Ensemble interpretation

Last updated

The ensemble interpretation of quantum mechanics considers the quantum state description to apply only to an ensemble of similarly prepared systems, rather than supposing that it exhaustively represents an individual physical system. [1]

Contents

The advocates of the ensemble interpretation of quantum mechanics claim that it is minimalist, making the fewest physical assumptions about the meaning of the standard mathematical formalism. It proposes to take to the fullest extent the statistical interpretation of Max Born, for which he won the Nobel Prize in Physics in 1954. [2] On the face of it, the ensemble interpretation might appear to contradict the doctrine proposed by Niels Bohr, that the wave function describes an individual system or particle, not an ensemble, though he accepted Born's statistical interpretation of quantum mechanics. It is not quite clear exactly what kind of ensemble Bohr intended to exclude, since he did not describe probability in terms of ensembles. The ensemble interpretation is sometimes, especially by its proponents, called "the statistical interpretation", [1] but it seems perhaps different from Born's statistical interpretation.

As is the case for "the" Copenhagen interpretation, "the" ensemble interpretation might not be uniquely defined. In one view, the ensemble interpretation may be defined as that advocated by Leslie E. Ballentine, Professor at Simon Fraser University. [3] His interpretation does not attempt to justify, or otherwise derive, or explain quantum mechanics from any deterministic process, or make any other statement about the real nature of quantum phenomena; it intends simply to interpret the wave function. It does not propose to lead to actual results that differ from orthodox interpretations. It makes the statistical operator primary in reading the wave function, deriving the notion of a pure state from that. In the opinion of Ballentine, perhaps the most notable supporter of such an interpretation was Albert Einstein:

The attempt to conceive the quantum-theoretical description as the complete description of the individual systems leads to unnatural theoretical interpretations, which become immediately unnecessary if one accepts the interpretation that the description refers to ensembles of systems and not to individual systems.

Albert Einstein [4]

History

In his 1926 [5] paper introducing the concept of quantum scattering theory Max Born proposed to view "the motion of the particle follows the laws of probability, but the probability itself propagates in accord with causal laws", where the causal laws are Schrodinger's equations. As related in his 1954 Nobel Prize in Physics lecture [6] Born viewed the statistical character of quantum mechanics as an empirical observation with philosophical implications.

Einstein maintained consistently that the quantum mechanics only supplied a statistical view. In 1936 he wrote "“the function does not in any way describe a condition which could be that of a single system; it relates rather to many systems, to ‘an ensemble of systems’ in the sense of statistical mechanics.” [7] However Einstein did not provide a detailed study of the ensemble, ultimately because he considered quantum mechanics itself to be incomplete primarily because it was only an ensemble theory. [8] Einstein believed quantum mechanics was correct in the same sense that thermodynamics is correct, but that it was insufficient as means of unifying physics. [9]

Also in the years around 1936, Karl Popper published philosophical studies countering the work of Heisenberg and Bohr. Popper considered their work as essentially subjectivist, unfalsifiable, and thus unscientific. He held that the quantum state represented statistical assertions which have no predictive power for individual particles. [10] Popper described "propensities" as the correct notion of probability for quantum mechanics.

Although several other notable physicists championed the ensemble concept, including John C. Slater, Edwin C. Kemble, and Dmitrii Blokhintsev, [10] Leslie Ballentine's 1970 paper 'The statistical interpretation of quantum mechanics" [11] and his textbook [12] have become the main sources. [8] [10] Ballentine followed up with axiomatic development of propensity theory, [13] analysis of decoherence in the ensemble interpretation [14] and other papers spanning 40 years.

States, systems, and ensembles

Perhaps the first expression of an ensemble interpretation was that of Max Born. [5] In a 1968 article, he used the German words 'gleicher Haufen', which are often translated into English, in this context, as 'ensemble' or 'assembly'. The atoms in his assembly were uncoupled, meaning that they were an imaginary set of independent atoms that defines its observable statistical properties. Born did not develop a more detailed specification of ensembles to complete his scattering theory work.

Although Einstein described quantum mechanics as clearly an ensemble theory he did present a formal definition of an ensemble. [15] Einstein sought a theory of individual entities, which he argued was not quantum mechanics.

Ballentine distinguish his particular ensemble interpretation the Statistical Interpretation. According to Ballentine, the distinguishing difference between many of the Copenhagen-like interpretations (CI) and the Statistical Interpretation (EI) is the following: [11]

CI: A pure state provides a complete description of an individual system, e.g. an electron.

EI: A pure state describes the statistical properties of an ensemble of identically prepared systems.

Ballentine defines a quantum state as an ensemble of similarly prepared systems. For example, the system may be a single electron, then the ensemble will "the set of all single electrons which subjected to the same state preparation technique." He uses the example of a low-intensity electron beam prepared with a narrow range of momenta. Each prepared electron is a system, the ensemble consists of many such systems.

Ballentine emphasizes that the meaning of the "Quantum State" or "State Vector" may be described, essentially, by a one-to-one correspondence to the probability distributions of measurement results, not the individual measurement results themselves. [16] A mixed state is a description only of the probabilities, and of positions, not a description of actual individual positions. A mixed state is a mixture of probabilities of physical states, not a coherent superposition of physical states.

Probability; propensity

Quantum observations are inherently statistical. For example, the electrons in a low-intensity double slit experiment arrive at random times and seemingly random places and yet eventually show an interference pattern.

