Popper's experiment

Last updated

Popper's experiment is an experiment proposed by the philosopher Karl Popper to test aspects of the uncertainty principle in quantum mechanics.

Contents

History

In fact, as early as 1934, Popper started criticising the increasingly more accepted Copenhagen interpretation, a popular subjectivist interpretation of quantum mechanics. Therefore, in his most famous book Logik der Forschung he proposed a first experiment alleged to empirically discriminate between the Copenhagen Interpretation and a realist ensemble interpretation, which he advocated. Einstein, however, wrote a letter to Popper about the experiment in which he raised some crucial objections [1] and Popper himself declared that this first attempt was "a gross mistake for which I have been deeply sorry and ashamed of ever since". [2]

Popper, however, came back to the foundations of quantum mechanics from 1948, when he developed his criticism of determinism in both quantum and classical physics. [3] As a matter of fact, Popper greatly intensified his research activities on the foundations of quantum mechanics throughout the 1950s and 1960s developing his interpretation of quantum mechanics in terms of real existing probabilities (propensities), [4] also thanks to the support of a number of distinguished physicists (such as David Bohm). [5]

In 1980, Popper proposed perhaps his more important, yet overlooked, contribution to QM: a "new simplified version of the EPR experiment". [6]

The experiment was however published only two years later, in the third volume of the Postscript to the Logic of Scientific Discovery . [7]

The most widely known interpretation of quantum mechanics is the Copenhagen interpretation put forward by Niels Bohr and his school. It maintains that observations lead to a wavefunction collapse, thereby suggesting the counter-intuitive result that two well separated, non-interacting systems require action-at-a-distance. Popper argued that such non-locality conflicts with common sense, and would lead to a subjectivist interpretation of phenomena, depending on the role of the 'observer'.

While the EPR argument was always meant to be a thought experiment, put forward to shed light on the intrinsic paradoxes of QM, Popper proposed an experiment which could have been experimentally implemented and participated at a physics conference organised in Bari in 1983, to present his experiment and propose to the experimentalists to carry it out.

The actual realisation of Popper's experiment required new techniques which would make use of the phenomenon of spontaneous parametric down-conversion but had not yet been exploited at that time, so his experiment was eventually performed only in 1999, five years after Popper had died.

Description

Popper's experiment of 1980 exploits couples of entangled particles, in order to put to the test Heisenberg's uncertainty principle. [6] [8]

Indeed, Popper maintains:

"I wish to suggest a crucial experiment to test whether knowledge alone is sufficient to create 'uncertainty' and, with it, scatter (as is contended under the Copenhagen interpretation), or whether it is the physical situation that is responsible for the scatter." [9]

Popper's proposed experiment consists of a low-intensity source of particles that can generate pairs of particles traveling to the left and to the right along the x-axis. The beam's low intensity is "so that the probability is high that two particles recorded at the same time on the left and on the right are those which have actually interacted before emission." [9]

There are two slits, one each in the paths of the two particles. Behind the slits are semicircular arrays of counters which can detect the particles after they pass through the slits (see Fig. 1). "These counters are coincident counters [so] that they only detect particles that have passed at the same time through A and B." [10]

Fig.1 Experiment with both slits equally wide. Both the particles should show equal scatter in their momenta. Popper-experiment-1.png
Fig.1 Experiment with both slits equally wide. Both the particles should show equal scatter in their momenta.

Popper argued that because the slits localize the particles to a narrow region along the y-axis, from the uncertainty principle they experience large uncertainties in the y-components of their momenta. This larger spread in the momentum will show up as particles being detected even at positions that lie outside the regions where particles would normally reach based on their initial momentum spread.

Popper suggests that we count the particles in coincidence, i.e., we count only those particles behind slit B, whose partner has gone through slit A. Particles which are not able to pass through slit A are ignored.

