Log in

Is the Aharonov–Bohm phase shift for a non-closed path a measurable quantity?

  • Regular Article
  • Published:
The European Physical Journal Plus Aims and scope Submit manuscript

Abstract

There recently appear some interesting attempts to explain the AB-effect through the interaction between the charged particle and the solenoid current mediated by the exchange of a virtual photon. A vital assumption of this approach is that AB-phase shift is proportional to the change of the interaction energy between the charged particle and solenoid along the path of the moving charge. Accordingly, they insist that the AB-phase change along a path does not depend on the gauge choice so that the AB-phase shift for a non-closed path is in principle measurable. We however notice the existence of two fairly different discussions on the interaction energy between the solenoid and a charge particle, the one is due to Boyer and the other is due to Saldanha and others. In the present paper, based on a self-contained quantum mechanical treatment of the combined system of a solenoid, a charged particle, and the quantized electromagnetic fields, we show that both interaction energies of Boyer and of Saldanha are in fact gauge invariant at least for non-singular gauge transformations but they are destined to cancel each other. Our analysis rather shows that the origin of the AB-phase can be traced back to other part of our effective Hamiltonian. Furthermore, based on the path-integral formalism with our effective Lagrangian, we explicitly demonstrate that the AB-phase shift for a non-closed path is not a gauge-variant quantity, which means that it would not correspond to direct experimental observables.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Subscribe and save

Springer+ Basic
EUR 32.99 /Month
  • Get 10 units per month
  • Download Article/Chapter or Ebook
  • 1 Unit = 1 Article or 1 Chapter
  • Cancel anytime
Subscribe now

Buy Now

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Similar content being viewed by others

Data Availability Statement

No Data associated in the manuscript.

Notes

  1. Nowadays, it is widely accepted that there is an unavoidable nonlocality in quantum mechanics as exemplified by the famous Einstein-Podolsky-Rosen paradox [16] and Bell’s inequality [17]. However, it is also believed that this type of nonlocality has little to do with the above-mentioned nonlocality in the interpretation of the AB effect.

  2. The deformation here is possible, under the nonrelativistic treatment of the charged particle moving with a constant velocity. In fact, in the general case, the velocity vector depends on the retarded time \(v = v (t_r)\) and extracting it from volume integrations is not possible, as it depends on the field point \(\varvec{x}\) through \(t_r = t - \vert \varvec{x} - \varvec{x}^\prime \vert / c\).

  3. There is some delicacy concerning the convergence of the integral given by Eq. (9). Strictly speaking, this integral diverges for an infinitely-long solenoid. On the other hand, for a very long but finite-length solenoid as assumed by Boyer, the magnetic field outside the solenoid does not exactly vanish and the form of \(\varvec{A}^{(S)}\) appearing in Boyer’s paper is justified only approximately. Since the discussion of this problem is a little involved, it is forwarded to an Appendix (A).

  4. The superscript (S) of \(\varvec{A}^{(S)} (\varvec{x})\) has double meanings. On the one hand, it means the Special solution of the above Poisson equation (B9). On the other hand, it means that the vector potential \(\varvec{A}^{(S)} (\varvec{x})\) is uniquely given as a definite convolution integral of the source current \(\varvec{j} (\varvec{x})\).

  5. We point out that, even restricting to regular gauge transformations, there is some special physical problem in which the transverse component of the vector potential cannot be uniquely determined. This happens, for example, in the famous Landau problem that handles the motion of an electron in a magnetic field uniformly spreading over the whole 2-dimensional plane. In such special circumstances, there exist plural forms of vector potential satisfying the transverse condition. In fact, one can easily confirm that the vector potentials with the three typical gauge choices, i.e. the symmetric gauge and the two Landau gauges, all satisfy the transverse condition [39, 40].

References

  1. W. Ehrenberg, R.E. Siday, The refractive index in electron optics and the principles of dynamics. Proc. Phys. Soc. London B62, 8 (1949)

    Article  ADS  Google Scholar 

  2. Y. Aharonov, D. Bohm, Significance of electromagnetic potentials in quantum theory. Phys. Rev. 115, 485 (1959)

    Article  MathSciNet  ADS  Google Scholar 

  3. M. Peshkin, The Aharonov–Bohm effect: why it cannot be eliminated from quantum mechanics. Phys. Rep. 80, 375 (1981)

    Article  MathSciNet  ADS  Google Scholar 

  4. S. Olariu, I.I. Popescu, The quantum effects of electromagnetic fluxes. Rev. Mod. Phys. 57, 339 (1985)

    Article  ADS  Google Scholar 

  5. M. Peshkin, A. Tonomura, The Aharonov–Bohm effect. Lecture Phys. 340, 1 (1989)

    Article  MathSciNet  Google Scholar 

  6. M. Wakamatsu, Y. Kitadono, L. Zou, P.-M. Zhang, The role of electron orbital angular momentum in the Aharonov–Bohm effect revisited. Ann. Phys. 397, 259 (2018)

    Article  MathSciNet  CAS  ADS  Google Scholar 

  7. R. P. Feynman, R. B. Leighton, M. Sands, Feynman lectures on physics, Vol. II, (Addison-Wesley Pub. Co., c1963-1965), Ch.15.4

