Spherical harmonics

Last updated
Visual representations of the first few real spherical harmonics. Blue portions represent regions where the function is positive, and yellow portions represent where it is negative. The distance of the surface from the origin indicates the absolute value of
Y
l
m
(
th
,
ph
)
{\displaystyle Y_{\ell }^{m}(\theta ,\varphi )}
in angular direction
(
th
,
ph
)
{\displaystyle (\theta ,\varphi )}
. Spherical Harmonics.png
Visual representations of the first few real spherical harmonics. Blue portions represent regions where the function is positive, and yellow portions represent where it is negative. The distance of the surface from the origin indicates the absolute value of in angular direction .

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

Contents

Since the spherical harmonics form a complete set of orthogonal functions and thus an orthonormal basis, each function defined on the surface of a sphere can be written as a sum of these spherical harmonics. This is similar to periodic functions defined on a circle that can be expressed as a sum of circular functions (sines and cosines) via Fourier series. Like the sines and cosines in Fourier series, the spherical harmonics may be organized by (spatial) angular frequency, as seen in the rows of functions in the illustration on the right. Further, spherical harmonics are basis functions for irreducible representations of SO(3), the group of rotations in three dimensions, and thus play a central role in the group theoretic discussion of SO(3).

Spherical harmonics originate from solving Laplace's equation in the spherical domains. Functions that are solutions to Laplace's equation are called harmonics. Despite their name, spherical harmonics take their simplest form in Cartesian coordinates, where they can be defined as homogeneous polynomials of degree in that obey Laplace's equation. The connection with spherical coordinates arises immediately if one uses the homogeneity to extract a factor of radial dependence from the above-mentioned polynomial of degree ; the remaining factor can be regarded as a function of the spherical angular coordinates and only, or equivalently of the orientational unit vector specified by these angles. In this setting, they may be viewed as the angular portion of a set of solutions to Laplace's equation in three dimensions, and this viewpoint is often taken as an alternative definition. Notice, however, that spherical harmonics are not functions on the sphere which are harmonic with respect to the Laplace-Beltrami operator for the standard round metric on the sphere: the only harmonic functions in this sense on the sphere are the constants, since harmonic functions satisfy the Maximum principle. Spherical harmonics, as functions on the sphere, are eigenfunctions of the Laplace-Beltrami operator (see the section Higher dimensions below).

A specific set of spherical harmonics, denoted or , are known as Laplace's spherical harmonics, as they were first introduced by Pierre Simon de Laplace in 1782. [1] These functions form an orthogonal system, and are thus basic to the expansion of a general function on the sphere as alluded to above.

Spherical harmonics are important in many theoretical and practical applications, including the representation of multipole electrostatic and electromagnetic fields, electron configurations, gravitational fields, geoids, the magnetic fields of planetary bodies and stars, and the cosmic microwave background radiation. In 3D computer graphics, spherical harmonics play a role in a wide variety of topics including indirect lighting (ambient occlusion, global illumination, precomputed radiance transfer, etc.) and modelling of 3D shapes.

History

Pierre-Simon Laplace, 1749-1827 Laplace, Pierre-Simon, marquis de.jpg
Pierre-Simon Laplace, 1749–1827

Spherical harmonics were first investigated in connection with the Newtonian potential of Newton's law of universal gravitation in three dimensions. In 1782, Pierre-Simon de Laplace had, in his Mécanique Céleste, determined that the gravitational potential at a point x associated with a set of point masses mi located at points xi was given by

Each term in the above summation is an individual Newtonian potential for a point mass. Just prior to that time, Adrien-Marie Legendre had investigated the expansion of the Newtonian potential in powers of r = |x| and r1 = |x1|. He discovered that if rr1 then

where γ is the angle between the vectors x and x1. The functions are the Legendre polynomials, and they can be derived as a special case of spherical harmonics. Subsequently, in his 1782 memoir, Laplace investigated these coefficients using spherical coordinates to represent the angle γ between x1 and x. (See Applications of Legendre polynomials in physics for a more detailed analysis.)

In 1867, William Thomson (Lord Kelvin) and Peter Guthrie Tait introduced the solid spherical harmonics in their Treatise on Natural Philosophy , and also first introduced the name of "spherical harmonics" for these functions. The solid harmonics were homogeneous polynomial solutions of Laplace's equation

By examining Laplace's equation in spherical coordinates, Thomson and Tait recovered Laplace's spherical harmonics. (See the section below, "Harmonic polynomial representation".) The term "Laplace's coefficients" was employed by William Whewell to describe the particular system of solutions introduced along these lines, whereas others reserved this designation for the zonal spherical harmonics that had properly been introduced by Laplace and Legendre.

The 19th century development of Fourier series made possible the solution of a wide variety of physical problems in rectangular domains, such as the solution of the heat equation and wave equation. This could be achieved by expansion of functions in series of trigonometric functions. Whereas the trigonometric functions in a Fourier series represent the fundamental modes of vibration in a string, the spherical harmonics represent the fundamental modes of vibration of a sphere in much the same way. Many aspects of the theory of Fourier series could be generalized by taking expansions in spherical harmonics rather than trigonometric functions. Moreover, analogous to how trigonometric functions can equivalently be written as complex exponentials, spherical harmonics also possessed an equivalent form as complex-valued functions. This was a boon for problems possessing spherical symmetry, such as those of celestial mechanics originally studied by Laplace and Legendre.

