Skip to main content

The de Broglie–Bohm Theory

  • Chapter
  • First Online:
Making Sense of Quantum Mechanics

Abstract

The de Broglie-Bohm theory, which is a deterministic theory of matter in motion, is explained in this chapter. We first show what the trajectories look like and, then, how to derive the usual quantum predictions, both for the positions of particles and for other observables, using spin and momentum as examples. This derivation rests on the assumption of quantum equilibrium, which we shall explain and discuss. We also see how nonlocality appears in the de Broglie-Bohm theory and we shall present and answer several objections to that theory.

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

Access this chapter

Chapter
USD 29.95
Price excludes VAT (USA)
  • Available as PDF
  • Read on any device
  • Instant download
  • Own it forever
eBook
USD 49.99
Price excludes VAT (USA)
  • Available as EPUB and PDF
  • Read on any device
  • Instant download
  • Own it forever
Softcover Book
USD 64.99
Price excludes VAT (USA)
  • Compact, lightweight edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info
Hardcover Book
USD 69.99
Price excludes VAT (USA)
  • Durable hardcover edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info

Tax calculation will be finalised at checkout

Purchases are for personal use only

Institutional subscriptions

Notes

  1. 1.

    In this chapter, we will discuss only the non-relativistic de Broglie–Bohm theory. The issue of relativity (and of quantum fields) will be addressed in Sect. 5.2.2. By “relativity”, we mean the special theory of relativity, unless the general one is mentioned.

  2. 2.

    We saw such claims made in Chap. 1 and we will discuss them again in Chap. 7.

  3. 3.

    See, e.g., [8, 9, 474] for elementary introductions and [24, 55, 70, 152, 153, 231, 268, 359, 473] for more advanced ones. There are also pedagogical videos made by students in Munich, available at: https://cast.itunes.uni-muenchen.de/vod/playlists/URqb5J7RBr.html.

  4. 4.

    In Sects. 5.1.3, 5.1.4, and 5.2.1, we will use the presentation of the de Broglie–Bohm theory in Chap. 7 of David Albert’s book Quantum Mechanics and Experience [8].

  5. 5.

    We let N be the number of variables associated with the system. So if the system consists of M particles moving in three-dimensional space, we have \(N=3M\).

  6. 6.

    We will call below the pair \(\big (\Psi (t), \mathbf{X}(t)\big )\) the state of the system, to be distinguished from the quantum state \(\Psi (t)\), which is only part of it.

  7. 7.

    The equations here are (partial) differential equations, similar to those discussed in Appendix 2.A. The reader unfamiliar with such equations can skip them and proceed to the remarks. The consequences of these equations will be discussed and illustrated later.

  8. 8.

    In this chapter, we set \(\hbar =1\). See Appendix 2.A for more details on Schrödinger’s equation .

  9. 9.

    Of course, in that case, the Schrödinger equation must be replaced by Pauli’s equation , but we will not discuss that (see, e.g., [152, Sect. 8.4]).

  10. 10.

    The reader familiar with quantum mechanics will recognize the right-hand side of (5.1.1.4) as the quantum probability current divided by \(|\Psi |^2\). This formula is derived in Appendix 5.A.

  11. 11.

    A point that is clearly misunderstood by some, e.g., Gisin [214, p. 46]: “If there are hidden positions, there must therefore also be hidden velocities. But that contradicts Heisenberg’s uncertainty principle which is a key part of the quantum formalism [\(\ldots \)].”.

  12. 12.

    See, for example, Gouesbet [237, Chap. 17], for an explicit objection to the de Broglie–Bohm theory on that basis.

  13. 13.

    Setting aside the problems coming from the self-interaction of the particle with its own field.

  14. 14.

    However, if the wave function is a plane wave, viz., \(\Psi (x)= \exp (ikx)\), a rather idealized situation, then (5.1.1.3) yields straight line motion with velocity k / m. We will see in the next two subsections that this violation of Newton’s first law does not contradict conservation of momentum , which also holds in quantum mechanics. Indeed, this conservation law concerns results of measurements and the de Broglie–Bohm theory will predict, in a subtle but natural way, the same results of measurements as ordinary quantum mechanics.

  15. 15.

    For a picture of that potential, see [70, p. 34].

  16. 16.

    See [151, 362, 522] for a discussion of weak measurements in the context of the de Broglie–Bohm theory.

  17. 17.

    This does not contradict Heisenberg’s uncertainty relations , because we need to do many operations and take an average to get this result. Heisenberg’s uncertainty relations apply to the results of strong, i.e., ordinary, measurements.

  18. 18.

    Those theories will be discussed in Sect. 5.4.1.

  19. 19.

    See [386, pp. 272–273].

  20. 20.

    To be mathematically precise, one should add “measurable” or “Borel” here.

  21. 21.

    Here we use the fact that a function can be defined by giving the value of its integral on every (Borel) subset, if the measure defined by those integrals is absolutely continuous, and we assume here that the map \(\phi ^t({\mathbf{x}})\) is such that this measure has this property.

  22. 22.

    To be precise, this is rather a consequence of what is called equivariance in [141]. See the second part of that appendix.

  23. 23.

    See Appendix 2.C.

  24. 24.

    We will take for granted in this subsection that one can detect the positions of particles, and we will see how the de Broglie–Bohm theory accounts for the usual quantum predictions, given that assumption. However, this detection does of course necessitate the coupling of the particle with a macroscopic device, and that coupling, described in conventional terms in Appendix 2.D, will be explained in the context of the de Broglie–Bohm theory in Sect. 5.1.6 and Appendix 5.E.

  25. 25.

    It should be stressed that all the “experiments” are only meant to illustrate the theory, not to explain how real experiments are performed.

  26. 26.

    The time evolution is given by Pauli’s equation rather than Schrödinger’s equation . See [152, Sect. 8.4] or [360] for a detailed discussion. The solution given here is similar to the one in Appendix 2.D, except that here we consider the evolution of the wave function of the particle, while in that appendix we considered the evolution of the wave function of the pointer .

  27. 27.

    One can check that the presence of a magnetic field does not destroy the symmetry \(\Psi ( z)=\Psi ( -z)\).

  28. 28.

    In that part of the figure, as well as in Fig. 5.6, the holes are not put in a natural way (top and bottom), but that is just a convention, chosen for graphical convenience.

  29. 29.

    To be more precise, “blocking” here has to be understood as a collapse of the quantum state, whose meaning, in the de Broglie–Bohm theory, will be explained in the next subsection.

  30. 30.

    See [38, 69] for further discussion of the “delayed choice” experiment in the de Broglie–Bohm theory.

  31. 31.

    See Sect. 2.2 and Appendix 2.D.

  32. 32.

    Englert, Scully, Sussman, and Walther have suggested that the de Broglie–Bohm trajectories are “surrealistic” rather than realistic [175], because those trajectories can have counterintuitive aspects, like the ones discussed here. Originally, the situation described in [175] was presented as an argument against the existence of de Broglie–Bohm trajectories, but in fact those “surrealistic” trajectories can be perfectly understood, in a natural way, within the de Broglie–Bohm theory. See [28, 130, 143, 176, 265] for further discussion of these trajectories.

  33. 33.

    In our language, where we speak of electrons rather than photons, it would mean that the particles are detected after the arrow with spin up and spin down in the direction 1 half of the time. (Note by J.B.).

  34. 34.

    The word “contextual” sometimes has a different meaning, as we saw in Appendix 2.F, namely that the result of a measurement may depend on which other measurement is performed simultaneously with the first one. But since one sees here that the result of the “measurement” depends on the details of the apparatus, one can understand that it may also depend on whether another measurement is performed simultaneously. We will come back to this notion of contextuality in Sects. 5.1.5 and 5.3.4.

  35. 35.

