ART

In quantum field theory, the Casimir effect and the Casimir–Polder force are physical forces arising from a quantized field. They are named after the Dutch physicist Hendrik Casimir, who predicted them in 1948. It was not until 1997 that a direct experiment by S. Lamoreaux quantitatively measured the force to within 5% of the value predicted by the theory.[1]

The Casimir effect can be understood by the idea that the presence of conducting metals and dielectrics alters the vacuum expectation value of the energy of the second-quantized electromagnetic field.[2][3] Since the value of this energy depends on the shapes and positions of the conductors and dielectrics, the Casimir effect manifests itself as a force between such objects.

Any medium supporting oscillations has an analogue of the Casimir effect. For example, beads on a string[4][5] as well as plates submerged in turbulent water[6] or gas[7] illustrate the Casimir force.

In modern theoretical physics, the Casimir effect plays an important role in the chiral bag model of the nucleon; in applied physics it is significant in some aspects of emerging microtechnologies and nanotechnologies.[8]

Physical properties

The typical example is of the two uncharged conductive plates in a vacuum, placed a few nanometers apart. In a classical description, the lack of an external field means that there is no field between the plates, and no force would be measured between them.[9] When this field is instead studied using the quantum electrodynamic vacuum, it is seen that the plates do affect the virtual photons which constitute the field, and generate a net force[10] – either an attraction or a repulsion depending on the specific arrangement of the two plates. Although the Casimir effect can be expressed in terms of virtual particles interacting with the objects, it is best described and more easily calculated in terms of the zero-point energy of a quantized field in the intervening space between the objects. This force has been measured and is a striking example of an effect captured formally by second quantization.[11][12]

The treatment of boundary conditions in these calculations has led to some controversy. In fact, "Casimir's original goal was to compute the van der Waals force between polarizable molecules" of the conductive plates. Thus it can be interpreted without any reference to the zero-point energy (vacuum energy) of quantum fields.[13]

Because the strength of the force falls off rapidly with distance, it is measurable only when the distance between the objects is extremely small. On a submicron scale, this force becomes so strong that it becomes the dominant force between uncharged conductors. In fact, at separations of 10 nm – about 100 times the typical size of an atom – the Casimir effect produces the equivalent of about 1 atmosphere of pressure (the precise value depending on surface geometry and other factors).[11]
History

Dutch physicists Hendrik Casimir and Dirk Polder at Philips Research Labs proposed the existence of a force between two polarizable atoms and between such an atom and a conducting plate in 1947;[14] this special form is called the Casimir–Polder force. After a conversation with Niels Bohr, who suggested it had something to do with zero-point energy, Casimir alone formulated the theory predicting a force between neutral conducting plates in 1948.[15] This latter phenomenon is called the Casimir effect in the narrow sense.

Predictions of the force were later extended to finite-conductivity metals and dielectrics, and recent calculations have considered more general geometries. Experiments before 1997 had observed the force qualitatively, and indirect validation of the predicted Casimir energy had been made by measuring the thickness of liquid helium films. However it was not until 1997 that a direct experiment by S. Lamoreaux quantitatively measured the force to within 5% of the value predicted by the theory.[1] Subsequent experiments approach an accuracy of a few percent.
Main article: Vacuum energy

The causes of the Casimir effect are described by quantum field theory, which states that all of the various fundamental fields, such as the electromagnetic field, must be quantized at each and every point in space. In a simplified view, a "field" in physics may be envisioned as if space were filled with interconnected vibrating balls and springs, and the strength of the field can be visualized as the displacement of a ball from its rest position. Vibrations in this field propagate and are governed by the appropriate wave equation for the particular field in question. The second quantization of quantum field theory requires that each such ball-spring combination be quantized, that is, that the strength of the field be quantized at each point in space. At the most basic level, the field at each point in space is a simple harmonic oscillator, and its quantization places a quantum harmonic oscillator at each point. Excitations of the field correspond to the elementary particles of particle physics. However, even the vacuum has a vastly complex structure, so all calculations of quantum field theory must be made in relation to this model of the vacuum.

The vacuum has, implicitly, all of the properties that a particle may have: spin,[16] or polarization in the case of light, energy, and so on. On average, most of these properties cancel out: the vacuum is, after all, "empty" in this sense. One important exception is the vacuum energy or the vacuum expectation value of the energy. The quantization of a simple harmonic oscillator states that the lowest possible energy or zero-point energy that such an oscillator may have is

\( {\displaystyle {E}={\begin{matrix}{\frac {1}{2}}\end{matrix}}\hbar \omega \ .} \)

Summing over all possible oscillators at all points in space gives an infinite quantity. Since only differences in energy are physically measurable (with the notable exception of gravitation, which remains beyond the scope of quantum field theory), this infinity may be considered a feature of the mathematics rather than of the physics. This argument is the underpinning of the theory of renormalization. Dealing with infinite quantities in this way was a cause of widespread unease among quantum field theorists before the development in the 1970s of the renormalization group, a mathematical formalism for scale transformations that provides a natural basis for the process.

