Quantum Zeno effect

Last updated
With the increasing number of measurements the wave function tends to stay in its initial form. In the animation, a free time evolution of a wave function, depicted on the left, is in the central part interrupted by occasional position measurements that localize the wave function in one of nine sectors. On the right, a series of very frequent measurements leads to the quantum Zeno effect. Quantum Zeno effect animation.gif
With the increasing number of measurements the wave function tends to stay in its initial form. In the animation, a free time evolution of a wave function, depicted on the left, is in the central part interrupted by occasional position measurements that localize the wave function in one of nine sectors. On the right, a series of very frequent measurements leads to the quantum Zeno effect.

The quantum Zeno effect (also known as the Turing paradox) is a feature of quantum-mechanical systems allowing a particle's time evolution to be slowed down by measuring it frequently enough with respect to some chosen measurement setting. [1]

Contents

Sometimes this effect is interpreted as "a system cannot change while you are watching it". [2] One can "freeze" the evolution of the system by measuring it frequently enough in its known initial state. The meaning of the term has since expanded, leading to a more technical definition, in which time evolution can be suppressed not only by measurement: the quantum Zeno effect is the suppression of unitary time evolution in quantum systems provided by a variety of sources: measurement, interactions with the environment, stochastic fields, among other factors. [3] As an outgrowth of study of the quantum Zeno effect, it has become clear that applying a series of sufficiently strong and fast pulses with appropriate symmetry can also decouple a system from its decohering environment. [4]

The first rigorous and general derivation of the quantum Zeno effect was presented in 1974 by Degasperis, Fonda, and Ghirardi, [5] although it had previously been described by Alan Turing. [6] The comparison with Zeno's paradox is due to a 1977 article by Baidyanath Misra & E. C. George Sudarshan.The name comes by analogy to Zeno's arrow paradox, which states that because an arrow in flight is not seen to move during any single instant, it cannot possibly be moving at all. In the quantum Zeno effect an unstable state seems frozen – to not 'move' – due to a constant series of observations.

According to the reduction postulate, each measurement causes the wavefunction to collapse to an eigenstate of the measurement basis. In the context of this effect, an observation can simply be the absorption of a particle, without the need of an observer in any conventional sense. However, there is controversy over the interpretation of the effect, sometimes referred to as the "measurement problem" in traversing the interface between microscopic and macroscopic objects. [7] [8]

Another crucial problem related to the effect is strictly connected to the time–energy indeterminacy relation (part of the indeterminacy principle). If one wants to make the measurement process more and more frequent, one has to correspondingly decrease the time duration of the measurement itself. But the request that the measurement last only a very short time implies that the energy spread of the state in which reduction occurs becomes increasingly large. However, the deviations from the exponential decay law for small times is crucially related to the inverse of the energy spread, so that the region in which the deviations are appreciable shrinks when one makes the measurement process duration shorter and shorter. An explicit evaluation of these two competing requests shows that it is inappropriate, without taking into account this basic fact, to deal with the actual occurrence and emergence of Zeno's effect. [9]

Closely related (and sometimes not distinguished from the quantum Zeno effect) is the watchdog effect, in which the time evolution of a system is affected by its continuous coupling to the environment. [10] [11] [12] [13]

Description

Unstable quantum systems are predicted to exhibit a short-time deviation from the exponential decay law. [14] [15] This universal phenomenon has led to the prediction that frequent measurements during this nonexponential period could inhibit decay of the system, one form of the quantum Zeno effect. Subsequently, it was predicted that measurements applied more slowly could also enhance decay rates, a phenomenon known as the quantum anti-Zeno effect. [16]

In quantum mechanics, the interaction mentioned is called "measurement" because its result can be interpreted in terms of classical mechanics. Frequent measurement prohibits the transition. It can be a transition of a particle from one half-space to another (which could be used for an atomic mirror in an atomic nanoscope [17] ) as in the time-of-arrival problem, [18] [19] a transition of a photon in a waveguide from one mode to another, and it can be a transition of an atom from one quantum state to another. It can be a transition from the subspace without decoherent loss of a qubit to a state with a qubit lost in a quantum computer. [20] [21] In this sense, for the qubit correction, it is sufficient to determine whether the decoherence has already occurred or not. All these can be considered as applications of the Zeno effect. [22] By its nature, the effect appears only in systems with distinguishable quantum states, and hence is inapplicable to classical phenomena and macroscopic bodies.

