Measurement in quantum mechanics

Last updated

In quantum physics, a measurement is the testing or manipulation of a physical system to yield a numerical result. A fundamental feature of quantum theory is that the predictions it makes are probabilistic. The procedure for finding a probability involves combining a quantum state, which mathematically describes a quantum system, with a mathematical representation of the measurement to be performed on that system. The formula for this calculation is known as the Born rule. For example, a quantum particle like an electron can be described by a quantum state that associates to each point in space a complex number called a probability amplitude. Applying the Born rule to these amplitudes gives the probabilities that the electron will be found in one region or another when an experiment is performed to locate it. This is the best the theory can do; it cannot say for certain where the electron will be found. The same quantum state can also be used to make a prediction of how the electron will be moving, if an experiment is performed to measure its momentum instead of its position. The uncertainty principle implies that, whatever the quantum state, the range of predictions for the electron's position and the range of predictions for its momentum cannot both be narrow. Some quantum states imply a near-certain prediction of the result of a position measurement, but the result of a momentum measurement will be highly unpredictable, and vice versa. Furthermore, the fact that nature violates the statistical conditions known as Bell inequalities indicates that the unpredictability of quantum measurement results cannot be explained away as due to ignorance about "local hidden variables" within quantum systems.

Contents

Measuring a quantum system generally changes the quantum state that describes that system. This is a central feature of quantum mechanics, one that is both mathematically intricate and conceptually subtle. The mathematical tools for making predictions about what measurement outcomes may occur, and how quantum states can change, were developed during the 20th century and make use of linear algebra and functional analysis. Quantum physics has proven to be an empirical success and to have wide-ranging applicability. However, on a more philosophical level, debates continue about the meaning of the measurement concept.

Mathematical formalism

"Observables" as self-adjoint operators

In quantum mechanics, each physical system is associated with a Hilbert space, each element of which represents a possible state of the physical system. The approach codified by John von Neumann represents a measurement upon a physical system by a self-adjoint operator on that Hilbert space termed an "observable". [1] :17 These observables play the role of measurable quantities familiar from classical physics: position, momentum, energy, angular momentum and so on. The dimension of the Hilbert space may be infinite, as it is for the space of square-integrable functions on a line, which is used to define the quantum physics of a continuous degree of freedom. Alternatively, the Hilbert space may be finite-dimensional, as occurs for spin degrees of freedom. Many treatments of the theory focus on the finite-dimensional case, as the mathematics involved is somewhat less demanding. Indeed, introductory physics texts on quantum mechanics often gloss over mathematical technicalities that arise for continuous-valued observables and infinite-dimensional Hilbert spaces, such as the distinction between bounded and unbounded operators; questions of convergence (whether the limit of a sequence of Hilbert-space elements also belongs to the Hilbert space), exotic possibilities for sets of eigenvalues, like Cantor sets; and so forth. [2] :79 [3] These issues can be satisfactorily resolved using spectral theory; [2] :101 the present article will avoid them whenever possible.

Projective measurement

The eigenvectors of a von Neumann observable form an orthonormal basis for the Hilbert space, and each possible outcome of that measurement corresponds to one of the vectors comprising the basis. A density operator is a positive-semidefinite operator on the Hilbert space whose trace is equal to 1. [1] [2] For each measurement that can be defined, the probability distribution over the outcomes of that measurement can be computed from the density operator. The procedure for doing so is the Born rule, which states that

where is the density operator, and is the projection operator onto the basis vector corresponding to the measurement outcome . The average of the eigenvalues of a von Neumann observable, weighted by the Born rule probabilities, is the expectation value of that observable. For an observable , the expectation value given a quantum state is

A density operator that is a rank-1 projection is known as a pure quantum state, and all quantum states that are not pure are designated mixed. Pure states are also known as wavefunctions. Assigning a pure state to a quantum system implies certainty about the outcome of some measurement on that system (i.e., for some outcome ). Any mixed state can be written as a convex combination of pure states, though not in a unique way. [4] The state space of a quantum system is the set of all states, pure and mixed, that can be assigned to it.

The Born rule associates a probability with each unit vector in the Hilbert space, in such a way that these probabilities sum to 1 for any set of unit vectors comprising an orthonormal basis. Moreover, the probability associated with a unit vector is a function of the density operator and the unit vector, and not of additional information like a choice of basis for that vector to be embedded in. Gleason's theorem establishes the converse: all assignments of probabilities to unit vectors (or, equivalently, to the operators that project onto them) that satisfy these conditions take the form of applying the Born rule to some density operator. [5] [6] [7]

Generalized measurement (POVM)

In functional analysis and quantum measurement theory, a positive-operator-valued measure (POVM) is a measure whose values are positive semi-definite operators on a Hilbert space. POVMs are a generalisation of projection-valued measures (PVMs) and, correspondingly, quantum measurements described by POVMs are a generalisation of quantum measurement described by PVMs. In rough analogy, a POVM is to a PVM what a mixed state is to a pure state. Mixed states are needed to specify the state of a subsystem of a larger system (see Schrödinger–HJW theorem); analogously, POVMs are necessary to describe the effect on a subsystem of a projective measurement performed on a larger system. POVMs are the most general kind of measurement in quantum mechanics, and can also be used in quantum field theory. [8] They are extensively used in the field of quantum information.

In the simplest case, of a POVM with a finite number of elements acting on a finite-dimensional Hilbert space, a POVM is a set of positive semi-definite matrices on a Hilbert space that sum to the identity matrix, [9] :90

In quantum mechanics, the POVM element is associated with the measurement outcome , such that the probability of obtaining it when making a measurement on the quantum state is given by

,

where is the trace operator. When the quantum state being measured is a pure state this formula reduces to

.

State change due to measurement

A measurement upon a quantum system will generally bring about a change of the quantum state of that system. Writing a POVM does not provide the complete information necessary to describe this state-change process. [10] :134 To remedy this, further information is specified by decomposing each POVM element into a product:

The Kraus operators , named for Karl Kraus, provide a specification of the state-change process. [lower-alpha 1] They are not necessarily self-adjoint, but the products are. If upon performing the measurement the outcome is obtained, then the initial state is updated to

An important special case is the Lüders rule, named for Gerhart Lüders. [16] [17] If the POVM is itself a PVM, then the Kraus operators can be taken to be the projectors onto the eigenspaces of the von Neumann observable:

If the initial state is pure, and the projectors have rank 1, they can be written as projectors onto the vectors and , respectively. The formula simplifies thus to

Lüders rule has historically been known as the "reduction of the wave packet" or the "collapse of the wavefunction". [17] [18] [19] The pure state implies a probability-one prediction for any von Neumann observable that has as an eigenvector. Introductory texts on quantum theory often express this by saying that if a quantum measurement is repeated in quick succession, the same outcome will occur both times. This is an oversimplification, since the physical implementation of a quantum measurement may involve a process like the absorption of a photon; after the measurement, the photon does not exist to be measured again. [9] :91