  8. E.J. Konopinski, What the electromagnetic potential describes? Am. J. Phys. bf 46, 499 (1978)

    Article  MathSciNet  ADS  Google Scholar 

  9. M.D. Semon, J.R. Taylor, Thoughts on the magnetic vector potential. Am. J. Phys. 64, 1361 (1996)

    Article  ADS  Google Scholar 

  10. A. Tonomura, N. Osakabe, T. Matsuda, T. Kawasaki, J. Endo, Evidence for Aharonov–Bohm effect with magnetic field completely shielded from electron wave. Phy. Rev. Lett. 56, 792 (1986)

    Article  CAS  ADS  Google Scholar 

  11. N. Osakabe, T. Matsuda, T. Kawasaki, J. Endo, A. Tonomura, S. Yano, H. Yamada, Experimental confirmation of Aharonov–Bohm effect using a toroidal magnet field confined by a superconductor. Phys. Rev. A 34, 815 (1986)

    Article  CAS  ADS  Google Scholar 

  12. R. Healey, Nonlocality and the Aharonov–Bohm Effect. Philos. Sci. 64, 18 (1997)

    Article  MathSciNet  Google Scholar 

  13. Y. Aharonov, E. Cohen, D. Rhorlich, Nonlocality in the Aharonov–Bohm effect. Phys. Rev. A 93, 42110 (2016)

    Article  ADS  Google Scholar 

  14. J.A. Heras, R. Heras, Topology, nonlocality and duality in classical electrodynamics. Eur. Phys. J. Plus 137, 157 (2022)

    Article  Google Scholar 

  15. R. Heras, The Aharonov–Bohm effect in a closed line. Eur. Phys. J. Plus 137, 641 (2022)

    Article  Google Scholar 

  16. A. Einstein, B. Podolsky, N. Rosen, Can quantum-mechanical description of physical reality be considered complete? Phys. Rev. 48, 696 (1935)

    Article  Google Scholar 

  17. J.S. Bell, On the Einstein Podolsky Rosen paradox. Physics 1, 195 (1964)

    Article  MathSciNet  Google Scholar 

  18. T. Adachi, T. Inagaki, M. Ozaki, K. Sasabe, The Vector Potential Revisited, Electrical Engineering in Japan, 114(6), 1993. Translated from Denki Gakkai Ronbunnshi A112, 763 (1992)

    Google Scholar 

  19. A.M. Stewart, Vector potential of the Coulomb gauge. Eur. J. Phys. 24, 519 (2003)

    Article  MathSciNet  Google Scholar 

  20. J.-F. Li, Y. Jiang, W.-M. Sun, H.-S. Zong, F. Wang, New application of decomposition of \(U(1)\) gauge potential: Aharonov–Bohm effect and Anderson-Higgs mechanism. Mod. Phys. Lett. B 26, 1250124 (2012)

    Article  ADS  Google Scholar 

  21. T.H. Boyer, Classical electromagnetic interaction of a charged particle with a constant-current solenoid. Phys. Rev. 8, 1667 (1971)

    Google Scholar 

  22. L. Vaidman, Role of potentials in the Aharonov–Bohm effect. Phys. Rev. A 86, 040101(R) (2012)

    Article  ADS  Google Scholar 

  23. K. Kang, Proposal for locality test of the Aharonov–Bohm effect via Andreev interferometer without a loop. Phys. Soc. 71, 565 (2017)

    Google Scholar 

  24. C. Marletto, V. Vedral, Aharonov–Bohm phase is locally generated like all other quantum phases. Phys. Rev. Lett. 125, 040401 (2020)

    Article  MathSciNet  CAS  PubMed  ADS  Google Scholar 

  25. P.L. Saldanha, Local description of the Aharonov–Bohm effect with a quantum electromagnetic field. Found. Phys 51, 6 (2021)

    Article  MathSciNet  ADS  Google Scholar 

  26. E. Santos, I. Gonzalo, Microscopic theory of the Aharonov–Bohm effect. Europhys. Lett. 45, 418 (1999)

    Article  CAS  ADS  Google Scholar 

  27. X. Li, T.H. Hansson, W. Ku, Gauge-independent description of the Aharonov–Bohm effect. Phys. Rev. A 106, 032217 (2022)

    Article  MathSciNet  CAS  ADS  Google Scholar 

  28. D.H. Kobe, Berry phase, Aharonov-Bohm effect and topology. J. Phys. A : Math Gen. 24, 3551 (1991)

    Article  MathSciNet  ADS  Google Scholar 

  29. E. Cohen, H. Larocque, F. Bouchard, F. Nejadsattari, Y. Gefen, E. Karimi, Geometric phase from Aharonov–Bohm to Pancharatnam-Berry and beyond. Nat. Rev. Phys. 1, 437 (2019)

    Article  Google Scholar 

  30. O. Heaviside, The electromagnetic effect of a moving charge. The Electrician 22, 147 (1888)

    Google Scholar 

  31. O.D. Jefimenko, Direct calculation of the electric and magnetic fields of an electric point charge moving with constant velocity. Am. J. Phys. 62, 79 (1994)

    Article  ADS  Google Scholar 

  32. B. Rothenstein, S. Propescu, G. J. Spix, Relativistic derivations of the electric and magnetic fields generated by an electric point charge moving with constant velocity, ar**v:physics/0601028 (2006)

