Coulomb wave function

Irregular Coulomb wave function G plotted from 0 to 20 with repulsive and attractive interactions in Mathematica 13.1
Irregular Coulomb wave function G plotted from 0 to 20 with repulsive and attractive interactions in Mathematica 13.1
image of complex plot of regular Coulomb wave function added

In mathematics, a Coulomb wave function is a solution of the Coulomb wave equation, named after Charles-Augustin de Coulomb. They are used to describe the behavior of charged particles in a Coulomb potential and can be written in terms of confluent hypergeometric functions or Whittaker functions of imaginary argument.

Coulomb wave equation

The Coulomb wave equation for a single charged particle of mass m {\displaystyle m} is the Schrödinger equation with Coulomb potential[1]

( 2 2 2 m + Z c α r ) ψ k ( r ) = 2 k 2 2 m ψ k ( r ) , {\displaystyle \left(-\hbar ^{2}{\frac {\nabla ^{2}}{2m}}+{\frac {Z\hbar c\alpha }{r}}\right)\psi _{\vec {k}}({\vec {r}})={\frac {\hbar ^{2}k^{2}}{2m}}\psi _{\vec {k}}({\vec {r}})\,,}

where Z = Z 1 Z 2 {\displaystyle Z=Z_{1}Z_{2}} is the product of the charges of the particle and of the field source (in units of the elementary charge, Z = 1 {\displaystyle Z=-1} for the hydrogen atom), α {\displaystyle \alpha } is the fine-structure constant, and 2 k 2 / ( 2 m ) {\displaystyle \hbar ^{2}k^{2}/(2m)} is the energy of the particle. The solution, which is the Coulomb wave function, can be found by solving this equation in parabolic coordinates

ξ = r + r k ^ , ζ = r r k ^ ( k ^ = k / k ) . {\displaystyle \xi =r+{\vec {r}}\cdot {\hat {k}},\quad \zeta =r-{\vec {r}}\cdot {\hat {k}}\qquad ({\hat {k}}={\vec {k}}/k)\,.}

Depending on the boundary conditions chosen, the solution has different forms. Two of the solutions are[2][3]

ψ k ( ± ) ( r ) = Γ ( 1 ± i η ) e π η / 2 e i k r M ( i η , 1 , ± i k r i k r ) , {\displaystyle \psi _{\vec {k}}^{(\pm )}({\vec {r}})=\Gamma (1\pm i\eta )e^{-\pi \eta /2}e^{i{\vec {k}}\cdot {\vec {r}}}M(\mp i\eta ,1,\pm ikr-i{\vec {k}}\cdot {\vec {r}})\,,}

where M ( a , b , z ) 1 F 1 ( a ; b ; z ) {\displaystyle M(a,b,z)\equiv {}_{1}\!F_{1}(a;b;z)} is the confluent hypergeometric function, η = Z m c α / ( k ) {\displaystyle \eta =Zmc\alpha /(\hbar k)} and Γ ( z ) {\displaystyle \Gamma (z)} is the gamma function. The two boundary conditions used here are

ψ k ( ± ) ( r ) e i k r ( k r ± ) , {\displaystyle \psi _{\vec {k}}^{(\pm )}({\vec {r}})\rightarrow e^{i{\vec {k}}\cdot {\vec {r}}}\qquad ({\vec {k}}\cdot {\vec {r}}\rightarrow \pm \infty )\,,}

which correspond to k {\displaystyle {\vec {k}}} -oriented plane-wave asymptotic states before or after its approach of the field source at the origin, respectively. The functions ψ k ( ± ) {\displaystyle \psi _{\vec {k}}^{(\pm )}} are related to each other by the formula

ψ k ( + ) = ψ k ( ) . {\displaystyle \psi _{\vec {k}}^{(+)}=\psi _{-{\vec {k}}}^{(-)*}\,.}

Partial wave expansion

The wave function ψ k ( r ) {\displaystyle \psi _{\vec {k}}({\vec {r}})} can be expanded into partial waves (i.e. with respect to the angular basis) to obtain angle-independent radial functions w ( η , ρ ) {\displaystyle w_{\ell }(\eta ,\rho )} . Here ρ = k r {\displaystyle \rho =kr} .

