sentence1
stringlengths
8
10.6k
sentence2
stringlengths
8
10.6k
label
int64
0
1
In matrix mechanics, observables such as position and momentum are represented by self-adjoint operators. When considering pairs of observables, an important quantity is the "commutator". For a pair of operators and , one defines their commutator as
Suppose, for the sake of proof by contradiction, that formula_13 is also a right eigenstate of momentum, with constant eigenvalue . If this were true, then one could write
1
In matrix mechanics, observables such as position and momentum are represented by self-adjoint operators. When considering pairs of observables, an important quantity is the "commutator". For a pair of operators and , one defines their commutator as
Throughout the main body of his original 1927 paper, written in German, Heisenberg used the word "Ungenauigkeit" ("indeterminacy"),
0
In matrix mechanics, observables such as position and momentum are represented by self-adjoint operators. When considering pairs of observables, an important quantity is the "commutator". For a pair of operators and , one defines their commutator as
On the other hand, the above canonical commutation relation requires that
1
In matrix mechanics, observables such as position and momentum are represented by self-adjoint operators. When considering pairs of observables, an important quantity is the "commutator". For a pair of operators and , one defines their commutator as
There is reason to believe that violating the uncertainty principle also strongly implies the violation of the second law of thermodynamics. See Gibbs paradox.
0
In matrix mechanics, observables such as position and momentum are represented by self-adjoint operators. When considering pairs of observables, an important quantity is the "commutator". For a pair of operators and , one defines their commutator as
When a state is measured, it is projected onto an eigenstate in the basis of the relevant observable. For example, if a particle's position is measured, then the state amounts to a position eigenstate. This means that the state is "not" a momentum eigenstate, however, but rather it can be represented as a sum of multiple momentum basis eigenstates. In other words, the momentum must be less precise. This precision may be quantified by the standard deviations,
1
In matrix mechanics, observables such as position and momentum are represented by self-adjoint operators. When considering pairs of observables, an important quantity is the "commutator". For a pair of operators and , one defines their commutator as
formulae are, beyond all doubt, derivable "statistical formulae" of the quantum theory. But they have been "habitually misinterpreted" by those quantum theorists who said that these formulae can be interpreted as determining some upper limit to the "precision of our measurements". [original emphasis
0
In the case of position and momentum, the commutator is the canonical commutation relation
In matrix mechanics, observables such as position and momentum are represented by self-adjoint operators. When considering pairs of observables, an important quantity is the "commutator". For a pair of operators and , one defines their commutator as
1
In the case of position and momentum, the commutator is the canonical commutation relation
The many-worlds interpretation originally outlined by Hugh Everett III in 1957 is partly meant to reconcile the differences between Einstein's and Bohr's views by replacing Bohr's wave function collapse with an ensemble of deterministic and independent universes whose "distribution" is governed by wave functions and the Schrödinger equation. Thus, uncertainty in the many-worlds interpretation follows from each observer within any universe having no knowledge of what goes on in the other universes.
0
In the case of position and momentum, the commutator is the canonical commutation relation
On the other hand, the above canonical commutation relation requires that
1
In the case of position and momentum, the commutator is the canonical commutation relation
but it was not always obvious what formula_69 precisely meant. The problem is that the time at which the particle has a given state is not an operator belonging to the particle, it is a parameter describing the evolution of the system. As Lev Landau once joked "To violate the time–energy uncertainty relation all I have to do is measure the energy very precisely and then look at my watch!"
0
In the case of position and momentum, the commutator is the canonical commutation relation
When a state is measured, it is projected onto an eigenstate in the basis of the relevant observable. For example, if a particle's position is measured, then the state amounts to a position eigenstate. This means that the state is "not" a momentum eigenstate, however, but rather it can be represented as a sum of multiple momentum basis eigenstates. In other words, the momentum must be less precise. This precision may be quantified by the standard deviations,
1
In the case of position and momentum, the commutator is the canonical commutation relation
Karl Popper approached the problem of indeterminacy as a logician and metaphysical realist. He disagreed with the application of the uncertainty relations to individual particles rather than to ensembles of identically prepared particles, referring to them as "statistical scatter relations". In this statistical interpretation, a "particular" measurement may be made to arbitrary precision without invalidating the quantum theory. This directly contrasts with the Copenhagen interpretation of quantum mechanics, which is non-deterministic but lacks local hidden variables.
0
The physical meaning of the non-commutativity can be understood by considering the effect of the commutator on position and momentum eigenstates. Let formula_13 be a right eigenstate of position with a constant eigenvalue . By definition, this means that formula_14 Applying the commutator to formula_13 yields
When a state is measured, it is projected onto an eigenstate in the basis of the relevant observable. For example, if a particle's position is measured, then the state amounts to a position eigenstate. This means that the state is "not" a momentum eigenstate, however, but rather it can be represented as a sum of multiple momentum basis eigenstates. In other words, the momentum must be less precise. This precision may be quantified by the standard deviations,
1
The physical meaning of the non-commutativity can be understood by considering the effect of the commutator on position and momentum eigenstates. Let formula_13 be a right eigenstate of position with a constant eigenvalue . By definition, this means that formula_14 Applying the commutator to formula_13 yields
He imagines an experimenter trying to measure the position and momentum of an electron by shooting a photon at it.
