mathbf{k}) e^{i\mathbf{k} \cdot \mathbf{r}} dV_k </math>


== Extensions to fields not decaying at infinity ==
The Fourier transform of a scalar field is a scalar field, and the Fourier transform of a vector field is a vector field of same dimension.


Most textbooks only deal with vector fields decaying faster then <math>|\mathbf{r}|^{-\delta}</math> with <math>\delta > 1</math> at infinity.<ref name="petrascheck2015" /><ref name="petrascheck2017"/> However, [[Otto Blumenthal]] showed in 1905 that an adapted integration kernel can be used to integrate fields decaying faster than <math>|\vec{x}|^{-\delta}</math> with <math>\delta > 0</math>, which is substantially less strict.
Now consider the following scalar and vector fields:
To achieve this, the kernel <math>K(\mathbf{r}, \mathbf{r}')</math> in the convolution integrals has to be replaced by <math>K'(\mathbf{r}, \mathbf{r}') = K(\mathbf{r}, \mathbf{r}') - K(0, \mathbf{r}')</math>.<ref name="blumenthal1905" />
<math display="block">\begin{align}
With even more complex integration kernels, solutions can be found even for divergent functions that need not grow faster than polynomial.<ref name="trancong1993" /><ref name="petrascheck2017"/><ref name="glotzl2020" /><ref name="gurtin1962"/>
G_\Phi(\mathbf{k}) &= i \frac{\mathbf{k} \cdot \mathbf{G}(\mathbf{k})}{\|\mathbf{k}\|^2} \\
\mathbf{G}_\mathbf{A}(\mathbf{k}) &= i \frac{\mathbf{k} \times \mathbf{G}(\mathbf{k})}{\|\mathbf{k}\|^2} \\ [8pt]
\Phi(\mathbf{r}) &= \iiint G_\Phi(\mathbf{k}) e^{i \mathbf{k} \cdot \mathbf{r}} dV_k \\
\mathbf{A}(\mathbf{r}) &= \iiint \mathbf{G}_\mathbf{A}(\mathbf{k}) e^{i \mathbf{k} \cdot \mathbf{r}} dV_k
\end{align} </math>


For all [[Analytic function|analytic]] vector fields that need not go to zero even at infinity, methods based on [[partial integration|partial integration]] and the [[Cauchy formula for repeated integration]]<ref name="cauchy1823" /> can be used to compute closed-form solutions of the rotation and scalar potentials, as in the case of [[multivariate polynomial|multivariate polynomial]], [[sine]], [[cosine]], and [[exponential function]]s.<ref name="glotzl2023" />
Hence

== Uniqueness of the solution ==

In general, the Helmholtz decomposition is not uniquely defined.
A [[harmonic function]] <math>H(\mathbf{r})</math> is a function that satisfies <math>\Delta H(\vec{x}) = 0</math>.
By adding <math>H(\mathbf{r})</math> to the scalar potential <math>\Phi(\mathbf{r})</math>, a different Helmholtz decomposition can be obtained:


<math display="block">\begin{align}
<math display="block">\begin{align}
\mathbf{G}(\mathbf{k}) &= - i \mathbf{k} G_\Phi(\mathbf{k}) + i \mathbf{k} \times \mathbf{G}_\mathbf{A}(\mathbf{k}) \\ [6pt]
\mathbf{G}'(\mathbf{r}) &= \nabla (\Phi(\mathbf{r}) + H(\mathbf{r})) = \mathbf{G}(\mathbf{r}) + \nabla H(\mathbf{r}),\\
\mathbf{F}(\mathbf{r}) &= -\iiint i \mathbf{k} G_\Phi(\mathbf{k}) e^{i \mathbf{k} \cdot \mathbf{r}} dV_k + \iiint i \mathbf{k} \times \mathbf{G}_\mathbf{A}(\mathbf{k}) e^{i \mathbf{k} \cdot \mathbf{r}} dV_k \\
\mathbf{R}'(\mathbf{r}) &= \mathbf{R}(\mathbf{r}) - \nabla H(\mathbf{r}).
&= - \nabla \Phi(\mathbf{r}) + \nabla \times \mathbf{A}(\mathbf{r})
\end{align}</math>
\end{align}</math>


For vector fields <math>\mathbf{F}</math>, decaying at infinity, it is a plausible choice that scalar and rotation potentials also decay at infinity.
==Fields with prescribed divergence and curl==
Because <math>H(\vec{x}) = 0</math> is the only harmonic function with this property, which follows from [[Liouville's theorem (complex analysis)|Liouville's theorem]], this guarantees the uniqueness of the gradient and rotation fields.<ref name="axler1992" />
The term "Helmholtz theorem" can also refer to the following. Let {{math|'''C'''}} be a [[solenoidal vector field]] and ''d'' a scalar field on {{math|'''R'''<sup>3</sup>}} which are sufficiently smooth and which vanish faster than {{math|1/''r''<sup>2</sup>}} at infinity. Then there exists a vector field {{math|'''F'''}} such that


This uniqueness does not apply to the potentials: In the three-dimensional case, the scalar and vector potential jointly have four components, whereas the vector field has only three. The vector field is invariant to gauge transformations and the choice of appropriate potentials known as [[gauge fixing]] is the subject of [[gauge theory]]. Important examples from physics are the [[Lorenz gauge condition]] and the [[Coulomb gauge]]. An alternative is to use the [[poloidal–toroidal decomposition]].
<math display="block">\nabla \cdot \mathbf{F} = d \quad \text{ and } \quad \nabla \times \mathbf{F} = \mathbf{C};</math>


== Applications ==
if additionally the vector field {{math|'''F'''}} vanishes as {{math|''r'' → ∞}}, then {{math|'''F'''}} is unique.<ref name="griffiths"/>
=== Electrodynamics ===


The Helmholtz theorem is of particular interest in [[electrodynamics]], since it can be used to write [[Maxwell's equations]] in the potential image and solve them more easily. The Helmholtz decomposition can be used to prove that, given [[electric current density|electric current density]] and [[charge density]], the [[electric field]] and the [[magnetic flux density]] can be determined. They are unique if the densities vanish at infinity and one assumes the same for the potentials.<ref name="petrascheck2015" />
In other words, a vector field can be constructed with both a specified divergence and a specified curl, and if it also vanishes at infinity, it is uniquely specified by its divergence and curl. This theorem is of great importance in [[electrostatics]], since [[Maxwell's equations]] for the electric and magnetic fields in the static case are of exactly this type.<ref name="griffiths"/> The proof is by a construction generalizing the one given above: we set


=== Fluid dynamics ===
<math display="block">\mathbf{F} = - \nabla(\mathcal{G} (d)) + \nabla \times (\mathcal{G}(\mathbf{C})),</math>


In [[fluid dynamics]], the Helmholtz projection plays an important role, especially for the solvability theory of the [[Navier-Stokes equations]]. If the Helmholtz projection is applied to the linearized incompressible Navier-Stokes equations, the [[Stokes equation]] is obtained. This depends only on the velocity of the particles in the flow, but no longer on the static pressure, allowing the equation to be reduced to one unknown. However, both equations, the Stokes and linearized equations, are equivalent. The operator <math>P\Delta</math> is called the [[Stokes operator]].<ref name="chorin1990" />
where <math>\mathcal{G}</math> represents the [[Newtonian potential]] operator. (When acting on a vector field, such as {{math|∇ × '''F'''}}, it is defined to act on each component.)


