<<

REVIEWS OF MODERN , VOLUME 81, OCTOBER–DECEMBER 2009

The Casimir between real materials: Experiment and theory

G. L. Klimchitskaya* Institute for , Leipzig University, Postfach 100920, Leipzig D-04009, Germany and North-West Technical University, Millionnaya Street 5, St. Petersburg 191065, Russia

U. Mohideen† Department of Physics and Astronomy, University of California, Riverside, California 92521, USA

V. M. Mostepanenko‡ Institute for Theoretical Physics, Leipzig University, Postfach 100920, Leipzig D-04009, Germany and Noncommercial Partnership “Scientific Instruments,” Tverskaya Street 11, Moscow 103905, Russia ͑Published 21 December 2009͒

The physical origin of the Casimir force is connected with the existence of zero-point and thermal fluctuations. The is very general and finds applications in various fields of physics. This review is limited to the rapid progress at the intersection of experiment and theory that has been achieved in the last few years. It includes a critical assessment of the proposed approaches to the resolution of the puzzles arising in the applications of the Lifshitz theory of the van der Waals and Casimir to real materials. All the primary experiments on the measurement of the Casimir force between macroscopic bodies and the Casimir-Polder force between an atom and a wall that have been performed in the last decade are reviewed, including the theory needed for their interpretation. The methodology for the comparison between experiment and theory in the force- measurements is presented. The experimental and theoretical results described here provide a deeper understanding of the phenomenon of forces in real materials and offer guidance for the application of the Lifshitz theory to the interpretation of the measurement results.

DOI: 10.1103/RevModPhys.81.1827 PACS number͑s͒: 12.20.Fv, 12.20.Ds, 78.20.Ci, 85.85.ϩj

CONTENTS 2. Real 1840 III. How to Compare Theory and Experiment 1842 A. Modeling of the optical properties of real materials 1842 I. Introduction 1828 1. Kramers-Kronig relations 1842 A. Fluctuations and the physical origin of the van der 2. Calculation for Au 1843 Waals and Casimir forces 1828 3. Problem of spatial dispersion 1844 B. Material and geometric properties in the theory of B. Corrections to the Casimir force due to the dispersion forces 1829 imperfect geometry of interacting bodies 1845 C. Modern Casimir force experiments 1830 1. Surface roughness 1845 D. Applications of the Casimir effect from fundamental 2. Finite size and thickness 1847 physics to 1832 C. Quantitative comparison between experiment and E. Structure of the review 1832 theory in force-distance measurements 1848 II. Lifshitz’s Theory of the Thermal van der Waals and 1. Experimental errors and precision 1848 Casimir Forces 1833 2. Comparison of experiment with theory 1849 A. Casimir interaction between two planar plates 1833 IV. Casimir Experiments with Metallic Test Bodies 1850 B. Approximations for nonplanar boundary surfaces 1834 A. Measurements of the Casimir force using an atomic C. Casimir-Polder atom-plate interaction and force microscope 1850 Bose-Einstein condensation 1835 B. Precise determination of the Casimir using D. Puzzles in the application of the Lifshitz theory to a micromachined oscillator 1854 real materials 1836 C. Other experiments on the Casimir force between 1. Real metals 1836 metal bodies 1858 1. Torsion pendulum experiment 1858 2. Micromechanical devices actuated by the *[email protected] Casimir force 1858 †[email protected] 3. The experiment using the parallel plate ‡[email protected] configuration 1859

0034-6861/2009/81͑4͒/1827͑59͒ 1827 ©2009 The American Physical Society 1828 Klimchitskaya, Mohideen, and Mostepanenko: The Casimir force between real materials: …

4. The Casimir force between thin metallic gian, 2005͒, the generic name for both the van der Waals films 1860 and Casimir forces. 5. Dynamic measurement using an atomic For atomic separations between a few angstroms and force microscope 1860 a few nanometers ͑which are much smaller than the 6. Ambient Casimir force measurements 1860 characteristic absorption wavelength͒ the retardation ef- 7. Related measurements 1861 fects are negligible. In this separation region the disper- D. Future prospects to measure the thermal Casimir sion force is usually called the . This force 1862 is a nonrelativistic phenomenon and its theory V. Casimir Force Between a Metallic Sphere and a was pioneered by London ͑1930͒. In the separation re- Semiconductor Plate 1863 gion of the van der Waals interaction, a virtual A. Motivation for use of semiconductors 1863 emitted by one atom can be absorbed by another atom B. Optically modulated Casimir force 1863 during the lifetime of this photon as determined by the C. Doped semiconductors with different charge carrier Heisenberg uncertainty relation. At relatively large densities 1867 atomic separations, of order of or larger than the char- 1. p-type silicon 1867 acteristic absorption wavelength, relativistic retardation 2. n-type silicon 1868 effects play an important role. At such separations the D. Silicon plate with rectangular trenches 1870 dispersion forces are usually called Casimir-Polder ͑for E. Future prospects to measure the Casimir force with atom-atom and atom-wall interactions͒ or Casimir ͑for semiconductor surfaces 1871 interaction between two macroscopic bodies͒ forces. 1. The -metal transition 1871 These are both relativistic and quantum-mechanical 2. Casimir forces between a sphere and a phenomena first described by Casimir and Polder ͑1948͒ plate with patterned geometry 1872 and by Casimir ͑1948͒, respectively. For atomic separa- ͉ ͉ 3. Pulsating Casimir force 1872 tions r2 −r1 of order of the characteristic absorption VI. Experiments on the Casimir-Polder Force 1873 wavelength, the virtual photon emitted by one atom can- A. Demonstration of the thermal Casimir-Polder force 1873 not reach the second one during its lifetime. However, 1. The force in thermal equilibrium 1873 the operators of the quantized electric field at the points 2. The force out of thermal equilibrium 1874 r1 and r2 in the state are correlated such that B. Future prospects to measure the Casimir-Polder ͗ ͑ ͒͘ ͗ ͑ ͒ ͑ ͒͘  ͑ ͒ force in quantum reflection 1875 Ek t,r1,2 = 0 but Ek t,r1 Ej t,r2 0. 1 C. Casimir-Polder interaction of atoms with carbon nanotubes 1876 As a result, the atoms situated at points r1 and r2 are VII. Lateral Casimir Force and Casimir Torques 1877 characterized by fluctuating dipole moments and these A. Lateral Casimir force between corrugated surfaces 1877 fluctuations are correspondingly correlated, leading to B. Demonstration of the lateral Casimir force 1877 the Casimir-Polder and Casimir forces. C. The Casimir torque 1879 The theoretical approach to the atom-wall and wall- VIII. Conclusions and Outlook 1879 wall interactions developed by Casimir and Polder used Acknowledgments 1880 ideal metal walls at zero temperature. The finite Casimir References 1880 per unit of two infinitely large parallel ideal metal walls separated by a distance a ͑and the respective pressure͒ was found as a difference between the ener- gies of zero-point ͑vacuum͒ of the electro- I. INTRODUCTION magnetic field in the presence and in the absence of walls as A. Fluctuations and the physical origin of the van der Waals and Casimir forces ␲2 បc ␲2 បc E ͑a͒ =− , P ͑a͒ =− . ͑2͒ 0 720 a3 0 240 a4 Long-range forces that are different from gravity but act between electrically neutral atoms or between an Ideal metals are characterized by perfect reflectivity at atom and a macrobody or between two macrobodies all frequencies which means that the absorption wave- have been discussed for centuries. However, only after length is zero. Thus, the results ͑2͒ are universal and the development of quantum and quantum valid at any separation distance. They do not transform field theory has the physical picture of these forces be- to the nonrelativistic London forces at short separations. come clear and the first quantitative results obtained. Due to the difference in these early theoretical ap- The origin of both the van der Waals and Casimir forces proaches to the description of the dispersion forces, the is connected with the existence of quantum fluctuations. van der Waals and Casimir-Polder ͑Casimir͒ forces were For a nonpolar atom the mean value of the of thought of as two different kinds of force rather than the dipole moment in the is equal to zero. two limiting cases of a single physical phenomenon, as However, due to quantum fluctuations, the mean value they are presently understood. of the square of the dipole moment is not equal to zero. A unified theory of both the van der Waals and Ca- This leads to the existence of what are referred to as simir forces between plane parallel material plates in dispersion forces ͑Mahanty and Ninham, 1976; Parse- thermal equilibrium separated by a vacuum gap was de-

Rev. Mod. Phys., Vol. 81, No. 4, October–December 2009 Klimchitskaya, Mohideen, and Mostepanenko: The Casimir force between real materials: … 1829 veloped by Lifshitz ͑1956͒. Lifshitz’s theory describes las and more general results in different media where dispersion forces between dissipative media as a physical the dispersion force is viewed as a vacuum quantum ef- phenomenon caused by the fluctuating electromagnetic fect ͑Tomaš, 2002; Raabe and Welsch, 2006͒. The com- field that is always present in both the interior and the mon roots in the theory of electromagnetic oscillations exterior of any medium. According to the fluctuation- relate the Casimir effect to other fluctuation phenom- dissipation theorem, there is a connection between the ena, such as the radiative heat transfer through a spectrum of fluctuations of the physical quantity in an vacuum gap ͑Volokitin and Persson, 2007͒. However, its equilibrium dissipative medium and the generalized sus- origin in quantum field theory relates the Casimir effect ceptibilities of this medium which describe its reaction to other quantum vacuum effects like the to an external influence. Using the fluctuation- and the anomalous magnetic moment of an electron, dissipation theorem, Lifshitz derived the general formu- where virtual particles play an important role ͑Jaffe, las for the free energy and force of the dispersion inter- 2005͒. action. In the limit of dilute bodies these formulas During the last few years, far-reaching generalizations describe the dispersion forces acting between atoms and of the Lifshitz formulas were obtained which express the molecules. In the framework of the Lifshitz theory ma- Casimir energy and force between two separated bodies terial properties are represented by the frequency- of arbitrary shape in terms of matrices of infinite dimen- dependent dielectric permittivities and atomic proper- sions. This is often referred to as the representation ties by the dynamic atomic polarizabilities. In the of the Casimir energy in terms of functional determi- limiting cases of small and large separation , in nants ͑Emig et al., 2006; Kenneth and Klich, 2006, 2008͒ comparison with the characteristic absorption wave- or in terms of scattering matrices ͑Bulgac et al., 2006; length, the Lifshitz theory reproduces the results ob- Lambrecht et al., 2006͒. Several applications of new rep- tained by London and by Casimir and Polder, respec- resentations related to the subject of this review are dis- tively. It also describes the transition region between the cussed in the respective sections. nonrelativistic and relativistic . In fact, the fluctuat- ing electromagnetic field in the Lifshitz theory is some ͑ ͒ B. Material and geometric properties in the theory of classical analog of vacuum zero-point oscillations in dispersion forces the field-theoretical approach developed by Casimir. van Kampen et al. ͑1968͒, Ninham et al. ͑1970͒, Gerlach Both Casimir and Lifshitz considered dispersion ͑1971͒, and Schram ͑1973͒ narrowed the distinction be- forces in the configuration of two closely spaced plane tween the Casimir and Lifshitz approaches. They ob- parallel plates. However, real bodies may possess quite tained the Lifshitz formulas for the free energy and different geometrical properties. The case of an isolated force between two nondissipative material plates as the body is also of much interest for the physics of disper- difference between the free of zero-point and sion forces. One unique feature of the Casimir force that thermal oscillations in the presence and in the absence has attracted widespread attention is its exotic geometry of plates. The eigenfrequencies of these oscillations dependence. For example, it was first realized that the were found from the standard continuity boundary con- Casimir force can be repulsive for an ideal metal spheri- ditions for the electric field and magnetic induction on cal shell ͑Boyer, 1968͒. Another example of the geomet- the surfaces of the dielectric plates. Later Barash and ric dependence was found in ideal metal rectangular Ginzburg ͑1975͒ generalized this approach for the case boxes, where the repulsive or attractive of the of plates made of dissipative materials in thermal equi- force depends on the ratio of the size of the sides ͑Lu- librium with a heat reservoir. This generalization is pre- kosz, 1971͒. Both these results led to an explosion in sented by Milonni ͑1994͒ and by Mostepanenko and theoretical studies of the shape dependence of the Ca- Trunov ͑1997͒. The applicability of the Lifshitz formula simir force with ideal boundaries extensively reviewed to dissipative materials was also demonstrated using the ͑Milonni, 1994; Mostepanenko and Trunov, 1997; Mil- scattering approach ͑Genet, Lambrecht, and Reynaud, ton, 2001͒. This subject is, however, outside the scope of 2003͒. The assumption of a thermal equilibrium is basic our review, which deals only with effects pertaining to for the Lifshitz theory. We repeatedly discuss the role of real material bodies. this assumption below. In addition to the shape dependence mentioned Thus, the theoretical foundations of the Casimir inter- above, there are more staid but no less important prop- action are based on two approaches. In one case the erties of interest for experimental measurements of the theory of the Casimir effect is based on the theory of Casimir force between real bodies. Chief among these equilibrium electromagnetic fluctuations in the media. are the effect of the surface roughness on Casimir force, In the second case the Casimir effect is a vacuum quan- the role of curvature in the most often used sphere-plate tum effect resulting from the influence of external con- configuration, the finite-size effects of the bodies used in ditions and described by quantum field theory. In this the measurement, and the lateral Casimir force from case, boundary conditions are imposed ͑in place of ma- corrugated surfaces. The last in particular exhibits dif- terial boundaries͒ which restrict the vol- fractionlike effects, which can be experimentally ob- ume and affect the spectrum of zero-point and thermal served. All these aspects are covered in the review. oscillations. In fact the two different approaches can be Another feature of real material bodies differentiating reconciled. There are derivations of the Lifshitz formu- them from ideal metal plates at zero temperature, as

Rev. Mod. Phys., Vol. 81, No. 4, October–December 2009 1830 Klimchitskaya, Mohideen, and Mostepanenko: The Casimir force between real materials: … considered by Casimir, concerns the account of realistic shitz theory inconsistent with thermodynamics. More- material properties and thermal effects. As mentioned over, a large thermal correction resulting from the in Sec. I.A, in the Lifshitz theory real material proper- simple inclusion of the dc conductivity was excluded ex- ties are taken into account through the frequency- perimentally in the experiments by Chen et al. ͑2007a, dependent dielectric permittivity ␧͑␻͒. In its turn ␧͑␻͒ is 2007b͒ and Obrecht et al. ͑2007͒ ͑see Secs. I.C, V. B , and expressed in terms of the complex index of refraction as VI.A͒. Below we illustrate by many examples that the dc ␧͑␻͒=n2͑␻͒. Both quantities are measurable over wide conductivity of dielectric materials is unrelated to dis- frequency regions. The magnitude of ␧͑␻͒ depends on persion forces and its inclusion contradicts the applica- the nature and structure of the material, including its bility conditions of the Lifshitz theory. electrical and optical properties. In the Lifshitz theory of For the Casimir force between real metals the situa- dispersion forces the free energy and other thermody- tion is closely analogous to that of the dielectrics. Bos- ␧͑ ␰ ͒ tröm and Sernelius ͑2000͒ applied the Lifshitz theory to namic quantities are expressed as functionals of i l , ␰ ␲ ប describe the free energy and force in the configuration where the Matsubara frequencies l =2 kBTl/ are de- fined for all l=0,1,2,... ͑k is the Boltzmann constant of two parallel metal plates at a temperature T. The B dielectric permittivity of the metal was obtained using and T is the temperature͒. In the alternative but math- the tabulated optical data for the complex index of re- ematically equivalent formulation, the free energy is ␧͑␻͒ fraction extrapolated to zero frequency by means of the represented as a functional of defined along the ͑ ͒ ͑ ͒ Drude model see Sec. III.A . This takes into account real frequency axis see Sec. II.A . The Lifshitz theory both the drift current of the conduction electrons includ- does not make a distinction among the physical pro- ing relaxation inherent to it and displacement currents cesses contributing to the value of the dielectric permit- connected with plasma oscillations of free electrons and tivity. However, the applicability conditions of this interband transitions of core electrons. As a result, an theory, in particular thermal equilibrium, must be fol- enormously large thermal correction to the Casimir free lowed. energy and force was obtained from the Lifshitz theory ͑ As is clear from the above, we call materials real real at short separations below 1 ␮m, in qualitative disagree- ͒ metals, real dielectrics, etc. to underline that this review ment with the case of ideal metal plates. The Casimir deals with physical surfaces rather than with idealized force between plates made of ideal metals at nonzero ones such as ideal metals or dielectrics with constant temperature was first treated independently by Brown dielectric permittivity and perfect geometric shape. and Maclay ͑1969͒ using thermal quantum field theory. In the last ten years significant progress has been Schwinger et al. ͑1978͒ demonstrated that the Lifshitz made in the measurement of the Casimir and Casimir- theory is in agreement with the case of ideal metals if Polder forces. This has added substantial information to the limit ␧͑i␰͒→ϱ is taken first before the limit ␰→0. the application of the Lifshitz theory to real materials. Bezerra et al. ͑2002b, 2004͒ showed that the inclusion of Historically the Lifshitz theory was proposed for ideal relaxation processes of conduction electrons in the Lif- dielectrics, i.e., for materials having zero conductivity at shitz theory violates the Nernst heat theorem for metals any temperature. At nonzero temperature dielectrics are with perfect crystal lattices. This is also in violation of characterized by a small but physically real dc conduc- the classical limit at large separations ͑Feinberg et al., tivity. From the start, the first papers ͑Lifshitz, 1956; 2001; Scardicchio and Jaffe, 2006͒. The large thermal Dzyaloshinskii et al., 1961͒ neglected this conductivity in correction predicted by Boström and Sernelius ͑2000͒ the Lifshitz theory, and at zero frequency dielectric ma- was found to be inconsistent with the measurement data terials were characterized by a finite static dielectric per- of the experiments by Lamoreaux ͑1997͒,byDecca, ␧ ␧͑ ͒ mittivity 0 = 0 . If, however, the dc conductivity of di- Fischbach, et al. ͑2003͒,byDecca, López, Fischbach, et electrics at T0 is included, ␧͑␻͒ goes to infinity in the al. ͑2005͒, Decca et al. ͑2007a, 2007b͒ ͑see Secs. I.C, IV.B, limit of zero frequency ͑see Sec. II.D͒. Geyer et al. and IV.C.1͒. At the same time, the use of the free elec- ͑2005b͒ showed that the inclusion of even a negligibly tron plasma model for the characterization of a metal small dc conductivity for the configuration of two paral- ͑Dzyaloshinskii et al., 1961; Hargreaves, 1965; Schwinger lel dielectric plates leads to an enormously large thermal et al., 1978; Mostepanenko and Trunov, 1985͒ leads to correction to the Casimir free energy and pressure. If a small thermal corrections at short separations in qualita- physical effect is negligibly small, its inclusion in a gen- tive agreement with the case of ideal metal plates ͑Bor- eral theory must not significantly influence the results. dag et al., 2000b; Genet et al., 2000͒. The results obtained Thus, this thermal correction is surprising and it suggests by neglecting the role of the conduction electrons con- that perhaps something is being done incorrectly. One nected with the drift current are in agreement with ther- might think that the dc conductivity of dielectric mate- modynamics and with the classical limit ͑Bezerra et al., rials is not a minor effect that can be neglected in the 2004͒. As argued below, the inclusion of a drift current is theory of dispersion forces and that the discovered large beyond the applicability conditions of the Lifshitz thermal correction is physically real but was overlooked theory. by the founders of the Lifshitz theory. However, as C. Modern Casimir force experiments shown by Geyer et al. ͑2005b͒, the inclusion of the dc conductivity in the model of the dielectric response vio- There were many measurements of the van der Waals lates the Nernst heat theorem and thus makes the Lif- and Casimir forces made before 1990, of which only the

Rev. Mod. Phys., Vol. 81, No. 4, October–December 2009 Klimchitskaya, Mohideen, and Mostepanenko: The Casimir force between real materials: … 1831 experiment by van Blockland and Overbeek ͑1978͒ with lations of a device element entirely due to the presence metallic surfaces can be considered as an unambiguous of the zero-point vacuum oscillations. demonstration ͓a historical survey has been made by The most precise measurements of the Casimir force Bordag et al. ͑2001͔͒. The modern stage in Casimir ex- between metallic surfaces were performed in a series of periments is characterized by the use of a new genera- experiments by Decca, Fischbach, et al. ͑2003͒,by tion of precise laboratory techniques for the measure- Decca, López, et al. ͑2003͒,byDecca et al. ͑2004, 2007a, ment of small forces and short separation distances, 2007b͒, and by Decca, López, Fischbach, et al. ͑2005͒ permitting one to determine the experimental preci- using a micromechanical torsional oscillator. This new sion and compare data with theory. The experiment by technique was first used in Casimir experiments by Chan Lamoreaux ͑1997͒ using a torsion balance to measure et al. ͑2001a, 2001b͒ as a demonstration of the actuation the Casimir force between a gold-coated spherical lens of micromechanical and nanomechanical devices by the and a plate was the first in this more recent series. The Casimir force ͑see Sec. IV.C.2͒. Although Decca et al. precision from 5% to 10% that might be achieved at exploited the sphere-plate configuration, the application separations of about 1 ␮m ͑see Sec. IV.C.1 for more de- of the dynamic measurement mode and the proximity tails͒ leads to the conclusion that the Drude model is an force approximation allowed them to determine the Ca- inadequate description in the theory of the thermal Ca- simir pressure between two parallel plates. In the last of ͑ ͒ simir force. this series of experiments Decca et al., 2007a, 2007b a A series of measurements with increased sensitivity of 0.2% total experimental error was reported at a separa- the Casimir force between a metallized sphere and a tion of 160 nm. The systematic error was shown to domi- plate was performed using the atomic force microscope nate the random one in the total experimental error. ͑Mohideen and Roy, 1998; Klimchitskaya et al., 1999; This was achieved here for the first time in Casimir force Roy et al., 1999; Harris et al., 2000; Chen, Klimchitskaya, measurements. The measurements of the Casimir pres- Mohideen, and Mostepanenko, 2004͒. These experi- sure using the micromechanical torsional oscillator ex- ments introduced the idea of using metallized polysty- clude the prediction of large thermal effects at a 99.9% confidence level for a wide separation region below rene spheres which have very low mass, as one of the ␮ ͑ ͒ interacting surfaces. Even the first experiment in this se- 1 m see Sec. IV.B . ries ͑Mohideen and Roy, 1998͒ demonstrated the role of The next series of recent Casimir experiments was de- voted to the investigation of the Casimir interaction be- the skin depth and the surface roughness corrections to tween a metallized sphere and a semiconductor plate the Casimir force at separations from 120 to 300 nm. ͑Chen et al., 2005, 2007a, 2007b; Chen, Klimchitskaya, et Note that the corrections to the skin depth are often al., 2006; Chen, Mohideen, et al., 2006͒. In the early referred to as the finite conductivity corrections.Itwas stages of this research it was demonstrated that a change shown that the data are in disagreement with the ideal of the charge carrier concentration in a semiconductor metal Casimir force whereas the inclusion of the finite plate changes the value of the Casimir force. In the ex- conductivity and roughness corrections leads to excel- periments by Chen et al. ͑2007a, 2007b͒ the Casimir lent agreement between the data and theory ͑thermal ͒ force between the sphere and the plate was measured in corrections being negligibly small in this experiment . the presence and in the absence of incident laser on The experimental error in the measurement of the Ca- the plate. This allowed a fundamental test of the role of simir force between 1% and 2% depending on the con- semiconductor conductivity properties in the Casimir fidence level was justified at the shortest separations force to be performed. The data were found to be in an ͑ Chen, Klimchitskaya, Mohideen, and Mostepanenko, excellent agreement with the Lifshitz theory if the dc ͒ 2004 . conductivity of the semiconductor plate in the dark The gradient of the Casimir pressure between two phase ͑no incident light͒ is disregarded. If, however, the parallel metallic plates was measured dynamically by dc conductivity of the plate in the dark phase is in- ͑ ͒ Bressi et al. 2002 . In this experiment small oscillations cluded, the theory was found to be inconsistent with were induced on one of the plates at the resonant fre- data at a confidence level of 95% ͑see Sec. V. B ͒. quency and the frequency shift due to the Casimir force Another important experiment was performed by was measured. This frequency shift is proportional to Obrecht et al. ͑2007͒. This is the first measurement of the the derivative of the Casimir force with respect to the thermal Casimir-Polder force made both in thermal separation distance between the plates ͑see Sec. IV.C.3͒. equilibrium and in the nonequilibrium case. The Chen et al. ͑2002a, 2002b͒ measured the lateral Ca- Casimir-Polder force between Rb atoms and a dielectric simir force between a metallized sphere and a plate cov- substrate changes the frequency of dipole oscillations ered with longitudinal coaxial sinusoidal corrugations of excited in a Bose-Einstein condensate separated from a equal periods ͑Sec. VII.B͒. This force, which was theo- wall by a distance of a few micrometers. The fractional retically predicted by Golestanian and Kardar ͑1997, difference of this frequency calculated on the basis of 1998͒, is a harmonic function of the phase shift between the Lifshitz theory by disregarding the dc conductivity of the corrugations on the sphere and the plate. The dem- the wall material was found to be in excellent agreement onstrated phenomenon of the lateral Casimir force is with data ͑Obrecht et al., 2007͒. Klimchitskaya and promising for nanotechnology where, together with the Mostepanenko ͑2008b͒ showed, however, that with in- normal Casimir force, it allows one to actuate any trans- clusion of the small but physically real dc conductivity

Rev. Mod. Phys., Vol. 81, No. 4, October–December 2009 1832 Klimchitskaya, Mohideen, and Mostepanenko: The Casimir force between real materials: … for the dielectric substrate the theory is inconsistent with retical methods allowing the calculation of the Casimir data ͑see Sec. VI.A͒. force between real material bodies used in experiments. Geometry effects beyond the proximity force approxi- Below we demonstrate that the Lifshitz theory can be mation were experimentally demonstrated by Chan et al. adapted to accomplish these ends. ͑2008͒ where the gradient of the Casimir force acting between a Au-coated sphere and Si plate with deep rect- angular trenches was measured ͑Sec. V. D ͒. E. Structure of the review D. Applications of the Casimir effect from fundamental physics to nanotechnology This review is devoted to the experimental and theo- retical results obtained in the last ten years and related Both the van der Waals and Casimir forces find many to the Casimir force between real materials. It is not applications spanning the range from the purely scien- intended to cover not less important but alternative re- tific to the mostly technological. The reason lies in the sults and methods related to the Casimir effect in ideal fluctuation nature of dispersion forces and the universal configurations and more complicated geometries. The role played by fluctuations in different physical phenom- latter will be touched on only if they provide important ena. Here we list only some of the main areas where the guidelines for the experimentally oriented theoretical Casimir effect is important. For a detailed discussion of approaches applicable to real materials. the applications of the Casimir effect, see Plunien et al. In Sec. II we present the main results of the Lifshitz ͑1986͒, Mostepanenko and Trunov ͑1988͒, Krech ͑1994͒, theory for the configurations of two planar plates and an Milonni ͑1994͒, Mostepanenko and Trunov ͑1997͒, Kar- atom near a plate. Some approximate methods appli- dar and Golestanian ͑1999͒, Bordag et al. ͑2001͒, Milton cable to nonplanar boundary surfaces are also consid- ͑ ͒ ͑ ͒ ͑ ͒ 2001, 2004 , Lamoreaux 2005 , Parsegian 2005 , Klim- ered. Section III discusses the problem of how to com- ͑ ͒ ͑ ͒ chitskaya and Mostepanenko 2006 , and Milonni 2007 . pare theory and experiment in the Casimir effect. Here A number of applications of the Casimir effect belong we consider the modeling of the optical properties of ͑ ͒ to the field of condensed matter physics Krech, 1994 real materials and corrections to the Casimir force due ͑ and nanotechnology Buks and Roukes, 2001, 2002; to surface roughness and the finite size of the interacting ͒ Chumak et al., 2004 . Areas of impact are multilayered bodies. Special attention is paid to the quantitative mea- structures, wetting phenomena, colloids, critical systems, sure of agreement between experiment and theory in adhesion of microelements in nanoelectromechanical the force-distance measurements. Section IV presents systems, and absorption of different gases by nanostruc- the main Casimir experiments with metallic test bodies, tures. In elementary the role of the Ca- including those using an atomic force microscope and a simir effect is very important. It is included in the calcu- micromechanical oscillator. The results obtained are lation of hadron masses and provides an effective compared with different theoretical computations using mechanism for the compactification of extra spatial di- the Lifshitz theory. The prospects for measuring the mensions in multidimensional physics ͑Mostepanenko and Trunov, 1997͒. In gravitation and cosmology the Ca- thermal Casimir force between metals are also dis- simir effect results in a nonzero in spaces cussed. In Sec. V the measurements of the Casimir force with non-Euclidean topology, can drive the inflation pro- between a metallic sphere and a semiconductor plate are cess, and leads to interesting effects in brane models of presented. The discussion starts with the crucial experi- the Universe ͑Saharian, 2006͒. The Casimir effect has ment on the optically modulated Casimir force. The ex- been actively used for obtaining stronger constraints on periments with heavily doped semiconductors having the hypothetical long-range interactions predicted in different charge carrier densities are also discussed. many theoretical schemes beyond the Then we consider the recent experiment involving a ͑Bordag et al., 1998, 1999, 2000a; Long et al., 1999; semiconductor plate with deep rectangular trenches. Fischbach et al., 2001; Mostepanenko and Novello, 2001; The proposed experiments on the change in Casimir Decca, Fischbach, et al., 2003, Decca, López, Chan, et al., forces for a dielectric-metal phase transition, between a 2005, Decca, López, Fischbach, et al., 2005; Decca et al., sphere and a plate with patterned geometry, and with 2007a, 2007b; Mostepanenko et al., 2008͒. the pulsating Casimir force are considered. Section VI is Additional physical phenomena where the Casimir ef- devoted to the measurements of the Casimir-Polder fect is important are the quantum reflection of atoms on force. Here the experiment on the first measurement of different surfaces and Bose-Einstein condensation. the thermal Casimir-Polder force at large separations is Casimir-Polder forces depend on real atomic and mate- presented. Section VII deals with the lateral Casimir rial properties and influence scattering amplitudes and force and Casimir torques. After an introduction to the condensation conditions ͑Antezza et al., 2004; Babb et subject, the first experiment on the measurement of the al., 2004͒. Even in biophysics, a proper account of the lateral Casimir force between corrugated surfaces is con- Casimir force is required for understanding the interac- sidered. This is followed by consideration of the Casimir tion of biological membranes through a liquid layer. torque between corrugated surfaces at small angles and The diverse applications of the Casimir effect in both between anisotropic surfaces. Section VIII contains our fundamental and demand reliable theo- conclusions.

Rev. Mod. Phys., Vol. 81, No. 4, October–December 2009 Klimchitskaya, Mohideen, and Mostepanenko: The Casimir force between real materials: … 1833

II. LIFSHITZ’S THEORY OF THE THERMAL VAN DER ␧͑n͒͑ ͒ ͑n͒ ͑n͒ ix q − k WAALS AND CASIMIR FORCES r ͑ix,kЌ͒ = , TM ␧͑n͒͑ix͒q + k͑n͒ A. Casimir interaction between two planar plates ͑n͒ Consider two thick dissimilar plane parallel plates ͑n͒ q − k r ͑ix,kЌ͒ = ͑ ͒ , ͑6͒ ͑semispaces͒ in thermal equilibrium at equal tempera- TE q + k n ture T, separated by an empty gap of width a. The Lif- shitz formula ͑Lifshitz, 1956; Dzyaloshinskii et al., 1961͒ where represents the van der Waals and Casimir free energy ͑ ͒ and force per unit area i.e., the pressure in terms of the ͑n͒ ϵ ͑n͒͑ ͒ ͱ 2 ␧͑n͒͑ ͒ 2 2 ͑ ͒ ͑n͒ ͑␻ ͒ ͑n͒͑␻ ͒ k k ix,kЌ = kЌ + ix x /c . 7 reflection coefficients rTM ,kЌ and rTE ,kЌ for the two independent polarizations of the electromagnetic We underline that the reflection coefficients ͑6͒ are the field. Here ␻ is the frequency and kЌ is the magnitude of the projection of the wave vector onto the plane of the standard Fresnel reflection coefficients calculated, how- plates ͑the z axis is perpendicular to the plates͒. The ever, along the imaginary axis ix, where they are real. index n=1,2 labels the plates. Transverse magnetic They are derived from the standard continuity condi- ͑TM͒ means that the magnetic field is per- tions for the tangential and normal components of the electric field, magnetic induction, and electric displace- pendicular to the plane formed by kЌ and the z axis, while for transverse electric ͑TE͒ polarization the elec- ment on the boundary planes, E1t =E2t, B1n =B2n, D1n ͑ tric field is perpendicular to this plane. There are many =D2n, and B1t =B2t where the subscripts t and n denote ͒ different derivations of the Lifshitz formula in the litera- the tangential and normal components, respectively . ture based on different approaches: in the framework of Note that these conditions can be identically reformu- quantum statistical physics, thermal quantum field lated in Casimir spirit such that one deals with fields ͑ ͒ theory in the Matsubara formulation, , only external to the plates Emig and Büscher, 2004 . etc. ͓see, e.g., Schram ͑1973͒, Milonni ͑1994͒, Zhou and The reflection properties of the electromagnetic Spruch ͑1995͒, Bordag et al. ͑2001͒, and Genet, Lam- waves on metal surfaces are often described in terms of ͑ ͒ brecht, and Reynaud ͑2003͔͒. In all the derivations the the Leontovich surface impedance Landau et al., 1984 . condition of thermal equilibrium is used. The final result The corresponding expressions for the reflection coeffi- is represented in one of two equivalent forms: as a sum- cients along the imaginary frequency axis can be found ͑ ͒ ͑ ͒ mation over the Matsubara frequencies along the imagi- in Bezerra et al. 2001 and Geyer et al. 2003 . nary frequency axis or as an over real frequen- For plates of finite thickness or plates made of several cies. The first representation is used more often as it is layers, the reflection coefficients have a more compli- ͓ more convenient for computations. Here the Casimir cated form depending on layer thicknesses see, e.g., ͑ ͒ ͑ ͒ free energy per unit area is given by Zhou and Spruch 1995 , Klimchitskaya et al. 2000 , and Tomaš ͑2002͔͒. In the literature, one can find the reflec- ϱ k T tion coefficients for planar multilayer magnetodielectrics F͑a,T͒ = B ͚Ј⌽ ͑␰ ͒, ͑3͒ ␮ ͑ ͒ 2␲ E l with 1 Buhmann et al., 2005; Tomaš, 2005 . The re- l=0 flection coefficients have a more complicated form if the where the prime on the summation sign means that the plate material is anisotropic. term for l=0 has to be multiplied by 1/2 and The Casimir pressure between two plates in thermal ϱ equilibrium at a temperature T is determined from Eq. ͑ ͒ ⌽ ͑ ͒ϵ͵ ͓ ͑1͒͑ ͒ 3 , E x kЌdkЌ͚ ln 1−r␣ ix,kЌ 0 ␣ ϱ F͑a,T͒ k Tץ ͒ ͑ ϫ ͑2͒͑ ͒ −2aq͔ r␣ ix,kЌ e . 4 ͑ ͒ B ͚Ј⌽ ͑␰ ͒ ͑ ͒ ␲ P l , 8 −= ץ −= P a,T ␣ ␰ a l=0 Here denotes TM or TE, l are the Matsubara fre- quencies, and where ͱ 2 2 2 q ϵ q͑ix,kЌ͒ = kЌ + x /c . ͑5͒ ϱ ⌽ ͑ ͒ e2aq −1 Note that the function E in Eq. 3 depends on the real ⌽ ͑ ͒ϵ͵ ͩ ͪ x kЌdkЌq͚ −1 . ␰ P ͑1͒͑ ͒ ͑2͒͑ ͒ argument x= l. In the Lifshitz theory, the material me- 0 ␣ r␣ ix,kЌ r␣ ix,kЌ dia are described by the dielectric permittivity, depend- ͑ ͒ ing only on the frequency ͑Lifshitz, 1956; Dzyaloshinskii 9 et al., 1961; Lifshitz and Pitaevskii, 1980͒. The descrip- As mentioned, Eqs. ͑3͒ and ͑8͒ are useful for practical tion of the dielectric properties by ␧͑␻͒ fully accounts computations. For some purposes ͑for example, to inves- for temporal dispersion but neglects the possible contri- tigate the comparative role of propagating and evanes- butions of spatial dispersion to the van der Waals and cent waves͒, the following form of the Casimir free en- Casimir forces. In the case of homogeneous nonmag- ergy expressed as along the real frequency axis netic media, the reflection coefficients are is convenient ͑Lifshitz, 1956; Bezerra et al., 2007͒:

Rev. Mod. Phys., Vol. 81, No. 4, October–December 2009 1834 Klimchitskaya, Mohideen, and Mostepanenko: The Casimir force between real materials: …

ប ϱ ប␻ mixed quantities without a definite physical meaning F͑ ͒ ͵ ␻ ⌽ ͑ ␻͒ ͑ ͒ a,T = d coth Im E − i . 10 ͑Bezerra et al., 2002b͒. They do not match F͑a,T͒ and 4␲2 2k T 0 B P͑a,0͒, respectively. As a consequence, the thermal cor- ⌽ ͑ ͒ ␻ ⌬ F͑ ͒ ⌬ ͑ ͒ Here the function E is defined by Eq. 4 with x=−i , rections T a,T and TP a,T do not coincide with where ␻ is the real frequency. The propagating waves ⌬F͑a,T͒ and ⌬P͑a,T͒ in Eq. ͑14͒. Thus, the so-called ͑ ͒ ͑ ͒ correspond to kЌ Ͻ␻/c, i.e., to a purely imaginary q. zero-temperature Lifshitz formulas 12 and 13 can be The evanescent waves correspond to kЌ ജ␻/c, i.e., to a used for the approximate calculation of the free energy real q. These problems, however, are beyond the scope and pressure only at rather small separation distances, ͑ ͒ ͑ ͒ of this review. They are considered, for instance, by where the computational results using Eqs. 3 and 12 ͓ ͑ ͒ ͑ ͒ ͔ Henkel et al. ͑2004͒, Intravaia and Lambrecht ͑2005͒, practically coincide Eqs. 8 and 13 , respectively . Bimonte ͑2006a, 2006b͒, and Intravaia et al. ͑2007͒. Usually this is the case for the calculation of the nonrel- ͑ ͒ Using the Abel-Plana formula ͑Mostepanenko and ativistic van der Waals forces Parsegian, 2005 . Trunov, 1997͒ or the Poisson summation formula ͑ ͒ Mehra, 1967; Schwinger et al., 1978 , one can present B. Approximations for nonplanar boundary surfaces the free energy per unit area and the pressure in the following forms: The Lifshitz theory was formulated for the case of two parallel plates. Experimentally it is hard to maintain two F͑a,T͒ = E͑a͒ + ⌬F͑a,T͒, plates parallel at short separations. Because of this, most ͑ ͒ ͑ ͒ ͑ ͒ ⌬ ͑ ͒ ͑ ͒ experiments see Secs. IV, V, and VII have been per- P a,T = P a + P a,T . 11 formed using the configuration of a sphere above a Here E͑a͒ is given by plate. The configuration of a cylinder above a plate also ϱ presents some advantages in comparison with the case ប of two parallel plates ͑see Sec. IV.D͒. Unfortunately, for E͑a͒ = ͵ d␰ ⌽ ͑␰͒. ͑12͒ ␲2 E 4 0 many years it was not possible to obtain exact expres- sions for the Casimir force in these configurations. Thus, Similarly, the quantity P͑a͒ in the second equality in Eq. the approximate method of Derjaguin ͑1934͒, later ͑11͒ is called the proximity force approximation ͑PFA͒͑Blocki ប ϱ et al., 1977͒, was used to compare experiment with P͑a͒ =− ͵ d␰ ⌽ ͑␰͒. ͑13͒ theory. According to this method, the Casimir energy in ␲2 P 2 0 the gap between two smooth curved surfaces at short The second terms on the right-hand sides of equalities separation can be calculated approximately as a sum of ͑11͒ are given by energies between pairs of small parallel plates corre- sponding to the curved geometry of the gap. Specifically, ϱ ik T ⌽ ͑i␰ t͒ − ⌽ ͑− i␰ t͒ under the condition aӶR, where R is the sphere or cyl- ⌬F͑a,T͒ = B ͵ dt E 1 E 1 , ␲ 2␲t inder , the Casimir forces between the sphere or 2 0 e −1 cylinder ͑per unit length͒ and a plate are given by ϱ ⌽ ͑ ␰ ͒ ⌽ ͑ ␰ ͒ ikBT P i 1t − P − i 1t 15␲ 2R ⌬P͑a,T͒ =− ͵ dt . ͑14͒ ͑ ͒ ␲ F͑ ͒ ͑ ͒ ͱ F͑ ͒ ␲ 2␲t Fs a,T =2 R a,T , Fc a,T = a,T , 0 e −1 16 a Note that the quantities E͑a͒ and P͑a͒ in Eqs. ͑12͒ and ͑16͒ ͑ ͒ 13 are often referred to as the Casimir energy per unit respectively, where F͑a,T͒ is the Casimir free energy ⌬F͑ ͒ area and pressure at zero temperature, and a,T and per unit area in the configuration of two parallel plates, ⌬ ͑ ͒ ͑ ͒ P a,T in Eq. 14 are referred to as the thermal cor- as defined in Eq. ͑3͒. rections to them. This terminology is, however, correct Within the PFA it is not possible to control the error only for plate materials with temperature-independent of the approximation in Eq. ͑16͒. From dimensional con- properties. In this case the Casimir free energy and pres- siderations it is evident ͑Bordag et al., 2001͒ that the sure depend on the temperature only through the Mat- relative error in Eq. ͑16͒ should be of the order of a/R, subara frequencies and the thermal corrections defined but the numerical coefficient of this ratio is unknown. In as fact, a rigorous determination of the error introduced by ⌬ F͑ ͒ϵF͑ ͒ F͑ ͒ the application of the PFA requires a comparison of the T a,T a,T − a,0 , approximate results in Eq. ͑16͒ with the exact analytic ⌬ P͑a,T͒ϵP͑a,T͒ − P͑a,0͒͑15͒ results or with some precise numerical computations in T the respective configurations. are the same as ⌬F͑a,T͒ and ⌬P͑a,T͒ in Eq. ͑11͒. In this As discussed in Sec. I.B, during the last few years the case E͑a͒=F͑a,0͒ and P͑a͒=P͑a,0͒. If, however, the finite representation of the Casimir energy for two sepa- properties of the medium ͑for instance, the dielectric rated bodies A and B in terms of the functional deter- permittivity͒ depend on the temperature, then Eqs. minants was obtained. In this representation the Casimir ͑12͒–͑14͒ also contain a parametric dependence on the energy can be written as ͑Kenneth and Klich, 2006, temperature. Thus, the quantities E͑a͒ and P͑a͒ are 2008͒