The mathematician Robin Gandy recalled Turing's formulation of the quantum Zeno effect in a letter to fellow mathematician Max Newman, shortly after Turing's death:

[I]t is easy to show using standard theory that if a system starts in an eigenstate of some observable, and measurements are made of that observable N times a second, then, even if the state is not a stationary one, the probability that the system will be in the same state after, say, one second, tends to one as N tends to infinity; that is, that continual observations will prevent motion. Alan and I tackled one or two theoretical physicists with this, and they rather pooh-poohed it by saying that continual observation is not possible. But there is nothing in the standard books (e.g., Dirac's) to this effect, so that at least the paradox shows up an inadequacy of Quantum Theory as usually presented.

Quoted by Andrew Hodges in Mathematical Logic, R. O. Gandy and C. E. M. Yates, eds. (Elsevier, 2001), p. 267.

As a result of Turing's suggestion, the quantum Zeno effect is also sometimes known as the Turing paradox. The idea is implicit in the early work of John von Neumann on the mathematical foundations of quantum mechanics, and in particular the rule sometimes called the reduction postulate . [23] It was later shown that the quantum Zeno effect of a single system is equivalent to the indetermination of the quantum state of a single system. [24] [25] [26]

Various realizations and general definition

The treatment of the Zeno effect as a paradox is not limited to the processes of quantum decay. In general, the term Zeno effect is applied to various transitions, and sometimes these transitions may be very different from a mere "decay" (whether exponential or non-exponential).

One realization refers to the observation of an object (Zeno's arrow, or any quantum particle) as it leaves some region of space. In the 20th century, the trapping (confinement) of a particle in some region by its observation outside the region was considered as nonsensical, indicating some non-completeness of quantum mechanics. [27] Even as late as 2001, confinement by absorption was considered as a paradox. [28] Later, similar effects of the suppression of Raman scattering was considered an expected effect, [29] [30] [31] not a paradox at all. The absorption of a photon at some wavelength, the release of a photon (for example one that has escaped from some mode of a fiber), or even the relaxation of a particle as it enters some region, are all processes that can be interpreted as measurement. Such a measurement suppresses the transition, and is called the Zeno effect in the scientific literature.

In order to cover all of these phenomena (including the original effect of suppression of quantum decay), the Zeno effect can be defined as a class of phenomena in which some transition is suppressed by an interaction – one that allows the interpretation of the resulting state in the terms 'transition did not yet happen' and 'transition has already occurred', or 'The proposition that the evolution of a quantum system is halted' if the state of the system is continuously measured by a macroscopic device to check whether the system is still in its initial state. [32]

Periodic measurement of a quantum system

Consider a system in a state , which is the eigenstate of some measurement operator. Say the system under free time evolution will decay with a certain probability into state . If measurements are made periodically, with some finite interval between each one, at each measurement, the wave function collapses to an eigenstate of the measurement operator. Between the measurements, the system evolves away from this eigenstate into a superposition state of the states and . When the superposition state is measured, it will again collapse, either back into state as in the first measurement, or away into state . However, its probability of collapsing into state after a very short amount of time is proportional to , since probabilities are proportional to squared amplitudes, and amplitudes behave linearly. Thus, in the limit of a large number of short intervals, with a measurement at the end of every interval, the probability of making the transition to goes to zero.

According to decoherence theory, the collapse of the wave function is not a discrete, instantaneous event. A "measurement" is equivalent to strongly coupling the quantum system to the noisy thermal environment for a brief period of time, and continuous strong coupling is equivalent to frequent "measurement". The time it takes for the wave function to "collapse" is related to the decoherence time of the system when coupled to the environment. The stronger the coupling is, and the shorter the decoherence time, the faster it will collapse. So in the decoherence picture, a perfect implementation of the quantum Zeno effect corresponds to the limit where a quantum system is continuously coupled to the environment, and where that coupling is infinitely strong, and where the "environment" is an infinitely large source of thermal randomness.

Experiments and discussion

Experimentally, strong suppression of the evolution of a quantum system due to environmental coupling has been observed in a number of microscopic systems.