When the scope of the physics is widened to include gravity, the interpretation of this formally infinite quantity remains problematic. There is currently no compelling explanation as to why it should not result in a cosmological constant that is many orders of magnitude larger than observed.[17] However, since we do not yet have any fully coherent quantum theory of gravity, there is likewise no compelling reason as to why it should instead actually result in the value of the cosmological constant that we observe.[18]

The Casimir effect for fermions can be understood as the spectral asymmetry of the fermion operator \( {\displaystyle (-1)^{F}} \), where it is known as the Witten index.
Relativistic van der Waals force

Alternatively, a 2005 paper by Robert Jaffe of MIT states that "Casimir effects can be formulated and Casimir forces can be computed without reference to zero-point energies. They are relativistic, quantum forces between charges and currents. The Casimir force (per unit area) between parallel plates vanishes as alpha, the fine structure constant, goes to zero, and the standard result, which appears to be independent of alpha, corresponds to the alpha approaching infinity limit," and that "The Casimir force is simply the (relativistic, retarded) van der Waals force between the metal plates."[13] Casimir and Polder's original paper used this method to derive the Casimir–Polder force. In 1978, Schwinger, DeRadd, and Milton published a similar derivation for the Casimir effect between two parallel plates.[19] In fact, the description in terms of van der Waals forces is the only correct description from the fundamental microscopic perspective,[20][21] while other descriptions of Casimir force are merely effective macroscopic descriptions.
Effects

Casimir's observation was that the second-quantized quantum electromagnetic field, in the presence of bulk bodies such as metals or dielectrics, must obey the same boundary conditions that the classical electromagnetic field must obey. In particular, this affects the calculation of the vacuum energy in the presence of a conductor or dielectric.

Consider, for example, the calculation of the vacuum expectation value of the electromagnetic field inside a metal cavity, such as, for example, a radar cavity or a microwave waveguide. In this case, the correct way to find the zero-point energy of the field is to sum the energies of the standing waves of the cavity. To each and every possible standing wave corresponds an energy; say the energy of the nth standing wave is \( E_{n} \). The vacuum expectation value of the energy of the electromagnetic field in the cavity is then

\( {\displaystyle \langle E\rangle ={\frac {1}{2}}\sum _{n}E_{n}} \)

with the sum running over all possible values of n enumerating the standing waves. The factor of 1/2 is present because the zero-point energy of the n'th mode is \( {\displaystyle E_{n}/2} \), where \( E_{n} \) is the energy increment for the n'th mode. (It is the same 1/2 as appears in the equation \( E=\hbar \omega/2.) \)Written in this way, this sum is clearly divergent; however, it can be used to create finite expressions.

In particular, one may ask how the zero-point energy depends on the shape s of the cavity. Each energy level \( E_{n} \) depends on the shape, and so one should write \( E_n(s) \) for the energy level, and\( \langle E(s) \rangle \) for the vacuum expectation value. At this point comes an important observation: the force at point p on the wall of the cavity is equal to the change in the vacuum energy if the shape s of the wall is perturbed a little bit, say by δ s {\displaystyle \delta s} \delta s, at point p. That is, one has

\( {\displaystyle F(p)=-\left.{\frac {\delta \langle E(s)\rangle }{\delta s}}\right\vert _{p}.\,} \)

This value is finite in many practical calculations.[22]

Attraction between the plates can be easily understood by focusing on the one-dimensional situation. Suppose that a moveable conductive plate is positioned at a short distance a from one of two widely separated plates (distance L apart). With a << L, the states within the slot of width a are highly constrained so that the energy E of any one mode is widely separated from that of the next. This is not the case in the large region L, where there is a large number (numbering about L/a) of states with energy evenly spaced between E and the next mode in the narrow slot – in other words, all slightly larger than E. Now on shortening a by da (< 0), the mode in the narrow slot shrinks in wavelength and therefore increases in energy proportional to −da/a, whereas all the L/a states that lie in the large region lengthen and correspondingly decrease their energy by an amount proportional to da/L (note the denominator). The two effects nearly cancel, but the net change is slightly negative, because the energy of all the L/a modes in the large region are slightly larger than the single mode in the slot. Thus the force is attractive: it tends to make a slightly smaller, the plates attracting each other across the thin slot.
Derivation of Casimir effect assuming zeta-regularization

See Wikiversity for an elementary calculation in one dimension.

In the original calculation done by Casimir, he considered the space between a pair of conducting metal plates at distance a {\displaystyle a} a apart. In this case, the standing waves are particularly easy to calculate, because the transverse component of the electric field and the normal component of the magnetic field must vanish on the surface of a conductor. Assuming the plates lie parallel to the xy-plane, the standing waves are

\( {\displaystyle \psi _{n}(x,y,z;t)=e^{-i\omega _{n}t}e^{ik_{x}x+ik_{y}y}\sin(k_{n}z)} \)

where \( \psi \) stands for the electric component of the electromagnetic field, and, for brevity, the polarization and the magnetic components are ignored here. Here, \( k_x \) and \( k_{y} \) are the wave numbers in directions parallel to the plates, and

\( {\displaystyle k_{n}={\frac {n\pi }{a}}} \)

is the wave-number perpendicular to the plates. Here, n is an integer, resulting from the requirement that ψ vanish on the metal plates. The frequency of this wave is

\( {\displaystyle \omega _{n}=c{\sqrt {{k_{x}}^{2}+{k_{y}}^{2}+{\frac {n^{2}\pi ^{2}}{a^{2}}}}}} \)

where c is the speed of light. The vacuum energy is then the sum over all possible excitation modes. Since the area of the plates is large, we may sum by integrating over two of the dimensions in k-space. The assumption of periodic boundary conditions yields,

\( {\displaystyle \langle E\rangle ={\frac {\hbar }{2}}\cdot 2\int {\frac {Adk_{x}dk_{y}}{(2\pi )^{2}}}\sum _{n=1}^{\infty }\omega _{n}} \)

where A is the area of the metal plates, and a factor of 2 is introduced for the two possible polarizations of the wave. This expression is clearly infinite, and to proceed with the calculation, it is convenient to introduce a regulator (discussed in greater detail below). The regulator will serve to make the expression finite, and in the end will be removed. The zeta-regulated version of the energy per unit-area of the plate is

\( {\displaystyle {\frac {\langle E(s)\rangle }{A}} =\hbar \int {\frac {dk_{x}dk_{y}}{(2\pi )^{2}}}\sum _{n=1}^{\infty }\omega _{n}\vert \omega _{n}\vert ^{-s}.} \)