  33. C. Cohen-Tannoudji, J. Dupont-Roc, G. Grynberg, Photons and atoms (Wiley, New York, 1989)

    Google Scholar 

  34. R.P. Feynman, A.R. Hibbs, Quantum mechanics and path integral (McGraw-Hill Companies Inc, NewYork, 1965)

    Google Scholar 

  35. S. Horvat, P.A. Guérin, L. Apadula, F. Del Santo, Probing quantum coherence at a distance and Aharonov–Bohm non-locality. Phys. Rev. A102, 062214 (2020)

    Article  ADS  Google Scholar 

  36. D. Singleton, E.C. Vagenas, The covariant, time-dependent Aharonov–Bohm effect Phys. Lett. B723, 241 (2013)

    ADS  Google Scholar 

  37. A. Shadowitz, The electromagnetic field (Dover Publication Inc, NewYork, 1975)

    Google Scholar 

  38. M. Babiker, R. Lowdon, Gauge invariance of the Aharonov–Bohm effect. J. Phys. A: Math. Gen. 17, 2973 (1984)

    Article  MathSciNet  ADS  Google Scholar 

  39. M. Wakamatsu, Y. Kiadono, P.-M. Zhang, The issue of gauge choice in the Landau problem and the physics of canonical and mechanical orbital angular momenta. Ann. Phys. 392, 287 (2018)

    Article  MathSciNet  CAS  ADS  Google Scholar 

  40. M. Wakamatsu, A. Hayashi, Physical symmetries and gauge choices in the Landau problem. Eur. Phys. J. A 58, 121 (2022)

    Article  CAS  ADS  Google Scholar 

  41. P. Bocchieri, A. Loinger, Nonexistence of the Aharonov–Bohm effect. Nuovo Cimento 47A, 475 (1978)

    Article  MathSciNet  ADS  Google Scholar 

  42. M. Kreizschmar, Must quantal wave functions be single-valued? Zeitschrift Für Physik 185, 73 (1965)

    Article  ADS  Google Scholar 

  43. M. Bawin, A. Burnel, Aharonov–Bohm effect and gauge invariance. J. Phys. A: Math. Gen. 16, 2173 (1983)

    Article  MathSciNet  ADS  Google Scholar 

  44. H. Miyazawa, T. Miyazawa, The physical meaning of the Ehrenberg–Siday–Aharonov–Bohm effect” (in Japanese), http://www.miyazawa1.sakura.ne.jp/papers/esba.pdf

Download references

Acknowledgements

The author of the paper would like to thank Akihisa Hayashi for his valuable advises and enlightening discussions. Without his help, the paper would have never been accomplished. The author would also like to thank the anonymous reviewer for many critical but constructive suggestions, which enhanced the quality of the paper.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Masashi Wakamatsu.

Appendices

Appendix A: On the convergence of the integral given by Eq. (9)

As indicated in the main text, Boyer as well as Saldanha have implicitly assumed the convergence of Eq. (9) for a very long solenoid without formally demonstrating it. We point out that more rigorous argument was already given in the paper by Babiker and Lowdon [38]. They derived the following expression for the vector potential of a finite solenoid of length \(2 \,L\) and a radius of R, as measured in the x-y plane. The expression can be written as

$$\begin{aligned} \varvec{A} \ = \ \frac{2 \,\Phi _{AB} \, \varvec{e}_\phi }{\pi ^2 \,R} \,\int _0^L \, \frac{(2 - k^2) \,\mathbb {K} (k) - 2 \,\mathbb {E} (k)}{k^2 \,\sqrt{(\rho + R)^2 + z^2}} \,d z, \end{aligned}$$
(A1)

where \(\Phi _{AB}\) is the AB-flux corresponding to an infinitely long solenoid, \(k^2 = 4 \,\pi \,R / [(\rho + R)^2 + z^2]\), and \(\mathbb {K} (k)\) and \(\mathbb {E} (k)\) are the complete elliptical integrals of the first and second kind, respectively. Following them, it can be shown that along the regions \(\rho \gg R\) and \(\rho \ll R\), one can approximate Eq. (A1) as

$$\begin{aligned} \varvec{A} = \frac{\Phi _{AB} \,\varvec{e}_\phi }{2 \,\pi \,\rho } \frac{L}{\sqrt{\rho ^2 + L^2}} \ \ (\rho \gg R), \varvec{A} = \frac{\Phi _{AB} \,\rho \,\varvec{e}_\phi }{\pi \,R^2} \frac{L}{\sqrt{R^2 + L^2} } \ \ (\rho \ll R). \end{aligned}$$
(A2)

Clearly, these potentials do not correspond with those in the AB effect since we are dealing with a finite-length solenoid, which is consistent with the fact that outside the solenoid the magnetic field is non-vanishing \(\nabla \times \varvec{A} \ne 0\). Accordingly, there would be a Lorentz force outside the solenoid acting on the charge. However, from Eq. (A2) it follows that such force would be negligible compared with other forces acting on the charge (e.g., a mechanical force) as L increases. In this regard, Eq. (A2) gives a qualitative idea of what is meant by a “very long solenoid”. In fact, in the limit of an infinitely long solenoid (\(L \rightarrow \infty )\) we obtain the expected result

$$\begin{aligned} \lim _{L \rightarrow \infty } \,\varvec{A} = \frac{\Phi _{AB} \,\varvec{e}_\phi }{2 \,\pi \,\rho } \ \ (\rho > R), \lim _{L \rightarrow \infty } \,\varvec{A} = \frac{\Phi _{AB} \,\rho \,\varvec{e}_\phi }{\pi \,R^2} \ \ (\rho < R). \end{aligned}$$
(A3)