=== Dynamical systems theory ===
==Solution space==
For two Helmholtz decompositions <math>(\Phi_1, {\mathbf A_1})</math> <math>(\Phi_2, {\mathbf A_2})</math> of <math>\mathbf F</math>, there holds
:<math>\Phi_1-\Phi_2 = \lambda,\quad {\mathbf{A}_1 - \mathbf{A}_2} ={\mathbf A}_\lambda + \nabla \varphi,</math>
:where
:* <math> \lambda</math> is an [[harmonic function|harmonic scalar field]],
:* <math> {\mathbf A}_\lambda </math> is a vector field determined by <math>\lambda</math>,
:* <math> \varphi </math> is any scalar field.


In the theory of [[dynamical system]]s, Helmholtz decomposition can be used to determine "quasipotentials" as well as to compute [[Lyapunov function]]s in some cases.<ref name="suda2019" /><ref name="suda2020" /><ref name="zhou2012" />
Proof:
Setting <math>\lambda = \Phi_2 - \Phi_1</math> and <math>{\mathbf B = A_2 - A_1}</math>, one has, according to the
definition of the Helmholtz decomposition,
:<math> -\nabla \lambda + \nabla \times \mathbf B = 0 </math>.
Taking the divergence of each member of this equation yields
<math>\nabla^2 \lambda = 0</math>, hence <math>\lambda</math> is harmonic.


For some dynamical systems such as the [[Lorenz system]] ([[Edward N. Lorenz]], 1963<ref name="lorenz1963" />), a simplified model for [[atmosphere|atmospheric]] [[convection]], a [[closed-form expression]] of the Helmholtz decomposition can be obtained:
Conversely, given any harmonic function <math>\lambda</math>,
<math display="block">\dot \vec{x} = \vec{f}(\vec{x}) = \big[a (x_2-x_1), x_1 (b-x_3)-x_2, x_1 x_2-c x_3 \big].</math>
<math>\nabla \lambda </math> is solenoidal since
The Helmholtz decomposition of <math>\vec{f}(\vec{x})</math>, with the scalar potential <math>\Phi(\vec{x}) = - \tfrac{a}{2} x_1^2 - \tfrac{1}{2} x_2^2 - \tfrac{c}{2} x_3^2</math> is given as:
:<math>\nabla\cdot (\nabla \lambda) = \nabla^2 \lambda = 0.</math>
Thus, according to the above section, there exists a vector field <math>{\mathbf A}_\lambda</math> such that
<math>\nabla \lambda = \nabla\times {\mathbf A}_\lambda</math>.
If <math>{\mathbf A'}_\lambda</math> is another such vector field,
then <math>\mathbf C = {\mathbf A}_\lambda - {\mathbf A'}_\lambda</math>
fulfills <math>\nabla \times {\mathbf C} = 0</math>, hence <math>C = \nabla \varphi</math>
for some scalar field <math>\varphi</math> (and conversely).


<math display="block">\vec{g}(\vec{x}) = \big[-a x_1, -x_2, -c x_3 \big],</math>
==Differential forms==
<math display="block">\vec{r}(\vec{x}) = \big[+ a x_2, b x_1 - x_1 x_3, x_1 x_2 \big].</math>
The [[Hodge decomposition#Hodge decomposition|Hodge decomposition]] is closely related to the Helmholtz decomposition, generalizing from vector fields on '''R'''<sup>3</sup> to [[differential forms]] on a [[Riemannian manifold]] ''M''. Most formulations of the Hodge decomposition require ''M'' to be [[compact space|compact]].<ref>{{cite journal| jstor=2695643| title=Vector Calculus and the Topology of Domains in 3-Space| first1=Jason |last1=Cantarella |first2=Dennis |last2=DeTurck | first3=Herman|last3=Gluck|journal=The American Mathematical Monthly|volume=109|issue=5|year=2002 |pages=409–442 | doi=10.2307/2695643 }}</ref> Since this is not true of '''R'''<sup>3</sup>, the Hodge decomposition theorem is not strictly a generalization of the Helmholtz theorem. However, the compactness restriction in the usual formulation of the Hodge decomposition can be replaced by suitable decay assumptions at infinity on the differential forms involved, giving a proper generalization of the Helmholtz theorem.


The quadratic scalar potential provides motion in the direction of the coordinate origin, which is responsible for the stable [[fixed point (mathematics)|fix point]] for some parameter range. For other parameters, the rotation field ensures that a [[strange attractor]] is created, causing the model to exhibit a [[butterfly effect]].<ref name="glotzl2023" /><ref name="peitgen1992" />
==Weak formulation==
The Helmholtz decomposition can also be generalized by reducing the regularity assumptions (the need for the existence of strong derivatives). Suppose {{math|Ω}} is a bounded, simply-connected, [[Lipschitz domain]]. Every [[square-integrable]] vector field {{math|'''u''' ∈ (''L''<sup>2</sup>(Ω))<sup>3</sup>}} has an [[orthogonality|orthogonal]] decomposition:


=== Computer animation and robotics ===
<math display="block">\mathbf{u}=\nabla\varphi+\nabla \times \mathbf{A}</math>


The Helmholtz decomposition is also used in the field of computer engineering. This includes robotics, image reconstruction but also computer animation, where the decomposition is used for realistic visualization of fluids or vector fields.<ref name="bhatia2013" /><ref name="bhatia2014" />
where {{mvar|φ}} is in the [[Sobolev space]] {{math|''H''<sup>1</sup>(Ω)}} of square-integrable functions on {{math|Ω}} whose partial derivatives defined in the [[distribution (mathematics)|distribution]] sense are square integrable, and {{math|'''A''' ∈ ''H''(curl, Ω)}}, the Sobolev space of vector fields consisting of square integrable vector fields with square integrable curl.