ψ k ( r ) = 4 π r = 0 m = i w ( η , ρ ) Y m ( r ^ ) Y m ( k ^ ) . {\displaystyle \psi _{\vec {k}}({\vec {r}})={\frac {4\pi }{r}}\sum _{\ell =0}^{\infty }\sum _{m=-\ell }^{\ell }i^{\ell }w_{\ell }(\eta ,\rho )Y_{\ell }^{m}({\hat {r}})Y_{\ell }^{m\ast }({\hat {k}})\,.}

A single term of the expansion can be isolated by the scalar product with a specific spherical harmonic

ψ k m ( r ) = ψ k ( r ) Y m ( k ^ ) d k ^ = R k ( r ) Y m ( r ^ ) , R k ( r ) = 4 π i w ( η , ρ ) / r . {\displaystyle \psi _{k\ell m}({\vec {r}})=\int \psi _{\vec {k}}({\vec {r}})Y_{\ell }^{m}({\hat {k}})d{\hat {k}}=R_{k\ell }(r)Y_{\ell }^{m}({\hat {r}}),\qquad R_{k\ell }(r)=4\pi i^{\ell }w_{\ell }(\eta ,\rho )/r.}

The equation for single partial wave w ( η , ρ ) {\displaystyle w_{\ell }(\eta ,\rho )} can be obtained by rewriting the laplacian in the Coulomb wave equation in spherical coordinates and projecting the equation on a specific spherical harmonic Y m ( r ^ ) {\displaystyle Y_{\ell }^{m}({\hat {r}})}

d 2 w d ρ 2 + ( 1 2 η ρ ( + 1 ) ρ 2 ) w = 0 . {\displaystyle {\frac {d^{2}w_{\ell }}{d\rho ^{2}}}+\left(1-{\frac {2\eta }{\rho }}-{\frac {\ell (\ell +1)}{\rho ^{2}}}\right)w_{\ell }=0\,.}

The solutions are also called Coulomb (partial) wave functions or spherical Coulomb functions. Putting z = 2 i ρ {\displaystyle z=-2i\rho } changes the Coulomb wave equation into the Whittaker equation, so Coulomb wave functions can be expressed in terms of Whittaker functions with imaginary arguments M i η , + 1 / 2 ( 2 i ρ ) {\displaystyle M_{-i\eta ,\ell +1/2}(-2i\rho )} and W i η , + 1 / 2 ( 2 i ρ ) {\displaystyle W_{-i\eta ,\ell +1/2}(-2i\rho )} . The latter can be expressed in terms of the confluent hypergeometric functions M {\displaystyle M} and U {\displaystyle U} . For Z {\displaystyle \ell \in \mathbb {Z} } , one defines the special solutions [4]

H ( ± ) ( η , ρ ) = 2 i ( 2 ) e π η / 2 e ± i σ ρ + 1 e ± i ρ U ( + 1 ± i η , 2 + 2 , 2 i ρ ) , {\displaystyle H_{\ell }^{(\pm )}(\eta ,\rho )=\mp 2i(-2)^{\ell }e^{\pi \eta /2}e^{\pm i\sigma _{\ell }}\rho ^{\ell +1}e^{\pm i\rho }U(\ell +1\pm i\eta ,2\ell +2,\mp 2i\rho )\,,}

where

σ = arg Γ ( + 1 + i η ) {\displaystyle \sigma _{\ell }=\arg \Gamma (\ell +1+i\eta )}

is called the Coulomb phase shift. One also defines the real functions

F ( η , ρ ) = 1 2 i ( H ( + ) ( η , ρ ) H ( ) ( η , ρ ) ) , {\displaystyle F_{\ell }(\eta ,\rho )={\frac {1}{2i}}\left(H_{\ell }^{(+)}(\eta ,\rho )-H_{\ell }^{(-)}(\eta ,\rho )\right)\,,}
Regular Coulomb wave function F plotted from 0 to 20 with repulsive and attractive interactions in Mathematica 13.1
Regular Coulomb wave function F plotted from 0 to 20 with repulsive and attractive interactions in Mathematica 13.1
G ( η , ρ ) = 1 2 ( H ( + ) ( η , ρ ) + H ( ) ( η , ρ ) ) . {\displaystyle G_{\ell }(\eta ,\rho )={\frac {1}{2}}\left(H_{\ell }^{(+)}(\eta ,\rho )+H_{\ell }^{(-)}(\eta ,\rho )\right)\,.}