On the other hand, when \(L \gg \rho \) and \(L \gg R\), we obtain the approximate expressions

$$\begin{aligned} \varvec{A} \vert _{L \gg \rho } \simeq \frac{\Phi _{AB} \,\varvec{e}_\phi }{2 \,\pi \rho } \ \ (\rho \gg R), \varvec{A} \vert _{L \gg R} \simeq \frac{\Phi _{AB} \,\rho \,\varvec{e}_\phi }{\pi \,R^2} \ \ (\rho \ll R). \end{aligned}$$
(A4)

For example, if \(L = 10^3 \,\rho \) and \(\rho = 10^3 \,R\), then outside the solenoid the vector potential takes the form

$$\begin{aligned} \varvec{A} = 0.99 \times \left( \frac{\Phi _{AB} \,\varvec{e}_\phi }{2 \,\pi \,\rho } \right) \simeq \frac{\Phi _{AB} \,\varvec{e}_\phi }{2 \,\pi \rho }, \end{aligned}$$
(A5)

This clarifies in what sense the expression of the vector potential given in Boyer’s paper is justified.

Appendix B: On gauge-choice independence of Boyer’s interaction energy

Here, we address to the question whether the interaction energy of Boyer has truly a gauge-invariant meaning beyond the Coulomb gauge. To this end, we ask ourselves the following question:

  • What is the most general form of vector potential generated by the steady current of an extremely long solenoid ?

To answer this question, we start with one of the (time-dependent) Maxwell equation,

$$\begin{aligned} \nabla \times \varvec{B} (\varvec{x}, t) \ - \ \frac{\partial \varvec{E} (\varvec{x}, t)}{\partial t} \ = \ \varvec{j} (\varvec{x}, t). \end{aligned}$$
(B6)

If one introduces the electromagnetic potentials \(\phi \) and \(\varvec{A}\) by the familiar relations \(\varvec{B} = \nabla \times \varvec{A}\) and \(\varvec{E} = - \,\nabla \phi - \frac{\partial \varvec{A}}{\partial t}\), the above Maxwell equation becomes

$$\begin{aligned} \nabla \left( \nabla \cdot \varvec{A} (\varvec{x}, t) \ + \, \frac{\partial \phi (\varvec{x}, t)}{\partial t} \right) \ - \ \Delta \varvec{A} (\varvec{x}, t) \ - \ \frac{\partial ^2 \varvec{A} (\varvec{x}, t)}{\partial t^2} \ = \ \varvec{j} (\varvec{x}, t), \end{aligned}$$
(B7)

Since the magnetic field inside the solenoid is thought to be generated by a steady (time-independent) surface current \(\varvec{j} (\varvec{x})\) of the solenoid, we can think that \(\phi \) and \(\varvec{A}\) are both time-independent without loss of generality. Under such circumstances, the above equation reduces to

$$\begin{aligned} \Delta \varvec{A} (\varvec{x}) \ - \,\nabla \,( \nabla \cdot \varvec{A} (\varvec{x})) \ = \ - \,\varvec{j} (\varvec{x}). \end{aligned}$$
(B8)

independently of the gauge condition for \(\varvec{A} (\varvec{x})\). The solution of the above equation can most easily be found if we impose the Coulomb gauge condition \(\nabla \cdot \varvec{A}^{(C)} (\varvec{x}) = 0\) for the vector potential. In fact, under this condition, the above equation reduces to the familiar Poisson equation [37]:

$$\begin{aligned} \Delta \varvec{A}^{(C)} (\varvec{x}) \ = \ - \,\varvec{j} (\varvec{x}), \end{aligned}$$
(B9)

Here, \(\varvec{A}^{(C)} (\varvec{x}) \) stands for the vector potential satisfying the above Poisson equation together with the Coulomb gauge condition. As described in standard textbooks, the general solution to this Poisson equation is given as

$$\begin{aligned} \varvec{A}^{(C)} (\varvec{x}) \ = \ \varvec{A}^{(S)} (\varvec{x}) \ + \ \nabla \chi ^{(H)} (\varvec{x}), \end{aligned}$$
(B10)

where \(\varvec{A}^{(S)} (\varvec{x})\) is given byFootnote 4

$$\begin{aligned} \varvec{A}^{(S)} (\varvec{x}) \ = \ \frac{1}{4 \,\pi } \,\int \,d^3 x^\prime \,\, \frac{\varvec{j} (\varvec{x}^\prime )}{\vert \varvec{x} - \varvec{x}^\prime \vert }. \end{aligned}$$
(B11)

while \(\chi ^{(H)} (\varvec{x})\) is any scalar function satisfying the Laplace equation \(\Delta \chi ^{(H)} (\varvec{x}) = 0\). The last condition \(\Delta \chi ^{(H)} = 0\) follows, because \(\varvec{A}^{(C)}\) must satisfy the Coulomb gauge condition, which dictates that

$$\begin{aligned} 0 \ = \ \nabla \cdot \varvec{A}^{(C)} \ = \ \nabla \cdot \varvec{A}^{(S)} \ + \ \Delta \chi ^{(H)} \ = \ \Delta \chi ^{(H)}. \end{aligned}$$
(B12)