    Within the de Broglie–Bohm theory, one can introduce, if one wants, “hidden” spin variables, but their value is a function of the quantum state and the positions of the particles (hence, they are redundant) and the actually measured spin values will be affected by the measuring devices and thus not pre-exist their “measurement”. See [268, Chap. 9] or [70, Chap. 10] and references therein for more details on these spin variables.

  36. 36.

    This is only approximate (hence the use of the symbol \(\approx \) below), because of the spreading of the wave functions under free evolution, which implies that, even if the initial wave function has a bounded support, this will not be true at later times. One says that the time evolution produces “tails” of the wave function, namely regions where it is small but nonzero. However, in practice, the probability of finding the particle in one of the tails of the wave function is exceedingly small.

  37. 37.

    Or, which amounts to the same, with a phase S that is independent of \((x_1,\ldots , x_N)\). Ground state wave functions are, in general, real.

  38. 38.

    See Appendix 2.C.

  39. 39.

    In [70, Sect. 6.3], Bohm and Hiley give a general analysis of such an interaction. We will discuss only a concrete example.

  40. 40.

    For another reply, see Gondran and Hoblos [234, 235] who, following Born [77], consider a non-stationary initial wave function and then obtain a de Broglie–Bohm motion close to the classical one.

  41. 41.

    However, one can even have measurements of the position operator that do not measure positions. See [147, Sect. 7.5] for an example of this situation.

  42. 42.

    See [147] for a more detailed discussion.

  43. 43.

    The last sentence is needed because in ordinary quantum mechanics the exact time when the collapse occurs is not well defined. It is defined by the “observation” but, as we discussed before, that notion itself is imprecise.

  44. 44.

    See Bell [38] for a discussion, similar to the one given here, of the effective collapse of the quantum state in the de Broglie–Bohm theory, but in the example of the double-slit experiment.

  45. 45.

    It is sometimes thought that, for such an effective collapse to occur, one needs the measuring device to interact with an environment, such as the air molecules surrounding it, and ultimately the entire universe . But that is not true: any sufficiently macroscopic device suffices, even if the latter were perfectly isolated from the rest of the universe (which is never the case, but that is not relevant since what we say would be true even if perfect isolation were possible).

  46. 46.

    For more details, see, e.g., [80, 225, 307].

  47. 47.

    There is also the crucial question, which we put aside for now, of what it could mean for the world to be a wave function (see Sect. 6.1).

  48. 48.

    See [185, Chap. 5] and [389, Chap. 7] for pedagogical discussions of this idea.

  49. 49.

    Assuming that the eigenvalue that is observed is non-degenerate .

  50. 50.

    We follow the notation of [141] and do not use boldface letters, even though xyXY generally belong to a space of more than one dimension. This is because we will use boldface letters below for coordinates of several copies of the same system.

  51. 51.

    In the situations where the decomposition (5.1.7.1) does not hold, there is a more general notion, the conditional wave function [141]:

    $$ \Psi (x)=\Psi (x,Y)\;, $$

    where x are the generic variables of the system and Y is the actual position of the environment. Of course, \(\Psi (x,Y)=\psi (x) \Phi (Y)\) when the decomposition (5.1.7.1) holds.

  52. 52.

    We have \(P(X =x|Y=y)= P(X=x,Y=y)/P(Y=y)\), and \(P(Y=y)= \int P(X=x,Y=y)dx = \int |\Psi (x,y)|^2 dx\). By (5.1.7.1), if \(y\in {\mathrm{supp}\,}\Phi \), the latter is equal to \(\int | \psi (x)\Phi (y)|^2 dx= |\Phi (y)|^2\), since \(\int | \psi (x)|^2 dx=1\). So finally, \(P(X=x,Y=y)/P(Y=y)=| \psi (x)\Phi (y)|^2/|\Phi (y)|^2=|\psi (x)|^2\).

  53. 53.

    Nor necessarily at the same time either, although the treatment of measurements at different times is more subtle, see [141, Sects. 8–10].

  54. 54.

    See Sects. 3.4.4 and 3.4.5 and Appendix 3.A for a definition of those notions.

  55. 55.

    This convergence to equilibrium cannot be true in full generality since, for any real wave function \(\Psi ({\mathbf{x}}) \) nothing moves, hence all distributions are stationary in that situation and do not converge to anything. There have also been some attempts to prove convergence to equilibrium theoretically [67, 481–484].

  56. 56.

    See [484, 485, 488] for a discussion of a possible non-equilibrium distribution in the early universe.

  57. 57.

    See Appendix 2.C.

  58. 58.

    Indeed, if we perform, say, a position measurement and, as a result we obtain a wave function with a relatively narrow support \(\psi (x)\), then, of course, a subsequent measurement of momentum will start from that wave function and will have a variance \(\text{ Var }(p)\) related by (2.C.1.4) to the variance \(\text{ Var }(x)\) of the original x measurement (see Appendix 2.C for the computation of those variances for Gaussian wave functions). On the other hand, if we consider the example of a particle in a box , discussed in Sect. 5.1.4 and Appendix 5.D, the initial momentum is zero and therefore has a variance equal to zero. But, since what we “measure” is not that initial momentum but the asymptotic position of the particle after removing the walls of the box, and thus setting the particle in motion, the result of that “measurement” will coincide with the ordinary quantum predictions (see Appendix 5.D), and will therefore also satisfy the Heisenberg uncertainty relations.

  59. 59.

    See [141, Sects. 12–14] for more details.

  60. 60.

    At least according to Wigner , who told the story as though von Neumann was his “friend”, see [516, Note 1, p. 1009].

  61. 61.

    Here, the probability refers to the Lebesgue measure on the interval [0, 1[, which amounts to considering all the symbols \(a_i\) to be independently distributed and giving equal probability to each possible value of \( a_i= 0, \dots , 9\).

  62. 62.

    Of course, this assumes a perfect symmetry between the two parts of Fig. 5.11.

  63. 63.

    The Schrödinger equation (5.1.1.1) is also first order in time, but that does not imply that the paths followed by the wave functions cannot cross each other. Indeed, in order for that crossing to be forbidden, one would need the right-hand side of (5.1.1.2) to coincide for the two wave functions when they cross, i.e., the wave functions and their second derivatives would have to coincide, and there is no reason why that should happen.

  64. 64.

    The fact that a non-equilibrium distribution would allow us to send messages instantaneously is in fact quite general [486].

  65. 65.

    And, because of the relativity of simultaneity , discussed in the next subsection, one could in principle send them into one’s own past, at least when one does not introduce an additional structure, such as a preferred foliation into spacetime. See [319, Chap. 4] and [48] for a discussion of the relationship between nonlocality and the sending of messages.

  66. 66.

    Readers may wonder what “first” and “later” mean in a relativistic framework. This will be discussed in the next subsection.

  67. 67.

    See [37, 158, 208], [70, p. 139], and [319, Chap. 4] for a general argument.

  68. 68.

    This is not the way quantum field theory is usually presented, but it allows us to make a straightforward connection with ordinary quantum mechanics.

  69. 69.

    See [145, 146, 148, 150]. For reviews and references to the original papers, see [464, 465, 477].

  70. 70.

    See, for example, [216] for an introduction to a rigorous approach to quantum field theories. One can always put a cutoff on the range of momenta to which the theory applies in order to make it well defined, although that trick does not answer the real question, namely, what happens if one removes the cutoff.

  71. 71.

    This is, of course, an idealization, but it is good enough to allow us to identify approximate inertial frames of reference.

  72. 72.

    This is because the speed of light enters into those equations. But then the equations cannot be invariant under Galilean transformations, since those transformations change the speed of any object when one passes from one frame to another.

  73. 73.

    For a good introduction to the theory of relativity, see [469], and for a careful conceptual discussion, see [326].

