Solid harmonics – Wikipedia

before-content-x4

In physics and mathematics, the solid harmonics are solutions of the Laplace equation in spherical polar coordinates, assumed to be (smooth) functions

R3C{displaystyle mathbb {R} ^{3}to mathbb {C} }
after-content-x4

. There are two kinds: the regular solid harmonics

Rm(r){displaystyle R_{ell }^{m}(mathbf {r} )}

, which are well-defined at the origin and the irregular solid harmonics

Im(r){displaystyle I_{ell }^{m}(mathbf {r} )}

, 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:

Derivation, relation to spherical harmonics[edit]

Introducing r, θ, and φ for the spherical polar coordinates of the 3-vector r, and assuming that

Φ{displaystyle Phi }

is a (smooth) function

after-content-x4
R3C{displaystyle mathbb {R} ^{3}to mathbb {C} }

, we can write the Laplace equation in the following form

where l2 is the square of the nondimensional angular momentum operator,

It is known that spherical harmonics Ym
are eigenfunctions of l2:

Substitution of Φ(r) = F(r) Ym
into the Laplace equation gives, after dividing out the spherical harmonic function, the following radial equation and its general solution,

The particular solutions of the total Laplace equation are regular solid harmonics:

and irregular solid harmonics:

The regular solid harmonics correspond to harmonic homogeneous polynomials, i.e. homogeneous polynomials which are solutions to Laplace’s equation.

Racah’s normalization[edit]

Racah’s normalization (also known as Schmidt’s semi-normalization) is applied to both functions

(and analogously for the irregular solid harmonic) instead of normalization to unity. This is convenient because in many applications the Racah normalization factor appears unchanged throughout the derivations.

Addition theorems[edit]

The translation of the regular solid harmonic gives a finite expansion,

where the Clebsch–Gordan coefficient is given by

The similar expansion for irregular solid harmonics gives an infinite series,

with

|r||a|{displaystyle |r|leq |a|,}

. The quantity between pointed brackets is again a Clebsch-Gordan coefficient,

The addition theorems were proved in different manners by several authors.[1][2]

Complex form[edit]

The regular solid harmonics are homogeneous, polynomial solutions to the Laplace equation

ΔR=0{displaystyle Delta R=0}

. Separating the indeterminate

z{displaystyle z}

and writing

R=apa(x,y)za{textstyle R=sum _{a}p_{a}(x,y)z^{a}}

, the Laplace equation is easily seen to be equivalent to the recursion formula

so that any choice of polynomials

p0(x,y){displaystyle p_{0}(x,y)}

of degree

{displaystyle ell }

and

p1(x,y){displaystyle p_{1}(x,y)}

of degree

1{displaystyle ell -1}

gives a solution to the equation. One particular basis of the space of homogeneous polynomials (in two variables) of degree

k{displaystyle k}

is

{(x2+y2)m(x±iy)k2m0mk/2}{displaystyle left{(x^{2}+y^{2})^{m}(xpm iy)^{k-2m}mid 0leq mleq k/2right}}

. Note that it is the (unique up to normalization) basis of eigenvectors of the rotation group

SO(2){displaystyle SO(2)}

: The rotation

ρα{displaystyle rho _{alpha }}

of the plane by

α[0,2π]{displaystyle alpha in [0,2pi ]}

acts as multiplication by

e±i(k2m)α{displaystyle e^{pm i(k-2m)alpha }}

on the basis vector

(x2+y2)m(x+iy)k2m{displaystyle (x^{2}+y^{2})^{m}(x+iy)^{k-2m}}

.

If we combine the degree

{displaystyle ell }

basis and the degree

1{displaystyle ell -1}

basis with the recursion formula, we obtain a basis of the space of harmonic, homogeneous polynomials (in three variables this time) of degree

{displaystyle ell }

consisting of eigenvectors for

SO(2){displaystyle SO(2)}

(note that the recursion formula is compatible with the

SO(2){displaystyle SO(2)}

-action because the Laplace operator is rotationally invariant). These are the complex solid harmonics:

and in general

for

0m{displaystyle 0leq mleq ell }

.