Here, use has been made of the relation \(\nabla \cdot \varvec{A}^{(S)} = 0\), which can be easily verified by using the conservation law \(\nabla \cdot \varvec{j} (\varvec{x}) = 0\) for the steady current. Unfortunately, the above form of the vector potential holds only in the Coulomb gauge. In fact, \(\varvec{A}^{(C)}\) does not satisfy more general field equation (B8) for the vector potential. A question is whether one can find more general solution that satisfies Eq.(B8), which holds in arbitrary gauge. Actually, this is not so difficult. In fact, one can easily confirm that the following form of the vector potential satisfies the equation (B8),

$$\begin{aligned} \varvec{A} (\varvec{x}) \ = \ \varvec{A}^{(S)} (\varvec{x}) \ + \ \nabla \chi (\varvec{x}), \end{aligned}$$
(B13)

where \(\chi (\varvec{x})\) is an arbitrary scalar function not restricted to a Harmonic function. In fact, it holds that

$$\begin{aligned} \Delta \varvec{A} (\varvec{x}) \, - \, \nabla (\nabla \cdot \varvec{A} (\varvec{x}))&= {} \Delta \,(\varvec{A}^{(S)} (\varvec{x}) + \nabla \chi (\varvec{x})) \, - \, \nabla \,(\nabla \cdot \varvec{A}^{(S)} (\varvec{x}) + \Delta \chi (\varvec{x})) \nonumber \\&= {} \Delta \,\varvec{A}^{(S)} (\varvec{x}) \ = \ \Delta \,\frac{1}{4 \,\pi } \, \int \,d^3 x^\prime \,\,\frac{\varvec{j} (\varvec{x}^\prime )}{\vert \varvec{x} - \varvec{x}^\prime \vert } \nonumber \\ &= {} \frac{1}{4 \,\pi } \,\int \,d^3 x^\prime \,\,( - \,4 \,\pi ) \,\, \delta (\varvec{x} - \varvec{x}^\prime ) \,\varvec{j} (\varvec{x}^\prime ) \ = \ - \,\varvec{j} (\varvec{x}). \end{aligned}$$
(B14)

We thus find that the vector potential generated by the steady current of a very long solenoid can always be expressed in the form (B13) independently of the gauge choice. The part \(\varvec{A}^{(S)} (\varvec{x})\) looks clearly gauge invariant since it is expressed as a definite convolution integral of the source current. On the other hand, the part \(\nabla \chi (\varvec{x})\), which carries the gauge degrees of freedom of \(\varvec{A} (\varvec{x})\), is totally arbitrary. Here, a natural guess is that \(\varvec{A}^{(S)} (\varvec{x})\) and \(\nabla \chi (\varvec{x})\) can, respectively, be identified with the transverse part \(\varvec{A}_\perp (\varvec{x})\) and the longitudinal part \(\varvec{A}_\parallel (\varvec{x})\) of the vector potential.Footnote 5

Unfortunately. there still remains a delicate problem in this identification. In fact, since the function \(\chi \) is an arbitrary scalar function, one can in principle consider the following singular or multi-valued gauge function,

$$\begin{aligned} \chi (\varvec{x}) \ = \ - \,\Phi \,\phi \ = \ - \,\Phi \,\arctan \left( \frac{y}{x} \right) , \end{aligned}$$
(B15)

with \(\varvec{x} = (\varvec{\rho }, z) = (\rho , \phi , z)\). As shown below, the rotation of \(\nabla \chi \) does not vanish, i.e. \(\nabla \times (\nabla \chi ) \ne 0\). Rather, it can be shown that \(\nabla \chi \) satisfies the transverse condition, \(\nabla \cdot (\nabla \chi ) = 0\). This means that the transverse-longitudinal decomposition, or more precisely, the identification of the transverse component of the vector potential is not unique, once the multi-valued gauge function as above is allowed.

Still, it seems to us that such a multi-valued gauge transformation is a little unnatural as well as artificial from a physical viewpoint. The reason will be explained below. We already know that, assuming an appropriate limiting procedure, the vector potential defined by

$$\begin{aligned} \varvec{A}^{(S)} (\varvec{x}) \ = \ \int \,\frac{\varvec{J}_{ext} (\varvec{x}^\prime )}{\vert \varvec{x} - \varvec{x}^\prime \vert } \,d^3 x^\prime \end{aligned}$$

with the surface current of the solenoid given by

$$\begin{aligned} \varvec{J}_{ext} (\varvec{x}) \ = \ B \,\delta (\rho - R) \,\varvec{e}_\phi \end{aligned}$$

takes the following axially-symmetric form

$$\begin{aligned} \varvec{A}^{(S)} (\varvec{x}) \ = \ \frac{\Phi }{2 \,\pi } \, \left[ \,\frac{\rho \,\theta (R - \rho )}{R^2} \ - \ \frac{\theta (\rho - R)}{\rho } \,\right] \,\varvec{e}_\phi , \end{aligned}$$

where \(\varvec{x} = (\varvec{\rho }, z) = (\rho , \phi , z)\). Suppose that we define a new vector potential \(\varvec{A}^\prime (\varvec{x})\) through the multi-valued gauge transformation

$$\begin{aligned} \varvec{A}^\prime (\varvec{x}) \ = \ \varvec{A}^{(S)} (\varvec{x}) \ + \ \nabla \chi (\varvec{x}), \end{aligned}$$

with the singular gauge function \(\chi \) given by Eq. (B15). Since \(\partial ^2 \chi / \partial x \,\partial y \ne \partial ^2 \chi / \partial y \,\partial x\) for the singular gauge function \(\chi \), \(\nabla \times (\nabla \chi )\) does not vanish. In fact, we find that

$$\begin{aligned} \nabla \times (\nabla \chi ) \ = \ - \,\Phi \,\delta (x) \,\delta (y) \,\varvec{e}_z \end{aligned}$$