In the end, the limit s → 0 {\displaystyle s\to 0} s\to 0 is to be taken. Here s is just a complex number, not to be confused with the shape discussed previously. This integral/sum is finite for s real and larger than 3. The sum has a pole at s=3, but may be analytically continued to s=0, where the expression is finite. The above expression simplifies to:

\( {\displaystyle {\frac {\langle E(s)\rangle }{A}}= {\frac {\hbar c^{1-s}}{4\pi ^{2}}}\sum _{n}\int _{0}^{\infty }2\pi qdq\left\vert q^{2}+{\frac {\pi ^{2}n^{2}}{a^{2}}}\right\vert ^{(1-s)/2},} \)

where polar coordinates \( {\displaystyle q^{2}=k_{x}^{2}+k_{y}^{2}} \) were introduced to turn the double integral into a single integral. The q {\displaystyle q} q in front is the Jacobian, and the \( 2\pi \) comes from the angular integration. The integral converges if Re[s] > 3, resulting in

⟨\( {\displaystyle {\frac {\langle E(s)\rangle }{A}}=-{\frac {\hbar c^{1-s}\pi ^{2-s}}{2a^{3-s}}}{\frac {1}{3-s}}\sum _{n}\vert n\vert ^{3-s}=-{\frac {\hbar c^{1-s}\pi ^{2-s}}{2a^{3-s}(3-s)}}\sum _{n}{\frac {1}{\left|n\right|^{s-3}}}.} \)

The sum diverges at s in the neighborhood of zero, but if the damping of large-frequency excitations corresponding to analytic continuation of the Riemann zeta function to s=0 is assumed to make sense physically in some way, then one has

\( {\displaystyle {\frac {\langle E\rangle }{A}}=\lim _{s\to 0}{\frac {\langle E(s)\rangle }{A}}=-{\frac {\hbar c\pi ^{2}}{6a^{3}}}\zeta (-3).} \)

But

\( {\displaystyle \zeta (-3)={\frac {1}{120}}} \)

and so one obtains

\( {\displaystyle {\frac {\langle E\rangle }{A}}={\frac {-\hbar c\pi ^{2}}{720a^{3}}}.} \)

The analytic continuation has evidently lost an additive positive infinity, somehow exactly accounting for the zero-point energy (not included above) outside the slot between the plates, but which changes upon plate movement within a closed system. The Casimir force per unit area \( F_c / A \) for idealized, perfectly conducting plates with vacuum between them is

\( {\displaystyle {F_{c} \over A}=-{\frac {d}{da}}{\frac {\langle E\rangle }{A}}=-{\frac {\hbar c\pi ^{2}}{240a^{4}}}} \)

where

\( \hbar \) (hbar, ħ) is the reduced Planck constant,
c is the speed of light,
a is the distance between the two plates

The force is negative, indicating that the force is attractive: by moving the two plates closer together, the energy is lowered. The presence of ℏ {\displaystyle \hbar } \hbar shows that the Casimir force per unit area \( F_c / A \) is very small, and that furthermore, the force is inherently of quantum-mechanical origin.

By integrating the equation above it is possible to calculate the energy required to separate to infinity the two plates as:

\( {\displaystyle U_{E}(a)=\int F(a)\,da=\int -\hbar c\pi ^{2}{\frac {A}{240a^{4}}}\,da} \)

\( {\displaystyle U_{E}(a)=-\hbar c\pi ^{2}{\frac {A}{720a^{3}}}} \)

where

\( \hbar \) (hbar, ħ) is the reduced Planck constant,
c is the speed of light,
A is the area of one of the plates,
a is the distance between the two plates

In Casimir's original derivation,[15] a moveable conductive plate is positioned at a short distance a from one of two widely separated plates (distance L apart). The 0-point energy on both sides of the plate is considered. Instead of the above ad hoc analytic continuation assumption, non-convergent sums and integrals are computed using Euler–Maclaurin summation with a regularizing function (e.g., exponential regularization) not so anomalous as \( \vert\omega_n\vert^{-s} \) in the above.[23]
More recent theory

Casimir's analysis of idealized metal plates was generalized to arbitrary dielectric and realistic metal plates by Lifshitz and his students.[24][25] Using this approach, complications of the bounding surfaces, such as the modifications to the Casimir force due to finite conductivity, can be calculated numerically using the tabulated complex dielectric functions of the bounding materials. Lifshitz's theory for two metal plates reduces to Casimir's idealized 1/a4 force law for large separations a much greater than the skin depth of the metal, and conversely reduces to the 1/a3 force law of the London dispersion force (with a coefficient called a Hamaker constant) for small a, with a more complicated dependence on a for intermediate separations determined by the dispersion of the materials.[26]

Lifshitz's result was subsequently generalized to arbitrary multilayer planar geometries as well as to anisotropic and magnetic materials, but for several decades the calculation of Casimir forces for non-planar geometries remained limited to a few idealized cases admitting analytical solutions.[27] For example, the force in the experimental sphere–plate geometry was computed with an approximation (due to Derjaguin) that the sphere radius R is much larger than the separation a, in which case the nearby surfaces are nearly parallel and the parallel-plate result can be adapted to obtain an approximate R/a3 force (neglecting both skin-depth and higher-order curvature effects).[27][28] However, in the 2000s a number of authors developed and demonstrated a variety of numerical techniques, in many cases adapted from classical computational electromagnetics, that are capable of accurately calculating Casimir forces for arbitrary geometries and materials, from simple finite-size effects of finite plates to more complicated phenomena arising for patterned surfaces or objects of various shapes.[27][29]

Measurement

One of the first experimental tests was conducted by Marcus Sparnaay at Philips in Eindhoven (Netherlands), in 1958, in a delicate and difficult experiment with parallel plates, obtaining results not in contradiction with the Casimir theory,[30][31] but with large experimental errors.