The Heisenberg scatter for both the beams of particles going to the right and to the left, is tested "by making the two slits A and B wider or narrower. If the slits are narrower, then counters should come into play which are higher up and lower down, seen from the slits. The coming into play of these counters is indicative of the wider scattering angles which go with a narrower slit, according to the Heisenberg relations." [10]

Fig.2 Experiment with slit A narrowed, and slit B wide open. Should the two particle show equal scatter in their momenta? If they do not, Popper says, the Copenhagen interpretation is wrong. If they do, it indicates action at a distance, says Popper. Popper-experiment-2.png
Fig.2 Experiment with slit A narrowed, and slit B wide open. Should the two particle show equal scatter in their momenta? If they do not, Popper says, the Copenhagen interpretation is wrong. If they do, it indicates action at a distance, says Popper.

Now the slit at A is made very small and the slit at B very wide. Popper wrote that, according to the EPR argument, we have measured position "y" for both particles (the one passing through A and the one passing through B) with the precision , and not just for particle passing through slit A. This is because from the initial entangled EPR state we can calculate the position of the particle 2, once the position of particle 1 is known, with approximately the same precision. We can do this, argues Popper, even though slit B is wide open. [10]

Therefore, Popper states that "fairly precise "knowledge"" about the y position of particle 2 is made; its y position is measured indirectly. And since it is, according to the Copenhagen interpretation, our knowledge which is described by the theory and especially by the Heisenberg relations — it should be expected that the momentum of particle 2 scatters as much as that of particle 1, even though the slit A is much narrower than the widely opened slit at B.

Now the scatter can, in principle, be tested with the help of the counters. If the Copenhagen interpretation is correct, then such counters on the far side of B that are indicative of a wide scatter (and of a narrow slit) should now count coincidences: counters that did not count any particles before the slit A was narrowed.

To sum up: if the Copenhagen interpretation is correct, then any increase in the precision in the measurement of our mere knowledge of the particles going through slit B should increase their scatter. [11]

Popper was inclined to believe that the test would decide against the Copenhagen interpretation, as it is applied to Heisenberg's uncertainty principle. If the test decided in favor of the Copenhagen interpretation, Popper argued, it could be interpreted as indicative of action at a distance.

The debate

Many viewed Popper's experiment as a crucial test of quantum mechanics, and there was a debate on what result an actual realization of the experiment would yield.

In 1985, Sudbery pointed out that the EPR state, which could be written as , already contained an infinite spread in momenta (tacit in the integral over k), so no further spread could be seen by localizing one particle. [12] [13] Although it pointed to a crucial flaw in Popper's argument, its full implication was not understood. Kripps theoretically analyzed Popper's experiment and predicted that narrowing slit A would lead to momentum spread increasing at slit B. Kripps also argued that his result was based just on the formalism of quantum mechanics, without any interpretational problem. Thus, if Popper was challenging anything, he was challenging the central formalism of quantum mechanics. [14]

In 1987 there came a major objection to Popper's proposal from Collet and Loudon. [15] They pointed out that because the particle pairs originating from the source had a zero total momentum, the source could not have a sharply defined position. They showed that once the uncertainty in the position of the source is taken into account, the blurring introduced washes out the Popper effect.

Furthermore, Redhead analyzed Popper's experiment with a broad source and concluded that it could not yield the effect that Popper was seeking. [16]

Realizations

Fig.3 Schematic diagram of Kim and Shih's experiment based on a BBO crystal which generates entangled photons. The lens LS helps create a sharp image of slit A on the location of slit B. Kimshih.png
Fig.3 Schematic diagram of Kim and Shih's experiment based on a BBO crystal which generates entangled photons. The lens LS helps create a sharp image of slit A on the location of slit B.
Fig.4 Results of the photon experiment by Kim and Shih, aimed at realizing Popper's proposal. The diffraction pattern in the absence of slit B (red symbols) is much narrower than that in the presence of a real slit (blue symbols). Kimshih-result.png
Fig.4 Results of the photon experiment by Kim and Shih, aimed at realizing Popper's proposal. The diffraction pattern in the absence of slit B (red symbols) is much narrower than that in the presence of a real slit (blue symbols).

Kim–Shih's experiment


Popper's experiment was realized in 1999 by Yoon-Ho Kim & Yanhua Shih using a spontaneous parametric down-conversion photon source. They did not observe an extra spread in the momentum of particle 2 due to particle 1 passing through a narrow slit. They write:

"Indeed, it is astonishing to see that the experimental results agree with Popper’s prediction. Through quantum entanglement one may learn the precise knowledge of a photon’s position and would therefore expect a greater uncertainty in its momentum under the usual Copenhagen interpretation of the uncertainty relations. However, the measurement shows that the momentum does not experience a corresponding increase in uncertainty. Is this a violation of the uncertainty principle?" [17]

Rather, the momentum spread of particle 2 (observed in coincidence with particle 1 passing through slit A) was narrower than its momentum spread in the initial state.