Introducing the notation

$$\begin{aligned} \varvec{A}^\prime (\varvec{x}) \ = \ \varvec{A}^{(S)} (\varvec{x}) \ + \ \nabla \chi (\varvec{x}) \ \equiv \ \varvec{A}^{(S)} (\varvec{x}) \ + \ \varvec{A}_{string} (\varvec{x}), \end{aligned}$$

we therefore find that

$$\begin{aligned} \varvec{B}^\prime (\varvec{x}) &= {} \nabla \times \varvec{A}^\prime (\varvec{x}) \\ &= {} \nabla \times \varvec{A}^{(S)} (\varvec{x}) \ + \ \nabla \times \varvec{A}_{string} (\varvec{x}) \\ &= {} B \,\theta (R - \rho ) \,\varvec{e}_z \ - \ \Phi \,\delta (x) \,\delta (y) \,\varvec{e}_z \\ &\equiv {} \varvec{B}^{(S)} (\varvec{x}) \ + \ \varvec{B}_{string} (\varvec{x}). \end{aligned}$$

Note that the total magnetic flux penetrating the solenoid vanishes, since

$$\begin{aligned} \Phi ^\prime&\equiv {} \int \,\varvec{B}^{(S)} (\varvec{x}) \cdot d \varvec{S} \ + \ \int \,\varvec{B}_{string} (\varvec{x}) \cdot d \varvec{S} \\ &= {} \pi \,R^2 \ - \ \Phi \ = \ 0. \end{aligned}$$

Also widely-known is the fact that the vector potential outside the solenoid can be completely eliminated by the above multi-valued gauge transformation, i.e.

$$\begin{aligned} \varvec{A}^\prime (\varvec{x}) \ = \ 0 \ \ \ \text{ for } \ \ \ \rho > R \end{aligned}$$
(B16)

Based on this last observation, Bocchieri and Loisinger once erroneously claimed that the Aharonov–Bohm effect does not exist [41]. As shown by several researchers, the AB-effect remains to exist even after such a multi-valued gauge transformation, if one properly takes account of the change of the \(2 \,\pi \) periodic boundary condition for the electron wave function [42,43,44].

Although mathematically allowable, however, we point out that very peculiar nature of the above-mentioned multi-valued gauge transformation. To explain it, let us recall the basic definition of the familiar transverse-longitudinal decomposition of the vector field \(\varvec{F}\) is given as

$$\begin{aligned} \varvec{F} \ = \ \varvec{F}_\perp \ + \ \varvec{F}_\parallel . \end{aligned}$$

where the transverse component \(\varvec{F}_\perp \) and the longitudinal component \(\varvec{F}_\parallel \) are, respectively, demanded to satisfy the following transverse condition and the longitudinal condition:

$$\begin{aligned} \nabla \cdot \varvec{F}_\perp \ = \ 0, \ \ \ \nabla \times \varvec{F}_\parallel \ = \ 0. \end{aligned}$$

Note that, since \(\nabla \times \varvec{A}_{string} \ne 0\), the part \(\varvec{A}_{string} = \nabla \chi \) with \(\chi = - \,\Phi \,\arctan (y / x)\) is not categorized into a longitudinal component any more, but it is a transverse component. In fact, one can verify that it satisfies the transverse condition,

$$\begin{aligned} \nabla \cdot \varvec{A}_{string} (\varvec{x}) \ = \ 0. \end{aligned}$$

This means that the transverse part of the vector potential is neither gauge invariant nor unique, if multi-valued gauge transformation is allowed. However, we continue to argue unphysical nature of such multi-valued gauge transformation. Suppose that we calculate rotation of the magnetic field \(\varvec{B}^{(S)} (\varvec{x})\) and \(\varvec{B}_{string} (\varvec{x})\), the sum of which gives the new magnetic field \(\varvec{B}^\prime (\varvec{x}) = \varvec{B}^{(S)} (\varvec{x}) + \varvec{B}_{string} (\varvec{x})\) after singular gauge transformation. We find that

$$\begin{aligned} \nabla \times \varvec{B}^{(S)} (\varvec{x})= & {} B \,\delta (\rho - R) \,\varvec{e}_\phi \ = \ \varvec{J}_{ext} (\varvec{x}), \\ \nabla \times \varvec{B}_{string} (\varvec{x})= & {} \Phi \,\left( \, \delta (x) \,\delta ^\prime (y) \,\varvec{e}_x \ - \ \delta ^\prime (x) \,\delta (y) \,\varvec{e}_y \,\right) \ \equiv \ \varvec{J}_{string} (\varvec{x}). \end{aligned}$$