Rev. Mod. Phys., Vol. 81, No. 4, October–December 2009 Klimchitskaya, Mohideen, and Mostepanenko: The Casimir force between real materials: … 1835

ϱ 1 ͑2008͒ for a/Rജ0.15. In both cases the extrapolation of ͑ ͒ ͵ ␰ ͑ TAG͑0͒ TBG͑0͒ ͒ E a = d Tr ln 1− ␰,AB ␰,BA the obtained results to smaller a/R leads to a coefficient 2␲ 0 near a/R approximately equal to 1.4. ϱ 1 In addition, the validity of the PFA for a sphere above ͵ ␰ ͑ TAG͑0͒ TBG͑0͒ ͒ ͑ ͒ = d ln det 1− ␰ ␰ . 17 ͑ 2␲ ,AB ,BA a plate has been estimated experimentally Krause et al., 0 2007͒ and the error introduced from the use of this ap- G͑0͒ proximation was shown to be less than a/R ͑see Sec. Here ␰,AB is the operator for the free space Green’s ͑ ͒ ͒ function with the matrix elements ͗r͉G 0 ͉rЈ͘, where r IV.B for details . This is in disagreement with the ex- ␰,AB ͑ ͒ A B trapolations made by Emig 2008 and Maia Neto et al. belongs to the body A and rЈ to B. T ͑T ͒ is the opera- ͑ ͒ ͑ ͒ 2008 . To solve this contradiction, it is desirable to find tor of the T matrix for body A body B . The latter is the analytical form of the first correction beyond the widely used in light scattering theory, where it is the PFA for a sphere above a plane, as in Eq. ͑18͒ for the basic object for expressing the properties of the scatter- cylinder-plane configuration. ers ͑Bohren and Huffmann, 2004͒. Using such represen- tations, Emig et al. ͑2006͒ obtained the analytic results for the electromagnetic Casimir energy for an ideal C. Casimir-Polder atom-plate interaction and Bose-Einstein metal cylinder above an ideal metal plane. Eventually, condensation the result is expressed through the determinant of an infinite matrix with elements given in terms of the Bessel The interaction of atoms with a wall has long been functions. The analytic asymptotic behavior of the exact investigated in different physical, chemical, and biologi- Casimir energy at short separations was found by Bor- ͑ ͒ cal processes including absorption and scattering from dag 2006 . It results in the following expression for the ͓ ͑ ͒ Ӷ various surfaces see, e.g., Mahanty and Ninham 1976 Casimir force at a R: and Israelashvili ͑1992͔͒. The general expression for the free energy of the atom-wall interaction can be obtained ␲3 R បc 1 20 7 a F ͑a,0͒ =− ͱ ͫ1− ͩ − ͪ ͬ. ͑18͒ from Eq. ͑3͒ by considering one of the plates as a rar- c ͱ 3 ␲2 384 2 a a 5 12 R efied dielectric ͑Lifshitz, 1956; Lifshitz and Pitaevskii, 1980; Milonni, 1994͒. In doing so one expands the dielec- The PFA result in this case matches the first term on the tric permittivity of a rarefied dielectric in powers of the right-hand side of Eq. ͑18͒. It can be obtained by replac- number of atoms per unit volume N, keeping only the ing the free energy F͑a,0͒ in the second equality in Eq. first-order contribution, ͑16͒ with the zero-temperature Casimir energy in the configuration of two ideal metal plates, E ͑a͒ from Eq. 0 ␧͑ ␰͒ ␲␣͑ ␰͒ ͑ 2͒ ͑ ͒ ͑2͒. i =1+4 i N + O N . 20 Equation ͑18͒ is very important. It demonstrates that the relative error of the electromagnetic Casimir force Here ␣͑␻͒ is the dynamic polarizability of an atom. We between a cylinder and a plate calculated using the PFA consider only atoms in the ground ͑or metastable͒ state. is equal to 0.2886a/R. Thus, for typical experimental pa- The discussion of the interaction of excited atoms with a rameters of R=100 ␮m and a=100 nm, this error is ap- wall ͓see, e.g., Buhmann and Welsch ͑2007͒, and refer- proximately equal to only 0.03%. ences therein͔ exceeds the scope of this review. For a sphere above a plate made of ideal metals at As a result, the free energy of atom-wall interaction T=0 PFA ͓Eq. ͑16͔͒ leads to the result with the wall temperature T in thermal equilibrium can be presented in the form ͑Caride et al., 2005; ͑ ͒ ␲3 ប 3 ͑ ͒ Fs a,0 =− R c/360a . 19 Mostepanenko, Babb, et al., 2006͒

For this configuration the exact analytic solution in the ϱ ϱ kЌdkЌ electromagnetic case has not yet been obtained. The sca- F͑ ͒ Ј␣͑ ␰ ͒͵ ⌽ ͑␰ ͒ ͑ ͒ a,T =−kBT͚ i l A l,kЌ , 21 lar Casimir energy for a sphere above a plate was found l=0 0 ql by Bordag ͑2006͒ and by Bulgac et al. ͑2006͒. The scalar Casimir energies for both a sphere and a cylinder above where a plate have also been computed numerically using the wordline algorithms ͑Gies and Klingmüller, 2006a, ␰2 ͒ l 2006b , but it was noted that the Casimir energies for the ⌽ ͑␰ ͒ϵ −2aqlͫͩ 2 ͪ ͑ ␰ ͒ ,kЌ e 2q − r i ,kЌ Dirichlet scalar field should not be taken as an estimate A l l c2 TM l for those in the electromagnetic case. Another numeri- ␰2 cal approach applicable in the electromagnetic case was l ͑ ␰ ͒ͬ ͑ ͒ − 2 rTE i l,y . 22 developed by Rodrigues, Ibanescu, et al. ͑2007͒. This ap- c proach has not yet been applied at short separation dis- ͑ ͒ ͑ ͒ tances of experimental interest. For an ideal metal The reflection coefficient rTM 0,kЌ in Eq. 22 takes dif- sphere above an ideal metal plane a correction of order ferent values for different wall materials. For dielectrics ␧ a/R beyond the PFA was computed numerically by with finite static dielectric permittivity 0 we get ͑ ͒ ജ ͑ ͒ ͑␧ ͒ ͑␧ ͒ ͑ ͒ Emig 2008 for a/R 0.075 and by Maia Neto et al. rTM 0,kЌ = 0 −1 / 0 +1 , and for metals rTM 0,kЌ =1.

Rev. Mod. Phys., Vol. 81, No. 4, October–December 2009 1836 Klimchitskaya, Mohideen, and Mostepanenko: The Casimir force between real materials: …

At zero temperature Eq. ͑21͒ results in Oberst, Taashiro, et al. ͑2005͒, and references therein͔.

ϱ ϱ Results on Casimir-Polder forces from such experiments ប kЌdkЌ ͑ ͒ ͵ ␰ ␣͑ ␰͒͵ ⌽ ͑␰ ͒ ͑ ͒ are discussed in Sec. VI.B. E a =− ␲ d i A ,kЌ . 23 2 0 0 q The Casimir-Polder interaction leads also to a change ␻ of center-of-mass frequency 0 of a Bose- In the nonrelativistic limit Eq. ͑23͒ leads to the well- Einstein condensate ͑Antezza et al., 2004͒, known result ͑Lifshitz and Pitaevskii, 1980͒, ␲ ␻ ␻ 2 / 0 2 2 0 ϱ ␻ − ␻ =− ͵ d␶ cos͑␻ ␶͒ ប ␧͑i␰͒ −1 C 0 z ␲Am 0 E͑a͒Ϸ− ͵ ␣͑i␰͒ d␰ ϵ − 3 . ͑24͒ a 0 ␲ 3 ␧͑ ␰͒ 3 4 a 0 i +1 a Rz ϫ ͵ ͑ ͒ ͓ ͑␻ ␶͒ ͔ dznz z F a + z + A cos 0 ,T . If we consider an atom characterized by the frequency- −Rz ͑ ͒ ␣͑ ␰ ͒ ␣͑ ͒ independent static polarizability i l = 0 near an ͑28͒ ideal metal wall at T=0, Eq. ͑23͒ results in Here the Casimir-Polder force in thermal equilibrium is 3បc C given by Eq. ͑27͒; m is the mass of atoms of the Bose- E͑a͒ =− ␣͑0͒ϵ− 4 . ͑25͒ a 8␲a4 a4 Einstein condensate under consideration. The averaging procedure includes the averaging of time over the pe- This is the famous result obtained by Casimir and Polder riod of oscillations in the z direction with an amplitude ͑1948͒. A ͑z is perpendicular to the plate͒ and the density of gas Calculations performed using the exact expression with a distribution function ͑21͒ showed that the approximation ͓Eq. ͑24͔͒ is valid 15 2 2 only at very short separation distances aϽ3 nm. To cal- ͑ ͒ ͩ z ͪ ͑ ͒ nz z = 1− 2 , 29 culate C3 in this case one should use the highly accurate 16Rz Rz N-oscillator model ͑Shih and Parsegian, 1975͒, where Rz is the Thomas-Fermi radius in the z direction. 2 N The measurement of the change in the center-of-mass e f0n ͑ ␣͑i␰͒ = ͚ , ͑26͒ oscillation frequency Harber et al., 2005; Obrecht et al., ␻2 ␰2 ͒ m n=1 0n + 2007 is a sensitive test of the Casimir-Polder force. Re- sults of the recent experiment ͑Obrecht et al., 2007͒, for the dynamic polarizability of an atom ͑Caride et al., where the thermal effect was measured for the first time 2005͒. In Eq. ͑26͒ m and e are the electron mass and in Casimir physics, are discussed in Sec. VI.A. ␻ charge and f0n and 0n are the oscillator strength and transition frequency from the nth to the ground-state transition, respectively. With the increase D. Puzzles in the application of the Lifshitz theory to real in the atom-wall distance the retardation effects become materials important. Starting from distances of several tens of na- 1. Real metals nometers, however, a more simplified single-oscillator model for the atomic polarizability is applicable. The In the framework of the Lifshitz theory of the van der description of an atom with the help of a frequency- Waals and Casimir forces, discussed above, the calcula- independent static polarizability works well only at tional results depend strongly on the model of dielectric rather large separation distances aϾ2 ␮m ͑Babb et al., permittivity used to describe real material. Different 2004͒. At room temperature, however, the temperature problems arise for metals and dielectrics. The source of corrections in Eq. ͑21͒ come into play at aജ3 ␮m. Thus, the discrepancy is in the different contributions from the the approximation ͑25͒ at room temperature works well zero frequency term ͓i.e., from the term with l=0 in Eq. only within a rather narrow separation region from ͑3͔͒. For metals, problems result from the contribution 2to3␮m. of the transverse electric mode. For ideal metal plates From Eq. ͑21͒, the Casimir-Polder force acting on an the Casimir free energy was found independently of the atom situated near a wall a distance a apart can be rep- Lifshitz theory within the frames of thermal quantum resented in the form ͑Babb et al., 2004͒ field theory with electrodynamic boundary conditions E =B =0 on the surface of plates ͑Mehra, 1967; Brown ϱ t n ϱ and Maclay, 1969͒. As mentioned in the Introduction, ͑ ͒ Ј␣͑ ␰ ͒͵ ⌽ ͑␰ ͒ ͑ ͒ F a,T =−2kBT͚ i l kЌdkЌ A l,kЌ . 27 the Lifshitz theory agrees with the case of an ideal metal l=0 0 if one takes the limit of infinite dielectric permittivity, ␧͑ ␰ ͒ ϱ Calculations of the energy of atom-wall interactions i l = , before putting the frequency equal to zero in and of the Casimir-Polder force play an important role the temperature sum ͓Eq. ͑3͔͒ ͑Schwinger et al., 1978͒. for the interpretation of experiments on quantum reflec- Then from Eqs. ͑6͒ and ͑7͒ one obtains tion, i.e., above-barrier reflection of slow atoms, with r ͑i␰ ,kЌ͒ =1, r ͑i␰ ,kЌ͒ =−1 ͑30͒ incident kinetic energy exceeding the barrier height ͓see, TM l TE l e.g., Côté et al. ͑1998͒, Druzhinina and DeKieviet ͑2003͒, for any l including l=0. After the substitution into Eq. Pasquini et al. ͑2004͒, Oberst, Kouznetsov, et al. ͑2005͒; ͑3͒ this leads to the same result for the Casimir free

Rev. Mod. Phys., Vol. 81, No. 4, October–December 2009 Klimchitskaya, Mohideen, and Mostepanenko: The Casimir force between real materials: … 1837

␻ →ϱ ͑ ͒ energy, as obtained independently using thermal quan- metal p and Eq. 33 reproduces the low- tum field theory. The sequence of limiting transitions temperature asymptotic behavior obtained with thermal used to obtain Eq. ͑30͒ is called the Schwinger prescrip- quantum field theory ͑Brown and Maclay, 1969͒.Inthe ӷ ͑ tion. high-temperature limit T Teff at room temperature T The ideal metal, however, can be obtained as the lim- =300 K it is achieved at aജ6 ␮m͒ the asymptotic ex- iting case of real metals when the conductivity goes to pression for the Casimir free energy is ͑Bordag et al., infinity. It is well known ͑Landau et al., 1984͒ that for 2000b͒ real metals ␧͑i␰͒ϳ1/␰ when ␰→0 ͑the Drude model is ␨͑ ͒ ͒ ͑ ͒ kBT 3 2c one particular case . Then from Eq. 6 we get for real F͑a,T͒ =− ͩ1− + ͪ. ͑34͒ ␲ 2 ␻ ¯ metals 8 a pa ͑ ͒ ͑ ͒ ͑ ͒ ␻ ␦ ␦ rTM 0,kЌ =1, rTE 0,kЌ =0. 31 Here c/ p = 0 is the skin depth and 0 /a is a small pa- rameter. The finite conductivity correction in Eq. ͑34͒ is In the limit of infinite conductivity one again obtains Eq. ͑ ͒ ജ negligibly small. Thus, the high-temperature limit for 30 for the reflection coefficients with l 1, but for the real metals described by the plasma model dielectric ͑ ͒ reflection coefficients with l=0 Eq. 31 remains valid. permittivity ͑32͒ coincides with that obtained for ideal This would lead to a different result for ideal metals metals ͑Brown and Maclay, 1969; Feinberg et al., 2001͒. than is obtained from the Schwinger prescription. As The agreement between the Lifshitz theory combined shown below, this contradiction arises from the use of with the plasma model and the case of ideal metals, as ϳ ␰ the dielectric permittivity 1/ which is outside the ap- described by thermal quantum field theory, has a simple plicability conditions of the Lifshitz theory. explanation. It is because for the plasma model with Conceptually the Lifshitz theory provides a way for ␻ →ϱ ͑ ͒ ͑ ͒ p , it is not Eq. 31 but Eq. 30 that is satisfied for obtaining all the necessary results for any real material. the reflection coefficients at zero frequency. One may describe the free electrons in metals by the A different situation occurs when the Drude dielectric ͓ plasma model as suggested in the first papers devoted function to the calculation of the finite conductivity corrections to ͑ ␧ ͑␻͒ ␻2 ␻͓␻ ␥͑ ͔͒ ͑ ͒ the Casimir result Lifshitz, 1956; Hargreaves, 1965; D =1− p/ + i T , 35 Schwinger et al., 1978͔͒, where ␥͑T͒ is the relaxation parameter, is substituted ␧͑␻͒ ␻2 ␻2 ͑ ͒ ͑ =1− p/ . 32 into the Lifshitz formula Boström and Sernelius, 2000; Høye et al., 2003, 2006; Brevik et al., 2005͒. In this case ␻ Here p is the plasma frequency. The dielectric permit- the zero-frequency values of the reflection coefficients ͑ ͒ tivity 32 is applicable in the frequency region of infra- are given by Eq. ͑31͒, and the high-temperature limit is ͑ ͒ red optics Lifshitz and Pitaevskii, 1981 . At room tem- equal to one-half of the corresponding value for ideal perature all nonzero Matsubara frequencies belong to metals, regardless of how large a conductivity is used for this region. Because of this, the plasma model leads to the real metal ͑Klimchitskaya and Mostepanenko, 2001͒. rather accurate results for the Casimir force at separa- The low-temperature thermal correction, calculated tion distances above the plasma wavelength. Below, the with the dielectric permittivity ͑35͒, can be presented in generalized plasmalike dielectric permittivity is also con- the form ͑Bezerra et al., 2004͒ sidered, which takes into account dissipation due to in- ⌬ F͑D͒͑ ͒ terband transitions of core electrons and can be applied T a,T at shorter separations ͑see Sec. III.A.1͒. The results ob- ϱ k T tained for real metals using the Lifshitz theory with the = ⌬ F͑p͒͑a,T͒ + F͑␥͒͑a,T͒ − B ͵ ydy T ␲ 2 permittivity ͑32͒ are qualitatively close to those obtained 16 a 0

for ideal metals using thermal quantum field theory ͱ 2 2 2 2 2 ͑ ͒ cy − 4a ␻ + c y Bordag et al., 2000b; Genet et al., 2000 . Thus, in the ϫlnͫ1−ͩ p ͪ e−yͬ. ͑36͒ Ӷ ͑ ប ͒ ͱ 2␻2 2 2 low-temperature limit T Teff kBTeff= c/2a , the cy + 4a p + c y asymptotic expression for the thermal correction to the With the condition that energy for two similar plates, as defined in Eq. ͑15͒,is ͑ ͒ ␥͑ ͒ Ͻ ␰ ͑ ͒ given by Bordag et al., 2000b; Geyer et al., 2001 T 1, 37 ␲2បc 45␨͑3͒ T 3 T 4 ⌬ F͑p͒͑ ͒ ͭ ͩ ͪ ͩ ͪ the low-temperature behavior of the second term on the a,T =− 1+ − ͑␥͒ T 3 ␲3 ͑ ͒ F ͑ ͒ϳ␥͑ ͒ ͑ ͒ 720a Teff Teff right-hand side of Eq. 36 is a,T T ln T/Teff ͑Bezerra et al., 2004͒. For metals with perfect crystal lat- ␨͑ ͒ 3 4 2c 45 3 T 1 T ␥͑ ͒→ 2 ͑ ͒ + ͫ ͩ ͪ − ͩ ͪ ͬ tices T 0asT Lifshitz et al., 1973 and the inequal- ␻ ␲3 pa Teff 2 Teff ity ͑37͒ is satisfied down to T=0. For metals with impu- ␥  ͑ rities there exists some residual value res 0 Kittel, ͮ ͑ ͒ ͒ −4 + ¯ . 33 1996 . As a result, at a T of about 10 K typical residual ␰ relaxation becomes equal to 1 and for smaller T Eq. Here ␨͑z͒ is the . Note that in this ͑37͒ is violated. case the quantity ⌬F͑a,T͒, as defined in Eq. ͑14͒, coin- For room temperature, at separation distances below cides with the thermal correction ͑15͒. For an ideal 1 ␮m the third term on the right-hand side of Eq. ͑36͒ is

Rev. Mod. Phys., Vol. 81, No. 4, October–December 2009 1838 Klimchitskaya, Mohideen, and Mostepanenko: The Casimir force between real materials: … dominant. It describes a rather large linear-in- strated that the application region of the obtained ex- temperature thermal correction to the Casimir free en- pressions is at temperatures below T=10−5 K. ergy and pressure, as predicted by the Drude model. According to Høye et al. ͑2007͒ and Brevik et al. This thermal correction is positive, i.e., it decreases the ͑2008͒, the formal satisfaction of the Nernst theorem for magnitude of the Casimir free energy and pressure. As the Drude model, as applied to metals with impurities, shown in Sec. IV.B, such a thermal correction is ex- resolves the problem of thermodynamic consistency of cluded by experiments at a high confidence level. that model combined with the Lifshitz theory. Klim- From a theoretical basis the thermal correction ͑36͒ is chitskaya and Mostepanenko ͑2008a͒ remarked that a not appropriate because it is inconsistent with thermo- perfect crystal lattice is a truly equilibrium system with a dynamics for metals with perfect crystal lattices. To nondegenerate dynamical state of the lowest energy. At verify the consistency with thermodynamics, one should zero temperature any part of the system must be in the consider the Casimir entropy ͑Bezerra et al., 2002a, ground ͑Landau and Lifshitz, 1980; 2002b, 2004͒ Rumer and Ryvkin, 1980͒. As a result, the Casimir en- tropy computed for a perfect crystal lattice must be ͒ ͑ ץ ͒ F͑ץ ͒ ͑ S a,T =− a,T / T. 38 equal to zero, a direct consequence of quantum statisti- cal physics. Thus, they pointed out that the Drude model In the case of metals described by the plasma model ͑ ͒ combined with the Lifshitz formula violates the Nernst dielectric permittivity ͑32͒ S p ͑a,T͒ϳT2 →0 when T heat theorem for a perfect crystal lattice and is in con- →0 ͑Bezerra et al., 2002b͒. In the case of Drude metals tradiction with quantum statistical physics. with perfect crystal lattices, the Casimir entropy has a Høye et al. ͑2008͒ argued that the Drude model com- nonzero negative value at T=0 following from Eq. ͑36͒ bined with the Lifshitz formula is thermodynamically ͑Bezerra et al., 2002a, 2002b, 2004, 2006; Mostepanenko, consistent in the case of a perfect crystal lattice also. To Bezerra, et al., 2006͒ justify this statement they introduced the definition of a ϱ perfect lattice as the limiting case of a crystal lattice with ͑ ͒ kB S D ͑a,0͒ = ͵ ydy nonzero residual relaxation ␥ when ␥ →0. However, 16␲a2 res res 0 in condensed matter physics ͑Kittel, 1996͒, a perfect lat- cy − ͱ4a2␻2 + c2y2 2 tice is commonly defined as a lattice with perfect crystal ϫlnͫ1−ͩ p ͪ e−yͬ structure without impurities, i.e., with ␥ =0 from the ͱ 2␻2 2 2 res cy + 4a p + c y outset. This common definition allows one to develop ␨͑ ͒ 2 the quantum theory of electron-phonon interactions, kB 3 4c c =− ͩ1− +12 − ͪ which is based on the properties of the lattice. 16␲a2 ␻ a ␻2 2 ¯ p pa Specifically, the relaxation parameter turns out to be Ͻ 0. ͑39͒ temperature dependent and vanishes with vanishing temperature. The definition of a perfect lattice intro- Note that the use of the Drude model in the Lifshitz duced by Høye et al. ͑2008͒ violates lattice symmetry formula is usually justified by the fact that it provides properties. For example, it would not be possible to re- smooth transition between the region of the normal skin produce the standard results of the theory of elementary effect and that of infrared optics. This ignores the region excitations on the basis of this definition. Recently, In- of the anomalous where the Lifshitz formula travaia and Henkel ͑2008͒ independently verified that in terms of ␧͑␻͒ is not applicable. With the decrease of T the Drude model used in the Lifshitz formula violates the application region of the normal skin effect becomes the Nernst heat theorem for metals with perfect crystal more narrow and the application region of the anoma- lattices as commonly understood ͓␥͑T͒→0atT→0͔,in lous skin effect widens. However, at any TϾ0, there ex- accordance with original statement by Bezerra et al. ists a frequency region including zero frequency where ͑2002a, 2002b, 2004͒. the normal skin effect is applicable. Thus, it is worth- Note that the substitution of the Drude dielectric while to use the Drude model at low T when discussing function ͑35͒ into the Lifshitz formula ͑3͒ for metal the thermodynamic consistency of the Lifshitz theory. plates of finite size is somewhat contradictory ͑Parse- It was shown ͑Boström and Sernelius, 2004; Høye et gian, 2005; Geyer et al., 2007; Mostepanenko and Geyer, al., 2007; Brevik et al., 2008͒ that for Drude metals with 2008͒. As pointed out by Parsegian ͑2005͒, the Drude impurities the Casimir entropy jumps abruptly to zero at model is derived from the Maxwell equations applied to TϽ10−4 K starting from a negative value ͓Eq. ͑39͔͒. The an infinite metallic medium ͑semispace͒ with no external corresponding analytic expression for the low- sources, with zero induced charge density, and with non- temperature behavior of the entropy has been obtained zero drift current. In such a medium there are no walls ͑Høye et al., 2007; Brevik et al., 2008͒ under the condi- limiting the flow of charges. For real metal plates of fi- ␨ Ӷ␥ ͑ ͒ tion l res, which is opposite to the inequality 37 . The nite size, however, the applicability conditions of the application region and the coefficients of this analytic Drude model are violated. If one admits that the electric expression were determined incorrectly because an field of the zero-point oscillations with vanishingly small ␥ ␥͑ ͒ ͑ ͒ overestimated value of res= T=300 K was used. The frequencies almost zero creates a short-lived current of correct values of the coefficients were found by Klim- conduction electrons, this leads to the formation of al- chitskaya and Mostepanenko ͑2008a͒, who also demon- most constant charge densities ±⌺ on the sidewalls of

Rev. Mod. Phys., Vol. 81, No. 4, October–December 2009 Klimchitskaya, Mohideen, and Mostepanenko: The Casimir force between real materials: … 1839

the Casimir plates. As a result, both the electric field and efficient in accordance with Eq. ͑40͒ leads to the viola- the current inside the plate will vanish due to screening, tion of the Nernst heat theorem for those dielectric ma- in a very short time interval of order of 10−18 s ͑Geyer et terials whose charge carrier density does not vanish al., 2007͒. Because of this behavior, for electromagnetic when T→0 and conductivity vanishes due to the vanish- fields at very low frequency, the finite metal plates ing mobility ͑Klimchitskaya, Mohideen, and Mostepa- should be described not by the Drude dielectric function nenko, 2008; Klimchitskaya, 2009͒. For metals with per- ͑35͒, but by the plasma model, which allows for displace- fect crystal lattices this approach violates the Nernst ment currents only ͑see Sec. III.A͒. The role of finite theorem as discussed above ͑Mostepanenko et al., 2009͒. size effects of the conductors was also illustrated in the Note that Dalvit and Lamoreaux ͑2008͒ applied their case of two wires of finite length described by the Drude approach to intrinsic semiconductor media. Bearing in model and interacting through the inductive coupling mind, however, that the classical Drude dielectric per- between Johnson currents ͑Bimonte, 2007͒.Itwas mittivity is exploited, the method can be applied for shown that in the thermal interaction between the wires metals and doped semiconductors as well by substituting the Nernst theorem is followed or violated depending on the corresponding screening length. As shown in Secs. whether the capacitive effects associated with the end IV.B and V. B , the computational results using the modi- points of the wires are taken or not taken into account. fied TM reflection coefficient are excluded by the mea- Another approach to the resolution of the puzzles surement data of two precise experiments ͑Chen et al., arising in the application of the Lifshitz theory to real 2007a, 2007b; Decca et al., 2007a, 2007b͒. materials includes the effect of screening on the reflec- As shown above, the description of real metals at low tion coefficients ͑Dalvit and Lamoreaux, 2008͒. This ap- frequencies by the seemingly obvious dielectric permit- proach takes into account both the drift and diffusion tivity of the Drude model leads to serious difficulties. currents of free charge carriers through the use of the Attempts at solving this problem by taking into account transport Boltzmann equation. It uses the standard Lif- the screening effects due to finite size of the plates or shitz formulas ͑3͒ and ͑21͒ for the free energy of wall- diffusion currents of free charge carriers also failed. The wall and atom-wall interactions with the TE reflection latter approach, if successful, would in fact be a recogni- ͑ ͒ coefficient rTE as defined in Eq. 6 within the Drude tion that the standard reflection coefficients, as used in model approach, but with the modified TM reflection the Lifshitz theory, expressed in terms of the dielectric coefficient permittivity are incorrect and one of them must be 2 modified using an additional microscopic quantity, the kЌ ␧͑ix͒ − ␧ ͑ix͒ ␧͑ ͒ c density of free charge carriers. ix q − k − ␩͑ ͒ ␧ ͑ ͒ mod ix c ix r ͑ix,kЌ͒ = . ͑40͒ The origin of the problem can be understood if one TM k2 ␧͑ix͒ − ␧ ͑ix͒ ␧͑ ͒ Ќ c takes into consideration the fact that the inclusion of the ix q + k + ␩͑ ͒ ␧ ͑ ͒ ix c ix drift current, as described by the Drude model, violates the basic applicability condition of the Lifshitz theory. Here the dielectric permittivity is given by This is the demand that the dielectric media and the ␧͑ ͒ ␧ ͑ ͒ ␻2 ͑ ␥͒ ͑ ͒ fluctuating electromagnetic field should be in thermal ix = c ix + p/x x + , 41 equilibrium with the heat reservoir. The point is that the ␧ ͑ ͒ where c ix is the permittivity of the bound core elec- existence of the drift electric current leads to a violation trons, of time-reversal symmetry and the introduction of the interaction between the system and the heat reservoir gj ␧ ͑ix͒ =1+͚ , ͑42͒ violating the state of thermal equilibrium ͑Bryksin and c ␻2 + x2 + ␥ x j j j Petrov, 2008͒. This is seen from the fact that the Drude ␻  model is derived from Maxwell equations with the drift with nonzero oscillator frequencies j 0, oscillator strengths g , and damping parameters ␥ . The quantity current of conduction electrons on the right-hand side j j ͑ ͒ ␩͑ix͒ is defined as see above . In its turn, the drift current leads to the emergence of Joule losses, i.e., to a heating of the crystal ␧ ͑0͒ ␧͑ix͒ 1/2 ͑ ͒ ␩͑ ͒ ͩ 2 ␬2 c ͪ ͑ ͒ lattice Geyer et al., 2003 . In order to preserve the tem- ix = kЌ + ␧ ͑ ͒ ␧͑ ͒ ␧ ͑ ͒ , 43 c ix ix − c ix perature constant, the unidirectional flux of heat from the Casimir plates to the heat reservoir must inevitably ␬ where 1/ is the screening length. If the charge carriers arise. Such an interaction between the system and the of density n are described by the classical Maxwell- heat reservoir cannot exist in the state of thermal equi- ͓ ͑ ͔͒ Boltzmann statistics Dalvit and Lamoreaux 2008 , one librium as it is in contradiction with its definition ␬ gets the Debye-Hückel screening length RD =1/ ͑Rumer and Ryvkin, 1980; Kondepudi and Prigogine, ͑␧ ␲ 2 ͒1/2 = 0kBT/4 e n . Assuming Fermi-Dirac statistics, 1998͒. one arrives at the Thomas-Fermi screening length RTF Thus not just any dielectric permittivity can be consid- ␬ ͑␧ ␲ 2 ͒1/2 ͑ =1/ = 0EF /6 e n Landau and Lifshitz, 1980; ered in combination with the Lifshitz formula. Specifi- ͒ ␧ Chazalviel, 1999 . Here e is the electron charge, 0 cally, the dielectric permittivity of the Drude model ͑or ␧ ͑ ͒ = c 0 is the dielectric constant due to the core elec- any other that is inversely proportional to the first power ប␻ ͒ trons, and EF = p is the Fermi energy. It was shown, of the frequency at low frequencies does not belong to however, that the modification of the TM reflection co- the application region of the Lifshitz theory. In a similar

Rev. Mod. Phys., Vol. 81, No. 4, October–December 2009 1840 Klimchitskaya, Mohideen, and Mostepanenko: The Casimir force between real materials: … way, the approach based on the transport Boltzmann ␧͑␻ ͒ ␧͑␻͒ ␲␴ ͑ ͒ ␻ ͑ ͒ ˜ ,T = + i4 0 T / , 47 equation, in addition to the drift current, introduces the ␴ ͑ ͒  ␧͑␻͒ diffusion current determined by a nonequilibrium distri- where 0 T is the dc conductivity at T 0 and is bution of free charge carriers. This exacerbates the vio- given by Eq. ͑44͒͑Palik, 1985; Zurita-Sánches et al., lation of thermal equilibrium. Thus, the modified reflec- 2004͒. Then, instead of Eq. ͑46͒, the reflection coeffi- tion coefficient ͑40͒ obtained from this approach cannot cients at ␰=0 are given by Eq. ͑31͒. be substituted into the Lifshitz formula. On the other The analysis of the low-temperature asymptotic be- hand, the dielectric permittivity of the plasma model havior of the Casimir free energy ͑3͒ calculated with the ͓Eq. ͑32͔͒ and its generalized version taking into account dielectric permittivity ͑44͒ for two similar plates shows the interband transitions ͓see Eq. ͑58͒ in Sec. III.A.1͔ do that ͑Geyer et al., 2005b, 2006; Klimchitskaya and Geyer, not admit the drift current but only the displacement 2008͒ one. Thus, they are compatible with the demand of ther- បc b͑a͒Li ͑r2͒ T 2 mal equilibrium and can be substituted into the Lifshitz F͑a,T͒ = E͑a͒ − ͫ 2 0 ͩ ͪ 3 ͑␧2 ͒ formulas. 8a 3 0 −1 Teff ␨͑3͒r2͑␧ +1͒ 3 0 0 ͩ T ͪ ͬ ͑ ͒ + ␲ , 48 2. Real dielectrics 4 Teff ͑ ͒ As mentioned in the Introduction, puzzling results where Lin z is the polylogarithm function, and the fol- were obtained in the application of the Lifshitz theory lowing notation is used: to real dielectrics. It is well known that all dielectric ma- g ␥ c terials ͑insulators and intrinsic semiconductors charac- b͑a͒ = ͚ j j . ͑49͒ ␻4 terized by a band gap of different width, Mott-Hubbard j 2a j dielectrics, and doped semiconductors with doping con- From Eq. ͑48͒ the Casimir entropy is given by centration below the critical value͒ have a zero value ͓ ͑ ͒ ͑ 2͒ of the dc conductivity at zero temperature see, e.g., kB T 2b a Li2 r0 ͑ ͒ ͑ ͔͒ S͑a,T͒ = ͩ Shklovskii and Efros 1984 and Mott 1990 . At non- 4a2 T 3͑␧2 −1͒ zero temperature these materials have a nonzero value eff 0 3␨͑3͒r2͑␧ +1͒ T of dc conductivity. Sometimes this conductivity is rather 0 0 ͪ ͑ ͒ + ␲ . 50 large, as, for instance, for doped semiconductors with a 4 Teff doping concentration only slightly below the critical value. In the commonly used application of the Lifshitz As is seen from Eq. ͑50͒, S͑a,T͒ goes to zero when T theory the dc conductivity of dielectrics is neglected. →0, i.e., the Nernst heat theorem is satisfied when the Their conductivity is assumed to be equal to zero at any dielectric permittivity is given by Eq. ͑44͒ with a finite temperature. In this case the dielectric permittivity is static value ͓Eq. ͑45͔͒. represented in the form ͓see, e.g., Parsegian ͑2005͔͒ A completely different result is obtained if one takes into account the conductivity of dielectrics at zero fre- Ͼ gj quency that arises at T 0. Note that in all cases this ␧͑␻͒ = ␧ ͑␻͒ =1+͚ , ͑44͒ ␴ ͑ ͒ϳ ͑ ͒ c ␻2 − ␻2 − i␥ ␻ conductivity vanishes exponentially, 0 T exp −C/T , j j j when T→0 ͑Shklovskii and Efros, 1984; Mott, 1990͒. The reflection coefficients at zero frequency calculated where ␻ 0 are the oscillator frequencies, g are the j j with the dielectric permittivity ͑47͒ do not satisfy Eq. oscillator strengths, and ␥ are the damping parameters j ͑46͒ but Eq. ͑31͒. This means that the TM reflection co- ͓note that when ␧͑i␰͒ is calculated along the imaginary efficient has a discontinuity when the dc conductivity of frequency axis the damping parameters are often ne- the dielectric material is taken into account. Substituting glected and some approximate formulas are used the dielectric permittivity ͑47͒ into the Lifshitz formula ͑Hough and White, 1980; Bergström, 1997; Parsegian, ͑3͒, we arrive at ͑Geyer et al., 2005b͒ 2005͔͒. From Eq. ͑44͒, the dielectric permittivity at zero frequency is given by k T F˜ ͑a,T͒ = F͑a,T͒ − B ͓␨͑3͒ −Li ͑r2͔͒ + R͑T͒, ͑51͒ 16␲a2 3 0 g ␧ ϵ ␧͑0͒ =1+͚ j Ͻϱ ͑45͒ 0 ␻2 where the low-temperature behavior of F͑a,T͒ is given j j by Eq. ͑48͒ and R͑T͒ decreases exponentially when T and the reflection coefficients at zero frequency are vanishes. From Eq. ͑51͒, the Casimir entropy at T=0 is given by ␧ −1 ͑ ͒ 0 ϵ ͑ ͒ ͑ ͒ S˜ ͑a,0͒ = ͑k /16␲a2͓͒␨͑3͒ −Li ͑r2͔͒ Ͼ 0, ͑52͒ rTM 0,kЌ = ␧ r0, rTE 0,kЌ =0. 46 B 3 0 0 +1 in violation of the Nernst theorem. Thus, the inclusion However, if one takes into account the existence of free of a nonzero conductivity arising at TϾ0 into the model charge carriers at nonzero temperature, the dielectric of the dielectric response leads to a contradiction be- permittivity can be represented by tween the Lifshitz theory and thermodynamics for all

Rev. Mod. Phys., Vol. 81, No. 4, October–December 2009 Klimchitskaya, Mohideen, and Mostepanenko: The Casimir force between real materials: … 1841

materials which have a zero conductivity at zero tem- all nonzero Matsubara frequencies. If one takes into ac- perature, i.e., for all dielectrics. The same happens if we count the fact that the reflection coefficient ͑53͒ ob- include the dc conductivity of the dielectric at TϾ0 tained by Pitaevskii ͑2008͒ is just the reflection coeffi- when describing the Casimir force between a metal and cient ͑40͒ at x=0, this concludes the proof of the above a dielectric plate ͑Geyer et al., 2005a, 2006; Klim- assumption. chitskaya, Geyer, and Mostepanenko, 2006; Geyer, The reflection coefficient ͑53͒ can be obtained by the Klimchitskaya, and Mostepanenko, 2008͒. In this case formal replacement of the dielectric permittivities de- the Casimir entropy at T=0 is given by Eq. ͑52͒ where pending only on frequency with the permittivities de- ͑ 2͒ ͑ ͒ Li3 r0 should be replaced with Li3 r0 . The above puz- pending on both frequency and wave vector in the zling results are in line with those obtained for real met- Fresnel reflection coefficient of an uniaxial crystal. To als described by the Drude model. They are explained confirm that this is the case, one should introduce dis- ␧ ␧ ␧ by the fact that the dielectric permittivity ͑47͒ is outside similar permittivities x = y and z and replace the coef- ͑ ͒ the application region of the Lifshitz theory. Thus, the ficient rTM in Eq. 6 with Ͼ nonzero conductivity arising at T 0 for all dielectric ͱ␧ ͑ ͒␧ ͑ ͒ mod x ix z ix q − kz materials must be disregarded in theoretical computa- r ͑ix,kЌ͒ = . ͑54͒ ͑ ͒ TM ͱ␧ ͑ ͒␧ ͑ ͒ tions using the Lifshitz formula 3 . This rule has already x ix z ix q + kz ͑ been confirmed experimentally see Secs. V. B and ͑ ͒ ͒ Here kz is defined by Eq. 7 with the replacement of VI.A . ␧͑ ͒ ␧ ͑ ͒ ␧ ␧ ix for z ix . With x and z depending only on the Similarly to the case of two dielectric plates, for an ͑ ͒ atom near a dielectric wall the Lifshitz theory is thermo- frequency, Eq. 54 is commonly used for the description of uniaxial crystals ͑Greenaway et al., 1969; Blagov et al., dynamically consistent if the dc conductivity is ne- ͒ ͑ ͒ ͑ ͒ glected. However, the inclusion of the dc conductivity of 2005 . Then Eq. 53 follows from Eq. 54 if one sets the wall material in accordance with Eq. ͑47͒ leads to the ␧ ͑ ͒ ␧ ␧ ͑ ͒ ␧ ͑ ͒ ␧ ͑ ␬2 2 ͒ ͑ ͒ x 0 = 0, z 0 = z 0,kЌ = 0 1+ /kЌ , 55 violation of the Nernst heat theorem ͑Klimchitskaya, ␧ ͑ ͒ Mohideen, and Mostepanenko, 2008͒. An attempt to where now z 0 depends on the wave vector. Thus, to solve this problem beyond the scope of the standard Lif- obtain Eq. ͑53͒ one substitutes into the reflection coeffi- shitz theory was undertaken by Pitaevskii ͑2008͒. This cient ͑54͒ derived from the standard continuity bound- attempt makes use of the effect of Debye-Hückel ary conditions the dielectric permittivity depending on a screening mentioned above in connection with metals. wave vector ͑see Sec. III.A.3 for reasons why this is in- The effect of screening is taken into account only for the appropriate͒. static field. It leads to the replacement of the transverse It was verified ͑Klimchitskaya, Mohideen, and magnetic reflection coefficient at zero frequency, as Mostepanenko, 2008͒ that for pure insulators and intrin- given by Eq. ͑46͒, with sic semiconductors ͓i.e., for materials for which n͑T͒ ex- ponentially decays to zero with vanishing temperature͔ mod͑ ͒ ͑␧ ͱ 2 ␬2 ͒ ͑␧ ͱ 2 ␬2 ͒ rTM 0,kЌ = 0 kЌ + − kЌ / 0 kЌ + + kЌ , the Lifshitz theory with the modified reflection coeffi- ͑ ͒ ͑53͒ cient 53 satisfies the Nernst heat theorem. However, if n does not go to zero when T goes to zero ͑this is true, ␬ where =1/RD is defined in Sec. II.D.1. When the total for instance, for dielectric materials, such as semicon- density of charge carriers n=0, Eq. ͑53͒ leads to the ductors doped below critical doping concentration or same result as Eq. ͑46͒. For n→ϱ, at fixed T0, solids with ionic conductivity͒, the Lifshitz theory with mod͑ ͒ ͑ ͒ rTM 0,kЌ =1, as in the standard Lifshitz theory with the the reflection coefficient 53 leads to a positive Casimir- inclusion of the dc conductivity. This solves the peculiar- Polder entropy at T=0, i.e., violates the Nernst heat ity of the small-␴ limit, i.e., the abrupt jump from Eq. theorem ͑Klimchitskaya, Mohideen, and Mostepanenko, ͑46͒ to Eq. ͑31͒ arising from an arbitrarily small conduc- 2008͒. Note here that the entropy of the fluctuating elec- tivity. According to Pitaevskii ͑2008͒, “the above- tromagnetic field is nonzero at T=0 and depends on mentioned peculiarity of the small ␴ limit exists only for separation, whereas the entropy of the plates is separa- the l=0 term.” This permits one to assume that all the tion independent. Because of this, the entropy of the ͑ ␰ ͒ ജ coefficients rTM,TE i l ,kЌ with l 1 under some condi- plates at T=0 cannot compensate the nonzero entropy tions remain unchanged. This assumption was confirmed of the fluctuating field so as to make the total entropy by Dalvit and Lamoreaux ͑2008͒, who obtained both the independent of the parameters of the system in accor- TM and TE reflection coefficients with account of the dance with the Nernst theorem. In fact, the conductivity ␴ ͑ ͒ ͉ ͉␮ ␮ screening effects at arbitrary Matsubara frequencies. is given by 0 T =n e , where is the mobility of The TE reflection coefficient was found to coincide with charge carriers ͑Ashcroft and Mermin, 1976͒. Although ␴ ͑ ͒ the standard one obtained using the Drude model ap- 0 T goes to zero exponentially fast for all dielectrics proach and the modified TM reflection coefficient is when T goes to zero, for most of them this happens not given by Eq. ͑40͒. It is easily seen that for materials with due to the vanishing n but due to the vanishing mobility. ͑ ͑ ͒ sufficiently small charge carrier density for instance, for One such example is fused silica SiO2 considered in Si with nഛ1017 cm−3͒ the modified reflection coefficient Sec. VI.A in connection with experiments on the ͓Eq. ͑40͔͒ at room temperature T=300 K takes approxi- Casimir-Polder force. Its conductivity is ionic in nature mately the same values as the standard one in Eq. ͑6͒ at and is determined by the concentration of impurities