  74. 74.

    Under the Lorentz transformations, the speed of light is constant and the laws of electromagnetism are invariant.

  75. 75.

    At least, since one cannot check experimentally that this simultaneity is absolute, much faster than the speed of light .

  76. 76.

    Technically, this is called a foliation of spacetime.

  77. 77.

    For example, when one derives conservation laws , such as conservation of energy , momentum , and angular momentum , one always assumes that the systems to which those laws apply are isolated, otherwise there would be no way to derive those laws. But if the conservation laws held only for the universe, they would have no practical applications.

  78. 78.

    In (2.A.2.24), one sees that the larger the mass, the slower the spreading, since time enters in (2.A.2.24) only through the factor t / m.

  79. 79.

    For particles in the box, this corresponds to large index n in the wave functions \(\Psi _n\) (see Appendix 5.D).

  80. 80.

    Here are some remarks that may help to understand the quote. The many-worlds theory will be discussed in Sect. 6.1. The von Neumann theorem will be discussed in Sect. 7.4. The Kochen –Specker theorem is a version of the theorem in Sect. 2.5. The curved trajectories in the absence of forces are illustrated in Fig. 5.1. The stationary quantum states referred to here are those were the wave function is real, whence the particles do not move, according to (5.1.1.3). In Appendix 5.D, we give examples of such states, even with high energies.

  81. 81.

    Since their positions vary with time, they also have velocities but, as we discussed in Sect. 5.1.4, when one “measures” the “momentum ”, one does not get the actual value of those velocities. So in the de Broglie–Bohm theory, there are no “hidden variables ” for the momenta. We will discuss this important property of the de Broglie–Bohm theory again in Sect. 5.3.3.

  82. 82.

    Like the particles in the box whose states are discussed in Appendix 5.D, or certain electrons , whose wave function is also real, for example, those in an atomic “ground state”.

  83. 83.

    See Appendix 5.A for the definition of the probability current. (Note by J.B.).

  84. 84.

    See [344] for a detailed discussion of this objection, and see also [89].

  85. 85.

    This holds for any given time t, which we suppress here in the arguments of \(\Psi (x)\) and \(\hat{\Psi }(p)\).

  86. 86.

    See Appendix 2.C.2.

  87. 87.

    Of course, both the Hamiltonian and the quantum state depend on the existence and the properties of the particles via the potential term in the Hamiltonian or in the Schrödinger equation .

  88. 88.

    Writing \(\Psi =Re^{iS}\), we get \(\log \Psi = \log R+ iS\) and \(|\exp (2 \log \Psi )|=R^2 =|\Psi |^2\).

  89. 89.

    The title of the next subsection was inspired by [89].

  90. 90.

    This follows from the observation that, if one adds any term of the form \({\mathbf{j}}({\mathbf{x}}, t)/|\Psi ({\mathbf{x}}, t)) |^2\) to the right-hand side of (5.1.1.4), where \(\nabla \cdot {\mathbf{j}}=0\), then the \(|\Psi ({\mathbf{x}}, t)|^2\) distribution density remains equivariant, because the left-hand side of (5.A.1) in Appendix 5.A is not modified by the addition of a current of the form (5.A.6) with \(\nabla \cdot {\mathbf{j}}=0\).

  91. 91.

    See [464, 465] for a summary of the different possibilities and for references to the original papers.

  92. 92.

    Here we introduce a distinction between bosons and fermions which is essential in quantum field theory, but which is not discussed elsewhere in this book. Without entering into a detailed discussion, one can think of bosons, e.g., photons , as mediating interactions between particles, and fermions, e.g., electrons , protons, and neutrons, as constituents of ordinary matter.

  93. 93.

    In [145, 146, 148, 150], one introduces a stochastic process with creation of pairs of particles at random times and a deterministic evolution in-between. In [99], one introduces a deterministic model with a “Dirac sea”.

  94. 94.

    One could also introduce only fermionic variables and not bosonic ones. This is actually Bell’s approach in [42].

  95. 95.

    For a discussion of the situation for relativistic de Broglie–Bohm theories, see [269, 270].

  96. 96.

    This was in Born’s book Natural Philosophy of Cause and Chance [76], since Bell did not read German and von Neumann’s book [496] was only translated into English in 1955. We will discuss Born and von Neumann’s views in Sects. 7.2 and 7.4.

  97. 97.

    This equation was first written by Erwin Madelung , who gave it a “hydrodynamic” interpretation [315, 316].

  98. 98.

    This property is what is called equivariance in [141].

  99. 99.

    We assume, as in (5.1.3.1), that the integrals (5.C.1.4) are absolutely continuous, so that they define the function \(\rho ({\mathbf{x}},t)\). Equation (5.C.1.4) is similar to (5.1.3.1), but here it applies to a probability distribution on \( \mathbb {R}^N\), with N the number of variables of the system, whereas in (5.1.3.1) it referred to the empirical density in \( \mathbb {R}^3\). These concepts are different, even if the formulas look similar. The connection between them is made at the end of this appendix.

  100. 100.

    We consider here each \(x_i \in \mathbb {R}\), but one could also consider \({\mathbf{x}}_i \in \mathbb {R}^3\).

  101. 101.

    This wave function could be the effective wave function of those non-interacting identical systems, defined by (5.1.7.5).

  102. 102.

    For more details on this example and on similar ones, see Holland [268, Chap. 4] or Bohm and Hiley [70, Sect. 3.6].

  103. 103.

    This Appendix is based on [268, Sect. 6.5].

  104. 104.

    Of course, these are eigenvectors of the Hamiltonian , and the corresponding E are its eigenvalues .

  105. 105.

    The theory of Fourier series implies that the set \(\big (\Psi _n(x)\big )^\infty _{n=1}\) forms a basis of \(L^2([0,L],dx)\), but we will not use this fact.

  106. 106.

    This may seem artificial as a limit, but if we put back factors of \(\hbar \) in our equations, it simply means that we take L large compared to the de Broglie wavelength of the particle.

  107. 107.

    See Sect. 5.1.4 and [172].

  108. 108.

    One can also check that, in this limiting situation, the acceleration of the particles vanishes, since the velocity is given by \(k_n\), and the quantum potential defined by (5.B.6) is constant, whence its contribution to (5.1.1.5) vanishes.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Jean Bricmont .

Appendices

Appendices

5.A The Continuity Equation for the \(|\Psi |^2\) Distribution

Let us consider Schrödinger’s equation (5.1.1.1) with H given by (5.1.1.2). We will first prove the following continuity equation:

$$\begin{aligned} \frac{\partial |\Psi ({\mathbf{x}}, t)|^2}{ \partial t} + \nabla \cdot {\mathbf{j}}({\mathbf{x}}, t)= 0\;, \end{aligned}$$
(5.A.1)

where \(\nabla = (\partial / \partial x_1, \ldots , \partial / \partial x_N)\) is the gradient, the dot denotes the scalar product, and \({\mathbf{j}}({\mathbf{x}},t)= (j_k ({\mathbf{x}}, t))_{k=1}^N\) (we index the components here by k instead of j) is the “probability current” or the “quantum flux” defined by

$$\begin{aligned} j_k ({\mathbf{x}}, t)&= \frac{1}{ 2i m_k } \left[ \Psi ^*({\mathbf{x}}, t) \frac{\partial }{ \partial x_k} \Psi ({\mathbf{x}}, t) - \Psi ({\mathbf{x}}, t)\frac{\partial }{ \partial x_{k}} \Psi ^*({\mathbf{x}}, t)\right] \\&= \frac{1}{m_k } \mathrm{Im}\left[ \Psi ^*({\mathbf{x}}, t)\frac{\partial }{ \partial x_k}\Psi ({\mathbf{x}}, t)\right] \;. \end{aligned}$$
(5.A.2)