Matter wave double slit diffraction pattern building up electron by electron. Each white dot represents a single electron hitting a detector; with a statistically large number of electrons interference fringes appear. Electron buildup movie from "Controlled double-slit electron diffraction" Roger Bach et al 2013 New J. Phys. 15 033018.gif
Matter wave double slit diffraction pattern building up electron by electron. Each white dot represents a single electron hitting a detector; with a statistically large number of electrons interference fringes appear.

The theory of quantum mechanics offer only statistical results. Given that we have prepared a system in a state , the theory predicts a result as a probability distribution:

.

Different approaches to probability can be applied to connect the probability distribution of theory to the observed randomness.

Popper, [18] Ballentine, [13] Paul Humphreys, [19] and others [20] point to propensity as the correct interpretation of probability in science. Propensity, a form of causality that is weaker than determinism, is the tendency of a physical system to produce a result. [21] Thus the mathematical statement

means the propensity for event to occur given the physical scenario is . The physical scenario is view as weakly causal condition.

The weak causation invalidates Bayes' theorem and correlation is no longer symmetric. [19] As noted by Paul Humphreys, many physical examples show the lack of reciprocal correlation, for example, the propensity for smokers to get lung cancer does not imply lung cancer has a propensity to cause smoking.

Propensity closely matches the application of quantum theory: single event probability can be predicted by theory but only verified by repeated samples in experiment. Popper explicitly developed propensity theory to eliminate subjectivity in quantum mechanics. [20]

Preparative and observing devices as origins of quantum randomness

An isolated quantum mechanical system, specified by a wave function, evolves in time in a deterministic way according to the Schrödinger equation that is characteristic of the system. Though the wave function can generate probabilities, no randomness or probability is involved in the temporal evolution of the wave function itself. This is agreed, for example, by Born, [22] Dirac, [23] von Neumann, [24] London & Bauer, [25] Messiah, [26] and Feynman & Hibbs. [27] An isolated system is not subject to observation; in quantum theory, this is because observation is an intervention that violates isolation.

The system's initial state is defined by the preparative procedure; this is recognized in the ensemble interpretation, as well as in the Copenhagen approach. [28] [29] [30] [31] The system's state as prepared, however, does not entirely fix all properties of the system. The fixing of properties goes only as far as is physically possible, and is not physically exhaustive; it is, however, physically complete in the sense that no physical procedure can make it more detailed. This is stated clearly by Heisenberg in his 1927 paper. [32] It leaves room for further unspecified properties. [33] For example, if the system is prepared with a definite energy, then the quantum mechanical phase of the wave function is left undetermined by the mode of preparation. The ensemble of prepared systems, in a definite pure state, then consists of a set of individual systems, all having one and the same definite energy, but each having a different quantum mechanical phase, regarded as probabilistically random. [34] The wave function, however, does have a definite phase, and thus specification by a wave function is more detailed than specification by state as prepared. The members of the ensemble are logically distinguishable by their distinct phases, though the phases are not defined by the preparative procedure. The wave function can be multiplied by a complex number of unit magnitude without changing the state as defined by the preparative procedure.

The preparative state, with unspecified phase, leaves room for the several members of the ensemble to interact in respectively several various ways with other systems. An example is when an individual system is passed to an observing device so as to interact with it. Individual systems with various phases are scattered in various respective directions in the analyzing part of the observing device, in a probabilistic way. In each such direction, a detector is placed, in order to complete the observation. When the system hits the analyzing part of the observing device, that scatters it, it ceases to be adequately described by its own wave function in isolation. Instead it interacts with the observing device in ways partly determined by the properties of the observing device. In particular, there is in general no phase coherence between system and observing device. This lack of coherence introduces an element of probabilistic randomness to the system–device interaction. It is this randomness that is described by the probability calculated by the Born rule. There are two independent originative random processes, one that of preparative phase, the other that of the phase of the observing device. The random process that is actually observed, however, is neither of those originative ones. It is the phase difference between them, a single derived random process.

The Born rule describes that derived random process, the observation of a single member of the preparative ensemble. In the ordinary language of classical or Aristotelian scholarship, the preparative ensemble consists of many specimens of a species. The quantum mechanical technical term 'system' refers to a single specimen, a particular object that may be prepared or observed. Such an object, as is generally so for objects, is in a sense a conceptual abstraction, because, according to the Copenhagen approach, it is defined, not in its own right as an actual entity, but by the two macroscopic devices that should prepare and observe it. The random variability of the prepared specimens does not exhaust the randomness of a detected specimen. Further randomness is injected by the quantum randomness of the observing device. It is this further randomness that makes Bohr emphasize that there is randomness in the observation that is not fully described by the randomness of the preparation. This is what Bohr means when he says that the wave function describes "a single system". He is focusing on the phenomenon as a whole, recognizing that the preparative state leaves the phase unfixed, and therefore does not exhaust the properties of the individual system. The phase of the wave function encodes further detail of the properties of the individual system. The interaction with the observing device reveals that further encoded detail. It seems that this point, emphasized by Bohr, is not explicitly recognized by the ensemble interpretation, and this may be what distinguishes the two interpretations. It seems, however, that this point is not explicitly denied by the ensemble interpretation.