0
The physical meaning of the non-commutativity can be understood by considering the effect of the commutator on position and momentum eigenstates. Let formula_13 be a right eigenstate of position with a constant eigenvalue . By definition, this means that formula_14 Applying the commutator to formula_13 yields
On the other hand, the above canonical commutation relation requires that
1
The physical meaning of the non-commutativity can be understood by considering the effect of the commutator on position and momentum eigenstates. Let formula_13 be a right eigenstate of position with a constant eigenvalue . By definition, this means that formula_14 Applying the commutator to formula_13 yields
The uncertainty principle is not readily apparent on the macroscopic scales of everyday experience. So it is helpful to demonstrate how it applies to more easily understood physical situations. Two alternative frameworks for quantum physics offer different explanations for the uncertainty principle. The wave mechanics picture of the uncertainty principle is more visually intuitive, but the more abstract matrix mechanics picture formulates it in a way that generalizes more easily.
0
The physical meaning of the non-commutativity can be understood by considering the effect of the commutator on position and momentum eigenstates. Let formula_13 be a right eigenstate of position with a constant eigenvalue . By definition, this means that formula_14 Applying the commutator to formula_13 yields
In matrix mechanics, observables such as position and momentum are represented by self-adjoint operators. When considering pairs of observables, an important quantity is the "commutator". For a pair of operators and , one defines their commutator as
1
The physical meaning of the non-commutativity can be understood by considering the effect of the commutator on position and momentum eigenstates. Let formula_13 be a right eigenstate of position with a constant eigenvalue . By definition, this means that formula_14 Applying the commutator to formula_13 yields
Some scientists including Arthur Compton and Martin Heisenberg have suggested that the uncertainty principle, or at least the general probabilistic nature of quantum mechanics, could be evidence for the two-stage model of free will. One critique, however, is that apart from the basic role of quantum mechanics as a foundation for chemistry, nontrivial biological mechanisms requiring quantum mechanics are unlikely, due to the rapid decoherence time of quantum systems at room temperature. Proponents of this theory commonly say that this decoherence is overcome by both screening and decoherence-free subspaces found in biological cells.
0
Suppose, for the sake of proof by contradiction, that formula_13 is also a right eigenstate of momentum, with constant eigenvalue . If this were true, then one could write
As in the wave mechanics interpretation above, one sees a tradeoff between the respective precisions of the two, quantified by the uncertainty principle.
1
Suppose, for the sake of proof by contradiction, that formula_13 is also a right eigenstate of momentum, with constant eigenvalue . If this were true, then one could write
The Robertson–Schrödinger uncertainty relation may be generalized in a straightforward way to describe mixed states.,
0
Suppose, for the sake of proof by contradiction, that formula_13 is also a right eigenstate of momentum, with constant eigenvalue . If this were true, then one could write
When a state is measured, it is projected onto an eigenstate in the basis of the relevant observable. For example, if a particle's position is measured, then the state amounts to a position eigenstate. This means that the state is "not" a momentum eigenstate, however, but rather it can be represented as a sum of multiple momentum basis eigenstates. In other words, the momentum must be less precise. This precision may be quantified by the standard deviations,
1
Suppose, for the sake of proof by contradiction, that formula_13 is also a right eigenstate of momentum, with constant eigenvalue . If this were true, then one could write
The Robertson–Schrödinger uncertainty relation may be generalized in a straightforward way to describe mixed states.,
0
Suppose, for the sake of proof by contradiction, that formula_13 is also a right eigenstate of momentum, with constant eigenvalue . If this were true, then one could write
On the other hand, the above canonical commutation relation requires that
1
Suppose, for the sake of proof by contradiction, that formula_13 is also a right eigenstate of momentum, with constant eigenvalue . If this were true, then one could write
The Robertson–Schrödinger uncertainty relation may be generalized in a straightforward way to describe mixed states.,
0
On the other hand, the above canonical commutation relation requires that
This implies that no quantum state can simultaneously be both a position and a momentum eigenstate.
1
On the other hand, the above canonical commutation relation requires that
The many-worlds interpretation originally outlined by Hugh Everett III in 1957 is partly meant to reconcile the differences between Einstein's and Bohr's views by replacing Bohr's wave function collapse with an ensemble of deterministic and independent universes whose "distribution" is governed by wave functions and the Schrödinger equation. Thus, uncertainty in the many-worlds interpretation follows from each observer within any universe having no knowledge of what goes on in the other universes.
0
On the other hand, the above canonical commutation relation requires that
In the case of position and momentum, the commutator is the canonical commutation relation
1
On the other hand, the above canonical commutation relation requires that
The principle is quite counter-intuitive, so the early students of quantum theory had to be reassured that naive measurements to violate it were bound always to be unworkable. One way in which Heisenberg originally illustrated the intrinsic impossibility of violating the uncertainty principle is by utilizing the observer effect of an imaginary microscope as a measuring device.
0
On the other hand, the above canonical commutation relation requires that
Suppose, for the sake of proof by contradiction, that formula_13 is also a right eigenstate of momentum, with constant eigenvalue . If this were true, then one could write
1
On the other hand, the above canonical commutation relation requires that
Wolfgang Pauli called Einstein's fundamental objection to the uncertainty principle "the ideal of the detached observer" (phrase translated from the German):
0
This implies that no quantum state can simultaneously be both a position and a momentum eigenstate.