In 1989, David J. Wineland and his group at NIST [33] observed the quantum Zeno effect for a two-level atomic system that was interrogated during its evolution. Approximately 5,000 9 Be + ions were stored in a cylindrical Penning trap and laser-cooled to below 250 mK. A resonant RF pulse was applied, which, if applied alone, would cause the entire ground-state population to migrate into an excited state. After the pulse was applied, the ions were monitored for photons emitted due to relaxation. The ion trap was then regularly "measured" by applying a sequence of ultraviolet pulses during the RF pulse. As expected, the ultraviolet pulses suppressed the evolution of the system into the excited state. The results were in good agreement with theoretical models.

In 2001, Mark G. Raizen and his group at the University of Texas at Austin observed the quantum Zeno effect for an unstable quantum system, [34] as originally proposed by Sudarshan and Misra. [1] They also observed an anti-Zeno effect. Ultracold sodium atoms were trapped in an accelerating optical lattice, and the loss due to tunneling was measured. The evolution was interrupted by reducing the acceleration, thereby stopping quantum tunneling. The group observed suppression or enhancement of the decay rate, depending on the regime of measurement.

In 2015, Mukund Vengalattore and his group at Cornell University demonstrated a quantum Zeno effect as the modulation of the rate of quantum tunnelling in an ultracold lattice gas by the intensity of light used to image the atoms. [35]

The quantum Zeno effect is used in commercial atomic magnetometers and proposed to be part of birds' magnetic compass sensory mechanism (magnetoreception). [36]

It is still an open question how closely one can approach the limit of an infinite number of interrogations due to the Heisenberg uncertainty involved in shorter measurement times. It has been shown, however, that measurements performed at a finite frequency can yield arbitrarily strong Zeno effects. [37] In 2006, Streed et al. at MIT observed the dependence of the Zeno effect on measurement pulse characteristics. [38]

The interpretation of experiments in terms of the "Zeno effect" helps describe the origin of a phenomenon. Nevertheless, such an interpretation does not bring any principally new features not described with the Schrödinger equation of the quantum system. [39] [40]

Even more, the detailed description of experiments with the "Zeno effect", especially at the limit of high frequency of measurements (high efficiency of suppression of transition, or high reflectivity of a ridged mirror) usually do not behave as expected for an idealized measurement. [17]

It was shown that the quantum Zeno effect persists in the many-worlds and relative-states interpretations of quantum mechanics. [41]

See also

Related Research Articles

<span class="mw-page-title-main">Einstein–Podolsky–Rosen paradox</span> Historical critique of quantum mechanics

The Einstein–Podolsky–Rosen (EPR) paradox is a thought experiment proposed by physicists Albert Einstein, Boris Podolsky and Nathan Rosen which argues that the description of physical reality provided by quantum mechanics is incomplete. In a 1935 paper titled "Can Quantum-Mechanical Description of Physical Reality be Considered Complete?", they argued for the existence of "elements of reality" that were not part of quantum theory, and speculated that it should be possible to construct a theory containing these hidden variables. Resolutions of the paradox have important implications for the interpretation of quantum mechanics.

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

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

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

<span class="mw-page-title-main">Quantum decoherence</span> Loss of quantum coherence

Quantum decoherence is the loss of quantum coherence. Quantum decoherence has been studied to understand how quantum systems convert to systems 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.

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.

In quantum mechanics, the measurement problem is the problem of definite outcomes: quantum systems have superpositions but quantum measurements only give one definite result.

A Bell test, also known as Bell inequality test or Bell experiment, is a real-world physics experiment designed to test the theory of quantum mechanics in relation to Albert Einstein's concept of local realism. Named for John Stewart Bell, the experiments test whether or not the real world satisfies local realism, which requires the presence of some additional local variables to explain the behavior of particles like photons and electrons. As of 2015, all Bell tests have found that the hypothesis of local hidden variables is inconsistent with the way that physical systems behave.

In the interpretation of quantum mechanics, a local hidden-variable theory is a hidden-variable theory that satisfies the principle of locality. These models attempt to account for the probabilistic features of quantum mechanics via the mechanism of underlying, but inaccessible variables, with the additional requirement that distant events be statistically independent.