The prevalence of spherical harmonics already in physics set the stage for their later importance in the 20th century birth of quantum mechanics. The (complex-valued) spherical harmonics are eigenfunctions of the square of the orbital angular momentum operator

and therefore they represent the different quantized configurations of atomic orbitals.

Laplace's spherical harmonics

Real (Laplace) spherical harmonics
Y
l
m
{\displaystyle Y_{\ell m}}
for
l
=
0
,
...
,
4
{\displaystyle \ell =0,\dots ,4}
(top to bottom) and
m
=
0
,
...
,
l
{\displaystyle m=0,\dots ,\ell }
(left to right). Zonal, sectoral, and tesseral harmonics are depicted along the left-most column, the main diagonal, and elsewhere, respectively. (The negative order harmonics
Y
l
(
-
m
)
{\displaystyle Y_{\ell (-m)}}
would be shown rotated about the z axis by
90
[?]
/
m
{\displaystyle 90^{\circ }/m}
with respect to the positive order ones.) Rotating spherical harmonics.gif
Real (Laplace) spherical harmonics for (top to bottom) and (left to right). Zonal, sectoral, and tesseral harmonics are depicted along the left-most column, the main diagonal, and elsewhere, respectively. (The negative order harmonics would be shown rotated about the z axis by with respect to the positive order ones.)
Alternative picture for the real spherical harmonics
Y
l
m
{\displaystyle Y_{\ell m}}
. Sphericalfunctions.svg
Alternative picture for the real spherical harmonics .

Laplace's equation imposes that the Laplacian of a scalar field f is zero. (Here the scalar field is understood to be complex, i.e. to correspond to a (smooth) function .) In spherical coordinates this is: [2]

Consider the problem of finding solutions of the form f(r, θ, φ) = R(r) Y(θ, φ). By separation of variables, two differential equations result by imposing Laplace's equation:

The second equation can be simplified under the assumption that Y has the form Y(θ, φ) = Θ(θ) Φ(φ). Applying separation of variables again to the second equation gives way to the pair of differential equations

for some number m. A priori, m is a complex constant, but because Φ must be a periodic function whose period evenly divides 2π, m is necessarily an integer and Φ is a linear combination of the complex exponentials e± imφ. The solution function Y(θ, φ) is regular at the poles of the sphere, where θ = 0, π. Imposing this regularity in the solution Θ of the second equation at the boundary points of the domain is a Sturm–Liouville problem that forces the parameter λ to be of the form λ = ( + 1) for some non-negative integer with ≥ |m|; this is also explained below in terms of the orbital angular momentum. Furthermore, a change of variables t = cos θ transforms this equation into the Legendre equation, whose solution is a multiple of the associated Legendre polynomial Pm
(cos θ)
. Finally, the equation for R has solutions of the form R(r) = A r + B r − 1; requiring the solution to be regular throughout R3 forces B = 0. [3]

Here the solution was assumed to have the special form Y(θ, φ) = Θ(θ) Φ(φ). For a given value of , there are 2 + 1 independent solutions of this form, one for each integer m with m. These angular solutions are a product of trigonometric functions, here represented as a complex exponential, and associated Legendre polynomials:

which fulfill

Here is called a spherical harmonic function of degree and order m, is an associated Legendre polynomial, N is a normalization constant, [4] and θ and φ represent colatitude and longitude, respectively. In particular, the colatitude θ, or polar angle, ranges from 0 at the North Pole, to π/2 at the Equator, to π at the South Pole, and the longitude φ, or azimuth, may assume all values with 0 ≤ φ < 2π. For a fixed integer , every solution Y(θ, φ), , of the eigenvalue problem

is a linear combination of . In fact, for any such solution, r Y(θ, φ) is the expression in spherical coordinates of a homogeneous polynomial that is harmonic (see below), and so counting dimensions shows that there are 2 + 1 linearly independent such polynomials.

The general solution to Laplace's equation in a ball centered at the origin is a linear combination of the spherical harmonic functions multiplied by the appropriate scale factor r,

where the are constants and the factors r Ym are known as (regular) solid harmonics . Such an expansion is valid in the ball

For , the solid harmonics with negative powers of (the irregular solid harmonics ) are chosen instead. In that case, one needs to expand the solution of known regions in Laurent series (about ), instead of the Taylor series (about ) used above, to match the terms and find series expansion coefficients .

Orbital angular momentum

In quantum mechanics, Laplace's spherical harmonics are understood in terms of the orbital angular momentum [5]

The ħ is conventional in quantum mechanics; it is convenient to work in units in which ħ = 1. The spherical harmonics are eigenfunctions of the square of the orbital angular momentum

Laplace's spherical harmonics are the joint eigenfunctions of the square of the orbital angular momentum and the generator of rotations about the azimuthal axis:

These operators commute, and are densely defined self-adjoint operators on the weighted Hilbert space of functions f square-integrable with respect to the normal distribution as the weight function on R3:

Furthermore, L2 is a positive operator.