As in the wave mechanics interpretation above, one sees a tradeoff between the respective precisions of the two, quantified by the uncertainty principle.
1
This implies that no quantum state can simultaneously be both a position and a momentum eigenstate.
In particular, the above Kennard bound is saturated for the ground state , for which the probability density is just the normal distribution.
0
This implies that no quantum state can simultaneously be both a position and a momentum eigenstate.
When a state is measured, it is projected onto an eigenstate in the basis of the relevant observable. For example, if a particle's position is measured, then the state amounts to a position eigenstate. This means that the state is "not" a momentum eigenstate, however, but rather it can be represented as a sum of multiple momentum basis eigenstates. In other words, the momentum must be less precise. This precision may be quantified by the standard deviations,
1
This implies that no quantum state can simultaneously be both a position and a momentum eigenstate.
In his celebrated 1927 paper, "Über den anschaulichen Inhalt der quantentheoretischen Kinematik und Mechanik" ("On the Perceptual Content of Quantum Theoretical Kinematics and Mechanics"), Heisenberg established this expression as the minimum amount of unavoidable momentum disturbance caused by any position measurement, but he did not give a precise definition for the uncertainties Δx and Δp. Instead, he gave some plausible estimates in each case separately. In his Chicago lecture he refined his principle:
0
This implies that no quantum state can simultaneously be both a position and a momentum eigenstate.
On the other hand, the above canonical commutation relation requires that
1
This implies that no quantum state can simultaneously be both a position and a momentum eigenstate.
The Robertson–Schrödinger uncertainty relation may be generalized in a straightforward way to describe mixed states.,
0
When a state is measured, it is projected onto an eigenstate in the basis of the relevant observable. For example, if a particle's position is measured, then the state amounts to a position eigenstate. This means that the state is "not" a momentum eigenstate, however, but rather it can be represented as a sum of multiple momentum basis eigenstates. In other words, the momentum must be less precise. This precision may be quantified by the standard deviations,
This implies that no quantum state can simultaneously be both a position and a momentum eigenstate.
1
When a state is measured, it is projected onto an eigenstate in the basis of the relevant observable. For example, if a particle's position is measured, then the state amounts to a position eigenstate. This means that the state is "not" a momentum eigenstate, however, but rather it can be represented as a sum of multiple momentum basis eigenstates. In other words, the momentum must be less precise. This precision may be quantified by the standard deviations,
Throughout the main body of his original 1927 paper, written in German, Heisenberg used the word "Ungenauigkeit" ("indeterminacy"),
0
When a state is measured, it is projected onto an eigenstate in the basis of the relevant observable. For example, if a particle's position is measured, then the state amounts to a position eigenstate. This means that the state is "not" a momentum eigenstate, however, but rather it can be represented as a sum of multiple momentum basis eigenstates. In other words, the momentum must be less precise. This precision may be quantified by the standard deviations,
As in the wave mechanics interpretation above, one sees a tradeoff between the respective precisions of the two, quantified by the uncertainty principle.
1
When a state is measured, it is projected onto an eigenstate in the basis of the relevant observable. For example, if a particle's position is measured, then the state amounts to a position eigenstate. This means that the state is "not" a momentum eigenstate, however, but rather it can be represented as a sum of multiple momentum basis eigenstates. In other words, the momentum must be less precise. This precision may be quantified by the standard deviations,
The first of Einstein's thought experiments challenging the uncertainty principle went as follows:
0
When a state is measured, it is projected onto an eigenstate in the basis of the relevant observable. For example, if a particle's position is measured, then the state amounts to a position eigenstate. This means that the state is "not" a momentum eigenstate, however, but rather it can be represented as a sum of multiple momentum basis eigenstates. In other words, the momentum must be less precise. This precision may be quantified by the standard deviations,
The physical meaning of the non-commutativity can be understood by considering the effect of the commutator on position and momentum eigenstates. Let formula_13 be a right eigenstate of position with a constant eigenvalue . By definition, this means that formula_14 Applying the commutator to formula_13 yields
1
When a state is measured, it is projected onto an eigenstate in the basis of the relevant observable. For example, if a particle's position is measured, then the state amounts to a position eigenstate. This means that the state is "not" a momentum eigenstate, however, but rather it can be represented as a sum of multiple momentum basis eigenstates. In other words, the momentum must be less precise. This precision may be quantified by the standard deviations,
Some scientists including Arthur Compton and Martin Heisenberg have suggested that the uncertainty principle, or at least the general probabilistic nature of quantum mechanics, could be evidence for the two-stage model of free will. One critique, however, is that apart from the basic role of quantum mechanics as a foundation for chemistry, nontrivial biological mechanisms requiring quantum mechanics are unlikely, due to the rapid decoherence time of quantum systems at room temperature. Proponents of this theory commonly say that this decoherence is overcome by both screening and decoherence-free subspaces found in biological cells.
0
As in the wave mechanics interpretation above, one sees a tradeoff between the respective precisions of the two, quantified by the uncertainty principle.
