Skip to main content
Log in

A Flea on Schrödinger’s Cat

  • Published:
Foundations of Physics Aims and scope Submit manuscript

‘Another secondary readership is made up of those philosophers and physicists who—again like myself—are puzzled by so-called foundational issues: what the strange quantum formalism implies about the nature of the world it so accurately describes. (…) My presentation is suffused with a perspective on the quantum theory that is very close to the venerable but recently much reviled Copenhagen interpretation. Those with a taste for such things may be startled to see how well quantum computation resonates with the Copenhagen point of view. Indeed, it had been my plan to call this book Copenhagen Computation until the excellent people at Cambridge University Press and my computer-scientist friends persuaded me that virtually no members of my primary readership would then have any idea what it was about.’

David Mermin, Quantum Computer Science: An Introduction [87] (Preface)

Abstract

We propose a technical reformulation of the measurement problem of quantum mechanics, which is based on the postulate that the final state of a measurement is classical; this accords with experimental practice as well as with Bohr’s views. Unlike the usual formulation (in which the post-measurement state is a unit vector in Hilbert space), our version actually opens the possibility of admitting a purely technical solution within the confines of conventional quantum theory (as opposed to solutions that either modify this theory, or introduce unusual and controversial interpretative rules and/or ontologies).

To that effect, we recall a remarkable phenomenon in the theory of Schrödinger operators (discovered in 1981 by Jona-Lasinio, Martinelli, and Scoppola), according to which the ground state of a symmetric double-well Hamiltonian (which is paradigmatically of Schrödinger’s Cat type) becomes exponentially sensitive to tiny perturbations of the potential as ħ→0. We show that this instability emerges also from the textbook wkb approximation, extend it to time-dependent perturbations, and study the dynamical transition from the ground state of the double well to the perturbed ground state (in which the cat is typically either dead or alive, depending on the details of the perturbation).

Numerical simulations show that adiabatically arising perturbations may (quite literally) cause the collapse of the wave-function in the classical limit. Thus, at least in the context of a simple mathematical model, we combine the technical and conceptual virtues of decoherence (which fails to solve the measurement problem but launches the key idea that perturbations may come from the environment) with those of dynamical collapse models à la grw (which do solve the measurement problem but are ad hoc), without sharing their drawbacks: single measurement outcomes are obtained (instead of merely diagonal reduced density matrices), and no modification of quantum mechanics is needed.

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

Access this article

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

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8

Similar content being viewed by others

Notes

  1. We regard these as Wittgensteinian “hinge propositions” [114], on which modern physics is based.

  2. This is sometimes called Bohr’s doctrine of classical concepts [102], which accurately describes experimental practice! See also [75] for a detailed analysis.

  3. For our work it is a moot point whether Bohr endorsed this second point as well; it is hard to say.

  4. The original expositions [28, 51, 59, 65, 66, 111] might be hard to follow for non-mathematicians (see, however, [22]). Applications in chemistry and solid-state physics may be found in [23, 67, 84].

  5. This notion needs to be quantified, of course [2, 19, 78]. See also [75] and the present paper.

  6. This is true in particular for the Many-Worlds (aka “Everett”) Interpretation [101], or the Modal Interpretation of quantum mechanics [36], in which radical changes are proposed in the ontology and/or the usual interpretative rules of the theory, without clarifying in any way what is really going on during measurements (a question one indeed is not supposed to ask, according to received wisdom). Bohmian mechanics (as a modern incarnation of de Broglie’s pilot-wave theory) does a better job here [33, 38, 39], but its narrow applicability (at least in its current form), focusing as it does on position as the only physical observable, makes it unattractive to many (including the authors). See some kinship in Sect. 5.4, however.

  7. This applies, for example, to the famous paper by Danieri, Loinger, and Prosperi [34], to early papers on decoherence [117], and to much of the mathematical physics literature on the measurement problem, including the work of the senior author [40, 60, 72, 73, 109]. We now regard such papers as mathematically interesting but conceptually misguided, at least on this point. On the other hand, it is to the credit of especially the Swiss school that it drew attention to the idea that measurement involves limiting procedures, so that solutions of the measurement problem should at least incorporate the appropriate limits.

  8. In this sense even von Neumann’s book [89] is misleading, since he suggested that the act of observation may be identified with a cut in the chain now named after him. What is right about this idea is that observation is linked to a voluntary change of information, but it would have been preferable to point out that such a change, in so far as it defines measurement, should be a loss of quantum information.

  9. Furthermore, despite its outspoken ambition to derive classical physics from quantum theory [68, 103, 117], decoherence hardly (if at all) invokes limits like Planck’s constant going to zero, which are needed, for one thing, to derive the correct classical equations of motion (cf. [75, 76]).

  10. An introduction for philosophers to the material in this section may be found in [75, Chaps. 4 and 5].

  11. A C-algebra is a complex algebra A that is complete in a norm ∥⋅∥ satisfying ∥ab∥≤∥a∥∥b∥ for all a,bA, and has an involution aa such that ∥a a∥=∥a2.

  12. Positivity of ω means that ω(a a)≥0 for all aA. If A has a unit 1, then a state may equivalently be defined as a positive linear functional ω:A→ℂ that satisfies ω(1)=1.

  13. A self-adjoint operator a on a Hilbert space is compact just in case it has a spectral decomposition a=∑ i λ i |e i 〉〈e i | (where |e i 〉〈e i | is the orthogonal projection onto the ray ℂe i ), where the eigenvectors (e i ) form an orthonormal basis of H, each nonzero eigenvalue λ i has finite multiplicity, and if the eigenvalues are listed in decreasing order of their absolute value (i.e., ∥a∥=|λ 1|≥|λ 2|≥⋯), then lim α→∞|λ α |=0.

  14. Although unbounded operators play a major practical role in physics (think of position and momentum), they may be awkward to deal with due to domain issues and in any case they can theoretically be avoided without any loss of generality. Indeed, unbounded self-adjoint operators a bijectively correspond to bounded operators (or constructions involving those) in at least four different ways [97]. First, one may pass to the unitary Cayley transform (ai)(a+i)−1. Second, one may construct the associated one-parameter unitary group t↦exp(ita) by Stone’s Theorem. Third, one may work with the bounded spectral projections, from which the operator may be reconstructed by the spectral theorem. Fourth, one could take the resolvents (aλ)−1, λ∉ℝ (whose typical integral kernels are Green’s functions).

  15. Classically, one could work with the (essentially) bounded integrable functions L (ℝ2), whilst in quantum theory one could take the algebra B(L 2(ℝ)) of all bounded operators on L 2(ℝ) (as opposed to merely the compact ones). These broader classes can be constructed from the C 0-functions or compact operators by taking pointwise and weak (or strong) limits, respectively. Enlarging the class of observables like that also brings in new continuity conditions on the states (namely σ-additivity and σ-weak continuity, respectively). Imposing these leads to the same states as discussed above; without them, one obtains more.