They concluded that:

"Popper and EPR were correct in the prediction of the physical outcomes of their experiments. However, Popper and EPR made the same error by applying the results of two-particle physics to the explanation of the behavior of an individual particle. The two-particle entangled state is not the state of two individual particles. Our experimental result is emphatically NOT a violation of the uncertainty principle which governs the behavior of an individual quantum." [17]

This led to a renewed heated debate, with some even going to the extent of claiming that Kim and Shih's experiment had demonstrated that there is no non-locality in quantum mechanics. [18]

Unnikrishnan (2001), discussing Kim and Shih's result, wrote that the result:

"is a solid proof that there is no state-reduction-at-a-distance. ... Popper's experiment and its analysis forces us to radically change the current held view on quantum non-locality." [19]

Short criticized Kim and Shih's experiment, arguing that because of the finite size of the source, the localization of particle 2 is imperfect, which leads to a smaller momentum spread than expected. [20] However, Short's argument implies that if the source were improved, we should see a spread in the momentum of particle 2. [ citation needed ]

Sancho carried out a theoretical analysis of Popper's experiment, using the path-integral approach, and found a similar kind of narrowing in the momentum spread of particle 2, as was observed by Kim and Shih. [21] Although this calculation did not give them any deep insight, it indicated that the experimental result of Kim-Shih agreed with quantum mechanics. It did not say anything about what bearing it has on the Copenhagen interpretation, if any.

Ghost diffraction

Popper's conjecture has also been tested experimentally in the so-called two-particle ghost interference experiment. [22] This experiment was not carried out with the purpose of testing Popper's ideas, but ended up giving a conclusive result about Popper's test. In this experiment two entangled photons travel in different directions. Photon 1 goes through a slit, but there is no slit in the path of photon 2. However, Photon 2, if detected in coincidence with a fixed detector behind the slit detecting photon 1, shows a diffraction pattern. The width of the diffraction pattern for photon 2 increases when the slit in the path of photon 1 is narrowed. Thus, increase in the precision of knowledge about photon 2, by detecting photon 1 behind the slit, leads to increase in the scatter of photons 2.

Predictions according to quantum mechanics

Tabish Qureshi has published the following analysis of Popper's argument. [23] [24]

The ideal EPR state is written as , where the two labels in the "ket" state represent the positions or momenta of the two particle. This implies perfect correlation, meaning, detecting particle 1 at position will also lead to particle 2 being detected at . If particle 1 is measured to have a momentum , particle 2 will be detected to have a momentum . The particles in this state have infinite momentum spread, and are infinitely delocalized. However, in the real world, correlations are always imperfect. Consider the following entangled state

where represents a finite momentum spread, and is a measure of the position spread of the particles. The uncertainties in position and momentum, for the two particles can be written as

The action of a narrow slit on particle 1 can be thought of as reducing it to a narrow Gaussian state:

.

This will reduce the state of particle 2 to

.

The momentum uncertainty of particle 2 can now be calculated, and is given by

If we go to the extreme limit of slit A being infinitesimally narrow (), the momentum uncertainty of particle 2 is , which is exactly what the momentum spread was to begin with. In fact, one can show that the momentum spread of particle 2, conditioned on particle 1 going through slit A, is always less than or equal to (the initial spread), for any value of , and . Thus, particle 2 does not acquire any extra momentum spread than it already had. This is the prediction of standard quantum mechanics. So, the momentum spread of particle 2 will always be smaller than what was contained in the original beam. This is what was actually seen in the experiment of Kim and Shih. Popper's proposed experiment, if carried out in this way, is incapable of testing the Copenhagen interpretation of quantum mechanics.