Einstein perhaps sometimes seemed to interpret the probabilistic "ensemble" as a preparative ensemble, recognizing that the preparative procedure does not exhaustively fix the properties of the system; therefore he said that the theory is "incomplete". Bohr, however, insisted that the physically important probabilistic "ensemble" was the combined prepared-and-observed one. Bohr expressed this by demanding that an actually observed single fact should be a complete "phenomenon", not a system alone, but always with reference to both the preparing and the observing devices. The Einstein–Podolsky–Rosen criterion of "completeness" is clearly and importantly different from Bohr's. Bohr regarded his concept of "phenomenon" as a major contribution that he offered for quantum theoretical understanding. [35] [36] The decisive randomness comes from both preparation and observation, and may be summarized in a single randomness, that of the phase difference between preparative and observing devices. The distinction between these two devices is an important point of agreement between Copenhagen and ensemble interpretations. Though Ballentine claims that Einstein advocated "the ensemble approach", a detached scholar would not necessarily be convinced by that claim of Ballentine. There is room for confusion about how "the ensemble" might be defined.

"Each photon interferes only with itself"

Niels Bohr famously insisted that the wave function refers to a single individual quantum system. He was expressing the idea that Dirac expressed when he famously wrote: "Each photon then interferes only with itself. Interference between different photons never occurs.". [37] Dirac clarified this by writing: "This, of course, is true only provided the two states that are superposed refer to the same beam of light, i.e. all that is known about the position and momentum of a photon in either of these states must be the same for each." [38] Bohr wanted to emphasize that a superposition is different from a mixture. He seemed to think that those who spoke of a "statistical interpretation" were not taking that into account. To create, by a superposition experiment, a new and different pure state, from an original pure beam, one can put absorbers and phase-shifters into some of the sub-beams, so as to alter the composition of the re-constituted superposition. But one cannot do so by mixing a fragment of the original unsplit beam with component split sub-beams. That is because one photon cannot both go into the unsplit fragment and go into the split component sub-beams. Bohr felt that talk in statistical terms might hide this fact.

The physics here is that the effect of the randomness contributed by the observing apparatus depends on whether the detector is in the path of a component sub-beam, or in the path of the single superposed beam. This is not explained by the randomness contributed by the preparative device.

Measurement and collapse

Bras and kets

The ensemble interpretation is notable for its relative de-emphasis on the duality and theoretical symmetry between bras and kets. The approach emphasizes the ket as signifying a physical preparation procedure. [39] There is little or no expression of the dual role of the bra as signifying a physical observational procedure. The bra is mostly regarded as a mere mathematical object, without very much physical significance. It is the absence of the physical interpretation of the bra that allows the ensemble approach to by-pass the notion of "collapse". Instead, the density operator expresses the observational side of the ensemble interpretation. It hardly needs saying that this account could be expressed in a dual way, with bras and kets interchanged, mutatis mutandis. In the ensemble approach, the notion of the pure state is conceptually derived by analysis of the density operator, rather than the density operator being conceived as conceptually synthesized from the notion of the pure state.

An attraction of the ensemble interpretation is that it appears to dispense with the metaphysical issues associated with reduction of the state vector, Schrödinger cat states, and other issues related to the concepts of multiple simultaneous states. The ensemble interpretation postulates that the wave function only applies to an ensemble of systems as prepared, but not observed. There is no recognition of the notion that a single specimen system could manifest more than one state at a time, as assumed, for example, by Dirac. [40] Hence, the wave function is not envisaged as being physically required to be "reduced". This can be illustrated by an example:

Consider a quantum die. If this is expressed in Dirac notation, the "state" of the die can be represented by a "wave" function describing the probability of an outcome given by:

Where the "+" sign of a probabilistic equation is not an addition operator, it is the standard probabilistic Boolean operator OR. The state vector is inherently defined as a probabilistic mathematical object such that the result of a measurement is one outcome OR another outcome.

It is clear that on each throw, only one of the states will be observed, but this is not expressed by a bra. Consequently, there appears to be no requirement for a notion of collapse of the wave function/reduction of the state vector, or for the die to physically exist in the summed state. In the ensemble interpretation, wave function collapse would make as much sense as saying that the number of children a couple produced, collapsed to 3 from its average value of 2.4.

The state function is not taken to be physically real, or be a literal summation of states. The wave function, is taken to be an abstract statistical function, only applicable to the statistics of repeated preparation procedures. The ket does not directly apply to a single particle detection, but only the statistical results of many. This is why the account does not refer to bras, and mentions only kets.

Diffraction

The ensemble approach differs significantly from the Copenhagen approach in its view of diffraction. The Copenhagen interpretation of diffraction, especially in the viewpoint of Niels Bohr, puts weight on the doctrine of wave–particle duality. In this view, a particle that is diffracted by a diffractive object, such as for example a crystal, is regarded as really and physically behaving like a wave, split into components, more or less corresponding to the peaks of intensity in the diffraction pattern. Though Dirac does not speak of wave–particle duality, he does speak of "conflict" between wave and particle conceptions. [41] He indeed does describe a particle, before it is detected, as being somehow simultaneously and jointly or partly present in the several beams into which the original beam is diffracted. So does Feynman, who speaks of this as "mysterious". [42]

The ensemble approach points out that this seems perhaps reasonable for a wave function that describes a single particle, but hardly makes sense for a wave function that describes a system of several particles. The ensemble approach demystifies this situation along the lines advocated by Alfred Landé, accepting Duane's hypothesis. In this view, the particle really and definitely goes into one or other of the beams, according to a probability given by the wave function appropriately interpreted. There is definite quantal transfer of translative momentum between particle and diffractive object. [43] This is recognized also in Heisenberg's 1930 textbook, [44] though usually not recognized as part of the doctrine of the so-called "Copenhagen interpretation". This gives a clear and utterly non-mysterious physical or direct explanation instead of the debated concept of wave function "collapse". It is presented in terms of quantum mechanics by other present day writers also, for example, Van Vliet. [45] [46] For those who prefer physical clarity rather than mysterianism, this is an advantage of the ensemble approach, though it is not the sole property of the ensemble approach. With a few exceptions, [44] [47] [48] [49] [50] [51] [52] this demystification is not recognized or emphasized in many textbooks and journal articles.