We can define a linear, trace-preserving, completely positive map, by summing over all the possible post-measurement states of a POVM without the normalisation:

It is an example of a quantum channel, [10] :150 and can be interpreted as expressing how a quantum state changes if a measurement is performed but the result of that measurement is lost. [10] :159

Examples

Bloch sphere representation of states (in blue) and optimal POVM (in red) for unambiguous quantum state discrimination on the states
|
ps
> 
=
|
0
> 
{\displaystyle |\psi \rangle =|0\rangle }
and
|
ph
> 
=
(
|
0
> 
+
|
1
> 
)
/
2
{\displaystyle |\varphi \rangle =(|0\rangle +|1\rangle )/{\sqrt {2}}}
. Note that on the Bloch sphere orthogonal states are antiparallel. Bloch sphere representation of optimal POVM and states for unambiguous quantum state discrimination.svg
Bloch sphere representation of states (in blue) and optimal POVM (in red) for unambiguous quantum state discrimination on the states and . Note that on the Bloch sphere orthogonal states are antiparallel.

The prototypical example of a finite-dimensional Hilbert space is a qubit, a quantum system whose Hilbert space is 2-dimensional. A pure state for a qubit can be written as a linear combination of two orthogonal basis states and with complex coefficients:

A measurement in the basis will yield outcome with probability and outcome with probability , so by normalization,

An arbitrary state for a qubit can be written as a linear combination of the Pauli matrices, which provide a basis for self-adjoint matrices: [10] :126

where the real numbers are the coordinates of a point within the unit ball and

POVM elements can be represented likewise, though the trace of a POVM element is not fixed to equal 1. The Pauli matrices are traceless and orthogonal to one another with respect to the Hilbert–Schmidt inner product, and so the coordinates of the state are the expectation values of the three von Neumann measurements defined by the Pauli matrices. [10] :126 If such a measurement is applied to a qubit, then by the Lüders rule, the state will update to the eigenvector of that Pauli matrix corresponding to the measurement outcome. The eigenvectors of are the basis states and , and a measurement of is often called a measurement in the "computational basis." [10] :76 After a measurement in the computational basis, the outcome of a or measurement is maximally uncertain.

A pair of qubits together form a system whose Hilbert space is 4-dimensional. One significant von Neumann measurement on this system is that defined by the Bell basis, [21] :36 a set of four maximally entangled states:

Probability density
P
n
(
x
)
{\displaystyle P_{n}(x)}
for the outcome of a position measurement given the energy eigenstate
|
n
> 
{\displaystyle |n\rangle }
of a 1D harmonic oscillator. Aufenthaltswahrscheinlichkeit harmonischer Oszillator.png
Probability density for the outcome of a position measurement given the energy eigenstate of a 1D harmonic oscillator.

A common and useful example of quantum mechanics applied to a continuous degree of freedom is the quantum harmonic oscillator. [22] :24 This system is defined by the Hamiltonian

where , the momentum operator and the position operator are self-adjoint operators on the Hilbert space of square-integrable functions on the real line. The energy eigenstates solve the time-independent Schrödinger equation:

These eigenvalues can be shown to be given by

and these values give the possible numerical outcomes of an energy measurement upon the oscillator. The set of possible outcomes of a position measurement on a harmonic oscillator is continuous, and so predictions are stated in terms of a probability density function that gives the probability of the measurement outcome lying in the infinitesimal interval from to .

History of the measurement concept

The "old quantum theory"

The old quantum theory is a collection of results from the years 1900–1925 [23] which predate modern quantum mechanics. The theory was never complete or self-consistent, but was rather a set of heuristic corrections to classical mechanics. [24] The theory is now understood as a semi-classical approximation [25] to modern quantum mechanics. [26] Notable results from this period include Planck's calculation of the blackbody radiation spectrum, Einstein's explanation of the photoelectric effect, Einstein and Debye's work on the specific heat of solids, Bohr and van Leeuwen's proof that classical physics cannot account for diamagnetism, Bohr's model of the hydrogen atom and Arnold Sommerfeld's extension of the Bohr model to include relativistic effects.

Stern-Gerlach experiment: Silver atoms travelling through an inhomogeneous magnetic field, and being deflected up or down depending on their spin; (1) furnace, (2) beam of silver atoms, (3) inhomogeneous magnetic field, (4) classically expected result, (5) observed result Stern-Gerlach experiment svg.svg
Stern–Gerlach experiment: Silver atoms travelling through an inhomogeneous magnetic field, and being deflected up or down depending on their spin; (1) furnace, (2) beam of silver atoms, (3) inhomogeneous magnetic field, (4) classically expected result, (5) observed result

The Stern–Gerlach experiment, proposed in 1921 and implemented in 1922, [27] [28] [29] became a prototypical example of a quantum measurement having a discrete set of possible outcomes. In the original experiment, silver atoms were sent through a spatially varying magnetic field, which deflected them before they struck a detector screen, such as a glass slide. Particles with non-zero magnetic moment are deflected, due to the magnetic field gradient, from a straight path. The screen reveals discrete points of accumulation, rather than a continuous distribution, owing to the particles' quantized spin. [30] [31] [32]

Transition to the “new” quantum theory

A 1925 paper by Heisenberg, known in English as "Quantum theoretical re-interpretation of kinematic and mechanical relations", marked a pivotal moment in the maturation of quantum physics. [33] Heisenberg sought to develop a theory of atomic phenomena that relied only on "observable" quantities. At the time, and in contrast with the later standard presentation of quantum mechanics, Heisenberg did not regard the position of an electron bound within an atom as "observable". Instead, his principal quantities of interest were the frequencies of light emitted or absorbed by atoms. [33]

The uncertainty principle dates to this period. It is frequently attributed to Heisenberg, who introduced the concept in analyzing a thought experiment where one attempts to measure an electron's position and momentum simultaneously. However, Heisenberg did not give precise mathematical definitions of what the "uncertainty" in these measurements meant. The precise mathematical statement of the position-momentum uncertainty principle is due to Kennard, Pauli, and Weyl, and its generalization to arbitrary pairs of noncommuting observables is due to Robertson and Schrödinger. [34] [35]

Writing and for the self-adjoint operators representing position and momentum respectively, a standard deviation of position can be defined as

and likewise for the momentum:

The Kennard–Pauli–Weyl uncertainty relation is

This inequality means that no preparation of a quantum particle can imply simultaneously precise predictions for a measurement of position and for a measurement of momentum. [36] The Robertson inequality generalizes this to the case of an arbitrary pair of self-adjoint operators and . The commutator of these two operators is

and this provides the lower bound on the product of standard deviations:

Substituting in the canonical commutation relation , an expression first postulated by Max Born in 1925, [37] recovers the Kennard–Pauli–Weyl statement of the uncertainty principle.