The Casimir effect was measured more accurately in 1997 by Steve K. Lamoreaux of Los Alamos National Laboratory,[1] and by Umar Mohideen and Anushree Roy of the University of California, Riverside.[32] In practice, rather than using two parallel plates, which would require phenomenally accurate alignment to ensure they were parallel, the experiments use one plate that is flat and another plate that is a part of a sphere with a very large radius.

In 2001, a group (Giacomo Bressi, Gianni Carugno, Roberto Onofrio and Giuseppe Ruoso) at the University of Padua (Italy) finally succeeded in measuring the Casimir force between parallel plates using microresonators.[33]

In 2013, a conglomerate of scientists from Hong Kong University of Science and Technology, University of Florida, Harvard University, Massachusetts Institute of Technology, and Oak Ridge National Laboratory demonstrated a compact integrated silicon chip that can measure the Casimir force.[34]
Regularization

In order to be able to perform calculations in the general case, it is convenient to introduce a regulator in the summations. This is an artificial device, used to make the sums finite so that they can be more easily manipulated, followed by the taking of a limit so as to remove the regulator.

The heat kernel or exponentially regulated sum is
\( {\displaystyle \langle E(t)\rangle ={\frac {1}{2}}\sum _{n}\hbar |\omega _{n}|\exp(-t|\omega _{n}|)} \)

where the limit \( t\to 0^+ \) is taken in the end. The divergence of the sum is typically manifested as

\( {\displaystyle \langle E(t)\rangle ={\frac {C}{t^{3}}}+{\textrm {finite}}\,} \)

for three-dimensional cavities. The infinite part of the sum is associated with the bulk constant C which does not depend on the shape of the cavity. The interesting part of the sum is the finite part, which is shape-dependent. The Gaussian regulator

\( {\displaystyle \langle E(t)\rangle ={\frac {1}{2}}\sum _{n}\hbar |\omega _{n}|\exp(-t^{2}|\omega _{n}|^{2})} \)

is better suited to numerical calculations because of its superior convergence properties, but is more difficult to use in theoretical calculations. Other, suitably smooth, regulators may be used as well. The zeta function regulator

\( {\displaystyle \langle E(s)\rangle ={\frac {1}{2}}\sum _{n}\hbar |\omega _{n}||\omega _{n}|^{-s}} \)

is completely unsuited for numerical calculations, but is quite useful in theoretical calculations. In particular, divergences show up as poles in the complex s plane, with the bulk divergence at s=4. This sum may be analytically continued past this pole, to obtain a finite part at s=0.

Not every cavity configuration necessarily leads to a finite part (the lack of a pole at s=0) or shape-independent infinite parts. In this case, it should be understood that additional physics has to be taken into account. In particular, at extremely large frequencies (above the plasma frequency), metals become transparent to photons (such as X-rays), and dielectrics show a frequency-dependent cutoff as well. This frequency dependence acts as a natural regulator. There are a variety of bulk effects in solid state physics, mathematically very similar to the Casimir effect, where the cutoff frequency comes into explicit play to keep expressions finite. (These are discussed in greater detail in Landau and Lifshitz, "Theory of Continuous Media".)
Generalities

The Casimir effect can also be computed using the mathematical mechanisms of functional integrals of quantum field theory, although such calculations are considerably more abstract, and thus difficult to comprehend. In addition, they can be carried out only for the simplest of geometries. However, the formalism of quantum field theory makes it clear that the vacuum expectation value summations are in a certain sense summations over so-called "virtual particles".

More interesting is the understanding that the sums over the energies of standing waves should be formally understood as sums over the eigenvalues of a Hamiltonian. This allows atomic and molecular effects, such as the van der Waals force, to be understood as a variation on the theme of the Casimir effect. Thus one considers the Hamiltonian of a system as a function of the arrangement of objects, such as atoms, in configuration space. The change in the zero-point energy as a function of changes of the configuration can be understood to result in forces acting between the objects.

In the chiral bag model of the nucleon, the Casimir energy plays an important role in showing the mass of the nucleon is independent of the bag radius. In addition, the spectral asymmetry is interpreted as a non-zero vacuum expectation value of the baryon number, cancelling the topological winding number of the pion field surrounding the nucleon.

A "pseudo-Casimir" effect can be found in liquid crystal systems, where the boundary conditions imposed through anchoring by rigid walls give rise to a long-range force, analogous to the force that arises between conducting plates.[35]
Dynamical Casimir effect

The dynamical Casimir effect is the production of particles and energy from an accelerated moving mirror. This reaction was predicted by certain numerical solutions to quantum mechanics equations made in the 1970s.[36] In May 2011 an announcement was made by researchers at the Chalmers University of Technology, in Gothenburg, Sweden, of the detection of the dynamical Casimir effect. In their experiment, microwave photons were generated out of the vacuum in a superconducting microwave resonator. These researchers used a modified SQUID to change the effective length of the resonator in time, mimicking a mirror moving at the required relativistic velocity. If confirmed this would be the first experimental verification of the dynamical Casimir effect.[37] [38] In March 2013 an article appeared on the PNAS scientific journal describing an experiment that demonstrated the dynamical Casimir effect in a Josephson metamaterial.[39]
Analogies

A similar analysis can be used to explain Hawking radiation that causes the slow "evaporation" of black holes (although this is generally visualized as the escape of one particle from a virtual particle-antiparticle pair, the other particle having been captured by the black hole).[40]

Constructed within the framework of quantum field theory in curved spacetime, the dynamical Casimir effect has been used to better understand acceleration radiation such as the Unruh effect.
Repulsive forces