  16. Often Weyl quantization \(Q^{W}_{\hbar}\) is used instead of Berezin quantization Q ħ , as in [9], but for Schwartz functions f on phase space these have the same asymptotic properties as ħ→0 [74]. The advantage of Berezin quantization is that it is well defined also for continuous functions vanishing at infinity, in that for any unit vector ΨL 2(ℝ) the map f↦〈Ψ|Q ħ (f)|Ψ〉 defines a probability measure on phase space. In contrast, the Wigner function defined by \(f\mapsto\langle\varPsi| Q^{W}_{\hbar}(f)|\varPsi\rangle\) may fail to be positive, as is well known.

  17. In (2.18) we regard classical states as probability measures on phase space; hence the addition on the right-hand side is a convex sum of measures, which has nothing to do with addition in the particular phase space ℝ2 (whose linear structure is accidental and irrelevant).

  18. This appeal to “practice” does not mean that we are resigned to fapp (i.e., “for all practical purposes”) solutions to the measurement problem. As in [73], we remain convinced that the classical description of a measurement apparatus is a purely epistemic move, relative to which outcomes are defined. So even if it were possible to study a cat as a quantum system, there would be no measurement problem, since in that case there would be innumerable superpositions but not a single (undesirable) mixture of classical states.

  19. We share this rejection with the Bohmians [32]. The folk wisdom (shared by the Founding Fathers) that the Copenhagen Interpretation has no measurement problem relies on these secondary Copenhagenian claims, which indeed sweep the problem under the rug. Incidentally, these claims seem much more popular than Bohr’s doctrine of classical concepts, which is generally not well understood, and/or mistaken for the idea that the goal of physics is to explain experiments, or that reality does not exist, et cetera.

  20. The analogy with the thermodynamic limit will be discussed in Sect. 5.2. As to the limit ħ→0, we repeat [75, pp. 471–472] that although ħ is a dimensionful constant, in practice one studies the (semi)classical regime of a given quantum theory by forming a dimensionless combination of ħ and other parameters; this combination then re-enters the theory as if it were a dimensionless version of ħ that can indeed be varied. The oldest example is Planck’s radiation formula, with the associated limit ħν/kT→0, and another example is the Schrödinger operator (2.10), with mass reinserted, where one may pass to a dimensionless parameter \(\hbar /\lambda \sqrt{2m\epsilon}\), where λ and ϵ are typical length and energy scales, respectively.

  21. Paraphrasing Bell [6]: the difference between \(\rho_{0}^{\pm}\) and \(\psi_{\hbar}^{(0)}\) can be made ‘as big as you do not like.’.

  22. Families of unit vectors like \(\varPsi^{(i)}_{\hbar}\), where i=0,1,+,−, typically do not have a limit as unit vectors (or even as density matrices, including one-dimensional projections).

  23. As explained above, the nonlinearity inherent in the limit ħ→0 makes it impossible to find the limit of this ground state \(\varPsi^{(0)}_{\hbar}\) by just adding the results for two localized wave-functions like \(\varPsi^{+}_{\hbar}\) and \(\varPsi^{-}_{\hbar}\).

  24. The “Flea on the Elephant” terminology used in [111] for the phenomenon in question evidently motivated the title of the present paper, which has identified the proper host animal at last!

  25. Some of the details in this section depend on the latter assumption, but our overall scenario in Sect. 4 does not. For example, if the value and/or the curvature of one of the minima is decreased, then the ground state wave-function will localize above that minimum, as follows from standard minimax techniques taking single harmonic eigenfunctions as trial states [51, 98]. So collapse is actually easier in that case.

  26. Symmetric perturbations are excluded by 3, as these would satisfy \(d_{V}'=d_{V}''\).

  27. If δV has support on both sides of the real axis (which is possible in the case (3.6)), a more detailed analysis of its shape is necessary in order to predict the direction of collapse [21].

  28. Compare [98, p. 35], [111] for such arguments. Nonetheless, the effect of the flea is counterintuitive even from the point of view of quantum-mechanical tunneling: for example, with a perturbation of the kind displayed in Figs. 4, 5 and 6, which falls under case (3.6), one would expect tunneling from the right into the left-handed well to be discouraged, even increasingly so as ħ→0, because the potential barrier through which to tunnel has been heightened, but in fact the right-handed peak of the unperturbed ground state tunnels to the left so as to localize the ground state wave-function. See Sect. 5.3 for further discussion.

  29. Cf. [7] (for mathematicians) or [55] (for physicists), and references therein.

  30. This approximation is extremely well known also in physics [79], but has hardly been studied in the present context. It is too simple to display the behaviour (2.22)–(2.24), though. See also Sect. 5.1.

  31. A corresponding movie may be found on www.math.ru.nl/~landsman/flea.avi.

  32. This states that under an adiabatic perturbation δV=δV(t/T) the unperturbed ground state moving under the perturbed Hamiltonian converges to the ground state of the latter as tT→∞ [52, 54].

  33. For potentials with n identical wells, this basis may be chosen to consist of the eigenfunctions of the n individual wells with Dirichlet boundary conditions [58, 59]. For the quantum Ising model (cf. [70]), where n=2, one may take the state with all spins up and the state with all spins down, etc.

  34. In this respect our approach has an edge over Bohmian mechanics and the grw theory, although it remains to be shown that typical environments induce the right perturbations for the mechanism to apply.

  35. Like the measurement problem, this seemingly paradoxical situation does not seem to bother physicists very much, although their Higgs mechanism relies on a resolution of it: apparently, in any finite volume the system refuses to choose a ground state (or vacuum), although all perturbative calculations underlying the successful Standard Model of elementary particle physics rely on such a choice. But it has been the subject of recent discussions in the philosophy of science [4, 20, 80, 86, 92], in which some claim that this “discontinuity” in passing from N<∞ to N=∞ is crucial for the possibility of emergence (‘More is Different’), whilst others try to find arguments for continuity and hence defend some form of reductionism.

  36. If it happens to be true that measurement outcomes emerge adiabatically, it would be a marked break with tradition, starting with von Neumann’s model, in which both the measurement interaction and the alleged collapse take place instantly [89]. However, the recent notion of a “weak measurement” seems to support our approach on this point [69, 82, 100]. We are indebted to Jos Groot for these references.

  37. See [26] for a recent discussion of randomness amplification, which focuses on the way experiments may be construed to amplify the randomness inherent in the (alleged) “free” choice of an experimentalist.