Criticism

David Mermin sees the ensemble interpretation as being motivated by an adherence ("not always acknowledged") to classical principles.

"[...] the notion that probabilistic theories must be about ensembles implicitly assumes that probability is about ignorance. (The 'hidden variables' are whatever it is that we are ignorant of.) But in a non-deterministic world probability has nothing to do with incomplete knowledge, and ought not to require an ensemble of systems for its interpretation".

However, according to Einstein and others, a key motivation for the ensemble interpretation is not about any alleged, implicitly assumed probabilistic ignorance, but the removal of "…unnatural theoretical interpretations…". A specific example being the Schrödinger cat problem stated above, but this concept applies to any system where there is an interpretation that postulates, for example, that an object might exist in two positions at once.

Mermin also emphasises the importance of describing single systems, rather than ensembles.

"The second motivation for an ensemble interpretation is the intuition that because quantum mechanics is inherently probabilistic, it only needs to make sense as a theory of ensembles. Whether or not probabilities can be given a sensible meaning for individual systems, this motivation is not compelling. For a theory ought to be able to describe as well as predict the behavior of the world. The fact that physics cannot make deterministic predictions about individual systems does not excuse us from pursuing the goal of being able to describe them as they currently are." [53]

Schrödinger's cat

The ensemble interpretation states that superpositions are nothing but subensembles of a larger statistical ensemble. That being the case, the state vector would not apply to individual cat experiments, but only to the statistics of many similar prepared cat experiments. Proponents of this interpretation state that this makes the Schrödinger's cat paradox a trivial non-issue. However, the application of state vectors to individual systems, rather than ensembles, has claimed explanatory benefits, in areas like single-particle twin-slit experiments and quantum computing (see Schrödinger's cat applications). As an avowedly minimalist approach, the ensemble interpretation does not offer any specific alternative explanation for these phenomena.

The frequentist probability variation

The claim that the wave functional approach fails to apply to single particle experiments cannot be taken as a claim that quantum mechanics fails in describing single-particle phenomena. In fact, it gives correct results within the limits of a probabilistic or stochastic theory.

Probability always requires a set of multiple data, and thus single-particle experiments are really part of an ensemble an ensemble of individual experiments that are performed one after the other over time. In particular, the interference fringes seen in the double-slit experiment require repeated trials to be observed.

The quantum Zeno effect

Leslie Ballentine promoted the ensemble interpretation in his book Quantum Mechanics, A Modern Development. In it, [54] he described what he called the "Watched Pot Experiment". His argument was that, under certain circumstances, a repeatedly measured system, such as an unstable nucleus, would be prevented from decaying by the act of measurement itself. He initially presented this as a kind of reductio ad absurdum of wave function collapse. [55]

The effect has been shown to be real. Ballentine later wrote papers claiming that it could be explained without wave function collapse. [56]

See also

Related Research Articles

The Copenhagen interpretation is a collection of views about the meaning of quantum mechanics, stemming from the work of Niels Bohr, Werner Heisenberg, Max Born, and others. The term "Copenhagen interpretation" was apparently coined by Heisenberg during the 1950s to refer to ideas developed in the 1925–1927 period, glossing over his disagreements with Bohr. Consequently, there is no definitive historical statement of what the interpretation entails.

<span class="mw-page-title-main">Many-worlds interpretation</span> Interpretation of quantum mechanics that denies the collapse of the wavefunction

The many-worlds interpretation (MWI) is a philosophical position about how the mathematics used in quantum mechanics relates to physical reality. It asserts that the universal wavefunction is objectively real, and that there is no wave function collapse. This implies that all possible outcomes of quantum measurements are physically realized in some "world" or universe. In contrast to some other interpretations, the evolution of reality as a whole in MWI is rigidly deterministic and local. Many-worlds is also called the relative state formulation or the Everett interpretation, after physicist Hugh Everett, who first proposed it in 1957. Bryce DeWitt popularized the formulation and named it many-worlds in the 1970s.

The mathematical formulations of quantum mechanics are those mathematical formalisms that permit a rigorous description of quantum mechanics. This mathematical formalism uses mainly a part of functional analysis, especially Hilbert spaces, which are a kind of linear space. Such are distinguished from mathematical formalisms for physics theories developed prior to the early 1900s by the use of abstract mathematical structures, such as infinite-dimensional Hilbert spaces, and operators on these spaces. In brief, values of physical observables such as energy and momentum were no longer considered as values of functions on phase space, but as eigenvalues; more precisely as spectral values of linear operators in Hilbert space.

<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 describes the behavior 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">Schrödinger's cat</span> Thought experiment in quantum mechanics

In quantum mechanics, Schrödinger's cat is a thought experiment, sometimes described as a paradox, of quantum superposition. In the thought experiment, a hypothetical cat may be considered simultaneously both alive and dead, while it is unobserved in a closed box, as a result of its fate being linked to a random subatomic event that may or may not occur. This thought experiment was devised by physicist Erwin Schrödinger in 1935 in a discussion with Albert Einstein to illustrate what Schrödinger saw as the problems of the Copenhagen interpretation of quantum mechanics.