In quantum mechanics, einselections, short for "environment-induced superselection", is a name coined by Wojciech H. Zurek for a process which is claimed to explain the appearance of wavefunction collapse and the emergence of classical descriptions of reality from quantum descriptions. In this approach, classicality is described as an emergent property induced in open quantum systems by their environments. Due to the interaction with the environment, the vast majority of states in the Hilbert space of a quantum open system become highly unstable due to entangling interaction with the environment, which in effect monitors selected observables of the system. After a decoherence time, which for macroscopic objects is typically many orders of magnitude shorter than any other dynamical timescale, a generic quantum state decays into an uncertain state which can be expressed as a mixture of simple pointer states. In this way the environment induces effective superselection rules. Thus, einselection precludes stable existence of pure superpositions of pointer states. These 'pointer states' are stable despite environmental interaction. The einselected states lack coherence, and therefore do not exhibit the quantum behaviours of entanglement and superposition.

Quantum Darwinism is a theory meant to explain the emergence of the classical world from the quantum world as due to a process of Darwinian natural selection induced by the environment interacting with the quantum system; where the many possible quantum states are selected against in favor of a stable pointer state. It was proposed in 2003 by Wojciech Zurek and a group of collaborators including Ollivier, Poulin, Paz and Blume-Kohout. The development of the theory is due to the integration of a number of Zurek's research topics pursued over the course of 25 years, including pointer states, einselection and decoherence.

<span class="mw-page-title-main">Trapped-ion quantum computer</span> Proposed quantum computer implementation

A trapped-ion quantum computer is one proposed approach to a large-scale quantum computer. Ions, or charged atomic particles, can be confined and suspended in free space using electromagnetic fields. Qubits are stored in stable electronic states of each ion, and quantum information can be transferred through the collective quantized motion of the ions in a shared trap. Lasers are applied to induce coupling between the qubit states or coupling between the internal qubit states and the external motional states.

Quantum cloning is a process that takes an arbitrary, unknown quantum state and makes an exact copy without altering the original state in any way. Quantum cloning is forbidden by the laws of quantum mechanics as shown by the no cloning theorem, which states that there is no operation for cloning any arbitrary state perfectly. In Dirac notation, the process of quantum cloning is described by:

Objective-collapse theories, also known spontaneous collapse models or dynamical reduction models, are proposed solutions to the measurement problem in quantum mechanics. As with other interpretations of quantum mechanics, they are possible explanations of why and how quantum measurements always give definite outcomes, not a superposition of them as predicted by the Schrödinger equation, and more generally how the classical world emerges from quantum theory. The fundamental idea is that the unitary evolution of the wave function describing the state of a quantum system is approximate. It works well for microscopic systems, but progressively loses its validity when the mass / complexity of the system increases.

In physics, the observer effect is the disturbance of an observed system by the act of observation. This is often the result of utilizing instruments that, by necessity, alter the state of what they measure in some manner. A common example is checking the pressure in an automobile tire, which causes some of the air to escape, thereby changing the pressure to observe it. Similarly, seeing non-luminous objects requires light hitting the object to cause it to reflect that light. While the effects of observation are often negligible, the object still experiences a change. This effect can be found in many domains of physics, but can usually be reduced to insignificance by using different instruments or observation techniques.

In quantum information theory, quantum discord is a measure of nonclassical correlations between two subsystems of a quantum system. It includes correlations that are due to quantum physical effects but do not necessarily involve quantum entanglement.

In quantum mechanics, the cat state, named after Schrödinger's cat, is a quantum state composed of two diametrically opposed conditions at the same time, such as the possibilities that a cat is alive and dead at the same time.

In quantum mechanics, weak measurements are a type of quantum measurement that results in an observer obtaining very little information about the system on average, but also disturbs the state very little. From Busch's theorem the system is necessarily disturbed by the measurement. In the literature weak measurements are also known as unsharp, fuzzy, dull, noisy, approximate, and gentle measurements. Additionally weak measurements are often confused with the distinct but related concept of the weak value.

Quantum foundations is a discipline of science that seeks to understand the most counter-intuitive aspects of quantum theory, reformulate it and even propose new generalizations thereof. Contrary to other physical theories, such as general relativity, the defining axioms of quantum theory are quite ad hoc, with no obvious physical intuition. While they lead to the right experimental predictions, they do not come with a mental picture of the world where they fit.

A generalized probabilistic theory (GPT) is a general framework to describe the operational features of arbitrary physical theories. A GPT must specify what kind of physical systems one can find in the lab, as well as rules to compute the outcome statistics of any experiment involving labeled preparations, transformations and measurements. The framework of GPTs has been used to define hypothetical non-quantum physical theories which nonetheless possess quantum theory's most remarkable features, such as entanglement or teleportation. Notably, a small set of physically motivated axioms is enough to single out the GPT representation of quantum theory.