To prove (5.A.1), write

$$\begin{aligned} \frac{\partial |\Psi ({\mathbf{x}}, t)|^2}{ \partial t} = \Psi ^*({\mathbf{x}}, t)\frac{\partial \Psi ({\mathbf{x}}, t)}{ \partial t} + \Psi ({\mathbf{x}}, t) \frac{\partial \Psi ^*({\mathbf{x}}, t)}{ \partial t}\;, \end{aligned}$$
(5.A.3)

take the complex conjugate of (5.1.1.1), viz.,

$$\begin{aligned} -i \frac{\partial \Psi ^*}{ \partial t} ({\mathbf{x}}, t) = H \Psi ^*({\mathbf{x}}, t)\;, \end{aligned}$$
(5.A.4)

and insert this and (5.1.1.1) into (5.A.3). The potential term cancels out and we get

$$\begin{aligned} \frac{\partial |\Psi ({\mathbf{x}}, t)|^2}{ \partial t} = -\frac{1}{ i} \sum ^N_{k=1} \frac{1}{ 2 m_k} \left[ \Psi ^*({\mathbf{x}}, t) \frac{\partial ^2 }{ \partial x_k^2} \Psi ({\mathbf{x}}, t) - \Psi ({\mathbf{x}}, t) \frac{\partial ^2 }{ \partial x_k^2} \Psi ^*({\mathbf{x}}, t)\right] \;. \end{aligned}$$
(5.A.5)

It is enough to observe that

$$\begin{aligned}&\frac{\partial }{ \partial x_k} \left[ \Psi ^*({\mathbf{x}}, t) \frac{\partial }{ \partial x_k} \Psi ({\mathbf{x}}, t) - \Psi ({\mathbf{x}}, t) \frac{\partial }{ \partial x_k} \Psi ^*({\mathbf{x}}, t)\right] =\Psi ^*({\mathbf{x}}, t) \frac{\partial ^2 }{ \partial x_k^2} \Psi ({\mathbf{x}}, t) \\&\quad - \Psi ({\mathbf{x}}, t)\frac{\partial ^2 }{ \partial x_k^2} \Psi ^*({\mathbf{x}}, t) \end{aligned}$$

to obtain (5.A.1).

Now, if we consider (5.1.1.4), we see that the function \({\mathbf{j}}({\mathbf{x}},t)= (j_k ({\mathbf{x}}, t))_{k=1}^N\) [whose arguments are the generic variables \({\mathbf{x}}\) rather than the actual positions of the particles as in (5.1.1.4)] is of the form

$$\begin{aligned} {\mathbf{j}}({\mathbf{x}})={\mathbf{V}}_{\Psi } ({\mathbf{x}}, t) \ |\Psi ({\mathbf{x}}, t)|^2\;, \end{aligned}$$
(5.A.6)

and we can write (5.A.1) as

$$\begin{aligned} \frac{\partial |\Psi ({\mathbf{x}},t)|^2}{ \partial t} + \nabla \cdot \Big [{\mathbf{V}}_{\Psi }({\mathbf{x}}, t) |\Psi ({\mathbf{x}},t) |^2\Big ]=0\;, \end{aligned}$$
(5.A.7)

which is the continuity equationFootnote 97 one would obtain for a “fluid” of density \(\rho ({\mathbf{x}},t)=|\Psi ({\mathbf{x}},t)|^2\), with local velocity \({\mathbf{V}}_{\Psi } ({\mathbf{x}},t)\). This will be useful when we come to prove the equivariance of the \(|\Psi ({\mathbf{x}},t)|^2\) distribution in Appendix 5.C.

5.B A Second Order Dynamics and the Quantum Potential

Here we derive (5.1.1.5) from (5.1.1.1)–(5.1.1.3). We insert \(\Psi ({\mathbf{x}},t)=R({\mathbf{x}},t)e^{iS({\mathbf{x}},t)}\) into Schrödinger’s equation (5.1.1.1), where \({\mathbf{x}}= (x_1, \dots , x_N)\). Suppressing here the arguments \(({\mathbf{x}}, t)\), we get

$$\begin{aligned} \frac{i \partial R}{ \partial t} e^{iS} -\frac{\partial S}{\partial t} \Psi = - \sum _{j=1}^N \frac{1}{ 2 m_j}\left[ \frac{\partial ^2 R}{ \partial x_j^2} - R \left( \frac{\partial S}{ \partial x_j}\right) ^2 +2i \frac{\partial R}{ \partial x_j} \frac{\partial S}{ \partial x_j} + iR\frac{\partial ^2 S}{ \partial x_j^2} \right] e^{iS} + V\Psi \;. \end{aligned}$$
(5.B.1)

Dividing both sides by \(e^{iS}\) and separating the real and imaginary parts give rise to two equations:

$$\begin{aligned}&\frac{R\partial S}{ \partial t} = \sum _{j=1}^N \frac{1}{ 2 m_j} \left[ \frac{\partial ^2R}{ \partial x_j^2} - R\left( \frac{\partial S }{ \partial x_j}\right) ^2 \right] - VR \;,\end{aligned}$$
(5.B.2)
$$\begin{aligned}&\frac{\partial R}{ \partial t} + \sum _{j=1}^N \frac{1}{ 2 m_j} \left[ 2 \frac{\partial R}{ \partial x_j} \frac{\partial S}{ \partial x_j} + R\frac{\partial ^2 S}{ \partial x_j^2} \right] =0\;. \end{aligned}$$
(5.B.3)

Dividing (5.B.2) by R gives

$$\begin{aligned} \frac{\partial S}{ \partial t} = \sum _{j=1}^N \frac{1}{ 2 m_j} \left[ \frac{1}{ R} \frac{\partial ^2 R}{ \partial x_j^2} - \left( \frac{\partial S}{ \partial x_j}\right) ^2\right] - V\;. \end{aligned}$$
(5.B.4)

Multiplying (5.B.3) by 2R gives

$$ \frac{\partial R^2}{ \partial t} + \sum _{j=1}^N \frac{1}{ m_j} \frac{\partial }{ \partial x_j} \left( R^2 \frac{\partial S}{ \partial x_j} \right) =0\;, $$

which, with \(R^2=|\Psi |^2\) and using (5.A.6), is the same as (5.A.1).

Now, if we consider (5.1.1.3) and take its time derivative, remembering that S depends on time through \(\big (X_1(t),\ldots ,X_N(t)\big )\) and t, we get

$$\begin{aligned} m_k\frac{d^2}{ dt^2} X_k (t) = \sum _{j=1}^N \frac{\partial ^2 S}{ \partial x_j \partial x_k} \dot{X}_j + \frac{\partial ^2S}{\partial x_k \partial t}\;. \end{aligned}$$
(5.B.5)

If we use (5.1.1.3) to replace \(\dot{X}_j\) by \((1/m_j)(\partial S/\partial x_j)\) in (5.B.5), and insert (5.B.4) in the second term of (5.B.5), evaluating the functions at \(\mathbf{X}\) instead of \(\mathbf{x}\), the terms

$$ \sum _{j=1}^N \frac{1}{ m_j}\frac{\partial ^2 S}{ \partial x_j \partial x_k} \frac{\partial }{\partial x_j} S \quad \text{ and }\quad -\frac{\partial }{\partial x_k} \sum _{j=1}^N \frac{1}{2 m_k} \left( \frac{\partial S}{\partial x_j}\right) ^2 $$

cancel each other and we get

$$\begin{aligned} m_k\frac{d^2}{ dt^2} X_k (t) =-\frac{\partial }{\partial x_k} (Q+V)\;, \end{aligned}$$

where the quantum potential Q is given by

$$\begin{aligned} Q=- \sum _{j=1}^N \frac{1}{ 2 m_j} \frac{1}{R} \frac{\partial ^2 R}{\partial x_j^2}\;. \end{aligned}$$
(5.B.6)

This proves (5.1.1.5).