If Y is a joint eigenfunction of L2 and Lz, then by definition

for some real numbers m and λ. Here m must in fact be an integer, for Y must be periodic in the coordinate φ with period a number that evenly divides 2π. Furthermore, since

and each of Lx, Ly, Lz are self-adjoint, it follows that λm2.

Denote this joint eigenspace by Eλ,m, and define the raising and lowering operators by

Then L+ and L commute with L2, and the Lie algebra generated by L+, L, Lz is the special linear Lie algebra of order 2, , with commutation relations

Thus L+ : Eλ,mEλ,m+1 (it is a "raising operator") and L : Eλ,mEλ,m−1 (it is a "lowering operator"). In particular, Lk
+
 : Eλ,mEλ,m+k
must be zero for k sufficiently large, because the inequality λm2 must hold in each of the nontrivial joint eigenspaces. Let YEλ,m be a nonzero joint eigenfunction, and let k be the least integer such that

Then, since

it follows that

Thus λ = ( + 1) for the positive integer = m + k.

The foregoing has been all worked out in the spherical coordinate representation, but may be expressed more abstractly in the complete, orthonormal spherical ket basis.

Harmonic polynomial representation

The spherical harmonics can be expressed as the restriction to the unit sphere of certain polynomial functions . Specifically, we say that a (complex-valued) polynomial function is homogeneous of degree if

for all real numbers and all . We say that is harmonic if

where is the Laplacian. Then for each , we define

For example, when , is just the 3-dimensional space of all linear functions , since any such function is automatically harmonic. Meanwhile, when , we have a 5-dimensional space:

For any , the space of spherical harmonics of degree is just the space of restrictions to the sphere of the elements of . [6] As suggested in the introduction, this perspective is presumably the origin of the term “spherical harmonic” (i.e., the restriction to the sphere of a harmonic function).

For example, for any the formula

defines a homogeneous polynomial of degree with domain and codomain , which happens to be independent of . This polynomial is easily seen to be harmonic. If we write in spherical coordinates and then restrict to , we obtain

which can be rewritten as

After using the formula for the associated Legendre polynomial , we may recognize this as the formula for the spherical harmonic [7] (See the section below on special cases of the spherical harmonics.)

Conventions

Orthogonality and normalization

Several different normalizations are in common use for the Laplace spherical harmonic functions . Throughout the section, we use the standard convention that for (see associated Legendre polynomials)

which is the natural normalization given by Rodrigues' formula.

Plot of the spherical harmonic
Y
l
m
(
th
,
ph
)
{\displaystyle Y_{\ell }^{m}(\theta ,\varphi )}
with
l
=
2
{\displaystyle \ell =2}
and
m
=
1
{\displaystyle m=1}
and
ph
=
p
{\displaystyle \varphi =\pi }
in the complex plane from
-
2
-
2
i
{\displaystyle -2-2i}
to
2
+
2
i
{\displaystyle 2+2i}
with colors created with Mathematica 13.1 function ComplexPlot3D Plot of the spherical harmonic Y l^m(theta,phi) with n=2 and m=1 and phi=pi in the complex plane from -2-2i to 2+2i with colors created with Mathematica 13.1 function ComplexPlot3D.svg
Plot of the spherical harmonic with and and in the complex plane from to with colors created with Mathematica 13.1 function ComplexPlot3D

In acoustics, [8] the Laplace spherical harmonics are generally defined as (this is the convention used in this article)

while in quantum mechanics: [9] [10]

where are associated Legendre polynomials without the Condon–Shortley phase (to avoid counting the phase twice).

In both definitions, the spherical harmonics are orthonormal

where δij is the Kronecker delta and dΩ = sin(θ) . This normalization is used in quantum mechanics because it ensures that probability is normalized, i.e.,

The disciplines of geodesy [11] and spectral analysis use

which possess unit power

The magnetics [11] community, in contrast, uses Schmidt semi-normalized harmonics

which have the normalization

In quantum mechanics this normalization is sometimes used as well, and is named Racah's normalization after Giulio Racah.

It can be shown that all of the above normalized spherical harmonic functions satisfy

where the superscript * denotes complex conjugation. Alternatively, this equation follows from the relation of the spherical harmonic functions with the Wigner D-matrix.

Condon–Shortley phase

One source of confusion with the definition of the spherical harmonic functions concerns a phase factor of , commonly referred to as the Condon–Shortley phase in the quantum mechanical literature. In the quantum mechanics community, it is common practice to either include this phase factor in the definition of the associated Legendre polynomials, or to append it to the definition of the spherical harmonic functions. There is no requirement to use the Condon–Shortley phase in the definition of the spherical harmonic functions, but including it can simplify some quantum mechanical operations, especially the application of raising and lowering operators. The geodesy [12] and magnetics communities never include the Condon–Shortley phase factor in their definitions of the spherical harmonic functions nor in the ones of the associated Legendre polynomials. [13]

Real form

A real basis of spherical harmonics can be defined in terms of their complex analogues by setting

The Condon–Shortley phase convention is used here for consistency. The corresponding inverse equations defining the complex spherical harmonics in terms of the real spherical harmonics are

The real spherical harmonics are sometimes known as tesseral spherical harmonics. [14] These functions have the same orthonormality properties as the complex ones above. The real spherical harmonics with m > 0 are said to be of cosine type, and those with m < 0 of sine type. The reason for this can be seen by writing the functions in terms of the Legendre polynomials as

The same sine and cosine factors can be also seen in the following subsection that deals with the Cartesian representation.