This means that the new magnetic field \(\varvec{B}^\prime (\varvec{x})\) satisfies the following equation

$$\begin{aligned} \nabla \times \varvec{B}^\prime (\varvec{x}) \ = \ \varvec{J}_{ext} (\varvec{x}) \ + \ \varvec{J}_{string} (\varvec{x}), \end{aligned}$$

which is different from the original Maxwell equation

$$\begin{aligned} \nabla \times \varvec{B}^{(S)} (\varvec{x}) \ = \ \varvec{J}_{ext} (\varvec{x}). \end{aligned}$$

In this way, we now realize that the multi-valued gauge transformation changes the magnetic field distribution inside the solenoid, and formally this change of the magnetic field distribution is thought to be generated by a peculiar effective source current \(\varvec{J}_{string} (\varvec{x})\) induced by the above singular gauge transformation. A logical consequence of this consideration is that it might also affect Boyer’s interaction energy between the solenoid and the charged particle.

To confirm it, let us first remember the expression of Boyer’s interaction energy written as

$$\begin{aligned} \Delta \varepsilon (Boyer) \ = \ \frac{1}{4 \,\pi } \,\int \,d^3 x \, \varvec{B}^s \cdot (\varvec{v} \times \varvec{E}^\prime ) \ = \ \varvec{v} \cdot \frac{1}{4 \,\pi } \,\int \,d^3 x \,\varvec{E}^\prime \times \varvec{B}^s. \end{aligned}$$

with

$$\begin{aligned} \varvec{E}^\prime (\varvec{x}, t)= & {} - \,\nabla \,\frac{e}{\vert \varvec{x} - \varvec{x}^\prime \vert }, \varvec{B}^s (\varvec{x}) \ = \ B \,\theta (R - \rho ) \,\varvec{e}_z. \end{aligned}$$

As already pointed out, under the multi-valued gauge transformation, the magnetic field distribution changes as

$$\begin{aligned} \varvec{B}^s (\varvec{x}) \ \rightarrow \ \varvec{B}^s (\varvec{x}) \ + \ \varvec{B}_{string} (\varvec{x}). \end{aligned}$$

This means that Boyer’s energy also changes as

$$\begin{aligned} \Delta \varepsilon (Boyer) \ \rightarrow \ \Delta \varepsilon ^\prime (Boyer) \ = \ \Delta \varepsilon (Boyer) \ + \ \Delta \varepsilon (string), \end{aligned}$$

where

$$\begin{aligned} \Delta \varepsilon (string) \ = \ \varvec{v} \cdot \frac{1}{4 \,\pi } \,\int \,d^3 x \,\varvec{E}^\prime \times \varvec{B}_{string}. \end{aligned}$$

After some algebra, we can show that

$$\begin{aligned} \Delta \varepsilon (string) \ = \ - \,e \,\varvec{v} \cdot \left( \,\frac{\Phi }{2 \,\pi } \, \frac{1}{\rho ^\prime } \right) \,\varvec{e}_{\phi ^\prime }. \end{aligned}$$

Comparing this with the expression of \(\Delta \varepsilon (Boyer)\) given as

$$\begin{aligned} \Delta \varepsilon (Boyer) \ = \ e \,\varvec{v} \cdot \frac{\Phi }{2 \,\pi } \, \left[ \frac{\rho ^\prime \,\theta (R - \rho ^\prime )}{R^2} \ + \ \frac{\theta (\rho ^\prime - R)}{\rho ^\prime } \right] \,\varvec{e}_{\phi ^\prime }, \end{aligned}$$

we find that

$$\begin{aligned} \Delta \varepsilon ^\prime (Boyer) \ = \ \Delta \varepsilon (Boyer) \ + \ \Delta \varepsilon (string) \ = \ 0. \end{aligned}$$

outside the solenoid, i.e. for \(\rho ^\prime > R\), which is the region where the charged particle is making a rectilinear motion. That is, Boyer’s interaction energy becomes zero after the multi-valued gauge transformation. As is discussed in the main text, the expression of Saldanha’s energy is just the same as Boyer’s energy except for the sign difference, so that it is clear that the Saldanha’s energy also becomes zero after the multi-valued gauge transformation. This means that the cancellation between the energies of Boyer and of Saldanha remains intact even after the multi-valued gauge transformation.

In this way, we realize that that the interaction energies of Boyer (and also of Saldanha) is not gauge-invariant, if a multi-valued gauge transformation is allowed. However, remember that the multi-valued gauge transformation induces peculiar effective current and it changes the form of the Maxwell equation for the magnetic field. Then, even though such a multi-valued gauge transformation is mathematically allowed, we feel it very unnatural from a physical viewpoint. On the other hand, if we confine to regular gauge transformation, we can say that the transverse part of the gauge potential is gauge-invariant as well as unique, and the energies of Boyer and Saldanha are certainly expressed with this transverse component, even though these two energies are destined to cancel each other out.