5.C Proof of the Equivariance of the \(|\Psi ({\mathbf{x}},t)|^2\) Distribution

5.1.1 5.C.1 Equivariance in \(\mathbb {R}^N\)

We will first proveFootnote 98 that, for a system in \(\mathbb {R}^N\) and \(\forall A \subset \mathbb {R}^N\),

$$\begin{aligned} P\big ({\mathbf{X}}(t)\in A\big )=\int _A |\Psi ({\mathbf{x}},t)|^2 d{\mathbf{x}}\;, \end{aligned}$$
(5.C.1.1)

provided that, at \(t=0\), \(\forall A \subset \mathbb {R}^N\),

$$\begin{aligned} P\big ({\mathbf{X}}(0)\in A\big ) = \int _A |\Psi ({\mathbf{x}},0)|^2 d{\mathbf{x}}\;. \end{aligned}$$
(5.C.1.2)

In (5.C.1.1), the left-hand side involves \({\mathbf{X}}(t)\), the solution of (5.1.1.4), while on the right-hand side, \(\Psi ({\mathbf{x}},t)\) is the solution of Schrödinger’s equation (5.1.1.1).

Consider first a general differential equation

$$\begin{aligned} \frac{d{\mathbf{X}}(t)}{ dt} = {\mathbf{V}}\big ({\mathbf{X}}(t),t\big )\;, \end{aligned}$$
(5.C.1.3)

with \({\mathbf{X}}:\mathbb {R}\rightarrow \mathbb {R}^N\). We assume that \({\mathbf{V}}\) is such that (5.C.1.3) has solutions for all times. Let \(\phi ^t({\mathbf{x}})\) denote the solution at time t of (5.C.1.3) with initial condition \({\mathbf{X}}(0)= {\mathbf{x}}\) at time 0. Then, given any probability density \(\rho ({\mathbf{x}}, 0)\) on the initial conditions \({\mathbf{X}}(0)={\mathbf{x}}\), defineFootnote 99 \(\rho ({\mathbf{x}},t)\), \(\forall A \subset \mathbb {R}^N\), by

$$\begin{aligned} \int _A \rho ({\mathbf{x}},t) d{\mathbf{x}}= P\big ({\mathbf{X}}(t)\in A\big ) = \int {\mathbbm {1}}_A \big (\phi ^t ({\mathbf{x}})\big )\rho ({\mathbf{x}}, 0)d{\mathbf{x}}\;, \end{aligned}$$
(5.C.1.4)

where \({\mathbbm {1}}_A\) is the indicator function of the set A.

Then we have

$$\begin{aligned} \frac{\partial \rho ({\mathbf{x}},t)}{ \partial t} + \mathbbm {\nabla } \cdot \Big [{\mathbf{V}}({\mathbf{x}},t)\rho ({\mathbf{x}},t)\Big ]=0\;. \end{aligned}$$
(5.C.1.5)

To prove (5.C.1.5), observe that, by approximating integrals by sums, for any smooth function with compact support, we get from (5.C.1.4):

$$\begin{aligned} \int F({\mathbf{x}}) \rho ({\mathbf{x}},t)d{\mathbf{x}}= \int F \big (\phi ^t({\mathbf{x}})\big ) \rho ({\mathbf{x}}, 0)d{\mathbf{x}}\;. \end{aligned}$$
(5.C.1.6)

Then, differentiating both sides of (5.C.1.6) with respect to t, and using the fact that (5.C.1.3) implies \(\partial \phi ^t ({\mathbf{x}})/\partial t = {\mathbf{V}}\big (\phi ^t ({\mathbf{x}}), t\big )\), we have

$$\begin{aligned} \int F({\mathbf{x}}) \frac{\partial \rho ({\mathbf{x}},t)}{ \partial t} d{\mathbf{x}}= \int \mathbbm {\nabla } F \big (\phi ^t ({\mathbf{x}})\big ) \cdot {\mathbf{V}}\big (\phi ^t ({\mathbf{x}}),t\big ) \rho ({\mathbf{x}}, 0)d{\mathbf{x}}\;. \end{aligned}$$

Using (5.C.1.6) with \(F (\cdot )\) replaced by \({\mathbbm {\nabla }} F (\cdot ) \cdot {\mathbf{V}}(\cdot ,t)\), the right-hand side of this equals

$$\begin{aligned} \int {\mathbbm {\nabla }} F ({\mathbf{x}}) \cdot {\mathbf{V}}({\mathbf{x}},t) \rho ({\mathbf{x}},t)d{\mathbf{x}}\;. \end{aligned}$$

Integrating by parts, this is

$$\begin{aligned} - \int F({\mathbf{x}}) \nabla \cdot \Big [{\mathbf{V}}({\mathbf{x}},t)\rho ({\mathbf{x}},t)\Big ] d{\mathbf{x}}\;. \end{aligned}$$
(5.C.1.7)

Since (5.C.1.7) is valid for any F, it implies (5.C.1.5).

Now, if we consider the differential equation (5.C.1.3) with \({\mathbf{V}}({\mathbf{X}},t)={\mathbf{V}}_{\Psi } ({\mathbf{X}}, t)\), with \({\mathbf{V}}_{\Psi } ({\mathbf{X}}, t)\) given by (5.1.1.4), and \(\rho ({\mathbf{x}},t)=|\Psi ({\mathbf{x}},t)|^2\), we get, using (5.C.1.4), (5.C.1.5), and (5.A.7),

$$\begin{aligned} \frac{d}{ dt} P\big ({\mathbf{X}}(t)\in A\big )&= \int _A \frac{\partial \rho ({\mathbf{x}},t)}{\partial t} d{\mathbf{x}} \nonumber \\&= - \int _A \nabla \cdot \big [{\mathbf{V}}_{\Psi } ({\mathbf{x}}, t)\rho ({\mathbf{x}},t)\big ] d{\mathbf{x}} \nonumber \\&= \frac{d}{dt} \int _A |\Psi ({\mathbf{x}},t)|^2 d{\mathbf{x}}\;, \end{aligned}$$
(5.C.1.8)

which proves (5.C.1.1) if we assume that (5.C.1.1) holds at \(t=0\), which is (5.C.1.2).

5.1.2 5.C.2 Proof of Equivariance of the Empirical Distributions

To finish the proof of (5.1.3.2), consider a large number N of copies of non-interacting identical systems,Footnote 100 whose initial wave functionFootnote 101 is

$$\begin{aligned} \Psi ({\mathbf{x}}, 0)=\Psi (x_1, \dots , x_N, 0)= \prod _{i=1}^N \psi (x_i, 0) \;. \end{aligned}$$
(5.C.2.1)

The density of the associated probability distribution is

$$\begin{aligned} |\Psi (x_1, \dots , x_N, 0)|^2= \prod _{i=1}^N |\psi (x_i, 0)|^2\;, \end{aligned}$$
(5.C.2.2)

which means that the variables \(x_i\) are independent identically distributed random variables with distribution \(|\psi (x, 0)|^2\). This implies, by the law of large numbers (see Appendix 3.A), that the empirical density distribution \(\rho (x,0)\) of the N particles will be given, as \(N\rightarrow \infty \), by \(\rho (x,0)= |\psi (x, 0)|^2\).

By (5.C.1.1), the probability distribution density of \((x_1, \dots , x_N)\) at time \(t>0\) will be

$$\begin{aligned} |\Psi (x_1, \dots , x_N, t)|^2= \prod _{i=1}^N |\psi (x_i, t)|^2\;, \end{aligned}$$
(5.C.2.3)

and the empirical density distribution \(\rho (x,t)\) of the N particles at time t will be given, as \(N\rightarrow \infty \), by \(\rho (x,t)= |\psi (x, t)|^2\). This completes the proof of (5.1.3.2).