In physics, statistical mechanics is a mathematical framework that applies statistical methods and probability theory to large assemblies of microscopic entities. It does not assume or postulate any natural laws, but explains the macroscopic behavior of nature from the behavior of such ensembles.

An interpretation of quantum mechanics is an attempt to explain how the mathematical theory of quantum mechanics might correspond to experienced reality. Although quantum mechanics has held up to rigorous and extremely precise tests in an extraordinarily broad range of experiments, there exist a number of contending schools of thought over their interpretation. These views on interpretation differ on such fundamental questions as whether quantum mechanics is deterministic or stochastic, local or non-local, which elements of quantum mechanics can be considered real, and what the nature of measurement is, among other matters.

In quantum mechanics, wave function collapse occurs when a wave function—initially in a superposition of several eigenstates—reduces to a single eigenstate due to interaction with the external world. This interaction is called an observation, and is the essence of a measurement in quantum mechanics, which connects the wave function with classical observables such as position and momentum. Collapse is one of the two processes by which quantum systems evolve in time; the other is the continuous evolution governed by the Schrödinger equation. Collapse is a black box for a thermodynamically irreversible interaction with a classical environment.

<span class="mw-page-title-main">Quantum superposition</span> Principle of quantum mechanics

Quantum superposition is a fundamental principle of quantum mechanics. In classical mechanics, things like position or momentum are always well-defined. It may not be known what they are at any given time, but that is an issue of understanding and not an issue of the physical system. A quantum system interacts in ways that can be explained with superposition of different discrete states. Measurements of quantum systems give a statistical result corresponding to any one of the possible states appearing at random.

In philosophy, philosophy of physics deals with conceptual and interpretational issues in modern physics, many of which overlap with research done by certain kinds of theoretical physicists. Philosophy of physics can be broadly divided into three areas:

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

In quantum physics, a wave function is a mathematical description of the quantum state of an isolated quantum system. The most common symbols for a wave function are the Greek letters ψ and Ψ. Wave functions are composed of complex numbers. For example, a wave function might assign a complex number to each point in a region of space. The Born rule provides the means to turn these complex probability amplitudes into actual probabilities. In one common form, it says that the squared modulus of a wave function that depends upon position is the probability density of measuring a particle as being at a given place. The integral of a wavefunction's squared modulus over all the system's degrees of freedom must be equal to 1, a condition called normalization. Since the wave function is complex-valued, only its relative phase and relative magnitude can be measured; its value does not, in isolation, tell anything about the magnitudes or directions of measurable observables. One has to apply quantum operators, whose eigenvalues correspond to sets of possible results of measurements, to the wave function ψ and calculate the statistical distributions for measurable quantities.

In physics, a hidden-variable theory is a deterministic physical model which seeks to explain the probabilistic nature of quantum mechanics by introducing additional variables.

<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 square of the modulus of this quantity represents a probability density.

In quantum mechanics, the measurement problem is the problem of how, or whether, wave function collapse occurs. The inability to observe such a collapse directly has given rise to different interpretations of quantum mechanics and poses a key set of questions that each interpretation must answer.

<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.

Quantum mechanics is the study of matter and its interactions with energy on the scale of atomic and subatomic particles. By contrast, classical physics explains matter and energy only on a scale familiar to human experience, including the behavior of astronomical bodies such as the moon. Classical physics is still used in much of modern science and technology. However, towards the end of the 19th century, scientists discovered phenomena in both the large (macro) and the small (micro) worlds that classical physics could not explain. The desire to resolve inconsistencies between observed phenomena and classical theory led to a revolution in physics, a shift in the original scientific paradigm: the development of quantum mechanics.

A first quantization of a physical system is a possibly semiclassical treatment of quantum mechanics, in which particles or physical objects are treated using quantum wave functions but the surrounding environment is treated classically.

The history of quantum mechanics is a fundamental part of the history of modern physics. The major chapters of this history begin with the emergence of quantum ideas to explain individual phenomena -- blackbody radiation, the photoelectric effect, solar emission spectra -- an era called the Old or Older quantum theories. Building on the technology developed in classical mechanics, the invention of wave mechanics by Erwin Schrödinger and expanded by many others triggers the "modern" era beginning around 1925. Paul Dirac's relativistic quantum theory work lead him to explore quantum theories of radiation, culminating in quantum electrodynamics, the first quantum field theory. The history of quantum mechanics continues in the history of quantum field theory. The history of quantum chemistry, theoretical basis of chemical structure, reactivity, and bonding, interlaces with the events discussed in this article.

In quantum physics, a quantum state is a mathematical entity that embodies the knowledge of a quantum system. Quantum mechanics specifies the construction, evolution, and measurement of a quantum state. The result is a quantum mechanical prediction for the system represented by the state. Knowledge of the quantum state, and the quantum mechanical rules for the system's evolution in time, exhausts all that can be known about a quantum system.

The von Neumann–Wigner interpretation, also described as "consciousness causes collapse", is an interpretation of quantum mechanics in which consciousness is postulated to be necessary for the completion of the process of quantum measurement.