Rev. Mod. Phys., Vol. 81, No. 4, October–December 2009 1842 Klimchitskaya, Mohideen, and Mostepanenko: The Casimir force between real materials: …

͑alkali-metal ions͒ which are always present as trace ͓see, e.g., Hough and White ͑1980͒, Bergström ͑1997͒, constituents. Two more examples are doped semicon- and Parsegian ͑2005͔͒ suggested for the calculation of ductors with dopant concentration below critical and di- the at short separation distances in electriclike semimetals. the region of the nonretarded van der Waals force. As shown in Sec. V. B , the modification of the trans- These expressions can also be used for calculations of verse magnetic reflection coefficient at zero frequency in the retarded Casimir forces using the Lifshitz formula accordance with Eq. ͑53͒ is in contradiction with the ͑3͒͑Parsegian, 2005͒. data of the experiment on the optically modulated Ca- The description of the dielectric properties of metals simir forces ͑Chen et al., 2007a, 2007b͒. This is in line is much more complicated. In calculations of the zero- with the fact that, being the particular case of Eq. ͑40͒, temperature Casimir energy and pressure ͓Eqs. ͑12͒ and the modified reflection coefficient ͑53͒ violates the appli- ͑13͔͒ the plasma model dielectric permittivity ͑32͒ is of- cability conditions of the Lifshitz theory. ten used, following Lifshitz ͑1956͒. This model is, how- ever, applicable only at separation distances larger than the plasma wavelength, as it does not take into account III. HOW TO COMPARE THEORY AND EXPERIMENT the absorption bands due to the core electrons. To ad- A. Modeling of the optical properties of real materials dress this problem, there are two approaches suggested in the literature. In the first approach the tabulated op- 1. Kramers-Kronig relations tical data for the complex index of refraction of the To compare theory and experiment one first needs re- metal under consideration are used to obtain the imagi- liable theoretical results. These results can be obtained nary part of the dielectric permittivity. The latter is ex- ͑ ͒ trapolated to the region of low frequencies with the help using the Lifshitz formula 3 with an appropriate dielec- ͑ tric permittivity for the plate material along the imagi- of the Drude dielectric function Lamoreaux, 1999; Klimchitskaya et al., 2000; Lambrecht and Reynaud, nary frequency axis. In general, it can be obtained from ͒ the imaginary part of complex dielectric permittivity 2000 . In the second approach the core electrons are de- ␧͑␻͒=␧Ј͑␻͒+i␧Љ͑␻͒ and the Kramers-Kronig relations. scribed by a set of oscillators with nonzero oscillator frequencies ͓similar to Eq. ͑44͒͑Jackson, 1999͔͒ but the The imaginary part of the permittivity is ␧Љ͑␻͒ free conduction electrons are described using the plasma =2nЈ͑␻͒nЉ͑␻͒, where nЈ͑␻͒ and nЉ͑␻͒ are the real and model ͑Decca et al., 2007b; Geyer et al. 2007; Klim- imaginary parts of the complex index of refraction. The chitskaya et al., 2007a; Mostepanenko and Geyer, 2008͒. tabulated optical data for nЈ͑␻͒ and nЉ͑␻͒ for the differ- In this case we obtain the so-called generalized plasma- ent materials can be found in Palik ͑1985͒. It should be like dielectric permittivity, noted that the form of Kramers-Kronig relations is dif- ␻2 ␻2 ferent depending on the analytic properties of the di- p p gj ␧͑␻͒ ␧ ͑␻͒ = ␧ ͑␻͒ − =1− + ͚ . electric permittivity under consideration. If is a gp c ␻2 ␻2 ␻2 − ␻2 − i␥ ␻ regular function at ␻=0 ͓the case of dielectrics with di- j j j electric permittivity given by Eq. ͑44͔͒, the Kramers- ͑58͒ ͑ ͒ Kronig relations take the simplest form Jackson, 1999 , The values of the oscillator parameters of core electrons ϱ 1 ␧Љ͑␰͒ are found from the tabulated optical data for the com- ␧Ј͑␻͒ =1+ P͵ d␰, plex index of refraction ͑Palik, 1985͒. Since the Casimir ␲ ␰ − ␻ −ϱ effect is a broadband phenomenon, it is of utmost im- ϱ portance to use as exact data as possible, determined 1 ␧Ј͑␰͒ ␧Љ͑␻͒ ͵ ␰ ͑ ͒ over a wide frequency range in the interband absorption =−␲P ␰ ␻d , 56 −ϱ − region ␻Ͼ2.5 eV. Presently such data ͑and also data for lower frequencies͒ can be obtained using an ellipsom- where the integrals are understood as a principal value. eter. Thus, Svetovoy et al. ͑2008͒ found precise optical The third dispersion relation, expressing the dielectric properties of several Au films in the frequency region permittivity along the imaginary frequency axis, in this from 0.73 to 8.86 eV using a vacuum ultraviolet ellip- case is someter ͑and from 0.038 to 0.65 eV using an infrared ϱ 2 ␻␧Љ͑␻͒ variable-angle spectroscopic ellipsometer͒. ␧͑i␰͒ =1+ ͵ d␻. ͑57͒ ␲ ␰2 + ␻2 In both approaches the dielectric permittivity along 0 the imaginary frequency axis can be obtained by means The optical data for dielectrics are available in a wide of the Kramers-Kronig relations. Regardless of which ͓␻ ␻ ͔͑ ͒ ␻ frequency range min, max Palik, 1985 . Near min, approach is used, the dielectric permittivity of a metal is ␧ ͑␻͒ ␻ ␻ Љ is almost zero. Bearing in mind that min is usually not regular at =0 but has a pole. This changes the form below all the relevant resonances of the respective di- of the Kramers-Kronig relations ͑56͒.If␧͑␻͒ has a ␻ ␧͑␻͒Ϸ ␲ ␴ ␻ electric medium, it is possible to extrapolate ␧Љ͑␻͒ as simple pole at =0, 4 i 0 / , the first two ഛ␻ഛ␻ ␻ ͑ ͒ zero in the region 0 min. Regarding max, it is usu- Kramers-Kronig relations are Landau et al., 1984 ally sufficiently high that no extrapolation to region ␻ ϱ 1 ␧Љ͑␰͒ Ͼ␻ is required. There are also many approximate ␧Ј͑␻͒ ͵ ␰ max =1+␲P ␰ ␻d , analytic expressions for the dielectric permittivity ͑44͒ −ϱ −

Rev. Mod. Phys., Vol. 81, No. 4, October–December 2009

Klimchitskaya, Mohideen, and Mostepanenko: The Casimir force between real materials: … 1843





Ð Ó    ´   µ 

 Ù ¼ ½ 1 7 0.9 6 0.8 5 0.7 4 0.6

3 0.5

 µ 2 0.4 ´ 1 0.3

0.2 0.4 0.6 0.8 1

µ ´  Ñ

14 15 16 17 



½

È

Ð Ó    ´ Ö × µ  ¼ ½ 1 FIG. 1. The dielectric permittivity of Au along the imaginary 0.9 frequency axis. The solid line is obtained using the generalized 0.8 plasmalike model taking into account the optical data related 0.7 to the core electrons. The dashed line is obtained using the 0.6 optical data extrapolated to low frequencies by the Drude 0.5

model.

 µ 0.4 ´ 0.3 ϱ 1 ␧Ј͑␰͒ 4␲␴ 0.2 0.4 0.6 0.8 1

␧Љ͑␻͒ ␰ 0 ͑ ͒

´  Ñ µ ͵  =−␲P ␰ ␻d + ␻ . 59 −ϱ − FIG. 2. Correction factors to ͑a͒ the Casimir energy per unit The third dispersion relation in this case is given by Eq. ͑ ͒ ͑57͒, i.e., remains the same as for a dielectric permittivity area and b the pressure vs separation between Au plates. Bold and fine solid lines are computed using the generalized regular at ␻=0. plasmalike model at T=300 K and 0, respectively. Long- and We now consider the generalized plasmalike dielectric ͑ ͒ short-dashed lines are calculated using the extrapolation of op- permittivity 58 having a second-order pole at zero fre- tical data by the Drude model at T=300 K and 0, respectively. ͓ ␧͑␻͒Ϸ ␻2 ␻2 quency i.e., an asymptotic behavior − p / when ␻→0͔. The standard derivation ͑Landau et al., 1984; Jackson, 1999͒, when applied to this case, leads to one the complete tabulated optical data extrapolated to low more form of the Kramers-Kronig relations ͑Klim- frequencies by the imaginary part of the Drude dielec- ␻ ␥ ͑ chitskaya et al., 2007a͒, tric function with p =9.0 eV and =0.035 eV Klim- ͒ ϱ chitskaya et al., 2000; Lambrecht and Reynaud, 2000 . 1 ␧Љ͑␰͒ ␻2 ␧Ј͑␻͒ =1+ P͵ d␰ − p , The solid line in Fig. 1 is obtained using the Kramers- ␲ ␰ ␻ ␻2 ͑ ͒ −ϱ − Kronig relation 61 adapted for the generalized plasma- like model with ␧Љ͑␻͒ describing the contribution of core ␻2 electrons. In this case ␧Љ͑␻͒ is given by the tabulated ␧ ͑␰͒ p ϱ Ј + data with the contribution of free conduction electrons 1 ␰2 ␧Љ͑␻͒ ͵ ␰ ͑ ͒ subtracted ͑Decca et al., 2007b͒ using ␻ =9.0 eV. Note =−␲P ␰ ␻ d . 60 p −ϱ − that for frequencies ␰Ͻ1016 rad/s the simple analytic In this case the third Kramers-Kronig relation ͑57͒ is six-oscillator model can also be used for the description ͑ also replaced with ͑Klimchitskaya et al., 2007a͒, of core electrons Decca et al., 2007b; Mostepanenko and Geyer, 2008͒. As is seen in Fig. 1, the computational ϱ 2 2 ␻␧Љ͑␻͒ ␻ results for ␧͑i␰͒ using the two approaches coincide at ␰ ␧͑i␰͒ =1+ ͵ d␻ + p , ͑61͒ ␲ ␰2 ␻2 ␰2 ജ 15 0 + 10 rad/s but have different behavior at lower fre- quencies. i.e., it acquires an additional term. In Fig. 2 we present the computational results for the correction factors to the Casimir energy per unit ͑a͒ area 2. Calculation for Au and ͑b͒ pressure obtained using the dielectric permittivi- ties presented in Fig. 1 by the solid and dashed lines. We now discuss the application of the above two These correction factors are defined as approaches to calculations of the dielectric permittivity along the imaginary frequency axis for Au, a metal F͑a,T͒ P͑a,T͒ ␬ ͑a,T͒ = , ␬ ͑a,T͒ = , ͑62͒ widely used in the recent Casimir force measurements E E ͑a͒ P P ͑a͒ ͑see Secs. IV–VI͒. As discussed, both approaches use the 0 0 ͑ ͒ ͑ ͒ tabulated optical data for the complex index of refrac- where E0 a and P0 a are the Casimir energy and pres- tion ͑Palik, 1985͒. The calculation results are presented sure between ideal metal planes defined in Eq. ͑2͒. The in Fig. 1. The dashed line in Fig. 1 is obtained using solid lines represent the computational results using the the Kramers-Kronig relation ͑57͒ with ␧Љ͑␻͒ given by generalized plasmalike model ͓bold lines are obtained

Rev. Mod. Phys., Vol. 81, No. 4, October–December 2009 1844 Klimchitskaya, Mohideen, and Mostepanenko: The Casimir force between real materials: …

using Eqs. ͑3͒ and ͑10͒ at T=300 K whereas fine lines the frequency ͑Lifshitz et al., 1973͒. As discussed, the using the zero-temperature Eqs. ͑12͒ and ͑13͒;inthe drift current of conduction electrons is not compatible case of the Casimir pressure bold and fine solid lines with the Lifshitz theory of dispersion forces ͑Parsegian, practially coincide at all aഛ1 ␮m͔. The dashed lines cor- 2005; Geyer et al., 2007; Mostepanenko and Geyer, respond to computations using the tabulated optical 2008͒. Because of this, sample-to-sample variations of data extrapolated by the Drude model ͓the long-dashed scattering processes connected with this current should lines are obtained using Eqs. ͑3͒ and ͑10͒ at T=300 K be disregarded in the theory of the Casimir force. and the short-dashed lines using the zero-temperature Lifshitz formulas ͑12͒ and ͑13͒ with the dielectric permit- tivity at room temperature͔. As is seen in Figs. 2͑a͒ and 3. Problem of spatial dispersion 2͑b͒, there are large deviations between the zero- Another problematic issue is connected with the role temperature short-dashed and T=300 K long-dashed of spatial dispersion. In particular, it is well known that lines at all separation distances above 100 nm. This in the frequency region of the anomalous skin effect means that the zero-temperature formulas ͑12͒ and ͑13͒ metals cannot be described by the dielectric permittivity with substituted room-temperature values of the dielec- depending only on the frequency ͑Lifshitz et al., 1973; tric permittivity accounting for the relaxation processes Lifshitz and Pitaevskii, 1981͒. The Lifshitz formulas ͑3͒, of conduction electrons can only be approximately used ͑6͒, ͑10͒, ͑12͒, and ͑13͒, however, are derived only for at short separation distances, i.e., in the region of the materials that can be described by ␧͑␻͒. The application van der Waals forces and in the beginning of the transi- range of the Lifshitz theory can be widened using the tion region to the Casimir forces ͑Parsegian, 2005͒.At Leontovich surface impedance boundary condition larger separation distances Eqs. ͑12͒ and ͑13͒ do not re- ͑Landau et al., 1984͒. Kats ͑1977͒ pioneered the applica- produce the values of the Casimir free energy. Using this tion of this approach to describe metals in the region of model of the dielectric permittivity, all calculations at the anomalous skin effect at zero temperature. The Le- separations above 100 nm at room temperature should ontovich approach makes it possible to take into ac- be done with the thermal Lifshitz formulas ͑3͒ and ͑10͒ count real material properties without considering the for the Casimir free energy per unit area and pressure. interior of the plates. In this approach the standard con- We emphasize, however, that the theoretical results ob- tinuity boundary conditions applicable to interfaces be- tained in this way are excluded by an experiment at a tween spatially local media are not followed. The Leon- high confidence level ͑see Sec. IV.B͒. tovich impedance approach now appears interesting for The preceding discussion sheds light on the problem the description of the thermal Casimir force. It was sug- of sample-to-sample dependence of the optical data. Ac- gested ͑Bezerra et al., 2001; Geyer et al., 2003͒ to ex- tually, this dependence is mostly determined by the re- trapolate the impedance for the frequency region ␻ ͑ ͒ laxation processes of free conduction electrons at infra- around the characteristic frequency c =c/ 2a to all fre- red and optical frequencies ͓e.g., due to different quencies. For separation distances related to infrared characteristic sizes of grains in thin films ͑Sotelo et al., optics, some, however, have used the form of impedance 2003͔͒. Uncertainty of the Casimir pressure due to the for the normal ͑Torgerson and Lamoreaux, 2004͒ or the variation of optical properties was investigated by anomalous ͑Svetovoy and Lokhanin, 2003͒ skin effect. Pirozhenko et al. ͑2006͒ and Svetovoy et al. ͑2008͒ using In these latter cases the theoretical results turn out to be the zero-temperature Lifshitz formula ͑13͒. For this pur- in violation of the Nernst heat theorem and in contra- pose both the real and imaginary parts of the dielectric diction with experiment ͑Geyer et al., 2003, 2004; permittivity of Au, as given by the optical data in the Bezerra et al., 2007͒. infrared region collected by others, were approximated There is another approach to account for spatial dis- by the real and imaginary parts of the Drude permittiv- persion by the substitution of the dielectric permittivity ity with an additional polarization term. It is easily seen depending on both the frequency and wave vector, that all sets of the Drude parameters obtained are ex- ␧͑␻,k͒, into the usual Lifshitz formulas ͑3͒ and ͑8͒ or cluded by the experiment on the measurement of the ͑12͒ and ͑13͒ with some modified reflection coefficients. Casimir pressure ͓see the general method of the verifi- This approach is, however, unjustified. The review by cation of hypotheses, as applied to this case, in Sec. Barash and Ginzburg ͑1975͒ contains a few references to III.C.2 and Fig. 11͑b͒ in Sec. IV.B͔. In contrast, the gen- incorrect results obtained by others using such substitu- eralized plasmalike dielectric permittivity is based on tions. In the last few years, more papers have appeared the characteristics of materials determined by the struc- ͑Esquivel et al., 2003; Esquivel and Svetovoy, 2004; ͑␻ ͒ ture of a crystal cell p is one such factor . These char- Contreras-Reyes and Mochán, 2005; Sernelius, 2005; acteristics are not sensitive to sample-to-sample varia- Svetovoy and Esquivel, 2005͒ employing the same ap- tions of the optical properties of the deposited proach in an attempt to solve the problem resulting from ␻ polycrystal films. Large variations in the values of p using the Drude dielectric permittivity in the Lifshitz obtained by Pirozhenko et al. ͑2006͒ for different Au formula. It was demonstrated ͑Klimchitskaya and samples can be explained by the fact that the authors Mostepanenko, 2007͒ that these attempts are not war- used the Drude model with frequency-independent re- ranted because the modified reflection coefficients are laxation parameter, whereas at high frequencies the re- derived using phenomenological ͑additional͒ boundary laxation parameter of conduction electrons depends on conditions and dielectric permittivities ␧͑␻,k͒ which are

Rev. Mod. Phys., Vol. 81, No. 4, October–December 2009 Klimchitskaya, Mohideen, and Mostepanenko: The Casimir force between real materials: … 1845 ill-defined quantities in the presence of a gap between appropriately chosen origin on the z axis leads to the plates as it violates the assumption of translational ͗ ͑s͒͘ ͗ ͑ ͒͘ invariance in the z direction. In the presence of spatial z1 = A1 f1 x1,y1 =0, dispersion, the boundary conditions following from the ͗ ͑s͒͘ ͗ ͑ ͒͘ ͑ ͒ Maxwell equations differ from the standard continuity z2 = a + A2 f2 x2,y2 = a. 64 boundary conditions due to the inclusion of the induced charge and current densities ͑Agranovich and Ginzburg, As mentioned, we consider small deviations from the Ӷ 1984; Ginzburg, 1985͒. It was also shown that in the basic geometry, i.e., Ai a. ͑ ͒ presence of the boundary surface the dielectric permit- The topography of the surfaces 63 can be obtained ͑ ͒ tivities depending on ␻,k can be approximately used experimentally as an atomic force microscope AFM image of a rather large area ͑see, e.g., Fig. 8 in Sec. for the description of the remainder of the medium ex- ͒ cept for a layer adjacent to the boundary surface ͑Agra- IV.A . AFM images of the surfaces of test bodies used in novich and Ginzburg, 1984͒. However, it is unlikely that the experiments described in Secs. IV.A, IV.B, V, and this phenomenological approach would be applicable for VII show that the roughness is represented by stochas- the calculation of the Casimir force between metallic tically distributed distortions of different heights with a surfaces where the boundary effects are of prime impor- typical lateral correlation length of about 200–400 nm. If the correlation effects can be neglected ͑below we tance. The generalization of the Lifshitz formula in ͒ terms of the functional determinants and scattering ma- discuss why they are small , the surface topography is trices ͓see Eq. ͑17͔͒ opens opportunities to include the approximately characterized by a discrete set of pairs ͑ ͑1,2͒ ͑1,2͒͒ ͑1,2͒ effects of spatial dispersion. For the configuration of two vi ,hi , where vi is the fraction of the surface ͑1,2͒ plane parallel plates the matrix elements of the operator area with height hi on the surface 1 or 2, respectively. G͑0͒ ͑ ͒ These data allow one to determine the zero-roughness ␰,AB reduce to a factor exp −qa and the same holds for ͑ ͒ ͑0͒ 1,2 G ͑Lambrecht et al., 2006͒. To describe the plates in- levels H0 relative to which the mean values of the ␰,BA ͑ ͒ cluding the spatial dispersion, one should find the re- functions f1,2 in Eq. 63 are equal to zero, TA TB ͑ ͒ spective T matrices of the operators and . In prin- K k ciple, this could be done by solving Maxwell’s equations ͑ ͑k͒ ͑k͒͒ ͑k͒ ͑ ͒ ͚ H0 − hi vi =0, k = 1,2. 65 for a given Casimir configuration made of a nonlocally i=1 responding material without using the dielectric permit- tivity depending on ␻ and k. This problem awaits its Note that the separation distances between the test bod- resolution. ies in the Casimir force measurements are measured just between these zero-roughness levels. The Casimir pressure between the two ideal metal B. Corrections to the Casimir force due to the imperfect plates with rough surfaces ͓Eq. ͑63͔͒ at zero temperature geometry of interacting bodies can be represented in the form

1. Surface roughness ͑ ͒ ͑␲2ប 4͒␬͑r͒ ͑ ͒ P a =− c/240a P , 66 The problem of roughness corrections to dispersion ͑r͒ where ␬ is the correction factor due to surface rough- forces has long attracted the attention of researchers P ness. In the framework of the pairwise summation ͓see, e.g., Maradudin and Mazur ͑1980͒, Mazur and Ma- method, which is based on the additive summation of radudin ͑1981͒, Derjaguin et al. ͑1987͒, and Rabinovich the Casimir-Polder interatomic potentials with the sub- and Churaev ͑1989͒, where roughness corrections to the sequent normalization of the obtained interaction con- nonretarded van der Waals force were investigated͔. stant, this correction factor calculated up to the fourth These corrections can be calculated using perturbation perturbation order in relative amplitudes A /a is ͑Bor- theory and Green’s functions ͑Balian and Duplantier, i dag et al., 1995a͒ 1977, 1978͒, ͑Golestanian and Kardar, 1997, 1998; Emig et al., 2001, 2003͒,orbythe 10 ␬͑r͒ ͑͗ 2͘ 2 ͗ ͘ ͗ 2͘ 2͒ phenomenological methods of pairwise summation or P =1+ f 1 A1 −2 f1f2 A1A2 + f 2 A2 a2 geometrical averaging ͑Bordag et al., 1994, 1995a, 1995b; Klimchitskaya et al., 1999͒. In doing so some small pa- 20 + ͑͗f 3͘A3 −3͗f 2f ͘A2A +3͗f f 2͘A A2 rameters should be introduced which characterize the a3 1 1 1 2 1 2 1 2 1 2 deviation from basic geometry. 35 Consider two plates whose surfaces possess small de- ͗ 3͘ 3͒ ͑͗ 4͘ 4 ͗ 3 ͘ 3 − f A + f A −4 f f2 A A2 viations from the plane geometry. The surfaces of these 2 2 a4 1 1 1 1 plates can be described by ͗ 2 2͘ 2 2 ͗ 3͘ 3 ͗ 4͘ 4͒ ͑ ͒ +6 f 1f 2 A1A2 −4 f1f 2 A1A2 + f 2 A2 . 67 ͑ ͒ ͑ ͒ z s = A f ͑x ,y ͒, z s = a + A f ͑x ,y ͒, ͑63͒ 1 1 1 1 1 2 2 2 2 2 The mixed terms in Eq. ͑67͒ represent interference be- where a is the mean value of the separation distance havior. between the plates. The values of the amplitudes are For example, in the case of two plane plates with a ͉ ͑ ͉͒ ␣ ͑ ͒ chosen in such a way that max fi xi ,yi =1. The averag- small angle between them, it follows from Eq. 67 ing of Eqs. ͑63͒ over the total area of the plates with that

Rev. Mod. Phys., Vol. 81, No. 4, October–December 2009 1846 Klimchitskaya, Mohideen, and Mostepanenko: The Casimir force between real materials: …

␬͑r͒ 10 ͑␣ ͒2 ͑␣ ͒4 ͑ ͒ sary to check carefully that the curvature radius of the P =1+ L/a +7 L/a , 68 3 sphere is constant. The role of roughness in atom-wall where 2L is the characteristic length of the plate. As is interaction was investigated by Bezerra, Klimchitskaya, easily seen from Eq. ͑68͒, for plates with LϷ1mmata and Romero ͑2000͒. separation aϷ500 nm the correction due to nonparallel- The dependence of the Casimir force between metal- ism of the plates is less than 1% if ␣ does not exceed lic test bodies on the penetration depth of the electro- 2.7ϫ10−5 rad. This is the reason why the experiments magnetic oscillations into the metal ͑i.e., on the skin with two plane plates demand a high degree of parallel- depth͒ was first demonstrated by Mohideen and Roy ity. Note also that the small average separations needed ͑1998͒. This raised the question of how to take into ac- for measurements cannot be achieved if the plates are count both the conductivity and roughness corrections. tilted, as they might make contact at the edges. For the The simplest approach is a multiplicative one: to calcu- above example the contact of the plate edges occurs at a late the Casimir free energy ͑3͒ or the Casimir pressure Ј ϫ −4 ␬͑r͒ ͑ ͒ tilt angle of 1.72 =5 10 rad. The correction factor E 10 using real material properties, including the role of to the Casimir energy between the two ideal metal the skin depth, and multiply the obtained results by the ␬͑r͒ ␬͑r͒ plates, correction factor for ideal metal plates, E or P , re- spectively ͓see, e.g., Mohideen and Roy ͑1998͒ and Chan ͑ ͒ ͑␲2ប 3͒␬͑r͒ ͑ ͒ E a =− c/720a E , 69 et al. ͑2001a͔͒. This approach corresponds to the addi- is connected to Eq. ͑67͒ by tion of the different errors if they are small. Thus, it is not applicable when at least one of the corrections is ␬͑r͒ ␬͑r͒ 1 ␬͑r͒ ͑ ͒ P = E − 3 d E /da. 70 large. At short separation distances a, where the conduc- In practical computations it is often convenient not to tivity corrections are not small and strongly depend on a ͓ ͑ ͒ ͑ ͔͒ consider the complete roughness topography of surfaces see Figs. 2 a and 2 b , more sophisticated methods are ͑63͒ but to describe the roughness as a stochastic process needed. by its rms variance, The nonmultiplicative approach of the geometrical averaging was elaborated by Klimchitskaya et al. ͑1999͒. ͑k͒ K According to this method, the Casimir pressure between ␦2 ͑ ͑k͒ ͑k͒͒2 ͑k͒ ͑ ͒ st,k = ͚ H0 − hi vi . 71 the two rough plates made of real materials can be cal- i=1 culated as ͑ ͒ ͑ ͒ If roughness can be characterized by the rms variance, K 1 K 2 the correction factor to the Casimir pressure between ͑r͒͑ ͒ ͚ ͚ ͑1͒ ͑2͒ ͑ P a,T = vi vj the two parallel plates takes a simpler form Bordag et i=1 j=1 al., 1995b͒ ϫ ͑ ͑1͒ ͑2͒ ͑1͒ ͑2͒ ͒ ͑ ͒ P a + H0 + H0 − hi − hj ,T . 74 ͑ ͒ 10 105 ␬ r =1+ ͑␦2 + ␦2 ͒ + ͑␦2 + ␦2 ͒2. ͑72͒ Here P͑z,T͒ is calculated using Eq. ͑8͒ including real P a2 st,1 st,2 a4 st,1 st,2 material properties. Note that Eq. ͑74͒ is not reduced to For specially prepared surfaces in modern Casimir ex- a simple multiplication of the correction factors due to periments, the roughness corrections calculated using finite conductivity and surface roughness but takes into Eqs. ͑67͒ and ͑72͒ usually lead to the same result. The account their combined ͑nonmultiplicative͒ effect as correction factor to the Casimir energy due to stochastic well. roughness takes the form ͑Bordag et al., 1995b͒ In a similar way, the Casimir free energy for two rough plates can be calculated as ͑ ͒ 6 45 r 2 2 2 2 2 ͑ ͒ ͑ ͒ ␬ =1+ ͑␦ + ␦ ͒ + ͑␦ + ␦ ͒ . ͑73͒ K 1 K 2 E a2 st,1 st,2 a4 st,1 st,2 F͑r͒͑ ͒ ͑1͒ ͑2͒ a,T = ͚ ͚ vi vj For the configuration of a sphere ͑spherical lens͒ i=1 j=1 above a plate, which is preferred from an experimental ϫF͑ ͑1͒ ͑2͒ ͑1͒ ͑2͒ ͒ ͑ ͒ point of view, the investigation of roughness corrections a + H0 + H0 − hi − hj ,T , 75 to the Casimir force in the framework of the pairwise where F͑z,T͒ is given in Eq. ͑3͒. The method of geo- summation method was performed by Klimchitskaya metrical averaging is widely used when comparing theo- ͑ ͒ and Pavlov 1996 . It was demonstrated that if the char- retical and experimental results in experiments on the acteristic lateral scales of surface roughness on a plate measurements of the Casimir force ͓see, e.g., Klim- ⌳ ⌳ ⌳ ⌳ Ӷͱ p and on a sphere s are rather small, p , s aR, the chitskaya et al. ͑1999͒, Decca, Fischbach, et al. ͑2003͒, correction factor to the Casimir force between a sphere Decca, López, et al. ͑2003͒, Chen, Klimchitskaya, Mo- and a plate made of ideal metals is the same as for the hideen, and Mostepanenko ͑2004͒, Decca, López, Fisch- Casimir energy between the two ideal metal plates, i.e., bach, et al. ͑2005͒, Decca et al. ͑2007a, 2007b͒, Lisanti et ␬͑r͒ ͑ ͒ ͑ ͔͒ is equal to E . al. 2005 , and Munday and Capasso 2007 . If the above inequalities are not fulfilled ͑e.g., there Equations ͑74͒ and ͑75͒ are based, however, on the are significant deviations of the lens surface from a proximity force approximation and do not take into ac- spherical shape͒, an additional term of the order ϳ1/a count the diffractionlike correlation effects. In first- results in the correction factor. That is why it is neces- order perturbation theory, these effects have been inves-

Rev. Mod. Phys., Vol. 81, No. 4, October–December 2009 Klimchitskaya, Mohideen, and Mostepanenko: The Casimir force between real materials: … 1847 tigated for ideal metal plates covered with periodic bodies on the calculational results. In fact, Eqs. ͑3͒ and roughness ͑Emig et al., 2001, 2003͒ and for metals de- ͑10͒ for the Casimir free energy per unit area and pres- scribed by the plasma model covered with stochastic sure are obtained for plates of infinite area. Using the roughness ͑Genet, Lambrecht, Maia Neto, and Rey- proximity force approximation, we obtain from Eq. ͑3͒ naud, 2003; Maia Neto et al., 2005͒. As demonstrated, the Casimir force between a sphere of radius Rӷa the correlation effects noticeably contribute to the above an infinitely large plate. In experimental configu- roughness correction only at relatively large separations rations, however, the plate always has a finite size, and, of order or much larger than the roughness period ⌳ ͑or instead of a sphere, a spherical lens of some thickness the correlation length͓͒see also van Zwol et al. ͑2007͔͒. HϽR may be used. It should be noted that recently The special analysis ͑Chen, Klimchitskaya, Mohideen, rigorous methods based on quantum field theory first and Mostepanenko, 2004; Decca, López, Fischbach, et principles have been developed which allow one to cal- al., 2005͒, based on the results by Emig et al. ͑2001, 2003͒ culate the scalar and electromagnetic Casimir forces be- showed that at separation distances a a few times tween arbitrarily shaped compact objects ͓see, e.g., Gies smaller than ⌳ the role of correlation effects is negligibly and Klingmüller ͑2006a, 2006b, 2006c͒ and Emig et al. small ͑about 0.03–0.04 % change in the magnitude of the ͑2007, 2008͔͒. However, the application of these methods roughness correction for the roughness topography of for calculation of small corrections due to the finite sizes test bodies used in the experiments͒. With an increase of of the test bodies in experimental configurations, where separation distance the correlation effects become more the separation distances are rather small, still remains a important, but at such large separations the complete problem ͑especially in the electromagnetic case͒. For roughness correction including the correlation effects is this reason, finite size corrections still need to be esti- negligibly small. Thus, an accurate account of surface mated using the simple method of pairwise summation. roughness corrections would need more sophisticated For a spherical lens of radius R and thickness H placed methods beyond the pairwise summation approach or at a distance a above a disk of radius L the finite size geometrical averaging for test bodies with roughness correction factor to the Casimir force is given by having a small correlation length. In this case there ͑Bezerra et al., 1997͒ would be large correlation effects even at relatively ␬ ͑ 3 3͒͑ ͒−3 ͑ ͒ short separations where the role of roughness is signifi- fs =1− a /R 1−T . 76 cant. The situation may become even more complicated if the roughness amplitudes are not small and one has to Here Tϵmax͓R/ͱR2 +L2 ,͑R−H͒/R͔. Using Eq. ͑76͒ go beyond the lowest-order perturbation theory. With one can estimate that in all experiments performed to respect to the roughness corrections, to date, however, date ͑see Secs. IV and V͒ the correction due to the finite there are no measurements of the Casimir force which size of the test bodies is negligibly small. Thus, it is rea- fall into this category. The first experimental evidence sonable to compare experimental results with the theo- for the effects beyond the proximity force approxima- retical computations using Eq. ͑3͒ or ͑10͒. tion was obtained by Chan et al. ͑2008͒ measuring the The finite thickness of plates can be taken into ac- Casimir interaction between a Au sphere and a Si plate count by the replacement of reflection coefficients ͑6͒ with deep rectangular trenches. This experiment is dis- with those obtained for multilayered structures ͑Zhou cussed in Sec. V. D . and Spruch, 1995; Klimchitskaya et al., 2000͒. For metal There is one more effect connected with surface plates of larger thickness than the skin depth in the fre- ͑␦ ␻ Ϸ roughness which should be accounted for in Casimir quency region of infrared optics 0 =c/ p 22 nm for force measurements. It is conceivable that the spatial Au͒ the calculational results are the same as those for variations of the surface potentials due to grains of the semispaces. Thus, for typical metal films of about polycrystalline metal film ͑the so-called patch potentials͒ 100 nm thickness deposited on the surfaces of test bod- simulate the Casimir force. Speake and Trenkel ͑2003͒ ies in different experiments, the Lifshitz formulas for obtained general expressions for the electrostatic free semispaces can be reliably used ͑Klimchitskaya et al., energy and pressure in the configuration of two parallel 2000͒. As an example, if in one case we have two Au plates, which result from random variations of patch semispaces at a separation of 200 nm and in the other a potentials. These expressions can be used to analyze the Au semispace at the same separation from a role of patch effects in any performed experiment. In 100-nm-thick Au film made on a Si semispace substrate, ͑ particular, it was shown Chen, Klimchitskaya, Mo- the relative difference in the Casimir energy is less than hideen, and Mostepanenko, 2004; Decca, López, Fisch- 0.01%. Recent experiments on the measurement of the ͒ bach, et al., 2005 that in the most precise experiments Casimir force between a metallic sphere and semicon- measuring the Casimir force between metallic test bod- ductor ͑Si͒ plate ͑Chen et al., 2005, 2007a, 2007b; Chen, ͑ ͒ ies see Secs. IV.A and IV.B the patch effects are negli- Klimchitskaya, et al., 2006; Chen, Mohideen, et al., 2006͒, gibly small even at the shortest separation distances. have created an interest in the role of the thickness of the Si plate ͑Duraffourg and Andreucci, 2006; Lambre- cht et al., 2007; Pirozhenko and Lambrecht, 2008a͒.In 2. Finite size and thickness these papers the Lifshitz formula at zero temperature is The last point to be considered in this section is the used in the computations and, in the latter two, large influence of the finite size and a thickness of interacting separations up to 1 mm were considered. The depen-

Rev. Mod. Phys., Vol. 81, No. 4, October–December 2009

1848 Klimchitskaya, Mohideen, and Mostepanenko: The Casimir force between real materials: …

 µ ´

   orous approach to the comparison of experiment and

theory in Casimir force measurements was presented ½

1 ͑Decca, López, Fischbach, et al., 2005; Chen, Mohideen, ¾ 0.8 et al., 2006; Klimchitskaya, Chen, et al., 2006͒ using sta- 0.6 tistical methods. The first step in the application of this approach is to 0.4 characterize how precise the experimental data are with- 0.2 out any relation to the theory. To do this, the absolute and relative total experimental errors should be calcu- 10 20 30 40 50 lated as a combination of systematic and random errors.