From uncertainty to no-hidden-variables

The existence of the uncertainty principle naturally raises the question of whether quantum mechanics can be understood as an approximation to a more exact theory. Do there exist "hidden variables", more fundamental than the quantities addressed in quantum theory itself, knowledge of which would allow more exact predictions than quantum theory can provide? A collection of results, most significantly Bell's theorem, have demonstrated that broad classes of such hidden-variable theories are in fact incompatible with quantum physics.

Bell published the theorem now known by his name in 1964, investigating more deeply a thought experiment originally proposed in 1935 by Einstein, Podolsky and Rosen. [38] [39] According to Bell's theorem, if nature actually operates in accord with any theory of local hidden variables, then the results of a Bell test will be constrained in a particular, quantifiable way. If a Bell test is performed in a laboratory and the results are not thus constrained, then they are inconsistent with the hypothesis that local hidden variables exist. Such results would support the position that there is no way to explain the phenomena of quantum mechanics in terms of a more fundamental description of nature that is more in line with the rules of classical physics. Many types of Bell test have been performed in physics laboratories, often with the goal of ameliorating problems of experimental design or set-up that could in principle affect the validity of the findings of earlier Bell tests. This is known as "closing loopholes in Bell tests". To date, Bell tests have found that the hypothesis of local hidden variables is inconsistent with the way that physical systems behave. [40] [41]

Quantum systems as measuring devices

The Robertson–Schrödinger uncertainty principle establishes that when two observables do not commute, there is a tradeoff in predictability between them. The Wigner–Araki–Yanase theorem demonstrates another consequence of non-commutativity: the presence of a conservation law limits the accuracy with which observables that fail to commute with the conserved quantity can be measured. [42] Further investigation in this line led to the formulation of the Wigner–Yanase skew information. [43]

Historically, experiments in quantum physics have often been described in semiclassical terms. For example, the spin of an atom in a Stern–Gerlach experiment might be treated as a quantum degree of freedom, while the atom is regarded as moving through a magnetic field described by the classical theory of Maxwell's equations. [2] :24 But the devices used to build the experimental apparatus are themselves physical systems, and so quantum mechanics should be applicable to them as well. Beginning in the 1950s, Rosenfeld, von Weizsäcker and others tried to develop consistency conditions that expressed when a quantum-mechanical system could be treated as a measuring apparatus. [44] One proposal for a criterion regarding when a system used as part of a measuring device can be modeled semiclassically relies on the Wigner function, a quasiprobability distribution that can be treated as a probability distribution on phase space in those cases where it is everywhere non-negative. [2] :375

Decoherence

A quantum state for an imperfectly isolated system will generally evolve to be entangled with the quantum state for the environment. Consequently, even if the system's initial state is pure, the state at a later time, found by taking the partial trace of the joint system-environment state, will be mixed. This phenomenon of entanglement produced by system-environment interactions tends to obscure the more exotic features of quantum mechanics that the system could in principle manifest. Quantum decoherence, as this effect is known, was first studied in detail during the 1970s. [45] (Earlier investigations into how classical physics might be obtained as a limit of quantum mechanics had explored the subject of imperfectly isolated systems, but the role of entanglement was not fully appreciated. [44] ) A significant portion of the effort involved in quantum computing is to avoid the deleterious effects of decoherence. [46] [21] :239

To illustrate, let denote the initial state of the system, the initial state of the environment and the Hamiltonian specifying the system-environment interaction. The density operator can be diagonalized and written as a linear combination of the projectors onto its eigenvectors:

Expressing time evolution for a duration by the unitary operator , the state for the system after this evolution is

which evaluates to

The quantities surrounding can be identified as Kraus operators, and so this defines a quantum channel. [45]

Specifying a form of interaction between system and environment can establish a set of "pointer states," states for the system that are (approximately) stable, apart from overall phase factors, with respect to environmental fluctuations. A set of pointer states defines a preferred orthonormal basis for the system's Hilbert space. [2] :423

Quantum information and computation

Quantum information science studies how information science and its application as technology depend on quantum-mechanical phenomena. Understanding measurement in quantum physics is important for this field in many ways, some of which are briefly surveyed here.

Measurement, entropy, and distinguishability

The von Neumann entropy is a measure of the statistical uncertainty represented by a quantum state. For a density matrix , the von Neumann entropy is

writing in terms of its basis of eigenvectors,

the von Neumann entropy is

This is the Shannon entropy of the set of eigenvalues interpreted as a probability distribution, and so the von Neumann entropy is the Shannon entropy of the random variable defined by measuring in the eigenbasis of . Consequently, the von Neumann entropy vanishes when is pure. [10] :320 The von Neumann entropy of can equivalently be characterized as the minimum Shannon entropy for a measurement given the quantum state , with the minimization over all POVMs with rank-1 elements. [10] :323

Many other quantities used in quantum information theory also find motivation and justification in terms of measurements. For example, the trace distance between quantum states is equal to the largest difference in probability that those two quantum states can imply for a measurement outcome: [10] :254

Similarly, the fidelity of two quantum states, defined by

expresses the probability that one state will pass a test for identifying a successful preparation of the other. The trace distance provides bounds on the fidelity via the Fuchs–van de Graaf inequalities: [10] :274

Quantum circuits

Circuit representation of measurement. The single line on the left-hand side stands for a qubit, while the two lines on the right-hand side represent a classical bit. Qcircuit measure-arrow.svg
Circuit representation of measurement. The single line on the left-hand side stands for a qubit, while the two lines on the right-hand side represent a classical bit.