On the other hand, if slit A is gradually narrowed, the momentum spread of particle 2 (conditioned on the detection of particle 1 behind slit A) will show a gradual increase (never beyond the initial spread, of course). This is what quantum mechanics predicts. Popper had said

"...if the Copenhagen interpretation is correct, then any increase in the precision in the measurement of our mere knowledge of the particles going through slit B should increase their scatter."

This particular aspect can be experimentally tested.

Faster-than-light signalling

The expected additional momentum scatter which Popper wrongly attributed to the Copenhagen interpretation would allow faster-than-light communication, which is excluded by the no-communication theorem in quantum mechanics. Note however that both Collet and Loudon [15] and Qureshi [23] [24] compute that scatter decreases with decreasing the size of slit A, contrary to the increase predicted by Popper. There was some controversy about this decrease also allowing superluminal communication. [25] [26] But the reduction is of the standard deviation of the conditional distribution of the position of particle 2 knowing that particle 1 did go through slit A, since we are only counting coincident detection. The reduction in conditional distribution allows for the unconditional distribution to remain the same, which is the only thing that matters to exclude superluminal communication. Also note that the conditional distribution would be different from the unconditional distribution in classical physics as well. But measuring the conditional distribution after slit B requires the information on the result at slit A, which has to be communicated classically, so that the conditional distribution cannot be known as soon as the measurement is made at slit A but is delayed by the time required to transmit that information.

See also

Related Research Articles

<span class="mw-page-title-main">Particle in a box</span> Physical model in quantum mechanics which is analytically solvable

In quantum mechanics, the particle in a box model describes a particle free to move in a small space surrounded by impenetrable barriers. The model is mainly used as a hypothetical example to illustrate the differences between classical and quantum systems. In classical systems, for example, a particle trapped inside a large box can move at any speed within the box and it is no more likely to be found at one position than another. However, when the well becomes very narrow, quantum effects become important. The particle may only occupy certain positive energy levels. Likewise, it can never have zero energy, meaning that the particle can never "sit still". Additionally, it is more likely to be found at certain positions than at others, depending on its energy level. The particle may never be detected at certain positions, known as spatial nodes.

<span class="mw-page-title-main">Quantum mechanics</span> Description of physical properties at the atomic and subatomic scale

Quantum mechanics is a fundamental theory in physics that provides a description of the physical properties of nature at the scale of atoms and subatomic particles. It is the foundation of all quantum physics including quantum chemistry, quantum field theory, quantum technology, and quantum information science.

<span class="mw-page-title-main">Uncertainty principle</span> Foundational principle in quantum physics

In quantum mechanics, the uncertainty principle is any of a variety of mathematical inequalities asserting a fundamental limit to the product of the accuracy of certain related pairs of measurements on a quantum system, such as position, x, and momentum, p. Such paired-variables are known as complementary variables or canonically conjugate variables.

Wave–particle duality is the concept in quantum mechanics according to which quantum entities exhibit both particle and a wave properties depending on the experimental circumstances. It expresses the inability of the classical concepts "particle" or "wave" to fully describe the behaviour of quantum-scale objects alone. As Albert Einstein wrote:

It seems as though we must use sometimes the one theory and sometimes the other, while at times we may use either. We are faced with a new kind of difficulty. We have two contradictory pictures of reality; separately neither of them fully explains the phenomena of light, but together they do.

The de Broglie–Bohm theory, also known as the pilot wave theory, Bohmian mechanics, Bohm's interpretation, and the causal interpretation, is an interpretation of quantum mechanics. In addition to the wavefunction, it also postulates an actual configuration of particles exists even when unobserved. The evolution over time of the configuration of all particles is defined by a guiding equation. The evolution of the wave function over time is given by the Schrödinger equation. The theory is named after Louis de Broglie (1892–1987) and David Bohm (1917–1992).

Bell's theorem is a term encompassing a number of closely related results in physics, all of which determine that quantum mechanics is incompatible with local hidden-variable theories, given some basic assumptions about the nature of measurement. "Local" here refers to the principle of locality, the idea that a particle can only be influenced by its immediate surroundings, and that interactions mediated by physical fields cannot propagate faster than the speed of light. "Hidden variables" are putative properties of quantum particles that are not included in quantum theory but nevertheless affect the outcome of experiments. In the words of physicist John Stewart Bell, for whom this family of results is named, "If [a hidden-variable theory] is local it will not agree with quantum mechanics, and if it agrees with quantum mechanics it will not be local."