5.D Examples of de Broglie–Bohm Dynamics

5.1.1 5.D.1 The Gaussian Wave Function

Consider the time evolution of a Gaussian wave function (2.A.2.23), with initial condition \( \Psi (x, 0)=\pi ^{-1/4}\exp (- x^2/2)\), which we copy hereFootnote 102:

$$\begin{aligned} \Psi (x,t) = \frac{1}{(1+it/m)^{1/2}} \frac{1}{\pi ^{1/4}}\exp \left[ - \frac{x^2}{2 (1+it/m)}\right] \;. \end{aligned}$$
(5.D.1.1)

It is easy to see that, if we write \(\Psi (x,t)= R (x,t) \exp \big [iS (x,t)\big ]\), we have (up to a constant term)

$$ S(x,t)= \frac{t x^2}{2m (1+t^2/m^2)}\;, $$

so that (5.1.1.3) gives rise to the equation of motion

$$\begin{aligned} \frac{d}{dt} X(t)= \frac{t X(t)}{m^2+t^2}\;, \end{aligned}$$
(5.D.1.2)

which can be easily integrated by a separation of variables:

$$\begin{aligned} X(t)=\frac{X(0)}{m} \sqrt{m^2+t^2}= X(0) \sqrt{1+\left( \frac{t}{m}\right) ^2}\;. \end{aligned}$$
(5.D.1.3)

This gives the explicit dependence of the position at time t on the initial condition and shows that the particle does not move if it is initially at \(X(0)=0\), but moves asymptotically, when \(t\rightarrow \infty \), according to \(X(t)\sim X(0)t/m\).

Equation (2.A.3.3) can be checked explicitly here. Indeed, putting \(m=1\) as in Appendix 2.A.3, we get from (2.A.2.24)

$$\begin{aligned} t|\Psi ( pt, t)|^2= t\frac{1}{\sqrt{\pi (1+t^2)} }\exp \left( - \frac{p^2t^2}{ 1+t^2}\right) \;, \end{aligned}$$
(5.D.1.4)

whose limit as \(t \rightarrow \infty \) is \(\pi ^{-1/2}\exp (- p^2) =|\widehat{\Psi }(p,0)|^2\).

It is also easy to verify explicitly the property (5.1.3.2) of equivariance. From (5.D.1.3), we get \(\phi ^t(x)= x \sqrt{1+(t/m)^2}\). If we make the change of variable \(y= \phi ^t(x)\) in the integral on the right-hand side of (5.1.3.1), we get, with the initial distribution \(\rho (x,0)= | \Psi (x,0)|^2= \pi ^{-1/2}\exp (- x^2)\),

$$\begin{aligned} \rho (y,t)= \frac{1}{\sqrt{\pi \big [1+(t/m)^2\big ]} }\exp \left[ - \frac{y^2}{ 1+(t/m)^2}\right] \;. \end{aligned}$$
(5.D.1.5)

Changing the y variable back to x, this coincides with \(| \Psi (x,t)|^2\) given by (2.A.2.24), with the prefactor \(1/\sqrt{1+(t/m)^2}\) coming from \(dx= dy/\sqrt{1+(t/m)^2}\).

5.1.2 5.D.2 The Particle in a Box

Consider a particle in a box,Footnote 103 in one dimension, of size L, chosen to be the interval [0, L]. The particle is free in the box, but constrained to remain in it. Let us look for solutions of the Schrödinger equation (5.1.1.1), with \(N=1\), of the form \(\exp (-iEt) \Psi (x) \), whose associated probability distribution \(|\exp (-iEt) \Psi (x)|^2= | \Psi (x)|^2 \) is independent of time.Footnote 104

We model the presence of the box by setting the potential in (5.1.1.2) such that \(V=0\) inside the box and \(V=\infty \) outside the box, which means that \(\Psi (x)\) vanishes outside V. So we have to solve

$$\begin{aligned} \frac{1}{2} \frac{d^2}{dx^2} \Psi (x) =- E \Psi (x)\;, \end{aligned}$$
(5.D.2.1)

with \(\Psi (0) = \Psi (L)=0\). We have set the mass \(m=1\).

The general solution of (5.D.2.1) is

$$\begin{aligned} A \exp \big (i \sqrt{2E} x\big ) + B \exp \big (-i \sqrt{2E} x\big )\;, \end{aligned}$$
(5.D.2.2)

and the conditions \(\Psi (0)=\Psi (L)=0\) imply \(A=-B\) and \(E_n=n^2\pi ^2/ L^2\), for \(n\in \mathbbm {Z}\), \(n \ne 0\). From (5.D.2.2), we get the eigenvectors

$$\begin{aligned} \Psi _n (x) = \sqrt{\frac{2}{L}} \sin (k_n x) {\mathbbm {1}}_{[0, L]} (x)\;, \end{aligned}$$
(5.D.2.3)

where \({\mathbbm {1}}_{[0, L]} \) is the indicator function of [0, L], and \(k_n = n\pi /L\). The sine function being odd, we can restrict ourselves to \(n=1,2,\ldots , \infty \), since the solution should be nonzero and is defined up to a multiplicative factor which does not affect the motion of the particle, as can be seen from (5.1.1.4).Footnote 105 The factor \(\sqrt{2/L}\) ensures that \(\int _{\mathbbm {R}} |\Psi _n(x)|^2 dx =1\).

If we take a function \(\Psi _n(x)\) as initial condition in (5.1.1.1), then the time-dependent solution is

$$\begin{aligned} \Psi _n (x,t) = \sqrt{\frac{2}{L}} e^{-iE_nt} \sin (k_n x) {\mathbbm {1}}_{[0, L]} (x)\;. \end{aligned}$$
(5.D.2.4)

Now comes the apparent paradox: the phase \(S_n\) of \(\Psi _n (x,t)\) is constant in x, so \(\partial S_n / \partial x=0\), and by (5.1.1.3), the particle is at rest in the box. Let us see how to solve that paradox. As we said, one method to measure the momentum is to remove the walls of the box and to detect the position x of the particle after a certain time t, and give x / t as a “measure” of momentum.

To see what happens when the walls of the box are removed, we have to compute the time evolution of the wave function with initial condition (5.D.2.3). Since the walls are removed, the time evolution is governed by the free Schrödinger equation (5.1.1.1). The solution is therefore obtained from (2.A.2.21) and we need to compute \(\widehat{\Psi }(p,0)\) from (5.D.2.3), leaving out the index n on \(\Psi \). Thus,

$$\begin{aligned} \widehat{\Psi }(p,0) =&\frac{1}{ \sqrt{2\pi }} \int _{{\mathbbm {R}}} e^{-ipx} \sqrt{\frac{2}{ L}} \sin (k_n x) {\mathbbm {1}}_{[0, L]} (x) dx\\ {}&= \frac{1}{ \sqrt{2\pi }} \int ^L_0 e^{-ipx} \sqrt{\frac{2}{ L}} \sin ( k_n x) dx\;. \end{aligned}$$
(5.D.2.5)

Writing \(\sin (k_n x) = (e^{ik_n x}-e^{-ik_n x})/2i\), making the change of variable \(x=x'+L/ 2\), and integrating over \(x'\in [-L/ 2,L/ 2]\), we obtain

$$\begin{aligned} \widehat{\Psi }(p,0) = \frac{-i}{\sqrt{\pi L}} \left[ e^{i(k_n-p)L/ 2}\frac{\sin \big [(p-k_n)L/ 2\big ]}{ p-k_n} - e^{-i(k_n+p)L/ 2}\frac{\sin \big [(p+k_n)L/ 2\big ]}{ p+k_n}\right] \;. \end{aligned}$$
(5.D.2.6)