There are few instances wherein the Casimir effect can give rise to repulsive forces between uncharged objects. Evgeny Lifshitz showed (theoretically) that in certain circumstances (most commonly involving liquids), repulsive forces can arise.[41] This has sparked interest in applications of the Casimir effect toward the development of levitating devices. An experimental demonstration of the Casimir-based repulsion predicted by Lifshitz was carried out by Munday et al.[42] who described it as "quantum levitation". Other scientists have also suggested the use of gain media to achieve a similar levitation effect,[43][44] though this is controversial because these materials seem to violate fundamental causality constraints and the requirement of thermodynamic equilibrium (Kramers–Kronig relations). Casimir and Casimir-Polder repulsion can in fact occur for sufficiently anisotropic electrical bodies; for a review of the issues involved with repulsion see Milton et al.[45] More on tunable repulsive Casimir effect.[46]
Speculative applications

It has been suggested that the Casimir forces have application in nanotechnology,[47] in particular silicon integrated circuit technology based micro- and nanoelectromechanical systems, and so-called Casimir oscillators.[48]

The Casimir effect shows that quantum field theory allows the energy density in certain regions of space to be negative relative to the ordinary vacuum energy, and it has been shown theoretically that quantum field theory allows states where the energy can be arbitrarily negative at a given point.[49] Many physicists such as Stephen Hawking,[50] Kip Thorne,[51] and others[52][53][54] therefore argue that such effects might make it possible to stabilize a traversable wormhole.
See also

Casimir pressure
Negative energy
Scharnhorst effect
Van der Waals force
Squeezed vacuum