References

  1. 1 2 Sudarshan, E. C. G.; Misra, B. (1977). "The Zeno's paradox in quantum theory". Journal of Mathematical Physics . 18 (4): 756–763. Bibcode:1977JMP....18..756M. doi:10.1063/1.523304. OSTI   7342282.
  2. https://phys.org/news/2015-10-zeno-effect-verifiedatoms-wont.html. Archived 2018-09-25 at the Wayback Machine
  3. Nakanishi, T.; Yamane, K.; Kitano, M. (2001). "Absorption-free optical control of spin systems: the quantum Zeno effect in optical pumping". Physical Review A . 65 (1): 013404. arXiv: quant-ph/0103034 . Bibcode:2001PhRvA..65a3404N. doi:10.1103/PhysRevA.65.013404. S2CID   56052019.
  4. Facchi, P.; Lidar, D. A.; Pascazio, S. (2004). "Unification of dynamical decoupling and the quantum Zeno effect". Physical Review A . 69 (3): 032314. arXiv: quant-ph/0303132 . Bibcode:2004PhRvA..69c2314F. doi:10.1103/PhysRevA.69.032314. S2CID   38253718.
  5. Degasperis, A.; Fonda, L.; Ghirardi, G. C. (1974). "Does the lifetime of an unstable system depend on the measuring apparatus?". Il Nuovo Cimento A . 21 (3): 471–484. Bibcode:1974NCimA..21..471D. doi:10.1007/BF02731351. S2CID   120279111.
  6. Hofstadter, D. (2004). Teuscher, C. (ed.). Alan Turing: Life and Legacy of a Great Thinker. Springer. p. 54. ISBN   978-3-540-20020-8.
  7. Greenstein, G.; Zajonc, A. (2005). The Quantum Challenge: Modern Research on the Foundations of Quantum Mechanics. Jones & Bartlett Publishers. p. 237. ISBN   978-0-7637-2470-2.
  8. Facchi, P.; Pascazio, S. (2002). "Quantum Zeno subspaces". Physical Review Letters . 89 (8): 080401. arXiv: quant-ph/0201115 . Bibcode:2002PhRvL..89h0401F. doi:10.1103/PhysRevLett.89.080401. PMID   12190448. S2CID   29178016.
  9. Ghirardi, G. C.; Omero, C.; Rimini, A.; Weber, T. (1979). "Small Time Behaviour of Quantum Nondecay Probability and Zeno's Paradox in Quantum Mechanics". Il Nuovo Cimento A . 52 (4): 421. Bibcode:1979NCimA..52..421G. doi:10.1007/BF02770851. S2CID   124911216.
  10. Kraus, K. (1981-08-01). "Measuring processes in quantum mechanics I. Continuous observation and the watchdog effect". Foundations of Physics. 11 (7–8): 547–576. Bibcode:1981FoPh...11..547K. doi:10.1007/bf00726936. ISSN   0015-9018. S2CID   121902392.
  11. Belavkin, V.; Staszewski, P. (1992). "Nondemolition observation of a free quantum particle". Phys. Rev. A. 45 (3): 1347–1356. arXiv: quant-ph/0512138 . Bibcode:1992PhRvA..45.1347B. doi:10.1103/PhysRevA.45.1347. PMID   9907114. S2CID   14637898.
  12. Ghose, P. (1999). Testing Quantum Mechanics on New Ground. Cambridge University Press. p. 114. ISBN   978-0-521-02659-8.
  13. Auletta, G. (2000). Foundations and Interpretation of Quantum Mechanics. World Scientific. p. 341. ISBN   978-981-02-4614-3.
  14. Khalfin, L. A. (1958). "Contribution to the decay theory of a quasi-stationary state". Soviet Physics JETP . 6: 1053. Bibcode:1958JETP....6.1053K. OSTI   4318804.
  15. Raizen, M. G.; Wilkinson, S. R.; Bharucha, C. F.; Fischer, M. C.; Madison, K. W.; Morrow, P. R.; Niu, Q.; Sundaram, B. (1997). "Experimental evidence for non-exponential decay in quantum tunnelling" (PDF). Nature . 387 (6633): 575. Bibcode:1997Natur.387..575W. doi:10.1038/42418. S2CID   4246387. Archived from the original (PDF) on 2010-03-31.