Quantum circuits are a model for quantum computation in which a computation is a sequence of quantum gates followed by measurements. [21] :93 The gates are reversible transformations on a quantum mechanical analog of an n-bit register. This analogous structure is referred to as an n-qubit register. Measurements, drawn on a circuit diagram as stylized pointer dials, indicate where and how a result is obtained from the quantum computer after the steps of the computation are executed. Without loss of generality, one can work with the standard circuit model, in which the set of gates are single-qubit unitary transformations and controlled NOT gates on pairs of qubits, and all measurements are in the computational basis. [21] :93 [47]

Measurement-based quantum computation

Measurement-based quantum computation (MBQC) is a model of quantum computing in which the answer to a question is, informally speaking, created in the act of measuring the physical system that serves as the computer. [21] :317 [48] [49]

Quantum tomography

Quantum state tomography is a process by which, given a set of data representing the results of quantum measurements, a quantum state consistent with those measurement results is computed. [50] It is named by analogy with tomography, the reconstruction of three-dimensional images from slices taken through them, as in a CT scan. Tomography of quantum states can be extended to tomography of quantum channels [50] and even of measurements. [51]

Quantum metrology

Quantum metrology is the use of quantum physics to aid the measurement of quantities that, generally, had meaning in classical physics, such as exploiting quantum effects to increase the precision with which a length can be measured. [52] A celebrated example is the introduction of squeezed light into the LIGO experiment, which increased its sensitivity to gravitational waves. [53] [54]

Laboratory implementations

The range of physical procedures to which the mathematics of quantum measurement can be applied is very broad. [55] In the early years of the subject, laboratory procedures involved the recording of spectral lines, the darkening of photographic film, the observation of scintillations, finding tracks in cloud chambers, and hearing clicks from Geiger counters. [lower-alpha 2] Language from this era persists, such as the description of measurement outcomes in the abstract as "detector clicks". [57]

The double-slit experiment is a prototypical illustration of quantum interference, typically described using electrons or photons. The first interference experiment to be carried out in a regime where both wave-like and particle-like aspects of photon behavior are significant was G. I. Taylor's test in 1909. Taylor used screens of smoked glass to attenuate the light passing through his apparatus, to the extent that, in modern language, only one photon would be illuminating the interferometer slits at a time. He recorded the interference patterns on photographic plates; for the dimmest light, the exposure time required was roughly three months. [58] [59] In 1974, the Italian physicists Pier Giorgio Merli, Gian Franco Missiroli, and Giulio Pozzi implemented the double-slit experiment using single electrons and a television tube. [60] A quarter-century later, a team at the University of Vienna performed an interference experiment with buckyballs, in which the buckyballs that passed through the interferometer were ionized by a laser, and the ions then induced the emission of electrons, emissions which were in turn amplified and detected by an electron multiplier. [61]

Modern quantum optics experiments can employ single-photon detectors. For example, in the "BIG Bell test" of 2018, several of the laboratory setups used single-photon avalanche diodes. Another laboratory setup used superconducting qubits. [40] The standard method for performing measurements upon superconducting qubits is to couple a qubit with a resonator in such a way that the characteristic frequency of the resonator shifts according to the state for the qubit, and detecting this shift by observing how the resonator reacts to a probe signal. [62]

Interpretations of quantum mechanics

Niels Bohr and Albert Einstein, pictured here at Paul Ehrenfest's home in Leiden (December 1925), had a long-running collegial dispute about what quantum mechanics implied for the nature of reality. Niels Bohr Albert Einstein4 by Ehrenfest cr.jpg
Niels Bohr and Albert Einstein, pictured here at Paul Ehrenfest's home in Leiden (December 1925), had a long-running collegial dispute about what quantum mechanics implied for the nature of reality.

Despite the consensus among scientists that quantum physics is in practice a successful theory, disagreements persist on a more philosophical level. Many debates in the area known as quantum foundations concern the role of measurement in quantum mechanics. Recurring questions include which interpretation of probability theory is best suited for the probabilities calculated from the Born rule; and whether the apparent randomness of quantum measurement outcomes is fundamental, or a consequence of a deeper deterministic process. [63] [64] [65] Worldviews that present answers to questions like these are known as "interpretations" of quantum mechanics; as the physicist N. David Mermin once quipped, "New interpretations appear every year. None ever disappear." [66]

A central concern within quantum foundations is the "quantum measurement problem," though how this problem is delimited, and whether it should be counted as one question or multiple separate issues, are contested topics. [56] [67] Of primary interest is the seeming disparity between apparently distinct types of time evolution. Von Neumann declared that quantum mechanics contains "two fundamentally different types" of quantum-state change. [68] :§V.1 First, there are those changes involving a measurement process, and second, there is unitary time evolution in the absence of measurement. The former is stochastic and discontinuous, writes von Neumann, and the latter deterministic and continuous. This dichotomy has set the tone for much later debate. [69] [70] Some interpretations of quantum mechanics find the reliance upon two different types of time evolution distasteful and regard the ambiguity of when to invoke one or the other as a deficiency of the way quantum theory was historically presented. [71] To bolster these interpretations, their proponents have worked to derive ways of regarding "measurement" as a secondary concept and deducing the seemingly stochastic effect of measurement processes as approximations to more fundamental deterministic dynamics. However, consensus has not been achieved among proponents of the correct way to implement this program, and in particular how to justify the use of the Born rule to calculate probabilities. [72] [73] Other interpretations regard quantum states as statistical information about quantum systems, thus asserting that abrupt and discontinuous changes of quantum states are not problematic, simply reflecting updates of the available information. [55] [74] Of this line of thought, Bell asked, "Whose information? Information about what?" [71] Answers to these questions vary among proponents of the informationally-oriented interpretations. [64] [74]

See also

Notes

  1. Hellwig and Kraus [11] [12] originally introduced operators with two indices, , such that . The extra index does not affect the computation of the measurement outcome probability, but it does play a role in the state-update rule, with the post-measurement state being now proportional to . This can be regarded as representing as a coarse-graining together of multiple outcomes of a more fine-grained POVM. [13] [14] [15] Kraus operators with two indices also occur in generalized models of system-environment interaction. [9] :364
  2. The glass plates used in the Stern–Gerlach experiment did not darken properly until Stern breathed on them, accidentally exposing them to sulfur from his cheap cigars. [31] [56]

Related Research Articles

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 teleportation</span> Physical phenomenon

Quantum teleportation is a technique for transferring quantum information from a sender at one location to a receiver some distance away. While teleportation is commonly portrayed in science fiction as a means to transfer physical objects from one location to the next, quantum teleportation only transfers quantum information. The sender does not have to know the particular quantum state being transferred. Moreover, the location of the recipient can be unknown, but to complete the quantum teleportation, classical information needs to be sent from sender to receiver. Because classical information needs to be sent, quantum teleportation cannot occur faster than the speed of light.

<span class="mw-page-title-main">Quantum entanglement</span> Correlation between quantum systems

Quantum entanglement is the phenomenon of a group of particles being generated, interacting, or sharing spatial proximity in such a way that the quantum state of each particle of the group cannot be described independently of the state of the others, including when the particles are separated by a large distance. The topic of quantum entanglement is at the heart of the disparity between classical and quantum physics: entanglement is a primary feature of quantum mechanics not present in classical mechanics.

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

The uncertainty principle, also known as Heisenberg's indeterminacy principle, is a fundamental concept in quantum mechanics. It states that there is a limit to the precision with which certain pairs of physical properties, such as position and momentum, can be simultaneously known. In other words, the more accurately one property is measured, the less accurately the other property can be known.

