Green's identities

Last updated

In mathematics, Green's identities are a set of three identities in vector calculus relating the bulk with the boundary of a region on which differential operators act. They are named after the mathematician George Green, who discovered Green's theorem.

Contents

Green's first identity

This identity is derived from the divergence theorem applied to the vector field F = ψφ while using an extension of the product rule that ∇ ⋅ (ψX ) = ∇ψX + ψ ∇⋅X: Let φ and ψ be scalar functions defined on some region URd, and suppose that φ is twice continuously differentiable, and ψ is once continuously differentiable. Using the product rule above, but letting X = ∇φ, integrate ∇⋅(ψφ) over U. Then [1]

where ∆ ≡ ∇2 is the Laplace operator, U is the boundary of region U, n is the outward pointing unit normal to the surface element dS and dS = ndS is the oriented surface element.

This theorem is a special case of the divergence theorem, and is essentially the higher dimensional equivalent of integration by parts with ψ and the gradient of φ replacing u and v.

Note that Green's first identity above is a special case of the more general identity derived from the divergence theorem by substituting F = ψΓ,

Green's second identity

If φ and ψ are both twice continuously differentiable on UR3, and ε is once continuously differentiable, one may choose F = ψεφφεψ to obtain

For the special case of ε = 1 all across UR3, then,

In the equation above, φ/∂n is the directional derivative of φ in the direction of the outward pointing surface normal n of the surface element dS,

Explicitly incorporating this definition in the Green's second identity with ε = 1 results in

In particular, this demonstrates that the Laplacian is a self-adjoint operator in the L2 inner product for functions vanishing on the boundary so that the right hand side of the above identity is zero.

Green's third identity

Green's third identity derives from the second identity by choosing φ = G, where the Green's function G is taken to be a fundamental solution of the Laplace operator, ∆. This means that:

For example, in R3, a solution has the form

Green's third identity states that if ψ is a function that is twice continuously differentiable on U, then

A simplification arises if ψ is itself a harmonic function, i.e. a solution to the Laplace equation. Then 2ψ = 0 and the identity simplifies to

The second term in the integral above can be eliminated if G is chosen to be the Green's function that vanishes on the boundary of U (Dirichlet boundary condition),

This form is used to construct solutions to Dirichlet boundary condition problems. Solutions for Neumann boundary condition problems may also be simplified, though the Divergence theorem applied to the differential equation defining Green's functions shows that the Green's function cannot integrate to zero on the boundary, and hence cannot vanish on the boundary. See Green's functions for the Laplacian or [2] for a detailed argument, with an alternative.

It can be further verified that the above identity also applies when ψ is a solution to the Helmholtz equation or wave equation and G is the appropriate Green's function. In such a context, this identity is the mathematical expression of the Huygens principle, and leads to Kirchhoff's diffraction formula and other approximations.

On manifolds

Green's identities hold on a Riemannian manifold. In this setting, the first two are

where u and v are smooth real-valued functions on M, dV is the volume form compatible with the metric, is the induced volume form on the boundary of M, N is the outward oriented unit vector field normal to the boundary, and Δu = div(grad u) is the Laplacian.

Green's vector identity

Green's second identity establishes a relationship between second and (the divergence of) first order derivatives of two scalar functions. In differential form

where pm and qm are two arbitrary twice continuously differentiable scalar fields. This identity is of great importance in physics because continuity equations can thus be established for scalar fields such as mass or energy. [3]

In vector diffraction theory, two versions of Green's second identity are introduced.

One variant invokes the divergence of a cross product [4] [5] [6] and states a relationship in terms of the curl-curl of the field

This equation can be written in terms of the Laplacians,

However, the terms

could not be readily written in terms of a divergence.

The other approach introduces bi-vectors, this formulation requires a dyadic Green function. [7] [8] The derivation presented here avoids these problems. [9]

Consider that the scalar fields in Green's second identity are the Cartesian components of vector fields, i.e.,

Summing up the equation for each component, we obtain

The LHS according to the definition of the dot product may be written in vector form as

The RHS is a bit more awkward to express in terms of vector operators. Due to the distributivity of the divergence operator over addition, the sum of the divergence is equal to the divergence of the sum, i.e.,

Recall the vector identity for the gradient of a dot product,

which, written out in vector components is given by

This result is similar to what we wish to evince in vector terms 'except' for the minus sign. Since the differential operators in each term act either over one vector (say ’s) or the other (’s), the contribution to each term must be

These results can be rigorously proven to be correct through evaluation of the vector components. Therefore, the RHS can be written in vector form as

Putting together these two results, a result analogous to Green's theorem for scalar fields is obtained,

Theorem for vector fields:

The curl of a cross product can be written as

Green's vector identity can then be rewritten as

Since the divergence of a curl is zero, the third term vanishes to yield Green's vector identity:

With a similar procedure, the Laplacian of the dot product can be expressed in terms of the Laplacians of the factors

As a corollary, the awkward terms can now be written in terms of a divergence by comparison with the vector Green equation,

This result can be verified by expanding the divergence of a scalar times a vector on the RHS.