References

Lamoreaux, S. K. (1997). "Demonstration of the Casimir Force in the 0.6 to 6 μm Range". Physical Review Letters. 78 (1): 5–8. Bibcode:1997PhRvL..78....5L. doi:10.1103/PhysRevLett.78.5. S2CID 25323874.
E. L. Losada" Functional Approach to the Fermionic Casimir Effect Archived 31 May 2011 at the Wayback Machine"
Michael Bordag; Galina Leonidovna Klimchitskaya; Umar Mohideen (2009). "Chapter I; § 3: Field quantization and vacuum energy in the presence of boundaries". Advances in the Casimir effect. Oxford University Press. pp. 33 ff. ISBN 978-0-19-923874-3. Reviewed in Lamoreaux, Steve K. (2010). "Advances in the Casimir Effect Advances in the Casimir Effect, M. Bordag, G. L. Klimchitskaya, U. Mohideen, and V. M. Mostepanenko Oxford U. Press, New York, 2009. $150.00 (749 pp.). ISBN 978-0-19-923874-3". Physics Today. 63 (8): 50–51. Bibcode:2010PhT....63h..50B. doi:10.1063/1.3480079.
Griffiths, D. J.; Ho, E. (2001). "Classical Casimir effect for beads on a string". American Journal of Physics. 69 (11): 1173. Bibcode:2001AmJPh..69.1173G. doi:10.1119/1.1396620.
Cooke, J. H. (1998). "Casimir force on a loaded string". American Journal of Physics. 66 (7): 569–572. Bibcode:1998AmJPh..66..569C. doi:10.1119/1.18907.
Denardo, B. C.; Puda, J. J.; Larraza, A. S. (2009). "A water wave analog of the Casimir effect". American Journal of Physics. 77 (12): 1095. Bibcode:2009AmJPh..77.1095D. doi:10.1119/1.3211416.
Larraza, A. S.; Denardo, B. (1998). "An acoustic Casimir effect". Physics Letters A. 248 (2–4): 151. Bibcode:1998PhLA..248..151L. doi:10.1016/S0375-9601(98)00652-5.
Astrid Lambrecht, Serge Reynaud and Cyriaque Genet (2007) "Casimir In The Nanoworld" Archived 22 November 2009 at the Wayback Machine
Genet, C.; Intravaia, F.; Lambrecht, A.; Reynaud, S. (2004). "Electromagnetic vacuum fluctuations, Casimir and Van der Waals forces" (PDF). Annales de la Fondation Louis de Broglie. 29 (1–2): 311–328. arXiv:quant-ph/0302072. Bibcode:2003quant.ph..2072G.
The Force of Empty Space, Physical Review Focus, 3 December 1998
Lambrecht, A. (1 September 2002). "The Casimir effect: a force from nothing". Physics World. Retrieved 17 July 2009.
American Institute of Physics News Note 1996
Jaffe, R. (2005). "Casimir effect and the quantum vacuum". Physical Review D. 72 (2): 021301. arXiv:hep-th/0503158. Bibcode:2005PhRvD..72b1301J. doi:10.1103/PhysRevD.72.021301. S2CID 13171179.
Casimir, H. B. G.; Polder, D. (15 February 1948). "The Influence of Retardation on the London-van der Waals Forces". Physical Review. 73 (4): 360–372. Bibcode:1948PhRv...73..360C. doi:10.1103/PhysRev.73.360. ISSN 0031-899X.
Casimir, H. B. G. (1948). "On the attraction between two perfectly conducting plates" (PDF). Proc. Kon. Ned. Akad. Wet. 51: 793.
Du, Z. Z.; Liu, H. M.; Xie, Y. L.; Wang, Q. H.; Liu, J.-M. (7 December 2015). "Spin Casimir effect in noncollinear quantum antiferromagnets: Torque equilibrium spin wave approach". Physical Review B. 92 (21): 214409. arXiv:1506.05211. Bibcode:2015arXiv150605211D. doi:10.1103/PhysRevB.92.214409. ISSN 1098-0121.
SE Rugh, H Zinkernagel; Zinkernagel (2002). "The quantum vacuum and the cosmological constant problem". Studies in History and Philosophy of Science Part B: Studies in History and Philosophy of Modern Physics. 33 (4): 663–705. arXiv:hep-th/0012253. Bibcode:2002SHPMP..33..663R. doi:10.1016/S1355-2198(02)00033-3. S2CID 9007190.
Bianchi, Eugenio; Rovelli, Carlo (2010). "Why all these prejudices against a constant?". arXiv:1002.3966 [astro-ph.CO].
Schwinger, Julian; DeRaad, Lester L.; Milton, Kimball A. (1978). "Casimir effect in dielectrics". Annals of Physics. 115 (1): 1–23. Bibcode:1978AnPhy.115....1S. doi:10.1016/0003-4916(78)90172-0.
Nikolić, Hrvoje (10 October 2016). "Proof that Casimir force does not originate from vacuum energy". Physics Letters B. 761: 197–202. arXiv:1605.04143. Bibcode:2016PhLB..761..197N. doi:10.1016/j.physletb.2016.08.036. S2CID 119265677.
Nikolić, Hrvoje (August 2017). "Is zero-point energy physical? A toy model for Casimir-like effect". Annals of Physics. 383: 181–195. arXiv:1702.03291. Bibcode:2017AnPhy.383..181N. doi:10.1016/j.aop.2017.05.013. S2CID 118883930.
For a brief summary, see the introduction in Passante, R.; Spagnolo, S. (2007). "Casimir-Polder interatomic potential between two atoms at finite temperature and in the presence of boundary conditions". Physical Review A. 76 (4): 042112. arXiv:0708.2240. Bibcode:2007PhRvA..76d2112P. doi:10.1103/PhysRevA.76.042112. S2CID 119651683.
Ruggiero, Zimerman; Villani (1977). "Application of Analytic Regularization to the Casimir Forces" (PDF). Revista Brasileira de Física. 7 (3).
Dzyaloshinskii, I E; Lifshitz, E M; Pitaevskii, Lev P (1961). "General Theory of van der Waals' Forces". Soviet Physics Uspekhi. 4 (2): 153. Bibcode:1961SvPhU...4..153D. doi:10.1070/PU1961v004n02ABEH003330.
Dzyaloshinskii, I E; Kats, E I (2004). "Casimir forces in modulated systems". Journal of Physics: Condensed Matter. 16 (32): 5659. arXiv:cond-mat/0408348. Bibcode:2004JPCM...16.5659D. doi:10.1088/0953-8984/16/32/003.
V. A. Parsegian, Van der Waals Forces: A Handbook for Biologists, Chemists, Engineers, and Physicists (Cambridge Univ. Press, 2006).
Rodriguez, A. W.; Capasso, F.; Johnson, Steven G. (2011). "The Casimir effect in microstructured geometries". Nature Photonics. 5 (4): 211–221. Bibcode:2011NaPho...5..211R. doi:10.1038/nphoton.2011.39. Review article.
B. V. Derjaguin, I. I. Abrikosova, and E. M. Lifshitz, Quarterly Reviews, Chemical Society, vol. 10, 295–329 (1956).