In quantum mechanics, a density matrix is a matrix that describes the quantum state of a physical system. It allows for the calculation of the probabilities of the outcomes of any measurement performed upon this system, using the Born rule. It is a generalization of the more usual state vectors or wavefunctions: while those can only represent pure states, density matrices can also represent mixed states. Mixed states arise in quantum mechanics in two different situations:

  1. when the preparation of the system is not fully known, and thus one must deal with a statistical ensemble of possible preparations, and
  2. when one wants to describe a physical system that is entangled with another, without describing their combined state; this case is typical for a system interacting with some environment.
<span class="mw-page-title-main">Quantum decoherence</span> Loss of quantum coherence

Quantum decoherence is the loss of quantum coherence, the process in which a system's behaviour changes from that which can be explained by quantum mechanics to that which can be explained by classical mechanics. Beginning out of attempts to extend the understanding of quantum mechanics, the theory has developed in several directions and experimental studies have confirmed some of the key issues. Quantum computing relies on quantum coherence and is the primary practical applications of the concept.

<span class="mw-page-title-main">Quantum tomography</span> Reconstruction of quantum states based on measurements

Quantum tomography or quantum state tomography is the process by which a quantum state is reconstructed using measurements on an ensemble of identical quantum states. The source of these states may be any device or system which prepares quantum states either consistently into quantum pure states or otherwise into general mixed states. To be able to uniquely identify the state, the measurements must be tomographically complete. That is, the measured operators must form an operator basis on the Hilbert space of the system, providing all the information about the state. Such a set of observations is sometimes called a quorum. The term tomography was first used in the quantum physics literature in a 1993 paper introducing experimental optical homodyne tomography.

In physics, the von Neumann entropy, named after John von Neumann, is an extension of the concept of Gibbs entropy from classical statistical mechanics to quantum statistical mechanics. For a quantum-mechanical system described by a density matrix ρ, the von Neumann entropy is

In functional analysis and quantum information science, a positive operator-valued measure (POVM) is a measure whose values are positive semi-definite operators on a Hilbert space. POVMs are a generalization of projection-valued measures (PVM) and, correspondingly, quantum measurements described by POVMs are a generalization of quantum measurement described by PVMs.

The Born rule is a postulate of quantum mechanics that gives the probability that a measurement of a quantum system will yield a given result. In its simplest form, it states that the probability density of finding a system in a given state, when measured, is proportional to the square of the amplitude of the system's wavefunction at that state. It was formulated and published by German physicist Max Born in July, 1926.

In quantum mechanics, notably in quantum information theory, fidelity is a measure of the "closeness" of two quantum states. It expresses the probability that one state will pass a test to identify as the other. The fidelity is not a metric on the space of density matrices, but it can be used to define the Bures metric on this space.

In mathematical physics, Gleason's theorem shows that the rule one uses to calculate probabilities in quantum physics, the Born rule, can be derived from the usual mathematical representation of measurements in quantum physics together with the assumption of non-contextuality. Andrew M. Gleason first proved the theorem in 1957, answering a question posed by George W. Mackey, an accomplishment that was historically significant for the role it played in showing that wide classes of hidden-variable theories are inconsistent with quantum physics. Multiple variations have been proven in the years since. Gleason's theorem is of particular importance for the field of quantum logic and its attempt to find a minimal set of mathematical axioms for quantum theory.

In quantum mechanics, the expectation value is the probabilistic expected value of the result (measurement) of an experiment. It can be thought of as an average of all the possible outcomes of a measurement as weighted by their likelihood, and as such it is not the most probable value of a measurement; indeed the expectation value may have zero probability of occurring. It is a fundamental concept in all areas of quantum physics.

<span class="mw-page-title-main">SIC-POVM</span> Type of measurement in quantum mechanics

In the context of quantum mechanics and quantum information theory, symmetric, informationally complete, positive operator-valued measures (SIC-POVMs) are a particular type of generalized measurement (POVM). SIC-POVMs are particularly notable thanks to their defining features of (1) being informationally complete; (2)having the minimal number of outcomes compatible with informational completeness, and (3) being highly symmetric. In this context, informational completeness is the property of a POVM of allowing to fully reconstruct input states from measurement data.

In quantum mechanics, and especially quantum information theory, the purity of a normalized quantum state is a scalar defined as

In quantum mechanics, and especially quantum information and the study of open quantum systems, the trace distanceT is a metric on the space of density matrices and gives a measure of the distinguishability between two states. It is the quantum generalization of the Kolmogorov distance for classical probability distributions.

The Koopman–von Neumann (KvN) theory is a description of classical mechanics as an operatorial theory similar to quantum mechanics, based on a Hilbert space of complex, square-integrable wavefunctions. As its name suggests, the KvN theory is loosely related to work by Bernard Koopman and John von Neumann in 1931 and 1932, respectively. As explained in this entry, however, the historical origins of the theory and its name are complicated.

Generalized relative entropy is a measure of dissimilarity between two quantum states. It is a "one-shot" analogue of quantum relative entropy and shares many properties of the latter quantity.

In quantum information theory and quantum optics, the Schrödinger–HJW theorem is a result about the realization of a mixed state of a quantum system as an ensemble of pure quantum states and the relation between the corresponding purifications of the density operators. The theorem is named after physicists and mathematicians Erwin Schrödinger, Lane P. Hughston, Richard Jozsa and William Wootters. The result was also found independently by Nicolas Gisin, and by Nicolas Hadjisavvas building upon work by Ed Jaynes, while a significant part of it was likewise independently discovered by N. David Mermin. Thanks to its complicated history, it is also known by various other names such as the GHJW theorem, the HJW theorem, and the purification theorem.

This glossary of quantum computing is a list of definitions of terms and concepts used in quantum computing, its sub-disciplines, and related fields.