See here for a list of real spherical harmonics up to and including , which can be seen to be consistent with the output of the equations above.

Use in quantum chemistry

As is known from the analytic solutions for the hydrogen atom, the eigenfunctions of the angular part of the wave function are spherical harmonics. However, the solutions of the non-relativistic Schrödinger equation without magnetic terms can be made real. This is why the real forms are extensively used in basis functions for quantum chemistry, as the programs don't then need to use complex algebra. Here, it is important to note that the real functions span the same space as the complex ones would.

For example, as can be seen from the table of spherical harmonics, the usual p functions () are complex and mix axis directions, but the real versions are essentially just x, y, and z.

Spherical harmonics in Cartesian form

The complex spherical harmonics give rise to the solid harmonics by extending from to all of as a homogeneous function of degree , i.e. setting

It turns out that is basis of the space of harmonic and homogeneous polynomials of degree . More specifically, it is the (unique up to normalization) Gelfand-Tsetlin-basis of this representation of the rotational group and an explicit formula for in cartesian coordinates can be derived from that fact.

The Herglotz generating function

If the quantum mechanical convention is adopted for the , then

Here, is the vector with components , , and

is a vector with complex coordinates:

The essential property of is that it is null:

It suffices to take and as real parameters. In naming this generating function after Herglotz, we follow Courant & Hilbert 1962 , §VII.7, who credit unpublished notes by him for its discovery.

Essentially all the properties of the spherical harmonics can be derived from this generating function. [15] An immediate benefit of this definition is that if the vector is replaced by the quantum mechanical spin vector operator , such that is the operator analogue of the solid harmonic , [16] one obtains a generating function for a standardized set of spherical tensor operators, :

The parallelism of the two definitions ensures that the 's transform under rotations (see below) in the same way as the 's, which in turn guarantees that they are spherical tensor operators, , with and , obeying all the properties of such operators, such as the Clebsch-Gordan composition theorem, and the Wigner-Eckart theorem. They are, moreover, a standardized set with a fixed scale or normalization.

Separated Cartesian form

The Herglotzian definition yields polynomials which may, if one wishes, be further factorized into a polynomial of and another of and , as follows (Condon–Shortley phase):

and for m = 0:

Here

and

For this reduces to

The factor is essentially the associated Legendre polynomial , and the factors are essentially .

Examples

Using the expressions for , , and listed explicitly above we obtain:

It may be verified that this agrees with the function listed here and here.

Real forms

Using the equations above to form the real spherical harmonics, it is seen that for only the terms (cosines) are included, and for only the terms (sines) are included:

and for m = 0:

Special cases and values

  1. When , the spherical harmonics reduce to the ordinary Legendre polynomials:
  2. When ,
    or more simply in Cartesian coordinates,
  3. At the north pole, where , and is undefined, all spherical harmonics except those with vanish:

Symmetry properties

The spherical harmonics have deep and consequential properties under the operations of spatial inversion (parity) and rotation.

Parity

The spherical harmonics have definite parity. That is, they are either even or odd with respect to inversion about the origin. Inversion is represented by the operator . Then, as can be seen in many ways (perhaps most simply from the Herglotz generating function), with being a unit vector,

In terms of the spherical angles, parity transforms a point with coordinates to . The statement of the parity of spherical harmonics is then

(This can be seen as follows: The associated Legendre polynomials gives (−1)+m and from the exponential function we have (−1)m, giving together for the spherical harmonics a parity of (−1).)

Parity continues to hold for real spherical harmonics, and for spherical harmonics in higher dimensions: applying a point reflection to a spherical harmonic of degree changes the sign by a factor of (−1).

Rotations

The rotation of a real spherical function with m = 0 and l = 3. The coefficients are not equal to the Wigner D-matrices, since real functions are shown, but can be obtained by re-decomposing the complex functions Rotation of octupole vector function.svg
The rotation of a real spherical function with m = 0 and = 3. The coefficients are not equal to the Wigner D-matrices, since real functions are shown, but can be obtained by re-decomposing the complex functions

Consider a rotation about the origin that sends the unit vector to . Under this operation, a spherical harmonic of degree and order transforms into a linear combination of spherical harmonics of the same degree. That is,

where is a matrix of order that depends on the rotation . However, this is not the standard way of expressing this property. In the standard way one writes,

where is the complex conjugate of an element of the Wigner D-matrix. In particular when is a rotation of the azimuth we get the identity,

The rotational behavior of the spherical harmonics is perhaps their quintessential feature from the viewpoint of group theory. The 's of degree provide a basis set of functions for the irreducible representation of the group SO(3) of dimension . Many facts about spherical harmonics (such as the addition theorem) that are proved laboriously using the methods of analysis acquire simpler proofs and deeper significance using the methods of symmetry.