  16. Chaudhry, Adam Zaman (2016-07-13). "A general framework for the Quantum Zeno and anti-Zeno effects". Scientific Reports . 6 (1): 29497. arXiv: 1604.06561 . Bibcode:2016NatSR...629497C. doi:10.1038/srep29497. ISSN   2045-2322. PMC   4942788 . PMID   27405268.
  17. 1 2 Kouznetsov, D.; Oberst, H.; Neumann, A.; Kuznetsova, Y.; Shimizu, K.; Bisson, J.-F.; Ueda, K.; Brueck, S. R. J. (2006). "Ridged atomic mirrors and atomic nanoscope". Journal of Physics B . 39 (7): 1605–1623. Bibcode:2006JPhB...39.1605K. doi:10.1088/0953-4075/39/7/005. S2CID   16653364.
  18. Allcock, J. (1969). "The time of arrival in quantum mechanics I. Formal considerations". Annals of Physics . 53 (2): 253–285. Bibcode:1969AnPhy..53..253A. doi:10.1016/0003-4916(69)90251-6.
  19. Echanobe, J.; Del Campo, A.; Muga, J. G. (2008). "Disclosing hidden information in the quantum Zeno effect: Pulsed measurement of the quantum time of arrival". Physical Review A . 77 (3): 032112. arXiv: 0712.0670 . Bibcode:2008PhRvA..77c2112E. doi:10.1103/PhysRevA.77.032112. S2CID   118335567.
  20. Stolze, J.; Suter, D. (2008). Quantum computing: a short course from theory to experiment (2nd ed.). Wiley-VCH. p. 99. ISBN   978-3-527-40787-3.[ permanent dead link ]
  21. "Quantum computer solves problem, without running". Phys.Org. 22 February 2006. Retrieved 2013-09-21.
  22. Franson, J.; Jacobs, B.; Pittman, T. (2006). "Quantum computing using single photons and the Zeno effect". Physical Review A . 70 (6): 062302. arXiv: quant-ph/0408097 . Bibcode:2004PhRvA..70f2302F. doi:10.1103/PhysRevA.70.062302. S2CID   119071343.
  23. von Neumann, J. (1932). Mathematische Grundlagen der Quantenmechanik. Springer. Chapter V.2. ISBN   978-3-540-59207-5. See also von Neumann, J. (1955). Mathematical Foundations of Quantum Mechanics . Princeton University Press. p.  366. ISBN   978-0-691-02893-4.); Menskey, M. B. (2000). Quantum Measurements and Decoherence. Springer. §4.1.1, pp. 315 ff. ISBN   978-0-7923-6227-2.; Wunderlich, C.; Balzer, C. (2003). Bederson, B.; Walther, H. (eds.). Quantum Measurements and New Concepts for Experiments with Trapped Ions. Advances in Atomic, Molecular, and Optical Physics. Vol. 49. Academic Press. p. 315. ISBN   978-0-12-003849-7.
  24. Orly Alter and Yoshihisa Yamamoto (April 1997). "Quantum Zeno Effect and the Impossibility of Determining the Quantum State of a Single System". Phys. Rev. A. 55 (5): R2499–R2502. Bibcode:1997PhRvA..55.2499A. doi:10.1103/PhysRevA.55.R2499.
  25. Orly Alter and Yoshihisa Yamamoto (October 1996). "The quantum Zeno effect of a single system is equivalent to the indetermination of the quantum state of a single system" (PDF). In F. De Martini, G. Denardo and Y. Shih (ed.). Quantum Interferometry. Wiley-VCH. pp. 539–544.
  26. Orly Alter and Yoshihisa Yamamoto (2001). Quantum Measurement of a Single System (PDF). Wiley-Interscience. doi:10.1002/9783527617128. ISBN   9780471283089. Archived from the original (PDF) on 2021-12-04. Retrieved 2021-12-04.
  27. Mielnik, B. (1994). "The screen problem". Foundations of Physics . 24 (8): 1113–1129. Bibcode:1994FoPh...24.1113M. doi:10.1007/BF02057859. S2CID   121708226.