See also

Related Research Articles

<span class="mw-page-title-main">Laplace's equation</span> Second-order partial differential equation

In mathematics and physics, Laplace's equation is a second-order partial differential equation named after Pierre-Simon Laplace, who first studied its properties. This is often written as

<span class="mw-page-title-main">Navier–Stokes equations</span> Equations describing the motion of viscous fluid substances

The Navier–Stokes equations are partial differential equations which describe the motion of viscous fluid substances, named after French engineer and physicist Claude-Louis Navier and Irish physicist and mathematician George Gabriel Stokes. They were developed over several decades of progressively building the theories, from 1822 (Navier) to 1842-1850 (Stokes).

<span class="mw-page-title-main">Potential flow</span> Velocity field as the gradient of a scalar function

In fluid dynamics, potential flow describes the velocity field as the gradient of a scalar function: the velocity potential. As a result, a potential flow is characterized by an irrotational velocity field, which is a valid approximation for several applications. The irrotationality of a potential flow is due to the curl of the gradient of a scalar always being equal to zero.

In fluid dynamics, Stokes' law is an empirical law for the frictional force – also called drag force – exerted on spherical objects with very small Reynolds numbers in a viscous fluid. It was derived by George Gabriel Stokes in 1851 by solving the Stokes flow limit for small Reynolds numbers of the Navier–Stokes equations.

<span class="mw-page-title-main">Stream function</span>

The stream function is defined for incompressible (divergence-free) flows in two dimensions – as well as in three dimensions with axisymmetry. The flow velocity components can be expressed as the derivatives of the scalar stream function. The stream function can be used to plot streamlines, which represent the trajectories of particles in a steady flow. The two-dimensional Lagrange stream function was introduced by Joseph Louis Lagrange in 1781. The Stokes stream function is for axisymmetrical three-dimensional flow, and is named after George Gabriel Stokes.

<span class="mw-page-title-main">Green's theorem</span> Theorem in calculus relating line and double integrals

In vector calculus, Green's theorem relates a line integral around a simple closed curve C to a double integral over the plane region D bounded by C. It is the two-dimensional special case of Stokes' theorem.

In physics, an operator is a function over a space of physical states onto another space of physical states. The simplest example of the utility of operators is the study of symmetry. Because of this, they are useful tools in classical mechanics. Operators are even more important in quantum mechanics, where they form an intrinsic part of the formulation of the theory.

In the calculus of variations, a field of mathematical analysis, the functional derivative relates a change in a functional to a change in a function on which the functional depends.

In differential geometry, the four-gradient is the four-vector analogue of the gradient from vector calculus.

<span class="mw-page-title-main">Vector calculus identities</span> Mathematical identities

The following are important identities involving derivatives and integrals in vector calculus.

<span class="mw-page-title-main">Charge density</span> Electric charge per unit length, area or volume

In electromagnetism, charge density is the amount of electric charge per unit length, surface area, or volume. Volume charge density is the quantity of charge per unit volume, measured in the SI system in coulombs per cubic meter (C⋅m−3), at any point in a volume. Surface charge density (σ) is the quantity of charge per unit area, measured in coulombs per square meter (C⋅m−2), at any point on a surface charge distribution on a two dimensional surface. Linear charge density (λ) is the quantity of charge per unit length, measured in coulombs per meter (C⋅m−1), at any point on a line charge distribution. Charge density can be either positive or negative, since electric charge can be either positive or negative.

<span class="mw-page-title-main">Mathematical descriptions of the electromagnetic field</span> Formulations of electromagnetism

There are various mathematical descriptions of the electromagnetic field that are used in the study of electromagnetism, one of the four fundamental interactions of nature. In this article, several approaches are discussed, although the equations are in terms of electric and magnetic fields, potentials, and charges with currents, generally speaking.

In quantum mechanics, the Pauli equation or Schrödinger–Pauli equation is the formulation of the Schrödinger equation for spin-½ particles, which takes into account the interaction of the particle's spin with an external electromagnetic field. It is the non-relativistic limit of the Dirac equation and can be used where particles are moving at speeds much less than the speed of light, so that relativistic effects can be neglected. It was formulated by Wolfgang Pauli in 1927.