Spherical harmonics expansion

The Laplace spherical harmonics form a complete set of orthonormal functions and thus form an orthonormal basis of the Hilbert space of square-integrable functions . On the unit sphere , any square-integrable function can thus be expanded as a linear combination of these:

This expansion holds in the sense of mean-square convergence — convergence in L2 of the sphere — which is to say that

The expansion coefficients are the analogs of Fourier coefficients, and can be obtained by multiplying the above equation by the complex conjugate of a spherical harmonic, integrating over the solid angle Ω, and utilizing the above orthogonality relationships. This is justified rigorously by basic Hilbert space theory. For the case of orthonormalized harmonics, this gives:

If the coefficients decay in sufficiently rapidly — for instance, exponentially — then the series also converges uniformly to f.

A square-integrable function can also be expanded in terms of the real harmonics above as a sum

The convergence of the series holds again in the same sense, namely the real spherical harmonics form a complete set of orthonormal functions and thus form an orthonormal basis of the Hilbert space of square-integrable functions . The benefit of the expansion in terms of the real harmonic functions is that for real functions the expansion coefficients are guaranteed to be real, whereas their coefficients in their expansion in terms of the (considering them as functions ) do not have that property.

Spectrum analysis

Power spectrum in signal processing

The total power of a function f is defined in the signal processing literature as the integral of the function squared, divided by the area of its domain. Using the orthonormality properties of the real unit-power spherical harmonic functions, it is straightforward to verify that the total power of a function defined on the unit sphere is related to its spectral coefficients by a generalization of Parseval's theorem (here, the theorem is stated for Schmidt semi-normalized harmonics, the relationship is slightly different for orthonormal harmonics):

where

is defined as the angular power spectrum (for Schmidt semi-normalized harmonics). In a similar manner, one can define the cross-power of two functions as

where

is defined as the cross-power spectrum. If the functions f and g have a zero mean (i.e., the spectral coefficients f00 and g00 are zero), then Sff() and Sfg() represent the contributions to the function's variance and covariance for degree , respectively. It is common that the (cross-)power spectrum is well approximated by a power law of the form

When β = 0, the spectrum is "white" as each degree possesses equal power. When β < 0, the spectrum is termed "red" as there is more power at the low degrees with long wavelengths than higher degrees. Finally, when β > 0, the spectrum is termed "blue". The condition on the order of growth of Sff() is related to the order of differentiability of f in the next section.

Differentiability properties

One can also understand the differentiability properties of the original function f in terms of the asymptotics of Sff(). In particular, if Sff() decays faster than any rational function of as → ∞, then f is infinitely differentiable. If, furthermore, Sff() decays exponentially, then f is actually real analytic on the sphere.

The general technique is to use the theory of Sobolev spaces. Statements relating the growth of the Sff() to differentiability are then similar to analogous results on the growth of the coefficients of Fourier series. Specifically, if

then f is in the Sobolev space Hs(S2). In particular, the Sobolev embedding theorem implies that f is infinitely differentiable provided that

for all s.

Algebraic properties

Addition theorem

A mathematical result of considerable interest and use is called the addition theorem for spherical harmonics. Given two vectors r and r′, with spherical coordinates and , respectively, the angle between them is given by the relation

in which the role of the trigonometric functions appearing on the right-hand side is played by the spherical harmonics and that of the left-hand side is played by the Legendre polynomials.

The addition theorem states [17]

where P is the Legendre polynomial of degree . This expression is valid for both real and complex harmonics. [18] The result can be proven analytically, using the properties of the Poisson kernel in the unit ball, or geometrically by applying a rotation to the vector y so that it points along the z-axis, and then directly calculating the right-hand side. [19]

In particular, when x = y, this gives Unsöld's theorem [20]

which generalizes the identity cos2θ + sin2θ = 1 to two dimensions.

In the expansion ( 1 ), the left-hand side is a constant multiple of the degree zonal spherical harmonic. From this perspective, one has the following generalization to higher dimensions. Let Yj be an arbitrary orthonormal basis of the space H of degree spherical harmonics on the n-sphere. Then , the degree zonal harmonic corresponding to the unit vector x, decomposes as [21]

Furthermore, the zonal harmonic is given as a constant multiple of the appropriate Gegenbauer polynomial:

Combining ( 2 ) and ( 3 ) gives ( 1 ) in dimension n = 2 when x and y are represented in spherical coordinates. Finally, evaluating at x = y gives the functional identity

where ωn−1 is the volume of the (n−1)-sphere.

Contraction rule

Another useful identity expresses the product of two spherical harmonics as a sum over spherical harmonics [22]

Many of the terms in this sum are trivially zero. The values of and that result in non-zero terms in this sum are determined by the selection rules for the 3j-symbols.

Clebsch–Gordan coefficients

The Clebsch–Gordan coefficients are the coefficients appearing in the expansion of the product of two spherical harmonics in terms of spherical harmonics themselves. A variety of techniques are available for doing essentially the same calculation, including the Wigner 3-jm symbol, the Racah coefficients, and the Slater integrals. Abstractly, the Clebsch–Gordan coefficients express the tensor product of two irreducible representations of the rotation group as a sum of irreducible representations: suitably normalized, the coefficients are then the multiplicities.