⌸expt͑ ͒

´  Ñ µ  We use in the error analysis the notation a for the measured quantity, which is either the Casimir pressure FIG. 3. The ratio of the Casimir free energy for a Si plate of Pexpt͑a͒ in the configuration of two parallel plates or the thickness d at separations a=1 ͑lines labeled 1͒ and 5 ␮m Casimir force Fexpt͑a͒ between a sphere and a plate, as a ͑lines labeled 2͒ from Au semispace to the free energy in the function of separation distance a between them. configuration of Si and Au semispaces at the same separations. In each experiment, there are several sources Solid lines are computed at T=300 K. Dashed lines are com- ⌬s⌸expt͑ ͒ of the absolute systematic errors i a and puted at T=0. ␦s⌸expt͑ ͒ corresponding relative systematic errors i a ⌬s⌸expt͑ ͒ ͉⌸expt͑ ͉͒ ഛ ഛ ͑ = i a / a , where 1 i J see Secs. IV and dences of the Casimir force on the plate thickness d and V for the description of specific experiments͒. It is nec- the doping concentration are investigated. The results essary to stress that both in metrology and in all natural obtained are, however, not applicable at large separation sciences ͑physics, chemistry, biology, etc.͒ the term sys- distances at the laboratory temperature T=300 K. For tematic error is used in two different meanings example, we plot in Fig. 3 the ratio of the Casimir free ͑Rabinovich, 2000͒. According to the first meaning, a energy for the configuration of a Si plate of thickness d systematic error is some bias in the measurement which at a separation a from a Au semispace to the free energy always makes the measured value higher or lower than in the configuration of Si and Au semispaces, as a func- the true value. Such systematic errors in the measure- tion of Si plate thickness d. Solid lines 1 and 2 are com- ment results are usually removed using some known puted at T=300 K at a=1 and 5 ␮m, respectively. process, i.e., through a calibration. They can also be Dashed lines 1 and 2 are computed at T=0 at the same taken into account as corrections ͑see the description of separations. As is seen in the figure, at a=5 ␮m the de- the calibration procedure and an example of correction viation between the solid and dashed lines reaches 20%. in the measurement of the Casimir force discussed in Thus, the role of finite thickness of the plate at separa- Sec. IV.A͒. The systematic errors in this understanding tions above 1 ␮m should be investigated using the Lif- are often called systematic deviations. Below it is as- shitz formula at nonzero temperature. sumed that the experimental data under consideration are already free of such deviations. Another meaning, which is used below, defines the C. Quantitative comparison between experiment and theory in systematic errors as the errors of a calibrated measure- force-distance measurements ment device. The errors of some theoretical formula used to convert the directly measured quantity into the 1. Experimental errors and precision indirectly measured one ͑see Sec. IV.B͒ are also consid- When comparing experimental data from Casimir ered as systematic. In accordance with common under- force experiments with theoretical predictions, the im- standing, the error of a calibrated device is the smallest portant question on how to quantitatively characterize fractional division on the scale of this device. In the lim- the measure of agreement between them has to be ad- its of this range the systematic errors are considered as dressed. Previously the concept of the root-mean-square random quantities characterized by a uniform distribu- deviation between theory and experiment has been used tion. Because of this, the total relative systematic error is to quantify the precision of measurements and the ͑Rabinovich, 2000͒ agreement with theory ͓see, e.g., early experiments by ␦s⌸expt͑ ͒ Bordag et al. ͑2001͒ and also some later experiments a ͑Bressi et al., 2002; Decca, Fischbach, et al., 2003; Chen, J J Klimchitskaya, Mohideen, and Mostepanenko, 2004͔͒.It ͭ ␦s⌸expt͑ ͒ ͑J͒ͱ ͓␦s⌸expt͑ ͔͒2ͮ = min ͚ i a ,k␤ ͚ i a , is well known, however, that this method is not appro- i=1 i=1 priate for strongly nonlinear quantities, such as the Ca- ͑77͒ simir force, which changes rapidly with separation dis- ͑J͒ tance ͑Rabinovich, 2000͒. It was shown ͑Ederth, 2000͒ where ␤ is the chosen confidence level and k␤ is the that the calculation of the root-mean-square deviation tabulated coefficient depending on ␤ and on the number between experiment and theory leads in this case to dif- of sources of systematic errors J in the experiment under ferent results when applied in different measurements consideration ͑Rabinovich, 2000͒. In precise experi- ranges but no better method was suggested. Later, a rig- ments, errors should be determined at a confidence level

Rev. Mod. Phys., Vol. 81, No. 4, October–December 2009 Klimchitskaya, Mohideen, and Mostepanenko: The Casimir force between real materials: … 1849

␤ജ95%. However, in many Casimir force measure- ⌬s⌸expt͑a͒ calculated at 95% confidence. In the separa- ments only a 67% confidence level is used ͑see Sec. IV͒. tion region, where 0.8ഛr͑a͒ഛ8, a combination of errors Equation ͑77͒ can be written equivalently in terms of the is performed using the rule absolute errors. ⌬t⌸expt͑ ͒ ͑ ͓͒⌬r⌸expt͑ ͒ ⌬s⌸expt͑ ͔͒ ͑ ͒ In precise experiments ͑see Secs. IV and V͒ it is com- a = q␤ r a + a , 80 mon to perform several sets of measurements, say n, where the coefficient q␤͑a͒ at a ␤=0.95 confidence level ͑ ͒ within the same separation region amin,amax . This is varies between 0.71 and 0.81 depending on r͑a͒. Being ͑ ͒ done in order to decrease the random error and to nar- conservative, one can use q0.95 r =0.8. row the confidence interval. Each set consists of pairs It is important that the total experimental error ͓ ⌸expt͑ ͔͒ ഛ ഛ ͑ t expt ai , ai , where 1 i imax imax is different for dif- ⌬ ⌸ ͑a͒ and the corresponding total relative error ferent experiments͒. All measurement data can be rep- ␦t⌸expt͑a͒=⌬t⌸expt͑a͒/͉⌸expt͑a͉͒ characterize the preci- ͓ ⌸expt͑ ͔͒ ഛ ഛ resented as pairs aij, aij , where 1 j n. If sepa- sion of an experiment in its own without comparison rations with fixed i are approximately the same in all sets with any theory. Note also that in metrology an experi- ͑ Ϸ ͒ of measurement i.e., aij ai , the mean and the variance ment is called precise if the random error is much ͑ ͒ of the mean at each point ai are obtained in the standard smaller than the systematic error Rabinovich, 2000 .In way ͑Rabinovich, 2000͒, Casimir force measurements performed up to date, only ͑ ͒ n one experiment Decca et al., 2007a, 2007b satisfies this 1 ⌸¯ expt ⌸expt͑ ͒ rigorous criterion. i = ͚ aij , n j=1 2. Comparison of experiment with theory n 2 1 ⌸theor͑ ͒͑ s = ͚ ͓⌸expt͑a ͒ − ⌸¯ expt͔2. ͑78͒ The theoretical values a the pressure between ⌸¯ ͑ ͒ ij i ͒ i n n −1 j=1 the plates or the force between the sphere and the plate computed using the Lifshitz theory also contain some If separations with fixed i are different in different sets errors. If the Casimir force between a sphere and a plate of measurements ͑i.e., with different j͒, the calculation of is calculated, one of the theoretical errors is caused us- variance demands a more sophisticated procedure ing the proximity force approximation. To be conserva- ͑Decca, López, Fischbach, et al., 2005; Klimchitskaya, tive, one can set the respective relative error to a/R, i.e., Chen, et al., 2006͒. to its maximum possible value ͑see Sec. II.B͒. Another ⌸¯ expt Direct calculations show that the mean values i error is due to the uncertainty in the optical data used. It are uniform, i.e., change smoothly with the change of can be conservatively estimated to be equal to 0.5% ͑ i. The variances of the mean s⌸¯ are, however, not uni- over the entire measurement range Chen, Klim- i form. To smooth them, a special procedure is used in chitskaya, Mohideen, and Mostepanenko, 2004͒. The to- statistics ͑Cochran, 1954; Brownlee, 1965͒. The applica- tal theoretical error can be computed using Eq. ͑77͒ tion of this procedure allows one to replace s⌸¯ with the adapted for theoretical errors. Note that when using dif- i smooth function of separation s⌸¯ ͑a͒͑Klimchitskaya, ferent sets of optical data one can arrive at different Chen, et al., 2006͒. Then the random absolute error at a computational results for ⌸theor͑a͒. This has resulted in separation distance a can be calculated at a chosen con- statements in the literature ͓see, e.g., Pirozhenko et al. ͑ ͒ ͑ ͔͒ fidence level ␤, 2006 and Capasso et al. 2007 that if different sets of optical data obtained for different samples lead to devia- ⌬r⌸expt͑ ͒ ͑ ͒ ͑ ͒ ͑ ͒ a = s⌸¯ a t͑1+␤͒/2 n −1 . 79 tions in the Casimir force equal to, say, around 5%, then the comparison of theory with experiment on a better ͑ ͒ Here the value of tp f can be found in the tables for level is impossible. These statements are, however, in- ͓ ͑ ͔͒ Student’s t distribution see, e.g., Brandt 1976 . correct from the point of view of measurement theory ⌸expt͑ ͒ To find the total experimental error of a , one ͑Decca et al., 2007b͒. The point to note is that here we should combine the random and systematic errors. In so are dealing with methods for the verification of hypoth- doing it is assumed that the random error is character- eses. The procedures for such verifications are well de- ized by the Student distribution which approaches to the veloped in statistics. Different hypotheses correspond to normal distribution with the increasing number of the using different theoretical approaches ͑for instance, measurement sets. The systematic error is assumed to be based on the Drude or plasma model description of con- described by a uniform distribution. There are different duction electrons in metals or the inclusion or the ne- methods in statistics to obtain this combination glect of the small dc conductivity of dielectrics at non- ͑ ͒ Rabinovich, 2000 . A widely used method is based on zero temperature͒ with different sets of optical data s expt the value of the quantity r͑a͒=⌬ ⌸ ͑a͒/s⌸¯ ͑a͒. Accord- available. The comparison of different hypotheses with ing to this method, at all a, where r͑a͒Ͻ0.8, the contri- experiment can be performed by plotting the respective bution from the systematic error is negligible and theoretical bands as functions of separation ͑the width ⌬t⌸expt͑a͒=⌬r⌸expt͑a͒ at a 95% confidence level ͓in so of the band is determined by the total theoretical error͒ doing ⌬r⌸expt͑a͒ is also calculated at 95% confidence us- and the experimental data with their absolute errors on ing Eq. ͑79͔͒. If the inequality r͑a͒Ͼ8 is valid, the ran- one graph ͓see, e.g., Fig. 11͑a͒ in Sec. IV.B͔. If the theo- dom error is negligible and ⌬t⌸expt͑a͒=⌬s⌸expt͑a͒ with retical band does not overlap with the experimental

Rev. Mod. Phys., Vol. 81, No. 4, October–December 2009 1850 Klimchitskaya, Mohideen, and Mostepanenko: The Casimir force between real materials: … data, including their errors, over a wide separation one of these approaches or all of them are false. The range, it is experimentally excluded and the correspond- rejection of some of the experimentally consistent ap- ing hypothesis must be rejected. It is important to note proaches can be done on a theoretical basis only. For that the data and theory are not compared at just one example, if the measurement is performed at room tem- point but over a wide separation region where the dis- perature and both theoretical computations done at T tance dependence is nonlinear. The set of optical data =0 and 300 K are consistent with the data, the compu- which contradicts the experimental results, when com- tation at T=300 K can be considered as preferable. In bined with any theoretical approach, should be consid- fact, to reliably discriminate between the two experi- ered as irrelevant to the actual properties of the film mentally consistent theoretical approaches, more exact used in the measurement. measurements are desirable. From the above it becomes Another method to compare theory with experiment clear that the described statistical criteria are somewhat is the consideration of the confidence interval for the asymmetric. They provide solid grounds for rejection of ⌸theor͑ ͒ ⌸¯ expt͑ ͒͑ the experimentally inconsistent approaches at a high random quantity ai − ai Decca, López, Fischbach, et al., 2005; Chen, Mohideen, et al., 2006͒.In confidence level but do not permit the claim of the ex- doing so, one should take into account that this compari- perimental confirmation of some consistent approach at the same high confidence. These general criteria are il- son is done at the separation distances ai measured with some error ⌬a. Thus, the theoretical value ⌸theor͑a ͒ at lustrated in Secs. IV and V using the examples of differ- i ͑ this point can be known only with the relative error ent Casimir force measurements see Figs. 12, 13, and ͒ ␣⌬a/a ͑Iannuzzi, Gelfand, et al., 2004͒. Here ␣=3 for 21 . It would be interesting to apply more sophisticated i ͑ the force between a sphere and a plate and ␣=4 for the statistical methods, for instance, Bayesian Berger, 1993; ͒ pressure between the two plates. This error, although Carlin and Louis, 2000; Lehmann and Romano, 2005 not connected with the theoretical approach used, for the data analysis of force-distance relations resulting should be included as part of the theoretical errors in from the Casimir force measurements. Until now, how- Eq. ͑77͒. Note that at short separations it results in a ever, such methods have not been used devoted to the major contribution to the total theoretical error ⌬t⌸theor. experimental study of the Casimir effect. Then the absolute error of the difference between It is notable that conclusions concerning consistency or rejection of a hypothesis do not depend on the theory and experiment ⌶⌸͑a͒ at a 95% confidence can ͑ ͒ method of comparison used ͑see discussion of different be calculated using Eq. ͑77͒ with J=2 and k 2 =1.1, 0.95 situations in Secs. IV–VI͒. The half width of the confi- t expt t theor ⌶ ͑ ͒ ⌶⌸͑a͒ = min͕⌬ ⌸ ͑a͒ + ⌬ ⌸ ͑a͒, dence interval ⌸ a has very little dependence on the theoretical approach used ͑i.e., on the hypothesis to be ϫ ͱ͓⌬t⌸expt͑ ͔͒2 ͓⌬t⌸theor͑ ͔͒2͖ ͑ ͒ 1.1 a + a . 81 verified͒. It usually results from the experimental errors The confidence interval for the quantity ⌸theor͑a͒ in the measurement of the separation distance and the Casimir force or the pressure. −⌸¯ expt͑a͒ at a 95% confidence level is given by ͓−⌶⌸͑a͒,⌶⌸͑a͔͒. When one compares a theoretical ap- proach with experimental data, the differences between IV. CASIMIR EXPERIMENTS WITH METALLIC TEST the theoretical and mean experimental values of ⌸ may BODIES or may not belong to this interval. A theoretical ap- proach for which not less than 95% of the differences A. Measurements of the Casimir force using an atomic force microscope theor expt ⌸ ͑a͒−⌸¯ ͑a͒ belong to the interval ͓−⌶⌸͑a͒,⌶⌸͑a͔͒ ͓ ͔ within any separation subinterval a1 ,a2 of the entire Recent interest in the experimental investigation of measurement range is consistent with the experiment. In the Casimir effect has its beginning in the experiment by this case the measure of agreement between experiment Lamoreaux ͑1997͒. This experiment, however, contains expt and theory is given by ⌶⌸͑a͒/͉⌸¯ ͑a͉͒. On the contrary, several uncertainties that do not permit one to consider ͓ ͔ it as precise and definitive ͑see the discussion in Sec. if for some theoretical approach a subinterval a1 ,a2 IV.C.1͒. The first precise and definitive direct measure- exists, where almost all differences ⌸theor͑a͒−⌸¯ expt͑a͒ ments of the Casimir force between a metal-coated are outside of the confidence interval ͓−⌶⌸͑a͒,⌶⌸͑a͔͒, sphere and a plate were performed by Mohideen and this theoretical approach is excluded by experiment at Roy ͑1998͒, Roy et al. ͑1999͒, and Harris et al. ͑2000͒ in separations from a1 to a2 at a 95% confidence level. If ͑ ͒ three successive experiments using an AFM operated in the theoretical approach hypothesis is excluded by ex- vacuum. In these experiments the fundamental require- periment at a 95% confidence level, the probability that ments for the measurement of the Casimir force pro- it is true is at most 5%. It may happen that several the- posed by Sparnaay ͑1958, 1989͒, such as ͑i͒ clean sur- oretical approaches i=1,2,... are consistent with experi- faces for the test bodies, ͑ii͒ precise and reproducible ment, i.e., not less than 95% of the differences measurement of the separation distance between them, ⌸theor͑ ͒ ⌸¯ expt͑ ͒͑ ͒ ͑ ͒ i a − a i=1,2,... belong to the confidence and iii low electrostatic potential differences, were met interval ͓−⌶⌸͑a͒,⌶⌸͑a͔͒ ͑such situations are considered for the first time ͓historical survey of the Casimir force in Secs. IV and V͒. The statistical criteria used do not measurements performed before 1990 is given by Bor- allow an indication of the probability of the event that dag et al. ͑2001͔͒. They introduced the now standard use

Rev. Mod. Phys., Vol. 81, No. 4, October–December 2009 Klimchitskaya, Mohideen, and Mostepanenko: The Casimir force between real materials: … 1851

20 200300400500 0 a (nmnm) -20 N) 12 - -40

-60

-80

Casimir Force (10 -100

FIG. 4. Schematic diagram of the measurement of the Casimir -120 force using an atomic force microscope. FIG. 5. Comparison of experiment and theory. Measured mean Casimir force between Al surfaces versus separation is of metal-coated polystyrene spheres, which have the shown as open squares. The theoretical Casimir force with cor- unique advantages of being extremely light weight, per- ͑ ͒ rections due to skin-depth and surface roughnesses and for fectly smooth made from liquid phase , and having very ideal metal surfaces is shown by the solid and dashed lines, ͑ ͒ low eccentricity less than parts per thousand as one of respectively. the surfaces in Casimir force measurements. These mea- surements also pioneered the independent determina- tion of the roughness-induced average separation dis- fiber interferometer ͑Chen and Mohideen, 2001͒. As the tances at contact, the cantilever calibration, and the piezo was moved with continuous periodic triangular potential differences between the two surfaces using the voltages, the nonlinearities can be precisely accounted electrostatic force. This allowed a careful comparison to for through interferometric calibration. The last term in the theory and a precise study of the effect of the metal Eq. ͑83͒ is due to the decrease in the separation distance conductivity ͑also in terms of the skin depth͒ and the from the flexing of the cantilever in response to any surface roughness which previously could not be done. Ͻ force. Note that Sdef 0 for an attractive force. The de- A schematic diagram of the experimental setup for flection coefficient m was found from electrostatic cali- the measurement of the Casimir force using the AFM is bration. shown in Fig. 4. The Casimir force between a sphere of To measure the Casimir force between the sphere and radius R and a plate causes the cantilever to flex. This the plate, they are both grounded together with the flexing is detected by the deflection of a laser beam lead- AFM. The plate is then moved toward the sphere using ing to a difference signal Sdef between photodiodes A the piezo and the corresponding photodiode difference and B which was calibrated by means of the electrostatic signal is measured. It is converted into the respective force between the sphere and the plate. For this purpose deflection of the cantilever tip and by use of Hooke’s law various voltages V were applied to the plate while the into the values of the Casimir force Sdefkm, where k is sphere remained grounded. The measurements of the the spring constant ͑see below for the results of electro- electrostatic forces were performed at separations a static calibration and the values of the main parameters͒. larger than a few micrometers where the contribution In the first two measurements using the AFM by Mo- from the Casimir force is negligibly small. The exact ex- hideen and Roy ͑1998͒ and Roy et al. ͑1999͒, the sphere pression for the electrostatic force in sphere-plate con- and the plate were coated with about 300 nm of Al. To ͑ ͒ figuration is given by Smythe, 1950 prevent rapid oxidation of Al it was sputter coated with ϱ thin Au/Pd layers. In the first experiment ͑Mohideen coth ␣ − n coth n␣ ͑ ͒ ␲⑀ ͑ ͒2 ͚ and Roy, 1998; Klimchitskaya et al., 1999͒ it was demon- Fel a =2 0 V − V0 ␣ n=1 sinh n strated that the Casimir force between real metal sur- ϵ X͑a͒͑V − V ͒2, ͑82͒ faces deviates significantly from the Casimir prediction 0 made for ideal metal surfaces. In Fig. 5 the measured ␣ where cosh =1+a/R, V0 is the residual voltage on the data for the Casimir force, as a function of separation ⑀ sphere and 0 is the permittivity of the vacuum. From a distance a, are shown as open squares. The dashed line comparison of theoretical expression ͑82͒ with the ex- indicates the Casimir force ͑19͒ between an ideal metal perimental data, the separation on contact a0 and the plate and an ideal metal sphere of R=100±2 ␮m radius residual potential difference between the test bodies used in the experiment. The solid line shows the theo- were determined. Finally the absolute separations were retical Casimir force calculated with account of the cor- found from rections due to the skin depth and the surface rough- ͑ ͒ ness. This force was found using the fourth-order a = a0 + apiezo + Sdefm, 83 ␦ perturbation theory in the relative skin depth 0 /a de- ͑ where the movement apiezo of the piezoelectric tube on fined in Sec. II.D.1. The perturbation theory Klim- which the plate is mounted was calibrated by an optical chitskaya et al., 1999; Bezerra, Klimchitskaya, and

Rev. Mod. Phys., Vol. 81, No. 4, October–December 2009

1852 Klimchitskaya, Mohideen, and Mostepanenko: The Casimir force between real materials: …

©

§ Mostepanenko, 2000͒, when applied in the appropriate ¡ region of the plasma model, leads to approximately the 0 same results as the Lifshitz formula at zero temperature -100 ͑first measurements using the AFM were not of suffi- -200 cient precision to measure the thermal effect at T =300 K͒. The Casimir force including the effect of both -300 the skin depth and roughness corrections was calculated -400 using Eq. ͑75͒ and the PFA ͑16͒. As shown in Fig. 5, the

solid line is in good agreement with data, thus, demon- 100 150 200 250 300

¡ £ ¥ § strating the role of skin depth and roughness correc- tions. We now consider the most precise, third, measure- FIG. 6. Comparison of experiment with ideal metal theory. The measured mean Casimir force between Au surfaces versus ment of the Casimir force between metal surfaces by separation is shown as dots. Theoretical Casimir force for ideal means of an AFM ͑Harris et al., 2000͒. The use of a thin metal surfaces is shown by the solid line. Au/Pd coating on the top of the Al in the first two ex- periments prevented a complete theoretical treatment of the properties of the real metal surfaces. Because of this, basis of these measurements a reanalysis of the indepen- in the third experiment, Au layers of 86.6±2 nm thick- dence of V0 on separation in the measurements of the ness were coated on both the plate and the spherical Casimir force by means of an atomic force microscope ͑ ͒ surfaces. Such a coating is sufficient to reproduce the and a micromachined oscillator see Sec. IV.B was in- ͑ ͒ properties of an infinitely thick metal ͑see Sec. III.B.2͒. vited. Decca et al. 2009 demonstrated, however, that The sphere diameter ͑including the metal coating͒ was for a centimeter-size spherical lens at such short separa- measured using the scanning electron microscope to be tions from the plate, the electrostatic force law used by Kim et al. ͑2008͒ is not applicable due to the inevitable 191.3±0.5 ␮m ͑independent calibration of the scanning deviations from a perfect spherical shape of the me- electron microscope was done with interferometrically chanically polished and ground surface and the presence calibrated AFM piezo standards͒. The cantilever was of dust or other impurities. Because of this, the observed calibrated and the residual potential between the sur- anomalies in the electrostatic calibration are not directly faces was measured using the electrostatic force at sepa- relevant to the experimental results considered here. rations larger than 3 ␮m with voltages V from −3 to 3 V The averaged Casimir force measured from 27 scans is applied to the plate. As a result the following values reproduced in Fig. 6 as dots within the measurement were obtained: m=8.9±0.3 nm per unit deflection signal, range from 62 to 300 nm. In the same figure the solid km=0.386±0.003 nN per unit deflection signal, a0 line represents the Casimir force between ideal metal =32.7±0.8 nm, and the residual potential on the interacting surfaces. This once again demonstrates the grounded sphere V0 =3±3 mV. At separations of about difference between ideal metal surfaces and real materi- 60 nm this residual potential leads to forces which are als. only 0.075% of the Casimir force. At a=100 nm the con- The experimental errors in the experiment by Harris tribution of the residual electric force increases up to et al. ͑2000͒ were reanalyzed by Chen, Klimchitskaya, 0.17% and at a=160 nm up to 0.36% of the Casimir Mohideen, and Mostepanenko ͑2004͒. The maximum force. value of the systematic error at a 95% confidence level The magnitudes of the residual potential and separa- was shown to be ⌬sFexpt=2.7 pN and of the random er- tion on contact were shown to be independent of the ror ⌬rFexpt=5.8 pN over the entire measurement range. distances where the measurements of the electric force Using the combination rule in Eq. ͑80͒, the total experi- were performed. The electric force corresponding to the mental error of the force measurements is given by residual potential difference has been subtracted from ⌬tFexpt=6.8 pN. It does not depend on separation. The the measured total force to obtain the pure Casimir relative error varies from 1.5% to 2% when the separa- force. This is in fact a correction made to remove the tion increases from 63 to 72 nm. It increases to 4.8% systematic deviation due to the residual potential differ- and 30% when separation increases to 100 and 200 nm, ence, as discussed in Sec. III.C.1. From the above values respectively. The error in the measurements of absolute it becomes clear that the electrostatic residuals are neg- separations was found to be ⌬a=1 nm. All errors are ligible in comparison with the Casimir force. The electric determined at a 95% confidence level. The experimental force due to the patch potentials contributes only 0.23% data over the separation region from 63 to 100 nm are and 0.008% of the Casimir force at separations a=62 shown in Fig. 7 as crosses, where the sizes of the errors and 100 nm, respectively ͑Chen, Klimchitskaya, Mo- in the separation and force measurements are given in hideen, and Mostepanenko, 2004͒. Recently it was true scales. To make the figure readable, only every sec- claimed that the residual potential V0 from the electro- ond data point is presented. We underline that the mea- static calibration in sphere-plate configuration is separa- surement of the Casimir force described above is inde- tion dependent ͑Kim et al., 2008͒. They used a Au- pendent in the sense that no fit to any theory of coated sphere of 30.9 mm radius at separations of a few dispersion forces has been used. The above procedures tens of nanometers above a Au-coated plate. On the used a fit only to the well-understood electric force in

Rev. Mod. Phys., Vol. 81, No. 4, October–December 2009

Klimchitskaya, Mohideen, and Mostepanenko: The Casimir force between real materials: … 1853

©

§ ¡ plate and the sphere were found from Eq. ͑65͒ to be ͑ -150 H0 =2.734 nm Chen, Klimchitskaya, Mohideen, and Mostepanenko, 2004͒. The rms variance defined in Eq. -200 ͑ ͒ ␦ 71 is st=1.18 nm. The role of roughness in this experi- -250 ment is very small. Even at the shortest separation a =62 nm roughness contributes only 0.24% of the Ca- -300 simir force. The -type and correlation effects discussed in Sec. III.B.1 were shown to contribute less -350 than one-tenth of this value ͑Chen, Klimchitskaya, Mo- -400 hideen, and Mostepanenko, 2004͒. Comparison of the data with the Lifshitz theory at T -450 =0 taking the surface roughness into account demon-

65 70 75 80 85 90 95 100 strated good agreement over the entire measurement

¡ £ ¥ § range ͑Harris et al., 2000; Chen, Klimchitskaya, Mo- hideen, and Mostepanenko, 2004͒. However, as dis- FIG. 7. Comparison of experiment and theory. The measured cussed in Sec. II.D.1, the use of the Drude model at low mean Casimir force between Au surfaces versus separation is frequencies presents serious difficulties. In addition, ac- shown as crosses. The arms of the crosses represent in true cording to Sec. III.A.2, the use of the zero-temperature scale absolute errors determined at 95% confidence. Theoret- Lifshitz formula for the interpretation of the experiment ical Casimir force calculated using the generalized plasmalike performed at T=300 K is open to discussion. In Sec. model is shown by the solid band. III.A.1 a second approach to the calculation of Au di- electric permittivity along the imaginary frequency axis, order to find the values of some parameters such as the was described. It is based on the plasmalike dielectric separation on contact a0, the residual potential differ- permittivity with the inclusion of the relaxation due to ence V0, and deflection coefficient m. the core electrons alone ͑the solid line in Fig. 1͒. This A theoretical calculation of the Casimir force by Har- approach can be used in combination with the Lifshitz ris et al. ͑2000͒ was performed using the Lifshitz formula formula at T0 with no thermodynamic inconsistencies. at zero temperature ͓Eq. ͑12͔͒ and the PFA ͓Eq. ͑16͔͒. Calculations of the Casimir force in the experimental The dielectric permittivity of Au along the imaginary configuration of Harris et al. ͑2000͒ at T=300 K using frequency axis was obtained using the first approach de- the generalized plasma model ͑i.e., with the dielectric scribed in Sec. III.A.1, i.e., with the optical data extrapo- permittivity given by the solid line in Fig. 1͒ were per- lated to low frequencies by means of the Drude model formed by Klimchitskaya et al. ͑2007a͒. The Casimir ͑ ͒ the dashed line in Fig. 1 . The character of roughness force obtained, as a function of separation, is shown by was investigated from the analysis of the AFM images of the solid band in Fig. 7. The width of this band indicates the surfaces of Au films. A typical AFM image of sur- the total theoretical error calculated at a 95% confi- face topography is presented in Fig. 8. As shown in this dence level as described in Sec. III.C.2, taking into ac- figure, the roughness is mostly represented by the sto- count the error of the PFA and uncertainties in the op- chastically distributed distortions with typical heights of tical data of Au related to the interband transitions. As about 2–4 nm and rare pointlike peaks with typical shown in Fig. 7, the theoretical approach using the gen- heights up to 16 nm. Zero-roughness levels on both the eralized plasmalike permittivity is also in good agree- ment with the data. Within the separation region from 63 to 72 nm, where the relative total experimental error determined at a 95% confidence level varies from 1.5% to 2%, all experimental crosses overlap with the theoret- ical band. From this one can conclude that the measure of agreement between experiment and theory at separa- tion distances from 63 to 72 nm also varies from 1.5% to 2% of the measured force. Thus, both theoretical approaches using the complete optical data for Au extrapolated to low frequencies by the Drude model and the Lifshitz formula at T=0 or the generalized plasmalike dielectric permittivity and the Lifshitz formula at T=300 K are consistent with the ex- perimental data by Harris et al. ͑2000͒. This is because at separations below 100 nm, where the precision of this experiment is relatively high, the theoretical approaches used lead to almost coincident results. At such short FIG. 8. Typical atomic force microscope image of the Au coat- separations approximately the same theoretical results ing on the plate with roughness height h and lateral scale l. are also obtained using the complete optical data ex-

Rev. Mod. Phys., Vol. 81, No. 4, October–December 2009 1854 Klimchitskaya, Mohideen, and Mostepanenko: The Casimir force between real materials: …

trapolated to zero frequency by the Drude model and the Lifshitz formula at T=300 K. However, at larger separation distances the calculational results obtained using the above two approaches at nonzero temperature become significantly different. Because of this, high- precision experiments at separations of a few hundred nanometers, considered next, are important for the un- derstanding of the underlying physics.

B. Precise determination of the Casimir pressure using a micromachined oscillator

Microelectromechanical systems are well adapted for the investigation of small forces acting between closely spaced surfaces. One such system, a micromachined os- FIG. 9. Schematic diagram of the measurement of the Casimir cillator, was first used by Chan et al. ͑2001a, 2001b͒ to force using a micromachined oscillator. demonstrate the influence of the Casimir force on the static and dynamic properties of micromechanical sys- F ͑a,T͒ץ b2 ͒ ͑ tems see Sec. IV.C.2 . ␻2 = ␻2ͩ1− s ͪ, ͑84͒ ץ r 0 ␻2 Precise determination of the Casimir pressure be- I 0 a tween two parallel metallic plates by means of a micro- machined oscillator was performed in three successive where b is the lever arm, I is the moment of inertia, and 2 −1 experiments by Decca et al. These experiments do not b /I=1.2432±0.0005 ␮g . ␻ ␻ use the configuration of two parallel plates. The first The frequency shift r − 0 is directly measured and experiment was made with a Au-coated sphere and Cu- from Eq. ͑84͒ it leads to the calculated values of the ͒ ͑ ץ ץ coated plate ͑Decca, Fischbach, et al., 2003; Decca, Ló- Casimir force gradient Fs / a. Using the PFA 16 this pez, et al., 2003, Decca et al., 2004͒. The second experi- gradient can be expressed through the effective Casimir ment with several improvements used both a Au-coated pressure in the configuration of two parallel plates, sphere and a plate ͑Decca, López, Fischbach, et al., 2005; ͒ ͑ ץ ͒ F͑ץ ͒ a,T 1 Fs a,T Klimchitskaya et al., 2005 . Further improvements P͑a,T͒ =− =− . ͑85͒ aץ a 2␲Rץ implemented in the third experiment with a Au-coated sphere and Au-coated plate ͑Decca et al., 2007a, 2007b͒ made it the most precise and reliable measurement with Because of this, a direct measurement of the frequency metallic test bodies ever performed in the Casimir force shift caused by the Casimir force between a sphere and a measurements to date. Here we discuss the measure- plate results in an indirect measurement of the Casimir ment scheme and the main physical results following pressure in the configuration of two parallel plates. from the third experiment by Decca et al. ͑2007a, 2007b͒. The separation distance between the zero-roughness levels of Au layers on the plate and on the sphere ͑see In metrological terms this is an indirect measurement ͒ ͑Rabinovich, 2000͒ of the Casimir pressure between two Sec. III.B.1 is given by Au-coated parallel plates. Note that the results of direct ͑ ͒ ␪ ͑ ͒ a = zmeas − D1 + D2 − b . 86 measurements are found just from the experiment. The results of indirect measurements are obtained with the Here zmeas is the separation between the end of the ␪ help of calculations using the known equations which cleaved fiber and the platform and and D1,2 are de- relate the quantity under consideration with some quan- fined in Fig. 9. The lever arm b was measured optically. tities measured directly. The Casimir force per unit area The value of ␪ was determined by measuring the differ- ͑pressure͒ was determined in these experiments dynami- ence in capacitance between the plate and the right and cally by means of a micromechanical torsional oscillator left electrodes shown in Fig. 9. The value D1 +D2 was consisting of a 500ϫ500 ␮m2 heavily doped polysilicon measured as a part of the system calibration. This was plate suspended along one central planar axis by serpen- done by the application of voltages V to the sphere tine springs and a sphere of R=151.3 ␮m radius above while the plate was grounded. The electric force be- it attached to an optical fiber ͑see Fig. 9͒. During the tween a sphere and a plate was measured in the static measurements, the separation distance between the regime, with no harmonic variations of the separation sphere and the plate was varied harmonically, a˜͑t͒=a distance between them at large separations aϾ3 ␮m. ͑␻ ͒ ␻ +Az cos rt , where r is the resonant angular frequency The residual potential difference V0 was found to be of the oscillator under the influence of the Casimir force independent of separation ͓details of calibration proce- ͑ ͒ Ӷ ␻ Fs a,T from the sphere and Az /a 1. The frequency r dures are presented by Decca, López, Fischbach, et al. is related to the natural angular frequency of the oscil- ͑2005͒, and Decca et al. ͑2009͔͒. We emphasize that in ␻ ␲ϫ͑ ͒ ͑ ͑ ͒ lator 0 =2 713.25±0.02 Hz by Chan et al., 2001a, the second Decca, López, Fischbach, et al., 2005 and 2001b; Decca, Fischbach, et al., 2003; Decca, López, et third ͑Decca et al., 2007a, 2007b͒ experiments of this se- al., 2003͒ ries, contact between the sphere and the plate has not

Rev. Mod. Phys., Vol. 81, No. 4, October–December 2009

Klimchitskaya, Mohideen, and Mostepanenko: The Casimir force between real materials: … 1855

©

§ ¡ tical rule formulated in Sec. III.C.1. As a result, t expt s expt 0 ⌬ P ͑a͒=⌬ P ͑a͒ within the entire measurement -0.2 range, i.e., the total experimental error is determined by -0.4 only the systematic one. This means that the experiment by Decca et al. ͑2007a, 2007b͒ satisfies one of the main -0.6 requirements imposed on precise experiments in metrol- -0.8 ogy ͑Rabinovich, 2000͒. For now no other experiment -1 in Casimir physics satisfies this requirement. The total

200 300 400 500 600 700 relative experimental error varies from 0.19% at a

¡ £ ¥ § =162 nm to 0.9% at a=400 nm and to 9.0% at the larg- est separation a=746 nm. FIG. 10. The measured mean Casimir pressure between two To conclusively compare the experimental data with ͑ ͒ parallel plates vs separation dots . theory, the topography of metallic coatings both on the plate ͓body ͑1͔͒ and on the sphere ͓body ͑2͔͒ was inves- tigated using an AFM. From the AFM images, the frac- been achieved and was not needed for the determina- ͑1,2͒ ͑1,2͒ tion of absolute separations. tion of each surface area vi with the height hi was ͑ ͒ ͑1͒ A set of 120 curves of Fel a was then used to fit the determined. It was found that for the plate hi varies ͑2͒ quantity D1 +D2. Finally absolute separations a were from 0 to 18.35 nm and for the sphere hi varies from measured with an absolute error ⌬a=0.6 nm determined 0 to 10.94 nm. Using Eq. ͑65͒ the zero-roughness levels ͑1͒ at a 95% confidence. In contrast to previous measure- on the plate and on the sphere are H0 =9.66 nm and ͑ ͑2͒ ments Decca, Fischbach, et al., 2003; Decca, López, et H0 =5.01 nm. The contribution of correlation and dif- al., 2003; Decca, et al. 2004; Decca, López, Fischbach, et fractionlike effects in the roughness correction was al., 2005͒, in this experiment n=33 sets of measurements shown to be negligibly small ͑Decca, López, Fischbach, within a separation region from 162 to 746 nm were per- et al., 2005͒. The overall contribution of roughness was formed at almost the same intermediate separations ai shown to be only 0.5% of the Casimir pressure at a ͑1ഛiഛ293͒ in each set. This was made possible due to =162 nm and it decreases with the increase of separa- about 7% improvement in the vibrational noise, and an tion. To conclusively compare experimental data with improvement in the interferometric technique used to theory, the resistivity of the Au layers as a function of ͑ ͒ yield the distance zmeas see Fig. 9 . The use of a two- temperature was also measured in the region from T1 ⌬ color fiber interferometer yielded an error zmeas =3 to 400 K. This has led to a slightly modified values of ␻ ␥ =0.2 nm, and for every repetition of the Casimir pres- the Drude parameters, p =8.9±0.1 eV, =0.0357 eV, to ͑ ␻ sure measurement it was possible to reposition the be used in theoretical calculations compare with p ⌬ sample to within zmeas. =9.0 eV, ␥=0.035 eV in Sec. III.A.2͒. The mean values of the Casimir pressure ͑78͒ were The theoretical Casimir pressure as a function of determined from the frequency shifts using Eqs. ͑84͒ and separation between the plates was calculated using Eqs. ͑85͒ from data obtained in 33 sets of measurements. ͑74͒ and ͑8͒ with different approaches to the determina- They are plotted in Fig. 10 as a function of separation. tion of the dielectric permittivity along the imaginary The random experimental error was found from Eqs. frequency axis ͑see Sec. III.A.2͒ and using the Leonto- ͑78͒ and ͑79͒. It varies from ⌬rPexpt=0.46 mPa at a vich surface impedance ͑Sec. II.A͒. Comparison of ex- =162 nm to ⌬rPexpt=0.11 mPa at a=300 nm. This value periment with theory was performed in two different remains constant for separations up to a=746 nm. The ways described in Sec. III.C.2 ͑Decca et al., 2007a, systematic error in the indirect measurement of the 2007b͒. In Figs. 11͑a͒ and 11͑b͒ the mean experimental pressure is determined by errors in the measurements of are plotted as crosses within the separation the resonant frequency and of the radius of the sphere, region from 500 to 600 nm. For other separation regions and also by the error introduced using the PFA. The the situation is quite similar ͑Decca et al., 2007a͒. The latter is now related to experiment rather than to theory, arms of the crosses show the total experimental errors of because in this dynamic measurement PFA is a part of separation and Casimir pressures in true scales deter- the experimental procedure of the determination of the mined at a 95% confidence level. The light-gray band in Casimir pressure between the two parallel plates. Ac- Fig. 11͑a͒ shows the theoretical results computed using cording to the results presented in Sec. II.B, the error the generalized plasmalike dielectric permittivity, i.e., ␧ ͑ ␰͒ due to the use of PFA is less than a/R. Because of this, with Au i , as given by the solid line in Fig. 1. The Decca et al. ͑2007a, 2007b͒ estimated this error conser- width of the theoretical band in the vertical direction vatively as a/R. By combining all the above J=3 system- indicates the total theoretical error ͑Sec. III.C.2͒.Itis atic errors at a 95% confidence using the statistical rule equal to 0.5% of the calculated pressure and arises from ͑77͒, the resulting systematic error was obtained. It is the variations of the optical data of core electrons. given by ⌬sPexpt=2.12 mPa at a=162 nm, decreases to Other factors, such as patch potentials, were shown to 0.44 mPa at a=300 nm, and then to 0.31 mPa at a be negligible ͑Decca, López, Fischbach, et al., 2005͒. =746 nm. Finally, the total experimental error ⌬tPexpt͑a͒ Note that when the experimental data with the errors, at a 95% confidence level was obtained using the statis- shown as crosses, are compared with the theoretical

Rev. Mod. Phys., Vol. 81, No. 4, October–December 2009

1856 Klimchitskaya, Mohideen, and Mostepanenko: The Casimir force between real materials: …

©

   

©  © 

¡ ¥ §

 ¦ ¡ ¤ 

 ¦ 20 ¡

15

10

 # !

5

¡ £ ¥ §

0



     -5 -8

200 300 400 500 600 700

¡ £ ¤ ¦

-10

¨

©    ©   

¤   ¦

-12 ¡

!

-14 

¦ 20 ¡ -16

520 540 560 580 600 15

    

10 FIG. 11. Comparison of experiment and theory. The measured mean Casimir pressure together with the absolute errors in the separation and pressure vs separation is shown as crosses. ͑a͒ 5 The theoretical Casimir pressure computed using the general- ized plasmalike model and the optical data extrapolated by the 0 Drude model is shown by the light-gray and dark-gray bands, respectively. ͑b͒ The theoretical Casimir pressure computed -5 using different sets of optical data available vs separation is

200 300 400 500 600 700

¡ £ ¤ ¦ shown as the dark-gray band.