<span class="mw-page-title-main">Wave function</span> Mathematical description of the quantum state of a system

In quantum physics, a wave function, represented by the Greek letter Ψ, is a mathematical description of the quantum state of an isolated quantum system. In the Copenhagen interpretation of quantum mechanics, the wave function is a complex-valued probability amplitude; the probabilities for the possible results of the measurements made on a measured system can be derived from the wave function.

Matter waves are a central part of the theory of quantum mechanics, being half of wave–particle duality. All matter exhibits wave-like behavior. For example, a beam of electrons can be diffracted just like a beam of light or a water wave.

In physics, specifically in quantum mechanics, a coherent state is the specific quantum state of the quantum harmonic oscillator, often described as a state which has dynamics most closely resembling the oscillatory behavior of a classical harmonic oscillator. It was the first example of quantum dynamics when Erwin Schrödinger derived it in 1926, while searching for solutions of the Schrödinger equation that satisfy the correspondence principle. The quantum harmonic oscillator arise in the quantum theory of a wide range of physical systems. For instance, a coherent state describes the oscillating motion of a particle confined in a quadratic potential well. The coherent state describes a state in a system for which the ground-state wavepacket is displaced from the origin of the system. This state can be related to classical solutions by a particle oscillating with an amplitude equivalent to the displacement.

<span class="mw-page-title-main">Probability amplitude</span> Complex number whose squared absolute value is a probability

In quantum mechanics, a probability amplitude is a complex number used for describing the behaviour of systems. The modulus squared of this quantity represents a probability density.

<span class="mw-page-title-main">Wave packet</span> Short "burst" or "envelope" of restricted wave action that travels as a unit

In physics, a wave packet is a short burst of localized wave action that travels as a unit, outlined by an envelope. A wave packet can be analyzed into, or can be synthesized from, a potentially-infinite set of component sinusoidal waves of different wavenumbers, with phases and amplitudes such that they interfere constructively only over a small region of space, and destructively elsewhere. Any signal of a limited width in time or space requires many frequency components around a center frequency within a bandwidth inversely proportional to that width; even a gaussian function is considered a wave packet because its Fourier transform is a "packet" of waves of frequencies clustered around a central frequency. Each component wave function, and hence the wave packet, are solutions of a wave equation. Depending on the wave equation, the wave packet's profile may remain constant or it may change (dispersion) while propagating.

<span class="mw-page-title-main">Bohr–Einstein debates</span> Series of public disputes between physicists Niels Bohr and Albert Einstein

The Bohr–Einstein debates were a series of public disputes about quantum mechanics between Albert Einstein and Niels Bohr. Their debates are remembered because of their importance to the philosophy of science, insofar as the disagreements—and the outcome of Bohr's version of quantum mechanics becoming the prevalent view—form the root of the modern understanding of physics. Most of Bohr's version of the events held in Solvay in 1927 and other places was first written by Bohr decades later in an article titled, "Discussions with Einstein on Epistemological Problems in Atomic Physics". Based on the article, the philosophical issue of the debate was whether Bohr's Copenhagen Interpretation of quantum mechanics, which centered on his belief of complementarity, was valid in explaining nature. Despite their differences of opinion and the succeeding discoveries that helped solidify quantum mechanics, Bohr and Einstein maintained a mutual admiration that was to last the rest of their lives.

In physics, a free particle is a particle that, in some sense, is not bound by an external force, or equivalently not in a region where its potential energy varies. In classical physics, this means the particle is present in a "field-free" space. In quantum mechanics, it means the particle is in a region of uniform potential, usually set to zero in the region of interest since the potential can be arbitrarily set to zero at any point in space.

The Compton wavelength is a quantum mechanical property of a particle, defined as the wavelength of a photon whose energy is the same as the rest energy of that particle. It was introduced by Arthur Compton in 1923 in his explanation of the scattering of photons by electrons.

Quantum noise is noise arising from the indeterminate state of matter in accordance with fundamental principles of quantum mechanics, specifically the uncertainty principle and via zero-point energy fluctuations. Quantum noise is due to the apparently discrete nature of the small quantum constituents such as electrons, as well as the discrete nature of quantum effects, such as photocurrents.