== See also ==
For a slightly smoother vector field {{math|'''u''' ∈ ''H''(curl, Ω)}}, a similar decomposition holds:
* [[Clebsch representation]] for a related decomposition of vector fields
* [[Darwin Lagrangian]] for an application
* [[Poloidal–toroidal decomposition]] for a further decomposition of the divergence-free component <math> \nabla \times \mathbf{A} </math>.
* [[Scalar–vector–tensor decomposition]]
* [[Hodge theory]] generalizing Helmholtz decomposition
* [[Polar factorization theorem]]
* ''Helmholtz–Leray decomposition'' used for defining the [[Leray projection]]


== Notes ==
<math display="block">\mathbf{u}=\nabla\varphi+\mathbf{v}</math>
{{Reflist|30em|refs=


<ref name="amrouche1998">Cherif Amrouche, [[Christine Bernardi]], [[Monique Dauge]], [[Vivette Girault]]: ''Vector potentials in three dimensional non-smooth domains''. In: ''[[Mathematical Methods in the Applied Sciences]]'' 21(9), 1998, pp. 823–864, {{DOI|10.1002/(sici)1099-1476(199806)21:9<823::aid-mma976>3.0.co;2-b}}, {{bibcode|1998MMAS...21..823A }}.</ref>
where {{math|''φ'' ∈ ''H''<sup>1</sup>(Ω), '''v''' ∈ (''H''<sup>1</sup>(Ω))<sup>''d''</sup>}}.


<ref name="axler1992">Sheldon Axler, Paul Bourdon, Wade Ramey: ''Bounded Harmonic Functions''. In: ''Harmonic Function Theory'' (= Graduate Texts in Mathematics 137). Springer, New York 1992, pp. 31–44, {{DOI|10.1007/0-387-21527-1_2}}.</ref>
==Longitudinal and transverse fields==
A terminology often used in physics refers to the curl-free component of a vector field as the '''longitudinal component''' and the divergence-free component as the '''transverse component'''.<ref>[https://arxiv.org/abs/0801.0335 Stewart, A. M.; Longitudinal and transverse components of a vector field, Sri Lankan Journal of Physics 12, 33–42 (2011)]</ref> This terminology comes from the following construction: Compute the three-dimensional [[Fourier transform]] <math>\hat\mathbf{F}</math> of the vector field <math>\mathbf{F}</math>. Then decompose this field, at each point '''k''', into two components, one of which points longitudinally, i.e. parallel to '''k''', the other of which points in the transverse direction, i.e. perpendicular to '''k'''. So far, we have


<ref name="bhatia2013">Harsh Bhatia, Gregory Norgard, Valerio Pascucci, Peer-Timo Bremer: ''The Helmholtz-Hodge Decomposition – A Survey''. In: ''[[Institute of Electrical and Electronics Engineers|IEEE]] Transactions on Visualization and Computer Graphics'' 19.8, 2013, pp. 1386–1404, {{DOI|10.1109/tvcg.2012.316}}.</ref>
<math display="block">\hat\mathbf{F} (\mathbf{k}) = \hat\mathbf{F}_t (\mathbf{k}) + \hat\mathbf{F}_l (\mathbf{k})</math>
<math display="block">\mathbf{k} \cdot \hat\mathbf{F}_t(\mathbf{k}) = 0.</math>
<math display="block">\mathbf{k} \times \hat\mathbf{F}_l(\mathbf{k}) = \mathbf{0}.</math>


<ref name="bhatia2014">Hersh Bhatia, Valerio Pascucci, Peer-Timo Bremer: ''The Natural Helmholtz-Hodge Decomposition for Open-Boundary Flow Analysis''. In: ''[[Institute of Electrical and Electronics Engineers|IEEE]] Transactions on Visualization and Computer Graphics'' 20.11, Nov. 2014, pp. 1566–1578, Nov. 2014, {{DOI|10.1109/TVCG.2014.2312012}}.</ref>
Now we apply an inverse Fourier transform to each of these components. Using properties of Fourier transforms, we derive:


<ref name="bladel1958">Jean Bladel: ''On Helmholtz's Theorem in Finite Regions''. Midwestern Universities Research Association, 1958.</ref>
<math display="block">\mathbf{F}(\mathbf{r}) = \mathbf{F}_t(\mathbf{r})+\mathbf{F}_l(\mathbf{r})</math>
<math display="block">\nabla \cdot \mathbf{F}_t (\mathbf{r}) = 0</math>
<math display="block">\nabla \times \mathbf{F}_l (\mathbf{r}) = \mathbf{0}</math>


<ref name="blumenthal1905">[[Otto Blumenthal]]: ''Über die Zerlegung unendlicher Vektorfelder''. In: ''[[Mathematische Annalen]]'' 61.2, 1905, pp. 235–250, {{DOI|10.1007/BF01457564}}.</ref>
Since <math>\nabla\times(\nabla\Phi)=0</math> and <math>\nabla\cdot(\nabla\times\mathbf{A})=0</math>,


<ref name="cantarella2002">{{cite journal| jstor=2695643| title=Vector Calculus and the Topology of Domains in 3-Space| first1=Jason |last1=Cantarella |first2=Dennis |last2=DeTurck | first3=Herman|last3=Gluck|journal=The American Mathematical Monthly|volume=109|issue=5|year=2002 |pages=409–442 | doi=10.2307/2695643 }}</ref>
we can get


<ref name="cauchy1823">[[Augustin-Louis Cauchy]]: ''Trente-Cinquième Leçon''. In: ''Résumé des leçons données à l’École royale polytechnique sur le calcul infinitésimal''. Imprimerie Royale, Paris 1823, pp. 133–140 ([https://gallica.bnf.fr/ark:/12148/bpt6k62404287/f146.item gallica.bnf.fr]).</ref>
<math display="block">\mathbf{F}_t=\nabla\times\mathbf{A}=\frac{1}{4\pi}\nabla\times\int_V\frac{\nabla'\times\mathbf{F}}{\left|\mathbf{r}-\mathbf{r}'\right|}\mathrm{d}V'</math>
<math display="block">\mathbf{F}_l=-\nabla\Phi=-\frac{1}{4\pi}\nabla\int_V\frac{\nabla'\cdot\mathbf{F}}{\left|\mathbf{r}-\mathbf{r}'\right|}\mathrm{d}V'</math>