FIG. 12. Comparison of experiment and theory. The difference band computed for the entire measurement range, the between theoretical and mean experimental Casimir pressures errors ⌬a in the measurement of separations are irrel- vs separation ͑dots͒ and the 95% confidence intervals ͑solid evant to theory. The dark-gray band in Fig. 11͑a͒ shows lines͒. The theoretical Casimir pressures are computed ͑a͒ us- the theoretical results computed using the complete ing the generalized plasmalike model and ͑b͒ using the Leon- tabulated data extrapolated to low frequencies by means tovich surface impedance approach. ␧ ͑ ␰͒ of the Drude model, i.e., with Au i shown as the dashed line in Fig. 1. As shown in Fig. 11͑a͒, all data as the dark-gray band. Note that the values of the relax- crosses overlap with the light-gray band but are sepa- ation parameter only slightly influence the Casimir pres- rated by a gap from the dark-gray band. This means that sure. As shown in Fig. 11͑b͒, the use of any alternative the theory using the generalized plasmalike permittivity value of ␻ makes the disagreement between the Drude is consistent with experiment within the limits of the p model approach and the data even more acute. experimental error. At the same time, a theory using the Another way to compare theory with experiment complete optical data extrapolated by the Drude model considered in Sec. III.C.2 is illustrated in Fig. 12, where is excluded by the data at a 95% confidence level. theor͑ ͒ expt͑ ͒ As mentioned in Sec. III.C.2, the choice of different the differences P ai −P ai are shown as dots sets of Drude parameters may lead to an up to 5% varia- and solid lines indicate the 95% confidence interval ͓ ⌶ ͑ ͒ ⌶ ͑ ͔͒ theor͑ ͒ expt͑ ͒ tion in the theoretical values of the Casimir pressure − P a , P a for the quantity P ai −P ai ͑ ͒ theor ͑ ͒ ͑Pirozhenko et al., 2006͒. According to Sec. III.C.2, the Decca et al., 2007b . The values of P in Fig. 12 a hypothesis that the thin films used in the experiment are computed using the generalized plasmalike model, possess such unusual Drude parameters is not supported whereas the values of P¯ theor in Fig. 12͑b͒ using the Le- by the data. Nevertheless, the disagreement between the ontovich surface impedance ͑Decca, López, Fischbach, Drude model approach and the experimental data only et al., 2005͒. As shown in Fig. 12, both theoretical ap- increases if some of the alternative Drude parameters, proaches are consistent with the data. Note that the val- as considered by Pirozhenko et al. ͑2006͒, are used. As ues of theoretical pressures in Fig. 12 are computed at ͑ ͒ an illustration, in Fig. 11 b the computation results ob- the experimental separations ai. As a result, these values ␻ ⌬ tained using the Drude model approach with p varying are associated with the additional errors 4 a/ai dis- from 6.85 ͑Pirozhenko et al., 2006͒ to 9.0 eV are plotted cussed in Sec. III.C.2. They are the primary theoretical

Rev. Mod. Phys., Vol. 81, No. 4, October–December 2009

Klimchitskaya, Mohideen, and Mostepanenko: The Casimir force between real materials: … 1857

Ñ Ó 

  Ó Ö Ü Ô Ø Ø 



È È ´ Ñ È  µ

È È ´ Ñ È  µ

14 30 12 25 10 20 8 15 6 10 4 5 2

300 400 500 600 700

200 300 400 500 600 700

 ´ Ò Ñ µ

 ´ Ò Ñ µ FIG. 14. Comparison of experiment and theory. The difference FIG. 13. Comparison of experiment and theory. The difference between theoretical Casimir pressures using the effect of between theoretical Casimir pressure calculated with the opti- charge screening for the transverse magnetic mode and mean cal data extrapolated using the Drude model and mean expri- experimental data vs separation is shown as dots. The solid mental Casimir pressure vs separation. The solid and dashed and dashed lines indicate the borders of the 95% and 99.9% lines indicate the borders of the 95% and 99.9% confidence confidence intervals, respectively. intervals, respectively.

Lamoreaux, 2008͒. A charge carrier density of nϷ5.9 errors at short separations in this approach to comparing ϫ1022 cm−3 at T=300 K and the Thomas-Fermi screen- theory with experiment. The respective width of the ing length were used in the computations for Au. In Fig. confidence interval ͓−⌶ ͑a͒,⌶ ͑a͔͒ is overestimated. P P 14 the differences between the computed theoretical Ca- The actual agreement of the theoretical approach with simir pressures and the experimental data by Decca et data can be characterized by the deviation of the differ- al. ͑2007a, 2007b͒ are shown as dots. As shown, the data theor͑ ͒ expt͑ ͒ ences P ai −P ai from zero. At the shortest exclude the theoretical approach using the modification separations this deviation exceeds the total experimental ͑40͒ at a 95% confidence level within the entire mea- ͑ error but for the theory using the generalized plasma- surement region from 160 to 750 nm and at a 99.9% ͒ Ͼ like permittivity at a 300 nm both are approximately confidence level within the measurement region from equal. The measure of agreement between experiment 160 to 640 nm. Note that for metals the same theoretical ⌶ ͉ ¯ expt͉ and theory given by P / P , as defined in Sec. III.C.2, results, as in the approach by Dalvit and Lamoreaux is equal to 1.9% and 1.8% at separations a=162 and ͑2008͒, are obtained in the approaches by Pitaevskii 400 nm, respectively. ͑2008͒ and Svetovoy ͑2008͒. Thus, the last two ap- In Fig. 13 the same method of comparing theory with proaches are also inconsistent with the experimental experiment is applied to the theoretical approach based data by Decca et al. ͑2007a, 2007b͒. on the Drude model. The solid and dashed lines indicate Another experiment performed by means of the mi- the limits of the 95% and 99.9% confidence intervals, cromechanical torsional oscillator was the test of correc- respectively ͑Decca et al., 2007b͒. The differences tions to the PFA ͑Krause et al., 2007͒. Taking these cor- theor͑ ͒ expt͑ ͒ rections into account, the Casimir force between a PD ai −P ai are shown as dots. As shown in the figure, the theoretical approach using the extrapolation sphere of radius R and a plate can be presented in the ͑ ͒ of the optical data to low frequencies by means of the form Scardicchio and Jaffe, 2006 Drude model is experimentally excluded at a 95% con- ͑ ͒ ␲ ͑ ͓͒ ␤ ͑ 2 2͔͒ ͑ ͒ fidence level for the entire measurement range from Fs a,R =2 RE a 1+ a/R + O a /R . 87 theor͑ ͒ 162 to 746 nm. Also note that the differences PD ai expt͑ ͒ −P ai are outside the 99.9% confidence interval at Here E͑a͒ is the Casimir energy per unit area of two separations from 210 to 620 nm. Finally Decca et al. parallel plates and ␤ is a dimensionless parameter char- ͑2007a, 2007b͒ concluded that their experiments cannot acterizing the lowest-order deviation from the PFA. The be reconciled with the Drude model approach to the constraints on the parameter ␤ in Eq. ͑87͒ can be ob- thermal Casimir force. They also remarked that the ex- tained from the static measurements of the Casimir periments with the micromachined torsional oscillator force. In these measurements the separation distance are not of sufficient precision to measure the small ther- between the sphere and the plate is not varied harmoni- ͑ ͒ mal effect between the two metal bodies at separations cally and the force Fs a,R is a directly measured quan- of a few hundred nanometers, as predicted by the gen- tity. eralized plasmalike model. The proposed experiments Dynamic measurements of the Casimir pressure de- capable of this goal are considered in Sec. IV.D. scribed above in this section are more precise than the The experimental data by Decca et al. ͑2007a, 2007b͒ static ones. Substituting the Casimir force ͑87͒ into the were also compared with the theoretical results obtained right-hand side of Eq. ͑85͒, one obtains the following using the modification of the transverse magnetic reflec- expression for the effective Casimir pressure ͑Krause et tion coefficient in accordance with Eq. ͑40͒͑Dalvit and al., 2007͒:

Rev. Mod. Phys., Vol. 81, No. 4, October–December 2009 1858 Klimchitskaya, Mohideen, and Mostepanenko: The Casimir force between real materials: …

sidual potential difference V=430 mV was measured. Peff͑a,R͒ = P͑a͓͒1+␤˜ ͑a͒a/R + O͑a2/R2͔͒, ͑88͒ The corresponding large electrostatic force was compen- where P͑a͒ is the Casimir pressure between two parallel sated with application of voltage to the lens. However, plates and the dimensionless quantity ␤˜ ͑a͒ is given by there appears to have been a large residual electrostatic force even after this compensation, which was deter- ␤˜ ͑a͒ = ␤͓1−E͑a͒/aP͑a͔͒. ͑89͒ mined only by fitting to the total force including the Casimir force above 1 ␮m. “Typically, the Casimir force ˜ Note that for ideal metal bodies ␤͑a͒=2␤/3=const. had magnitude of at least 20% of the electrical force at To obtain constraints on ␤ and ␤˜ a series of both static the point of closest approach” ͑Lamoreaux, 1997͒. The and dynamic measurements has been performed with uncertainty in the measurement of absolute separations Au-coated plate and spheres with radii R=10.5, 31.4, ⌬a “was normally less than 0.1 ␮m” ͑Lamoreaux, 1997͒. 52.3, 102.8, and 148.2 ␮m. The static measurement of Data were compared with theory for the ideal metal the Casimir force between the sphere and the plate was lens and plate. The conclusion reached that there is an performed at separations from 160 to 750 nm in 10 nm agreement at the level of 5% in the 0.6–6 ␮m range steps. Dynamic determination of the effective Casimir ͑Lamoreaux, 1997͒ was found to be incompatible with pressure between two parallel plates was done at sepa- the magnitude of the thermal correction to the Casimir rations from 164 to 986 nm with 2 nm steps. The influ- force ͑Bordag et al., 1998͒. Thus, at separations of 4, 5, ences of the effects of the nonzero skin depth and sur- and 6 ␮m the thermal corrections are 86%, 129%, and face roughness on the dominant first-order correction to 174% of the zero-temperature force, respectively. The the PFA were estimated to be of order 10% and 1%, data, however, were found to be “not of sufficient accu- respectively. Comparison between data and theory at racy to demonstrate the finite temperature correction” ͑Lamoreaux, 1997͒. From this it follows that the agree- separations aϽ300 nm leads to the result ͉␤˜ ͑a͉͒Ͻ0.4 at a ment of the data with the theory at the level of 5–10% 95% confidence level. In the same separation region, may exist only at separations of about 1 ␮m. Here the ͉␤͉Ͻ0.6 was obtained ͑Krause et al., 2007͒. These con- thermal correction is relatively small, whereas the larger straints are compatible with the exact results for a corrections due to the skin depth and surface roughness cylinder-plate configuration but are not in agreement have the opposite sign and partially compensate each with the extrapolations made for a sphere above a plate other. Keeping in mind that theoretical forces calculated ͑see the discussion in Sec. II.B͒. at experimental separations are burdened with an addi- tional error of about 3⌬a/aϷ30% at a=1 ␮m ͓see Sec. C. Other experiments on the Casimir force between metal III.C.2 and Iannuzzi, Gelfand, et al. ͑2004͔͒, the errors in expt͑ ͒ theor͑ ͒ bodies the differences F ai −F ai , as shown in Fig. 4, bottom ͑Lamoreaux, 1997͒, are significantly underesti- 1. Torsion pendulum experiment mated. Chronologically, the first experiment, in the more re- The results of this experiment at about 1 ␮m separa- cent series of the Casimir force measurements, was per- tion were used ͑Torgerson and Lamoreaux, 2004; Lam- formed by Lamoreaux ͑1997͒. While this experiment re- oreaux, 2005͒ to exclude the theoretical approach to the kindled interest in the investigation of the Casimir force thermal Casimir force which uses the Drude model at and stimulated further development of the field, the re- low frequencies. The latter predicts −19% thermal cor- sults obtained contain several uncertainties. The Casimir rection at a=1 ␮m ͑to be compared with a 1.2% thermal force between a Au-coated spherical lens and a flat plate correction in the case of ideal metals͒ which was not was measured using a torsion pendulum. A lens with a experimentally observed. The conclusion that the Drude radius R=11.3±0.1 cm ͓later corrected to 12.5±0.3 cm model approach is inconsistent with data is in agreement ͑Lamoreaux, 1998͔͒ was mounted on a piezo stack and a with the same conclusion obtained at a high confidence plate on an arm of the torsion balance in vacuum. The level in the much more precise experiments of Decca, other arm of the torsion balance formed the center elec- Fischbach, et al. ͑2003͒, Decca et al. ͑2004, 2007a, 2007b͒, trode of a dual parallel plate capacitor. The positions of and Decca, López, Fischbach, et al. ͑2005͒ ͑see Sec. this arm and consequently the angle of the torsion pen- IV.B͒. dulum were controlled by the application of voltages to the plates of the dual capacitor. The Casimir force be- 2. Micromechanical devices actuated by the Casimir force tween the plate and the lens would result in a torque, leading to a change in the angle of the torsion balance. Chan et al. ͑2001a͒ used a micromachined torsional This change results in changes of the capacitances. Then device actuated by the Casimir force. A more advanced compensating voltages were applied to these capaci- version of this device shown in Fig. 9 was later used in tances to counteract the change in the angle of the tor- precise measurements of the Casimir force ͑see Sec. sion balance. These compensating voltages were a mea- IV.B͒. When the sphere in Fig. 9 was moved closer to the sure of the Casimir force. plate in a vacuum with a pressure of less than 1 mtorr, The calibration of the measurement system was done the Casimir force, acting on the plate, tilted it about its electrostatically. When the lens and plate surfaces were central axis toward the sphere. Thus, vacuum oscilla- grounded, a “shockingly large” ͑Lamoreaux, 1997͒ re- tions of the electromagnetic field led to the mechanical

Rev. Mod. Phys., Vol. 81, No. 4, October–December 2009 Klimchitskaya, Mohideen, and Mostepanenko: The Casimir force between real materials: … 1859

͒ ͑ ץ ͒ ͑ ץ␣ ␯2 ͑ ͒ ␯2 ␯2⌬

expt a = − 0 =− P a / a. 90  ␯ Here 0 =138.275 Hz is the natural frequency in the ab- ␣ ␲2 sence of the Casimir pressure P and =S/4 meff Ϸ0.0479 m2 /kg ͑S is the area of a capacitor formed by ͒ the plates and meff is the effective mass . FIG. 15. Schematic diagram for the model of oscillator actu- Since the measurement is a dynamic one, the directly ated by the Casimir force. measured quantity is the frequency shift ͑90͒ arising from the effect of the Casimir force. This frequency shift of the plate demonstrating the first microme- is related to the gradient of the Casimir pressure using ͑ ͒ chanical device driven by the Casimir force. Using a mi- Eq. 90 . Thus, though this experiment uses the configu- cromachined oscillator, Chan et al. ͑2001a͒ measured the ration of two parallel plates, it is an indirect measure- Casimir force between a Au-coated polystyrene sphere ment of the gradient of the Casimir pressure between and a plate. Comparison of the measurement data with the plates, similarly to the experiments using the sphere- the ideal metal Casimir result demonstrated again the plate configuration in the dynamic regime for the same ͑ ͒ ͑ ͒ influence of skin depth and surface roughness on the purpose see Sec. IV.B . Bressi et al. 2002 did not aim Casimir force between real material bodies ͓previously to restore the Casimir pressure from the pressure gradi- ⌬␯2 ͑ ͒ demonstrated by Mohideen and Roy ͑1998͒ and Harris ent. Instead, they fitted the experimental data expt a ⌬␯2 ͑ ͒ et al. ͑2000͒; see Figs. 5 and 6͔. to theor a computed from the theoretical dependence A similar device was used to investigate the influence of the Casimir pressure between two ideal metal plates 4 of the Casimir force on the oscillatory behavior of mi- −KC /a with a free parameter KC. The best fit resulted ͑ ͒ϫ −27 2 cromachines ͑Chan et al., 2001b͒. The simple model of in KC = 1.22±0.18 10 Nm . This was compared the Casimir oscillator is shown in Fig. 15. It consists of a with the exact Casimir coefficient for ideal metal plates ͑ ͒ ␲2ប ϫ −27 2 movable metallic plate which is subject to both the re- in Eq. 2 , KC = c/240=1.3 10 Nm . The conclu- storing force of a spring and the Casimir force from the sion was drawn that the related force coefficient is de- interaction with a fixed metallic sphere. The force from termined at the 15% precision level ͑Bressi et al., 2002͒. the spring is linear in the movement of the plate ⌬a, From the point of view of a general method for the whereas the Casimir force is nonlinear in ⌬a. The poten- comparison of experiment with theory, which was devel- tial energy of this microdevice possesses local and global oped after this experiment was performed, it would be ⌬␯2 ͑ ͒ minima separated by the potential barrier. The Casimir reasonable not to use any fit but to compare expt a ⌬␯2 ͑ ͒ force changes the resonant frequency of oscillations with the exact expression for theor a with no adjust- ͑ ͒ around the local minimum and makes the oscillations able parameters such as KC. In Fig. 16 a the results of anharmonic ͑Chan et al., 2001b͒. These properties may such comparison are presented where the experimental be useful in future micromechanical systems driven by data are shown as crosses and the solid line shows the Casimir force. ⌬␯2 theor computed from the exact expression for the Ca- simir pressure between two ideal metal plates, as given in Eq. ͑2͒. As shown, at separations below 1 ␮m the 3. The experiment using the parallel plate configuration experimental crosses only touch the solid line. This can The only experiment in the recent series of Casimir be explained by the role of the nonzero skin depth. If ͑ ͒ ͑ ͒ force measurements which uses the original configura- instead of P a =P0 a one uses the Casimir pressure tion of two plane plates was performed by Bressi et al. with the first and second order corrections due to the ͑ ͑2002͒. A Si cantilever and a thick plate rigidly con- nonzero skin depth Bezerra, Klimchitskaya, and ͒ nected to a frame ͑source͒ both covered with a Cr layer Mostepanenko, 2000 , with adjustable separation distance between them were used as two plates. The coarse separation distance was ␲2បc 16 c c2 adjusted with a dc motor and fine tuning was achieved P͑a͒ =− ͩ1− +24 ͪ, ͑91͒ 4 ␻ ␻2 2 using a linear piezoelectric transducer attached to the 240a 3 pa pa frame. Calibration was performed using the electrostatic force. As a result, the error in absolute separation was found to be ⌬a=35 nm. Small oscillations induced on the agreement with data improves. As an illustration, ⌬␯2 the source by the application of a sinusoidal voltage to theor is recalculated using the theoretical Casimir pres- ͑ ͒ ␻ Ϸ the piezo induce oscillations of the cantilever through sure 91 with the p 13 eV for Cr. The results are the Casimir force. Thus, this experiment is dynamical shown as the solid line in Fig. 16͑b͒. It is seen that the like those considered in Sec. IV.B. The motion of the Casimir pressure taking the skin depth into account is in resonator placed in the vacuum was detected by means much better agreement with the data than the ideal of a fiber optic interferometer. After subtracting the metal Casimir pressure. Suggestions on how to improve electrostatic forces, the residual frequency shift is given the sensitivity of this experiment are discussed in Sec. by IV.D.

Rev. Mod. Phys., Vol. 81, No. 4, October–December 2009

1860 Klimchitskaya, Mohideen, and Mostepanenko: The Casimir force between real materials: …

©

 § ¡ films using the Lifshitz theory. This result is not surpris- 0 ing because for films of small thickness the effects of spatial dispersion should be taken into account which -1000 are not included in the Lifshitz theory ͑Klimchitskaya et ͒

-2000 al., 2000 .

 § -3000 ¡ 5. Dynamic measurement using an atomic force microscope Jourdan et al. ͑2009͒ recently performed a measure-

0.5 1 1.5 2 2.5 3

¡ £ ¥ § ment of the Casimir force between a Au-coated sphere

of R=20 ␮m radius and a plate using an AFM. No direct

©

 § ¡ sphere-plate contact was used to determine absolute 0 separation distances between the surfaces. The AFM cantilever with the glued sphere was considered as a har- -1000 monic oscillator with the natural resonant frequency ␻ ␲ϫ -2000 0 =2 50 182 rad/s which is modified by the effect of

the Casimir force. In these two aspects ͑direct contact

 § -3000 ¡ between the two surfaces was avoided and the dymanic measurement mode was used͒ the experimental ap-

0.5 1 1.5 2 2.5 3 proach resembles the one used by Decca, López, Fisch-

¡ £ ¥ § bach, et al. ͑2005͒ and Decca et al. ͑2007a, 2007b͒ in the measurements of the Casimir pressure by means of a FIG. 16. Comparison of experiment and theory. The measured micromechanical torsional oscillator. The uncertainty in frequency shift vs separation together with the absolute errors the absolute separations ⌬a=2 nm was, however, more in separation and ⌬␯2 is shown as crosses. Solid lines indicate ͑ ͒ than three times larger than in experiments with a mi- the theoretical frequency shift calculated for a ideal metal ͑ ͒ plates and ͑b͒ real metals with inclusion of the skin-depth cor- cromachined oscillator see Sec. IV.B . The experimental rection. data for the force gradient were compared with the Lif- shitz theory at zero temperature. The dielectric permit- tivity along the imaginary frequency axis was found us- 4. The Casimir force between thin metallic films ing the complete optical data ͑Palik, 1985͒ extrapolated to zero frequency by means of the Drude model with Lisanti et al. ͑2005͒ reported the observation of the ␻ ␥ ͑ p =9.0 eV and =0.035 eV Lambrecht and Reynaud, skin-depth effect on the Casimir force between metallic 2000͒. The correction due to surface roughness was not surfaces. The Casimir force between a thick plate and a taken into account. The discrepancy between the force ␮ 100- m-radius polystyrene sphere coated with metallic gradient measurements and the results of the theoretical films of different thicknesses was measured. The sphere computations described above was found to be within was positioned above a micromachined torsional bal- 3% of the theoretical force at separations between 100 ance. The Casimir attraction between the sphere and the and 200 nm. The measured Casimir force gradient was top plate of the balance induced a rotation angle which also compared with that computed using ideal metal sur- was measured as a function of the separation between faces and a deviation was reported. On this basis it was the surfaces. Without the indication of errors and confi- concluded that the experimental data demonstrate again dence levels, it was reported that the Casimir attraction the finite conductivity ͑skin-depth͒ effects on the Ca- between the metallic plate and the sphere with coatings simir force. thinner than the skin depth is smaller than that of the same plate and a sphere with thick metal coating. Physi- 6. Ambient Casimir force measurements cally, this is in fact the same effect as is demonstrated in Figs. 5 and 6 in Sec. IV.A where the Casimir forces act- In this recent period two other ambient ͑open to air͒ ing between ideal and real metals were compared with experiments measuring the Casimir force were reported. corrections based on the plasma frequency ͑finite con- Both did not use a vacuum environment and reported ductivity corrections͒. Ideal metals are better reflectors the presence of water layers on the interacting surfaces. and the magnitude of the Casimir force between them is The first of these was the experiment by Ederth ͑2000͒, larger than between real metals. In a similar manner where the force was measured between two cylindrical thick real metal films are better reflectors than thin real template stripped gold surfaces with 0.4 nm roughness, metal films. in a distance range from 20 to 100 nm. The 200 nm gold Lisanti et al. ͑2005͒ compared their experimental re- films were fixed to 10 mm radius silica cylinders using a sults with the Lifshitz theory at zero temperature “soft glue.” In addition, a hydrocarbon layer of hexade- adapted for the description of layered structures ͑Zhou canethiol was applied to the interacting surfaces. It was and Spruch, 1995; Klimchitskaya et al., 2000; Tomaš, noted that this top hydrocarbon layer was necessary to 2002͒. It was shown that the experimental forces ob- preserve the purity of the gold surface in the ambient tained for films of thickness smaller than the skin depth environment. This hydrocarbon organic layer prevented have smaller magnitudes than those computed for such a direct measurement of the electrostatic forces or an

Rev. Mod. Phys., Vol. 81, No. 4, October–December 2009 Klimchitskaya, Mohideen, and Mostepanenko: The Casimir force between real materials: … 1861 independent determination of the surface separation. Capasso, 2007͒. The experimental data obtained were Also the presence of the hydrocarbon layer means that compared with the Lifshitz theory taking into account the experiment cannot be strictly classified as that be- the frequency dependence of the dielectric functions of tween two gold surfaces. One of the cylinders was Au and ethanol and the correction due to surface rough- mounted on a piezo and the other on bimorph deflection ness. Consistency of the obtained data with Lifshitz’s sensor. The charge produced by the bimorph in response theory was claimed although at separations below 50 nm to the deflection induced by a force on the interacting a disagreement has been observed which increases with surfaces was measured by an electrometer amplifier. The the decrease of separation. However, as discussed soft glue led to a deformation which was estimated to be ͑Geyer, Klimchitskaya, Mohideen, and Mostepanenko, 18–20 nm for glue thicknesses used. To compensate for 2008͒, the theoretical computations of the Casimir force this deformation the two measured force-distance curves between the smooth Au surfaces separated by ethanol were shifted, one by 9 nm and the other by 12 nm to done according to the method provided by Munday and overlap the calculations. They reported that “it is not Capasso ͑2007͒ ͓i.e., by the use of the Kramers-Kronig possible to establish with certainty” the validity of the relations and tabulated optical data ͑Palik, 1985͔͒ lead to displacement due to the deformation and that it “dimin- a discrepancy up to 25% with respect to the reported ishes the strength of the measurement as a test of the theoretical results. The latter can be reproduced if one Casimir force and also precludes a quantitative assess- uses at all imaginary frequencies the Drude dielectric ment of the agreement between theory and experi- function ͑35͒ for both the sphere and plate materials. A ment.” second drawback is that the effect of the residual poten- A second experiment done in an ambient environ- tial difference between the sphere and the plate was cal- ment using the AFM for separation distances between culated incorrectly and significantly underestimated by a 12 and 200 nm was recently reported ͑van Zwol et al., factor of 590. Finally, the possible interaction between 2008͒. Here a Au-coated sphere of R=20 ␮m radius was the double layer formed in liquids due to salt impurities, fixed to a gold-coated AFM cantilever. The Au-coated which would decrease the electrostatic force, was not plate was mounted on the piezo. Both the sphere and taken into account without any justification. The result- plate were coated with 100 nm of Au. The optical prop- ing electrostatic force is of the same magnitude as the erties of the Au-coated plate were measured with Casimir force to be measured. All this makes the inter- an ellipsometer in the wavelength region from 137 nm pretation of this experiment uncertain ͑Geyer, Klim- to 33 ␮m and fitted to obtain the plasma frequency chitskaya, Mohideen, and Mostepanenko, 2008͒.Inthe of 7.9±0.2 eV and a relaxation frequency of reply ͑Munday and Capasso, 2008͒ it was recognized 0.048±0.005 eV since the finite conductivity corrections that the original paper ͑Munday and Capasso, 2007͒ did for the separation range considered are large. The in fact use the Drude model. It was also recognized that roughness or the water layer was not taken into account the equation originally used to estimate the residual in the fit. The errors in the cantilever spring constant electrostatic force “is not strictly correct” and that salt and the deflection coefficient were reported to be 4% contaminants exist even in the purest solutions leading and 3%, respectively, which together were reported to to the screening of electrostatic interaction. In a later lead to errors of 4–10 %. The calibration errors were work ͑Munday et al., 2008͒ the effect of electrostatic reported to lead to an overall error of about 5–35%. The forces and Debye screening on the measurement of the electrostatically measured contact potential 10±10 mV Casimir force in fluids was further investigated. The was reported to lead to a 10% error. They reported that electrostatic force with account of Debye screening was they were not able to independently determine the sepa- calculated as outlined by Geyer, Klimchitskaya, Mo- ration on contact of the two surfaces due to the stiff hideen, and Mostepanenko ͑2008͒. The influence of the cantilevers employed. Based on the roughness they esti- concentration of salt impurities on the Debye screening mated a 1 nm error in the contact separation “leading was investigated. No measurable change in the force therefore to a 28% relative error at the smallest separa- with or without grounding of the sphere was reported. tions.” However, a general 10% agreement with the The Casimir force was measured 51 times with the stan- theory was reported below 100 nm separation. Given dard deviation which varies from 130 to 90 pN when the ambient nature of the experiment, “typically a few separation increases from 30 to 80 nm ͑the variances are nanometers,” water layers on both surfaces were present less by a factor of ͱ51͒. As a result, the relative random but not treated in the theoretical comparison or system- experimental error of the Casimir force measurements atic errors. Repeating the experiments in a vacuum en- in this experiment determined at a 67% confidence level vironment should allow for a more definitive compari- is 7% at a=30 nm and increases to 60% at a=80 nm. son. Bearing in mind that for aϽ30 nm they recognized de- viations between the theory and the measurement data, this experiment can be considered as only a qualitative 7. Related measurements demonstration of the Casimir force in fluids. An interesting preliminary test of the Lifshitz theory Another experiment used an adaptive holographic in- for three-layer systems is the measurement of the attrac- terferometer to measure periodical nonlinear deforma- tive Casimir force between a Au-coated sphere and a tions of a thin pellicle caused by an oscillating Casimir plate immersed in ethanol using an AFM ͑Munday and force due to a spherical lens ͑Petrov et al., 2006͒. Both

Rev. Mod. Phys., Vol. 81, No. 4, October–December 2009 1862 Klimchitskaya, Mohideen, and Mostepanenko: The Casimir force between real materials: …

test bodies were coated with a thin Al film and placed in serious experimental difficulties associated with the par- a vacuum chamber. The lens was mounted on a vibrating allelism of the plates͒ and a sphere above a plate. In piezodriver. As a result, the oscillations of the lens posi- addition, as discussed in Sec. II.B, the exact solution for tion led to a periodic modulation of the Casimir force. the ideal metal cylinder-plate configuration has been ob- The experimental data were found to be in qualitative tained recently. This gives the possibility to determine agreement with the theory based on ideal metal bound- the accuracy of the PFA and to apply it to real materials aries at separation distances of a few hundred nanom- with high reliability. Finally it was concluded that using eters. Corrections due to the finite skin depth and sur- face roughness were not provided. Also the use of Al the dynamic measuring scheme it is possible to measure whose surface undergoes rapid oxidation even in rela- the Casimir force between a plate and a cylinder at sepa- tively high vacuum adds uncertainty to the results of this rations of about 3 ␮m with a precision of a few percent experiment. However, the new measurement technique ͑Brown-Hayes et al., 2005͒. used may be promising for future measurements of the Another proposal suggests to use a highly sensitive Casimir force. torsion balance in the separation range from 1 to 10 ␮m to measure the Casimir force in the configuration of two parallel plates ͑Lambrecht et al., 2005͒. The construction D. Future prospects to measure the thermal Casimir force of the balance is similar to that used in the Eötvos-type experiments aimed to test the equivalence principle. It is As described in Secs. IV.A–IV.C, the measurements of planned to measure the thermal Casimir force with an the Casimir force between macroscopic bodies per- formed to date were not of sufficient precision to mea- accuracy of a few percent and to discriminate between sure the magnitude of the thermal effect. The experi- different theoretical approaches discussed. ments using the micromachined oscillator ͑Sec. IV.B͒ One more experiment exploiting the two-parallel- possess the highest experimental precision at separa- plate configuration is proposed at separations larger tions below 1 ␮m. They have been used to exclude ther- than a few micrometers ͑Antonini et al., 2006͒. The ex- mal effects, as predicted by models using the Drude perimental scheme is based on the use of a Michelson- model at low frequencies. However, thermal effects pre- type interferometer and the dynamic technique with one dicted by the generalized plasmalike model at short oscillating plate. Calibrations show that a force of 5 separations remain below the experimental sensitivity. ϫ10−11 N can be measured in this setup with the relative In this respect, large separation measurements of the error from about 10% to 20%. This would be sufficient Casimir force would be of great interest. At separations ␮ of a few micrometers the thermal regime is reached to measure the thermal effect at a separation of 5 m ͑ ͒ where the Casimir free energy is entirely of thermal ori- Antonini et al., 2006 . gin. Calculations using the plasma model or the general- The thermal effect in the Casimir force can be mea- ized plasmalike model result in Eq. ͑34͒. The Drude sured at short separations below 1 ␮m if the difference model approach to the thermal Casimir force leads to in the thermal forces ⌬F at different temperatures rather only one-half of the result for ideal metal plates ͓Eq. than the absolute value of the thermal Casimir force is ͑ ͒ ␻ →ϱ͔ 34 with p . Thus, large-separation measurements measured ͑Chen et al., 2003͒. For real metals, this differ- of the Casimir force would bring direct information to ence of the thermal Casimir forces ͑in contrast with the bear on the magnitude of the thermal effect between relative thermal correction͒ does not increase but de- macroscopic bodies ͑such a measurement in the atom- creases with increasing separation distance. This allows plate configuration has been performed already; see Sec. VI.A͒. the observation of the thermal effect on the Casimir In Sec. IV.C.1 it was shown that the experiment using force at small separations of about 0.5 ␮m where the the torsion pendulum within the region from relative thermal correction, as predicted by the general- 0.6 to 6 ␮m is in fact uncertain at separations above ized plasmalike model, is rather small. Preliminary esti- 2 ␮m where the thermal effects begin to make a sub- mation shows that with a sphere of R=2 mm radius at- stantial contribution. An analysis ͑Lamoreaux and But- tached to a cantilever of an AFM the measurable tler, 2005͒ shows that the torsion pendulum technique changes of the force amplitude of order 10−13 N are has the potential to measure the Casimir force between achievable from a 50 K change in the temperature. Such ␮ a plate and a spherical lens at a=4 m with a relative a temperature difference can be obtained by the illumi- ␮ error of 10%. Bearing in mind that at a=4 m the ther- nation of the sphere and plate surfaces with laser pulses mal correction contributes as much as 86% of the zero- of 10−2 s duration ͑Chen, Mohideen, and Milonni, 2004͒. temperature Casimir force, such an experiment, if suc- cessfully performed, holds great promise. The idea of exploiting difference force measurements to There is a proposal aimed at measuring the Casimir probe the thermal Casimir effect in superconducting force in the cylinder-plate configuration at separations cavities was proposed by Bimonte ͑2008͒. around 3 ␮m ͑Brown-Hayes et al., 2005͒. This geometry From the above discussion it is clear that experimen- can be considered as a compromise between the two- tal investigations of thermal effects in the Casimir force parallel-plate configuration ͑which is connected with appear feasible in the near future.

Rev. Mod. Phys., Vol. 81, No. 4, October–December 2009 Klimchitskaya, Mohideen, and Mostepanenko: The Casimir force between real materials: … 1863

Force difference

Vacuum 640 nm Optical filter chamber laser Photo-

Lock-in amplifier diodes Cantilever

Sphere

a Function generator 100Hz piezo piezo

514nm AOM Ar laser

FIG. 17. Schematic of the experimental setup on measuring the difference Casimir force between Au sphere and Si plate illuminated with laser pulses.

V. CASIMIR FORCE BETWEEN A METALLIC SPHERE force measurements. In addition, as semiconductors are AND A SEMICONDUCTOR PLATE the basic fabrication materials for nanotechnology, the use of semiconductor surfaces for the control of the Ca- A. Motivation for use of semiconductors simir force will lead to many applications. A vital issue in many applications of the Casimir ef- fect is how to control the magnitude of the force by B. Optically modulated Casimir force changing the parameters of the system. A natural method for this control is to change the material prop- The modification of the Casimir force through the ra- erties of the interacting bodies. Arnold et al. ͑1979͒ diation induced change in the carrier density was first ͑ ͒ made an attempt to modify the van der Waals and Ca- demonstrated by Chen et al. 2007a, 2007b . A high- simir forces between semiconductors with light. Attrac- vacuum-based AFM was employed to measure the tive forces were measured between a glass lens and a Si change in the Casimir force between a Au-coated sphere plate and also between a glass lens coated with amor- and a Si membrane in the presence and in the absence of phous Si and a Si plate. The glass lens, however, is an incident light. An oil-free vacuum chamber with a pres- ϫ −7 insulator and therefore the electric forces, such as due to sure of around 2 10 torr was used. A polystyrene ␮ potential differences, could not be con- sphere of diameter 2R=197.8±0.3 m coated with a Au trolled. This might explain why Arnold et al. ͑1979͒ layer of 82±2 nm thickness was mounted at the tip of a found no force change occurred on illumination at sepa- 320 ␮m conductive cantilever ͑the general scheme of the rations below 350 nm, where it should have been most experiment is shown in Fig. 17͒. A specially fabricated Si pronounced. One more attempt to modify the Casimir membrane ͓see Chen et al. ͑2007b͒ for preparation de- force was made by Iannuzzi, Lisanti, and Capasso ͑2004͒ tails͔ was mounted on top of the piezo which is used to when measuring the Casimir force acting between a change the separation distance a between the sphere plate and a sphere coated with a hydrogen-switchable and the membrane from contact to 6 ␮m. The excitation mirror that becomes transparent upon hydrogenation. of the carriers in the Si membrane was done with Despite expectations, no significant decrease of the Ca- 5-ms-wide light pulses ͑50% duty cycle͒. These pulses simir force owing to the increased transparency was ob- were obtained from a cw Ar ion laser light at 514 nm served. The negative result is explained by the Lifshitz wavelength modulated at a frequency of 100 Hz using an theory, which requires a change of the reflectivity prop- acousto-optic modulator ͑AOM͒. The laser pulses were erties within a wide range of frequencies in order to focused on the bottom surface of the Si membrane. The markedly affect the magnitude of the Casimir force. This Gaussian width of the focused beam on the membrane requirement is not satisfied by hydrogenation. was measured to be w=0.23±0.01 mm. The resulting The appropriate materials for the control, modifica- modification of the Casimir force in response to the car- tion, and fine tuning of the Casimir force are semicon- rier excitation was measured with a lock-in amplifier ductors. The reflectivity properties of semiconductor ͑see Fig. 17͒. The same function generator signal used to surfaces can be changed in a wide frequency range by generate the Ar laser pulses was also used as a reference changing the carrier density through the variation of for the lock-in amplifier. temperature, using different kinds of doping, or, alterna- The illumination of Si has to be done such that very tively, via illumination of the surface with laser light. At little if any light-impinges on the sphere, as this would the same time, semiconductor surfaces with reasonably lead to a light-induced force from the photon pressure. high conductivity avoid accumulation of excess charges As the Si membrane is illuminated from the bottom, and, thus, preserve the advantage of metals for Casimir care should be taken that the fraction of light transmit-

Rev. Mod. Phys., Vol. 81, No. 4, October–December 2009

1864 Klimchitskaya, Mohideen, and Mostepanenko: The Casimir force between real materials: …

©

§ ted through the membrane is negligibly small. The opti- ¡ cal absorption depth for Si at a wavelength of 514 nm is 1 1 ␮m. The measured thickness of the membrane used 0 was 4.0±0.3 ␮m. The calculations presented showed -1 that for the membrane thicknesses used, the force on the

sphere due to photon pressure varies from 2.7% to 8.7% -2

  of the difference of dispersion forces to be measured  -3 when separation increases from 100 to 200 nm. The calibration of the setup, the determination of the

150 200 250 300 350 400 450 500

¡ £ ¥ §

cantilever deflection coefficient, and the average separa-

 

   tion on contact between the test bodies are performed as  in earlier experiments with metal test bodies ͓see Sec. 1 IV.A and Harris et al. ͑2000͒; Chen, Mohideen, et al. 0 ͑2006͒; Chen et al. ͑2007b͒; Chiu et al. ͑2008͔͒. The de- flection coefficient was m=137.2±0.6 nm per unit deflec- -1 tion signal. The difference in the value of m from previ- -2

ous measurements ͑see Sec. IV.A͒ is due to the use of

  the 514 nm optical filter ͑see Fig. 17͒ which reduces the -3  cantilever deflection signal. This filter was used to pre-

vent the interference of the 514 nm excitation light with 150 200 250 300 350 400 450 500

     the cantilever deflection signal from the 640 nm laser ͑see Fig. 17͒. The separation distance on contact was FIG. 18. The difference of the Casimir forces in the presence a0 =97 nm. The uncertainty in the quantity a0 +mSdef was and in the absence of light vs separation for different absorbed ⌬ 1 nm. This leads to practically the same error a powers ͑a͒ 9.3 and ͑b͒ 4.7 mW. The measured differences =1 nm in the absolute separations ͑83͒ because the error ͗Fexpt͘ are shown as dots. See text for the description of theo- in piezocalibration is negligibly small. An independent retical solid, dotted, and dashed lines computed using the dif- measurement of the lifetime of the carriers excited in ferent approaches. the Si membrane by pulses from the Ar laser was per- formed using a noninvasive optical pump-probe tech- within the resolution error. It was stated that the inde- nique ͑Nagai and Kuwata-Gonokami, 2002; Sabbah and pendence of the residual potential difference of separa- Riffe, 2002͒. tion is a basic and necessary condition for every Casimir In the experiment under consideration it is not suffi- force measurement. The dependence of V of separation cient to determine the residual potential difference V 0 0 indicates the presence of electrostatic surface impurities, between the sphere and the membrane as described in space charge effects, and/or electrostatic inhomogene- Sec. IV.A. In the presence of a pulse train the values of ities on the sphere or plate surface ͑Chiu et al., 2008͒. the residual potential difference can be different: Vl and 0 The small change in the residual potential difference be- V0 during the bright and dark phases of a laser pulse tween the sphere and the membrane in the presence and train, respectively. Both values are generally different absence of excitation light is primarily due to the screen- from the residual potential difference determined in the ing of surface states by few of the optically excited elec- absence of pulse train. In the measurement procedure, l trons and holes. The value of this difference of around the voltages V and V are applied to the membrane dur- 78 mV is equal to the change in band bending at the ing the bright and dark phases, respectively. The differ- ͑ ͒ surface. It is consistent with the fact that almost flat ence in the total force electric and Casimir for the bands are obtained at the surface with the surface pas- states with and without carrier excitation was measured. ͑ ͒ sivation technique used for the preparation of the Si Using Eq. 82 , this difference can be written in the form membrane ͑Chen et al., 2007b͒. ͑ l ͒ ⌬F ͑a͒ = X͑a͓͒͑Vl − Vl ͒2 − ͑V − V ͒2͔ + ⌬F͑a͒. ͑92͒ Then other voltages V ,V were applied to the Si tot 0 0 ⌬ membrane and the difference in the total force Ftot was Here ⌬F͑a͒ is the difference of the Casimir forces for measured as a function of separation within the interval the states with and without light. By keeping V=const from 100 to 500 nm. From these measurement results, l ⌬ expt and changing V , the parabolic dependence of Ftot on the difference in the Casimir force ⌬F ͑a͒ was deter- Vl was measured. The value of Vl where this function mined from Eq. ͑92͒. This procedure was repeated with reaches a maximum ͓recall that X͑a͒Ͻ0 in Eq. ͑82͔͒ some number of pairs ͑J͒ of different applied voltages l l ͑ l ͒ ͗⌬ expt͑ ͒͘ is V0. Then by keeping V =const and changing V, V ,V and at each separation the mean value F a ⌬ the parabolic dependence of Ftot on V was mea- was found. The measurements were performed for dif- eff ͑ ͒ sured, which allows one to find V0. Both procedures ferent absorbed laser powers: P =9.3 mW J=31 , were repeated at different separations and the values 8.5 mW ͑J=41͒, and 4.7 mW ͑J=33͒, corresponding to l V0 =−0.303±0.002 V and V0 =−0.225±0.002 V were the incident powers of 15.0, 13.7, and 7.6 mW, respec- found to be independent of separation in the range from tively. The experimental data for Peff=9.3 and 4.7 mW 100 to 500 nm. These values were shown to change only are shown by dots in Figs. 18͑a͒ and 18͑b͒, respectively.