The wave–particle duality relation, often loosely referred to as the Englert–Greenberger–Yasin duality relation, or the Englert–Greenberger relation, relates the visibility, , of interference fringes with the definiteness, or distinguishability, , of the photons' paths in quantum optics. As an inequality:

Photon polarization is the quantum mechanical description of the classical polarized sinusoidal plane electromagnetic wave. An individual photon can be described as having right or left circular polarization, or a superposition of the two. Equivalently, a photon can be described as having horizontal or vertical linear polarization, or a superposition of the two.

The theoretical and experimental justification for the Schrödinger equation motivates the discovery of the Schrödinger equation, the equation that describes the dynamics of nonrelativistic particles. The motivation uses photons, which are relativistic particles with dynamics described by Maxwell's equations, as an analogue for all types of particles.

<span class="mw-page-title-main">Angle-resolved photoemission spectroscopy</span> Experimental technique to determine the distribution of electrons in solids

Angle-resolved photoemission spectroscopy (ARPES) is an experimental technique used in condensed matter physics to probe the allowed energies and momenta of the electrons in a material, usually a crystalline solid. It is based on the photoelectric effect, in which an incoming photon of sufficient energy ejects an electron from the surface of a material. By directly measuring the kinetic energy and emission angle distributions of the emitted photoelectrons, the technique can map the electronic band structure and Fermi surfaces. ARPES is best suited for the study of one- or two-dimensional materials. It has been used by physicists to investigate high-temperature superconductors, graphene, topological materials, quantum well states, and materials exhibiting charge density waves.

The Planck constant, or Planck's constant, is a fundamental physical constant of foundational importance in quantum mechanics. The constant gives the relationship between the energy of a photon and its frequency, and by the mass-energy equivalence, the relationship between mass and frequency. Specifically, a photon's energy is equal to its frequency multiplied by the Planck constant. The constant is generally denoted by . The reduced Planck constant, or Dirac constant, equal to divided by , is denoted by .