  38. What follows will hardly be new to specialists in these matters, but it needs to be stated clearly.

  39. Using the traditional scene where (Alice, Bob) are spacelike separated and perform experiments with settings (a,b) and outcomes (x,y), respectively, a stochastic hidden variable λ (or rather the corresponding theory) satisfies oi if Alice’s conditional probability P(xa,b,y,λ) of finding x given (a,b,y,λ) is independent of Bob’s outcome y, whereas the theory satisfies pi if her conditional probability P(xa,b,λ) of finding x given (a,b,λ) is independent of Bob’s setting b, and vice versa. These notions were originally introduced, in slightly different form, by Jarrett [64]. Controversies around this terminology and its use of the sort discussed in e.g. [18, 48, 85, 93, 106] seem irrelevant to our purposes.

  40. Muller also notes that this paper reveals the ‘empiricist presuppositions’ of the authors. For example, we are not worried about superpositions like Ψ=∑ i c i Ψ i in Sect. 2.1 as such, despite the fact that most quantum-mechanical observables, including the one (O) that is being measured, have no value in such a state. Compared to classical physics this is admittedly a curious feature of quantum theory, but it causes neither inconsistencies within the formalism nor disagreements with observation. Furthermore, it has nothing to do with Ψ being a superposition (which is a basis-dependent statement), and in our view it only leads to a problem (namely the measurement problem) if the device whose state Ψ represents is classical (e.g., in being macroscopic). For in that case quantum mechanics naively (i.e. without the flea mechanism) appears to assign numerous values to O, which seems unacceptable (rather than assigning no value, which for the stated reasons we can live with both as theorists and as alleged empiricists).

  41. As opposed to the extremely sophisticated and mathematically rigorous methods of Helffer and Sjöstrand [37, 58, 59], who somewhat confusingly suggest they use the ordinary wkb method.

  42. This result can also be found by applying the method of comparison equations, which is explained in [116]. Further references are [81] and [88].

  43. This can be explained by level crossing, i.e. certain energy levels of the two individual wells coincide.

  44. To keep the discussion straightforward, we ignored the special case δ∈{|k∈ℤ∖{0}} here.

References

  1. Bambusi, D., Graffi, S., Paul, T.: Long time semiclassical approximation of quantum flows: a proof of the Ehrenfest time. Asymptot. Anal. 21, 149–160 (1999)

    MathSciNet  MATH  Google Scholar 

  2. Bassi, A., Ghirardi, G.: Dynamical reduction models. Phys. Rep. 379, 257–426 (2003)

    Article  MathSciNet  ADS  MATH  Google Scholar 

  3. Bassi, A., Ghirardi, G.: The Conway-Kochen argument and relativistic GRW models. Found. Phys. 37, 169–185 (2007)

    Article  MathSciNet  ADS  MATH  Google Scholar 

  4. Batterman, R.W.: Emergence, singularities, and symmetry breaking. Found. Phys. 41, 1031–1050 (2011)

    Article  MathSciNet  ADS  MATH  Google Scholar 

  5. Bell, J.S.: On the Einstein–Podolsky–Rosen paradox. Physics 1, 195–200 (1964)

    Google Scholar 

  6. Bell, J.S.: On wave packet reduction in the Coleman-Hepp model. Helv. Phys. Acta 48, 93–98 (1975)

    MathSciNet  MATH  Google Scholar 

  7. Berglund, N.: Kramers’ law: validity, derivations and generalizations arXiv:1106.5799

  8. Berry, M.V.: The adiabatic limit and the semiclassical limit. J. Phys. A 17, 1225–1233 (1984)

    Article  MathSciNet  ADS  Google Scholar 

  9. Berry, M.V., Balazs, N.L.: Evolution of semiclassical quantum states in phase space. J. Phys. A 12, 625–642 (1979)

    Article  MathSciNet  ADS  MATH  Google Scholar 

  10. Bohr, N.: Discussion with Einstein on epistemological problems in atomic physics. In: Schlipp, P.A. (ed.) Albert Einstein: Philosopher-Scientist, pp. 201–241. Open Court, La Salle (1949)

    Google Scholar 

  11. Blanchard, Ph., Bolz, G., Cini, M., Angelis, G.F., Serva, M.: Localization stabilized by noise. J. Stat. Phys. 75, 749–755 (1994)

    Article  ADS  Google Scholar 

  12. Bonechi, F., De Bièvre, S.: Exponential mixing and |lnħ| time scales in quantized hyperbolic maps on the torus. Commun. Math. Phys. 211, 659–686 (2000)

    Article  ADS  MATH  Google Scholar 

  13. und, H., Born, M., Einstein, A.: Briefwechsel 1916–1955. Nymphemburger, München (1969)

    Google Scholar 

  14. Born, M.: Quantenmechanik der Stoßvorgänge. Z. Phys. 38, 803–827 (1926)

    Article  ADS  Google Scholar 

  15. Bouzouina, A., Robert, D.: Uniform semiclassical estimates for the propagation of quantum observables. Duke Math. J. 11, 223–252 (2002)

    MathSciNet  Google Scholar 

  16. Brown, H.R.: The insolubility proof of the quantum measurement problem. Found. Phys. 16, 857–870 (1986)

    Article  MathSciNet  ADS  Google Scholar 

  17. Brown, H., Svetlichny, G.: Nonlocality and Gleason’s lemma. Part I. Deterministic theories. Found. Phys. 20, 1379–1387 (1990)

    Article  MathSciNet  ADS  Google Scholar 

  18. Bub, J.: Interpreting the Quantum World. Cambridge University Press, Cambridge (1997)

    MATH  Google Scholar 

  19. Busch, P., Lahti, P.J., Mittelstaedt, P.: The Quantum Theory of Measurement. Springer, Berlin (1991)

    Google Scholar 

  20. Butterfield, J.: Less is different: emergence and reduction reconciled. Found. Phys. 41, 1065–1135 (2011)

    Article  MathSciNet  ADS  MATH  Google Scholar 

  21. Cesi, F.: An algorithm to study tunneling in a wide class of one-dimensional multiwell potentials. I. J. Phys. A 22, 1027–1052 (1989)

    Article  MathSciNet  ADS  MATH  Google Scholar 

  22. Cesi, F., Rossi, G.C., Testa, M.: Non-symmetric double well and Euclidean functional integral. Ann. Phys. 206, 318–333 (1991)

    Article  MathSciNet  ADS  MATH  Google Scholar 

  23. Claverie, P., Jona-Lasinio, G.: Instability of tunneling and the concept of molecular structure in quantum mechanics: the case of pyramidal molecules and the enantiomer problem. Phys. Rev. A 33, 2245–2253 (1986)