Visualization of the spherical harmonics

Schematic representation of
Y
l
m
{\displaystyle Y_{\ell m}}
on the unit sphere and its nodal lines.
R
[
Y
l
m
]
{\displaystyle \Re [Y_{\ell m}]}
is equal to 0 along m great circles passing through the poles, and along l-m circles of equal latitude. The function changes sign each time it crosses one of these lines. Spherical harmonics positive negative.svg
Schematic representation of on the unit sphere and its nodal lines. is equal to 0 along m great circles passing through the poles, and along m circles of equal latitude. The function changes sign each time it crosses one of these lines.
3D color plot of the spherical harmonics of degree n = 5. Note that n = l. Spherical harmonics.png
3D color plot of the spherical harmonics of degree n = 5. Note that n = .

The Laplace spherical harmonics can be visualized by considering their "nodal lines", that is, the set of points on the sphere where , or alternatively where . Nodal lines of are composed of circles: there are |m| circles along longitudes and −|m| circles along latitudes. One can determine the number of nodal lines of each type by counting the number of zeros of in the and directions respectively. Considering as a function of , the real and imaginary components of the associated Legendre polynomials each possess −|m| zeros, each giving rise to a nodal 'line of latitude'. On the other hand, considering as a function of , the trigonometric sin and cos functions possess 2|m| zeros, each of which gives rise to a nodal 'line of longitude'.

When the spherical harmonic order m is zero (upper-left in the figure), the spherical harmonic functions do not depend upon longitude, and are referred to as zonal . Such spherical harmonics are a special case of zonal spherical functions. When = |m| (bottom-right in the figure), there are no zero crossings in latitude, and the functions are referred to as sectoral. For the other cases, the functions checker the sphere, and they are referred to as tesseral.

More general spherical harmonics of degree are not necessarily those of the Laplace basis , and their nodal sets can be of a fairly general kind. [23]

List of spherical harmonics

Analytic expressions for the first few orthonormalized Laplace spherical harmonics that use the Condon–Shortley phase convention:

Higher dimensions

The classical spherical harmonics are defined as complex-valued functions on the unit sphere inside three-dimensional Euclidean space . Spherical harmonics can be generalized to higher-dimensional Euclidean space as follows, leading to functions . [24] Let P denote the space of complex-valued homogeneous polynomials of degree in n real variables, here considered as functions . That is, a polynomial p is in P provided that for any real , one has

Let A denote the subspace of P consisting of all harmonic polynomials:

These are the (regular) solid spherical harmonics. Let H denote the space of functions on the unit sphere

obtained by restriction from A

The following properties hold:

An orthogonal basis of spherical harmonics in higher dimensions can be constructed inductively by the method of separation of variables, by solving the Sturm-Liouville problem for the spherical Laplacian

where φ is the axial coordinate in a spherical coordinate system on Sn−1. The end result of such a procedure is [26]

where the indices satisfy |1|2 ≤ ⋯ ≤ n−1 and the eigenvalue is n−1(n−1 + n−2). The functions in the product are defined in terms of the Legendre function

Connection with representation theory

The space H of spherical harmonics of degree is a representation of the symmetry group of rotations around a point (SO(3)) and its double-cover SU(2). Indeed, rotations act on the two-dimensional sphere, and thus also on H by function composition

for ψ a spherical harmonic and ρ a rotation. The representation H is an irreducible representation of SO(3). [27]

The elements of H arise as the restrictions to the sphere of elements of A: harmonic polynomials homogeneous of degree on three-dimensional Euclidean space R3. By polarization of ψA, there are coefficients symmetric on the indices, uniquely determined by the requirement

The condition that ψ be harmonic is equivalent to the assertion that the tensor must be trace free on every pair of indices. Thus as an irreducible representation of SO(3), H is isomorphic to the space of traceless symmetric tensors of degree .

More generally, the analogous statements hold in higher dimensions: the space H of spherical harmonics on the n-sphere is the irreducible representation of SO(n+1) corresponding to the traceless symmetric -tensors. However, whereas every irreducible tensor representation of SO(2) and SO(3) is of this kind, the special orthogonal groups in higher dimensions have additional irreducible representations that do not arise in this manner.

The special orthogonal groups have additional spin representations that are not tensor representations, and are typically not spherical harmonics. An exception are the spin representation of SO(3): strictly speaking these are representations of the double cover SU(2) of SO(3). In turn, SU(2) is identified with the group of unit quaternions, and so coincides with the 3-sphere. The spaces of spherical harmonics on the 3-sphere are certain spin representations of SO(3), with respect to the action by quaternionic multiplication.

Connection with hemispherical harmonics

Spherical harmonics can be separated into two set of functions. [28] One is hemispherical functions (HSH), orthogonal and complete on hemisphere. Another is complementary hemispherical harmonics (CHSH).

Generalizations

The angle-preserving symmetries of the two-sphere are described by the group of Möbius transformations PSL(2,C). With respect to this group, the sphere is equivalent to the usual Riemann sphere. The group PSL(2,C) is isomorphic to the (proper) Lorentz group, and its action on the two-sphere agrees with the action of the Lorentz group on the celestial sphere in Minkowski space. The analog of the spherical harmonics for the Lorentz group is given by the hypergeometric series; furthermore, the spherical harmonics can be re-expressed in terms of the hypergeometric series, as SO(3) = PSU(2) is a subgroup of PSL(2,C).