The physical meaning of the non-commutativity can be understood by considering the effect of the commutator on position and momentum eigenstates. Let formula_13 be a right eigenstate of position with a constant eigenvalue . By definition, this means that formula_14 Applying the commutator to formula_13 yields
1
As in the wave mechanics interpretation above, one sees a tradeoff between the respective precisions of the two, quantified by the uncertainty principle.
to describe the basic theoretical principle. Only in the endnote did he switch to the word "Unsicherheit" ("uncertainty"). When the English-language version of Heisenberg's textbook, "The Physical Principles of the Quantum Theory", was published in 1930, however, the translation "uncertainty" was used, and it became the more commonly used term in the English language thereafter.
0
As in the wave mechanics interpretation above, one sees a tradeoff between the respective precisions of the two, quantified by the uncertainty principle.
In the case of position and momentum, the commutator is the canonical commutation relation
1
As in the wave mechanics interpretation above, one sees a tradeoff between the respective precisions of the two, quantified by the uncertainty principle.
Bohr's response was that the wall is quantum mechanical as well, and that to measure the recoil to accuracy , the momentum of the wall must be known to this accuracy before the particle passes through. This introduces an uncertainty in the position of the wall and therefore the position of the slit equal to , and if the wall's momentum is known precisely enough to measure the recoil, the slit's position is uncertain enough to disallow a position measurement.
0
As in the wave mechanics interpretation above, one sees a tradeoff between the respective precisions of the two, quantified by the uncertainty principle.
When a state is measured, it is projected onto an eigenstate in the basis of the relevant observable. For example, if a particle's position is measured, then the state amounts to a position eigenstate. This means that the state is "not" a momentum eigenstate, however, but rather it can be represented as a sum of multiple momentum basis eigenstates. In other words, the momentum must be less precise. This precision may be quantified by the standard deviations,
1
As in the wave mechanics interpretation above, one sees a tradeoff between the respective precisions of the two, quantified by the uncertainty principle.
This was later improved as follows: if formula_128 is such that
0
The most common general form of the uncertainty principle is the "Robertson uncertainty relation".
For an arbitrary Hermitian operator formula_22 we can associate a standard deviation
1
The most common general form of the uncertainty principle is the "Robertson uncertainty relation".
Bohr's response was that the wall is quantum mechanical as well, and that to measure the recoil to accuracy , the momentum of the wall must be known to this accuracy before the particle passes through. This introduces an uncertainty in the position of the wall and therefore the position of the slit equal to , and if the wall's momentum is known precisely enough to measure the recoil, the slit's position is uncertain enough to disallow a position measurement.
0
For an arbitrary Hermitian operator formula_22 we can associate a standard deviation
The most common general form of the uncertainty principle is the "Robertson uncertainty relation".
1
For an arbitrary Hermitian operator formula_22 we can associate a standard deviation
A coherent state is a right eigenstate of the annihilation operator,
0
where the brackets formula_24 indicate an expectation value. For a pair of operators formula_25 and formula_26, we may define their "commutator" as
For an arbitrary Hermitian operator formula_22 we can associate a standard deviation
1
where the brackets formula_24 indicate an expectation value. For a pair of operators formula_25 and formula_26, we may define their "commutator" as
In 1964, John Bell showed that this assumption can be falsified, since it would imply a certain inequality between the probabilities of different experiments. Experimental results confirm the predictions of quantum mechanics, ruling out Einstein's basic assumption that led him to the suggestion of his "hidden variables". These hidden variables may be "hidden" because of an illusion that occurs during observations of objects that are too large or too small. This illusion can be likened to rotating fan blades that seem to pop in and out of existence at different locations and sometimes seem to be in the same place at the same time when observed. This same illusion manifests itself in the observation of subatomic particles. Both the fan blades and the subatomic particles are moving so fast that the illusion is seen by the observer. Therefore, it is possible that there would be predictability of the subatomic particles behavior and characteristics to a recording device capable of very high speed tracking...Ironically this fact is one of the best pieces of evidence supporting Karl Popper's philosophy of invalidation of a theory by falsification-experiments. That is to say, here Einstein's "basic assumption" became falsified by experiments based on Bell's inequalities. For the objections of Karl Popper to the Heisenberg inequality itself, see below.
0
The Robertson uncertainty relation immediately follows from a slightly stronger inequality, the "Schrödinger uncertainty relation",
The most common general form of the uncertainty principle is the "Robertson uncertainty relation".
1
The Robertson uncertainty relation immediately follows from a slightly stronger inequality, the "Schrödinger uncertainty relation",
In quantum mechanics, the uncertainty principle (also known as Heisenberg's uncertainty principle) is any of a variety of mathematical inequalities asserting a fundamental limit to the accuracy with which the values for certain pairs of physical quantities of a particle, such as position, "x", and momentum, "p", can be predicted from initial conditions.
0
The Robertson–Schrödinger uncertainty relation can be trivial if the state of the system is chosen to be eigenstate of one of the observable. The stronger uncertainty relations proved by Maccone and Pati give non-trivial bounds on the sum of the variances for two incompatible observables. (Earlier works on uncertainty relations formulated as the sum of variances include, e.g., Ref. due to Huang.) For two non-commuting observables formula_31 and formula_32 the first stronger uncertainty relation is given by
is nonzero unless formula_43 is an eigenstate of formula_44. One may note that formula_37 can be an eigenstate of formula_46 without being an eigenstate of either
1
The Robertson–Schrödinger uncertainty relation can be trivial if the state of the system is chosen to be eigenstate of one of the observable. The stronger uncertainty relations proved by Maccone and Pati give non-trivial bounds on the sum of the variances for two incompatible observables. (Earlier works on uncertainty relations formulated as the sum of variances include, e.g., Ref. due to Huang.) For two non-commuting observables formula_31 and formula_32 the first stronger uncertainty relation is given by
Heisenberg's paper did not admit any unobservable quantities like the exact position of the electron in an orbit at any time; he only allowed the theorist to talk about the Fourier components of the motion. Since the Fourier components were not defined at the classical frequencies, they could not be used to construct an exact trajectory, so that the formalism could not answer certain overly precise questions about where the electron was or how fast it was going.