    Article  ADS  Google Scholar 

  24. Clifton, R.: Getting contextual and nonlocal elements-of-reality the easy way. Am. J. Phys. 61, 443–447 (1993)

    Article  ADS  Google Scholar 

  25. Clifton, R., Bub, J., Halvorson, H.: Characterizing quantum theory in terms of information-theoretic constraints. Found. Phys. 33, 1561–1591 (2003)

    Article  MathSciNet  MATH  Google Scholar 

  26. Colbeck, R., Renner, R.: Free randomness can be amplified. Nat. Phys. 8, 450–454 (2012)

    Article  Google Scholar 

  27. Colin de Verdière, Y., Parisse, B.: Équilibre instable en régime semi-classique I: concentration microlocale. Commun. Partial Differ. Equ. 19, 1535–1563 (1994)

    Article  Google Scholar 

  28. Combes, J.M., Duclos, P., Seiler, R.: Convergent expansions for tunneling. Commun. Math. Phys. 92, 229–245 (1983)

    Article  MathSciNet  ADS  MATH  Google Scholar 

  29. Combescure, M., Robert, D.: Semiclassical spreading of quantum wave packets and applications near unstable fixed points of the classical flow. Asymptot. Anal. 14, 377–404 (1997)

    MathSciNet  MATH  Google Scholar 

  30. Conway, J., Kochen, S.: The Free Will Theorem. Found. Phys. 36, 1441–1473 (2006)

    Article  MathSciNet  ADS  MATH  Google Scholar 

  31. Conway, J., Kochen, S.: The Strong Free Will Theorem. Not. Am. Math. Soc. 56, 226–232 (2009)

    MathSciNet  MATH  Google Scholar 

  32. Cushing, J.T.: Quantum Mechanics: Historical Contingency and the Copenhagen Hegemony. University of Chicago Press, Chicago (1994)

    MATH  Google Scholar 

  33. Cushing, J.T., Fine, A. (eds.) Bohmian Mechanics and Quantum Theory: an Appraisal (1996)

    Google Scholar 

  34. Danieri, A., Loinger, A., Prosperi, G.M.: Quantum theory of measurement and ergodic condition. Nucl. Phys. 33, 297–319 (1962)

    Article  Google Scholar 

  35. De Bièvre, S., Robert, D.: Semiclassical propagation on |logħ| time scales. Int. Math. Res. Not. 12, 667–696 (2003)

    Article  Google Scholar 

  36. Dieks, D., Vermaas, P.E. (eds.): The Modal Interpretation of Quantum Mechanics. Kluwer Academic, Norwell (1998)

    MATH  Google Scholar 

  37. Dimassi, M., Sjöstrand, J.: Spectral Asymptotics in the Semi-classical Limit. Cambridge University Press, Cambridge (1999)

    Book  MATH  Google Scholar 

  38. Dürr, D., Goldstein, S., Zanghi, N.: Quantum mechanics, randomness, and deterministic reality. Phys. Lett. A 172, 6–12 (1992)

    Article  MathSciNet  ADS  Google Scholar 

  39. Dürr, D., Teufel, S.: Bohmian Mechanics: The Physics and Mathematics of Quantum Theory. Springer, Berlin (2009)

    MATH  Google Scholar 

  40. Emch, G.G., Whitten-Wolfe, B.: Mechanical quantum measuring process. Helv. Phys. Acta 49, 45–55 (1976)

    MathSciNet  Google Scholar 

  41. Faure, F.: Semi-classical formula beyond the Ehrenfest time in quantum chaos. I. Trace formula. Ann. Inst. Fourier (Grenoble) 57, 2525–2599 (2007)

    Article  MathSciNet  MATH  Google Scholar 

  42. Fine, A.: Insolubility of the Quantum Measurement Problem. Phys. Rev. D 2, 2783–2787 (1970)

    Article  MathSciNet  ADS  MATH  Google Scholar 

  43. Fine, A.: The Shaky Game: Einstein, Realism, and the Quantum Theory. University of Chicago Press, Chicago (1986). Also in Commun. Math. Phys. 101, 21–46 (1985)

    Google Scholar 

  44. Fröman, N., Fröman, P.O.: Phase-Integral Method: Allowing Nearlying Transition Points. Springer, Berlin (1996)

    Book  MATH  Google Scholar 

  45. Fröman, N., Fröman, P.O.: Physical Problems Solved by the Phase-Integral Method. Cambridge University Press, Cambridge (2002)

    Book  MATH  Google Scholar 

  46. Garg, A.: Tunnel splittings for one-dimensional potential wells revisited. Am. J. Phys. 68, 430–437 (2000)

    Article  ADS  Google Scholar 

  47. Germinet, F., De Bièvre, S.: Dynamical localization for discrete and continuous random Schrödinger operators. Commun. Math. Phys. 194, 323–341 (1998)

    Article  ADS  MATH  Google Scholar 

  48. Ghirardi, G.C.: Does quantum nonlocality irremediably conflict with special relativity? Found. Phys. 40, 1379–1395 (2010)

    Article  MathSciNet  ADS  MATH  Google Scholar 

  49. Ghirardi, G.C., Grassi, R., Butterfield, J., Fleming, G.N.: Parameter dependence and outcome dependence in dynamical models for state vector reduction. Found. Phys. 23, 341–364 (1993)

    Article  MathSciNet  ADS  Google Scholar 

  50. Ghirardi, G.C., Rimini, A., Weber, T.: Unified dynamics for microscopic and macroscopic systems. Phys. Rev. D 34, 470–491 (1986)

    Article  MathSciNet  ADS  MATH  Google Scholar 

  51. Graffi, S., Grecchi, V., Jona-Lasinio, G.: Tunnelling instability via perturbation theory. J. Phys. A 17, 2935–2944 (1984)

    Article  MathSciNet  ADS  MATH  Google Scholar 

  52. Griffiths, D.J.: Introduction to Quantum Mechanics. Pearson Education, Upper Saddle River (1995)

    MATH  Google Scholar 

  53. Haag, R.: Local Quantum Physics: Fields, Particles, Algebras. Springer, Berlin (1992)

    Book  MATH  Google Scholar 

  54. Hagedorn, G.A., Joye, A.: Elementary exponential error estimates for the adiabatic approximation. J. Math. Anal. Appl. 267, 235–246 (2002)

    Article  MathSciNet  MATH  Google Scholar 

  55. Hänggi, P., Talkner, P., Borkovec, M.: Reaction-rate theory: fifty years after Kramers. Rev. Mod. Phys. 62, 251–341 (1990)