References

  1. 1 2 Ballentine, L. E. (1970-10-01). "The Statistical Interpretation of Quantum Mechanics". Reviews of Modern Physics. American Physical Society (APS). 42 (4): 358–381. Bibcode:1970RvMP...42..358B. doi:10.1103/revmodphys.42.358. ISSN   0034-6861. S2CID   120024263.
  2. "The statistical interpretation of quantum mechanics" (PDF). Nobel Lecture. December 11, 1954.
  3. Leslie E. Ballentine (1998). Quantum Mechanics: A Modern Development. World Scientific. Chapter 9. ISBN   981-02-4105-4.
  4. Einstein: Philosopher-Scientist, edited by Paul Arthur Schilpp (Tudor Publishing Company, 1957), p. 672.
  5. 1 2 Born, Max (1926). "Zur Quantenmechanik der Stoßvorgänge". Zeitschrift für Physik (in German). Springer Science and Business Media LLC. 37 (12): 863–867. Bibcode:1926ZPhy...37..863B. doi:10.1007/bf01397477. ISSN   1434-6001. S2CID   119896026.; English translation by Gunter Ludwig, pp. 206–225, 'On the quantum mechanics of collisions', in Wave Mechanics (1968), Pergamon, Oxford UK.
  6. "The statistical interpretation of quantum mechanics" (PDF). Nobel Lecture. December 11, 1954.
  7. Einstein, Albert (March 1936). "Physics and reality". Journal of the Franklin Institute. 221 (3): 349–382. Bibcode:1936FrInJ.221..349E. doi:10.1016/S0016-0032(36)91047-5.
  8. 1 2 Home, D; Whitaker, M.A.B (January 1992). "Ensemble interpretations of quantum mechanics. A modern perspective". Physics Reports. 210 (4): 223–317. Bibcode:1992PhR...210..223H. doi:10.1016/0370-1573(92)90088-H.
  9. Ballentine, L. E. (1972-12-01). "Einstein's Interpretation of Quantum Mechanics". American Journal of Physics. 40 (12): 1763–1771. Bibcode:1972AmJPh..40.1763B. doi: 10.1119/1.1987060 . ISSN   0002-9505.
  10. 1 2 3 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. S2CID   236412344.
  11. 1 2 Ballentine, L. E. (1970-10-01). "The Statistical Interpretation of Quantum Mechanics". Reviews of Modern Physics. 42 (4): 358–381. Bibcode:1970RvMP...42..358B. doi:10.1103/RevModPhys.42.358. ISSN   0034-6861. S2CID   120024263.
  12. Ballentine, Leslie E (2014). Quantum Mechanics: A Modern Development (2 ed.). WORLD SCIENTIFIC. doi:10.1142/9038. ISBN   978-981-4578-57-8.
  13. 1 2 3 Ballentine, Leslie E. (August 2016). "Propensity, Probability, and Quantum Theory". Foundations of Physics. 46 (8): 973–1005. Bibcode:2016FoPh...46..973B. doi:10.1007/s10701-016-9991-0. ISSN   0015-9018. S2CID   254508686.
  14. Ballentine, Leslie (October 2008). "Classicality without Decoherence: A Reply to Schlosshauer". Foundations of Physics. 38 (10): 916–922. Bibcode:2008FoPh...38..916B. doi: 10.1007/s10701-008-9242-0 . ISSN   0015-9018. S2CID   122887738.
  15. Home, D. (1997). Conceptual Foundations of Quantum Physics: An Overview from Modern Perspectives, Springer, New York, ISBN   978-1-4757-9810-4, p. 362: "Einstein's references to the ensemble interpretation remained in general rather sketchy."
  16. Quantum Mechanics, A Modern Development, p. 48.
  17. Bach, Roger; Pope, Damian; Liou, Sy-Hwang; Batelaan, Herman (2013-03-13). "Controlled double-slit electron diffraction". New Journal of Physics. IOP Publishing. 15 (3): 033018. arXiv: 1210.6243 . Bibcode:2013NJPh...15c3018B. doi:10.1088/1367-2630/15/3/033018. ISSN   1367-2630. S2CID   832961.
  18. Popper, Karl R. “The Propensity Interpretation of Probability.” The British Journal for the Philosophy of Science, vol. 10, no. 37, 1959, pp. 25–42. JSTOR, http://www.jstor.org/stable/685773. Accessed 27 Aug. 2023.
  19. 1 2 Humphreys, Paul (October 1985). "Why Propensities Cannot be Probabilities". The Philosophical Review. 94 (4): 557–570. doi:10.2307/2185246. JSTOR   2185246. S2CID   55871596.
  20. 1 2 Berkovitz, Joseph. “The Propensity Interpretation of Probability: A Re-Evaluation.” Erkenntnis (1975-) 80 (2015): 629–711. http://www.jstor.org/stable/24735118.
  21. For inferential probability, the arguments may, in principle, be any propositions. But for ensemble probability, the first argument A must be an event, and the second argument C must describe a repeatable procedure that can generate an ensemble of outcome events. [13]
  22. Born, Max (1951). "Physics in the Last Fifty Years". Nature. Springer Science and Business Media LLC. 168 (4276): 625–630. Bibcode:1951Natur.168..625B. doi:10.1038/168625a0. ISSN   0028-0836. S2CID   38153029.: 630: "We have accustomed ourselves to abandon deterministic causality for atomic events; but we have still retained the belief that probability spreads in space (multi-dimensional) and time according to deterministic laws in the form of differential equations."
  23. Dirac, Paul Adrien Maurice (1927). "The physical interpretation of the quantum dynamics". Proceedings of the Royal Society of London. Series A, Containing Papers of a Mathematical and Physical Character. The Royal Society. 113 (765): 621–641. Bibcode:1927RSPSA.113..621D. doi: 10.1098/rspa.1927.0012 . ISSN   0950-1207. S2CID   122855515.: "One can suppose that the initial state of a system determines definitely the state of the system at any subsequent time. ... The notion of probabilities does not enter into the ultimate description of mechanical processes."