References

  1. 1 2 Holevo, Alexander S. (2001). Statistical Structure of Quantum Theory. Lecture Notes in Physics. Springer. ISBN   3-540-42082-7. OCLC   318268606.
  2. 1 2 3 4 5 6 Peres, Asher (1995). Quantum Theory: Concepts and Methods . Kluwer Academic Publishers. ISBN   0-7923-2549-4.
  3. Tao, Terry (12 August 2014). "Avila, Bhargava, Hairer, Mirzakhani". What's New. Retrieved 9 February 2020.
  4. Kirkpatrick, K. A. (February 2006). "The Schrödinger-HJW Theorem". Foundations of Physics Letters . 19 (1): 95–102. arXiv: quant-ph/0305068 . Bibcode:2006FoPhL..19...95K. doi:10.1007/s10702-006-1852-1. ISSN   0894-9875. S2CID   15995449.
  5. Gleason, Andrew M. (1957). "Measures on the closed subspaces of a Hilbert space". Indiana University Mathematics Journal . 6 (4): 885–893. doi: 10.1512/iumj.1957.6.56050 . MR   0096113.
  6. Busch, Paul (2003). "Quantum States and Generalized Observables: A Simple Proof of Gleason's Theorem". Physical Review Letters . 91 (12): 120403. arXiv: quant-ph/9909073 . Bibcode:2003PhRvL..91l0403B. doi:10.1103/PhysRevLett.91.120403. PMID   14525351. S2CID   2168715.
  7. Caves, Carlton M.; Fuchs, Christopher A.; Manne, Kiran K.; Renes, Joseph M. (2004). "Gleason-Type Derivations of the Quantum Probability Rule for Generalized Measurements". Foundations of Physics . 34 (2): 193–209. arXiv: quant-ph/0306179 . Bibcode:2004FoPh...34..193C. doi:10.1023/B:FOOP.0000019581.00318.a5. S2CID   18132256.
  8. Peres, Asher; Terno, Daniel R. (2004). "Quantum information and relativity theory". Reviews of Modern Physics . 76 (1): 93–123. arXiv: quant-ph/0212023 . Bibcode:2004RvMP...76...93P. doi:10.1103/RevModPhys.76.93. S2CID   7481797.
  9. 1 2 3 Nielsen, Michael A.; Chuang, Isaac L. (2000). Quantum Computation and Quantum Information (1st ed.). Cambridge: Cambridge University Press. ISBN   978-0-521-63503-5. OCLC   634735192.
  10. 1 2 3 4 5 6 7 8 9 10 Wilde, Mark M. (2017). Quantum Information Theory (2nd ed.). Cambridge University Press. arXiv: 1106.1445 . doi:10.1017/9781316809976.001. ISBN   9781107176164. OCLC   973404322. S2CID   2515538.
  11. Hellwig, K. -E.; Kraus, K. (September 1969). "Pure operations and measurements". Communications in Mathematical Physics . 11 (3): 214–220. doi:10.1007/BF01645807. ISSN   0010-3616. S2CID   123659396.
  12. Kraus, Karl (1983). States, effects, and operations: fundamental notions of quantum theory. Lectures in mathematical physics at the University of Texas at Austin. Vol. 190. Springer-Verlag. ISBN   978-3-5401-2732-1. OCLC   925001331.
  13. Barnum, Howard; Nielsen, M. A.; Schumacher, Benjamin (1 June 1998). "Information transmission through a noisy quantum channel". Physical Review A . 57 (6): 4153–4175. arXiv: quant-ph/9702049 . Bibcode:1998PhRvA..57.4153B. doi:10.1103/PhysRevA.57.4153. ISSN   1050-2947. S2CID   13717391.
  14. Fuchs, Christopher A.; Jacobs, Kurt (16 May 2001). "Information-tradeoff relations for finite-strength quantum measurements". Physical Review A . 63 (6): 062305. arXiv: quant-ph/0009101 . Bibcode:2001PhRvA..63f2305F. doi:10.1103/PhysRevA.63.062305. ISSN   1050-2947. S2CID   119476175.
  15. Poulin, David (7 February 2005). "Macroscopic observables". Physical Review A . 71 (2): 022102. arXiv: quant-ph/0403212 . Bibcode:2005PhRvA..71b2102P. doi:10.1103/PhysRevA.71.022102. ISSN   1050-2947. S2CID   119364450.
  16. Lüders, Gerhart (1950). "Über die Zustandsänderung durch den Messprozeß". Annalen der Physik . 443 (5–8): 322. Bibcode:1950AnP...443..322L. doi:10.1002/andp.19504430510. Translated by K. A. Kirkpatrick as Lüders, Gerhart (3 April 2006). "Concerning the state-change due to the measurement process". Annalen der Physik . 15 (9): 663–670. arXiv: quant-ph/0403007 . Bibcode:2006AnP...518..663L. doi:10.1002/andp.200610207. S2CID   119103479.
  17. 1 2 Busch, Paul; Lahti, Pekka (2009), Greenberger, Daniel; Hentschel, Klaus; Weinert, Friedel (eds.), "Lüders Rule", Compendium of Quantum Physics, Springer Berlin Heidelberg, pp. 356–358, doi:10.1007/978-3-540-70626-7_110, ISBN   978-3-540-70622-9
  18. Jammer, Max (1979). "A Consideration of the Philosophical Implications of the New Physics". In Radnitzky, Gerard; Andersson, Gunnar (eds.). The Structure and Development of Science. Vol. 59. Dordrecht: Springer Netherlands. pp. 41–61. doi:10.1007/978-94-009-9459-1_3. ISBN   978-90-277-0995-0 . Retrieved 26 March 2024.
  19. Pessoa, Osvaldo (2022). "The Measurement Problem". In Freire, Olival (ed.). The Oxford Handbook of the History of Quantum Interpretations. Oxford University Press. pp. 281–302. doi:10.1093/oxfordhb/9780198844495.013.0012. ISBN   978-0-191-88008-7.
  20. Peres, Asher; Terno, Daniel R. (1998). "Optimal distinction between non-orthogonal quantum states". Journal of Physics A: Mathematical and General . 31 (34): 7105–7111. arXiv: quant-ph/9804031 . Bibcode:1998JPhA...31.7105P. doi:10.1088/0305-4470/31/34/013. ISSN   0305-4470. S2CID   18961213.
  21. 1 2 3 4 5 Rieffel, Eleanor G.; Polak, Wolfgang H. (4 March 2011). Quantum Computing: A Gentle Introduction. MIT Press. ISBN   978-0-262-01506-6.
  22. Weinberg, Steven (2015). Lectures on quantum mechanics (Second ed.). Cambridge, United Kingdom: Cambridge University Press. ISBN   978-1-107-11166-0. OCLC   910664598.
  23. Pais, Abraham (2005). Subtle is the Lord: The Science and the Life of Albert Einstein (illustrated ed.). Oxford University Press. p. 28. ISBN   978-0-19-280672-7.