    Article  ADS  Google Scholar 

  56. Harrell, E.W.: Double wells. Commun. Math. Phys. 75, 239–261 (1980)

    Article  MathSciNet  ADS  MATH  Google Scholar 

  57. Heisenberg, W.: Über den anschaulichen Inhalt der quantentheoretischen Kinematik und Mechanik. Z. Phys. 43, 172–198 (1927)

    Article  ADS  MATH  Google Scholar 

  58. Helffer, B.: Semi-classical Analysis for the Schrödinger Operator and Applications. Springer, Berlin (1988)

    MATH  Google Scholar 

  59. Helffer, B., Sjöstrand, J.: Puits multiples en limite semi-classique. II. Interaction moléculaire. Symétries. Perturbation. Ann. Inst. Henri Poincaré, a Phys. Théor. 42, 127–212 (1985)

    MATH  Google Scholar 

  60. Hepp, K.: Quantum theory of measurement and macroscopic observables. Helv. Phys. Acta 45, 237–248 (1972)

    Google Scholar 

  61. Heywood, P., Redhead, M.: Nonlocality and the Kochen–Specker paradox. Found. Phys. 13, 481–499 (1983)

    Article  MathSciNet  ADS  Google Scholar 

  62. Hislop, P.D., Sigal, I.M.: Introduction to Spectral Theory. Springer, Berlin (1996)

    Book  MATH  Google Scholar 

  63. Janssen, H.: Reconstructing reality: environment-induced decoherence, the measurement problem, and the emergence of definiteness in quantum mechanics. M.Sc. Thesis, Radboud University Nijmegen. Available at: philsci-archive.pitt.edu/archive/00004224/

  64. Jarrett, J.P.: On the physical significance of the locality conditions in the Bell arguments. Nous 18, 569–589 (1984)

    Article  MathSciNet  Google Scholar 

  65. Jona-Lasinio, G., Martinelli, F., Scoppola, E.: New approach to the semiclassical limit of quantum mechanics. Commun. Math. Phys. 80, 223–254 (1981)

    Article  MathSciNet  ADS  MATH  Google Scholar 

  66. Jona-Lasinio, G., Martinelli, F., Scoppola, E.: The semiclassical limit of quantum mechanics: a qualitative theory via stochastic mechanics. Phys. Rep. 77, 313–327 (1981)

    Article  MathSciNet  ADS  Google Scholar 

  67. Jona-Lasinio, G., Presilla, C., Toninelli, C.: Interaction induced localization in a gas of pyramidal molecules. Phys. Rev. Lett. 88, 123001 (2002)

    Article  ADS  Google Scholar 

  68. Joos, E., Zeh, H.D., Kiefer, C., Giulini, D., Kupsch, J., Stamatescu, I.-O.: Decoherence and the Appearance of a Classical World in Quantum Theory. Springer, Berlin (2003)

    Google Scholar 

  69. Kocsis, S., et al.: Observing the average trajectories of single photons in a two-slit interferometer. Science 332, 1170–1173 (2011)

    Article  ADS  Google Scholar 

  70. Koma, T., Tasaki, H.: Symmetry breaking and finite-size effects in quantum many-body systems. J. Stat. Phys. 76, 745–803 (1994)

    Article  MathSciNet  ADS  MATH  Google Scholar 

  71. Landau, L.D., Lifshitz, E.M.: Quantum Mechanics, 3rd edn. Pergamon, Elmsford (1977)

    Google Scholar 

  72. Landsman, N.P.: Algebraic theory of superselection sectors and the measurement problem in quantum mechanics. Int. J. Mod. Phys. A 6, 5349–5372 (1991)

    Article  MathSciNet  ADS  MATH  Google Scholar 

  73. Landsman, N.P.: Observation and superselection in quantum mechanics. Stud. Hist. Philos. Mod. Phys. 26, 45–73 (1995)

    Article  MathSciNet  MATH  Google Scholar 

  74. Landsman, N.P.: Mathematical Topics Between Classical and Quantum Mechanics. Springer, Berlin (1998)

    Book  Google Scholar 

  75. Landsman, N.P.: Between classical and quantum. In: Butterfield, J., Earman, J. (eds.) Handbook of the Philosophy of Science. Vol. 2: Philosophy of Physics, Part A, pp. 417–553 (2007)

    Google Scholar 

  76. Landsman, N.P.: Essay review: Decoherence and the quantum-to-classical transition by M. Schlosshauer. Stud. Hist. Philos. Mod. Phys. 40, 94–95 (2009)

    Article  Google Scholar 

  77. Landsman, N.P.: The Born rule and its interpretation. In: Greenberger, D., Hentschel, K., Weinert, F. (eds.) Compendium of Quantum Physics, pp. 64–70. Springer, Berlin (2009)

    Chapter  Google Scholar 

  78. Leggett, A.J.: Testing the limits of quantum mechanics: motivation, state of play, prospects. J. Phys. C 14, R415–R451 (2002)

    Google Scholar 

  79. Leggett, A.J., Chakravarty, S., Dorsey, A.T., Fisher, M.P.A., Garg, A., Zwerger, W.: Dynamics of the dissipative two-state system. Rev. Mod. Phys. 59, 1–85 (1987)

    Article  ADS  Google Scholar 

  80. Liu, C., Emch, G.G.: Explaining quantum spontaneous symmetry breaking. Stud. Hist. Philos. Mod. Phys. 36, 137–163 (2005)

    Article  MathSciNet  MATH  Google Scholar 

  81. Lynn, R.Y.S., Keller, J.B.: Uniform asymptotic solutions of second order linear ordinary differential equations with turning points. Commun. Pure Appl. Math. 23, 379–408 (1970)

    Article  MathSciNet  MATH  Google Scholar 

  82. Lund, A.P., Wiseman, H.M.: Measuring measurement–disturbance relationships with weak values. New J. Phys. 12, 093011 (2010)

    Article  ADS  Google Scholar 

  83. Maioli, M., Sacchetti, A.: Two level systems driven by a stochastic perturbation. J. Stat. Phys. 119, 1383–1396 (2005)

    Article  MathSciNet  ADS  MATH  Google Scholar 

  84. Martinelli, F., Scoppola, E.: Introduction to the mathematical theory of Anderson localization. Riv. Nuovo Cimento 10, 1–90 (1987)

    MathSciNet  Google Scholar 

  85. Maudlin, T.: Quantum Non-Locality and Relativity: Metaphysical Intimations of Modern Physics, 3rd edn. Wiley, New York (2011)

    Book  Google Scholar 

  86. Menon, T., Callender, C.: Turn and face the strange … Ch-ch-changes. In Batterman, R.W. (ed.): The Oxford Handbook of Philosophy of Physics. Oxford University Press, Oxford (2013, to appear)