This can be written as:

$$\begin{aligned} \widehat{\Psi }(p,0)&= \frac{-i}{\sqrt{\pi L}} e^{-ip L/ 2} e^{ik_nL/ 2} \left[ \frac{\sin \big [(p-k_n)L/ 2\big ]}{ p-k_n} - e^{-ik_nL}\frac{\sin \big [(p+k_n)L/ 2\big ]}{ p+k_n}\right] \end{aligned}$$
$$\begin{aligned}&= \frac{-i}{\sqrt{\pi L}} e^{-ip L/ 2} e^{ik_nL/ 2} \left[ \frac{\sin \big [(p-k_n)L/ 2\big ]}{ p-k_n} +(-1)^{n+1}\frac{\sin \big [(p+k_n)L/ 2\big ]}{ p+k_n}\right] \;, \end{aligned}$$
(5.D.2.7)

since \(e^{-ik_nL}= e^{-in\pi }=(-1)^n\). From (2.A.2.21) we get, using the variable \(x'=x - L/2\),

$$\begin{aligned}&\Psi (x',t) = \frac{-i}{\pi \sqrt{2 L}}e^{ik_nL/ 2}\int _{{\mathbbm {R}}} \exp \left( -it \frac{p^2}{2}+ipx'\right) \left[ \frac{\sin \big [(p-k_n)L/ 2\big ]}{ p-k_n}\right. \\&\quad \left. +\,(-1)^{n+1}\frac{\sin \big [(p+k_n)L/ 2\big ]}{ p+k_n}\right] dp\;. \end{aligned}$$
(5.D.2.8)

From Plancherel’s theorem [see also (2.A.2.26)], we know that

$$ \int _{{\mathbbm {R}}} |\Psi (x',t) |^2 dx'= \int _{{\mathbbm {R}}} |\widehat{\Psi }(p,t) |^2 dp= \int _{\mathbbm {R}} |\widehat{\Psi }(p,0) |^2 dp= \int _{\mathbbm {R}} |\Psi (x,0) |^2 dx=1\;. $$

Now, for \(t\ne 0\), if we write \(\Psi (x',t)=R (x',t) e^{iS(x',t)}\), there is no reason why we should have \(\partial S(x',t)/\partial x'=0\) and an analysis of (5.D.2.8) shows that we do indeed have \(\partial S(x',t)/ \partial x'\ne 0\) for \(t\ne 0\). In fact, we can see this indirectly using (2.A.3.2) and (2.A.3.3), which imply

$$\begin{aligned} \lim _{t\rightarrow \infty } \int _{At} |\Psi (x',t)|^2 dx'= \int _A |\widehat{\Psi }(p,0)|^2 dp\;. \end{aligned}$$
(5.D.2.9)

This means that the support of \(\Psi (x',t)\) spreads as \(t \rightarrow \infty \) (away from \(x'=0\), i.e., away from the center of the box \(x=L/ 2\)), and by equivariance (5.1.3.1), this means that the particles will move towards either \(+\infty \) or \(-\infty \).

To illustrate this, consider what happens in the limit of large nL (mathematically, \(n, L \rightarrow \infty \)) with \(k_n = n\pi / L\) fixed.Footnote 106 It is well known (from the theory of Fourier transforms) that

$$\begin{aligned} \frac{\sin (Ly/2)}{ y}=\frac{1}{2} \int _{-L/2}^{L/2} e^{isy} ds \;\longrightarrow \; \pi \delta (y)\;, \end{aligned}$$
(5.D.2.10)

as \(L\rightarrow \infty \), in the sense of distributions. So in that limit, we may change variables \(y=p-k_n\) or \(y=p+k_n\) in (5.D.2.8) and, to a good approximation, replace the integral in (5.D.2.8) by the value of the factor \(\exp (it p^2/2+ipx) \) at \(p=\pm k_n\). We then obtain

$$\begin{aligned} \Psi (x',t) \approx \frac{-i e^{ik_nL/2}}{ \sqrt{2 L}} \left[ e^{-it k^2_n/ 2+ik_n x'} +(-1)^{n+1} e^{-it k^2_n/ 2-ik_n x'} \right] \;, \end{aligned}$$
(5.D.2.11)

which represents two plane waves propagating in opposite directions with absolute velocity \(k_n\), since \(\partial (\pm k_n x')/\partial x'= \pm k_n\).

Of course, one has to be a bit careful with this limit, since we have \(L\rightarrow \infty \) and \(1/\sqrt{ L}\) in (5.D.2.11). But the function in square brackets in (5.D.2.11) is not square integrable over \(\mathbbm {R}\) (while we know that \(\int _{\mathbbm {R}} |\Psi (x',t) |^2 dx'=1\) for all t), and moreover, its phase is constant in \(x'\). So what the limit means is that, for large t, \(\Psi (x',t)\) is the sum of two square integrable functions, one whose Fourier transform is localized near \(p=k_n\), and the other whose Fourier transform is localized around \(p=-k_n\). And by (5.D.2.8) and equivariance, we find that the particles that are in the support of one of those two wave functions move with a speed approximately equal to \(\pm k_n\).

The result (5.D.2.11) is exactly what Einstein asked for, when he thought of the classical limit as being made of a motion going back and forth between the walls of the box,Footnote 107 except that here this only happens once we remove the walls of the box.Footnote 108

5.E The Effective Collapse of the Quantum State and Decoherence

Let us describe the measurement process of Appendix 2.D in more detail (here we follow [70, Sect. 6.1]). When the particle is deflected by the magnetic field , either up or down, its presence is detected by a macroscopic device. One example of such a device would be, for a charged particle, an ion chamber. A cascade of ionization of atoms will eventually produce a pulse that can be observed, at a macroscopic level, by a galvanometer.

Consider two measuring devices, one detecting the particle if it goes up, and one detecting it if it goes down. Let the initial wave function for the two devices be

$$\begin{aligned} \Phi _0 (x_1,\ldots ,x_N,y_1,\ldots ,y_N)= \Phi _0^{\text {{up}}} (x_1,\ldots ,x_N) \Phi _0^{\text {{down}}} (y_1,\ldots ,y_N)\;, \end{aligned}$$
(5.E.1)

where \(x_1 \ldots x_N\) are the coordinates of the particles making up the first device and \(y_1 \ldots y_N\) are those for the second device, while N is of the order of the Avogadro number \((N\sim 10^{23})\) . Let \(\Phi ^{\text {{up}}}_\mathrm{f} (x_1,\ldots ,x_N)\) and \( \Phi _\mathrm{f}^{\text {{down}}} (y_1,\ldots ,y_N)\) (f for “final”) denote the wave functions of the detector after the particle has been detected. Note that these wave functions have a support that is disjoint from the support of \(\Phi _0^{\text {{up}}}\), \(\Phi _0^{\text {{down}}}\) for each variable \(x_i\), \(y_i\). Indeed, if we think of an ion chamber, then the electrons lost by the ions will move towards an electric plate in \(\Phi _\mathrm{f}^{\text {{up}}}\), while they will remain bound to their atoms in \(\Phi _0^{\text {{up}}}\) (and the same holds for \(\Phi _0^{\text {{down}}}\), \(\Phi _\mathrm{f}^{\text {{down}}}\)). But this can occur only if the supports of the wave functions are disjoint, for each electron .