In particular one has

F ( η , ρ ) = 2 e π η / 2 | Γ ( + 1 + i η ) | ( 2 + 1 ) ! ρ + 1 e i ρ M ( + 1 + i η , 2 + 2 , 2 i ρ ) . {\displaystyle F_{\ell }(\eta ,\rho )={\frac {2^{\ell }e^{-\pi \eta /2}|\Gamma (\ell +1+i\eta )|}{(2\ell +1)!}}\rho ^{\ell +1}e^{i\rho }M(\ell +1+i\eta ,2\ell +2,-2i\rho )\,.}

The asymptotic behavior of the spherical Coulomb functions H ( ± ) ( η , ρ ) {\displaystyle H_{\ell }^{(\pm )}(\eta ,\rho )} , F ( η , ρ ) {\displaystyle F_{\ell }(\eta ,\rho )} , and G ( η , ρ ) {\displaystyle G_{\ell }(\eta ,\rho )} at large ρ {\displaystyle \rho } is

H ( ± ) ( η , ρ ) e ± i θ ( ρ ) , {\displaystyle H_{\ell }^{(\pm )}(\eta ,\rho )\sim e^{\pm i\theta _{\ell }(\rho )}\,,}
F ( η , ρ ) sin θ ( ρ ) , {\displaystyle F_{\ell }(\eta ,\rho )\sim \sin \theta _{\ell }(\rho )\,,}
G ( η , ρ ) cos θ ( ρ ) , {\displaystyle G_{\ell }(\eta ,\rho )\sim \cos \theta _{\ell }(\rho )\,,}

where

θ ( ρ ) = ρ η log ( 2 ρ ) 1 2 π + σ . {\displaystyle \theta _{\ell }(\rho )=\rho -\eta \log(2\rho )-{\frac {1}{2}}\ell \pi +\sigma _{\ell }\,.}

The solutions H ( ± ) ( η , ρ ) {\displaystyle H_{\ell }^{(\pm )}(\eta ,\rho )} correspond to incoming and outgoing spherical waves. The solutions F ( η , ρ ) {\displaystyle F_{\ell }(\eta ,\rho )} and G ( η , ρ ) {\displaystyle G_{\ell }(\eta ,\rho )} are real and are called the regular and irregular Coulomb wave functions. In particular one has the following partial wave expansion for the wave function ψ k ( + ) ( r ) {\displaystyle \psi _{\vec {k}}^{(+)}({\vec {r}})} [5]

ψ k ( + ) ( r ) = 4 π ρ = 0 m = i e i σ F ( η , ρ ) Y m ( r ^ ) Y m ( k ^ ) , {\displaystyle \psi _{\vec {k}}^{(+)}({\vec {r}})={\frac {4\pi }{\rho }}\sum _{\ell =0}^{\infty }\sum _{m=-\ell }^{\ell }i^{\ell }e^{i\sigma _{\ell }}F_{\ell }(\eta ,\rho )Y_{\ell }^{m}({\hat {r}})Y_{\ell }^{m\ast }({\hat {k}})\,,}

Properties of the Coulomb function

The radial parts for a given angular momentum are orthonormal. When normalized on the wave number scale (k-scale), the continuum radial wave functions satisfy [6][7]

0 R k ( r ) R k ( r ) r 2 d r = δ ( k k ) {\displaystyle \int _{0}^{\infty }R_{k\ell }^{\ast }(r)R_{k'\ell }(r)r^{2}dr=\delta (k-k')}