  87. Mermin, N.D.: Quantum Computer Science: An Introduction. Cambridge University Press, Cambridge (2007)

    Book  MATH  Google Scholar 

  88. Miller, S.C., Good, R.H.: A WKB-type approximation to the Schrödinger equation. Phys. Rev. 91, 174–179 (1953)

    Article  MathSciNet  ADS  MATH  Google Scholar 

  89. von Neumann, J.: Mathematische Grundlagen der Quantenmechanik. Springer, Berlin (1932)

    MATH  Google Scholar 

  90. Nieto, M.M., et al.: Resonances in quantum mechanical tunneling. Phys. Lett. B 163, 336–342 (1985)

    Article  MathSciNet  ADS  Google Scholar 

  91. Nonnenmacher, S., Voros, A.: Chaotic eigenfunctions in phase space. J. Stat. Phys. 92, 431–518 (1998)

    Article  MathSciNet  MATH  Google Scholar 

  92. Norton, J.D.: Approximation and idealization: why the difference matters. Philos. Sci. 79, 207–232 (2012)

    Article  MathSciNet  Google Scholar 

  93. Norsen, T.: Local causality and completeness: Bell vs. Jarrett. Found. Phys. 39, 273–294 (2009)

    Article  MathSciNet  ADS  MATH  Google Scholar 

  94. Olivieri, E., Vares, M.E.: Large Deviations and Metastability. Cambridge University Press, Cambridge (2005)

    Book  Google Scholar 

  95. Pearle, P.: Combining stochastic dynamical state-vector reduction with spontaneous localization. Phys. Rev. A 39, 2277–2289 (1989)

    Article  ADS  Google Scholar 

  96. Paul, T., Uribe, A.: On the pointwise behavior of semi-classical measures. Commun. Math. Phys. 175, 229–258 (1996)

    Article  MathSciNet  ADS  MATH  Google Scholar 

  97. Reed, M., Simon, B.: Methods of Modern Mathematical Physics. Vol I. Functional Analysis. Academic Press, San Diego (1972)

    Google Scholar 

  98. Reed, M., Simon, B.: Methods of Modern Mathematical Physics. Vol IV. Analysis of Operators. Academic Press, San Diego (1978)

    Google Scholar 

  99. Robert, D.: Autour de l’Approximation Semi-Classique. Birkhäuser, Basel (1987)

    MATH  Google Scholar 

  100. Rozema, L.A., et al.: Violation of Heisenberg’s measurement-disturbance relationship by weak measurements. Phys. Rev. Lett. 109, 100404 (2012)

    Article  ADS  Google Scholar 

  101. Saunders, S., Barrett, J., Kent, A., Wallace, D. (eds.): Many Worlds? Everett, Quantum Theory, and Reality. Oxford University Press, Oxford (2010)

    MATH  Google Scholar 

  102. Scheibe, E.: The Logical Analysis of Quantum Mechanics. Pergamon, Elmsford (1973)

    Google Scholar 

  103. Schlosshauer, M.A.: Decoherence and the Quantum-to-Classical Transition. Springer, Berlin (2008)

    Google Scholar 

  104. Schrödinger, E.: Die gegenwärtige Situation in der Quantenmechanik. Naturwissenschaften 23, 807–812, 823–828, 844–849 (1935)

    Article  ADS  Google Scholar 

  105. Schubert, R.: Semiclassical behaviour of expectation values in time evolved Lagrangian states for large times. Commun. Math. Phys. 256, 239–254 (2005)

    Article  MathSciNet  ADS  MATH  Google Scholar 

  106. Seevinck, M.P.: Parts and wholes: an inquiry into quantum and classical correlations. PhD Thesis, Utrecht University (2008). philsci-archive.pitt.edu/4583/

  107. Sewell, G.L.: Stability, equilibrium and metastability in statistical mechanics. Phys. Rep. 57, 307–342 (1980)

    Article  MathSciNet  ADS  Google Scholar 

  108. Sewell, G.L.: Quantum Theory of Collective Phenomena. Oxford University Press, Oxford (1986)

    Google Scholar 

  109. Sewell, G.L.: On the mathematical structure of quantum measurement theory. arXiv:math-ph/0505032v2

  110. Shifman, M.: Advanced Topics in Quantum Field Theory. Cambridge University Press, Cambridge (2012)

    MATH  Google Scholar 

  111. Simon, B.: Semiclassical analysis of low lying eigenvalues. IV. The flea on the elephant. J. Funct. Anal. 63, 123–136 (1985)

    Article  MathSciNet  MATH  Google Scholar 

  112. Stairs, A.: Quantum logic, realism, and value definiteness. Philos. Sci. 50, 578–602 (1983)

    Article  MathSciNet  MATH  Google Scholar 

  113. Wallace, D.: Decoherence and its role in the modern measurement problem. arXiv:1111.2187

  114. Wittgenstein, L.: On Certainty. Blackwell, Oxford (1975). Translated by Denis Paul and G.E.M. Anscombe

    Google Scholar 

  115. Zaslavsky, G.M.: Stochasticity in quantum systems. Phys. Rep. 80(3), 157–250 (1981)

    Article  MathSciNet  ADS  Google Scholar 

  116. Zauderer, E.: A uniform asymptotic turning point theory for second order linear ordinary differential equations. Proc. Am. Math. Soc. 31, 489–494 (1972)

    Article  MathSciNet  Google Scholar 

  117. Zurek, W.H.: Decoherence and the transition from quantum to classical. Phys. Today 44(10), 36–44 (1991).

    Article  Google Scholar 

Download references

Acknowledgements

The authors are indebted to Koen Reijnders for help with the appendix. They also wish to thank Jeremy Butterfield, Fred Muller, and an anonymous referee for their penetrating comments on the first draft of this paper, which have improved the current (final) version.Footnote 40

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to N. P. (Klaas) Landsman.

Appendix A: The Flea from wkb

Appendix A: The Flea from wkb

In this appendix, we study the “flea” type perturbation from the point of view of the wkb method of the physics textbooks (like [52, 71]).Footnote 41 As explained in [44, 45], the connection formulae stated in such books are actually correct only for simple potentials like a single well, but with due modifications (see below), the formalism will reproduce both the rigorous and the numerical results described in the main body of this paper.