Rev. Mod. Phys., Vol. 81, No. 4, October–December 2009

Klimchitskaya, Mohideen, and Mostepanenko: The Casimir force between real materials: … 1865

©

¡ §

 ´   µ

 Ë 0 25 -0.25 20 -0.5 

-0.75  15 -1 -1.25 10 -1.5

-1.75 5

 ½ 160 170 180 190 200

11 12 13 14 15 16 17

¡ £ ¥ §

Ð Ó    ´ Ö  » × µ 

½ ¼ FIG. 19. Comparison of experiment and theory. The experi- FIG. 20. The dielectric permittivity of the Si membrane along mental differences in the Casimir force with their experimental the imaginary frequency axis for Si with different concentra- errors are shown as crosses ͑the absorbed power is 4.7 mW͒. tion of charge carriers ͑see text for further discussion͒. The solid and dotted lines represent the theoretical differences computed at T=300 K using the model with a finite static per- mittivity of high-resistivity Si but different models for Si in the that the use of the analytic approximation for ␧ ͑Inui, ͑ ͒ Si presence of light see text for further discussion . The dashed 2003͒ leads to about 10% error in the magnitudes of the line represents the theoretical differences computed at the Casimir force ͑Chen, Mohideen, et al., 2006͒ and, thus, is same temperature including the dc conductivity. not suitable for the comparison with precise measure- ments. The dielectric permittivity of the high-resistivity As expected, the magnitude of the Casimir force differ- Si with dc conductivity included ͑the charge carrier den- ence has the largest values at the shortest separations sity n˜ Ϸ5ϫ1014 cm−3͒ is shown in Fig. 20 by the short- and decreases with decrease of the separation. It also dashed line. decreases with decrease of the absorbed laser power. On irradiation of the Si membrane with light, an equi- The analysis of the experimental errors performed in librium value of the carrier density is rapidly established line with Sec. III.C.1 ͑Chen et al., 2007b͒ shows that the during a period of time much shorter than the duration total experimental error in this experiment is equal to of a laser pulse ͑note that the lifetimes were indepen- the random one. The relative experimental errors at a dently measured͒. This allows one to assume that there 95% confidence level vary from 10% to 20% at a sepa- is an equilibrium concentration of pairs ͑electrons and ration a=100 nm and from 25% to 33% at a separation holes͒ when the light is incident. In this case the dielec- a=180 nm for different absorbed laser powers. As an tric permittivity along the imaginary frequency axis is example, in Fig. 19 the experimental data are shown commonly represented in the form ͓see, e.g., Palik with their absolute errors calculated at a 95% confi- ͑1985͒, Vogel et al. ͑1992͒, and Inui ͑2004, 2006͔͒ dence level for the absorbed power 4.7 mW in the sepa- ͑l͒ 2 2 ␧ ͑i␰͒ = ␧ ͑i␰͒ + ␻ ͑ ͒/␰͓␰ + ␥͑ ͔͒ + ␻ ͑ ͒/␰͓␰ + ␥͑ ͔͒, ration region from 150 to 200 nm ͓each third dot from Si Si p e e p p p Fig. 18͑b͒ is shown͔. Thus, the modulation of the disper- ͑93͒ sion force with light is demonstrated by Chen et al. ␻ ␥ ͑ ͒ where p͑e,p͒ and ͑e,p͒ are the plasma frequencies and 2007a, 2007b at a high reliability and confidence. the relaxation parameters of electrons and holes, respec- For the comparison of experimental results with 13 12 tively ͓␥͑ ͒ Ϸ1.8ϫ10 rad/s, ␥͑ ͒ Ϸ5.0ϫ10 rad/s ͑Vo- theory, the difference of dispersion forces between a e p gel et al., 1992͒; the values of the plasma frequencies are sphere and a plate was calculated according to the first given below͔. To avoid violation of the thermal equilib- equality in Eq. ͑16͒ with the free energy given by the rium arising from the use of the Drude model ͑see Sec. Lifshitz formula ͑3͒͑Chen et al., 2007a, 2007b͒. The sur- II.D.1͒, it was suggested ͑Mostepanenko and Geyer, face roughness was taken into account using geometrical 2008͒ that the influence of free charge carriers in metal- averaging ͑75͒. It was shown that the contribution from type semiconductors shold be included by means of the the roughness correction to the calculated value of ͑ ͒ ⌬ ͑ ͒ generalized plasmalike model 58 . This rule should also F a is small. At the shortest separation a=100 nm it be applied to all doped semiconductors with dopant con- is only 1.2% and decreases to 0.5% at a=150 nm. The centration above critical. Then, instead of Eq. ͑93͒, the calculations were done at the laboratory temperature dielectric permittivity is given by T=300 K. ␧͑l͒͑ ␰͒ ␧ ͑ ␰͒ ␻2 ␰2 ␻2 ␰2 ͑ ͒ Both cases of the dielectric permittivity of the high- Si i = Si i + p͑e͒/ + p͑p͒/ . 94 resistivity ͑10 ⍀ cm͒ Si in the absence of light were con- The plasma frequency can be calculated from sidered, i.e., the inclusion or the neglect of the dc con- ductivity ͑see Sec. II.D.2͒. The dielectric permittivity of ␻ ͑ 2 * ⑀ ͒1/2 ͑ ͒ p͑e,p͒ = ne /me,p 0 , 95 the dielectric Si ͑no light and the dc conductivity is dis- ͒ * * regarded along the imaginary frequency axis was found where the effective masses are mp =0.2063m and me from the tabulated optical data ͑Palik, 1985͒. It is shown =0.2588m, and n is the concentration of charge carriers ␧ ͑ ͒ in Fig. 20 as the long-dashed line. Here 0 =11.66. Note Vogel et al., 1992 . The value of n for the different ab-

Rev. Mod. Phys., Vol. 81, No. 4, October–December 2009

1866 Klimchitskaya, Mohideen, and Mostepanenko: The Casimir force between real materials: …

  !  ©       ©

# § " sorbed powers can be calculated with the knowledge ¡ that at equilibrium the number of charge carrier pairs 1 created per unit time per unit volume is equal to the 2 recombination rate of pairs per unit volume. This leads 0.5 to ͑Chen et al., 2007b͒ 1 0 eff␶ ប␻ ␲ ͑ ͒ n =4Pw / d w, 96

-0.5

A ? eff eff < where Pw =0.393P is the power in the central part of the Gaussian beam focused on the membrane ͑see -1

above͒ and ␶ is the measured lifetime of the charge car- 120 140 160 180 200 220 240

¡ £ ¥ §

riers. For the absorbed power Peff=9.3 mW ͓Fig. 18͑a͔͒

, - . / 0 1 2 3 + , / 4 5 - 6

͑ ͒ ͑ ͒ ͑ ͒ + 7 8 * it follows from Eqs. 96 and 95 that na = 2.1±0.4 ' ϫ 19 −3 ␻ ͑ ͒ϫ 14 ␻ 10 cm , p͑e͒ = 5.1±0.5 10 rad/s, and p͑p͒ =͑5.7±0.6͒ϫ1014 rad/s ͑Chen et al., 2007b͒. Here and 1 below all errors are found at a 95% confidence level. For 0.5

eff ͓ ͑ ͔͒ ; P =4.7 mW Fig. 18 b the same parameters are nb ͑ ͒ϫ 19 −3 ␻ ͑ ͒ϫ 14 = 1.4±0.3 10 cm , p͑e͒ = 4.1±0.4 10 rad/s, and 0 ␻ ͑ ͒ϫ 14 ͓ p͑p͒ = 4.6±0.4 10 rad/s. Note that Chen et al.

-0.5

= ? ͑2007b͒ contains two misprints corrected here.͔ The cor- < responding dielectric permittivities are shown in Fig. 20 -1 with the solid lines labeled a and b. It should be re-

marked that the dielectric permittivities ͑93͒ and ͑94͒ 120 140 160 180 200 220 240

' ( ) * almost overlap at all nonzero Matsubara frequencies. & The dielectric permittivity of Au along the imaginary frequency axis is presented in Fig. 1. For the dielectric FIG. 21. Comparison of experiment and theory. Theoretical minus experimental differences of the Casimir force vs separa- permittivity of Si in the dark phase and the bright phase tion are shown as dots. In the absence of the laser pulse the when Eq. ͑93͒ is used, the reflection coefficient is theoretical results for dots labeled 1 are computed with a finite Si ͑ ͒ rTE 0,kЌ =0. Thus, the calculation results do not de- static dielectric permittivity of high resistivity Si and for dots pend on whether the solid or the dashed line in Fig. 1 for labeled 2 taking the dc conductivity of high resistivity Si into ␧ ͑ ␰͒ ␧͑l͒͑ ␰͒ Au i is used. If in the presence of light Si i is given account. When the laser pulse is on, the charge carriers are by Eq. ͑94͒, then the Au dielectric permittivity in the described by ͑a͒ the Drude and ͑b͒ the plasma model. Solid framework of the generalized plasmalike model should lines show the 95% confidence interval. be used ͑the solid line in Fig. 1͒. The results of the numerical computations of the dif- ference Casimir force ⌬Ftheor͑a͒ between rough surfaces semiconductors more precise experiments are required with the dielectric permittivity ͑94͒ in the presence of ͑see Sec. V.D.2͒. light and by neglecting dc conductivity in the absence of The calculation of the theoretical differences light are shown by the solid lines in Fig. 18 ͑for different ⌬Ftheor͑a͒ using ͑in the absence of light͒ the dielectric absorbed powers͒. The results of the analogous calcula- permittivity with the inclusion of dc conductivity leads tions with the dielectric permittivity ͑93͒ in the presence to the results presented by the dashed lines in Figs. 18 of light are shown in Fig. 18 with the dotted lines. As and 19. As shown, this theoretical prediction is excluded seen from Fig. 18, both the solid and dotted lines are in by the data. The same can also be observed in Fig. 21͑a͒ good agreement with the experimental data shown by where the dots labeled 2 present the differences be- dots. Thus, the experiment under consideration does not tween the theoretical results calculated with the small dc allow one to discriminate between the dielectric permit- conductivity of the Si membrane included in the absence tivities ͑93͒ and ͑94͒ used above for the description of of light and the experimental data. Almost all dots la- charge carriers in metal-type semiconductors. This is beled 2 are outside the 95% confidence interval. This also seen in Fig. 21 where the dots present the values means that a model which includes the dc conductivity of the differences between ⌬Ftheor͑a͒ and ͗⌬Fexpt͑a͒͘ for of the high-resistivity Si is excluded by the experimental the absorbed power of 4.7 mW and the solid lines show data at a 95% confidence level within the separation the 95% confidence interval calculated as described in region from 100 to 200 nm. Sec. III.C.2 ͑Chen et al., 2007b; Mostepanenko and The results of this experiment were applied ͑Klim- Geyer, 2008͒. Dots labeled 1 in Figs. 21͑a͒ and 21͑b͒ are chitskaya, Mohideen, and Mostepanenko, 2008͒ to test ␧ ͑ ␰͒ obtained with Si i in the dark phase but with the di- the modification of the transverse magnetic reflection electric permittivities ͑93͒ or ͑94͒ in the bright phase, coefficient, as given by Eq. ͑40͒͑Dalvit and Lamoreaux, respectively. It can be observed that the dots labeled 1 2008͒ and Eq. ͑53͒ with the standard contributions of all are inside the confidence intervals in both Figs. 21͑a͒ nonzero Matsubara frequencies ͑Pitaevskii, 2008͒. The and 21͑b͒. To distinguish between the two models using experimental data for the difference of the Casimir the dielectric permittivities ͑93͒ and ͑94͒ of metal-type forces in the presence and absence of laser light are

Rev. Mod. Phys., Vol. 81, No. 4, October–December 2009

Klimchitskaya, Mohideen, and Mostepanenko: The Casimir force between real materials: … 1867

©

§ ¡ in the dark phase ͑the solid lines͒. The theoretical results 0 ͑ ͒ ␰ ͑ ͒ computed using Eq. 40 at all l or Eq. 53 with the -0.5 standard contributions of all nonzero Matsubara fre- quencies are excluded by data at a 70% confidence level -1 ͑Klimchitskaya, Mohideen, and Mostepanenko, 2008͒. According to Svetovoy ͑2008͒, the experimental data -1.5 for the difference Casimir force are equally consistent -2 with the nonlocal approach using the reflection coeffi- cient ͑53͒ and the Lifshitz theory with dc conductivity -2.5 neglected in the dark phase. To prove this, the experi-

mental data of Fig. 22͑a͒ at a 70% confidence level were

 § -3 ¡ used, but the dashed line was replaced with the theoret- ical band whose width was determined at a 95% confi- 125 150 175 200 225 250 275 300

dence level using the respective uncertainty in charge

¡ £ ¥ §

⌬ ϫ 19 −3

© carrier density n=0.4 10 cm . Such a mismatched

¡ § 0 comparison of experiment with theory is irregular. It can be seen that the theoretical bands related to the solid -0.5 and dashed lines in Fig. 22͑a͒ do not overlap if one uses in computations the uncertainty in charge carriers ⌬n -1 =0.3ϫ1019 cm−3 determined at the same 70% confi- -1.5 dence level as the experimental errors ͑Mostepanenko et al., 2009͒. -2 The experimental results presented here support the -2.5 conclusion made in Sec. II.D.2. In accordance with this

conclusion, one should disregard the role of free charge

 § -3 ¡ carriers in all dielectric materials ͑in particular, in semi- conductors with dopant concentration below critical͒

125 150 175 200 225 250 275 300 when calculating dispersion forces in the framework of

¡ £ ¥ § the Lifshitz theory.

FIG. 22. Difference of the Casimir forces between a Au- coated sphere and Si plate in the presence and absence of laser C. Doped semiconductors with different charge carrier light on the plate vs separation for the absorbed powers of ͑a͒ densities 9.3 and ͑b͒ 4.7 mW. The experimental data are shown as 1. p-type silicon crosses. The solid and dashed lines are computed using the standard Lifshitz theory with the dc conductivity of Si in the The most important materials used in nanotechnology dark phase neglected and taking into account the effect of are semiconductors with conductivity properties ranging charge screening for the transverse magnetic mode, respec- from the metallic to the dielectric. As mentioned in Sec. tively. V. A , semiconductors with a relatively high conductivity have an advantage that they avoid accumulation of re- shown in Fig. 22 for the absorbed powers ͑a͒ of 9.3 mW sidual charges but, at the same time, possess a typical and ͑b͒ of 4.7 mW. In contrast to Fig. 19, the experimen- dielectric dependence of the dielectric permittivity on tal errors are presented at a 70% confidence level. The frequency within a wide frequency range ͑see, e.g., Fig. computational results obtained on the basis of the stan- 20͒. This makes it possible to examine the influence of dard Lifshitz theory with the dc conductivity neglected doping concentration on the Casimir force. Chen et al. in the dark phase are shown by the solid lines. The ͑2005͒, and Chen, Mohideen, et al. ͑2006͒ first measured dashed lines are computed using Eq. ͑40͒ for the modi- the Casimir force between a Au-coated sphere of the ␰ ␮ ͑ fied TM reflection coefficient at all x= l with concentra- diameter 2R=202.6±0.3 m a Au layer of 105 nm tion of charge carriers n=n˜ and the Debye-Hückel thickness was used͒ and 5ϫ10 mm2 single crystal sili- ͗ ͘ length in the dark phase and n=2na or n=2nb in the con Si 100 plate. The resistivity of this p-type B-doped presence of light. The computation of the dashed lines plate measured using the four-probe technique was ␳ was repeated using Eq. ͑53͒ for the TM reflection coef- =0.0035 ⍀ cm. ficient at zero frequency. At all nonzero Matsubara fre- The Casimir force was measured with an improved quencies, the standard terms of the Lifshitz formula version of the setup previously used ͑Harris et al., 2000͒ were used. In both cases practically coinciding computa- for the two Au test bodies ͑see Sec. IV.A͒. The main tional resuts were obtained, shown as the dashed lines in improvements in the experimental setup were the use of Figs. 22͑a͒ and 22͑b͒. As shown in Figs. 22͑a͒ and 22͑b͒, much higher vacuum and the reduction of the uncer- the experimental data are consistent with the theoretical tainty in the determination of absolute separation a ͓see results computed on the basis of the standard Lifshitz Eq. ͑83͔͒. As in Sec. V. B , a much higher vacuum ͑2 theory with the dc conductivity of dielectric Si neglected ϫ10−7 torr͒ is needed to maintain the chemical purity of

Rev. Mod. Phys., Vol. 81, No. 4, October–December 2009

1868 Klimchitskaya, Mohideen, and Mostepanenko: The Casimir force between real materials: …

 

¡   §

  ´ Ô Æ µ 0  -50 220 -100

-150 200

-200 180 -250

-300 160 -350

100 150 200 250 300 350 140

¡ £ ¥ §

76 78 80 82 84 86 88 90

´ Ò Ñ µ FIG. 23. The mean measured Casimir force vs separation be-  tween a Au sphere and B-doped Si plate. FIG. 24. Comparison of experiment and theory. The magni- tude of the experimental Casimir force with the errors shown the Si surface which otherwise oxidizes rapidly to SiO2. as crosses vs separation. The solid line shows the theoretical The high-vacuum system consists of oil-free mechanical dependence calculated for the sample used in the experiment pumps, turbo pumps, and ion pumps. To maintain the and the dashed line for a dielectric Si. lowest pressure during data acquisition, only the ion pump is used. This helps to reduce the influence of me- perimental error varies from 0.87% at the shortest chanical noise. The absolute error in the determination separation to 3% at a=100 nm to 64% at a=300 nm. ⌬ of absolute separation was reduced to a=0.8 nm in In Fig. 24, the experimental data with their error bars ⌬ ͑ ͒ comparison to a=1 nm see Sec. IV.A . This was are plotted as crosses for a more narrow separation achieved using a piezo capable of traveling a distance range from 75 to 90 nm. The solid line in Fig. 24 shows ␮ 6 m from initial separation to contact. More details on the theoretical results calculated using the Lifshitz ͑ ͒ the improvements are provided by Chiu et al. 2008 . theory with the dielectric permittivity ͑93͓͒for the In contrast to Au, the Si surface is very reactive. Be- ␻ ␻ ϫ 14 ͑ p-type plate used p͑e͒ =0, p͑p͒ =7 10 rad/s Chen, cause of this, a special passivation procedure is needed Mohideen, et al., 2006͔͒. The dashed line in Fig. 24 pre- ͑ to prepare it for force measurements Chen, Mohideen, sents the calculation results for dielectric Si ͑the permit- ͒ et al., 2006 . To characterize the topography of both tivity is shown by the long-dashed line in Fig. 20͒.As samples, the Au coating on the sphere and the surface of seen from Fig. 24, the solid line is consistent with the the Si plate were investigated using an AFM. Images experiment whereas the dashed line is inconsistent. resulting from the surface scan of the Au coating show Thus, for the Si plate with a rather high doping concen- that the roughness is represented mostly by stochasti- tration ͑nϷ3ϫ1019 cm−3͒ the presence of free charge cally distributed distortions of about 8–22 nm height carriers markedly influences the Casimir force. ␦͑1͒ with st =3.446 nm. The surface scan of the Si surface shows much smoother distortions with typical heights ␦͑1͒ 2. n-type silicon from 0.4 to 0.6 nm and st =0.111 nm. All calibrations and determination of the residual Further investigation of the difference in the Casi- electrostatic force and the separation on contact were mir force for samples with different charge-carrier den- done immediately before the Casimir force measure- sities was performed for n-type Si doped with P ͑Chen, ments in the same high-vacuum apparatus ͑see the Klimchitskaya, et al., 2006͒. The above same high description in Sec. IV.A͒. The force calibration con- vacuum based AFM was used to measure the Casimir stant was 1.440±0.007 nN per unit cantilever deflec- force between a Au-coated sphere of a diameter 2R tion signal, and the residual potential difference was =201.8±0.6 ␮m and two 4ϫ7mm2 size Si plates with V0 =−0.114±0.002 V. different charge carrier densities placed next to each In Fig. 23, the mean measured Casimir force is pre- other. The thickness of the Au coating on the sphere was sented ͑average of 65 measurements͒. The variance of measured to be 92±2 nm. this mean is found to be approximately the same over The two Si samples chosen for this experiment were the entire measurement range 62.33ഛaഛ349.97 nm and identically polished single crystals of 500 ␮m thickness. equal to 1.5 pN. Thus, the random experimental error at The resistivity of the samples was measured using the ␳ ⍀ a 95% confidence level is 3.0 pN. The systematic error in four-probe technique to be a =0.43 cm. Thus, the Ϸ ϫ 16 −3 this experiment, determined at a 95% confidence level concentration of the charge carriers na 1.2 10 cm ͑in line with Sec. III.C.1͒, is 1.17 pN. The total experi- was well below the critical concentration corresponding mental error, calculated using Eq. ͑80͒ at a 95% confi- to the dielectric-metal transition in Si doped with P ⌬t exptϷ ͑ Ϸ ϫ 18 −3͒ dence level, is F 3.33 pN. Thus, the relative ex- ncr 1.3 10 cm . One of these samples was used as

Rev. Mod. Phys., Vol. 81, No. 4, October–December 2009

Klimchitskaya, Mohideen, and Mostepanenko: The Casimir force between real materials: … 1869

© © 

  

¡   §

¢   ¨

0 5 0 -100

-5

-200 -10 ¡

-300 -15 -20 -400

80 100 120 140 160 180 200 100 150 200 250 300 350 400

¡ £ ¥ §

¢ ¤ ¦ ¨

FIG. 25. The mean measured Casimir force vs separation be- FIG. 26. Comparison of experiment and theory. The differ- tween a Au sphere and two N-doped Si plates of ͑a͒ higher and ences of the mean measured Casimir forces acting between a ͑b͒ lower resistivities. Au sphere and lower and higher resistivity Si samples vs sepa- ration are shown as dots. The respective theoretical difference is shown by the solid line. the first Si plate in the experiment. The other one was subject to thermal diffusion doping to prepare the sec- ␳ Ϸ ϫ −4 ⍀ onstrating the effect of different charge-carrier densities ond, lower-resistivity plate, with b 6.7 10 cm ͑ Ϸ ϫ 20 −3͒ in the two Si plates used. nb 3.2 10 cm . Both plates of higher and lower resistivity were subjected to a special passivation proce- The error analysis performed, as described in Sec. dure to prepare their surfaces for the force measure- III.C.1, shows that the total experimental errors deter- ments. mined at a 95% confidence level are equal to the ran- ⌬t expt All calibration procedures were done as in the previ- dom ones in each measurement. Thus, Fa =8,6,and ous measurements ͑see Sec. IV.A͒ separately for the 4pNata=61.19 and 70 nm and aജ80 nm, respectively. samples of higher and lower resistivities. For the sample The measurement for the sample of lower resistivity is ⌬t expt of higher resistivity, the value of the residual potential slightly more noisy. Here Fa =11, 7, and 5 pN at a ജ difference V0 =−0.341±0.002 V was obtained. The canti- =61.19 and 70 nm and a 80 nm. From Fig. 25 it shown lever deflection constant was 1.646±0.004 nN per unit that the deviation between the two sets of data is larger deflection signal and the separation on contact was than the total experimental error in the separation re- a0 =32.4±1.0 nm. For the sample of lower resistivity V0 gion from 61.12 to 120 nm. =−0.337±0.002 V, a0 =32.3±1.0 nm, and the cantilever The force-distance relationship measured for the two deflection constant 1.700±0.004 nN per unit deflection Si samples was compared with theory. For the descrip- signal was found. Note that the expression for the elec- tion of the sample with lower resistivity the dielectric ͑ ͒ ͓ ␻ ␻ tric force ͑82͒ used in the calibration does not take into permittivity 93 was used with p͑p͒ =0, p͑e͒ =2.0 account possible influence of screening effects ͑space- ϫ1015 rad/s͔. The sample of higher resistivity was de- charge layer at the surface of Si͒. According to Bingqian scribed by the dielectric permittivity of a dielectric Si et al. ͑1999͒, for high-resistivity n-type Si with the con- ͑the long-dashed line in Fig. 20͒. centration of charge carriers of order 1016 cm−3, the im- In Fig. 26 the differences of the measured Casimir pact of this layer on the electrostatic force is negligible forces for the plates of the lower and higher resitivity ␮ expt͑ ͒ expt͑ ͒ at separations from 300–400 nm to 2.5 m where the Fb a −Fa a versus separation are shown as dots. In calibration fit was performed. Obviously, in the case of the same figure, the difference in the respective theoreti- lower resistivity Si the influence of space-charge layer is cally computed Casimir forces is shown by the solid line. even more negligible. These conclusions were experi- As seen from Fig. 26, the experimental and theoretical mentally confirmed through the demonstration that both results as functions of a are in good agreement. the residual potential difference V0 and the separation The experiments, described here, demonstrate the on contact a0 are the same regardless of the separation possibility of modifying the Casimir force by changing distance where the calibration fit is performed ͑Chen, the doping concentration of semiconductor materials. Klimchitskaya, et al., 2006; Chiu et al., 2008͒. Note that recently a theoretical investigation of the The Casimir force as a function of the separation was influence of Si doping concentration on the Casimir measured for both samples immediately after the cali- force in the separation region from 1 nm to 5 mm was bration procedure was done for each one. The mean performed using the zero-temperature Lifshitz formula values of the measured Casimir force F¯ expt are presented ͑Pirozhenko and Lambrecht, 2008a͒. The obtained re- in Fig. 25. Dots labeled a show the results for the sample sults, however, are reliable only at separations below of higher resistivity ͑average of 40 measurements͒. Dots 1 ␮m because the use of the zero-temperature Lifshitz labeled b correspond to the sample of lower resistivity formula with the room-temperature parameters at sepa- ͑average of 39 measurements͒. As seen from Fig. 25, rations above 1 ␮m is physically problematic ͑Bordag dots labeled a and b are distinct from each other dem- et al., 2001͒.

Rev. Mod. Phys., Vol. 81, No. 4, October–December 2009 1870 Klimchitskaya, Mohideen, and Mostepanenko: The Casimir force between real materials: …

angular corrugated and plane plates made of ideal met-

 als. The trenches were fabricated in p-doped silicon ͑the density of charge carriers 2ϫ1018 cm−3 was determined À from the dc conductivity which was 0.028 ⍀ cm͒. Silicon oxide was used as the mask. Deep uv lithography fol-

£ lowed by reactive ion etching was used to transfer the pattern. Trenches of approximate depth 1 ␮m were cre- FIG. 27. The periodical uniaxial rectangular corrugations on ated. Three types of samples were made on the same one of the plates. ⌳ ␮ wafer: sample A with A =1 m period, sample B with ⌳ B =400 nm period, and one with a flat surface. The D. Silicon plate with rectangular trenches fractional areas of solid volume were pA =0.478±0.002 and p =0.510±0.001 for samples A and B using a scan- ͑ ͒ B Chan et al. 2008 reported a measurement of the Ca- ning electron microscope. The residual hydrocarbons simir force between a gold-coated sphere of radius R were removed by oxygen plasma etching and the oxide and a silicon surface that had been structured with nano- mask was etched with HF. Samples of size 0.7 ͑ ͒ scale rectangular corrugations trenches . Measurements ϫ0.7 mm2 were used for the force measurement. The were performed in the dynamic regime using a microme- corrugated silicon surface was prepared as done by ͑ ͒ chanical torsional oscillator see Sec. IV.B . This means Chen, Klimchitskaya, et al. ͑2006͒ and Chen, Mohideen, that the immediately measured quantity was the change et al. ͑2006͒ using hydrogen passivation. However, after of the resonant frequency of the oscillator which is pro- this the silicon chip was baked to 120 °C to remove the portional to the derivative of the Casimir force with re- residual water from the bottom of the trenches, which spect to separation distance. In accordance with the ͑ ͒ might impair the passivation layer leading to patch po- proximity force approximation given by Eq. 85 , this tentials. derivative is The micromachined oscillator used in the measure- ␮ ␮ FЈ͑a͒ =−2␲RP͑a͒, ͑97͒ ments consists of a 3.5- m-thick, 500- m-square silicon s plate. Unlike in previous experiments with microma- where P͑a͒ is the Casimir pressure between one rectan- chined oscillators ͑see Sec. IV.B͒, the spheres were at- gular corrugated plate and one plane plate ͑see Fig. 27͒. tached to the plate. Two sputter gold coated with thick- Note that the experiment was done at room tempera- ness of about 400 nm glass spheres of radius R=50 ␮m ture, but the results obtained were compared with theo- were attached on top of each other to the torsional os- retical computations at T=0. cillator at a distance of b=210 ␮m using conductive ep- For an ideal metal case such a configuration was con- oxy. Two spheres are used to provide a large distance sidered by Büscher and Emig ͑2004͒ using exact meth- between the corrugated surface and the top of the tor- ͑␻ ␲ ods presented in Sec. II.B. It was shown that in the lim- sional oscillator. The resonant frequency 0 =2 iting case aӶ⌳ ͑recall that separation is measured ϫ1783 Hz, quality factor Q=32 000͒ was excited by ap- between the zero-corrugation levels of the lower plate plying voltage on one of the bottom electrodes. The os- and the upper plate͒ the Casimir pressure for the frac- cillations were detected with additional voltages with an tional surface area of solid volume equal to 1/2 is given amplitude of 100 mV and a frequency of 2␲ϫ102 kHz by which were applied to measure the capacitance change between the top plate and the bottom electrodes. A ␲2បc 1 1 1 P͑a͒ =− ͩ + ͪ. ͑98͒ phase-locked loop was used to detect the change in the 240 2 ͑a − H/2͒4 ͑a + H/2͒4 resonant frequency as a function of the distance be- tween the sphere and the corrugated plate. This is in fact the minimum value of the Casimir pres- The measurements were done at a vacuum of sure between the rectangular corrugated and plane 10−6 torr using a dry roughing pump and a turbo pump. plates. Under the condition aӶ⌳ for ideal metals the The residual potential difference V and the initial sepa- same result is obtained by the application of the prox- 0 ration between the surfaces a were determined and the imity force approximation or pairwise summation meth- 0 calibration was done using electrostatic forces. No value ods ͑see Sec. III.B.1͒. ӷ⌳ of a0 or errors in its determination were provided. A In the opposite limit a the exact result is given by ϳ ͑Büscher and Emig, 2004͒ residual potential difference V0 −0.43 V was found be- tween a sphere and a flat Si plate and noted to vary with ␲2បc 1 3 mV within the separation distance of 100 nm to 2 ␮m. P͑a͒ =− , ͑99͒ 240 ͑a − H/2͒4 If the same was found for the corrugated Si surfaces this was not mentioned. Voltages between V0 +245 mV and which is up to a factor of 2 larger in magnitude than the V0 +300 mV were applied and the calibration constant prediction of the proximity force approximation and pai- was found to be 628±5 m N−1 s−1. Since no analytic ex- wise summation in Eq. ͑98͒. This is the maximum value pression for the electrostatic force is available for the of the Casimir pressure in the configuration of the rect- trench geometry a two-dimensional ͑2D͒ numerical so-

Rev. Mod. Phys., Vol. 81, No. 4, October–December 2009 Klimchitskaya, Mohideen, and Mostepanenko: The Casimir force between real materials: … 1871

lution of the Poisson equation was used to calculate the than that expected for ideal metals. This discrepancy electrostatic energy between a flat plate and the trench was reported as quite natural due to the interplay of surface. This energy was then converted to a force be- nonzero skin depth and geometry effects. tween a sphere and the trench surface using the PFA. Thus, the experiment by Chan et al. ͑2008͒ reports the The Casimir force gradients between the flat plate and measurement of deviation resulting from geometry for samples A and B were measured after the application of corrugated rectangular trenches of relatively small peri- compensating voltages to the plates. The main uncer- ods. The depth of the corrugated trenches allowed good tainty in these measurements is reported as that coming comparison to the results obtained using the PFA taking from thermomechanical noise with a value of about into account only the fractional area of the top surface. 0.64 pN ␮m−1 at a=800 nm. The Casimir force gradient Deviations between 10–20% from the PFA were re- Ј ported. At present no theoretical computations exist between the flat plate and the gold sphere Fflat was first measured. Good agreement was found with the calcula- which would allow a definite comparison between ex- tion using the Lifshitz theory. The tabulated data for periment and theory for spherical and corrugated sur- gold and silicon ͑Palik, 1985͒ were used along with the faces made of real metals at room temperature ͓at T modification corresponding to the carrier density of sili- =0 such computations were performed by Lambrecht con ͑Chen, Klimchitskaya, et al., 2006͒. The roughness and Marachevsky ͑2008͔͒. correction was taken into account using rms roughnesses of 4 nm on the sphere and 0.6 nm on the silicon surface measured using an AFM ͑see Sec. III.B.1͒. Next the E. Future prospects to measure the Casimir force with Ј Ј semiconductor surfaces force gradients FA,expt and FB,expt were measured on the corrugated surfaces using the same gold sphere. 1. The dielectric-metal transition For the configurations used by Chan et al. ͑2008͒ one can neglect the contribution from the remote bottom The possibility for the modulation of the Casimir parts of trenches. Then the proximity force approxima- force due to a change of charge carrier density is offered tion leads to the following force gradients for samples A by semiconductor materials that undergo the dielectric- and B: metal transition with the increase of temperature. From a fundamental point of view, the modulation of the Ca- Ј ͑ ͒ ␲ ͑r͒͑ ͒ simir force due to the phase transitions of different kinds FA,PFA a =−2 RpAP a − H/2 , offers one more precision test of the role of conductivity and optical properties in the Lifshitz theory of the Ca- Ј ͑ ͒ ␲ ͑r͒͑ ͒ ͑ ͒ FB,PFA a =−2 RpBP a − H/2 , 100 simir force. An experiment was proposed ͑Castillo-Garza et al., ͑r͒ where P ͑a͒ is the Casimir pressure between two non- 2007͒ to measure the change of the Casimir force acting corrugated plates covered with a stochastic roughness between a Au-coated sphere and a vanadium dioxide calculated using the Lifshitz formula as described in Sec. ͑ ͒ VO2 film deposited on a sapphire substrate which un- III.B.1. dergoes the dielectric-metal transition with the increase To compare the experimental data with theory, Chan of temperature. It has been known that VO crystals and ͑ ͒ 2 et al. 2008 considered the ratios thin films undergo an abrupt transition from semicon- ducting monoclinic phase at room temperature to a me- ␳ Ј Ј ␳ Ј Ј ͑ ͒ A = FA,expt/FA,PFA, B = FB,expt/FB,PFA 101 tallic tetragonal phase at 68 °C ͑Zylbersztejn and Mott, 1975; Soltani et al., 2004; Suh et al., 2004͒. The phase It was shown that for sample A there are deviations of transition causes the resistivity of the sample to decrease ␳ 4 −3 A from unity up to 10% over the measurement range by a factor of 10 from 10 to 10 ⍀ cm. In addition, the from a=650 to 750 nm exceeding the experimental er- optical transmission for a wide region of wavelengths ␳ rors. For sample B there are deviations of B from unity extending from 1 ␮m to greater than 10 ␮m decreases up to 20% over the same measurement range. This dif- by more than a factor of 10–100. ference in the comparison of the experimental data for The increase of temperature necessary for the phase samples A and B with the PFA results is natural, as transition can be induced by laser light ͑Soltani et al., ⌳ ⌳ a/ A =0.7 and a/ B =1.75 at a typical separation dis- 2004; Suh et al., 2004͒. Thus, a setup similar to the tance considered a=700 nm. Thus, for sample B the ap- one employed in the demonstration of optically modu- plicability condition of the PFA, a/⌳Ӷ1, considered in lated dispersion forces ͑see Sec. V. B ͒ can be used. In the Sec. III.B.1 is violated to a larger extent than for sample initial stage of the experimental work, the procedures of A. film fabrication and their heating were investigated The measurements were repeated three times for each ͑Castillo-Garza et al., 2007͒. The preliminary theoretical sample and consistent results have been observed. The results are also obtained ͑Castillo-Garza et al., 2007; data were also compared to values from the exact calcu- Pirozhenko and Lambrecht, 2008a͒ based on the Lifshitz ͑ lations for ideal metal boundaries, which were converted theory and optical data for VO2 films Verleur et al., to the sphere-trenched plate case using the PFA. How- 1968͒. Calculations of the difference Casimir force be- ever, the measured deviations from PFA, as applied to tween a Au sphere and VO2 film on sapphire substrate the rectangular corrugations, were found to be 50% less after and before the phase transition showed that the

Rev. Mod. Phys., Vol. 81, No. 4, October–December 2009 1872 Klimchitskaya, Mohideen, and Mostepanenko: The Casimir force between real materials: …

measuring conductivity͒. The measurement of the differ- ence Casimir force is planned as follows. The Si plate is positioned such that the boundary is below the vertical ∆F diameter of the sphere ͑see Fig. 28͒. The distance be- patterned tween the sphere and Si plate a is kept fixed and the Si Si plate plate oscillates in the horizontal direction using the pi- ezo such that the sphere crosses the boundary in the perpendicular direction during each oscillation ͓a similar approach was exploited ͑Decca, López, Chan, et al., 2005͒ for constraining new forces from the oscillations of the Au-coated sphere above two dissimilar metals, Au and Ge͔. The Casimir force on the sphere changes as the FIG. 28. Schematic diagram of the experimental setup for the sphere crosses the boundary. This change corresponds to measurement of the difference Casimir force between a Au- the differential force, coated sphere and patterned Si plate. ⌬ ͑ ͒ ͑ ͒ ͑ ͒ ͑ ͒ F a = Fn˜ a − Fn a , 102 proposed experiment has much promise for the under- standing of the role of free charge carriers in the Lifshitz equal to the difference of the Casimir forces due to the theory of dispersion forces. different charge carrier densities n˜ and n, respectively. Interesting results can also be obtained when investi- This causes a difference in the deflection of the cantile- gating the change of the Casimir free energy in the ver. In order to reduce random noise by averaging, the phase transition of a metal to the superconducting state. periodic horizontal movement of the plate will be of an The variation of the Casimir free energy during this angular frequency ⍀ϳ0.1 Hz. The amplitude of plate transition is small ͑Mostepanenko and Trunov, 1997͒. oscillations is limited by the piezocharacteristics but can Nevertheless, the magnitude of this variation can be be of order 100 ␮m, much larger than the typical width comparable to the condensation energy of a semicon- of the transition region equal to 200 nm. ducting film and causes a measurable increase in the The proposed experiment holds promise for the inves- value of the critical magnetic field ͑Bimonte, Calloni, tigation of the possible variation of the Casimir force in Esposito, Milano, and Rosa, 2005; Bimonte, Calloni, Es- the dielectric-metal transition in semiconductors with posito, and Rosa, 2005͒. the increase of doping concentration ͑Klimchitskaya and Geyer, 2008͒. It has the potential to distinguish between 2. Casimir forces between a sphere and a plate with patterned the two models of the dielectric permittivity of semicon- geometry ductors with the concentration of charge carriers above the critical ͓see Eqs. ͑93͒ and ͑94͒ in Sec. V. B ͔. The difference force measurements are very sensitive to relatively small variations of the Casimir force ͑see ͒ Sec. V. B . Recently an experimental scheme was pro- 3. Pulsating Casimir force posed ͑Castillo-Garza et al., 2007͒ which promises a record sensitivity to a force difference at the level of 1 At present, a consensus has been reached that appli- fN. The patterned Si plate with two sections of different cations of the Casimir force in the design, fabrication, doping concentrations ͑see Fig. 28͒ is mounted on a pi- and actuation of micromechanical and nanomechanical ezo below a Au-coated sphere attached to the cantilever devices are very promising. When the characteristic sizes of an AFM. The piezo oscillates in the horizontal direc- of a device shrink below 1 ␮m, the Casimir force be- tion causing the flexing of the cantilever in response to comes larger than typical electric forces. Considerable the Casimir force above different regions of the plate. opportunities for micromechanical design would be Thus, the sphere is subject to the difference Casimir opened by pulsating Casimir plates moving back and force which can be measured using the static and dy- forth entirely due to the effect of the zero-point energy, namic techniques. The patterned plate is composed of a without the action of mechanical springs. This can be single crystal of Si specifically fabricated to have adja- achieved only through the use of both attractive and cent sections of two different charge carrier densities. A repulsive Casimir forces. In connection with this, it special procedure was developed for the preparation of should be noted that while the repulsive Casimir froces the Si sample with the two sections having different con- for a single cube or a sphere are still debated, the Ca- ductivities ͑Castillo-Garza et al., 2007͒ where the p- and simir repulsion between two parallel plates is well un- n-type dopants can be used ͑B and P, respectively͒. derstood. Repulsion occurs when the plates with dielec- ␧ ␧ Sharp transition boundaries between the two sections of tric permittivities 1 and 2 along the imaginary the Si plate of width, below 200 nm can be achieved. frequency axis are immersed inside a medium with the ␧ ␧ Ͻ␧ Ͻ␧ ␧ Ͻ␧ Identically prepared but unpatterned samples can be dielectric permittivity 0 such that 1 0 2 or 2 0 Ͻ␧ ͑ ͒ used to measure the properties which are needed for the 1 Mahanty and Ninham, 1976 . At short separations theoretical computations ͑Hall probes for measuring the in the nonretarded van der Waals regime this effect has charge carrier concentration, a four-probe technique for been long discussed and measurements have been re-

Rev. Mod. Phys., Vol. 81, No. 4, October–December 2009 Klimchitskaya, Mohideen, and Mostepanenko: The Casimir force between real materials: … 1873

P (mPa) VI. EXPERIMENTS ON THE CASIMIR-POLDER FORCE 60 40 2 A. Demonstration of the thermal Casimir-Polder force 20 1. The force in thermal equilibrium 0 -20 1 As mentioned in Sec. II.C, the Casimir-Polder interac- -40 tion between an atom and a wall leads to a change of the -60 center-of-mass oscillation frequency of the Bose- 100 150 200 250 300 Einstein condensate in the direction perpendicular to a (nm) the wall ͑Antezza et al., 2004͒. This frequency shift can be measured precisely, leading to an indirect measure- FIG. 29. The Casimir pressure vs separation in a three-layer ment of the Casimir-Polder force ͑Harber et al., 2005͒. ␣ ͑ ͒ system -Al5O2-ethanol-Si with no light on the Si plate line 1 Using this technique, the first measurement of the ther- ͑ ͒ and with the illuminated Si plate line 2 . mal Casimir-Polder force at large separations between an atom and a plate was performed by Obrecht et al. ͑2007͒. In that experiment the dipole oscillations with ported ͓see, e.g., Visser ͑1981͒ and Meurk et al. ͑1997͔͒. the frequency ␻ were excited in a 87Rb Bose-Einstein At separations of about 30 nm the same effect was mea- 0 ͑ ͒ condensate separated by a distance of a few microme- sured by Munday et al. 2009 . ters from a fused-silica substrate ͑wall͒. The Casimir- ͑ ͒ Recently it was shown that the illumination of one Si Polder force ͑27͒ between a Rb atom and a substrate plate in the three-layer systems Au-ethanol-Si, Si- changes the magnitude of the oscillation frequency mak- ethanol-Si, and ␣-Al O -ethanol-Si with laser pulses can ␻ ͑ 2 3 ing it equal to some z the z direction is perpendicular change the Casimir attraction to Casimir repulsion and to the wall͒. The use of the Bose-Einstein condensate is vice versa ͑Klimchitskaya et al., 2007b͒. The illumination convenient because it provides a spatially compact col- can be performed as described in Sec. V. B . Calculations lection of a relatively large number of atoms ͓2.5ϫ105 showed that in the system Au-ethanol-Si the force is atoms in the experiment by Obrecht et al. ͑2007͔͒.Itis repulsive at separations aϾ160 nm. The illumination of well characterized by the Thomas-Fermi density profile ͑ ͒ ␮ ͑ the Si plate, as in Sec. V. B , changes this repulsion to 29 with the radius Rz =2.69 m corresponding to a ␻ ␲ attraction. In the system Si-ethanol-Si the force between magnetic trap frequency in the radial direction 0 =2 ϫ ͒ the Si plates is attractive, however, with one Si plate 229 Hz . This profile is used for spatial averaging in ͑ ͒ illuminated the attraction is replaced with repulsion at Eq. 28 . The detailed description of the experimental setup, separations aϾ175 nm. calibration procedures and measurements is presented In the systems mentioned above, the magnitudes of by Harber et al. ͑2003, 2005͒ and McGuirk et al. ͑2004͒. the repulsive forces are several times less than the mag- The directly measured quantity was the relative fre- nitude of the attractive forces at the same separations. quency shift However, it is possible to design a system where the light-induced Casimir repulsion is of the same order of ␥ ͉␻ ␻ ͉ ␻ ͉␻2 ␻2͉ ␻2 ͑ ͒ z = 0 − z / 0 = 0 − z /2 0, 103 magnitude as the attraction. A good example is given by ␣ ␻2 ␻2 ͓ ͑ ͔͒ the three-layer system -Al2O3-ethanol-Si where the Si where 0 − z is connected see Eq. 28 with the spatial plate is illuminated with laser pulses. The computational and time-averaged Casimir-Polder force ͑Antezza et al., results for the Casimir pressure versus separation are 2004͒. presented in Fig. 29. The Casimir attraction ͑solid line 1͒ In thermal equilibrium the temperature of the fused changes to repulsion at separations aϾ70 nm when the silica substrate TS was equal to the environment tem- ͑ ͒ Si plate is illuminated ͑solid line 2͒. perature TE: TS =TE =310 K. The frequency shift 103 ͑ Note that the observation of the pulsating Casimir was measured for a number of wall-atom center of mass ͒ ␮ force requires that the plates be completely immersed in of the condensate separations from 7 to 11 m. In Fig. a liquid far away from any air-liquid interfaces. This pre- 30 the experimental data obtained by Obrecht et al. ͑ ͒ vents the occurrence of capillary forces. Surface prepa- 2007 in thermal equilibrium at separations below 10 ␮m are shown as crosses. Harber et al. ͑2005͒ esti- ration of the plates is necessary to bring about an inti- mated random, systematic, and total errors in the mea- mate contact between the plates and the liquid. The only sured values of ␥ in such experiments at a 66% confi- liquid-based force is the drag force due to the movement z dence level. Obrecht et al. ͑2007͒ performed analysis for of the plates in response to the change of the force. For each experimental point separately. The main sources of pressure values of around 10 mPa and typical spring systematic errors discussed by Harber et al. ͑2005͒ are constants of 0.02 N/m, the corresponding drag pressure connected with possible presence of spatially inhomoge- from the plate movement would be six orders of magni- neous electric or magnetic surface contaminations or tude less in value. Thus, in the near future one may uniform magnetic and electric fields. Special investiga- expect experimental confirmation of the possibility of tions were performed to obtain upper bounds on all sys- Casimir repulsion. tematic errors. In Fig. 30 the absolute total errors in the

Rev. Mod. Phys., Vol. 81, No. 4, October–December 2009

1874 Klimchitskaya, Mohideen, and Mostepanenko: The Casimir force between real materials: …



­ ¢ ½ ¼ Þ The inclusion of the dc conductivity of fused silica in ͑ ͒ 4 the model of the dielectric response, as in Eq. 47 , dra- matically affects the calculations. This changes the value ͑ ͒ 3 of the reflection coefficient rTM 0,kЌ and, consequently, the Casimir-Polder force ͑27͒ which leads to a change in 2 ␥ ͑ ͒ ͑ ͒ the magnitude of z computed using Eqs. 28 and 103 . ␥ 1 The computational results for z are shown in Fig. 30 as the dashed line. As shown, the first two experimental 0 points are in disagreement with theory taking into ac-

7 8 9 10 11 count the conductivity of fused silica.