0
The Robertson–Schrödinger uncertainty relation can be trivial if the state of the system is chosen to be eigenstate of one of the observable. The stronger uncertainty relations proved by Maccone and Pati give non-trivial bounds on the sum of the variances for two incompatible observables. (Earlier works on uncertainty relations formulated as the sum of variances include, e.g., Ref. due to Huang.) For two non-commuting observables formula_31 and formula_32 the first stronger uncertainty relation is given by
formula_47 or formula_48. However, when formula_41 is an eigenstate of one of the two observables the Heisenberg–Schrödinger uncertainty relation becomes trivial. But the lower bound in the new relation is nonzero
1
The Robertson–Schrödinger uncertainty relation can be trivial if the state of the system is chosen to be eigenstate of one of the observable. The stronger uncertainty relations proved by Maccone and Pati give non-trivial bounds on the sum of the variances for two incompatible observables. (Earlier works on uncertainty relations formulated as the sum of variances include, e.g., Ref. due to Huang.) For two non-commuting observables formula_31 and formula_32 the first stronger uncertainty relation is given by
Bohr was compelled to modify his understanding of the uncertainty principle after another thought experiment by Einstein. In 1935, Einstein, Podolsky and Rosen (see EPR paradox) published an analysis of widely separated entangled particles. Measuring one particle, Einstein realized, would alter the probability distribution of the other, yet here the other particle could not possibly be disturbed. This example led Bohr to revise his understanding of the principle, concluding that the uncertainty was not caused by a direct interaction.
0
where formula_34, formula_35, formula_36 is a normalized vector that is orthogonal to the state of the system formula_37 and one should choose the sign of formula_38 to make this real quantity a positive number.
The Robertson–Schrödinger uncertainty relation can be trivial if the state of the system is chosen to be eigenstate of one of the observable. The stronger uncertainty relations proved by Maccone and Pati give non-trivial bounds on the sum of the variances for two incompatible observables. (Earlier works on uncertainty relations formulated as the sum of variances include, e.g., Ref. due to Huang.) For two non-commuting observables formula_31 and formula_32 the first stronger uncertainty relation is given by
1
where formula_34, formula_35, formula_36 is a normalized vector that is orthogonal to the state of the system formula_37 and one should choose the sign of formula_38 to make this real quantity a positive number.
There is reason to believe that violating the uncertainty principle also strongly implies the violation of the second law of thermodynamics. See Gibbs paradox.
0
where formula_34, formula_35, formula_36 is a normalized vector that is orthogonal to the state of the system formula_37 and one should choose the sign of formula_38 to make this real quantity a positive number.
formula_47 or formula_48. However, when formula_41 is an eigenstate of one of the two observables the Heisenberg–Schrödinger uncertainty relation becomes trivial. But the lower bound in the new relation is nonzero
1
where formula_34, formula_35, formula_36 is a normalized vector that is orthogonal to the state of the system formula_37 and one should choose the sign of formula_38 to make this real quantity a positive number.
He imagines an experimenter trying to measure the position and momentum of an electron by shooting a photon at it.
0
The form of formula_42 implies that the right-hand side of the new uncertainty relation
is nonzero unless formula_43 is an eigenstate of formula_44. One may note that formula_37 can be an eigenstate of formula_46 without being an eigenstate of either
1
The form of formula_42 implies that the right-hand side of the new uncertainty relation
Wolfgang Pauli called Einstein's fundamental objection to the uncertainty principle "the ideal of the detached observer" (phrase translated from the German):
0
The form of formula_42 implies that the right-hand side of the new uncertainty relation
formula_47 or formula_48. However, when formula_41 is an eigenstate of one of the two observables the Heisenberg–Schrödinger uncertainty relation becomes trivial. But the lower bound in the new relation is nonzero
1
The form of formula_42 implies that the right-hand side of the new uncertainty relation
The many-worlds interpretation originally outlined by Hugh Everett III in 1957 is partly meant to reconcile the differences between Einstein's and Bohr's views by replacing Bohr's wave function collapse with an ensemble of deterministic and independent universes whose "distribution" is governed by wave functions and the Schrödinger equation. Thus, uncertainty in the many-worlds interpretation follows from each observer within any universe having no knowledge of what goes on in the other universes.
0
is nonzero unless formula_43 is an eigenstate of formula_44. One may note that formula_37 can be an eigenstate of formula_46 without being an eigenstate of either
The form of formula_42 implies that the right-hand side of the new uncertainty relation
1
is nonzero unless formula_43 is an eigenstate of formula_44. One may note that formula_37 can be an eigenstate of formula_46 without being an eigenstate of either
There is reason to believe that violating the uncertainty principle also strongly implies the violation of the second law of thermodynamics. See Gibbs paradox.