<ref name="chorin1990">Alexandre J. Chorin, Jerrold E. Marsden: ''A Mathematical Introduction to Fluid Mechanics'' (= Texts in Applied Mathematics 4). Springer US, New York 1990, {{DOI|10.1007/978-1-4684-0364-0}}.</ref>
so this is indeed the Helmholtz decomposition.<ref>[http://bohr.physics.berkeley.edu/classes/221/1112/notes/hamclassemf.pdf Online lecture notes by Robert Littlejohn]</ref>


<ref name="dautray1990">R. Dautray and J.-L. Lions. ''Spectral Theory and Applications,'' volume 3 of Mathematical Analysis and Numerical Methods for Science and Technology. Springer-Verlag, 1990.</ref>
==See also==
* [[Clebsch representation]] for a related decomposition of vector fields
* [[Darwin Lagrangian]] for an application
* [[Poloidal–toroidal decomposition]] for a further decomposition of the divergence-free component <math> \nabla \times \mathbf{A} </math>.
* [[Scalar–vector–tensor decomposition]]
* [[Hodge theory]] generalizing Helmholtz decomposition
* [[Polar factorization theorem]]
* ''Helmholtz–Leray decomposition'' used for defining the [[Leray projection]]


<ref name="edwards1922">Joseph Edwards: ''A Treatise on the Integral Calculus''. Volume 2. Chelsea Publishing Company, 1922.</ref>
==Notes==

{{Reflist|30em}}
<ref name="gibbs1901">[[J. W. Gibbs]], [[Edwin Bidwell Wilson]]: ''[https://archive.org/stream/117714283#page/236/mode/2up Vector Analysis]''. 1901, p. 237, link from [[Internet Archive]].</ref>

<ref name="girault1986">[[Vivette Girault|V. Girault]], P.A. Raviart: ''Finite Element Methods for Navier–Stokes Equations: Theory and Algorithms.'' Springer Series in Computational Mathematics. Springer-Verlag, 1986.</ref>

<ref name="glotzl2020">Erhard Glötzl, Oliver Richters: ''Helmholtz Decomposition and Rotation Potentials in n-dimensional Cartesian Coordinates''. 2020, {{arXiv|2012.13157}}, {{DOI|10.48550/arXiv.2012.13157}}.</ref>

<ref name="glotzl2023">Erhard Glötzl, Oliver Richters: ''Helmholtz decomposition and potential functions for n-dimensional analytic vector fields''. In: ''[[Journal of Mathematical Analysis and Applications]]'' 525(2), 127138, 2023, {{DOI|10.1016/j.jmaa.2023.127138}}, {{arXiv|2102.09556v3}}. ''Mathematica'' worksheet at {{DOI|10.5281/zenodo.7512798}}.</ref>

<ref name="gregory1996">R. Douglas Gregory: ''Helmholtz's Theorem when the domain is Infinite and when the field has singular points''. In: ''[[The Quarterly Journal of Mechanics and Applied Mathematics]]'' 49.3, 1996, pp. 439–450, {{DOI|10.1093/qjmam/49.3.439}}.</ref>

<ref name="griffiths1999">[[David J. Griffiths]]: ''Introduction to Electrodynamics''. Prentice-Hall, 1999, p. 556.</ref>

<ref name="gurtin1962">Morton E. Gurtin: ''On Helmholtz’s theorem and the completeness of the Papkovich-Neuber stress functions for infinite domains''. In: ''[[Archive for Rational Mechanics and Analysis]]'' 9.1, 1962, pp. 225–233, {{DOI|10.1007/BF00253346}}.</ref>

<ref name="heaviside1893">[[Oliver Heaviside]]: ''Electromagnetic theory''. Volume 1, "The Electrician" printing and publishing company, limited, 1893.</ref>

<ref name="helmholtz1858">[[Hermann von Helmholtz]]: ''Über Integrale der hydrodynamischen Gleichungen, welche den Wirbelbewegungen entsprechen''. In: ''[[Journal für die reine und angewandte Mathematik]]'' 55, 1858, pp. 25–55, {{DOI|10.1515/crll.1858.55.25}} ([http://resolver.sub.uni-goettingen.de/purl?GDZPPN002150212 sub.uni-goettingen.de], [https://www.digizeitschriften.de/dms/img/?PID=GDZPPN002150212 digizeitschriften.de]). On page 38, the components of the fluid's velocity (''u'',&nbsp;''v'',&nbsp;''w'') are expressed in terms of the gradient of a scalar potential P and the curl of a vector potential&nbsp;(''L'',&nbsp;''M'',&nbsp;''N'').</ref>

<ref name="johnson1881">[[William Woolsey Johnson]]: ''An Elementary Treatise on the Integral Calculus: Founded on the Method of Rates Or Fluxions''. John Wiley & Sons, 1881.<br />See also: [[Method of Fluxions]].</ref>

<ref name="koenigsberger1906">[[Leo Koenigsberger]]: ''Hermann von Helmholtz''. Clarendon Press, 1906, p. 357.</ref>

<ref name="kustepeli2016">Alp Kustepeli: ''On the Helmholtz Theorem and Its Generalization for Multi-Layers''. In: ''[[Electromagnetics]]'' 36.3, 2016, pp. 135–148, {{DOI|10.1080/02726343.2016.1149755}}.</ref>

<ref name="littlejohn">Robert Littlejohn: [http://bohr.physics.berkeley.edu/classes/221/1112/notes/hamclassemf.pdf ''The Classical Electromagnetic Field Hamiltonian'']. Online lecture notes, berkeley.edu.</ref>

<ref name="lorenz1963">[[Edward N. Lorenz]]: ''Deterministic Nonperiodic Flow''. In: ''[[Journal of the Atmospheric Sciences]]'' 20.2, 1963, pp. 130–141, {{DOI|10.1175/1520-0469(1963)020<0130:DNF>2.0.CO;2}}.</ref>

<ref name="murray1898">[[Daniel Murray (mathematician)|Daniel Alexander Murray]]: ''An Elementary Course in the Integral Calculus''. American Book Company, 1898. p. 8.</ref>

<ref name="peitgen1992">Heinz-Otto Peitgen, Hartmut Jürgens, Dietmar Saupe: ''Strange Attractors: The Locus of Chaos''. In: ''Chaos and Fractals''. Springer, New York, pp. 655–768. {{DOI|10.1007/978-1-4757-4740-9_13}}.</ref>

<ref name="petrascheck2015">Dietmar Petrascheck: ''The Helmholtz decomposition revisited''. In: ''[[European Journal of Physics]]'' 37.1, 2015, Artikel 015201, {{DOI|10.1088/0143-0807/37/1/015201}}.</ref>