 ´ Ñ µ

FIG. 30. The fractional change in the trap frequency vs sepa- 2. The force out of thermal equilibrium ration in thermal equilibrium with TS =TE =310 K computed Recently a nonequilibrium situation was considered by neglecting ͑solid line͒ and including ͑dashed line͒ the con- where the substrate is at a temperature TS but the envi- ductivity of the dielectric substrate. The experimental data are ͑ ͒ shown as crosses. ronment remote wall is at another temperature TE ͑Antezza et al., 2005͒. In this case the Casimir-Polder force has two additional terms due to the thermal fluc- ␥ measurement of the separations and z are presented in tuations from the nearest wall at a temperature TS and true scales at each individual data point. ͑ ͒ from remote walls environment at a temperature TE. Comparison between the experimental data and the The force acting between an atom and a substrate out theory is done as in Fig. 14 ͑see Sec. IV.C͒ where the of thermal equilibrium depends on both temperatures frequency shift due to the Casimir pressure between two and is given by ͑Antezza et al., 2005͒ parallel plates was shown. The measured quantity ͑in ͑ ͒ ͑ ͒ ͑ ͒ ͑ ͒ ͑ ͒ this case the relative frequency shift͒ is plotted as crosses F a,TS,TE = F a,TE + Fn a,TS − Fn a,TE . 104 on the same graph as the theoretical lines computed us- The first term on the right-hand side of Eq. ͑104͒ is the ing different approaches. The theoretical values of the Casimir-Polder force ͑27͒. The nonequilibrium contribu- relative frequency shift as a function of separation are tion F ͑a,T͒ obtained in the approximation of static ͑ ͒ ͑ ͒ n calculated using Eqs. 103 and 28 with the Casimir- atomic polarizability is defined as ͑Antezza et al., 2005͒ ͑ ͒ Polder force 27 . The mass of a Rb atom is m=1.443 ϱ ϱ ϫ −25 10 kg. The dymanic polarizability of a Rb atom in ͑ ͒ ͵ ␻͵ ͑␻ ͒ −2␻xa/c ͑ ͒ Fn a,T =−K d dxf ,x e , Eq. 27 can be considered as frequency-independent, 0 0 ␣͑ ␰ ͒Ϸ␣͑ ͒ ϫ −23 −3 and the static value i l 0 =4.73 10 cm was used in the computations. This allows one to obtain ␻4x2 ͑␻ ͒ ͓͉ ͑␻ ͉͒ ␧͑␻͒ 2͔1/2 highly accurate results for the separations under consid- f ,x = ប␻ p ,x +Re −1−x e /kBT −1 eration ͑Babb et al., 2004͒. 1 Fused silica is a good insulator. However, like any in- ϫͫ sulator, it possesses a nonzero dc conductivity at non- ͉ͱp͑␻,x͒ + ix͉2 zero temperature. Electrical conductivity in fused silica is ionic in nature and is determined by the concentration ͑2x2 +1͓͒x2 +1+͉p͑␻,x͉͔͒ ͑ ͒ + ͬ, of the impurities alkali ions which are always present ͉ͱp͑␻,x͒ + i␧͑␻͒x͉2 as trace constituents. At TS =TE =310 K this conductivity −9 2 −1 ͑ varies within a wide region from 10 to 10 s Bansal 2ͱ2ប␣͑0͒ ͒ K ϵ , p͑␻,x͒ϵ␧͑␻͒ −1−x2. ͑105͒ and Doremus, 1986; Shackelford and Alexander, 2001 . ␲ 4 When neglecting the dc conductivity, the dielectric per- c ␧͑ ␰ ͒ ␰ mittivity of fused silica i l as a function of l can be The frequency shift of the condensate oscillations out of calculated ͑Caride et al., 2005͒ using the tabulated opti- thermal equilibrium can be calculated using Eqs. ͑103͒ ͑ ͒ ͑ ͒ ͑ ͒ ͑ ͒ ͑ ͒ cal data Palik, 1985 and the dispersion relation 57 .In and 28 where F a,T is replaced with F a,TS ,TE given this case the static dielectric permittivity has a finite by Eqs. ͑104͒ and ͑105͒. ␧ ͑ ͒ value of 0 =3.81. The respective computational results In the experiment by Obrecht et al. 2007 the fused- ␥ for z with the dc conductivity of fused silica neglected silica substrate was heated by the absorption with laser are shown in Fig. 30 by the solid line ͑Obrecht et al., light. The experimental data obtained out of thermal ͒ ␧͑ ␰ ͒ ␧ 2007 . Note that the results computed using i l = 0 equilibrium for TE =310 K and two different values of ͑ ͒ Obrecht et al., 2007 and frequency-dependent dielec- substrate temperature, TS =479 and 605 K, are shown as tric permittivity ͑Antezza et al., 2004; Klimchitskaya and crosses in Figs. 31͑a͒ and 31͑b͒, respectively. As in Fig. Mostepanenko, 2008b͒ are almost coincident at large 30, the absolute errors are presented in true scales at ͑ ␥ separation distances only small deviations are observed each data point. The computational results for z in a at aϽ8 ␮m͒. The theoretical computations are in good nonequilibrium situation obtained by neglecting the dc agreement with the data, as stated by Obrecht et al. conductivity of fused silica ͑Obrecht et al., 2007͒ are pre- ͑2007͒. sented in Fig. 31 as the solid lines. Note that the fre-

Rev. Mod. Phys., Vol. 81, No. 4, October–December 2009

Klimchitskaya, Mohideen, and Mostepanenko: The Casimir force between real materials: … 1875



¼ ­ ¢ ½ Þ B. Future prospects to measure the Casimir-Polder force in

4 quantum reflection

µ ´  3 The magnitude and the distance dependence of the Casimir-Polder force were confirmed by Sukenik et al. 2 ͑1993͒ when studying the deflection of ground-state Na atoms passing through a micron-sized parallel-plate cav- 1 ity. The intensity of an atomic beam transmitted through 0 the cavity was measured as a function of the plate sepa- ration. Comparison of the experimental data with the

7 8 9 10 11

Ñ µ ´ 

 theoretical position-dependent potential for an atom be-



­ ¢ ½ ¼ Þ tween parallel ideal mirrors ͑Barton, 1987a, 1987b͒ al-

4 lowed one to investigate the atom-wall interaction at

 µ ´ separation distances below 3 ␮m. This experiment gave 3 impetus to the investigation of the Casimir-Polder forces 2 in scattering experiments. Of special interest are situa- tions when the wave nature of atom becomes dominant 1 with respect to its classical behavior as a particle. Such a pure quantum effect is what is referred to as quantum 0 reflection, i.e., a process in which a particle moving

7 8 9 10 11 through a classically allowed region is reflected by a po-

´  Ñ µ  tential without reaching a classical turning point. The possibility of reflection of an ultracold atom under the FIG. 31. The fractional change in the trap frequency vs sepa- ͑ ͒ influence of an attractive atom-wall interaction was pre- ration out of thermal equilibrium a with TS =479 K and TE ͓ =310 K and ͑b͒ with T =605 K and T =310 K. Computations dicted long ago on quantum-mechanical grounds see, S E ͑ ͔͒ are done by neglecting ͑solid line͒ and including ͑dashed line͒ e.g., Lennard-Jones and Devonshire 1936 . However, the conductivity of the dielectric substrate. The experimental the experimental observation of this phenomenon has data are shown as crosses. become possible only recently due to the success in the production of ultracold atoms. First it was investigated using liquid surfaces, as the reflection of He and H at- quency dependence of ␧͑␻͒ in Eq. ͑105͒ does not affect oms on liquid He ͑Nayak et al., 1983; Berkhout et al., the contributions to the frequency shift from the non- 1989͒ and from the sticking coefficient of H atoms on equilibrium terms in the total atom-wall force ͑104͒. liquid He ͑Doyle et al., 1991; Yu et al., 1993͒. Later a Direct computations show that in the nonequilibrium specular reflection of very slow metastable Ne atoms on situation the disagreement between the experimental Si and BK7 glass surfaces was studied ͑Shimizu, 2001͒. data and the theory with the dc conductivity of the sub- The observed velocity dependence was explained by the strate material included widens further. The correspond- quantum reflection which is caused by the attractive ing results are presented in Fig. 31 as the dashed lines Casimir-Polder interaction. ͑Klimchitskaya and Mostepanenko, 2008b͒. As shown in Quantum reflection becomes efficient when the mo- Fig. 31͑a͒, the three experimental points for T =479 K tion of the particle can no longer be treated semiclassi- S ͑ ͒ exclude the dashed line and the other two only touch it. cally Friedrich and Trost, 2004 . The behavior of the The dashed line in Fig. 31͑b͒ demonstrates that all data particle is of a quantum character when ͒ ͑ ജ ץ ͒ ͑ ␭ץ -for T =605 K exclude the theoretical prediction calcu S B z / z 1, 106 lated with the inclusion of the dc conductivity of fused ␭ ͑ ͒ ␲ប ͱ ͓ ͑ ͔͒ silica. Thus, the confidence at which the theoretical ap- where B z =2 / 2m E−V z is the local de Broglie proach based on Eq. ͑47͒ is excluded by data increases wavelength for a particle of mass m and initial kinetic with the increase of substrate temperature. energy E moving in the potential V͑z͒. The same can be It is notable that the inclusion of the dc conductivity formulated as a condition that the variation of the local ␲ ␭ ͑ ͒ of fused silica in the model of the dielectric response wave vector k=2 / B z , perpendicular to the surface, ͑47͒ does not affect the contributions to the frequency within the distance of the atomic de Broglie wavelength, ͑ ͒ is larger than k itself ͑Shimizu, 2001͒, shift arising from the nonequilibrium terms Fn a,T in ͑ ͒͑ ͒ Eq. 104 Klimchitskaya and Mostepanenko, 2008b . 1 dk Thus, the conductivity influences the computational re- ⌽ = Ͼ 1. ͑107͒ k2 dz sults only through the equilibrium Casimir-Polder force ͑see Sec. VI.A.1͒. The important role of the Obrecht et The reflection amplitude depends critically on the en- al. ͑2007͒ experiment is that, not only was the model of ergy V͑z͒ of the atom-wall interaction. This has at- the dielectric response taking the dc conductivity into tracted considerable attention to the theoretical investi- account excluded, but the thermal effect, as predicted by gation of the reflection amplitude depending on atomic the Lifshitz theory with the dc conductivity omitted, was energy and the form of the interaction potential measured for the first time. ͑Friedrich et al., 2002; Jurisch and Friedrich, 2004; Voro-

Rev. Mod. Phys., Vol. 81, No. 4, October–December 2009 1876 Klimchitskaya, Mohideen, and Mostepanenko: The Casimir force between real materials: … nin and Froelich, 2005; Voronin et al., 2005; Madroñero different nanostructures ͑Arnecke et al., 2007; Fermani and Friedrich, 2007͒. The reflection probability tends to et al., 2007͒. In this case, the probability of quantum unity, as the incident velocity tends to zero. Thus, a high reflection is sensitive to the shape of the potential probability of quantum reflection calls for small incident ͑Friedrich et al., 2002; Arnecke et al., 2007͒. Thus, it velocities, i.e., for cold atoms. needs to be ascertained if the general theory of disper- Advances in cooling techniques in the past decade sion forces can describe the interaction between micro- have made it possible to perform experiments with cold particles and nanostructures. An additional question to atoms interacting with solid surfaces. The quantum re- be answered is to what extent the macroscopic concept flection of 3He atoms in the scattering from an ␣-quartz of the dielectric permittivity can be used as an adequate crystal was observed at energies far from E→0 characteristic of nanostructure material properties. ͑Druzhinina and DeKieviet, 2003͒. The observation of As a first step in this direction, the Lifshitz theory of the large reflection amplitudes for the dilute Bose- the van der Waals force has been extended for the case Einstein condensate of 23Na atoms on a Si surface ͑Pas- of a microparticle interacting with a plane surface of a quini et al., 2004, 2006͒ allows the possibility of using uniaxial crystal or with a cylindrical shell made of such quantum reflection as a trapping mechanism. The corre- crystal ͑Blagov et al., 2005͒. An approximate expression sponding trap model has been developed theoretically for the free energy of a microparticle-cylinder interac- ͑Jurisch and Friedrich, 2006; Madroñero and Friedrich, tion was obtained using the PFA which is of high preci- 2007͒. The quantum reflection of He* atoms on a flat Si sion at microparticle-cylinder separations smaller than surface and on microfabricated surface structures was the cylinder radius. The reflection coefficients from the also investigated ͑Oberst, Kouznetsov, et al., 2005; surface of a uniaxial crystal ͑with the optical axis z per- Oberst, Taashiro, et al., 2005͒. The latter experiment has pendicular to the surface͒ can be expressed in terms of ␧ ͑␻͒ ␧ ͑␻͒ stimulated the development of adequate theoretical two dissimilar dielectric permittivities x and z . models for the description of the scattering of atomic For example, the multilayer graphite cylinder was con- matter waves on ridged surfaces ͑Kouznetsov and sidered as a simple model of a multiwalled carbon nano- Oberst, 2005͒. tube. Using this model, the Casimir-Polder interaction For theoretical calculations of the reflection ampli- between hydrogen atoms ͑molecules͒ and multiwalled tude, a simple attractive atom-wall interaction potential carbon nanotubes was investigated ͑Blagov et al., 2005; is commonly used ͑Friedrich et al., 2002; Druzhinina and Klimchitskaya, Blagov, and Mostepanenko, 2006͒. Later, DeKieviet, 2003; Oberst, Taashiro, et al., 2005͒ which Lifshitz-type formulas have been obtained which de- is an interpolation between the nonretarded van der scribe the van der Waals and Casimir-Polder interaction Waals energy ͑24͒ and the Casimir-Polder energy ͑25͒. between a sheet and a material plate or a mi- Comparison of computational results with the measure- croparticle ͑Bordag et al., 2006͒ and a microparticle and ment data for the reflection amplitudes allows one to a single-walled carbon nanotube ͑Blagov et al., 2007͒. estimate the parameters of the potential ͑Druzhinina For this purpose graphene was considered as an infini- and DeKieviet, 2003; Oberst, Taashiro, et al., 2005͒. The tesimally thin positive charged flat sheet, carrying a con- increased precision of the measurements opens new op- tinuous fluid with some mass and negative charge densi- portunities for comparison with the more exact Eq. ͑21͒ ties. This plasma sheet is characterized by a typical wave for the free energy of atom-wall interactions. Such a number K determined by the parameters of the hexago- comparison could yield new information on the role of nal structure of graphite. The interaction of the electro- the atomic and material properties in dispersion forces. magnetic oscillations with such a sheet was considered As shown by Bezerra et al. ͑2008͒, for an atom of meta- and the normal modes and reflection coefficients were stable He* and a Au wall at zero temperature, the larg- found ͑Barton, 2004, 2005͒. est deviations between the phenomenological potential A single-walled carbon nanotube can be approxi- and the exact interaction energy of about 10% are mately modeled as a cylindrical sheet carrying a two- achieved at separations of 300–500 nm. However, at dimensional free-electron gas with the same reflection room temperature the free energy deviates from the coefficients as for a plane sheet ͑Blagov et al., 2007͒. phenomenological potential by 31% at a=5 ␮m. There Note that the above approximate models do not take is also related theoretical research on Casimir physics in into account nanotube chirality ͑Saito et al., 1998͒. This atom-wall interaction devoted to the influence of non- can be included using the optical data for the nanotube zero temperature on atomic polarizability ͑Buhmann index of refraction. Regarding single-walled carbon and Scheel, 2008͒, the use of the lateral Casimir-Polder nanotubes, it remains unclear if it is possible to describe force between an atom, and a corrugated surface to nanotubes with surfaces such as metals or semiconduc- study nontrivial geometrical effects ͑Dalvit et al., 2008͒ tors by varying only one parameter, the typical wave and to the interaction of atoms with conducting micro- number K, in the reflection coefficients. structures ͑Eberlein and Zietal, 2007͒. By comparing the calculational results for the multi- walled nanotubes with those for single-walled, it was C. Casimir-Polder interaction of atoms with carbon nanotubes demonstrated that the macroscopic desciption using the concept of dielectric permittivity is applicable for nano- Experiments with ultracold atoms have generated in- tubes with only two or three layers at separation dis- terest on the scattering and trapping of such atoms by tances larger than 2 nm ͑Blagov et al., 2007; Klim-

Rev. Mod. Phys., Vol. 81, No. 4, October–December 2009 Klimchitskaya, Mohideen, and Mostepanenko: The Casimir force between real materials: … 1877 P chitskaya, Blagov, and Mostepanenko, 2008͒. It is well to hotodiodes Laser A bear in mind that the description of multiwalled nano- tubes by the graphite dielectric permittivity and single- B walled nanotubes as a two-dimensional free electron gas x Cantilever are only models and are not a complete description of all the nanotube properties. Nevertheless, they are good Sphere z

approximations for the estimation of microparticle- Plate Corrugated nanotube interaction on the basis of the Lifshitz theory.

VII. LATERAL CASIMIR FORCE AND CASIMIR Smooth Plate TORQUES

A. Lateral Casimir force between corrugated surfaces xandzpiezo electric tubes The Casimir force considered in the preceding sec- FIG. 32. Schematic of the experimental setup for measuring tions is the normal one, i.e., it leads to an attraction the lateral Casimir force. The x piezo and z piezo are indepen- perpendicular to the surfaces of the interacting bodies. dent. However, as mentioned in Sec. I.B, there is a geometry dependence for the Casimir force on the shape of the surface, especially for corrugated surfaces. Thus, the the axes of the corrugations, the phase shift along the x nontrivial boundary dependence of the normal Casimir axis becomes the periodic function of y with a period ⌳ ⌳ ␽ ␸͑ ͒ force acting between a plate with periodic uniaxial sinu- y = cot . Within one period y depends on y lin- soidal corrugations and a large sphere was demonstrated early, taking on values from 0 to 2␲. The resulting lat- ͑ ͒ ⌳ by Roy and Mohideen 1999 . It follows also from Eq. eral force should be averaged over the period y.Inthe ͑67͒ that the pressure between two corrugated plates case of an infinite plate this leads to a zero value at any with parallel uniaxial sinusoidal corrugations of the ␽0. For real bodies of finite size the lateral Casimir same period harmonically depends on the phase shift force will be measurable only for small deviations of the ͑ ͒ between the corrugations Bordag et al., 1995a . This corrugation axes from parallelity such that ⌳ cot ␽ is should result in the lateral Casimir force in addition to much larger than the size of the smallest body. For ex- the normal one. ample, for ⌳Ϸ1 ␮m and a body size of 10 ␮m the lat- Using the path integral method, Golestanian and Kar- eral Casimir force is only if ␽Ӷ0.1 rad. ͑ ͒ dar 1997, 1998 predicted the existence of the lateral The existence of the lateral Casimir force opens new Casimir force in the configuration of two corrugated opportunities for the application of the Casimir force in plates. Later the geometry dependence of the Casimir micromachines ͑Ashourvan et al., 2007a, 2007b; Emig, energy for sinusoidally corrugated ideal metal plates was 2007͒. It is also important that this force can be well studied by the path integral quantization of the electro- controlled by specific shapes of corrugations ͑Blagov et ͑ ͒ magnetic field Emig et al., 2001, 2003 . The general ex- al., 2004͒. Based on the path integral theory ͑Emig et al., pressions for both the normal and the lateral Casimir 2003͒ it appears that the lateral Casimir force is a prom- force were obtained in the lowest order of perturbation ising venue for the demonstration of shape dependences theory with respect to the relative corrugation ampli- and diffractionlike effects. For these reasons an experi- tudes A1 and A2. Here the interacting surfaces were de- mental investigation of the lateral Casimir force is of ͑ ͒ ͑ ͒ ͑ ␲ ⌳͒ ͑ ͒ scribed by Eq. 63 with f1 x,y =sin 2 x/ , f2 x,y much interest. ͓ ␲͑ ͒ ⌳͔ =sin 2 x+x0 / . A harmonic dependence of the lat- ␸ ␲ ⌳ eral force on the phase shift =2 x0 / was obtained ͑Emig et al., 2001, 2003͒. Later, the exact solution was B. Demonstration of the lateral Casimir force found for the lateral Casimir force in the configuration of two parallel ideal metal plates with laterally shifted The lateral Casimir force between a Au plate and a uniaxial rectangular gratings ͑Büscher and Emig, 2005͒. sphere, both sinusoidally corrugated, was measured for The influence of the lateral force acting on a sphere surface separations between 200 and 300 nm using an above a corrugated plate can explain ͑Klimchitskaya et AFM ͑Chen et al., 2002a, 2002b͒. The experiment was al., 2001͒ the nontrivial character of the normal Casimir performed at a pressure below 50 mtorr and at room force in this configuration experimentally demonstrated temperature. A schematic diagram of the experiment is by Roy and Mohideen ͑1999͒. shown in Fig. 32. To implement this experiment a dif- All theoretical results on the lateral Casimir force fraction grating with uniaxial sinusoidal corrugations of mentioned above were obtained for uniaxial corruga- period ⌳=1.2 ␮m and an amplitude of 90 nm was used tions of equal period. If the periods of corrugations are as the template. In order to obtain perfect orientation different, the lateral force becomes equal to zero. The and phase between the corrugated surfaces, a special im- uniaxial character of the corrugations is also important printing procedure was developed ͑Chen et al., 2002a, for the observation of the lateral Casimir force. Under 2002b͒. The amplitude of the corrugations on the plate the assumption that there is a nonzero angle ␽ between and that imprinted on the sphere were measured using

Rev. Mod. Phys., Vol. 81, No. 4, October–December 2009 1878 Klimchitskaya, Mohideen, and Mostepanenko: The Casimir force between real materials: …

͑ Ϸ ͒ A1 /a 0.27 . For this reason, it is not appropriate to 6 compare experimental results with theoretical predic- tions obtained for ideal metal test bodies in the lowest- 4 order perturbation theory. However, separations are

N) ⌳ 2 several times smaller than the corrugation period . Be- -1 3 cause of this, diffractionlike effects do not contribute 0 ͑see Sec. III.B͒. In this case the theoretical expression for the lateral Casimir force including corrections to the -2 skin depth can be obtained in the form -4 ␲4Rបc Flat͑a,␸͒ = A A G͑a,⌳,A ,A ͒sin ␸. ͑108͒ 120a4⌳ 1 2 1 2 Lateral force (10 -6

-8 The explicit form of the function G is different based on whether the proximity force approximation or the pair- -500 0 500 1000 1500 2000 wise summation method is used. In the framework of Lateral displacement of corrugated plate (nm) the PFA G does not depend on ⌳ ͑Chen et al., 2002a, 2002b͒. The resulting theoretical values for the ampli- FIG. 33. Comparison of experiment and theory. The average tude of the lateral Casimir force at a=221 nm are 0.27 measured lateral Casimir force as a function of the lateral dis- and 0.31 pN when calculated up to the second and placement is shown as solid squares. The solid line is the best fourth perturbation orders in the small parameters A /a, fit sine curve to the data leading to a lateral force amplitude of i 0.32 pN. respectively. Thus, the fourth-order term is 15% of the second-order one and cannot be neglected. The pairwise summation method is applicable to ⌳ the AFM to be 59±7 and 8±1 nm, respectively. The cali- larger ratios of a/ and the results obtained depend on bration of the cantilever and the measurement of the ⌳. Using this method, the explicit expression for the residual potential voltage between the sphere and the function G in Eq. ͑108͒ is plate were done by electrostatic means in a manner simi- 2 2 ͑ ͒ ͑ ͒ a 5 ͑ ͒ a A + A lar to experiments on measuring the normal Casimir G = ␬ sp ͑a͒ͫS 1 ͩ2␲ ͪ + S 2 ͩ2␲ ͪ 1 2 E pw ⌳ pw ⌳ 2 force ͑see Sec. IV.A͒. 2 a The corrugated plate was mounted on two piezos that ͑2͒ a A1A2 allowed independent movement of the plate in the ver- − 5 cos ␦ S ͩ4␲ ͪ ͬ. ͑109͒ pw ⌳ a2 tical and horizontal directions with the help of the x piezo and z piezo, respectively. Movement in the x di- Here the coefficients depending on a/⌳ are given by rection is necessary to achieve a lateral phase shift be- x2 x3 x4 tween the corrugations on the sphere and the plate. In- ͑1͒͑ ͒ −xͩ ͪ ⌫͑ ͒ Spw x = e 1+x + − + 0,x , dependent movement in the z direction is necessary for 6 6 6 control of the surface separation. The lateral Casimir 4 5 ͑ ͒ 31 13 x x force was measured at four different separations starting S 2 ͑x͒ = e−xͩ1+x + x2 + x3 − + ͪ from 221 nm with steps of 12 nm. At each separation the pw 80 240 480 480 measurement was repeated 60 times and the averaged x6 force was determined. The averaged lateral force mea- − ⌫͑0,x͒͑110͒ sured at a separation 221 nm is shown as solid squares in 480 Fig. 33. The sinusoidal oscillations as a function of the and ⌫͑␣,x͒ is the incomplete Gamma function. The cor- phase difference between the two corrugations are ␬͑sp͒ rection factor E to the Casimir energy due to the non- clearly observed. The periodicity of the lateral force os- zero skin depth in the configuration of a sphere above a cillations is in agreement with the corrugation period of plate is defined by the plate. A sine curve fit to the observed data is shown ϱ 2 in Fig. 33 as the solid line and corresponds to an ampli- ␲ បc ͑ ͒ − ͵ E͑aЈ͒daЈ = ␬ sp ͑a͒, ͑111͒ tude of 0.32 pN. Detailed error analysis ͑Chen et al., 2 E a 1440a 2002b͒ led to the total absolute experimental error of 0.077 pN at a 95% confidence level. The resulting preci- where E͑a͒ is the Casimir energy per unit area of two sion of the amplitude measurements at the closest sepa- parallel plates presented in Eq. ͑12͒. Using Eqs. ͑108͒ ration is around 24%. and ͑109͒ at a=221 nm, one obtains the amplitude of the The measurement data were compared with theoreti- lateral Casimir force equal to 0.24 and 0.30 pN in the cal results. Note that the measurements were performed second and fourth perturbation orders in Ai /a, respec- at separations from 200 to 300 nm where the corrections tively. Here the fourth-order term is 25% of the second- due to the skin depth are rather large ͑see Sec. IV.A͒.In order result. It is notable that the amplitudes of the lat- addition, the ratio of the corrugation amplitude on the eral Casimir force computed using both approximate plate to the separation distance is not a small parameter methods up to the fourth perturbation order in Ai /a are

Rev. Mod. Phys., Vol. 81, No. 4, October–December 2009 Klimchitskaya, Mohideen, and Mostepanenko: The Casimir force between real materials: … 1879 almost equal and in good agreement with the measure- ␽Ӷ⌳/L, where L is the size of the plate along the cor- ment result. The force amplitudes measured at different rugation axis. The optimum values of the corrugation separations demonstrated the a−4 dependence in agree- period ⌳ and plate separations were found in order to ment with theoretical predictions. get larger values of the Casimir torque in comparison Recently a theoretical approach was developed ͑Rod- with those predicted for anisotropic test bodies of the rigues et al., 2006a; Rodrigues, Maia Neto, et al., 2007͒ same area ͑Rodrigues et al., 2006b͒. which allows one to calculate the lateral Casimir force Both experiments, if successfully performed, will pro- between sinusoidally corrugated plates made of metals vide important new information about dispersion forces described by the plasma model ͓Eq. ͑32͔͒ without using between real materials. From an application standpoint, the PFA or pairwise summation but only in the lowest the Casimir torque may serve as one more mechanism of second perturbation order under the conditions A1 ,A2 control in micromachines. Ӷ␭ ⌳ p ,a, . These conditions are not satisfied in the ex- periment discussed above ͑specifically in that experi- Ϸ ␭ Ϸ ͒ VIII. CONCLUSIONS AND OUTLOOK ment A1 /a 0.27, A1 / p 0.43 . As we have seen, the contributions of the fourth perturbation orders in Ai /a In the foregoing, the intersection of experiment and are important. Because of this, comparison of the ex- theory in the investigation of the Casimir force between perimental data with the theoretical predictions ͑Rod- real material bodies has been reviewed. In the last ten rigues et al., 2006a; Rodrigues, Maia Neto, et al., 2007͒ is years the field has undergone many rapid advances. not appropriate ͑Chen et al., 2007͒. Many new results concerning the Casimir effect of both a fundamental and applied character have been ob- tained and were presented above. They are related to C. The Casimir torque condensed matter physics, statistical physics, atomic physics, and nanotechnology. This list should be supple- Another interesting effect that predicted long ago but mented with physics, quantum field has not yet been observed is the Casimir torque. Such a theory, gravitation and cosmology, and mathematical torque arises due to the zero-point electromagnetic os- physics if one keeps in mind that many important results cillations in the configuration of anisotropic or asymmet- on the Casimir effect are not directly connected with ric bodies. In the case of plates made of a uniaxial crys- experiments using real materials and, consequently, are ␧ ͑␻͒ tal described by the dielectric permittivities x and not covered in this review. ␧ ͑␻͒ ␧ ͑␻͒ y = z , the torque arises if there is a nonzero angle Although experimental Casimir physics is still a young ␽ between the optical axes of the plates. In the nonre- field, an important period in its development has been tarded limit the Casimir torque was investigated by Par- completed and the related conclusions can be formu- segian and Weiss ͑1972͒. The retardation effects were lated. The most important of these conclusions is that taken into account by Barash ͑1973͒. Recently these re- the Casimir force between real materials has been ex- sults were obtained using another method ͑Philbin and perimentally measured by many using different tech- Leonhardt, 2008͒. The torque leads to the rotation of the niques, and the data obtained demonstrate the role of plates until their optical axes are aligned. In the nonre- real material properties and the geometrical shape of ͉␧ ␧ ͉ ␧ Ӷ tarded limit and for small anisotropies x − y / y 1 the the test bodies ͑see Secs. IV.A–IV.C, V. B –V. D , and torque between the test bodies of area S at a separation VII.B͒. Although there are still different opinions on the a is Mϳ−S sin͑2␽͒/a2 ͑Munday et al., 2005͒.Inthe measure of agreement between experiment and theory relativistic limit Mϳ−S sin͑2␽͒/a3 ͑Mostepanenko and that has been achieved, statistical procedures for data Trunov, 1997͒. processing and for comparison between the experimen- Recently an interesting experimental scheme was pro- tal and theoretical results in precise force-distance mea- posed to observe this torque when a barium titanate surements have been developed ͑see Sec. III.C͒ and suc- plate is immersed in ethanol and an anisotropic disk is cessfully used in several experiments ͑see Secs. IV.A, placed above ͑Munday et al., 2005͒. The dielectric per- IV.B, V. B , and V. C ͒. Another important conclusion is mittivities are chosen in such a way that the Casimir that the application of the general Lifshitz theory of dis- force between the anisotropic bodies is repulsive ͑see persion forces to real materials at nonzero temperature Sec. V.E.3͒. The disk would float parallel to the plate at results in puzzles. These puzzles are of both theoretical a distance where its weight is counterbalanced by the and experimental character. On the theoretical side, an Casimir repulsion, being free to rotate in response to a enormously large thermal effect was predicted, arising small torque. Detailed numerical calculations are made when physically real but seemingly negligible properties demonstrating the feasibility of this experiment ͑Mun- of real dielectrics are taken into account. A related ther- day et al., 2005͒. mal effect also arises for Drude metals, and both lead to Another possibility to observe the Casimir torque was a thermodynamic inconsistency ͑Sec. II.D͒. Experimen- theoretically investigated in the configuration of two cor- tally, the predicted large thermal effects were excluded rugated plates with a small angle between the corruga- by measurements of different groups performed with tion directions ͑Rodrigues et al., 2006b͒. As discussed in metal, dielectric, and semiconductor test bodies ͑see Sec. 7.A, the lateral Casimir force and related torque Secs. IV.B, VI.A, and V. B , respectively͒. If some specific differ from zero for an angle between corrugation axes properties of the real bodies ͑the dc conductivity of di-

Rev. Mod. Phys., Vol. 81, No. 4, October–December 2009 1880 Klimchitskaya, Mohideen, and Mostepanenko: The Casimir force between real materials: … electrics, which vanishes with vanishing temperature, or fabricated metamaterials in the visible frequency range relaxation properties of conduction electrons in metals͒ have a very narrow frequency region where ␮Ͼ1 ͑Chet- are disregarded, the Lifshitz theory is found to be in tiar et al., 2007͒. Thus many engineering challenges re- good agreement with all performed experiments and main for design of appropriate materials. Nevertheless, with thermodynamics. As argued in Sec. II.D.1, the di- they show great potential for exploring new Casimir ef- electric permittivities used to describe the drift and dif- fect phenomena. Future prospects for measurement of fusion currents in dielectrics and metals are not compat- the thermal Casimir force, the Casimir force between ible with the Lifshitz theory because the existence of semiconductor surfaces and the Casimir-Polder force in these currents violates the state of thermal equilibrium. quantum reflection have already been discussed in Secs. This suggests that one should not include the drift and IV.D, V. E , and VI.B, respectively. This means we look diffusion currents in the model of the dielectric response forward to many exciting applications of the Casimir when applying the Lifshitz theory to real materials. force in the near future. There is no current consensus with respect to this infer- ence. ACKNOWLEDGMENTS One can conclude that the current Lifshitz theory, if applied in accordance with its basic postulates, is We are grateful to many colleagues for numerious consistent with the principles of thermodynamics and all helpful discussions. We express special gratitude to J. F. available experimental data. Some claims may be raised Babb, V. B. Bezerra, G. Bimonte, E. V. Blagov, M. Bor- because this theory does not include the conductivity dag, F. Chen, R. S. Decca, E. Fischbach, B. Geyer, D. E. properties of real materials. However, the experimental Krause, D. López, and C. Romero. We are also grateful data testify that these properties have nothing to do with to M. Antezza, G. Carugno, G. Jourdan, and R. Wild for dispersion forces. We believe that a future theory of dis- important additional information about their expe- persion forces will not use reflection coefficients ex- riments. This work was supported by the National Sci- pressed in terms of dielectric permittivity but will deal ence Foundation ͑Grant No. PHY0653657͒, Department with more general characteristics of scattering processes of Energy ͑Grant No. DE-FG02-04ER46131͒, and allowing, in particular, the inclusion of spatial nonlocal- Deutsche Forschungsgemeinschaft ͑Grant No. 436 RUS ity. 113/789/0-4͒. Many other future applications of dispersion forces in real systems have been discussed but have not yet been investigated experimentally. Because of this, they are REFERENCES not reflected in our review. One might mention the van der Waals and Casimir interatomic interaction in a mag- Agranovich, V. M., and V. L. Ginzburg, 1984, Crystal Optics ͑ ͒ netodielectric medium ͑Spagnolo et al., 2007͒, the with Spatial Dispersion Springer, Berlin . Casimir-Polder forces of excited atoms ͑Buhmann and Antezza, M., L. P. Pitaevskii, and S. Stringari, 2004, Phys. Rev. A 70, 053619. Welsch, 2008͒, and the Casimir interaction between Antezza, M., L. P. Pitaevskii, and S. Stringari, 2005, Phys. Rev. plates with dielectric permittivity and magnetic perme- ͑ Lett. 95, 113202. ability Kenneth et al., 2002; Pirozhenko and Lambrecht, Antonini, P., G. Bressi, G. Carugno, G. Galeazzi, G. Messineo, ͒ ͑ 2008b . Artificially microstructured metamaterials Pen- and G. Ruoso, 2006, New J. Phys. 8, 239. ͒ dry et al., 1999 are also promising for use in studies of Arnecke, F., H. Friedrich, and J. Madroñero, 2007, Phys. Rev. the Casimir effect. For example, the theoretical possibil- A 75, 042903. ity that metamaterials might lead to repulsive Casimir Arnold, W., S. Hunklinger, and K. Dransfeld, 1979, Phys. Rev. forces in some distance ranges has been investigated B 19, 6049. ͑Henkel and Joulain, 2005; Leonhardt and Philbin, 2007; Ashcroft, N. W., and N. D. Mermin, 1976, Solid State Physics Rosa et al., 2008͒. It was pointed out that a strong modi- ͑Saunders College, Philadelphia͒. fication of the Casimir force is possible for surface sepa- Ashourvan, A., M. Miri, and R. Golestanian, 2007a, Phys. Rev. rations around the resonance wavelength of the mag- Lett. 98, 140801. netic response. Note that natural materials are Ashourvan, A., M. Miri, and R. Golestanian, 2007b, Phys. Rev. physically limited to magnetic permeabilities ␮ of order E 75, 040103. of unity in the visible spectrum. However, artificially de- Babb, J. F., G. L. Klimchitskaya, and V. M. Mostepanenko, signed micrometer and submicrometer split-ring resona- 2004, Phys. Rev. A 70, 042901. Balian, R., and B. Duplantier, 1977, Ann. Phys. ͑N.Y.͒ 104, 300. tors and fish-net arrays with ␮Ͼ1 have been reported Balian, R., and B. Duplantier, 1978, Ann. Phys. ͑N.Y.͒ 112, 165. with resonances in the infrared and near-optical fre- ͑ ͒ ͑ ͒ Bansal, N. P., and R. H. Doremus, 1986, Handbook of Glass quencies Shalaev, 2007 . Rosa et al. 2008 simulated the Properties ͑Academic, New York͒. Casimir force between a metamaterial slab and a half Barash, Yu. S., 1973, Izv. Vyssh. Uchebn. Zaved., Radiofiz. 16, space of Ag. A repulsive Casimir force was found for a 1227 ͓ Sov. Radiophys. 16, 945 ͑1973͔͒. short band of separations only if there is no background Barash, Yu. S., and V. L. Ginzburg, 1975, Usp. Fiz. Nauk 116,5 metallic response component. It was pointed out that ͓Sov. Phys. Usp. 18, 305 ͑1975͔͒. the large dissipation that is present in metamaterials Barton, G., 1987a, Proc. R. Soc. London, Ser. A 410, 141. ͑Chettiar et al., 2007͒ would substantially reduce any Ca- Barton, G., 1987b, Proc. R. Soc. London, Ser. A 410, 175. simir repulsion. It should also be noted that presently Barton, G., 2004, J. Phys. A 37, 1011.