More generally, hypergeometric series can be generalized to describe the symmetries of any symmetric space; in particular, hypergeometric series can be developed for any Lie group. [29] [30] [31] [32]

See also

Notes

  1. A historical account of various approaches to spherical harmonics in three dimensions can be found in Chapter IV of MacRobert 1967. The term "Laplace spherical harmonics" is in common use; see Courant & Hilbert 1962 and Meijer & Bauer 2004.
  2. The approach to spherical harmonics taken here is found in ( Courant & Hilbert 1962 , §V.8, §VII.5).
  3. Physical applications often take the solution that vanishes at infinity, making A = 0. This does not affect the angular portion of the spherical harmonics.
  4. Weisstein, Eric W. "Spherical Harmonic". mathworld.wolfram.com. Retrieved 2023-05-10.
  5. Edmonds 1957 , §2.5
  6. Hall 2013 Section 17.6
  7. Hall 2013 Lemma 17.16
  8. Williams, Earl G. (1999). Fourier acoustics : sound radiation and nearfield acoustical holography. San Diego, Calif.: Academic Press. ISBN   0080506909. OCLC   181010993.
  9. Messiah, Albert (1999). Quantum mechanics : two volumes bound as one (Two vol. bound as one, unabridged reprint ed.). Mineola, NY: Dover. ISBN   9780486409245.
  10. Claude Cohen-Tannoudji; Bernard Diu; Franck Laloë (1996). Quantum mechanics. Translated by Susan Reid Hemley; et al. Wiley-Interscience: Wiley. ISBN   9780471569527.
  11. 1 2 Blakely, Richard (1995). Potential theory in gravity and magnetic applications . Cambridge England New York: Cambridge University Press. p.  113. ISBN   978-0521415088.
  12. Heiskanen and Moritz, Physical Geodesy, 1967, eq. 1-62
  13. Weisstein, Eric W. "Condon-Shortley Phase". mathworld.wolfram.com. Retrieved 2022-11-02.
  14. Whittaker & Watson 1927 , p. 392.
  15. See, e.g., Appendix A of Garg, A., Classical Electrodynamics in a Nutshell (Princeton University Press, 2012).
  16. Li, Feifei; Braun, Carol; Garg, Anupam (2013), "The Weyl-Wigner-Moyal Formalism for Spin", Europhysics Letters, 102 (6): 60006, arXiv: 1210.4075 , Bibcode:2013EL....10260006L, doi:10.1209/0295-5075/102/60006, S2CID   119610178
  17. Edmonds, A. R. (1996). Angular Momentum In Quantum Mechanics . Princeton University Press. p.  63.
  18. This is valid for any orthonormal basis of spherical harmonics of degree . For unit power harmonics it is necessary to remove the factor of 4π.
  19. Whittaker & Watson 1927 , p. 395
  20. Unsöld 1927
  21. Stein & Weiss 1971 , §IV.2
  22. Brink, D. M.; Satchler, G. R. Angular Momentum. Oxford University Press. p. 146.
  23. Eremenko, Jakobson & Nadirashvili 2007
  24. Solomentsev 2001; Stein & Weiss 1971 , §Iv.2
  25. Cf. Corollary 1.8 of Axler, Sheldon; Ramey, Wade (1995), Harmonic Polynomials and Dirichlet-Type Problems
  26. Higuchi, Atsushi (1987). "Symmetric tensor spherical harmonics on the N-sphere and their application to the de Sitter group SO(N,1)". Journal of Mathematical Physics. 28 (7): 1553–1566. Bibcode:1987JMP....28.1553H. doi:10.1063/1.527513.
  27. Hall 2013 Corollary 17.17
  28. Zheng Y, Wei K, Liang B, Li Y, Chu X (2019-12-23). "Zernike like functions on spherical cap: principle and applications in optical surface fitting and graphics rendering". Optics Express. 27 (26): 37180–37195. Bibcode:2019OExpr..2737180Z. doi: 10.1364/OE.27.037180 . ISSN   1094-4087. PMID   31878503.
  29. N. Vilenkin, Special Functions and the Theory of Group Representations, Am. Math. Soc. Transl., vol. 22, (1968).
  30. J. D. Talman, Special Functions, A Group Theoretic Approach, (based on lectures by E.P. Wigner), W. A. Benjamin, New York (1968).
  31. W. Miller, Symmetry and Separation of Variables, Addison-Wesley, Reading (1977).
  32. A. Wawrzyńczyk, Group Representations and Special Functions, Polish Scientific Publishers. Warszawa (1984).

Related Research Articles

<span class="mw-page-title-main">Polar coordinate system</span> Coordinates determined by distance and angle

In mathematics, the polar coordinate system is a two-dimensional coordinate system in which each point on a plane is determined by a distance from a reference point and an angle from a reference direction. The reference point is called the pole, and the ray from the pole in the reference direction is the polar axis. The distance from the pole is called the radial coordinate, radial distance or simply radius, and the angle is called the angular coordinate, polar angle, or azimuth. Angles in polar notation are generally expressed in either degrees or radians.