<ref name="petrascheck2017">D. Petrascheck, R. Folk: ''Helmholtz decomposition theorem and Blumenthal’s extension by regularization''. In: ''Condensed Matter Physics'' 20(1), 13002, 2017, {{DOI|10.5488/CMP.20.13002}}.</ref>

<ref name="shaw1922">James Byrnie Shaw: ''Vector Calculus: With Applications to Physics''. D. Van Nostrand, 1922, p. 205.<br />See also: [[Green's Theorem]].</ref>

<ref name="sprossig2009">Wolfgang Sprössig: ''On Helmholtz decompositions and their generalizations – An overview''. In: ''[[Mathematical Methods in the Applied Sciences]]'' 33.4, 2009, pp. 374–383, {{DOI|10.1002/mma.1212}}.</ref>

<ref name="stewart2011">A. M. Stewart: ''Longitudinal and transverse components of a vector field''. In: ''Sri Lankan Journal of Physics'' 12, pp. 33–42, 2011, {{DOI|10.4038/sljp.v12i0.3504}} {{arxiv|0801.0335}}</ref>

<ref name="stokes1849">[[George Gabriel Stokes]]: ''On the Dynamical Theory of Diffraction''. In: ''Transactions of the [[Cambridge Philosophical Society]]'' 9, 1849, pp. 1–62. {{DOI|10.1017/cbo9780511702259.015}}, see pp. 9–10.</ref>

<ref name="suda2019">Tomoharu Suda: ''Construction of Lyapunov functions using Helmholtz–Hodge decomposition''. In: ''Discrete & Continuous Dynamical Systems – A'' 39.5, 2019, pp. 2437–2454, {{DOI|10.3934/dcds.2019103}}.</ref>

<ref name="suda2020">Tomoharu Suda: ''Application of Helmholtz–Hodge decomposition to the study of certain vector fields''. In: ''[[Journal of Physics]] A: Mathematical and Theoretical'' 53.37, 2020, pp. 375703. {{DOI|10.1088/1751-8121/aba657}}.</ref>

<ref name="trancong1993">Ton Tran-Cong: ''On Helmholtz’s Decomposition Theorem and Poissons’s Equation with an Infinite Domain''. In: ''[[Quarterly of Applied Mathematics]]'' 51.1, 1993, pp. 23–35, {{JSTOR|43637902}}.</ref>

<ref name="vermont">{{cite web |url=http://www.cems.uvm.edu/~oughstun/LectureNotes141/Topic_03_(Helmholtz'%20Theorem).pdf |title=Helmholtz' Theorem |publisher=University of Vermont| access-date=2011-03-11 | archive-url=https://web.archive.org/web/20120813005804/http://www.cems.uvm.edu/~oughstun/LectureNotes141/Topic_03_(Helmholtz'%20Theorem).pdf| archive-date=2012-08-13| url-status=dead}}</ref>

<ref name="warner1983">Frank W. Warner: ''The Hodge Theorem''. In: ''Foundations of Differentiable Manifolds and Lie Groups''. (= Graduate Texts in Mathematics 94). Springer, New York 1983, {{DOI|10.1007/978-1-4757-1799-0_6}}.</ref>

<ref name="woolhouse1854">[[Wesley Stoker Barker Woolhouse]]: ''Elements of the differential calculus''. Weale, 1854.</ref>

<ref name="zhou2012">Joseph Xu Zhou, M. D. S. Aliyu, Erik Aurell, Sui Huang: ''Quasi-potential landscape in complex multi-stable systems''. In: ''[[Journal of The Royal Society Interface]]'' 9.77, 2012, pp. 3539–3553, {{DOI|10.1098/rsif.2012.0434}}.</ref>

}}


==References==
==References==


===General references===
{{refbegin}}
{{refbegin}}
* [[George B. Arfken]] and Hans J. Weber, ''Mathematical Methods for Physicists'', 4th edition, Academic Press: San Diego (1995) pp.&nbsp;92–93
* [[George B. Arfken]] and Hans J. Weber, ''Mathematical Methods for Physicists'', 4th edition, Academic Press: San Diego (1995) pp.&nbsp;92–93
Line 308: Line 463:
* [[Rutherford Aris]], ''Vectors, tensors, and the basic equations of fluid mechanics'', Prentice-Hall (1962), {{oclc|299650765}}, pp.&nbsp;70–72
* [[Rutherford Aris]], ''Vectors, tensors, and the basic equations of fluid mechanics'', Prentice-Hall (1962), {{oclc|299650765}}, pp.&nbsp;70–72
{{refend}}
{{refend}}

===References for the weak formulation===
{{refbegin}}
* {{cite journal | last1 = Amrouche | first1 = C. | last2 = Bernardi | first2 = C. | author2-link = Christine Bernardi | last3 = Dauge | first3 = M. | author3-link = Monique Dauge| last4 = Girault | first4 = V.|author4-link= Vivette Girault | title = Vector potentials in three dimensional non-smooth domains | journal = Mathematical Methods in the Applied Sciences | volume = 21 | pages = 823–864| year= 1998 | issue = 9 | doi=10.1002/(sici)1099-1476(199806)21:9<823::aid-mma976>3.0.co;2-b|bibcode = 1998MMAS...21..823A }}
* R. Dautray and J.-L. Lions. ''Spectral Theory and Applications,'' volume 3 of Mathematical Analysis and Numerical Methods for Science and Technology. Springer-Verlag, 1990.
* [[Vivette Girault|V. Girault]] and P.A. Raviart. ''Finite Element Methods for Navier–Stokes Equations: Theory and Algorithms.'' Springer Series in Computational Mathematics. Springer-Verlag, 1986.
{{refend}}

==External links==
*[http://mathworld.wolfram.com/HelmholtzsTheorem.html Helmholtz theorem] on [[MathWorld]]


{{Authority control}}
{{Authority control}}

Revision as of 13:26, 26 March 2023

In physics and mathematics, in the area of vector calculus, Helmholtz's theorem,[1][2] also known as the fundamental theorem of vector calculus,[3][4][5][6][7][8][9] states that any sufficiently smooth, rapidly decaying vector field in three dimensions can be resolved into the sum of an irrotational (curl-free) vector field and a solenoidal (divergence-free) vector field; this is known as the Helmholtz decomposition or Helmholtz representation. It is named after Hermann von Helmholtz.