1.1 A.1 Quantization Condition for an Asymmetric Double Well

We start by recalling some standard wkb formulas. The wkb wave-function in the classically allowed region without turning points (E>V(x)) can be written as

$$ \varPsi(x)\cong\frac{1}{\sqrt{p(x)}} \bigl[Ae^{\frac{i}{\hbar}\int ^{x}{p(y)dy}}+Be^{-\frac{i}{\hbar}\int^{x}{p(y)dy}} \bigr], $$
(A.1)

where

$$ p(x)=\left \{ \begin{array}{l@{\quad}l} \sqrt{ [E-V(x) ]} & \mbox{if } E\geq V(x)\\[4pt] \pm i\sqrt{ [V(x)-E ]} & \mbox{if } E< V(x) . \end{array} \right . $$
(A.2)

A similar formula holds for the classically forbidden region (E<V(x)), namely

$$ \varPsi(x)\cong\frac{1}{\sqrt{|p(x)|}} \bigl[Ce^{-\frac{1}{\hbar }\int ^{x}{|p(y)|dy}}+De^{\frac{1}{\hbar}\int^{x}{|p(y)|dy}} \bigr]. $$
(A.3)

These wave-functions can be connected across turning points via so-called connection formulas, stated in books like [52]. First, we need to distinguish between two kinds of turning points in the usual way: we use the coefficients A l , B l , C l and D l for a left-hand turning point and A r , B r , C r and D r for a right-hand one. The lower limit of the integrals in the above equations is always the coordinate of the turning point. The connection formulas for a left-hand turning point are given by

(A.4)

whilst those for a right-hand turning point are given by

(A.5)

Now consider a general asymmetric double well, as shown in Fig. 9. This figure also introduces part of the notation used.

Fig. 9
figure 9

An asymmetric double-well potential V. The minima are a and b. We assume that the particle has energy E. This provides us with turning points x 1, x 2, x 3 and x 4 and hence with five distinct regions. Four of these regions are named with Roman numerals

We need some more notation for the wkb coefficients used in our calculation. As in (A.1) and (A.3), A,B and C,D denote the coefficients of the wkb wave-function in the classically allowed region and the classically forbidden region, respectively. The number attached to a letter shows to which turning point it belongs, e.g. A 1 and B 1 are the coefficients of the wkb wave-function in region II with respect to x 1 (i.e. x 1 is the lower boundary of the integral in (A.1)). We also need the following three quantities:

$$ \theta_1=\frac{1}{\hbar}\int^{x_2}_{x_1}p(x)dx, \quad\quad \theta_2=\frac{1}{\hbar}\int^{x_4}_{x_3}p(x)dx, \quad\quad K=\frac{1}{\hbar}\int^{x_3}_{x_2}\big|p(x)\big|dx. $$
(A.6)

A final quantity we need is

$$ \tilde{\phi}=\arg{ \biggl[\varGamma \biggl(\frac{1}{2}+i\frac{K}{\pi } \biggr) \biggr]}+\frac{K}{\pi}-\frac{K}{\pi}\ln \biggl( \frac{K}{\pi } \biggr). $$
(A.7)

We are interested in the limit K→∞, since this implies that the barrier is very high and broad, which corresponds to the classical limit ħ→0. Note that \(\tilde{\phi}\rightarrow0\) as K→∞. Our goal is the following quantization condition for the general double well in Fig. 9:

$$ \bigl(1+e^{-2K} \bigr)^{1/2}=\frac{\cos(\theta_1-\theta_2)}{\cos (\theta_1+\theta_2-\pi+\tilde{\phi})} . $$
(A.8)

This condition can be derived in the following way:

  1. 1.

    We start out in region I (coefficients C 1 and D 1). The wave-function needs to be square integrable, so we immediately see that C 1=0.

  2. 2.

    Using the left connection matrix from (A.4), we move to region II (coefficients A 1 and B 1). We can then write the wkb wave-function with respect to x 2 by using

    $$ \left ( \begin{array}{c} A_2\\ B_2\\ \end{array} \right )= \left ( \begin{array}{c@{\quad}c} e^{i\theta_1} & 0\\ 0 & e^{-i\theta_1}\\ \end{array} \right ) \left ( \begin{array}{c} A_1\\ B_1\\ \end{array} \right ), $$
    (A.9)

    which can be proved by changing the lower boundary of the integrals in the wkb wave-function (A.1). The result is

    $$ \left ( \begin{array}{c} A_2\\ B_2\\ \end{array} \right )=e^{i\pi/4} \left ( \begin{array}{c} -i e^{i\theta_1}\\ e^{-i\theta_1}\\ \end{array} \right )D_1. $$
    (A.10)
  3. 3.

    In a similar way, we start in region IV (coefficients C 4 and D 4), and see that D 4=0. After moving to region III with a connection matrix and rewriting the wave-function with respect to x 3, we find

    $$ \left ( \begin{array}{c} A_3\\ B_3\\ \end{array} \right )=e^{i\pi/4} \left ( \begin{array}{c} e^{-i\theta_2}\\ -i e^{i\theta_2}\\ \end{array} \right )C_4 . $$
    (A.11)
  4. 4.

    We now use a result derived in [44] to jump over the barrier and connect the wkb wave-functions in region II and III, viz.Footnote 42

    $$ \left ( \begin{array}{c} A_2\\ B_2\\ \end{array} \right )= \left ( \begin{array}{c@{\quad}c} (1+e^{2K} )^{1/2}e^{-i\tilde{\phi}} & ie^K\\ -ie^K & (1+e^{2K} )^{1/2}e^{i\tilde{\phi}}\\ \end{array} \right ) \left ( \begin{array}{c} A_3\\ B_3\\ \end{array} \right ). $$
    (A.12)
  5. 5.

    Combining the above results (i.e. inserting (A.10) and (A.11) in (A.12)), we find

    (A.13)
    (A.14)
  6. 6.

    The equality of the above two equations leads to the quantization condition (A.8).

As will be discussed in the next two subsections, Eqs. (A.8) and (A.13) have implications for the energy levels and the wave-functions in an asymmetric double well.

1.2 A.2 Energy Splitting in an Asymmetric Double-Well Potential

Assume that for a certain (unperturbed) symmetric double well and given energy E, the constants θ 1 and θ 2 equal some value θ. As in Fig. 9, we introduce a perturbation in the right-hand well. For example, by (A.7), this means that θ=θ 1>θ 2 for a positive perturbation. We therefore write θ 1=θ, θ 2=θδ with δ∈ℝ (e.g. δ>0 in Fig. 9). The quantization condition (A.8) then becomes

$$ \bigl(1+e^{-2K} \bigr)^{1/2}=\frac{\cos(\delta)}{\cos(2\theta -\delta-\pi+\tilde{\phi})}. $$
(A.15)

We can solve for θ, yielding two solutions

$$ \theta_{\pm}=\biggl(n+\frac{1}{2}\biggr)\pi+\frac{1}{2} \delta-\frac {1}{2}\tilde{\phi} \pm \frac{1}{2}\arccos \biggl[ \frac{\cos(\delta)}{ (1+e^{-2K} )^{1/2}} \biggr]. $$
(A.16)