<span class="mw-page-title-main">Liénard–Wiechert potential</span> Electromagnetic effect of point charges

The Liénard–Wiechert potentials describe the classical electromagnetic effect of a moving electric point charge in terms of a vector potential and a scalar potential in the Lorenz gauge. Stemming directly from Maxwell's equations, these describe the complete, relativistically correct, time-varying electromagnetic field for a point charge in arbitrary motion, but are not corrected for quantum mechanical effects. Electromagnetic radiation in the form of waves can be obtained from these potentials. These expressions were developed in part by Alfred-Marie Liénard in 1898 and independently by Emil Wiechert in 1900.

The intent of this article is to highlight the important points of the derivation of the Navier–Stokes equations as well as its application and formulation for different families of fluids.

In mathematics, vector spherical harmonics (VSH) are an extension of the scalar spherical harmonics for use with vector fields. The components of the VSH are complex-valued functions expressed in the spherical coordinate basis vectors.

In fluid dynamics, the Oseen equations describe the flow of a viscous and incompressible fluid at small Reynolds numbers, as formulated by Carl Wilhelm Oseen in 1910. Oseen flow is an improved description of these flows, as compared to Stokes flow, with the (partial) inclusion of convective acceleration.

<span class="mw-page-title-main">Stokes' theorem</span> Theorem in vector calculus

Stokes' theorem, also known as the Kelvin–Stokes theorem after Lord Kelvin and George Stokes, the fundamental theorem for curls or simply the curl theorem, is a theorem in vector calculus on . Given a vector field, the theorem relates the integral of the curl of the vector field over some surface, to the line integral of the vector field around the boundary of the surface. The classical theorem of Stokes can be stated in one sentence: The line integral of a vector field over a loop is equal to the flux of its curl through the enclosed surface. It is illustrated in the figure, where the direction of positive circulation of the bounding contour ∂Σ, and the direction n of positive flux through the surface Σ, are related by a right-hand-rule. For the right hand the fingers circulate along ∂Σ and the thumb is directed along n.

Curvilinear coordinates can be formulated in tensor calculus, with important applications in physics and engineering, particularly for describing transportation of physical quantities and deformation of matter in fluid mechanics and continuum mechanics.

Lagrangian field theory is a formalism in classical field theory. It is the field-theoretic analogue of Lagrangian mechanics. Lagrangian mechanics is used to analyze the motion of a system of discrete particles each with a finite number of degrees of freedom. Lagrangian field theory applies to continua and fields, which have an infinite number of degrees of freedom.

References

  1. Strauss, Walter. Partial Differential Equations: An Introduction. Wiley.
  2. Jackson, John David (1998-08-14). Classical Electrodynamics. John Wiley & Sons. p. 39.
  3. Guasti, M Fernández (2004-03-17). "Complementary fields conservation equation derived from the scalar wave equation". Journal of Physics A: Mathematical and General. IOP Publishing. 37 (13): 4107–4121. Bibcode:2004JPhA...37.4107F. doi:10.1088/0305-4470/37/13/013. ISSN   0305-4470.
  4. Love, Augustus E. H. (1901). "I. The integration of the equations of propagation of electric waves". Philosophical Transactions of the Royal Society of London. Series A, Containing Papers of a Mathematical or Physical Character. The Royal Society. 197 (287–299): 1–45. doi: 10.1098/rsta.1901.0013 . ISSN   0264-3952.
  5. Stratton, J. A.; Chu, L. J. (1939-07-01). "Diffraction Theory of Electromagnetic Waves". Physical Review. American Physical Society (APS). 56 (1): 99–107. Bibcode:1939PhRv...56...99S. doi:10.1103/physrev.56.99. ISSN   0031-899X.
  6. Bruce, Neil C (2010-07-22). "Double scatter vector-wave Kirchhoff scattering from perfectly conducting surfaces with infinite slopes". Journal of Optics. IOP Publishing. 12 (8): 085701. Bibcode:2010JOpt...12h5701B. doi:10.1088/2040-8978/12/8/085701. ISSN   2040-8978. S2CID   120636008.
  7. Franz, W (1950-09-01). "On the Theory of Diffraction". Proceedings of the Physical Society. Section A. IOP Publishing. 63 (9): 925–939. Bibcode:1950PPSA...63..925F. doi:10.1088/0370-1298/63/9/301. ISSN   0370-1298.
  8. Chen-To Tai (1972). "Kirchhoff theory: Scalar, vector, or dyadic?". IEEE Transactions on Antennas and Propagation. Institute of Electrical and Electronics Engineers (IEEE). 20 (1): 114–115. Bibcode:1972ITAP...20..114T. doi:10.1109/tap.1972.1140146. ISSN   0096-1973.
  9. Fernández-Guasti, M. (2012). "Green's Second Identity for Vector Fields". ISRN Mathematical Physics. Hindawi Limited. 2012: 1–7. doi: 10.5402/2012/973968 . ISSN   2090-4681.