  24. ter Haar, D. (1967). The Old Quantum Theory . Pergamon Press. pp.  206. ISBN   978-0-08-012101-7.
  25. "Semi-classical approximation". Encyclopedia of Mathematics. Retrieved 1 February 2020.
  26. Sakurai, J. J.; Napolitano, J. (2014). "Quantum Dynamics". Modern Quantum Mechanics. Pearson. ISBN   978-1-292-02410-3. OCLC   929609283.
  27. Gerlach, W.; Stern, O. (1922). "Der experimentelle Nachweis der Richtungsquantelung im Magnetfeld". Zeitschrift für Physik . 9 (1): 349–352. Bibcode:1922ZPhy....9..349G. doi:10.1007/BF01326983. S2CID   186228677.
  28. Gerlach, W.; Stern, O. (1922). "Das magnetische Moment des Silberatoms". Zeitschrift für Physik . 9 (1): 353–355. Bibcode:1922ZPhy....9..353G. doi:10.1007/BF01326984. S2CID   126109346.
  29. Gerlach, W.; Stern, O. (1922). "Der experimentelle Nachweis des magnetischen Moments des Silberatoms". Zeitschrift für Physik . 8 (1): 110–111. Bibcode:1922ZPhy....8..110G. doi:10.1007/BF01329580. S2CID   122648402.
  30. Franklin, Allan; Perovic, Slobodan. "Experiment in Physics, Appendix 5". In Edward N. Zalta (ed.). The Stanford Encyclopedia of Philosophy (Winter 2016 ed.). Retrieved 14 August 2018.
  31. 1 2 Friedrich, B.; Herschbach, D. (2003). "Stern and Gerlach: How a Bad Cigar Helped Reorient Atomic Physics". Physics Today . 56 (12): 53. Bibcode:2003PhT....56l..53F. doi: 10.1063/1.1650229 . S2CID   17572089.
  32. Zhu, Guangtian; Singh, Chandralekha (May 2011). "Improving students' understanding of quantum mechanics via the Stern–Gerlach experiment". American Journal of Physics . 79 (5): 499–507. arXiv: 1602.06367 . Bibcode:2011AmJPh..79..499Z. doi:10.1119/1.3546093. ISSN   0002-9505. S2CID   55077698.
  33. 1 2 van der Waerden, B. L. (1968). "Introduction, Part II". Sources of Quantum Mechanics. Dover. ISBN   0-486-61881-1.
  34. Busch, Paul; Lahti, Pekka; Werner, Reinhard F. (17 October 2013). "Proof of Heisenberg's Error-Disturbance Relation". Physical Review Letters. 111 (16): 160405. arXiv: 1306.1565 . Bibcode:2013PhRvL.111p0405B. doi:10.1103/PhysRevLett.111.160405. ISSN   0031-9007. PMID   24182239. S2CID   24507489.
  35. Appleby, David Marcus (6 May 2016). "Quantum Errors and Disturbances: Response to Busch, Lahti and Werner". Entropy. 18 (5): 174. arXiv: 1602.09002 . Bibcode:2016Entrp..18..174A. doi: 10.3390/e18050174 .
  36. Landau, L.D.; Lifschitz, E.M. (1977). Quantum Mechanics: Non-Relativistic Theory. Vol. 3 (3rd ed.). Pergamon Press. ISBN   978-0-08-020940-1. OCLC   2284121.
  37. Born, M.; Jordan, P. (1925). "Zur Quantenmechanik". Zeitschrift für Physik . 34 (1): 858–888. Bibcode:1925ZPhy...34..858B. doi:10.1007/BF01328531. S2CID   186114542.
  38. Bell, J. S. (1964). "On the Einstein Podolsky Rosen Paradox" (PDF). Physics Physique Физика . 1 (3): 195–200. doi: 10.1103/PhysicsPhysiqueFizika.1.195 .
  39. Einstein, A; Podolsky, B; Rosen, N (15 May 1935). "Can Quantum-Mechanical Description of Physical Reality be Considered Complete?". Physical Review . 47 (10): 777–780. Bibcode:1935PhRv...47..777E. doi: 10.1103/PhysRev.47.777 .
  40. 1 2 The BIG Bell Test Collaboration (9 May 2018). "Challenging local realism with human choices". Nature . 557 (7704): 212–216. arXiv: 1805.04431 . Bibcode:2018Natur.557..212B. doi:10.1038/s41586-018-0085-3. PMID   29743691. S2CID   13665914.
  41. Wolchover, Natalie (7 February 2017). "Experiment Reaffirms Quantum Weirdness". Quanta Magazine . Retrieved 8 February 2020.
  42. See, for example:
  43. Luo, Shenlong (2003). "Wigner–Yanase Skew Information and Uncertainty Relations". Physical Review Letters . 91 (18): 180403. Bibcode:2003PhRvL..91r0403L. doi:10.1103/PhysRevLett.91.180403. PMID   14611271.
  44. 1 2 Camilleri, K.; Schlosshauer, M. (2015). "Niels Bohr as Philosopher of Experiment: Does Decoherence Theory Challenge Bohr's Doctrine of Classical Concepts?". Studies in History and Philosophy of Modern Physics. 49: 73–83. arXiv: 1502.06547 . Bibcode:2015SHPMP..49...73C. doi:10.1016/j.shpsb.2015.01.005. S2CID   27697360.
  45. 1 2 Schlosshauer, M. (2019). "Quantum Decoherence". Physics Reports. 831: 1–57. arXiv: 1911.06282 . Bibcode:2019PhR...831....1S. doi:10.1016/j.physrep.2019.10.001. S2CID   208006050.
  46. DiVincenzo, David; Terhal, Barbara (March 1998). "Decoherence: the obstacle to quantum computation". Physics World. 11 (3): 53–58. doi:10.1088/2058-7058/11/3/32. ISSN   0953-8585.
  47. Terhal, Barbara M. (7 April 2015). "Quantum error correction for quantum memories". Reviews of Modern Physics . 87 (2): 307–346. arXiv: 1302.3428 . Bibcode:2013arXiv1302.3428T. doi:10.1103/RevModPhys.87.307. ISSN   0034-6861. S2CID   118646257.
  48. Raussendorf, R.; Browne, D. E.; Briegel, H. J. (2003). "Measurement based Quantum Computation on Cluster States". Physical Review A . 68 (2): 022312. arXiv: quant-ph/0301052 . Bibcode:2003PhRvA..68b2312R. doi:10.1103/PhysRevA.68.022312. S2CID   6197709.
  49. Childs, Andrew M.; Leung, Debbie W.; Nielsen, Michael A. (17 March 2005). "Unified derivations of measurement-based schemes for quantum computation". Physical Review A . 71 (3): 032318. arXiv: quant-ph/0404132 . Bibcode:2005PhRvA..71c2318C. doi:10.1103/PhysRevA.71.032318. ISSN   1050-2947. S2CID   27097365.
  50. 1 2 Granade, Christopher; Combes, Joshua; Cory, D. G. (1 January 2016). "Practical Bayesian tomography". New Journal of Physics. 18 (3): 033024. arXiv: 1509.03770 . Bibcode:2016NJPh...18c3024G. doi:10.1088/1367-2630/18/3/033024. ISSN   1367-2630. S2CID   88521187.