Definition

For a vector field defined on a domain , a Helmholtz decomposition is a pair of vector fields and such that:

Here, is a scalar potential, is its gradient, and is the divergence. The irrotational vector field is called a gradient field and is called a solenoidal field or rotation field. This decomposition does not exist for all vector fields and is not unique.[10]

History

The Helmholtz decomposition in three dimensions was first described in 1849[11] by George Gabriel Stokes for a theory of diffraction. Hermann von Helmholtz published his paper on some hydrodynamic basic equations in 1858,[12][13] which was part of his research on the Helmholtz's theorems describing the motion of fluid in the vicinity of vortex lines.[13] Their derivation required the vector fields to decay sufficiently fast at infinity. Later, this condition could be relaxed, and the Helmholtz decomposition could be extended to higher dimensions.[10][14][15]

The decomposition has become an important tool for many problems in theoretical physics,[13][16] but has also found applications in animation, computer vision as well as robotics.[17] In this context, the Helmholtz decomposition has been extended to higher-dimensional spaces and applied to riemannian manifolds as the Helmholtz-Hodge decomposition using differential geometry and tensor calculus.[10][13][16][17]

Three-dimensional space

Many physics textbooks restrict the Helmholtz decomposition to the three-dimensional space and limit its application to vector fields that decay sufficiently fast at infinity or to bump function that are defined on a bounded domain. Then, a vector potential can be defined, such that the rotation field is given by , using the Curl of a vector field.[18]

Let be a vector field on a bounded domain , which is twice continuously differentiable inside , and let be the surface that encloses the domain . Then can be decomposed into a curl-free component and a divergence-free component as follows:[19]

where

and is the nabla operator with respect to , not .

If and is therefore unbounded, and vanishes at least as fast as as , then one has[20]

This holds in particular if is twice continuously differentiable in and of bounded support.

Derivation

Proof

Suppose we have a vector function of which we know the curl, , and the divergence, , in the domain and the fields on the boundary. Writing the function using delta function in the form

where is the Laplace operator, we have

where we have used the definition of the vector Laplacian:

differentiation/integration with respect to by and in the last line, linearity of function arguments:

Then using the vectorial identities

we get

Thanks to the divergence theorem the equation can be rewritten as

with outward surface normal .

Defining

we finally obtain

Fields with prescribed divergence and curl

The term "Helmholtz theorem" can also refer to the following. Let C be a solenoidal vector field and d a scalar field on R3 which are sufficiently smooth and which vanish faster than 1/r2 at infinity. Then there exists a vector field F such that

if additionally the vector field F vanishes as r → ∞, then F is unique.[20]

In other words, a vector field can be constructed with both a specified divergence and a specified curl, and if it also vanishes at infinity, it is uniquely specified by its divergence and curl. This theorem is of great importance in electrostatics, since Maxwell's equations for the electric and magnetic fields in the static case are of exactly this type.[20] The proof is by a construction generalizing the one given above: we set

where represents the Newtonian potential operator. (When acting on a vector field, such as ∇ × F, it is defined to act on each component.)

Weak formulation

The Helmholtz decomposition can be generalized by reducing the regularity assumptions (the need for the existence of strong derivatives). Suppose Ω is a bounded, simply-connected, Lipschitz domain. Every square-integrable vector field u ∈ (L2(Ω))3 has an orthogonal decomposition:[21][22][23]

where φ is in the Sobolev space H1(Ω) of square-integrable functions on Ω whose partial derivatives defined in the distribution sense are square integrable, and AH(curl, Ω), the Sobolev space of vector fields consisting of square integrable vector fields with square integrable curl.

For a slightly smoother vector field uH(curl, Ω), a similar decomposition holds:

where φH1(Ω), v ∈ (H1(Ω))d.

Derivation from the Fourier transform

Note that in the theorem stated here, we have imposed the condition that if is not defined on a bounded domain, then shall decay faster than . Thus, the Fourier transform of , denoted as , is guaranteed to exist. We apply the convention

The Fourier transform of a scalar field is a scalar field, and the Fourier transform of a vector field is a vector field of same dimension.

Now consider the following scalar and vector fields:

Hence

Longitudinal and transverse fields

A terminology often used in physics refers to the curl-free component of a vector field as the longitudinal component and the divergence-free component as the transverse component.[24] This terminology comes from the following construction: Compute the three-dimensional Fourier transform of the vector field . Then decompose this field, at each point k, into two components, one of which points longitudinally, i.e. parallel to k, the other of which points in the transverse direction, i.e. perpendicular to k. So far, we have

Now we apply an inverse Fourier transform to each of these components. Using properties of Fourier transforms, we derive:

Since and ,

we can get

so this is indeed the Helmholtz decomposition.[25]

Generalization to higher dimensions

Matrix approach

The generalization to dimensions cannot be done with a vector potential, since the rotation operator and the cross product are defined only in three dimensions.

Let be a vector field on a bounded domain which decays faster then for and .

The scalar potential is defined similar to the three dimensional case as:

where as the integration kernel is again the fundamental solution of Laplace's equation, but in d-dimensional space:
with the volume of the d-dimensional unit balls and the gamma function.

For , is just equal to , yielding the same prefactor as above. The rotational potential is an antisymmetric matrix with the elements:

Above the diagonal are entries which occur again mirrored at the diagonal, but with a negative sign. In the three-dimensional case, the matrix elements just correspond to the components of the vector potential . However, such a matrix potential can be written as a vector only in the three-dimensional case, because is valid only for .

As in the three-dimensional case, the gradient field is defined as

The rotational field, on the other hand, is defined in the general case as the row divergence of the matrix:
In three-dimensional space, this is equivalent to the rotation of the vector potential.[10][26]

Tensor approach

In a -dimensional vector space with , can be replaced by the appropriate Green's function for the Laplacian, defined by

where Einstein summation convention is used for the index . For example, in 2D.

Following the same steps as above, we can write

where is the Kronecker delta (and the summation convention is again used). In place of the definition of the vector Laplacian used above, we now make use of an identity for the Levi-Civita symbol ,
which is valid in dimensions, where is a -component multi-index. This gives

We can therefore write

where
Note that the vector potential is replaced by a rank- tensor in dimensions.

For a further generalization to manifolds, see the discussion of Hodge decomposition below.

Differential forms

The Hodge decomposition is closely related to the Helmholtz decomposition,[27] generalizing from vector fields on R3 to differential forms on a Riemannian manifold M. Most formulations of the Hodge decomposition require M to be compact.[28] Since this is not true of R3, the Hodge decomposition theorem is not strictly a generalization of the Helmholtz theorem. However, the compactness restriction in the usual formulation of the Hodge decomposition can be replaced by suitable decay assumptions at infinity on the differential forms involved, giving a proper generalization of the Helmholtz theorem.