Reid, M. T. H.; White, J.; Johnson, S. G. (2011). "Computation of Casimir interactions between arbitrary three-dimensional objects with arbitrary material properties". Physical Review A. 84 (1): 010503(R). arXiv:1010.5539. Bibcode:2011PhRvA..84a0503R. doi:10.1103/PhysRevA.84.010503.
Sparnaay, M. J. (1957). "Attractive Forces between Flat Plates". Nature. 180 (4581): 334–335. Bibcode:1957Natur.180..334S. doi:10.1038/180334b0. S2CID 4263111.
Sparnaay, M (1958). "Measurements of attractive forces between flat plates". Physica. 24 (6–10): 751–764. Bibcode:1958Phy....24..751S. doi:10.1016/S0031-8914(58)80090-7.
Mohideen, U.; Roy, Anushree (1998). "Precision Measurement of the Casimir Force from 0.1 to 0.9 µm". Physical Review Letters. 81 (21): 4549–4552. arXiv:physics/9805038. Bibcode:1998PhRvL..81.4549M. doi:10.1103/PhysRevLett.81.4549. S2CID 56132451.
Bressi, G.; Carugno, G.; Onofrio, R.; Ruoso, G. (2002). "Measurement of the Casimir Force between Parallel Metallic Surfaces". Physical Review Letters. 88 (4): 041804. arXiv:quant-ph/0203002. Bibcode:2002PhRvL..88d1804B. doi:10.1103/PhysRevLett.88.041804. PMID 11801108. S2CID 43354557.
Zao, J.; Marcet, Z.; Rodriguez, A. W.; Reid, M. T. H.; McCauley, A. P.; Kravchenko, I. I.; Lu, T.; Bao, Y.; Johnson, S. G.; Chan, H. B.; et al. (14 May 2013). "Casimir forces on a silicon micromechanical chip". Nature Communications. 4: 1845. arXiv:1207.6163. Bibcode:2013NatCo...4.1845Z. doi:10.1038/ncomms2842. PMID 23673630. S2CID 46359798.
Ajdari, A.; Duplantier, B.; Hone, D.; Peliti, L.; Prost, J. (March 1992). ""Pseudo-Casimir" effect in liquid crystals". Journal de Physique II. 2 (3): 487–501. Bibcode:1992JPhy2...2..487A. doi:10.1051/jp2:1992145. S2CID 55236741.
Fulling, S. A.; Davies, P. C. W. (1976). "Radiation from a Moving Mirror in Two Dimensional Space-Time: Conformal Anomaly". Proceedings of the Royal Society A. 348 (1654): 393. Bibcode:1976RSPSA.348..393F. doi:10.1098/rspa.1976.0045. S2CID 122176090.
"First Observation of the Dynamical Casimir Effect". Technology Review.
Wilson, C. M.; Johansson, G.; Pourkabirian, A.; Simoen, M.; Johansson, J. R.; Duty, T.; Nori, F.; Delsing, P. (2011). "Observation of the Dynamical Casimir Effect in a Superconducting Circuit". Nature. 479 (7373): 376–379. arXiv:1105.4714. Bibcode:2011Natur.479..376W. doi:10.1038/nature10561. PMID 22094697. S2CID 219735.
"Dynamical Casimir effect in a Josephson metamaterial". Proceedings of the National Academy of Sciences of the United States of America.
Sundermier, Ali (15 August 2016). "Evidence of Stephen Hawking's famous prediction about black holes was just observed for the first time". Business Insider. Retrieved 23 January 2017.
Dzyaloshinskii, I.E.; Lifshitz, E.M.; Pitaevskii, L.P. (1961). "The general theory of van der Waals forces†". Advances in Physics. 10 (38): 165. Bibcode:1961AdPhy..10..165D. doi:10.1080/00018736100101281.
Munday, J.N.; Capasso, F.; Parsegian, V.A. (2009). "Measured long-range repulsive Casimir-Lifshitz forces". Nature. 457 (7226): 170–3. Bibcode:2009Natur.457..170M. doi:10.1038/nature07610. PMC 4169270. PMID 19129843.
Highfield, Roger (6 August 2007). "Physicists have 'solved' mystery of levitation". The Daily Telegraph. London. Retrieved 28 April 2010.
Leonhardt, Ulf; Philbin, Thomas G. (August 2007). "Quantum levitation by left-handed metamaterials". New Journal of Physics. IOP Publishing and German Physical Society. 9 (8): 254. arXiv:quant-ph/0608115. Bibcode:2007NJPh....9..254L. doi:10.1088/1367-2630/9/8/254.
Milton, K. A.; Abalo, E. K.; Parashar, Prachi; Pourtolami, Nima; Brevik, Iver; Ellingsen, Simen A. (2012). "Repulsive Casimir and Casimir-Polder Forces". J. Phys. A. 45 (37): 4006. arXiv:1202.6415. Bibcode:2012JPhA...45K4006M. doi:10.1088/1751-8113/45/37/374006. S2CID 118364958.
Jiang, Qing-Dong; Wilczek, Frank (4 March 2019). "Chiral Casimir forces: Repulsive, enhanced, tunable". Physical Review B. 99 (12): 125403. arXiv:1805.07994. Bibcode:2019PhRvB..99l5403J. doi:10.1103/PhysRevB.99.125403. S2CID 67802144.
Capasso, F.; Munday, J.N.; Iannuzzi, D.; Chan, H.B. (2007). "Casimir forces and quantum electrodynamical torques: physics and nanomechanics". IEEE Journal of Selected Topics in Quantum Electronics. 13 (2): 400. Bibcode:2007IJSTQ..13..400C. doi:10.1109/JSTQE.2007.893082. S2CID 32996610.
Serry, F.M.; Walliser, D.; MacLay, G.J. (1995). "The anharmonic Casimir oscillator (ACO)-the Casimir effect in a model microelectromechanical system" (PDF). Journal of Microelectromechanical Systems. 4 (4): 193. doi:10.1109/84.475546.
Everett, Allen; Roman, Thomas (2012). Time Travel and Warp Drives. University of Chicago Press. p. 167. ISBN 978-0-226-22498-5.
"Space and Time Warps". Hawking.org.uk. Archived from the original on 10 February 2012. Retrieved 11 November 2010.
Morris, Michael; Thorne, Kip; Yurtsever, Ulvi (1988). "Wormholes, Time Machines, and the Weak Energy Condition" (PDF). Physical Review Letters. 61 (13): 1446–1449. Bibcode:1988PhRvL..61.1446M. doi:10.1103/PhysRevLett.61.1446. PMID 10038800.
Sopova, V.; Ford, L. H. (2002). "The Energy Density in the Casimir Effect". Physical Review D. 66 (4): 045026. arXiv:quant-ph/0204125. Bibcode:2002PhRvD..66d5026S. doi:10.1103/PhysRevD.66.045026. S2CID 10649139.
Ford, L. H.; Roman, Thomas A. (1995). "Averaged Energy Conditions and Quantum Inequalities". Physical Review D. 51 (8): 4277–4286. arXiv:gr-qc/9410043. Bibcode:1995PhRvD..51.4277F. doi:10.1103/PhysRevD.51.4277. PMID 10018903. S2CID 7413835.