  51. Lundeen, J. S.; Feito, A.; Coldenstrodt-Ronge, H.; Pregnell, K. L.; Silberhorn, Ch; Ralph, T. C.; Eisert, J.; Plenio, M. B.; Walmsley, I. A. (2009). "Tomography of quantum detectors". Nature Physics. 5 (1): 27–30. arXiv: 0807.2444 . Bibcode:2009NatPh...5...27L. doi:10.1038/nphys1133. ISSN   1745-2481. S2CID   119247440.
  52. Braunstein, Samuel L.; Caves, Carlton M. (30 May 1994). "Statistical distance and the geometry of quantum states". Physical Review Letters . 72 (22): 3439–3443. Bibcode:1994PhRvL..72.3439B. doi:10.1103/physrevlett.72.3439. PMID   10056200.
  53. Koberlein, Brian (5 December 2019). "LIGO Will Squeeze Light To Overcome The Quantum Noise Of Empty Space". Universe Today. Retrieved 2 February 2020.
  54. Ball, Philip (5 December 2019). "Focus: Squeezing More from Gravitational-Wave Detectors". Physics. 12. doi:10.1103/Physics.12.139. S2CID   216538409.
  55. 1 2 Peierls, Rudolf (1991). "In defence of "measurement"". Physics World . 4 (1): 19–21. doi:10.1088/2058-7058/4/1/19. ISSN   2058-7058.
  56. 1 2 Barad, Karen (2007). Meeting the Universe Halfway: Quantum Physics and the Entanglement of Matter and Meaning. Duke University Press. ISBN   978-0-8223-3917-5. OCLC   1055296186.
  57. Englert, Berthold-Georg (22 November 2013). "On quantum theory". The European Physical Journal D . 67 (11): 238. arXiv: 1308.5290 . Bibcode:2013EPJD...67..238E. doi:10.1140/epjd/e2013-40486-5. ISSN   1434-6079. S2CID   119293245.
  58. Taylor, G. I. (1909). "Interference fringes with feeble light". Proceedings of the Cambridge Philosophical Society . 15: 114–115.
  59. Gbur, Greg (25 August 2018). "Taylor sees the (feeble) light (1909)". Skulls in the Stars. Retrieved 24 October 2020.
  60. Merli, P G; Missiroli, G F; Pozzi, G (1976). "On the statistical aspect of electron interference phenomena". American Journal of Physics. 44 (3): 306–307. Bibcode:1976AmJPh..44..306M. doi:10.1119/1.10184.
  61. Arndt, Markus; Nairz, Olaf; Vos-Andreae, Julian; Keller, Claudia; Van Der Zouw, Gerbrand; Zeilinger, Anton (1999). "Wave–particle duality of C60 molecules". Nature. 401 (6754): 680–682. Bibcode:1999Natur.401..680A. doi:10.1038/44348. PMID   18494170. S2CID   4424892.
  62. Krantz, Philip; Bengtsson, Andreas; Simoen, Michaël; Gustavsson, Simon; Shumeiko, Vitaly; Oliver, W. D.; Wilson, C. M.; Delsing, Per; Bylander, Jonas (9 May 2016). "Single-shot read-out of a superconducting qubit using a Josephson parametric oscillator". Nature Communications . 7 (1): 11417. arXiv: 1508.02886 . Bibcode:2016NatCo...711417K. doi: 10.1038/ncomms11417 . ISSN   2041-1723. PMC   4865746 . PMID   27156732.
  63. Schlosshauer, Maximilian; Kofler, Johannes; Zeilinger, Anton (6 January 2013). "A Snapshot of Foundational Attitudes Toward Quantum Mechanics". Studies in History and Philosophy of Science Part B: Studies in History and Philosophy of Modern Physics. 44 (3): 222–230. arXiv: 1301.1069 . Bibcode:2013SHPMP..44..222S. doi:10.1016/j.shpsb.2013.04.004. S2CID   55537196.
  64. 1 2 Cabello, Adán (2017). "Interpretations of quantum theory: A map of madness". In Lombardi, Olimpia; Fortin, Sebastian; Holik, Federico; López, Cristian (eds.). What is Quantum Information?. Cambridge University Press. pp. 138–143. arXiv: 1509.04711 . Bibcode:2015arXiv150904711C. doi:10.1017/9781316494233.009. ISBN   9781107142114. S2CID   118419619.
  65. Schaffer, Kathryn; Barreto Lemos, Gabriela (24 May 2019). "Obliterating Thingness: An Introduction to the "What" and the "So What" of Quantum Physics". Foundations of Science. 26: 7–26. arXiv: 1908.07936 . doi:10.1007/s10699-019-09608-5. ISSN   1233-1821. S2CID   182656563.
  66. Mermin, N. David (1 July 2012). "Commentary: Quantum mechanics: Fixing the shifty split". Physics Today . 65 (7): 8–10. Bibcode:2012PhT....65g...8M. doi: 10.1063/PT.3.1618 . ISSN   0031-9228.
  67. Bub, Jeffrey; Pitowsky, Itamar (2010). "Two dogmas about quantum mechanics". Many Worlds?. Oxford University Press. pp. 433–459. arXiv: 0712.4258 . ISBN   9780199560561. OCLC   696602007.
  68. von Neumann, John (2018). Wheeler, Nicholas A. (ed.). Mathematical Foundations of Quantum Mechanics. New Edition. Translated by Robert T. Beyer. Princeton University Press. ISBN   9-781-40088-992-1. OCLC   1021172445.
  69. Wigner, E. P. (1995), "Review of the Quantum-Mechanical Measurement Problem", in Mehra, Jagdish (ed.), Philosophical Reflections and Syntheses, Springer Berlin Heidelberg, pp. 225–244, doi:10.1007/978-3-642-78374-6_19, ISBN   978-3-540-63372-3
  70. Faye, Jan (2019). "Copenhagen Interpretation of Quantum Mechanics". In Zalta, Edward N. (ed.). Stanford Encyclopedia of Philosophy. Metaphysics Research Lab, Stanford University.
  71. 1 2 Bell, John (1990). "Against 'measurement'". Physics World . 3 (8): 33–41. doi:10.1088/2058-7058/3/8/26. ISSN   2058-7058.
  72. Kent, Adrian (2010). "One world versus many: the inadequacy of Everettian accounts of evolution, probability, and scientific confirmation". Many Worlds?. Oxford University Press. pp. 307–354. arXiv: 0905.0624 . ISBN   9780199560561. OCLC   696602007.
  73. Barrett, Jeffrey (2018). "Everett's Relative-State Formulation of Quantum Mechanics". In Zalta, Edward N. (ed.). Stanford Encyclopedia of Philosophy. Metaphysics Research Lab, Stanford University.
  74. 1 2 Healey, Richard (2016). "Quantum-Bayesian and Pragmatist Views of Quantum Theory". In Zalta, Edward N. (ed.). Stanford Encyclopedia of Philosophy. Metaphysics Research Lab, Stanford University.

Further reading