Extensions to fields not decaying at infinity

Most textbooks only deal with vector fields decaying faster then with at infinity.[18][15] However, Otto Blumenthal showed in 1905 that an adapted integration kernel can be used to integrate fields decaying faster than with , which is substantially less strict. To achieve this, the kernel in the convolution integrals has to be replaced by .[29] With even more complex integration kernels, solutions can be found even for divergent functions that need not grow faster than polynomial.[14][15][26][30]

For all analytic vector fields that need not go to zero even at infinity, methods based on partial integration and the Cauchy formula for repeated integration[31] can be used to compute closed-form solutions of the rotation and scalar potentials, as in the case of multivariate polynomial, sine, cosine, and exponential functions.[10]

Uniqueness of the solution

In general, the Helmholtz decomposition is not uniquely defined. A harmonic function is a function that satisfies . By adding to the scalar potential , a different Helmholtz decomposition can be obtained:

For vector fields , decaying at infinity, it is a plausible choice that scalar and rotation potentials also decay at infinity. Because is the only harmonic function with this property, which follows from Liouville's theorem, this guarantees the uniqueness of the gradient and rotation fields.[32]

This uniqueness does not apply to the potentials: In the three-dimensional case, the scalar and vector potential jointly have four components, whereas the vector field has only three. The vector field is invariant to gauge transformations and the choice of appropriate potentials known as gauge fixing is the subject of gauge theory. Important examples from physics are the Lorenz gauge condition and the Coulomb gauge. An alternative is to use the poloidal–toroidal decomposition.

Applications

Electrodynamics

The Helmholtz theorem is of particular interest in electrodynamics, since it can be used to write Maxwell's equations in the potential image and solve them more easily. The Helmholtz decomposition can be used to prove that, given electric current density and charge density, the electric field and the magnetic flux density can be determined. They are unique if the densities vanish at infinity and one assumes the same for the potentials.[18]

Fluid dynamics

In fluid dynamics, the Helmholtz projection plays an important role, especially for the solvability theory of the Navier-Stokes equations. If the Helmholtz projection is applied to the linearized incompressible Navier-Stokes equations, the Stokes equation is obtained. This depends only on the velocity of the particles in the flow, but no longer on the static pressure, allowing the equation to be reduced to one unknown. However, both equations, the Stokes and linearized equations, are equivalent. The operator is called the Stokes operator.[33]

Dynamical systems theory

In the theory of dynamical systems, Helmholtz decomposition can be used to determine "quasipotentials" as well as to compute Lyapunov functions in some cases.[34][35][36]

For some dynamical systems such as the Lorenz system (Edward N. Lorenz, 1963[37]), a simplified model for atmospheric convection, a closed-form expression of the Helmholtz decomposition can be obtained:

The Helmholtz decomposition of , with the scalar potential is given as:

The quadratic scalar potential provides motion in the direction of the coordinate origin, which is responsible for the stable fix point for some parameter range. For other parameters, the rotation field ensures that a strange attractor is created, causing the model to exhibit a butterfly effect.[10][38]

Computer animation and robotics

The Helmholtz decomposition is also used in the field of computer engineering. This includes robotics, image reconstruction but also computer animation, where the decomposition is used for realistic visualization of fluids or vector fields.[17][39]