Rev. Mod. Phys., Vol. 81, No. 4, October–December 2009 Klimchitskaya, Mohideen, and Mostepanenko: The Casimir force between real materials: … 1881

Barton, G., 2005, J. Phys. A 38, 2997. Bordag, M., G. L. Klimchitskaya, and V. M. Mostepanenko, Berger, J., 1993, Statistical Decision Theory and Bayesian 1995b, Phys. Lett. A 200, 95. Analysis ͑Springer, New York͒. Bordag, M., U. Mohideen, and V. M. Mostepanenko, 2001, Bergström, L., 1997, Adv. Colloid Interface Sci. 70, 125. Phys. Rep. 353,1. Berkhout, J. J., O. J. Luiten, I. D. Setija, T. W. Hijmans, T. Boström, M., and B. E. Sernelius, 2000, Phys. Rev. Lett. 84, Mizusaki, and J. T. M. Walraven, 1989, Phys. Rev. Lett. 63, 4757. 1689. Boström, M., and B. E. Sernelius, 2004, Physica A 339, 53. Bezerra, V. B., G. Bimonte, G. L. Klimchitskaya, V. M. Boyer, T. H., 1968, Phys. Rev. 174, 1764. Mostepanenko, and C. Romero, 2007, Eur. Phys. J. C 52, 701. Brandt, S., 1976, Statistical and Computational Methods in Bezerra, V. B., R. S. Decca, E. Fischbach, B. Geyer, G. L. Data Analysis ͑North-Holland, Amsterdam͒. Klimchitskaya, D. E. Krause, D. López, V. M. Mostepanenko, Bressi, G., G. Carugno, R. Onofrio, and G. Ruoso, 2002, Phys. and C. Romero, 2006, Phys. Rev. E 73, 028101. Rev. Lett. 88, 041804. Bezerra, V. B., G. L. Klimchitskaya, and V. M. Mostepanenko, Brevik, I., J. B. Aarseth, J. S. Høye, and K. A. Milton, 2005, 2000, Phys. Rev. A 62, 014102. Phys. Rev. E 71, 056101. Bezerra, V. B., G. L. Klimchitskaya, and V. M. Mostepanenko, Brevik, I., S. A. Ellingsen, J. S. Høye, and K. A. Milton, 2008, 2002a, Phys. Rev. A 65, 052113. J. Phys. A: Math. Theor. 41, 164017. Bezerra, V. B., G. L. Klimchitskaya, and V. M. Mostepanenko, Brown, L. S., and G. J. Maclay, 1969, Phys. Rev. 184, 1272. 2002b, Phys. Rev. A 66, 062112. Brown-Hayes, M., D. A. R. Dalvit, F. D. Mazzitelli, W. J. Kim, Bezerra, V. B., G. L. Klimchitskaya, V. M. Mostepanenko, and and R. Onofrio, 2005, Phys. Rev. A 72, 052102. C. Romero, 2004, Phys. Rev. A 69, 022119. Brownlee, K. A., 1965, Statistical Theory and Methodology in Bezerra, V. B., G. L. Klimchitskaya, V. M. Mostepanenko, and Science and Engineering ͑Wiley, New York͒. C. Romero, 2008, Phys. Rev. A 78, 042901. Bryksin, V. V., and P. M. Petrov, 2008, Fiz. Tverd. Tela ͑St. Bezerra, V. B., G. L. Klimchitskaya, and C. Romero, 1997, Petersburg͒ 50, 222 ͓Phys. Solid State 50, 229 ͑2008͔͒. Mod. Phys. Lett. A 12, 2613. Buhmann, S. Y., and S. Scheel, 2008, Phys. Rev. Lett. 100, Bezerra, V. B., G. L. Klimchitskaya, and C. Romero, 2000, 253201. Phys. Rev. A 61, 022115. Buhmann, S. Y., and D.-G. Welsch, 2007, Prog. Quantum Elec- Bezerra, V. B., G. L. Klimchitskaya, and C. Romero, 2001, tron. 31, 51. Phys. Rev. A 65, 012111. Buhmann, S. Y., and D.-G. Welsch, 2008, Phys. Rev. A 77, Bimonte, G., 2006a, Phys. Rev. Lett. 96, 160401. 012110. Bimonte, G., 2006b, Phys. Rev. E 73, 048101. Buhmann, S. Y., D.-G. Welsch, and T. Kampf, 2005, Phys. Rev. Bimonte, G., 2007, New J. Phys. 9, 281. A 72, 032112. Bimonte, G., 2008, Phys. Rev. A 78, 062101. Buks, E., and M. L. Roukes, 2001, Phys. Rev. B 63, 033402. Bimonte, G., E. Calloni, G. Esposito, L. Milano, and L. Rosa, Buks, E., and M. L. Roukes, 2002, Nature ͑London͒ 419, 119. 2005, Phys. Rev. Lett. 94, 180402. Bulgac, A., P. Magierski, and A. Wirzba, 2006, Phys. Rev. D 73, Bimonte, G., E. Calloni, G. Esposito, and L. Rosa, 2005, Nucl. 025007. Phys. B 726, 441. Büscher, R., and T. Emig, 2004, Phys. Rev. A 69, 062101. Bingqian, L., Z. Changchun, and L. Junhua, 1999, J. Micro- Büscher, R., and T. Emig, 2005, Phys. Rev. Lett. 94, 133901. mech. Microeng. 9, 319. Capasso, F., J. N. Munday, D. Iannuzzi, and H. B. Chan, 2007, Blagov, E. V., G. L. Klimchitskaya, U. Mohideen, and V. M. IEEE J. Quantum Electron. 13, 400. Mostepanenko, 2004, Phys. Rev. A 69, 044103. Caride, A. O., G. L. Klimchitskaya, V. M. Mostepanenko, and Blagov, E. V., G. L. Klimchitskaya, and V. M. Mostepanenko, S. I. Zanette, 2005, Phys. Rev. A 71, 042901. 2005, Phys. Rev. B 71, 235401. Carlin, B. P., and T. A. Louis, 2000, Bayes and Empirical Bayes Blagov, E. V., G. L. Klimchitskaya, and V. M. Mostepanenko, Methods of Data Analysis ͑CRC, Boca Raton͒. 2007, Phys. Rev. B 75, 235413. Casimir, H. B. G., 1948, Proc. K. Ned. Akad. Wet. Ser. B: Phys. Blocki, J., J. Randrup, W. J. Swiatecki, and C. F. Tsang, 1977, Sci. 51, 793. Ann. Phys. ͑N.Y.͒ 105, 427. Casimir, H. B. G., and D. Polder, 1948, Phys. Rev. 73, 360. Bohren, C. F., and D. R. Huffmann, 1998, Absorption and Scat- Castillo-Garza, R., C.-C. Chang, D. Jimenez, G. L. Klim- tering of Light by Small Particles ͑Wiley, New York͒. chitskaya, V. M. Mostepanenko, and U. Mohideen, 2007, Bordag, M., 2006, Phys. Rev. D 73, 125018. Phys. Rev. A 75, 062114. Bordag, M., B. Geyer, G. L. Klimchitskaya, and V. M. Chan, H. B., V. A. Aksyuk, R. N. Kleiman, D. J. Bishop, and F. Mostepanenko, 1998, Phys. Rev. D 58, 075003. Capasso, 2001a, Science 291, 1941. Bordag, M., B. Geyer, G. L. Klimchitskaya, and V. M. Chan, H. B., V. A. Aksyuk, R. N. Kleiman, D. J. Bishop, and F. Mostepanenko, 1999, Phys. Rev. D 60, 055004. Capasso, 2001b, Phys. Rev. Lett. 87, 211801. Bordag, M., B. Geyer, G. L. Klimchitskaya, and V. M. Chan, H. B., Y. Bao, J. Zou, R. A. Cirelli, F. Clemens, W. M. Mostepanenko, 2000a, Phys. Rev. D 62, 011701͑R͒. Mansfield, and C. S. Pai, 2008, Phys. Rev. Lett. 101, 030401. Bordag, M., B. Geyer, G. L. Klimchitskaya, and V. M. Chazalviel, J.-N., 1999, Coulomb Screening of Mobile Charges: Mostepanenko, 2000b, Phys. Rev. Lett. 85, 503. Applications to Material Science, Chemistry and Biology Bordag, M., B. Geyer, G. L. Klimchitskaya, and V. M. ͑Birkhauser, Boston͒. Mostepanenko, 2006, Phys. Rev. B 74, 205431. Chen, F., G. L. Klimchitskaya, U. Mohideen, and V. M. Bordag, M., G. L. Klimchitskaya, and V. M. Mostepanenko, Mostepanenko, 2003, Phys. Rev. Lett. 90, 160404. 1994, Mod. Phys. Lett. A 9, 2515. Chen, F., G. L. Klimchitskaya, U. Mohideen, and V. M. Bordag, M., G. L. Klimchitskaya, and V. M. Mostepanenko, Mostepanenko, 2004, Phys. Rev. A 69, 022117. 1995a, Int. J. Mod. Phys. A 10, 2661. Chen, F., G. L. Klimchitskaya, V. M. Mostepanenko, and U.

Rev. Mod. Phys., Vol. 81, No. 4, October–December 2009 1882 Klimchitskaya, Mohideen, and Mostepanenko: The Casimir force between real materials: …

Mohideen, 2006, Phys. Rev. Lett. 97, 170402. 193202. Chen, F., G. L. Klimchitskaya, V. M. Mostepanenko, and U. Duraffourg, L., and P. Andreucci, 2006, Phys. Lett. A 359, 406. Mohideen, 2007a, Opt. Express 15, 4823. Dzyaloshinskii, I. E., E. M. Lifshitz, and L. P. Pitaevskii, 1961, Chen, F., G. L. Klimchitskaya, V. M. Mostepanenko, and U. Usp. Fiz. Nauk 73, 381 ͓ Adv. Phys. 38, 165 ͑1961͔͒. Mohideen, 2007b, Phys. Rev. B 76, 035338. Eberlein, C., and R. Zietal, 2007, Phys. Rev. A 75, 032516. Chen, F., and U. Mohideen, 2001, Rev. Sci. Instrum. 72, 3100. Ederth, T., 2000, Phys. Rev. A 62, 062104. Chen, F., U. Mohideen, G. L. Klimchitskaya, and V. M. Emig, T., 2007, Phys. Rev. Lett. 98, 160801. Mostepanenko, 2002a, Phys. Rev. Lett. 88, 101801. Emig, T., 2008, J. Stat. Mech.: Theory Exp. ͑2008͒, P04007. Chen, F., U. Mohideen, G. L. Klimchitskaya, and V. M. Emig, T., and R. Büscher, 2004, Nucl. Phys. B 696, 468. Mostepanenko, 2002b, Phys. Rev. A 66, 032113. Emig, T., N. Graham, R. L. Jaffe, and M. Kardar, 2007, Phys. Chen, F., U. Mohideen, G. L. Klimchitskaya, and V. M. Rev. Lett. 99, 170403. Mostepanenko, 2005, Phys. Rev. A 72, 020101͑R͒. Emig, T., N. Graham, R. L. Jaffe, and M. Kardar, 2008, Phys. Chen, F., U. Mohideen, G. L. Klimchitskaya, and V. M. Rev. D 77, 025005. Mostepanenko, 2006, Phys. Rev. A 74, 022103. Emig, T., A. Hanke, R. Golestanian, and M. Kardar, 2001, Chen, F., U. Mohideen, G. L. Klimchitskaya, and V. M. Phys. Rev. Lett. 87, 260402. Mostepanenko, 2007, Phys. Rev. Lett. 98, 068901. Emig, T., A. Hanke, R. Golestanian, and M. Kardar, 2003, Chen, F., U. Mohideen, and P. W. Milonni, 2004, in Quantum Phys. Rev. A 67, 022114. Theory Under the Influence of External Conditions, Emig, T., R. L. Jaffe, M. Kardar, and A. Scardicchio, 2006, ͑ ͒ edited by K. A. Milton Rinton, Princeton, NJ . Phys. Rev. Lett. 96, 080403. Chettiar, U. K., A. V. Kildishev, H.-K. Yuan, W.-S. Cai, S.-M. Esquivel, R., and V. B. Svetovoy, 2004, Phys. Rev. A 69, Xiao, V. P. Drachev, and V. M. Shalaev, 2007, Opt. Lett. 32, 062102. 1671. Esquivel, R., C. Villarreal, and W. L. Mochán, 2003, Phys. Rev. Chiu, H.-C., C.-C. Chang, R. Castillo-Garza, F. Chen, and U. A 68, 052103; 71, 029904͑E͒͑2005͒. Mohideen, 2008, J. Phys. A: Math. Theor. 41, 164022. Feinberg, J., A. Mann, and M. Revzen, 2001, Ann. Phys. ͑N.Y.͒ Chumak, A. A., P. W. Milonni, and G. P. Berman, 2004, Phys. 288, 103. Rev. B 70, 085407. Fermani, R., S. Scheel, and P. L. Knight, 2007, Phys. Rev. A 75, Cochran, W. G., 1954, Biometrics 10, 101. 062905. Contreras-Reyes, A. M., and W. L. Mochán, 2005, Phys. Rev. Fischbach, E., D. E. Krause, V. M. Mostepanenko, and M. No- A 72, 034102. vello, 2001, Phys. Rev. D 64, 075010. Côté, R., B. Segev, and M. G. Raizen, 1998, Phys. Rev. A 58, Friedrich, H., G. Jacoby, and C. G. Meister, 2002, Phys. Rev. A 3999. 65, 032902. Dalvit, D. A. R., and S. K. Lamoreaux, 2008, Phys. Rev. Lett. Friedrich, H., and J. Trost, 2004, Phys. Rep. 397, 359. 101, 163203. Genet, C., A. Lambrecht, P. Maia Neto, and S. Reynaud, 2003, Dalvit, D. A. R., P. A. Maia Neto, A. Lambrecht, and S. Rey- Europhys. Lett. 62, 484. naud, 2008, Phys. Rev. Lett. 100, 040405. Genet, C., A. Lambrecht, and S. Reynaud, 2000, Phys. Rev. A Decca, R. S., E. Fischbach, G. L. Klimchitskaya, D. E. Krause, 62, 012110. D. López, U. Mohideen, and V. M. Mostepanenko, 2009, Genet, C., A. Lambrecht, and S. Reynaud, 2003, Phys. Rev. A Phys. Rev. A 79, 026101. 67, 043811. Decca, R. S., E. Fischbach, G. L. Klimchitskaya, D. E. Krause, D. López, and V. M. Mostepanenko, 2003, Phys. Rev. D 68, Gerlach, E., 1971, Phys. Rev. B 4, 393. 116003. Geyer, B., G. L. Klimchitskaya, U. Mohideen, and V. M. Decca, R. S., D. López, H. B. Chan, E. Fischbach, D. E. Mostepanenko, 2008, Phys. Rev. A 77, 036102. Krause, and C. R. Jamell, 2005, Phys. Rev. Lett. 94, 240401. Geyer, B., G. L. Klimchitskaya, and V. M. Mostepanenko, Decca, R. S., D. López, E. Fischbach, G. L. Klimchitskaya, D. 2001, Int. J. Mod. Phys. A 16, 3291. E. Krause, and V. M. Mostepanenko, 2004, J. Low Temp. Geyer, B., G. L. Klimchitskaya, and V. M. Mostepanenko, Phys. 135, 63. 2003, Phys. Rev. A 67, 062102. Decca, R. S., D. López, E. Fischbach, G. L. Klimchitskaya, D. Geyer, B., G. L. Klimchitskaya, and V. M. Mostepanenko, E. Krause, and V. M. Mostepanenko, 2005, Ann. Phys. ͑N.Y.͒ 2004, Phys. Rev. A 70, 016102. 318, 37. Geyer, B., G. L. Klimchitskaya, and V. M. Mostepanenko, Decca, R. S., D. López, E. Fischbach, G. L. Klimchitskaya, D. 2005a, Phys. Rev. A 72, 022111. E. Krause, and V. M. Mostepanenko, 2007a, Phys. Rev. D 75, Geyer, B., G. L. Klimchitskaya, and V. M. Mostepanenko, 077101. 2005b, Phys. Rev. D 72, 085009. Decca, R. S., D. López, E. Fischbach, G. L. Klimchitskaya, D. Geyer, B., G. L. Klimchitskaya, and V. M. Mostepanenko, E. Krause, and V. M. Mostepanenko, 2007b, Eur. Phys. J. C 2006, Int. J. Mod. Phys. A 21, 5007. 51, 963. Geyer, B., G. L. Klimchitskaya, and V. M. Mostepanenko, Decca, R. S., D. López, E. Fischbach, and D. E. Krause, 2003, 2007, J. Phys. A: Math. Theor. 40, 13485. Phys. Rev. Lett. 91, 050402. Geyer, B., G. L. Klimchitskaya, and V. M. Mostepanenko, Derjaguin, B., 1934, Kolloid-Z. 69, 155. 2008, Ann. Phys. ͑N.Y.͒ 323, 291. Derjaguin, B., N. V. Churaev, and Ya. I. Rabinovich, 1987, Gies, H., and K. Klingmüller, 2006a, Phys. Rev. Lett. 96, Adv. Colloid Interface Sci. 28, 197. 220401. Doyle, J. M., J. C. Sandberg, I. A. Yu, C. L. Cesar, D. Klepp- Gies, H., and K. Klingmüller, 2006b, Phys. Rev. D 74, 045002. ner, and T. J. Greytak, 1991, Phys. Rev. Lett. 67, 603. Gies, H., and K. Klingmüller, 2006c, Phys. Rev. Lett. 97, Druzhinina, V., and M. DeKieviet, 2003, Phys. Rev. Lett. 91, 220405.

Rev. Mod. Phys., Vol. 81, No. 4, October–December 2009 Klimchitskaya, Mohideen, and Mostepanenko: The Casimir force between real materials: … 1883

Ginzburg, V. L., 1985, Physics and Astrophysics ͑Pergamon, Klimchitskaya, G. L., 2009, J. Phys.: Conf. Ser. 161, 012002. Oxford͒. Klimchitskaya, G. L., E. V. Blagov, and V. M. Mostepanenko, Golestanian, R., and M. Kardar, 1997, Phys. Rev. Lett. 78, 2006, J. Phys. A 39, 6481. 3421. Klimchitskaya, G. L., E. V. Blagov, and V. M. Mostepanenko, Golestanian, R., and M. Kardar, 1998, Phys. Rev. A 58, 1713. 2008, J. Phys. A: Math. Theor. 41, 164012. Greenaway, D. L., G. Harbeke, F. Bassani, and E. Tosatti, Klimchitskaya, G. L., F. Chen, R. S. Decca, E. Fischbach, D. E. 1969, Phys. Rev. 178, 1340. Krause, D. López, U. Mohideen, and V. M. Mostepanenko, Harber, D. M., J. M. McGuirk, J. M. Obrecht, and E. A. Cor- 2006, J. Phys. A 39, 6485. nell, 2003, J. Low Temp. Phys. 133, 229. Klimchitskaya, G. L., R. S. Decca, E. Fischbach, D. E. Krause, Harber, D. M., J. M. Obrecht, J. M. McGuirk, and E. A. Cor- D. López, and V. M. Mostepanenko, 2005, Int. J. Mod. Phys. nell, 2005, Phys. Rev. A 72, 033610. A 20, 2205. Hargreaves, C. M., 1965, Proc. K. Ned. Akad. Wet., Ser. B: Klimchitskaya, G. L., and B. Geyer, 2008, J. Phys. A: Math. Phys. Sci. 68, 231. Theor. 41, 164032. Harris, B. W., F. Chen, and U. Mohideen, 2000, Phys. Rev. A Klimchitskaya, G. L., B. Geyer, and V. M. Mostepanenko, 62, 052109. 2006, J. Phys. A 39, 6495. Henkel, C., and K. Joulain, 2005, Europhys. Lett. 72, 929. Klimchitskaya, G. L., U. Mohideen, and V. M. Mostepanenko, Henkel, C., K. Joulain, J.-P. Mulet, and J.-J. Greffet, 2004, 2000, Phys. Rev. A 61, 062107. Phys. Rev. A 69, 023808. Klimchitskaya, G. L., U. Mohideen, and V. M. Mostepanenko, Hough, D. B., and L. R. White, 1980, Adv. Colloid Interface 2007a, J. Phys. A: Math. Theor. 40, F339. Sci. 14,3. Klimchitskaya, G. L., U. Mohideen, and V. M. Mostepanenko, Høye, J. S., I. Brevik, J. B. Aarseth, and K. A. Milton, 2003, 2007b, J. Phys. A: Math. Theor. 40, F841. Phys. Rev. E 67, 056116. Klimchitskaya, G. L., U. Mohideen, and V. M. Mostepanenko, Høye, J. S., I. Brevik, J. B. Aarseth, and K. A. Milton, 2006, J. 2008, J. Phys. A: Math. Theor. 41, 432001. Phys. A 39, 6031. Klimchitskaya, G. L., and V. M. Mostepanenko, 2001, Phys. Høye, J. S., I. Brevik, S. A. Ellingsen, and J. B. Aarseth, 2007, Rev. A 63, 062118. Phys. Rev. E 75, 051127. Klimchitskaya, G. L., and V. M. Mostepanenko, 2006, Con- Høye, J. S., I. Brevik, S. A. Ellingsen, and J. B. Aarseth, 2008, temp. Phys. 47, 131. Phys. Rev. E 77, 023102. Klimchitskaya, G. L., and V. M. Mostepanenko, 2007, Phys. Iannuzzi, D., I. Gelfand, M. Lisanti, and F. Capasso, 2004, in Rev. B 75, 036101. under the Influence of External Con- Klimchitskaya, G. L., and V. M. Mostepanenko, 2008a, Phys. ditions, edited by K. A. Milton ͑Rinton, Princeton, NJ͒. Rev. E 77, 023101. Iannuzzi, D., M. Lisanti, and F. Capasso, 2004, Proc. Natl. Klimchitskaya, G. L., and V. M. Mostepanenko, 2008b, J. Phys. Acad. Sci. U.S.A. 101, 4019. A: Math. Theor. 41, 312002͑F͒. Intravaia, F., and C. Henkel, 2008, J. Phys. A: Math. Theor. 41, Klimchitskaya, G. L., and Yu. V. Pavlov, 1996, Int. J. Mod. 164018. Phys. A 11, 3723. Intravaia, F., C. Henkel, and A. Lambrecht, 2007, Phys. Rev. A Klimchitskaya, G. L., A. Roy, U. Mohideen, and V. M. 76, 033820. Mostepanenko, 1999, Phys. Rev. A 60, 3487. Intravaia, F., and A. Lambrecht, 2005, Phys. Rev. Lett. 94, Klimchitskaya, G. L., S. I. Zanette, and A. O. Caride, 2000, 110404. Phys. Rev. A 63, 014101. Inui, N., 2003, J. Phys. Soc. Jpn. 72, 2198. Kondepugi, D., and I. Prigogine, 1998, Modern Thermodynam- Inui, N., 2004, J. Phys. Soc. Jpn. 73, 332. ics ͑Wiley, New York͒. Inui, N., 2006, J. Phys. Soc. Jpn. 75, 024004. Kouznetsov, D., and H. Oberst, 2005, Phys. Rev. A 72, 013617. Israelashvili, J., 1992, Intermolecular and Surface Forces ͑Aca- Krause, D. E., R. S. Decca, D. López, and E. Fischbach, 2007, demic, New York͒. Phys. Rev. Lett. 98, 050403. Jackson, J. D., 1999, Classical Electrodynamics ͑Wiley, New Krech, M., 1994, The Casimir Effect in Critical Systems ͑World York͒. Scientific, Singapore͒. Jaffe, R. L., 2005, Phys. Rev. D 72, 021301. Lambrecht, A., P. A. Maia Neto, and S. Reynaud, 2006, New J. Jourdan, G., A. Lambrecht, F. Comin, and J. Chevrier, 2009, Phys. 8, 243. EPL 85, 31001. Lambrecht, A., and V. N. Marachevsky, 2008, Phys. Rev. Lett. Jurisch, A., and H. Friedrich, 2004, Phys. Rev. A 70, 032711. 101, 160403. Jurisch, A., and H. Friedrich, 2006, Phys. Lett. A 349, 230. Lambrecht, A., V. V. Nesvizhevsky, R. Onofrio, and S. Rey- Kardar, M., and R. Golestanian, 1999, Rev. Mod. Phys. 71, naud, 2005, Class. Quantum Grav. 22, 5397. 1233. Lambrecht, A., I. Pirozhenko, L. Duraffourg, and P. An- Kats, E. I., 1977, Zh. Eksp. Teor. Fiz. 73, 212 ͓ Sov. Phys. JETP dreucci, 2007, EPL 77, 44006; 81, 19901͑E͒͑2008͒. 46, 109 ͑1977͔͒. Lambrecht, A., and S. Reynaud, 2000, Eur. Phys. J. D 8, 309. Kenneth, O., and I. Klich, 2006, Phys. Rev. Lett. 97, 160401. Lamoreaux, S. K., 1997, Phys. Rev. Lett. 78,5. Kenneth, O., and I. Klich, 2008, Phys. Rev. B 78, 014103. Lamoreaux, S. K., 1998, Phys. Rev. Lett. 81, 5475͑E͒. Kenneth, O., I. Klich, A. Mann, and M. Revzen, 2002, Phys. Lamoreaux, S. K., 1999, Phys. Rev. A 59, R3149. Rev. Lett. 89, 033001. Lamoreaux, S. K., 2005, Rep. Prog. Phys. 68, 201. Kim, W. J., M. Brown-Hayes, D. A. R. Dalvit, J. H. Brownell, Lamoreaux, S. K., and W. T. Buttler, 2005, Phys. Rev. E 71, and R. Onofrio, 2008, Phys. Rev. A 78, 020101͑R͒. 036109. Kittel, C., 1996, Introduction to Solid State Physics ͑Wiley, New Landau, L. D., and E. M. Lifshitz, 1980, Statistical Physics ͑Per- York͒. gamon, Oxford͒, Pt. I.

Rev. Mod. Phys., Vol. 81, No. 4, October–December 2009 1884 Klimchitskaya, Mohideen, and Mostepanenko: The Casimir force between real materials: …

Landau, L. D., E. M. Lifshitz, and L. P. Pitaevskii, 1984, Elec- Effect and its Applications ͑Clarendon, Oxford͒. trodynamics of Continuous Media ͑Pergamon, Oxford͒. Mott, N. F., 1990, Metal-Insulator Transitions, 2nd ed. ͑Taylor Lehmann, E. L., and J. P. Romano, 2005, Testing Statistical & Francis, London͒. Hypotheses ͑Springer, New York͒. Munday, J. N., and F. Capasso, 2007, Phys. Rev. A 75, Lennard-Jones, J. E., and A. F. Devonshire, 1936, Proc. R. Soc. 060102͑R͒. London 156,6. Munday, J. N., and F. Capasso, 2008, Phys. Rev. A 77, 036103. Leonhardt, U., and T. G. Philbin, 2007, New J. Phys. 9, 254. Munday, J. N., F. Capasso, and V. A. Parsegian, 2009, Nature Lifshitz, E. M., 1956, Zh. Eksp. Teor. Fiz. 29,94͓ Sov. Phys. ͑London͒ 457, 170. JETP 2,73͑1956͔͒. Munday, J. N., F. Capasso, V. A. Parsegian, and S. M. Lifshitz, E. M., and L. P. Pitaevskii, 1980, Statistical Physics Bezrukov, 2008, Phys. Rev. A 78, 032109. ͑Pergamon, Oxford͒, Pt. II. Munday, J. N., D. Iannuzzi, Yu. Barash, and F. Capasso, 2005, Lifshitz, E. M., and L. P. Pitaevskii, 1981, Physical Kinetics Phys. Rev. A 71, 042102. ͑Pergamon, Oxford͒. Nagai, M., and M. Kuwata-Gonokami, 2002, J. Phys. Soc. Jpn. Lifshitz, I. M., M. Ya. Azbel’, and M. I. Kaganov, 1973, Elec- 71, 2276. tron Theory of Metals ͑Consultants Bureau, New York͒. Nayak, V. U., D. E. Edwards, and N. Masuhara, 1983, Phys. Lisanti, M., D. Iannuzzi, and F. Capasso, 2005, Proc. Natl. Rev. Lett. 50, 990. Acad. Sci. U.S.A. 102, 11989. Ninham, B. W., V. A. Parsegian, and J. H. Weiss, 1970, J. Stat. London, F., 1930, Z. Phys. 63, 245. Phys. 2, 323. Long, J. C., H. W. Chan, and J. C. Price, 1999, Nucl. Phys. B Oberst, H., D. Kouznetsov, K. Shimizu, J.-I. Fujita, and F. 539, 23. Shimizu, 2005, Phys. Rev. Lett. 94, 013203. Lukosz, W., 1971, Physica ͑Amsterdam͒ 56, 109. Oberst, H., Y. Taashiro, K. Shimizu, and F. Shimizu, 2005, Madroñero, J., and H. Friedrich, 2007, Phys. Rev. A 75, Phys. Rev. A 71, 052901. 022902. Obrecht, J. M., R. J. Wild, M. Antezza, L. P. Pitaevskii, S. Mahanty, J., and B. W. Ninham, 1976, Dispersion Forces ͑Aca- Stringari, and E. A. Cornell, 2007, Phys. Rev. Lett. 98, demic, New York͒. 063201. Maia Neto, P. A., A. Lambrecht, and S. Reynaud, 2005, Phys. Palik, E. D., 1985, Ed., Handbook of Optical Constants of Sol- Rev. A 72, 012115. ids ͑Academic, New York͒. Maia Neto, P. A., A. Lambrecht, and S. Reynaud, 2008, Phys. Parsegian, V. A., 2005, van der Waals Forces: A Handbook for Rev. A 78, 012115. Biologists, Chemists, Engineers, and Physicists ͑Cambridge Maradudin, A. A., and P. Mazur, 1980, Phys. Rev. B 22, 1677. University Press, Cambridge͒. Mazur, P., and A. A. Maradudin, 1981, Phys. Rev. B 23, 695. Parsegian, V. A., and G. H. Weiss, 1972, Adv. Colloid Interface McGuirk, J. M., D. M. Harber, J. M. Obrecht, and E. A. Cor- Sci. 40, 35. nell, 2004, Phys. Rev. A 69, 062905. Pasquini, T. A., M. Saba, G.-B. Jo, Y. Shin, W. Ketterle, and D. Mehra, J., 1967, Physica ͑Amsterdam͒ 37, 145. E. Pritchard, 2006, Phys. Rev. Lett. 97, 093201. Meurk, A., P. F. Luckham, and L. Bergström, 1997, Langmuir Pasquini, T. A., Y. Shin, C. Sanner, M. Saba, A. Schirotzek, D. 13, 3896. E. Pritchard, and W. Ketterle, 2004, Phys. Rev. Lett. 93, Milonni, P. W., 1994, . An Introduction 223201. to ͑Academic, San Diego͒. Pendry, J. B., A. J. Holden, D. J. Robbins, and W. J. Stewart, Milonni, P. W., 2007, Phys. Scr. 76, C167. 1999, IEEE Trans. Theory Tech. 47, 2075 Milton, K. A., 2001, The Casimir Effect: Physical Manifestation Petrov, V., M. Petrov, V. Bryksin, J. Petter, and T. Tschudi, of Zero-Point Energy ͑World Scientific, Singapore͒. 2006, Opt. Lett. 31, 3167. Milton, K. A., 2004, J. Phys. A 37, R209. Philbin, T. G., and U. Leonhardt, 2008, Phys. Rev. A 78, Mohideen, U., and A. Roy, 1998, Phys. Rev. Lett. 81, 4549. 042107. Mostepanenko, V. M., J. F. Babb, A. O. Caride, G. L. Klim- Pirozhenko, I., and A. Lambrecht, 2008a, Phys. Rev. A 77, chitskaya, and S. I. Zanette, 2006, J. Phys. A 39, 6583. 013811. Mostepanenko, V. M., V. B. Bezerra, R. S. Decca, E. Fisch- Pirozhenko, I., and A. Lambrecht, 2008b, J. Phys. A: Math. bach, B. Geyer, G. L. Klimchitskaya, D. E. Krause, D. López, Theor. 41, 164015. and C. Romero, 2006, J. Phys. A 39, 6589. Pirozhenko, I., A. Lambrecht, and V. B. Svetovoy, 2006, New Mostepanenko, V. M., R. S. Decca, E. Fischbach, G. Geyer, G. J. Phys. 8, 238. L. Klimchitskaya, D. E. Krause, D. López, and U. Mohideen, Pitaevskii, L. P., 2008, Phys. Rev. Lett. 101, 163202. 2009, Int. J. Mod. Phys. A 24, 1721. Plunien, G., B. Müller, and W. Greiner, 1986, Phys. Rep. 134, Mostepanenko, V. M., R. S. Decca, E. Fischbach, G. L. Klim- 87. chitskaya, D. E. Krause, and D. López, 2008, J. Phys. A: Raabe, C., and D.-G. Welsch, 2006, Phys. Rev. A 73, 063822. Math. Theor. 41, 164054. Rabinovich, S. G., 2000, Measurement Errors and Uncertain- Mostepanenko, V. M., and B. Geyer, 2008, J. Phys. A: Math. ties. Theory and Practice ͑Springer, New York͒. Theor. 41, 164014. Rabinovich, Ya. I., and N. V. Churaev, 1989, Kolloidn. Zh. 51, Mostepanenko, V. M., and M. Novello, 2001, Phys. Rev. D 63, 83 ͓Colloid J. USSR 51,65͑1989͔͒. 115003. Rodrigues, R. B., P. A. Maia Neto, A. Lambrecht, and S. Rey- Mostepanenko, V. M., and N. N. Trunov, 1985, Yad. Fiz. 42, naud, 2006a, Phys. Rev. Lett. 96, 100402. 1297 ͓ Sov. J. Nucl. Phys. 42, 818 ͑1985͔͒. Rodrigues, R. B., P. A. Maia Neto, A. Lambrecht, and S. Rey- Mostepanenko, V. M., and N. N. Trunov, 1988, Usp. Fiz. Nauk naud, 2006b, Europhys. Lett. 76, 822. 156, 385 ͓ Sov. Phys. Usp. 31, 965 ͑1988͔͒. Rodrigues, R. B., P. A. Maia Neto, A. Lambrecht, and S. Rey- Mostepanenko, V. M., and N. N. Trunov, 1997, The Casimir naud, 2007, Phys. Rev. A 75, 062108.

Rev. Mod. Phys., Vol. 81, No. 4, October–December 2009 Klimchitskaya, Mohideen, and Mostepanenko: The Casimir force between real materials: … 1885

Rodriguez, A., M. Ibanescu, D. Iannuzzi, J. D. Joannopoulos, Suh, J. Y., R. Lopez, L. C. Feldman, and R. F. Haglund, 2004, and S. G. Johnson, 2007, Phys. Rev. A 76, 032106. J. Appl. Phys. 96, 1209. Rosa, F. S. S., D. A. R. Dalvit, and P. W. Milonni, 2008, Phys. Sukenik, C. I., M. G. Boshier, D. Cho, V. Sandoghbar, and E. Rev. Lett. 100, 183602. A. Hinds, 1993, Phys. Rev. Lett. 70, 560. Roy, A., C.-Y. Lin, and U. Mohideen, 1999, Phys. Rev. D 60, Svetovoy, V. B., 2008, Phys. Rev. Lett. 101, 163603. 111101͑R͒. Svetovoy, V. B., and R. Esquivel, 2005, Phys. Rev. E 72, Roy, A., and U. Mohideen, 1999, Phys. Rev. Lett. 82, 4380. 036113. Rumer, Yu. B., and M. S. Ryvkin, 1980, Thermodynamics, Sta- Svetovoy, V. B., and M. V. Lokhanin, 2003, Phys. Rev. A 67, tistical Physics, and Kinetics ͑Mir, Moscow͒. 022113. Sabbah, A. J., and D. M. Riffe, 2002, Phys. Rev. B 66, 165217. Svetovoy, V. B., P. J. van Zwol, G. Palasantzas, and J. Th. M. Saharian, A. A., 2006, Phys. Rev. D 73, 064019. De Hosson, 2008, Phys. Rev. B 77, 035439. Saito, R., G. Dresselhaus, and M. S. Dresselhaus, 1998, Physi- Tomaš, M. S., 2002, Phys. Rev. A 66, 052103. ͑ cal Properties of Carbon Nanotubes Imperial College Press, Tomaš, M. S., 2005, Phys. Lett. A 342, 381. ͒ London . Torgerson, J. R., and S. K. Lamoreaux, 2004, Phys. Rev. E 70, Scardicchio, A., and R. L. Jaffe, 2006, Nucl. Phys. B 743, 249. 047102. Schram, K., 1973, Phys. Lett. 43A, 282. van Blockland, P. H. G. M., and J. T. G. Overbeek, 1978, J. Schwinger, J., L. L. DeRaad, and K. A. Milton, 1978, Ann. Chem. Soc., Faraday Trans. 1 74, 2637. Phys. ͑N.Y.͒ 115,1. Van Kampen, N. G., B. R. A. Nijboer, and K. Schram, 1968, Sernelius, B. E., 2005, Phys. Rev. B 71, 235114. Phys. Lett. 26A, 307. Shackelford, J. F., and W. Alexander, 2001, Eds., Material Sci- van Zwol, P. J., G. Palasantzas, and J. Th. M. De Hosson, 2007, ence and Engineering Handbook ͑CRC, Boca Raton͒. Appl. Phys. Lett. 91, 144108. Shalaev, V. M., 2007, Nature Photon. 1, 41. van Zwol, P. J., G. Palasantzas, M. van de Schootbrugge, and J. Shih, A., and V. A. Parsegian, 1975, Phys. Rev. A 12, 835. Shimizu, F., 2001, Phys. Rev. Lett. 86, 987. Th. M. De Hosson, 2008, Appl. Phys. Lett. 92, 054101. Shklovskii, B. I., and A. L. Efros, 1984, Electronic Properties of Verleur, H. W., A. S. Barker, and C. N. Berglund, 1968, Phys. Doped Semiconductors, Solid State Series Vol. 45 ͑Springer, Rev. 172, 788. Berlin͒. Visser, J., 1981, Adv. Colloid Interface Sci. 15, 157. Smythe, W. R., 1950, Electrostatics and Electrodynamics Vogel, T., G. Dodel, E. Holzhauer, H. Salzmann, and A. Theu- ͑McGraw-Hill, New York͒. rer, 1992, Appl. Opt. 31, 329. Soltani, M., M. Chaker, E. Haddad, R. V. Kruzelecky, and D. Volokitin, A. I., and B. N. J. Persson, 2007, Rev. Mod. Phys. 79, Nikanpour, 2004, J. Vac. Sci. Technol. A 22, 859. 1291. Sotelo, J., J. Ederth, and G. Niklasson, 2003, Phys. Rev. B 67, Voronin, A. Yu., and P. Froelich, 2005, J. Phys. B 38, L301. 195106. Voronin, A. Yu., P. Froelich, and B. Zygelman, 2005, Phys. Spagnolo, S., D. A. R. Dalvit, and P. W. Milonni, 2007, Phys. Rev. A 72, 062903. Rev. A 75, 052117. Yu, I. A., J. M. Doyle, J. C. Sandberg, C. L. Cesar, D. Klepp- Sparnaay, M. J., 1958, Physica ͑Amsterdam͒ 24, 751. ner, and T. J. Greytak, 1993, Phys. Rev. Lett. 71, 1589. Sparnaay, M. J., 1989, in Physics in the Making, edited by A. Zhou, F., and L. Spruch, 1995, Phys. Rev. A 52, 297. Sarlemijn and M. J. Sparnaay ͑North-Holland, Amsterdam͒. Zurita-Sánchez, J. R., J.-J. Greffet, and L. Novotny, 2004, Speake, C. C., and C. Trenkel, 2003, Phys. Rev. Lett. 90, Phys. Rev. A 69, 022902. 160403. Zylbersztejn, A., and N. F. Mott, 1975, Phys. Rev. B 11, 4383.

Rev. Mod. Phys., Vol. 81, No. 4, October–December 2009