Appendix C: On the identity for \(H_{EM} + H_j\)

With the use of Eqs. (90) and (97), we get

$$\begin{aligned} H_{EM} + H_j= & {} \tilde{H}_{EM} + \int d^3 x \,\, \tilde{\varvec{B}} (\varvec{x}) \cdot \varvec{B}_{ext} (\varvec{x}) \nonumber \\&- \int d^3 x \,\,\tilde{\varvec{A}} (\varvec{x}) \cdot \varvec{j}_{ext} (\varvec{x}) + \text{ constant }. \end{aligned}$$
(C17)

Note that the 2nd and 3rd terms on the r.h.s. of the above equation cancel out owing to the nontrivial identity (74) in the text. Thus, we have

$$\begin{aligned} H_{EM} + H_j &= {} \tilde{H}_{EM} + \text{ constant } \nonumber \\ &= {} \sum _{\lambda = 1}^2 \,\int d^3 k \,\,\omega \, \tilde{a}^\dagger (\varvec{k}, \lambda ) \,\tilde{a} (\varvec{k}, \lambda ) + \text{ constant }. \end{aligned}$$
(C18)

This means that the ground state of \(H_{EM} + H_j\) is the vacuum \(\vert \tilde{0} \rangle \) of the quanta \(\tilde{a} (\varvec{k}, \lambda )\), so that it holds that

$$\begin{aligned} \left( H_{EM} + H_j \right) \,\vert \tilde{0} \rangle \ = \ \tilde{E}_0 \,\vert \tilde{0} \rangle . \end{aligned}$$
(C19)

Here, \(\tilde{E}_0\) represents the ground-state energy of \(H_{EM} + H_j\), which may in principle contain self-energies of the solenoid and the moving charge.

Appendix D: Proof of Eq. (71).

Up to \(O (e^2)\), we have

$$\begin{aligned}&\langle \tilde{0}_1 \vert \,\tilde{\varvec{B}} (\varvec{x}) \vert \, \tilde{0}_1 \rangle \nonumber \\ &\quad = - \ \frac{e}{m} \,\langle \tilde{0} \,\vert \, \nabla \times \tilde{\varvec{A}} (\varvec{x}) \, \frac{1}{\tilde{E}_0 - (H_{EM} + H_j )} \, \varvec{p} \cdot \tilde{\varvec{A}} (\varvec{q}) \,\vert \tilde{0} \rangle \ + \ c.c \end{aligned}$$
(D20)

In view of the Fourier expansion of \(\tilde{\varvec{A}} (\varvec{x})\), this gives

$$\begin{aligned}&\langle \tilde{0}_1 \vert \,\tilde{\varvec{B}} (\varvec{x}) \vert \, \tilde{0}_1 \rangle \nonumber \\ &\quad = \ - \,2 \,\,\frac{e}{m} \,\int \frac{d^3 k}{(2 \,\pi )^3} \,\sum _{\lambda = 1}^2 \,\, \frac{\langle \tilde{0} \,\vert \,\nabla \times \tilde{\varvec{A}} (\varvec{x}) \, \vert \varvec{k}, \lambda \rangle \,\langle \varvec{k}, \lambda \,\vert \, \varvec{p} \cdot \tilde{\varvec{A}} (\varvec{q}) \,\vert \tilde{0} \rangle }{- \,\omega } \,\, \frac{1}{2 \,\omega }. \end{aligned}$$
(D21)

Now, we proceed as

$$\begin{aligned}&\int \frac{d^3 k}{(2 \,\pi )^3} \,\frac{1}{- \,\omega ^2} \, \sum _{\lambda = 1}^2 \,\langle \tilde{0} \,\vert \, \nabla \times \tilde{\varvec{A}} (\varvec{x}) \,\vert \varvec{k}, \lambda \rangle \, \langle \varvec{k}, \lambda \,\vert \varvec{p} \cdot \tilde{\varvec{A}} (\varvec{q}) \, \tilde{0} \rangle \nonumber \\ &\quad = - \, 2 \,i \,\int \frac{d^3 k}{(2 \,\pi )^3} \,\frac{1}{\omega ^2} \, \sum _{\lambda = 1}^2 \,\varvec{k} \times \varvec{\epsilon }_{\varvec{k}, \lambda } \, e^{\,i \,\varvec{k} \cdot \varvec{x}} \,\varvec{p} \cdot \varvec{\epsilon }_{\varvec{k}, \lambda } \,e^{\,- \,i \,\varvec{k} \cdot \varvec{q}}. \end{aligned}$$
(D22)

Then, with the use of the formula for the polarization sum, we get

$$\begin{aligned} \sum _{\lambda = 1}^2 \,\varvec{k} \times \varvec{\epsilon }_{\varvec{k}, \lambda } \,\, \varvec{p} \cdot \varvec{\epsilon }_{\varvec{k}, \,\lambda } \ = \ - \,\varvec{p} \times \varvec{k}, \end{aligned}$$
(D23)

we finally obtain

$$\begin{aligned} \langle \tilde{0}_1 \,\vert \tilde{\varvec{B}} (\varvec{x}) \,\vert \tilde{0}_1 \rangle \ = \ - \,\frac{\varvec{p}}{m} \times \nabla \,\frac{1}{4 \,\pi } \, \frac{1}{\vert \varvec{x} - \varvec{q} \vert } \ + \ O (e^2), \end{aligned}$$
(D24)

which proves Eq. (95) in the text.

Rights and permissions

Springer Nature or its licensor (e.g. a society or other partner) holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Wakamatsu, M. Is the Aharonov–Bohm phase shift for a non-closed path a measurable quantity?. Eur. Phys. J. Plus 139, 112 (2024). https://doi.org/10.1140/epjp/s13360-024-04922-9

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1140/epjp/s13360-024-04922-9

Navigation