Olum, Ken D. (1998). "Superluminal travel requires negative energies". Physical Review Letters. 81 (17): 3567–3570. arXiv:gr-qc/9805003. Bibcode:1998PhRvL..81.3567O. doi:10.1103/PhysRevLett.81.3567. S2CID 14513456.

Further reading
Introductory readings

Casimir effect description from University of California, Riverside's version of the Usenet physics FAQ.
A. Lambrecht, The Casimir effect: a force from nothing, Physics World, September 2002.
NASA Astronomy Picture of the Day: Casimir effect (17 December 2006)
Simpson, W. M. R; Leonhardt, U. (2015). Forces of the Quantum Vacuum: An introduction to Casimir physics. World Scientific. ISBN 978-981-4632-90-4.

Papers, books and lectures

Casimir, H. B. G.; Polder, D. (1948). "The Influence of Retardation on the London-van der Waals Forces". Physical Review. 73 (4): 360–372. Bibcode:1948PhRv...73..360C. doi:10.1103/PhysRev.73.360.
Casimir, H. B. G. (1948). "On the attraction between two perfectly conducting plates" (PDF). Proceedings of the Koninklijke Nederlandse Akademie van Wetenschappen. B51: 793–795.
Lamoreaux, S. K. (1997). "Demonstration of the Casimir Force in the 0.6 to 6 μm Range". Physical Review Letters. 78 (1): 5–8. Bibcode:1997PhRvL..78....5L. doi:10.1103/PhysRevLett.78.5. S2CID 25323874.
Bordag, M.; Mohideen, U.; Mostepanenko, V. M. (October 2001). "New developments in the Casimir effect". Physics Reports. 353 (1–3): 1–205. arXiv:quant-ph/0106045. Bibcode:2001PhR...353....1B. doi:10.1016/S0370-1573(01)00015-1. S2CID 119352552.
Milton, K. A. (2001). The Casimir Effect: Physical Manifestations of Zero-point Energy (Reprint ed.). World Scientific. ISBN 978-981-02-4397-5.
Dalvit, Diego; Milonni, Peter; Roberts, David; Da Rosa, Felipe (2011). Dalvit, Diego; Milonni, Peter W.; Roberts, David; da Rosa, Felipe (eds.). Casimir Physics. Lecture Notes in Physics LNP. Lecture Notes in Physics. 834. p. 210301. arXiv:1007.0966. Bibcode:2011LNP...834.....D. doi:10.1007/978-3-642-20288-9. ISBN 978-3-642-20287-2. ISSN 0075-8450. OCLC 844922239. PMID 25965028. S2CID 118655763.
Bressi, G.; Carugno, G.; Onofrio, R.; Ruoso, G. (2002). "Measurement of the Casimir Force between Parallel Metallic Surfaces". Physical Review Letters. 88 (4): 041804. arXiv:quant-ph/0203002. Bibcode:2002PhRvL..88d1804B. doi:10.1103/PhysRevLett.88.041804. PMID 11801108. S2CID 43354557.
Kenneth, O.; Klich, I.; Mann, A.; Revzen, M. (2002). "Repulsive Casimir Forces". Physical Review Letters. 89 (3): 033001. arXiv:quant-ph/0202114. Bibcode:2002PhRvL..89c3001K. doi:10.1103/PhysRevLett.89.033001. PMID 12144387. S2CID 20903628.
Barrow, J. D. (2005). "Much Ado About Nothing". Lecture at Gresham College. Archived from the original on 30 September 2007. (Includes discussion of French naval analogy.)
Barrow, J. D. (2000). The Book of Nothing: Vacuums, Voids, and the Latest Ideas About the Origins of the Universe. Pantheon Books. ISBN 978-0-09-928845-9. (Also includes discussion of French naval analogy.)
Downling, J. P. (1989). "The Mathematics of the Casimir Effect". Mathematics Magazine. 62 (5): 324–331. doi:10.1080/0025570X.1989.11977464.
Patent No. PCT/RU2011/000847 Author Urmatskih.

Temperature dependence

Measurements Recast Usual View of Elusive Force from NIST
Nesterenko, V. V.; Lambiase, G.; Scarpetta, G. (2005). "Calculation of the Casimir energy at zero and finite temperature: Some recent results". Rivista del Nuovo Cimento. 27 (6): 1–74. arXiv:hep-th/0503100. Bibcode:2004NCimR..27f...1N. doi:10.1393/ncr/i2005-10002-2. S2CID 14693485.

External links

Casimir effect article search on arxiv.org
G. Lang, The Casimir Force web site, 2002
J. Babb, bibliography on the Casimir Effect web site, 2009

vte

Quantum mechanics
Background

Introduction History
timeline Glossary Classical mechanics Old quantum theory

Fundamentals

Bra–ket notation Casimir effect Coherence Coherent control Complementarity Density matrix Energy level
degenerate levels excited state ground state QED vacuum QCD vacuum Vacuum state Zero-point energy Hamiltonian Heisenberg uncertainty principle Pauli exclusion principle Measurement Observable Operator Probability distribution Quantum Qubit Qutrit Scattering theory Spin Spontaneous parametric down-conversion Symmetry Symmetry breaking
Spontaneous symmetry breaking No-go theorem No-cloning theorem Von Neumann entropy Wave interference Wave function
collapse Universal wavefunction Wave–particle duality
Matter wave Wave propagation Virtual particle

Quantum

quantum coherence annealing decoherence entanglement fluctuation foam levitation noise nonlocality number realm state superposition system tunnelling Quantum vacuum state

Mathematics
Equations

Dirac Klein–Gordon Pauli Rydberg Schrödinger

Formulations

Heisenberg Interaction Matrix mechanics Path integral formulation Phase space Schrödinger

Other

Quantum
algebra calculus
differential stochastic geometry group Q-analog
List

Interpretations

Bayesian Consistent histories Cosmological Copenhagen de Broglie–Bohm Ensemble Hidden variables Many worlds Objective collapse Quantum logic Relational Stochastic Transactional

Experiments

Afshar Bell's inequality Cold Atom Laboratory Davisson–Germer Delayed-choice quantum eraser Double-slit Elitzur–Vaidman Franck–Hertz experiment Leggett–Garg inequality Mach-Zehnder inter. Popper Quantum eraser Quantum suicide and immortality Schrödinger's cat Stern–Gerlach Wheeler's delayed choice

Science

Measurement problem QBism

Quantum

biology chemistry chaos cognition complexity theory computing
Timeline cosmology dynamics economics finance foundations game theory information nanoscience metrology mind optics probability social science spacetime

Technologies

Quantum technology
links Matrix isolation Phase qubit Quantum dot
cellular automaton display laser single-photon source solar cell Quantum well
laser

Extensions

Dirac sea Fractional quantum mechanics Quantum electrodynamics
links Quantum geometry Quantum field theory
links Quantum gravity
links Quantum information science
links Quantum statistical mechanics Relativistic quantum mechanics De Broglie–Bohm theory Stochastic electrodynamics

Related

Quantum mechanics of time travel Textbooks

Physics Encyclopedia

World

Index

Hellenica World - Scientific Library

Retrieved from "http://en.wikipedia.org/"
All text is available under the terms of the GNU Free Documentation License