Skip to main content
Log in

Scarf for Lifshitz

  • Regular Article
  • Published:
The European Physical Journal Plus Aims and scope Submit manuscript

Abstract

Polarization of dispersive and dissipative dielectric media with smoothed-out inhomogeneities is studied with the goal to clarify the question of renormalizability of electromagnetic stress–energy tensor. The stress tensor is computed with the Lifshitz approach to van der Waals forces in the non-retarded limit, which accounts for dominant effects at the distances from the interface shorter than the absorption wavelength. After the substraction of the leading free space ultraviolet divergencies, there still remain two types of divergencies. First, contributions diverging in the sharp interface case become finite once it is smoothed out. Second, new subleading ultraviolet cut-off-dependent contributions appear due to the interface internal structure. The Hadamard expansion, based on the heat kernel method, is applied to systematically single out both finite and subleading contributions and to demonstrate incomplete renormalizability of the Lifshitz theory. The above approach also allows us to reveal the purely quantum mechanical nature of surface tension, which consists of finite cut-off-independent as well as cut-off-dependent contributions. The deduced theory of surface tension and its calculations for real dielectric media are favourably compared to the available experimental data. The problem of surface tension proves to be of a distinguished limit type because the sharp interface formulation loses the critical information about the internal structure of an interface. The general theory offered here is illustrated with an exactly solvable model representing a smooth transition between two different dielectric media, which relies upon a solution of the Schrödinger equation with the Scarf potential.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7

Similar content being viewed by others

Notes

  1. Because in the time domain the relation between the displacement vector \({\varvec{D}}\) and the electric field vector \({\varvec{E}}\) is nonlocal, \({\varvec{D}}(t,{\varvec{x}}) = {\varvec{E}}(t,{\varvec{x}}) + \int _{0}^{\infty }{f(\tau ) {\varvec{E}}(t-\tau ,{\varvec{x}}) \, \mathrm {d}\tau }\), while in the frequency-domain it becomes \(\widehat{{\varvec{D}}}(\omega ;{\varvec{x}}) = \varepsilon (\omega ) \widehat{{\varvec{E}}}(\omega ;{\varvec{x}}), \ \varepsilon (\omega ) = 1 + \int _{0}^{\infty }{f(\tau ) \, e^{\mathrm {i}\omega \tau } \, \mathrm {d}\tau }\).

  2. For signs of various Green’s functions here and later in the text, we refer reader to Fursaev and Vassilevich [55] for our choice of the metric signature. For other choices of the metric, one may consult [51].

  3. The sign convention [57, 58] used here is dictated by the need to regularize the Feynman path integral via an analytic continuation from real t to complex t with negative imaginary part. However, the opposite sign convention is often used [54], which also transforms the Lorentzian into the Euclidean metric.

  4. Consider deformation of a small volume with fixed number of molecules as an example.

  5. The power \(-12\) was chosen by Lennard-Jones for convenience: in reality, the actual potential may have a mixed polynomial-exponential form, cf. [67] and references therein. For example, in the case of helium atoms the repulsive interaction between them is caused by a depletion in electron density in the overlap region that descreens the nuclei from each other resulting in internuclear repulsion. In general, while the \(r^{-12}\) term is often attributed to the Pauli exclusion principle for fermions due to anti-symmetry of their wave functions resulting in an exchange interaction between identical particles (which also applies to bosons)—this interaction is different from Coulomb electrostatic repulsion and, in fact, stronger on shorter distances and ultimately responsible for the stability of matter [68, 69]—it could also originate from the Heisenberg principle of uncertainty [70].

  6. Physically, this absorption of electrostriction to the renormalized isotropic pressure follows from the chemical potential [14, 62], which must be constant for media in equilibrium. One may think of the electrostriction stress as analogous to the gravity in the ocean compensated by the mechanical stresses in water: if it is balanced in the r-direction, then due to isotropy it must be balanced in the \(\theta \)-direction as well. Notably, the \(\theta \)-dependence of the electrostriction stress leads to non-uniform compression of the matter which is stronger near the interface. However, due to isotropy, electrostriction does not contribute to ST (Sect. 4).

  7. Because from now on we will work mostly with the Fourier transforms of all quantities, we implicitly assume that they depend on the Matsubara frequencies \(\zeta _n\) and q, and, for simplicity, we omit these arguments in functions. Thus, in these notations we have \(\varepsilon (\mathrm {i}\zeta ;{\varvec{x}})\equiv \varepsilon ({\varvec{x}})\), \({\widehat{G}}(\zeta ;{\varvec{x}},{\varvec{x'}})\equiv {\widehat{G}}({\varvec{x}},{\varvec{x'}})\), etc. Since \(\varepsilon \) is already dimensionless, we use the same letter for the functions \(\varepsilon =\varepsilon ({\overline{z}})=\varepsilon (\mathrm {i}\zeta ;{w}{\overline{z}})\).

  8. This formula is usually deduced phenomenologically [73], but has a QM justification [74].

  9. Recall that the pressure is the force per unit area of the wall should it be bounding the medium, while normal stress is the force per unit area acting on the medium itself.

  10. It must be also mentioned that the electrostriction contributions, which would have to be added to (18) in the dielectric phase, cancel out in (69) due to their isotropic nature cf. (36a).