If the initial quantum state of the particle is given by (5.1.4.7), then the combined initial quantum state of the particle and the ion chamber is

$$ \Phi _0 = \Phi _0 (x_1,\ldots ,x_N,y_1,\ldots ,y_N) \Psi (z) \big (c_1 |1 \uparrow \rangle + c_2 |1 \downarrow \rangle \big )\;. $$

Then, the final quantum state [see (5.1.4.8)] will be:

$$\begin{aligned} \Phi ^{\text {{up}}}_\mathrm{f} \Phi ^{\text {{down}}}_0 \Psi (z-t) c_1 |1 \uparrow \rangle + \Phi ^{\text {{up}}}_0 \Phi ^{\text {{down}}}_\mathrm{f} \Psi (z+t) c_2 |1 \downarrow \rangle \;, \end{aligned}$$
(5.E.2)

where we have left out the arguments of the wave functions \(\Phi ^{\text {{up}}}_\mathrm{f} \Phi ^{\text {{down}}}_0 \) and \(\Phi ^{\text {{up}}}_0 \Phi ^{\text {{down}}}_\mathrm{f}\). This means that, for the \(\Psi (z-t) c_1 |1 \uparrow \rangle \) part of the quantum state, the up detector will be activated and the down part will remain in its original state, and vice versa for the \(\Psi (z+t) c_2 |1 \downarrow \rangle \) part.

Now, since the two parts in (5.E.2) have disjoint supports, the original particle can only be in one of them, and so will be guided only by that part of the quantum state in whose support it happens to lie. However, we cannot generally neglect the part of the wave function in the support of which the particle is not located, because the two parts could subsequently be brought together, in such a way that their supports overlap . This is exactly what causes the interference phenomena, exemplified by the Mach–Zehnder interferometer , as discussed in Sect. 2.2. In our example, the quantum state of the particle, viz.,

$$ \Psi (z) \big (c_1 |1 \uparrow \rangle + c_2 |1 \downarrow \rangle \big )\;, $$

evolves into

$$\begin{aligned} \Psi (z-t) c_1 |1 \uparrow \rangle +\Psi (z+t) c_2 |1 \downarrow \rangle \;, \end{aligned}$$
(5.E.3)

where \(\Psi (z-t)\) and \(\Psi (z+t)\) have disjoint supports (for t not too small). In a quantum state like (5.E.3), the particle is guided by \(\Psi (z-t)\) or \(\Psi (z+t)\), depending on where it is, but we cannot “reduce” the quantum state and keep only that part, because \(\Psi (z-t)\) and \(\Psi (z+t)\) may later evolve into wave functions with overlapping supports.

But how can one do that with a quantum state like (5.E.2) involving of the order of \(10^{23}\) particles? It is not enough to make \(\Psi (z-t)\) and \(\Psi (z+t)\) overlap, we need to do that also with \(\Phi ^{\text {{up}}}_\mathrm{f}\) \(\Phi ^{\text {{down}}}_0\) and \(\Phi ^{\text {{up}}}_0\) \(\Phi ^{\text {{down}}}_\mathrm{f}\). But this is impossible in practice. Let us see why.

Remember that these wave functions represent a collection of \(N\sim 10^{23}\) particles, all of which have different locations, some moving more or less freely and some being bound to their atoms. Hence, the wave functions must have different supports for each particle. Consider, for simplicity, a wave function that factorizes :

$$\begin{aligned} \Phi ^{\text {{up}}}_0(x_1, \ldots , x_N) = \prod ^N_{i=1} \Phi _{0 i} (x_i)\;, \end{aligned}$$
(5.E.4)

and similarly for \(\Phi ^{\text {{up}}}_\mathrm{f}\), \(\Phi ^{\text {{down}}}_0 \), and \(\Phi ^{\text {{down}}}_\mathrm{f}\). The support of \(\Phi _{0 i}\) is disjoint from the support of \(\Phi _{\mathrm{f} i}\) for every particle since, in the support of \(\Phi _{0 i}\), the particle is bound to the atom and, in the support of \(\Phi _{\mathrm{f} i}\), it is moving.

For the two wave functions to interfere again, we need the wave functions \(\Phi _{0 i}(x_i)\) of all the particles to overlap in the future. Indeed, \(\Phi ^{\text {{up}}}_0\) and \(\Phi ^{\text {{up}}}_\mathrm{f}\) have disjoint supports if

$$\begin{aligned} \Phi ^{\text {{up}}}_0(x_1, \ldots , x_N)\Phi ^{\text {{up}}}_\mathrm{f}(x_1, \ldots , x_N)=0\;, \end{aligned}$$
(5.E.5)

\(\forall (x_1, \ldots , x_N) \in {{\mathbb {R}}}^N\). Of course, the same equation also holds for \(\Phi ^{\text {{down}}}_0(x_1, \ldots , x_N)\Phi ^{\text {{down}}}_\mathrm{f}(x_1, \ldots , x_N)\). But for (5.E.5) to be true, it is enough to have \(\Phi _{0 i} (x_i) \Phi _{\mathrm{f} i}(x_i)=0\), \(\forall x_i \in {{\mathbb {R}}}\), for a single value of i. Hence, for interference to occur, we need to get the wave functions of every particle in (say) the ion chamber to overlap again.

It is clear that this is, in practice, impossible. The same is true for the particles constituting the live cat and the dead cat, the pointer pointing up or down, the exploded and unexploded bomb, etc. Now, of course, once the position of the original particle is in the support of one of the two terms in (5.E.2), we can simply reduce the quantum state to that term, since we can be sure that the other term will never again interfere with it.

But we can know in which of the supports of the two terms the particle happens to be located simply by looking at the (by definition) macroscopic result. Hence, in some sense, we do “collapse ” the quantum state after we look at the result of an experiment. However, this is an entirely practical matter. We can still consider, if we wish, that the true quantum state is and remains forever given by the time evolution of (5.E.2). Put simply, there is one of the terms that does not guide the motion of the particle, either now or ever again.

5.F Proof of the Factorization Formula

We now prove (5.1.7.5). Note that (5.1.7.4) implies that

$$\begin{aligned} \Psi (x_1,\ldots ,x_M,{\mathbf{Y}})=\psi (x_i) \Phi _i(\hat{x}_i, {\mathbf{Y}})\;, \end{aligned}$$
(5.F.1)

where \(\hat{x}_i = (x_1,\ldots ,x_M)\) with \(x_i\) missing and we have put in \(\Phi _i(\hat{x}_i, {\mathbf{Y}})\) the arguments that were not indicated in (5.1.7.4).

To prove (5.1.7.5), consider \(M=2\). By (5.F.1), we have

$$\begin{aligned} \Psi (x_1,x_2,{\mathbf{Y}})&= \psi (x_1) \Phi _1 (x_2,{\mathbf{Y}}) \end{aligned}$$
$$\begin{aligned}&=\psi (x_2) \Phi _2 (x_1,{\mathbf{Y}})\;. \end{aligned}$$
(5.F.2)

Dividing by \(\psi (x_1)\psi (x_2)\), we get

$$\begin{aligned} \frac{\Phi _1 (x_2,{\mathbf{Y}})}{ \psi (x_2)} = \frac{\Phi _2(x_1,{\mathbf{Y}})}{\psi (x_1)}\;, \end{aligned}$$
(5.F.3)

which means that both sides are functions of \({\mathbf{Y}}\) alone, which we can write \(\Phi ({\mathbf{Y}})\). Thus,

$$\begin{aligned} \frac{\Psi (x_1,x_2,{\mathbf{Y}})}{ \psi (x_1)\psi (x_2)}=\Phi ({\mathbf{Y}})\;, \end{aligned}$$

which is (5.1.7.5) for \(M=2\). The argument easily extends to all M.

Rights and permissions

Reprints and permissions

Copyright information

© 2016 Springer International Publishing Switzerland

About this chapter

Cite this chapter

Bricmont, J. (2016). The de Broglie–Bohm Theory. In: Making Sense of Quantum Mechanics. Springer, Cham. https://doi.org/10.1007/978-3-319-25889-8_5

Download citation

Publish with us

Policies and ethics