  24. J. von Neumann (1932). Mathematische Grundlagen der Quantenmechanik (in German). Berlin: Springer. Translated as J. von Neumann (1955). Mathematical Foundations of Quantum Mechanics . Princeton NJ: Princeton University Press. P. 349: "... the time dependent Schrödinger differential equation ... describes how the system changes continuously and causally."
  25. London, F., Bauer, E. (1939). La Théorie de l'Observation dans la Mécanique Quantique, issue 775 of Actualités Scientifiques et Industrielles, section Exposés de Physique Générale, directed by Paul Langevin, Hermann & Cie, Paris, translated by Shimony, A., Wheeler, J.A., Zurek, W.H., McGrath, J., McGrath, S.M. (1983), at pp. 217–259 in Wheeler, J.A., Zurek, W.H. editors (1983). Quantum Theory and Measurement, Princeton University Press, Princeton NJ; p. 232: "... the Schrödinger equation has all the features of a causal connection."
  26. Messiah, A. (1961). Quantum Mechanics, volume 1, translated by G.M. Temmer from the French Mécanique Quantique, North-Holland, Amsterdam, p. 61: "... specifying Ψ at a given initial instant uniquely defines its entire later evolution, in accord with the hypothesis that the dynamical state of the system is entirely determined once Ψ is given."
  27. Feynman, R.P., Hibbs, A. (1965). Quantum Mechanics and Path Integrals, McGraw–Hill, New York, p. 22: "the amultitudes φ are solutions of a completely deterministic equation (the Schrödinger equation)."
  28. Dirac, P.A.M. (1940). The Principles of Quantum Mechanics, fourth edition, Oxford University Press, Oxford UK, pages 11–12: "A state of a system may be defined as an undisturbed motion that is restricted by as many conditions or data as are theoretically possible without mutual interference or contradiction. In practice, the conditions could be imposed by a suitable preparation of the system, consisting perhaps of passing it through various kinds of sorting apparatus, such as slits and polarimeters, the system being undisturbed after preparation."
  29. Messiah, A. (1961). Quantum Mechanics, volume 1, translated by G.M. Temmer from the French Mécanique Quantique, North-Holland, Amsterdam, pp. 204–205: "When the preparation is complete, and consequently the dynamical state of the system is completely known, one says that one is dealing with a pure state, in contrast to the statistical mixtures which characterize incomplete preparations."
  30. L. E., Ballentine (1998). Quantum Mechanics: A Modern Development. Singapore: World Scientific. p. Chapter 9. ISBN   981-02-4105-4. P.  46: "Any repeatable process that yields well-defined probabilities for all observables may be termed a state preparation procedure."
  31. Jauch, J.M. (1968). Foundations of Quantum Mechanics, Addison–Wesley, Reading MA; p. 92: "Two states are identical if the relevant conditions in the preparation of the state are identical; p. 93: "Thus, a state of a quantum system can only be measured if the system can be prepared an unlimited number of times in the same state."
  32. Heisenberg, W. (1927). "Über den anschaulichen Inhalt der quantentheoretischen Kinematik und Mechanik". Zeitschrift für Physik (in German). Springer Science and Business Media LLC. 43 (3–4): 172–198. Bibcode:1927ZPhy...43..172H. doi:10.1007/bf01397280. ISSN   1434-6001. S2CID   122763326. Translation as 'The actual content of quantum theoretical kinematics and mechanics'. Also translated as 'The physical content of quantum kinematics and mechanics' at pp. 62–84 by editors John Wheeler and Wojciech Zurek, in Quantum Theory and Measurement (1983), Princeton University Press, Princeton NJ: "Even in principle we cannot know the present [state] in all detail."
  33. London, F., Bauer, E. (1939). La Théorie de l'Observation dans la Mécanique Quantique, issue 775 of Actualités Scientifiques et Industrielles, section Exposés de Physique Générale, directed by Paul Langevin, Hermann & Cie, Paris, translated by Shimony, A., Wheeler, J.A., Zurek, W.H., McGrath, J., McGrath, S.M. (1983), at pp. 217–259 in Wheeler, J.A., Zurek, W.H. editors (1983). Quantum Theory and Measurement, Princeton University Press, Princeton NJ; p. 235: "ignorance about the phases".
  34. Dirac, Paul (1926). "On the theory of quantum mechanics". Proceedings of the Royal Society of London. Series A, Containing Papers of a Mathematical and Physical Character. The Royal Society. 112 (762): 661–677. Bibcode:1926RSPSA.112..661D. doi: 10.1098/rspa.1926.0133 . ISSN   0950-1207. S2CID   122757433., p. 677: "The following argument shows, however, that the initial phases are of real physical importance, and that in consequence the Einstein coefficients are inadequate to describe the phenomena except in special cases."
  35. Bohr, Niels (1948). "On the notions of complementarity and causality". Dialectica. Verein philosophie. 2 (3–4): 312–319. doi:10.1111/j.1746-8361.1948.tb00703.x. ISSN   0012-2017.: "As a more appropriate way of expression, one may advocate limitation of the use of the word phenomenon to refer to observations obtained under specified circumstances, including an account of the whole experiment."
