Jump to content

User:Phenomenam/sandbox

From Wikipedia, the free encyclopedia

Introduction

[edit]

In order to capture the interference phenomena that occur during a scattering process that the classical theory fails to address properly, yet to avoid a fully quantum treatment that is complicated, a semiclassical approach may be employed (see Quantum Scattering Theory for a fully quantum treatment and Classical Scattering Theory for a classical treatment of a scattering process). Semiclassical scattering theory makes use of trajectories, impact parameters, and the like that are classical and the superposition of scattering waves that are quantum to approximate or model a scattering phenomenon fairly accurately under the condition where each classical trajectories carry unique phases.


Semiclassical Physics

[edit]

Definition

[edit]

Semiclassical physics is "a vast discipline of theoretical physics that attempts to compute spectra and wave functions of quantum (wave) systems on the basis of their classical trajectories (rays)" with the addition of quantum superposition principle and Planck's constant . [1] (see also Semiclassical physics)

Usefulness

[edit]

In general, semiclassical approach is said to be useful since very few analytically solvable quantum systems are known to date. Usually, semiclassical methods are incorporated as they allow us to compute spectra, i.e. energy levels, and wave functions of quantum systems based on classical trajectories even in the absence of the exact quantum approach. Note that the method still involves quantum phases and the quantum superposition principle.[1]

Background

[edit]

This section largely follows section 5.2 of the reference [1].

In order to obtain approximate solutions to a quantum system with potential , we assume stationary wave function

where and are chosen such that the wave function satisfies Schrödinger equation. Inserting the form into Schrödinger equation and solving for the equation generally, we obtain a "classical" (absence of ) equation

and a "quantum" equation

Solving for using the classical equation,

where is an arbitrary location in the classically allowed region, is the momentum , and is the classical action. The importance of the result lies in the fact that the quantum phase angle shift of the wave function is now directly related to the classical action , which serves as the backbone of semiclassical treatments.

Semiclassical Approximation

[edit]

Among many semiclassical approximations (see Semiclassical physics), we focus on Bohr-Sommerfeld Quantization and JWKB Approximation that are most useful in dealing with scattering problems.

Bohr-Sommerfeld Quantization

[edit]

Dating from the early 20th century, Bohr-Sommerfeld Quantization procedure [2] [3] forms the basis of the "old quantum mechanics''. The approximation taken in this procedure is to assume the amplitude to be constant, namely,

Computing the accumulated phase angle for the wave that takes off from the starting position, say , then returns to its initial position, we obtain

Since the wave function is stationary and unique,

where is a positive integer.

JWKB Approximation

[edit]

In the limit is small, which is justified by the fact that we are in semiclassical realm, the "quantum" equation (see Background section above) may be approximated to be

which has a solution of the form

where is a complex number. With found in Background section, the wave function in the classically allowed region may be written as

and the wave function in the classically forbidden region may be written as

where , , and are complex numbers. At this point, we notice that the obtained wave functions are not valid at the classical turning point due to their non-zero momenta. Using linear approximation of the shape of at the turning point, with the help of Airy functions and the boundary conditions on and (see WKB approximation, [4] [5] [6] [7] , or [1] for details), we obtain the wave function in the classically allowed region to be

where denotes the location of the classical turning point. For the sake of completeness, we note in passing that in the case where there are two classical turning points, say and , which would not be of interest for scattering purpose, one may show

as an analogous expression to the classical action found in Bohr-Sommerfeld Quantization. In the case where there is only one classical turning point , which is of interest for investigating a scattering system, a more involved approach is needed (see below).

Applications to scattering theory

[edit]

Applying the semiclassical approximation techniques to a scattering system is relatively straightforward. The theoretical treatment of a scattering system states (i) mass reduced mass , (ii) potential effective potential , and (iii) starting position far limit . Applying these changes to Bohr-Sommerfeld Quantization, we have a situation where the phase of the wave function oscilates on a nearly constant background provided by the potential, with the fictitious particle with mass coming in from , rebounding from the inner wall of the effective potential , and returning to . During this process, the particle or the wave function, then, picks up an extra phase angle whose 1-D analytical computation is shown above. When we apply the same changes to JWKB approximation, in analogy to the 1-D case shown above, we obtain the 3-D wave function at far limit to be

where

Rewriting the far limit wave function as

and comparing it with the form of the far limit of the exact quantum solution

where is a spherical Bessel function of the first kind, together with the large approximation of

we see that the accumulated or extra phase picked up via scattering in the limit is given by [8]

where we used

and

The usefulness of the semiclassical expression above is apparent given the wide use of in finding the following scattering related quantities. The list includes but is not limited to the total cross section

the scattering length

etc.


Connection to Classical and Quantum Scattering

[edit]

Previously, we have shown that the classical action in semiclassical physics corresponds to the integral of the momentum of a given system over displacement, which, in 3-D, is written as

where is the impact parameter. Dividing the coordinates into the radial and the angular parts, namely,

and using the fact that the difference between the classical action in the presence of the potential and the classical action in the absence of the potential is the scattering phase shift when integrated along a closed path, we write in general, with the deflection angle and the angular momentum of the system,

where

which has the name "classical phase", and

We note that, with

where is the asymptotic relative velocity of the two bodies participating in a scattering process, and

the first derivative of becomes [9]

so that

Applying the principle of least action, the action derivative, then, must read zero, meaning , hence it is the classically allowed path. Said differently, the waves that are on the paths near the classically allowed one constructively interfere with one another whereas those that are not interfere destructively.