  28. Yamane, K.; Ito, M.; Kitano, M. (2001). "Quantum Zeno effect in optical fibers". Optics Communications . 192 (3–6): 299–307. Bibcode:2001OptCo.192..299Y. doi:10.1016/S0030-4018(01)01192-0.
  29. Thun, K.; Peřina, J.; Křepelka, J. (2002). "Quantum Zeno effect in Raman scattering". Physics Letters A . 299 (1): 19–30. Bibcode:2002PhLA..299...19T. doi:10.1016/S0375-9601(02)00629-1.
  30. Peřina, J. (2004). "Quantum Zeno effect in cascaded parametric down-conversion with losses". Physics Letters A . 325 (1): 16–20. Bibcode:2004PhLA..325...16P. doi:10.1016/j.physleta.2004.03.026.
  31. Kouznetsov, D.; Oberst, H. (2005). "Reflection of Waves from a Ridged Surface and the Zeno Effect". Optical Review . 12 (5): 1605–1623. Bibcode:2005OptRv..12..363K. doi:10.1007/s10043-005-0363-9. S2CID   55565166.
  32. Panov, A. D. (2001). "Quantum Zeno effect in spontaneous decay with distant detector". Physics Letters A . 281 (1): 9. arXiv: quant-ph/0101031 . Bibcode:2001PhLA..281....9P. doi:10.1016/S0375-9601(01)00094-9. S2CID   18357530.
  33. Itano, W.; Heinzen, D.; Bollinger, J.; Wineland, D. (1990). "Quantum Zeno effect" (PDF). Physical Review A . 41 (5): 2295–2300. Bibcode:1990PhRvA..41.2295I. doi:10.1103/PhysRevA.41.2295. PMID   9903355. Archived from the original (PDF) on 2004-07-20.
  34. Fischer, M.; Gutiérrez-Medina, B.; Raizen, M. (2001). "Observation of the Quantum Zeno and Anti-Zeno Effects in an Unstable System". Physical Review Letters . 87 (4): 040402. arXiv: quant-ph/0104035 . Bibcode:2001PhRvL..87d0402F. doi:10.1103/PhysRevLett.87.040402. PMID   11461604. S2CID   11178428.
  35. Patil, Y. S.; Chakram, S.; Vengalattore, M. (2015). "Measurement-Induced Localization of an Ultracold Lattice Gas". Physical Review Letters. 115 (14): 140402. arXiv: 1411.2678 . Bibcode:2015PhRvL.115n0402P. doi:10.1103/PhysRevLett.115.140402. ISSN   0031-9007. PMID   26551797.
  36. Kominis, I. K. (2009). "Quantum Zeno effect explains magnetic-sensitive radical-ion-pair reactions". Phys. Rev. E. 80 (5): 056115. arXiv: 0806.0739 . Bibcode:2009PhRvE..80e6115K. doi:10.1103/PhysRevE.80.056115. PMID   20365051. S2CID   9848948.
  37. Layden, D.; Martin-Martinez, E.; Kempf, A. (2015). "Perfect Zeno-like effect through imperfect measurements at a finite frequency". Physical Review A. 91 (2): 022106. arXiv: 1410.3826 . Bibcode:2015PhRvA..91b2106L. doi:10.1103/PhysRevA.91.022106. S2CID   119628035.
  38. Streed, E.; Mun, J.; Boyd, M.; Campbell, G.; Medley, P.; Ketterle, W.; Pritchard, D. (2006). "Continuous and Pulsed Quantum Zeno Effect". Physical Review Letters . 97 (26): 260402. arXiv: cond-mat/0606430 . Bibcode:2006PhRvL..97z0402S. doi:10.1103/PhysRevLett.97.260402. PMID   17280408. S2CID   2414199.
  39. Petrosky, T.; Tasaki, S.; Prigogine, I. (1990). "Quantum zeno effect". Physics Letters A . 151 (3–4): 109. Bibcode:1990PhLA..151..109P. doi:10.1016/0375-9601(90)90173-L.
  40. Petrosky, T.; Tasaki, S.; Prigogine, I. (1991). "Quantum Zeno effect". Physica A . 170 (2): 306. Bibcode:1991PhyA..170..306P. doi:10.1016/0378-4371(91)90048-H.
  41. Home, D.; Whitaker, M. A. B. (1987). "The many-worlds and relative states interpretations of quantum mechanics, and the quantum Zeno paradox". Journal of Physics A . 20 (11): 3339–3345. Bibcode:1987JPhA...20.3339H. doi:10.1088/0305-4470/20/11/036.

Further reading