References

  1. K. Popper (1959). The Logic of Scientific Discovery. London: Hutchinson. appendix *xii. ISBN   0-415-27844-9.
  2. Popper, Karl (1982). Quantum Theory and the Schism in Physics. London: Hutchinson (from 1992 published by Routledge). pp.  27–29. ISBN   0-8476-7019-8.
  3. Popper, Karl R. (1950). "Indeterminism in quantum physics and in classical physics". British Journal for the Philosophy of Science. 1 (2): 117–133. doi:10.1093/bjps/I.2.117.
  4. Lomonosov Moscow State University; Pechenkin, Alexander (2021-05-28). "The Ensemble Interpretation of Quantum Mechanics and Scientific Realism" (PDF). Acta Baltica Historiae et Philosophiae Scientiarum. 9 (1): 5–17. doi:10.11590/abhps.2021.1.01.
  5. Del Santo, Flavio (2019). "Karl Popper's Forgotten Role in the Quantum Debate at the Edge between Philosophy and Physics in 1950s and 1960s". Studies in History and Philosophy of Science Part B: Studies in History and Philosophy of Modern Physics. 67: 78. arXiv: 1811.00902 . Bibcode:2019SHPMP..67...78D. doi:10.1016/j.shpsb.2019.05.002. S2CID   53626865.
  6. 1 2 Del Santo, Flavio (2017). "Genesis of Karl Popper's EPR-Like Experiment and its Resonance amongst the Physics Community in the 1980s". Studies in History and Philosophy of Science Part B: Studies in History and Philosophy of Modern Physics. 62: 56–70. arXiv: 1701.09178 . Bibcode:2018SHPMP..62...56D. doi:10.1016/j.shpsb.2017.06.001. S2CID   119491612.
  7. Popper, Karl (1985). "Realism in quantum mechanics and a new version of the EPR experiment". In Tarozzi, G.; van der Merwe, A. (eds.). Open Questions in Quantum Physics. Dordrecht: Reidel. pp. 3–25. doi:10.1007/978-94-009-5245-4_1. ISBN   978-94-010-8816-9.
  8. William M. Shields (2012). "A Historical Survey of Sir Karl Popper's Contribution to Quantum Mechanics". Quanta. 1 (1): 1–12. doi: 10.12743/quanta.v1i1.4 .
  9. 1 2 Popper (1982), p. 27.
  10. 1 2 3 Popper (1982), p. 28.
  11. Popper (1982), p.29.
  12. A. Sudbery (1985). "Popper's variant of the EPR experiment does not test the Copenhagen interpretation". Philosophy of Science. 52 (3): 470–476. doi:10.1086/289261. S2CID   121976400.
  13. A. Sudbery (1988). "Testing interpretations of quantum mechanics". In Tarozzi, G.; van der Merwe, A. (eds.). Microphysical Reality and Quantum Formalism. Dordrecht: Kluwer. pp. 470–476.
  14. H. Krips (1984). "Popper, propensities, and the quantum theory". British Journal for the Philosophy of Science. 35 (3): 253–274. doi:10.1093/bjps/35.3.253.
  15. 1 2 M. J. Collet; R. Loudon (1987). "Analysis of a proposed crucial test of quantum mechanics". Nature. 326 (6114): 671–672. Bibcode:1987Natur.326..671C. doi:10.1038/326671a0. S2CID   31007584.
  16. M. Redhead (1996). "Popper and the quantum theory". In O'Hear, A. (ed.). Karl Popper: Philosophy and Problems . Cambridge: Cambridge University Press. pp.  163–176. ISBN   9780521558150.
  17. 1 2 Y.-H. Kim & Y. Shih (1999). "Experimental realization of Popper's experiment: violation of the uncertainty principle?". Foundations of Physics. 29 (12): 1849–1861. doi:10.1023/A:1018890316979. S2CID   189841160.
  18. C.S. Unnikrishnan (2002). "Is the quantum mechanical description of physical reality complete? Proposed resolution of the EPR puzzle". Foundations of Physics Letters. 15: 1–25. doi:10.1023/A:1015823125892.
  19. C.S. Unnikrishnan (2001). "Resolution of the Einstein-Podolsky-Rosen non-locality puzzle". In Sidharth, B.G.; Altaisky, M.V. (eds.). Frontiers of Fundamental Physics 4 . New York: Springer. pp.  145–160. Bibcode:2001ffpf.book.....S. ISBN   9781461513391.
  20. A. J. Short (2001). "Popper's experiment and conditional uncertainty relations". Foundations of Physics Letters. 14 (3): 275–284. doi:10.1023/A:1012238227977. S2CID   117154579.
  21. P. Sancho (2002). "Popper's Experiment Revisited". Foundations of Physics. 32 (5): 789–805. doi:10.1023/A:1016009127074. S2CID   84178335.
  22. Tabish Qureshi (2012). "Analysis of Popper's Experiment and Its Realization". Progress of Theoretical Physics. 127 (4): 645–656. arXiv: quant-ph/0505158 . Bibcode:2012PThPh.127..645Q. doi:10.1143/PTP.127.645. S2CID   119484882.
  23. 1 2 Tabish Qureshi (2005). "Understanding Popper's Experiment". American Journal of Physics. 73 (6): 541–544. arXiv: quant-ph/0405057 . Bibcode:2005AmJPh..73..541Q. doi:10.1119/1.1866098. S2CID   119437948.
  24. 1 2 Tabish Qureshi (2012). "Popper's Experiment: A Modern Perspective". Quanta. 1 (1): 19–32. arXiv: 1206.1432 . doi:10.12743/quanta.v1i1.8. S2CID   59483612.
  25. E. Gerjuoy; A.M. Sessler (2006). "Popper's experiment and communication". American Journal of Physics. 74 (7): 643–648. arXiv: quant-ph/0507121 . Bibcode:2006AmJPh..74..643G. doi:10.1119/1.2190684. S2CID   117564757.
  26. Ghirardi, GianCarlo; Marinatto, Luca; de Stefano, Francesco (2007). "Critical analysis of Popper's experiment". Physical Review A. 75 (4): 042107. arXiv: quant-ph/0702242 . Bibcode:2007PhRvA..75d2107G. doi:10.1103/PhysRevA.75.042107. ISSN   1050-2947. S2CID   119506558.