This resembles the original quantization condition \(\theta=(n+\frac{1}{2} )\pi\) for a single well, which is derived using connection formulas in [52]. Here, the energy levels have split up in pairs around the original ones (where the minus sign in (A.16) corresponds to the lower energy by (A.7)). To see what this means, we will examine this equation for two special cases. We first set δ=0 and check if this reproduces known results for a symmetric double well:

$$ \theta_{\pm}=\biggl(n+\frac{1}{2}\biggr)\pi-\frac{1}{2} \tilde{\phi} \pm \frac{1}{2}\arccos \biggl[\frac{1}{ (1+e^{-2K} )^{1/2}} \biggr]. $$
(A.17)

Supposing that K is large, this means that

$$ \theta_{\pm}\approx\biggl(n+\frac{1}{2}\biggr)\pi\pm \frac{1}{2}e^{-K}, $$
(A.18)

since for K large, \(\tilde{\phi}\approx0\) and \(\arccos (\frac{1}{\sqrt{1+x^{2}}} )=\arctan{x}\approx x\) for small x. We find that the energy levels of the single well have split into two. As discussed in [45], this leads exactly to the familiar energy splitting for a symmetric double-well potential stated in texts like [71]. That means that our method for general double wells reproduces known results for a symmetric one. Now that this has been confirmed, let us look at (A.16) in the classical limit K→∞. Solving (A.16) for K→∞ (and so \(\tilde{\phi}\rightarrow0\)) gives

(A.19)
(A.20)

This differs from the symmetric well, which for K→∞ gives a twofold degeneracy for each energy level labeled by n. Equation (A.20) can be understood in the following way: in the classical limit, tunneling is suppressed. Therefore, the particle is localized in one of the wells, where it obeys the familiar quantization condition for a single well. If it is in the left well, then \(\theta_{1}=(n+\frac{1}{2})\pi=\theta_{-}\), but if it is in the right well, we have \(\theta_{2}=(n+\frac{1}{2})\pi =\theta_{+} -\delta\).

1.3 A.3 Localization in an Asymmetric Double-Well Potential

Now that we have analyzed the behaviour of the energy splitting, we turn to the wkb wave-function. With the notation used in the previous section, (A.13) leads to

$$ \frac{D_1}{C_4} =i \bigl[ \bigl(1+e^{2K} \bigr)^{1/2}e^{-i(2\theta _{\pm}-\delta+\tilde{\phi})}+e^Ke^{-i\delta} \bigr]. $$
(A.21)

Inserting (A.16), the reader can check that for δ∈[−π,π] one has

$$ \frac{D_1}{C_4} =\sin(\delta)e^{K} \mp\sqrt{\sin^2( \delta )e^{2K}+1}. $$
(A.22)

This allows us to derive localization of the wkb wave-function in the classical limit K→∞. As can be seen from (A.10), D 1 is a measure of the amplitude of the wkb wave-function in regions I and II in Fig. 9. In a similar way, (A.11) shows that C 4 is a measure of the amplitude of the wkb wave-function in regions III and IV. Therefore, the fraction D 1/C 4 indicates whether the wave-function is localized, and if so, where. Doing the same calculation again for δ∈[π,3π] gives the above result multiplied by −1. Of course, this can be generalized: for n∈ℤ and δ∈[(2n−1)π,(2n+1)π], the result (A.22) is correct for n even and should be multiplied by −1 for n odd. This will not affect our conclusions, as we will see. We consider some cases and check what (A.22) tells us:

  • For δ=0 (no perturbation), we find that \(\frac {D_{1}}{C_{4}}=\mp1\). The general double well has pairs of energy levels (labeled by n). Such a pair consists of a lower and higher lying level, corresponding to θ and θ + in (A.16), respectively. Here, we see that for the lower level D 1=C 4, i.e. the wkb wave-function is even. However, for the higher level we find D 1=−C 4, which means the wkb wave-function is odd. This is a well-known fact and it is nice to see our method reproducing it. Note that this conclusion is not only independent of n, but also of K, as expected.

  • For δ>0,δ∉{|k∈ℤ} (which corresponds to a positive perturbation in the right well, e.g. the potential in Fig. 9), we find, in the limit K→∞, that:

    $$ \frac{D_1}{C_4}\longrightarrow \left \{ \begin{array}{l@{\quad}l} \infty& \mbox{for $\theta_{-}$ in (A.16)}\quad \mbox{(lower energy)}\\[2pt] 0 & \mbox{for $\theta_{+}$ in (A.16)}\quad \mbox{(higher energy)} . \end{array} \right . $$

    Hence for low (high) energy, the wkb wave-function is localized on the left (right).

  • For δ<0, δ∉{|k∈ℤ}, i.e., a negative perturbation in the right well, we find

    $$ \frac{D_1}{C_4}\longrightarrow \left \{ \begin{array}{l@{\quad}l} 0 & \mbox{for $\theta_{-}$ in (A.16)}\quad \mbox{(lower energy)}\\ [2pt] \infty& \mbox{for $\theta_{+}$ in (A.16)}\quad \mbox{(higher energy)}. \end{array} \right . $$

    For the lower (higher) energy, the wkb wave-function is localized on the right (left).

  • For δ∈{|k∈ℤ∖{0}}, something peculiar happens, in that either \(\frac {D_{1}}{C_{4}}=\pm1\) or \(\frac{D_{1}}{C_{4}}=\mp1\). This implies that no localization takes place.Footnote 43

  • So far, we have interpreted δ as the result of a perturbation in the right well. However, our approach allows us to interpret a positive perturbation in the right-hand well as a negative one in the left-hand well, and vice versa. Therefore, the above results change places if we put the perturbation in the left-hand well.

Our method produces the results we would expect. However, to be precise, the above reasoning needs to be amended as follows. We have treated δ as a constant, but in reality it depends on K. The reason for this is that K affects θ 1 and θ 2, and therefore δ=θ 1θ 2, via the quantization condition. Now consider a fixed energy level (i.e. fixed n and fixed sign ± in (A.16)) in a given double-well potential that has a perturbation in one of the wells. In the limit of completely decoupled wells (K→∞), we know this energy level has some fixed limit higher than the minimum of the potential. As long as the perturbation is below this energy level, we know that θ 1θ 2≠0 by (A.7). This means that there exists some K 0 such that |θ 1θ 2|≠0 for any K>K 0. We may then apply the above reasoning to verify that our conclusions about localization are still correct.Footnote 44

Rights and permissions

Reprints and permissions

About this article

Cite this article

(Klaas) Landsman, N.P., Reuvers, R. A Flea on Schrödinger’s Cat. Found Phys 43, 373–407 (2013). https://doi.org/10.1007/s10701-013-9700-1

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10701-013-9700-1

Keywords

Navigation