Other common normalizations of continuum wave functions are on the reduced wave number scale ( k / 2 π {\displaystyle k/2\pi } -scale),

0 R k ( r ) R k ( r ) r 2 d r = 2 π δ ( k k ) , {\displaystyle \int _{0}^{\infty }R_{k\ell }^{\ast }(r)R_{k'\ell }(r)r^{2}dr=2\pi \delta (k-k')\,,}

and on the energy scale

0 R E ( r ) R E ( r ) r 2 d r = δ ( E E ) . {\displaystyle \int _{0}^{\infty }R_{E\ell }^{\ast }(r)R_{E'\ell }(r)r^{2}dr=\delta (E-E')\,.}

The radial wave functions defined in the previous section are normalized to

0 R k ( r ) R k ( r ) r 2 d r = ( 2 π ) 3 k 2 δ ( k k ) {\displaystyle \int _{0}^{\infty }R_{k\ell }^{\ast }(r)R_{k'\ell }(r)r^{2}dr={\frac {(2\pi )^{3}}{k^{2}}}\delta (k-k')}

as a consequence of the normalization

ψ k ( r ) ψ k ( r ) d 3 r = ( 2 π ) 3 δ ( k k ) . {\displaystyle \int \psi _{\vec {k}}^{\ast }({\vec {r}})\psi _{{\vec {k}}'}({\vec {r}})d^{3}r=(2\pi )^{3}\delta ({\vec {k}}-{\vec {k}}')\,.}

The continuum (or scattering) Coulomb wave functions are also orthogonal to all Coulomb bound states[8]

0 R k ( r ) R n ( r ) r 2 d r = 0 {\displaystyle \int _{0}^{\infty }R_{k\ell }^{\ast }(r)R_{n\ell }(r)r^{2}dr=0}

due to being eigenstates of the same hermitian operator (the hamiltonian) with different eigenvalues.

Further reading

  • Bateman, Harry (1953), Higher transcendental functions (PDF), vol. 1, McGraw-Hill, archived from the original (PDF) on 2011-08-11, retrieved 2011-07-30.
  • Jaeger, J. C.; Hulme, H. R. (1935), "The Internal Conversion of γ -Rays with the Production of Electrons and Positrons", Proceedings of the Royal Society of London. Series A, Mathematical and Physical Sciences, 148 (865): 708–728, Bibcode:1935RSPSA.148..708J, doi:10.1098/rspa.1935.0043, ISSN 0080-4630, JSTOR 96298
  • Slater, Lucy Joan (1960), Confluent hypergeometric functions, Cambridge University Press, MR 0107026.

References

  1. ^ Hill, Robert N. (2006), Drake, Gordon (ed.), Handbook of atomic, molecular and optical physics, Springer New York, pp. 153–155, doi:10.1007/978-0-387-26308-3, ISBN 978-0-387-20802-2
  2. ^ Landau, L. D.; Lifshitz, E. M. (1977), Course of theoretical physics III: Quantum mechanics, Non-relativistic theory (3rd ed.), Pergamon Press, p. 569
  3. ^ Messiah, Albert (1961), Quantum mechanics, North Holland Publ. Co., p. 485
  4. ^ Gaspard, David (2018), "Connection formulas between Coulomb wave functions", J. Math. Phys., 59 (11): 112104, arXiv:1804.10976, doi:10.1063/1.5054368
  5. ^ Messiah, Albert (1961), Quantum mechanics, North Holland Publ. Co., p. 426
  6. ^ Formánek, Jiří (2004), Introduction to quantum theory I (in Czech) (2nd ed.), Prague: Academia, pp. 128–130
  7. ^ Landau, L. D.; Lifshitz, E. M. (1977), Course of theoretical physics III: Quantum mechanics, Non-relativistic theory (3rd ed.), Pergamon Press, p. 121
  8. ^ Landau, L. D.; Lifshitz, E. M. (1977), Course of theoretical physics III: Quantum mechanics, Non-relativistic theory (3rd ed.), Pergamon Press, pp. 668–669