See also

Notes

  1. ^ Jean Bladel: On Helmholtz's Theorem in Finite Regions. Midwestern Universities Research Association, 1958.
  2. ^ Leo Koenigsberger: Hermann von Helmholtz. Clarendon Press, 1906, p. 357.
  3. ^ Daniel Alexander Murray: An Elementary Course in the Integral Calculus. American Book Company, 1898. p. 8.
  4. ^ J. W. Gibbs, Edwin Bidwell Wilson: Vector Analysis. 1901, p. 237, link from Internet Archive.
  5. ^ Oliver Heaviside: Electromagnetic theory. Volume 1, "The Electrician" printing and publishing company, limited, 1893.
  6. ^ Wesley Stoker Barker Woolhouse: Elements of the differential calculus. Weale, 1854.
  7. ^ William Woolsey Johnson: An Elementary Treatise on the Integral Calculus: Founded on the Method of Rates Or Fluxions. John Wiley & Sons, 1881.
    See also: Method of Fluxions.
  8. ^ James Byrnie Shaw: Vector Calculus: With Applications to Physics. D. Van Nostrand, 1922, p. 205.
    See also: Green's Theorem.
  9. ^ Joseph Edwards: A Treatise on the Integral Calculus. Volume 2. Chelsea Publishing Company, 1922.
  10. ^ a b c d e f Erhard Glötzl, Oliver Richters: Helmholtz decomposition and potential functions for n-dimensional analytic vector fields. In: Journal of Mathematical Analysis and Applications 525(2), 127138, 2023, doi:10.1016/j.jmaa.2023.127138, arXiv:2102.09556v3. Mathematica worksheet at doi:10.5281/zenodo.7512798.
  11. ^ George Gabriel Stokes: On the Dynamical Theory of Diffraction. In: Transactions of the Cambridge Philosophical Society 9, 1849, pp. 1–62. doi:10.1017/cbo9780511702259.015, see pp. 9–10.
  12. ^ Hermann von Helmholtz: Über Integrale der hydrodynamischen Gleichungen, welche den Wirbelbewegungen entsprechen. In: Journal für die reine und angewandte Mathematik 55, 1858, pp. 25–55, doi:10.1515/crll.1858.55.25 (sub.uni-goettingen.de, digizeitschriften.de). On page 38, the components of the fluid's velocity (uvw) are expressed in terms of the gradient of a scalar potential P and the curl of a vector potential (LMN).
  13. ^ a b c d Alp Kustepeli: On the Helmholtz Theorem and Its Generalization for Multi-Layers. In: Electromagnetics 36.3, 2016, pp. 135–148, doi:10.1080/02726343.2016.1149755.
  14. ^ a b Ton Tran-Cong: On Helmholtz’s Decomposition Theorem and Poissons’s Equation with an Infinite Domain. In: Quarterly of Applied Mathematics 51.1, 1993, pp. 23–35, JSTOR 43637902.
  15. ^ a b c D. Petrascheck, R. Folk: Helmholtz decomposition theorem and Blumenthal’s extension by regularization. In: Condensed Matter Physics 20(1), 13002, 2017, doi:10.5488/CMP.20.13002.
  16. ^ a b Wolfgang Sprössig: On Helmholtz decompositions and their generalizations – An overview. In: Mathematical Methods in the Applied Sciences 33.4, 2009, pp. 374–383, doi:10.1002/mma.1212.
  17. ^ a b c Harsh Bhatia, Gregory Norgard, Valerio Pascucci, Peer-Timo Bremer: The Helmholtz-Hodge Decomposition – A Survey. In: IEEE Transactions on Visualization and Computer Graphics 19.8, 2013, pp. 1386–1404, doi:10.1109/tvcg.2012.316.
  18. ^ a b c Dietmar Petrascheck: The Helmholtz decomposition revisited. In: European Journal of Physics 37.1, 2015, Artikel 015201, doi:10.1088/0143-0807/37/1/015201.
  19. ^ "Helmholtz' Theorem" (PDF). University of Vermont. Archived from the original (PDF) on 2012-08-13. Retrieved 2011-03-11.
  20. ^ a b c David J. Griffiths: Introduction to Electrodynamics. Prentice-Hall, 1999, p. 556.
  21. ^ Cherif Amrouche, Christine Bernardi, Monique Dauge, Vivette Girault: Vector potentials in three dimensional non-smooth domains. In: Mathematical Methods in the Applied Sciences 21(9), 1998, pp. 823–864, doi:10.1002/(sici)1099-1476(199806)21:9<823::aid-mma976>3.0.co;2-b, Bibcode:/abstract 1998MMAS...21..823A .
  22. ^ R. Dautray and J.-L. Lions. Spectral Theory and Applications, volume 3 of Mathematical Analysis and Numerical Methods for Science and Technology. Springer-Verlag, 1990.
  23. ^ V. Girault, P.A. Raviart: Finite Element Methods for Navier–Stokes Equations: Theory and Algorithms. Springer Series in Computational Mathematics. Springer-Verlag, 1986.
  24. ^ A. M. Stewart: Longitudinal and transverse components of a vector field. In: Sri Lankan Journal of Physics 12, pp. 33–42, 2011, doi:10.4038/sljp.v12i0.3504 arXiv:0801.0335
  25. ^ Robert Littlejohn: The Classical Electromagnetic Field Hamiltonian. Online lecture notes, berkeley.edu.
  26. ^ a b Erhard Glötzl, Oliver Richters: Helmholtz Decomposition and Rotation Potentials in n-dimensional Cartesian Coordinates. 2020, arXiv:2012.13157, doi:10.48550/arXiv.2012.13157.
  27. ^ Frank W. Warner: The Hodge Theorem. In: Foundations of Differentiable Manifolds and Lie Groups. (= Graduate Texts in Mathematics 94). Springer, New York 1983, doi:10.1007/978-1-4757-1799-0_6.
  28. ^ Cantarella, Jason; DeTurck, Dennis; Gluck, Herman (2002). "Vector Calculus and the Topology of Domains in 3-Space". The American Mathematical Monthly. 109 (5): 409–442. doi:10.2307/2695643. JSTOR 2695643.
  29. ^ Otto Blumenthal: Über die Zerlegung unendlicher Vektorfelder. In: Mathematische Annalen 61.2, 1905, pp. 235–250, doi:10.1007/BF01457564.
  30. ^ Morton E. Gurtin: On Helmholtz’s theorem and the completeness of the Papkovich-Neuber stress functions for infinite domains. In: Archive for Rational Mechanics and Analysis 9.1, 1962, pp. 225–233, doi:10.1007/BF00253346.
  31. ^ Augustin-Louis Cauchy: Trente-Cinquième Leçon. In: Résumé des leçons données à l’École royale polytechnique sur le calcul infinitésimal. Imprimerie Royale, Paris 1823, pp. 133–140 (gallica.bnf.fr).
  32. ^ Sheldon Axler, Paul Bourdon, Wade Ramey: Bounded Harmonic Functions. In: Harmonic Function Theory (= Graduate Texts in Mathematics 137). Springer, New York 1992, pp. 31–44, doi:10.1007/0-387-21527-1_2.
  33. ^ Alexandre J. Chorin, Jerrold E. Marsden: A Mathematical Introduction to Fluid Mechanics (= Texts in Applied Mathematics 4). Springer US, New York 1990, doi:10.1007/978-1-4684-0364-0.
  34. ^ Tomoharu Suda: Construction of Lyapunov functions using Helmholtz–Hodge decomposition. In: Discrete & Continuous Dynamical Systems – A 39.5, 2019, pp. 2437–2454, doi:10.3934/dcds.2019103.
  35. ^ Tomoharu Suda: Application of Helmholtz–Hodge decomposition to the study of certain vector fields. In: Journal of Physics A: Mathematical and Theoretical 53.37, 2020, pp. 375703. doi:10.1088/1751-8121/aba657.
  36. ^ Joseph Xu Zhou, M. D. S. Aliyu, Erik Aurell, Sui Huang: Quasi-potential landscape in complex multi-stable systems. In: Journal of The Royal Society Interface 9.77, 2012, pp. 3539–3553, doi:10.1098/rsif.2012.0434.
  37. ^ Edward N. Lorenz: Deterministic Nonperiodic Flow. In: Journal of the Atmospheric Sciences 20.2, 1963, pp. 130–141, doi:10.1175/1520-0469(1963)020<0130:DNF>2.0.CO;2.
  38. ^ Heinz-Otto Peitgen, Hartmut Jürgens, Dietmar Saupe: Strange Attractors: The Locus of Chaos. In: Chaos and Fractals. Springer, New York, pp. 655–768. doi:10.1007/978-1-4757-4740-9_13.
  39. ^ Hersh Bhatia, Valerio Pascucci, Peer-Timo Bremer: The Natural Helmholtz-Hodge Decomposition for Open-Boundary Flow Analysis. In: IEEE Transactions on Visualization and Computer Graphics 20.11, Nov. 2014, pp. 1566–1578, Nov. 2014, doi:10.1109/TVCG.2014.2312012.
Cite error: A list-defined reference named "gregory1996" is not used in the content (see the help page).

References

  • George B. Arfken and Hans J. Weber, Mathematical Methods for Physicists, 4th edition, Academic Press: San Diego (1995) pp. 92–93
  • George B. Arfken and Hans J. Weber, Mathematical Methods for Physicists – International Edition, 6th edition, Academic Press: San Diego (2005) pp. 95–101
  • Rutherford Aris, Vectors, tensors, and the basic equations of fluid mechanics, Prentice-Hall (1962), OCLC 299650765, pp. 70–72