Plugging in spherical coordinates

x=rcos(θ)sin(φ){displaystyle x=rcos(theta )sin(varphi )}

,

y=rsin(θ)sin(φ){displaystyle y=rsin(theta )sin(varphi )}

,

z=rcos(φ){displaystyle z=rcos(varphi )}

and using

x2+y2=r2sin(φ)2=r2(1cos(φ)2){displaystyle x^{2}+y^{2}=r^{2}sin(varphi )^{2}=r^{2}(1-cos(varphi )^{2})}

one finds the usual relationship to spherical harmonics

Rm=reimϕPm(cos(ϑ)){displaystyle R_{ell }^{m}=r^{ell }e^{imphi }P_{ell }^{m}(cos(vartheta ))}

with a polynomial

Pm{displaystyle P_{ell }^{m}}

, which is (up to normalization) the associated Legendre polynomial, and so

Rm=rYm(θ,φ){displaystyle R_{ell }^{m}=r^{ell }Y_{ell }^{m}(theta ,varphi )}

(again, up to the specific choice of normalization).

Real form[edit]

By a simple linear combination of solid harmonics of ±m these functions are transformed into real functions, i.e. functions

R3R{displaystyle mathbb {R} ^{3}to mathbb {R} }

. The real regular solid harmonics, expressed in Cartesian coordinates, are real-valued homogeneous polynomials of order

{displaystyle ell }

in x, y, z. The explicit form of these polynomials is of some importance. They appear, for example, in the form of spherical atomic orbitals and real multipole moments. The explicit Cartesian expression of the real regular harmonics will now be derived.

Linear combination[edit]

We write in agreement with the earlier definition

with

where

P(cosθ){displaystyle P_{ell }(cos theta )}

is a Legendre polynomial of order .
The m dependent phase is known as the Condon–Shortley phase.

The following expression defines the real regular solid harmonics:

m = 0:

Since the transformation is by a unitary matrix the normalization of the real and the complex solid harmonics is the same.

z-dependent part[edit]

Upon writing u = cos θ the m-th derivative of the Legendre polynomial can be written as the following expansion in u

with

Since z = r cos θ it follows that this derivative, times an appropriate power of r, is a simple polynomial in z,

(x,y)-dependent part[edit]

Consider next, recalling that x = r sin θ cos φ and y = r sin θ sin φ,

Likewise

Further

and

In total[edit]

List of lowest functions[edit]

We list explicitly the lowest functions up to and including = 5.
Here

Π¯m(z)[(2δm0)(m)!(+m)!]1/2Πm(z).{displaystyle {bar {Pi }}_{ell }^{m}(z)equiv left[{tfrac {(2-delta _{m0})(ell -m)!}{(ell +m)!}}right]^{1/2}Pi _{ell }^{m}(z).}

The lowest functions

Am(x,y){displaystyle A_{m}(x,y),}

and

Bm(x,y){displaystyle B_{m}(x,y),}

are:

m Am Bm
0
1
2
3
4
5

References[edit]

  1. ^ R. J. A. Tough and A. J. Stone, J. Phys. A: Math. Gen. Vol. 10, p. 1261 (1977)
  2. ^ M. J. Caola, J. Phys. A: Math. Gen. Vol. 11, p. L23 (1978)
  • Steinborn, E. O.; Ruedenberg, K. (1973). “Rotation and Translation of Regular and Irregular Solid Spherical Harmonics”. In Lowdin, Per-Olov (ed.). Advances in quantum chemistry. Vol. 7. Academic Press. pp. 1–82. ISBN 9780080582320.
  • Thompson, William J. (2004). Angular momentum: an illustrated guide to rotational symmetries for physical systems. Weinheim: Wiley-VCH. pp. 143–148. ISBN 9783527617838.

after-content-x4