0
is nonzero unless formula_43 is an eigenstate of formula_44. One may note that formula_37 can be an eigenstate of formula_46 without being an eigenstate of either
formula_47 or formula_48. However, when formula_41 is an eigenstate of one of the two observables the Heisenberg–Schrödinger uncertainty relation becomes trivial. But the lower bound in the new relation is nonzero
1
is nonzero unless formula_43 is an eigenstate of formula_44. One may note that formula_37 can be an eigenstate of formula_46 without being an eigenstate of either
Consider a one-dimensional quantum harmonic oscillator. It is possible to express the position and momentum operators in terms of the creation and annihilation operators:
0
formula_47 or formula_48. However, when formula_41 is an eigenstate of one of the two observables the Heisenberg–Schrödinger uncertainty relation becomes trivial. But the lower bound in the new relation is nonzero
The form of formula_42 implies that the right-hand side of the new uncertainty relation
1
formula_47 or formula_48. However, when formula_41 is an eigenstate of one of the two observables the Heisenberg–Schrödinger uncertainty relation becomes trivial. But the lower bound in the new relation is nonzero
When a state is measured, it is projected onto an eigenstate in the basis of the relevant observable. For example, if a particle's position is measured, then the state amounts to a position eigenstate. This means that the state is "not" a momentum eigenstate, however, but rather it can be represented as a sum of multiple momentum basis eigenstates. In other words, the momentum must be less precise. This precision may be quantified by the standard deviations,
0
formula_47 or formula_48. However, when formula_41 is an eigenstate of one of the two observables the Heisenberg–Schrödinger uncertainty relation becomes trivial. But the lower bound in the new relation is nonzero
is nonzero unless formula_43 is an eigenstate of formula_44. One may note that formula_37 can be an eigenstate of formula_46 without being an eigenstate of either
1
formula_47 or formula_48. However, when formula_41 is an eigenstate of one of the two observables the Heisenberg–Schrödinger uncertainty relation becomes trivial. But the lower bound in the new relation is nonzero
According to the de Broglie hypothesis, every object in the universe is a wave, i.e., a situation which gives rise to this phenomenon. The position of the particle is described by a wave function formula_1. The time-independent wave function of a single-moded plane wave of wavenumber "k"0 or momentum "p"0 is
0
In the phase space formulation of quantum mechanics, the Robertson–Schrödinger relation follows from a positivity condition on a real star-square function. Given a Wigner function formula_51 with star product ★ and a function "f", the following is generally true:
The non-negative eigenvalues then imply a corresponding non-negativity condition on the determinant,
1
In the phase space formulation of quantum mechanics, the Robertson–Schrödinger relation follows from a positivity condition on a real star-square function. Given a Wigner function formula_51 with star product ★ and a function "f", the following is generally true:
There is reason to believe that violating the uncertainty principle also strongly implies the violation of the second law of thermodynamics. See Gibbs paradox.
0
Since this positivity condition is true for "all" "a", "b", and "c", it follows that all the eigenvalues of the matrix are non-negative.
The non-negative eigenvalues then imply a corresponding non-negativity condition on the determinant,
1
Since this positivity condition is true for "all" "a", "b", and "c", it follows that all the eigenvalues of the matrix are non-negative.
The Robertson–Schrödinger uncertainty relation can be trivial if the state of the system is chosen to be eigenstate of one of the observable. The stronger uncertainty relations proved by Maccone and Pati give non-trivial bounds on the sum of the variances for two incompatible observables. (Earlier works on uncertainty relations formulated as the sum of variances include, e.g., Ref. due to Huang.) For two non-commuting observables formula_31 and formula_32 the first stronger uncertainty relation is given by
0
The non-negative eigenvalues then imply a corresponding non-negativity condition on the determinant,
In the phase space formulation of quantum mechanics, the Robertson–Schrödinger relation follows from a positivity condition on a real star-square function. Given a Wigner function formula_51 with star product ★ and a function "f", the following is generally true:
1
The non-negative eigenvalues then imply a corresponding non-negativity condition on the determinant,
In matrix mechanics, observables such as position and momentum are represented by self-adjoint operators. When considering pairs of observables, an important quantity is the "commutator". For a pair of operators and , one defines their commutator as
0
Since the Robertson and Schrödinger relations are for general operators, the relations can be applied to any two observables to obtain specific uncertainty relations. A few of the most common relations found in the literature are given below.
One "false" formulation of the energy–time uncertainty principle says that measuring the energy of a quantum system to an accuracy formula_70 requires a time interval formula_71. This formulation is similar to the one alluded to in Landau's joke, and was explicitly invalidated by Y. Aharonov and D. Bohm in 1961. The time formula_69 in the uncertainty relation is the time during which the system exists unperturbed, not the time during which the experimental equipment is turned on, whereas the position in the other version of the principle refers to where the particle has some probability to be and not where the observer might look.
1
Since the Robertson and Schrödinger relations are for general operators, the relations can be applied to any two observables to obtain specific uncertainty relations. A few of the most common relations found in the literature are given below.
The Copenhagen interpretation of quantum mechanics and Heisenberg's Uncertainty Principle were, in fact, seen as twin targets by detractors who believed in an underlying determinism and realism. According to the Copenhagen interpretation of quantum mechanics, there is no fundamental reality that the quantum state describes, just a prescription for calculating experimental results. There is no way to say what the state of a system fundamentally is, only what the result of observations might be.