<span class="mw-page-title-main">Spherical coordinate system</span> 3-dimensional coordinate system

In mathematics, a spherical coordinate system is a coordinate system for three-dimensional space where the position of a given point in space is specified by three numbers, : the radial distance of the radial liner connecting the point to the fixed point of origin ; the polar angle θ of the radial line r; and the azimuthal angle φ of the radial line r.

<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

<i>n</i>-sphere Generalized sphere of dimension n (mathematics)

In mathematics, an n-sphere or hypersphere is an n-dimensional generalization of the 1-dimensional circle and 2-dimensional sphere to any non-negative integer n. The n-sphere is the setting for n-dimensional spherical geometry.

<span class="mw-page-title-main">Ellipsoid</span> Quadric surface that looks like a deformed sphere

An ellipsoid is a surface that can be obtained from a sphere by deforming it by means of directional scalings, or more generally, of an affine transformation.

<span class="mw-page-title-main">Unit vector</span> Vector of length one

In mathematics, a unit vector in a normed vector space is a vector of length 1. A unit vector is often denoted by a lowercase letter with a circumflex, or "hat", as in .

In mechanics and geometry, the 3D rotation group, often denoted O(3), is the group of all rotations about the origin of three-dimensional Euclidean space under the operation of composition.

In probability theory, the Borel–Kolmogorov paradox is a paradox relating to conditional probability with respect to an event of probability zero. It is named after Émile Borel and Andrey Kolmogorov.

In rotordynamics, the rigid rotor is a mechanical model of rotating systems. An arbitrary rigid rotor is a 3-dimensional rigid object, such as a top. To orient such an object in space requires three angles, known as Euler angles. A special rigid rotor is the linear rotor requiring only two angles to describe, for example of a diatomic molecule. More general molecules are 3-dimensional, such as water, ammonia, or methane.

A multipole expansion is a mathematical series representing a function that depends on angles—usually the two angles used in the spherical coordinate system for three-dimensional Euclidean space, . Similarly to Taylor series, multipole expansions are useful because oftentimes only the first few terms are needed to provide a good approximation of the original function. The function being expanded may be real- or complex-valued and is defined either on , or less often on for some other .

In classical mechanics, the shell theorem gives gravitational simplifications that can be applied to objects inside or outside a spherically symmetrical body. This theorem has particular application to astronomy.

This is a table of orthonormalized spherical harmonics that employ the Condon-Shortley phase up to degree . Some of these formulas are expressed in terms of the Cartesian expansion of the spherical harmonics into polynomials in x, y, z, and r. For purposes of this table, it is useful to express the usual spherical to Cartesian transformations that relate these Cartesian components to and as

<span class="mw-page-title-main">Multiple integral</span> Generalization of definite integrals to functions of multiple variables

In mathematics (specifically multivariable calculus), a multiple integral is a definite integral of a function of several real variables, for instance, f(x, y) or f(x, y, z). Physical (natural philosophy) interpretation: S any surface, V any volume, etc.. Incl. variable to time, position, etc.

In physics, spherical multipole moments are the coefficients in a series expansion of a potential that varies inversely with the distance R to a source, i.e., as  Examples of such potentials are the electric potential, the magnetic potential and the gravitational potential.

In special functions, a topic in mathematics, spin-weighted spherical harmonics are generalizations of the standard spherical harmonics and—like the usual spherical harmonics—are functions on the sphere. Unlike ordinary spherical harmonics, the spin-weighted harmonics are U(1) gauge fields rather than scalar fields: mathematically, they take values in a complex line bundle. The spin-weighted harmonics are organized by degree l, just like ordinary spherical harmonics, but have an additional spin weights that reflects the additional U(1) symmetry. A special basis of harmonics can be derived from the Laplace spherical harmonics Ylm, and are typically denoted by sYlm, where l and m are the usual parameters familiar from the standard Laplace spherical harmonics. In this special basis, the spin-weighted spherical harmonics appear as actual functions, because the choice of a polar axis fixes the U(1) gauge ambiguity. The spin-weighted spherical harmonics can be obtained from the standard spherical harmonics by application of spin raising and lowering operators. In particular, the spin-weighted spherical harmonics of spin weight s = 0 are simply the standard spherical harmonics:

In physics, the Green's function for the Laplacian in three variables is used to describe the response of a particular type of physical system to a point source. In particular, this Green's function arises in systems that can be described by Poisson's equation, a partial differential equation (PDE) of the form

In physics, the Laplace expansion of potentials that are directly proportional to the inverse of the distance, such as Newton's gravitational potential or Coulomb's electrostatic potential, expresses them in terms of the spherical Legendre polynomials. In quantum mechanical calculations on atoms the expansion is used in the evaluation of integrals of the inter-electronic repulsion.

In physics and mathematics, the solid harmonics are solutions of the Laplace equation in spherical polar coordinates, assumed to be (smooth) functions . There are two kinds: the regular solid harmonics, which are well-defined at the origin and the irregular solid harmonics, which are singular at the origin. Both sets of functions play an important role in potential theory, and are obtained by rescaling spherical harmonics appropriately:

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.

References

Cited references

General references