Comparing the two momenta expressions above and in the previous section, we obtain

where

with defined in the previous section. The expression bears an importance as it connects the specific impact parameters with corresponding angular momentum quantum numbers, which may be used in the following ways.

Recalling the quantum mechanical expression for the total cross section above, we may now rewrite as

In order to see the difference that semiclassical methods make, suppose we have a hard sphere (radius ) scattering, for instance. The evaluation of the integral results in , which is twice the classical value. The existence of additional contribution to , then, demonstrates the ability of semiclassical method capturing the effect of diffraction of the incident plane wave around the target.

This time, we take a look at the JWKB phase shift . Instead of comparing the semiclassical wave function with the form of the exact quantum solution as in previous section, we write directly;

In the limit of large angular momentum, since the contribution from the potential would be small, the turning point would approximately be . Combining the two integrals with the common factor and using Taylor expansion up to first order in , one can show

which was obtained by Massey and Mohr in 1934 [10]. The expression may then be used to compute the total cross section given the long-range potential of the form , which is known as the Massey-Mohr formula.

Finally, consider the partial wave expansion of the (axially-symmetric) scattering amplitude, the amplitude of the scattered wave function whose radial dependence is ,

where is the extra phase picked up by the presence of potential in the scattering process associated with the angular quantum number and is the Legendre polynomials. Assuming those terms in the sum with are small, replacing in the sum with an approximate formula that holds for

results in

where we first replaced and with the correspondent and expressions respectively (see above) then used integral approximation on and

Now relaxing the positivity condition of , with , we obtain [11]

where we have for the repulsive trajectory with and for the attractive trajectory with . We note that we have replaced the argument of with since is a function of and . Incorporating the stationary phase or the steepest descent method, justified by the fact that the action must not change rapidly in if the integral were not to be 0 in consistency with the least action principle, now evaluates to

where we used

and denotes the stationary point. Obviously, if there are more than one , the different paths will cause interference phenomena. Solving for the spacing of the maxima given two different 's, namely and , we obtain

which is nothing but the spacing of interference maxima in Young's double-slit experiment, for which ; multiple amplitudes will lead to multiple sets of interference minima and maxima. In analogy to the double-slit experiment, we conclude that the greater the spacing between the 's are, the closer the spacing of the fringes, a.k.a. diffraction oscillations.

Limitations

[edit]

Semiclassical approximations taken here do have their own limitations. Among many other possible limitations, the notable one includes the breakdown of expression at , the rainbow angle. This is so, because the stationary phase approximation does not hold at the rainbow angle as the second derivative of action disappears. This failure of the primitive semiclassical approximation may be saved by

where is the rainbow impact parameter. Because now the integral that needs to be done includes an exponential in a cubic power of , the final form of the integral is an Airy function. [8]

References

[edit]
  1. ^ a b c d R. Bl\"umel (2011). Advanced Quantum Mechanics: The Classical-Quantum Connection. Jones & Bartlett.
  2. ^ Wilson, W. (1915). "LXXXIII. The quantum-theory of radiation and line spectra". Phil. Mag. 29 (174): 795–802. doi:10.1080/14786440608635362.
  3. ^ Sommerfeld, A. (1916). "Zur Quantentheorie der Spektrallinien". Ann. Der Physik. 51 (17): 1–94. doi:10.1002/andp.19163561702.
  4. ^ Jeffreys, H. (1924). Proc. Lond. Math. Soc. 23: 428. {{cite journal}}: Missing or empty |title= (help)
  5. ^ Wentzel, G. (1926). "Eine Verallgemeinerung der Quantenbedingungen f�r die Zwecke der Wellenmechanik". Z. Phys. 38 (6–7): 518–529. doi:10.1007/BF01397171. S2CID 120096571. {{cite journal}}: replacement character in |title= at position 48 (help)
  6. ^ Kramers, H. A. (1926). "Wellenmechanik und halbzahlige Quantisierung". Z. Phys. 39 (10–11): 828–840. doi:10.1007/BF01451751. S2CID 122955156.
  7. ^ Brillouin, L. (1926). C. R. Acad. Sci. 183: 24. {{cite journal}}: Missing or empty |title= (help)
  8. ^ a b Ford, K. W. (1959). "Semiclassical description of scattering". Ann. Phys. 7 (3): 259–286. doi:10.1016/0003-4916(59)90026-0. {{cite journal}}: Unknown parameter |coauthors= ignored (|author= suggested) (help)
  9. ^ Smith, F. T. (1965). "Classical and Quantal Scattering. I. The Classical Action". J. Chem. Phys. 65 (7): 2419–2426. doi:10.1063/1.1696310. hdl:2060/19660023885.
  10. ^ Massey, H. S. W. (1934). Proc. Roy. Soc. A. 144: 188. {{cite journal}}: Missing or empty |title= (help); Unknown parameter |coauthors= ignored (|author= suggested) (help)
  11. ^ Pritchard, D. E. (1970). "Interpretation of Interference Structure in Elastic Scattering Using the Semiclassical Action". Phys. Rev. A. 1 (4): 1120–1124. doi:10.1103/PhysRevA.1.1120.