0
Since the Robertson and Schrödinger relations are for general operators, the relations can be applied to any two observables to obtain specific uncertainty relations. A few of the most common relations found in the literature are given below.
but it was not always obvious what formula_69 precisely meant. The problem is that the time at which the particle has a given state is not an operator belonging to the particle, it is a parameter describing the evolution of the system. As Lev Landau once joked "To violate the time–energy uncertainty relation all I have to do is measure the energy very precisely and then look at my watch!"
1
Since the Robertson and Schrödinger relations are for general operators, the relations can be applied to any two observables to obtain specific uncertainty relations. A few of the most common relations found in the literature are given below.
Albert Einstein believed that randomness is a reflection of our ignorance of some fundamental property of reality, while Niels Bohr believed that the probability distributions are fundamental and irreducible, and depend on which measurements we choose to perform. Einstein and Bohr debated the uncertainty principle for many years.
0
but it was not always obvious what formula_69 precisely meant. The problem is that the time at which the particle has a given state is not an operator belonging to the particle, it is a parameter describing the evolution of the system. As Lev Landau once joked "To violate the time–energy uncertainty relation all I have to do is measure the energy very precisely and then look at my watch!"
In 1932 Dirac offered a precise definition and derivation of the time–energy uncertainty relation in a relativistic quantum theory of "events".
1
but it was not always obvious what formula_69 precisely meant. The problem is that the time at which the particle has a given state is not an operator belonging to the particle, it is a parameter describing the evolution of the system. As Lev Landau once joked "To violate the time–energy uncertainty relation all I have to do is measure the energy very precisely and then look at my watch!"
The first of Einstein's thought experiments challenging the uncertainty principle went as follows:
0
but it was not always obvious what formula_69 precisely meant. The problem is that the time at which the particle has a given state is not an operator belonging to the particle, it is a parameter describing the evolution of the system. As Lev Landau once joked "To violate the time–energy uncertainty relation all I have to do is measure the energy very precisely and then look at my watch!"
Since the Robertson and Schrödinger relations are for general operators, the relations can be applied to any two observables to obtain specific uncertainty relations. A few of the most common relations found in the literature are given below.
1
but it was not always obvious what formula_69 precisely meant. The problem is that the time at which the particle has a given state is not an operator belonging to the particle, it is a parameter describing the evolution of the system. As Lev Landau once joked "To violate the time–energy uncertainty relation all I have to do is measure the energy very precisely and then look at my watch!"
for some convenient polynomial and real positive definite matrix of type .
0
One "false" formulation of the energy–time uncertainty principle says that measuring the energy of a quantum system to an accuracy formula_70 requires a time interval formula_71. This formulation is similar to the one alluded to in Landau's joke, and was explicitly invalidated by Y. Aharonov and D. Bohm in 1961. The time formula_69 in the uncertainty relation is the time during which the system exists unperturbed, not the time during which the experimental equipment is turned on, whereas the position in the other version of the principle refers to where the particle has some probability to be and not where the observer might look.
Since the Robertson and Schrödinger relations are for general operators, the relations can be applied to any two observables to obtain specific uncertainty relations. A few of the most common relations found in the literature are given below.
1
One "false" formulation of the energy–time uncertainty principle says that measuring the energy of a quantum system to an accuracy formula_70 requires a time interval formula_71. This formulation is similar to the one alluded to in Landau's joke, and was explicitly invalidated by Y. Aharonov and D. Bohm in 1961. The time formula_69 in the uncertainty relation is the time during which the system exists unperturbed, not the time during which the experimental equipment is turned on, whereas the position in the other version of the principle refers to where the particle has some probability to be and not where the observer might look.
Although this result appears to violate the Robertson uncertainty principle, the paradox is resolved when we note that formula_84 is not in the domain of the operator formula_94, since multiplication by formula_73 disrupts the periodic boundary conditions imposed on formula_26. Thus, the derivation of the Robertson relation, which requires formula_97 and formula_98 to be defined, does not apply. (These also furnish an example of operators satisfying the canonical commutation relations but not the Weyl relations.)
0
One "false" formulation of the energy–time uncertainty principle says that measuring the energy of a quantum system to an accuracy formula_70 requires a time interval formula_71. This formulation is similar to the one alluded to in Landau's joke, and was explicitly invalidated by Y. Aharonov and D. Bohm in 1961. The time formula_69 in the uncertainty relation is the time during which the system exists unperturbed, not the time during which the experimental equipment is turned on, whereas the position in the other version of the principle refers to where the particle has some probability to be and not where the observer might look.
Another common misconception is that the energy–time uncertainty principle says that the conservation of energy can be temporarily violated—energy can be "borrowed" from the universe as long as it is "returned" within a short amount of time. Although this agrees with the "spirit" of relativistic quantum mechanics, it is based on the false axiom that the energy of the universe is an exactly known parameter at all times. More accurately, when events transpire at shorter time intervals, there is a greater uncertainty in the energy of these events. Therefore it is not that the conservation of energy is "violated" when quantum field theory uses temporary electron–positron pairs in its calculations, but that the energy of quantum systems is not known with enough precision to limit their behavior to a single, simple history. Thus the influence of "all histories" must be incorporated into quantum calculations, including those with much greater or much less energy than the mean of the measured/calculated energy distribution.