  36. Rosenfeld, L. (1967).'Niels Bohr in the thirties: Consolidation and extension of the conception of complementarity', pp. 114–136 in Niels Bohr: His life and work as seen by his friends and colleagues, edited by S. Rozental, North Holland, Amsterdam; p. 124: "As a direct consequence of this situation it is now highly necessary, in the definition of any phenomenon, to specify the conditions of its observation, the kind of apparatus determining the particular aspect of the phenomenon we wish to observe; and we have to face the fact that different conditions of observation may well be incompatible with each other to the extent indicated by indeterminacy relations of the Heisenberg type."
  37. Dirac, P.A.M., The Principles of Quantum Mechanics, (1930), 1st edition, p. 15; (1935), 2nd edition, p. 9; (1947), 3rd edition, p. 9; (1958), 4th edition, p. 9.
  38. Dirac, P.A.M., The Principles of Quantum Mechanics, (1930), 1st edition, p. 8.
  39. Ballentine, L.E. (1998). Quantum Mechanics: a Modern Development, World Scientific, Singapore, p. 47: "The quantum state description may be taken to refer to an ensemble of similarly prepared systems."
  40. Dirac, P.A.M. (1958). The Principles of Quantum Mechanics, 4th edition, Oxford University Press, Oxford UK, p. 12: "The general principle of superposition of quantum mechanics applies to the states, with either of the above meanings, of any one dynamical system. It requires us to assume that between these states there exist peculiar relationships such that whenever the system is definitely in one state we can consider it as being partly in each of two or more other states."
  41. Dirac, P.A.M. (1958). The Principles of Quantum Mechanics, 4th edition, Oxford University Press, Oxford UK, p. 8.
  42. Feynman, R.P., Leighton, R.B., Sands, M. (1965). The Feynman Lectures on Physics, volume 3, Addison-Wesley, Reading, MA, p. 1–1. Accessed 2020-04-29.
  43. Ballentine, L.E. (1998). Quantum Mechanics: a Modern Development, World Scientific, Singapore, ISBN   981-02-2707-8, p. 136.
  44. 1 2 Heisenberg, W. (1930). The Physical Principles of the Quantum Theory, translated by C. Eckart and F.C. Hoyt, University of Chicago Press, Chicago, pp. 77–78.
  45. Van Vliet, K. (1967). Linear momentum quantization in periodic structures, Physica, 35: 97–106, doi:10.1016/0031-8914(67)90138-3.
  46. Van Vliet, K. (2010). Linear momentum quantization in periodic structures ii, Physica A, 389: 1585–1593, doi:10.1016/j.physa.2009.12.026.
  47. Pauling, L.C., Wilson, E.B. (1935). Introduction to Quantum Mechanics: with Applications to Chemistry, McGraw-Hill, New York, pp. 34–36.
  48. Landé, A. (1951). Quantum Mechanics, Sir Isaac Pitman and Sons, London, pp. 19–22.
  49. Bohm, D. (1951). Quantum Theory, Prentice Hall, New York, pp. 71–73.
  50. Thankappan, V.K. (1985/2012). Quantum Mechanics, third edition, New Age International, New Delhi, ISBN   978-81-224-3357-9, pp. 6–7.
  51. Schmidt, L. Ph. H.; Lower, J.; Jahnke, T.; Schößler, S.; Schöffler, M. S.; et al. (2013-09-05). "Momentum Transfer to a Free Floating Double Slit: Realization of a Thought Experiment from the Einstein-Bohr Debates". Physical Review Letters. American Physical Society (APS). 111 (10): 103201. Bibcode:2013PhRvL.111j3201S. doi:10.1103/physrevlett.111.103201. ISSN   0031-9007. PMID   25166663.
  52. Wennerström, Håkan (2014). "Scattering and diffraction described using the momentum representation". Advances in Colloid and Interface Science. Elsevier BV. 205: 105–112. doi:10.1016/j.cis.2013.10.025. ISSN   0001-8686. PMID   24262675.
  53. Mermin, N.D. The Ithaca interpretation
  54. Leslie E. Ballentine (1998). Quantum Mechanics: A Modern Development. World Scientific. p. 342. ISBN   981-02-4105-4.
  55. "Like the old saying "A watched pot never boils", we have been led to the conclusion that a continuously observed system never changes its state! This conclusion is, of course false. The fallacy clearly results from the assertion that if an observation indicates no decay, then the state vector must be |y_u>. Each successive observation in the sequence would then "reduce" the state back to its initial value |y_u>, and in the limit of continuous observation there could be no change at all. Here we see that it is disproven by the simple empirical fact that [..] continuous observation does not prevent motion. It is sometimes claimed that the rival interpretations of quantum mechanics differ only in philosophy, and can not be experimentally distinguished. That claim is not always true. as this example proves". Ballentine, L. Quantum Mechanics, A Modern Development(p 342)
  56. Ballentine, L. E. (1991-05-01). "Comment on Quantum Zeno effect". Physical Review A. American Physical Society (APS). 43 (9): 5165–5167. Bibcode:1991PhRvA..43.5165B. doi:10.1103/physreva.43.5165. ISSN   1050-2947. PMID   9905643.: "The quantum Zeno effect is not a general characteristic of continuous measurements. In a recently reported experiment [Itano et al., Phys. Rev. A 41, 2295 (1990)], the inhibition of atomic excitation and deexcitation is not due to any collapse of the wave function, but instead is caused by a very strong perturbation due to the optical pulses and the coupling to the radiation field. The experiment should not be cited as providing empirical evidence in favor of the notion of wave-function collapse."