1
One "false" formulation of the energy–time uncertainty principle says that measuring the energy of a quantum system to an accuracy formula_70 requires a time interval formula_71. This formulation is similar to the one alluded to in Landau's joke, and was explicitly invalidated by Y. Aharonov and D. Bohm in 1961. The time formula_69 in the uncertainty relation is the time during which the system exists unperturbed, not the time during which the experimental equipment is turned on, whereas the position in the other version of the principle refers to where the particle has some probability to be and not where the observer might look.
A similar analysis with particles diffracting through multiple slits is given by Richard Feynman.
0
Another common misconception is that the energy–time uncertainty principle says that the conservation of energy can be temporarily violated—energy can be "borrowed" from the universe as long as it is "returned" within a short amount of time. Although this agrees with the "spirit" of relativistic quantum mechanics, it is based on the false axiom that the energy of the universe is an exactly known parameter at all times. More accurately, when events transpire at shorter time intervals, there is a greater uncertainty in the energy of these events. Therefore it is not that the conservation of energy is "violated" when quantum field theory uses temporary electron–positron pairs in its calculations, but that the energy of quantum systems is not known with enough precision to limit their behavior to a single, simple history. Thus the influence of "all histories" must be incorporated into quantum calculations, including those with much greater or much less energy than the mean of the measured/calculated energy distribution.
In 1932 Dirac offered a precise definition and derivation of the time–energy uncertainty relation in a relativistic quantum theory of "events".
1
Another common misconception is that the energy–time uncertainty principle says that the conservation of energy can be temporarily violated—energy can be "borrowed" from the universe as long as it is "returned" within a short amount of time. Although this agrees with the "spirit" of relativistic quantum mechanics, it is based on the false axiom that the energy of the universe is an exactly known parameter at all times. More accurately, when events transpire at shorter time intervals, there is a greater uncertainty in the energy of these events. Therefore it is not that the conservation of energy is "violated" when quantum field theory uses temporary electron–positron pairs in its calculations, but that the energy of quantum systems is not known with enough precision to limit their behavior to a single, simple history. Thus the influence of "all histories" must be incorporated into quantum calculations, including those with much greater or much less energy than the mean of the measured/calculated energy distribution.
There is reason to believe that violating the uncertainty principle also strongly implies the violation of the second law of thermodynamics. See Gibbs paradox.
0
Another common misconception is that the energy–time uncertainty principle says that the conservation of energy can be temporarily violated—energy can be "borrowed" from the universe as long as it is "returned" within a short amount of time. Although this agrees with the "spirit" of relativistic quantum mechanics, it is based on the false axiom that the energy of the universe is an exactly known parameter at all times. More accurately, when events transpire at shorter time intervals, there is a greater uncertainty in the energy of these events. Therefore it is not that the conservation of energy is "violated" when quantum field theory uses temporary electron–positron pairs in its calculations, but that the energy of quantum systems is not known with enough precision to limit their behavior to a single, simple history. Thus the influence of "all histories" must be incorporated into quantum calculations, including those with much greater or much less energy than the mean of the measured/calculated energy distribution.
Since the Robertson and Schrödinger relations are for general operators, the relations can be applied to any two observables to obtain specific uncertainty relations. A few of the most common relations found in the literature are given below.
1
Another common misconception is that the energy–time uncertainty principle says that the conservation of energy can be temporarily violated—energy can be "borrowed" from the universe as long as it is "returned" within a short amount of time. Although this agrees with the "spirit" of relativistic quantum mechanics, it is based on the false axiom that the energy of the universe is an exactly known parameter at all times. More accurately, when events transpire at shorter time intervals, there is a greater uncertainty in the energy of these events. Therefore it is not that the conservation of energy is "violated" when quantum field theory uses temporary electron–positron pairs in its calculations, but that the energy of quantum systems is not known with enough precision to limit their behavior to a single, simple history. Thus the influence of "all histories" must be incorporated into quantum calculations, including those with much greater or much less energy than the mean of the measured/calculated energy distribution.
Some scientists including Arthur Compton and Martin Heisenberg have suggested that the uncertainty principle, or at least the general probabilistic nature of quantum mechanics, could be evidence for the two-stage model of free will. One critique, however, is that apart from the basic role of quantum mechanics as a foundation for chemistry, nontrivial biological mechanisms requiring quantum mechanics are unlikely, due to the rapid decoherence time of quantum systems at room temperature. Proponents of this theory commonly say that this decoherence is overcome by both screening and decoherence-free subspaces found in biological cells.
0
In 1932 Dirac offered a precise definition and derivation of the time–energy uncertainty relation in a relativistic quantum theory of "events".
but it was not always obvious what formula_69 precisely meant. The problem is that the time at which the particle has a given state is not an operator belonging to the particle, it is a parameter describing the evolution of the system. As Lev Landau once joked "To violate the time–energy uncertainty relation all I have to do is measure the energy very precisely and then look at my watch!"
1
In 1932 Dirac offered a precise definition and derivation of the time–energy uncertainty relation in a relativistic quantum theory of "events".
In a quantum harmonic oscillator of characteristic angular frequency ω, place a state that is offset from the bottom of the potential by some displacement "x"0 as
0