  11. In the case of the Scarf model, \(\gamma ^{ \text{(div) }}\) can be calculated explicitly:

    $$\begin{aligned} \gamma ^{ \text{(div) }}={23\over 6 (4\pi )^{3/2} \beta \lambda _{0}w}\sum _n \frac{(\ln \varepsilon _{\infty })^2}{(\ln \varepsilon _{\infty })^2+4\pi ^2}\frac{1+\sqrt{\varepsilon _{\infty }}}{\sqrt{\varepsilon _{\infty }}}. \end{aligned}$$
  12. First, this leads to the Keesom effect [82] contributing to vdW forces due to attraction of permanent dipoles with moments \(\mathrm {p}\)—in the case of water the ratio of the Keesom to London energies of interaction \(\varphi _{K}/\varphi _{L} = 4 \mathrm {p}^{4}/(9 h_{{ \text{ B }}} T \alpha _{w}^{2} \omega _{a}) \approx 6.32\), where \(\alpha _{w}\) is the water molecular polarizability; most importantly, the hydrogen bond energy \(3.82 \cdot 10^{-20} \, \mathrm {J}\) per bond is comparable to that of the latent heat of evaporation per molecule \(6.78 \cdot 10^{-20} \, \mathrm {J}\) and much larger than \(k_{{ \text{ B }}} T \approx 4.14 \cdot 10^{-21} \, \mathrm {J}\) thus indeed being capable of affecting the value of ST. Second, the water molecule polarity may give rise to a local structure in the bulk and potentially at the interface thus, in particular, invalidating assumptions behind the isotropy of the dielectric permittivity upon which the Lifshitz theory is based—if one thinks of water molecules as permanent dipoles, they may orient themselves along the interface thus increasing ST as opposed to, say, surfactants, which, when added to water, stick their polar heads in water and thus orient themselves perpendicularly to the interface, thus decreasing ST due to repulsion between themselves.

  13. This is typical for non-closed theories, e.g. the Navier–Stokes equations require the empirically measured viscosity coefficient.

References

  1. I.Z. Fisher, Statistical Theory of Liquids (University of Chicago Press, Chicago, 1964)

    Google Scholar 

  2. R. Zallen, The Physics of Amorphous Solids (Wiley, New York, 2004)

    Google Scholar 

  3. J.P. Hansen, I.R. McDonald, Theory of Simple Liquids (Academic Press, Boca Raton, 2006)

    MATH  Google Scholar 

  4. J. Israelachvili, Intermolecular and Surface Forces (Academic Press, Boca Raton, 2011)

    Google Scholar 

  5. B.V. Derjaguin, M.M. Kusakov, The properties of thin layers of liquids. Proc. Acad. Sci. USSR Chem. Ser. 5, 741 (1936)

    Google Scholar 

  6. H.B.G. Casimir, D. Polder, The influence of retardation on the London–van der Waals forces. Phys. Rev. 73, 360 (1948)

    Article  ADS  MATH  Google Scholar 

  7. H.B.G. Casimir, On the attraction between two perfectly conducting planes. Proc. K. Ned. Akad. Wet. 51, 793 (1948)

    MATH  Google Scholar 

  8. B.V. Derjaguin, A.S. Titijevskaia, I.I. Abricossova, A.D. Malkina, Investigations of the forces of interaction of surfaces in different media and their application to the problem of colloid stability. Disc. Faraday Soc. 18, 24 (1954)

    Article  Google Scholar 

  9. B.V. Deryaguin, I.I. Abrikosova, E.M. Lifshitz, Direct measurement of molecular attraction between solids separated by a narrow gap. Q. Rev. 10, 295 (1956)

    Article  Google Scholar 

  10. D. Tabor, R.H.S. Winterton, Direct measurement of normal and retarded van der Waals forces. Nature (London) 219, 1120 (1968)

    Article  ADS  Google Scholar 

  11. B. Derjaguin, Untersuchungen über die Reibung und Adhäsion. IV. Kolloid Z. 69, 155 (1934)

    Article  Google Scholar 

  12. H.C. Hamaker, The London—van der Waals attraction between spherical particles. Physica 4, 1058 (1937)

    Article  ADS  Google Scholar 

  13. E.M. Lifshitz, The theory of molecular attractive forces between solids. Sov. Phys. JETP 2, 73 (1956)

    Google Scholar 

  14. I.E. Dzyaloshinskii, E.M. Lifshitz, L.P. Pitaevskii, General theory of Van der Waals’ forces. Sov. Phys. Uspekhi 4, 153 (1961)

    Article  ADS  MATH  Google Scholar 

  15. M.J. Sparnaay, Measurement of attractive forces between flat plates. Physica 24, 751 (1958)

    Article  ADS  Google Scholar 

  16. S.K. Lamoreaux, Demonstration of the Casimir force in the 0.6 to 6 \(\mu {{\rm m}}\) range. Phys. Rev. Lett. 78, 5 (1997)

    Article  ADS  Google Scholar 

  17. R. Decca, V.A.D. López, Casimir force in micro and nano electro mechanical systems, in Casimir Physics, ed. by D. Dalvit, P. Milonni, D. Roberts, F. da Rosa (Springer, New York, 2011), pp. 287–309

    Chapter  Google Scholar 

  18. K.A. Milton, S.A. Fulling, P. Parashar, P. Kalauni, T. Murphy, Stress tensor for a scalar field in a spatially varying background potential: Divergences, “renormalization,” anomalies, and Casimir forces. Phys. Rev. D 93, 085017 (2016)

  19. L.S. Brown, G.J. Maclay, Vacuum stress between conducting plates: an image solution. Phys. Rev. 184, 1272 (1969)

    Article  ADS  Google Scholar 

  20. D. Deutsch, P. Candelas, Boundary effects in quantum field theory. Phys. Rev. D 20, 3063 (1979)

    Article  ADS  MathSciNet  Google Scholar 

  21. T.G. Philbin, C. Xiong, U. Leonhardt, Casimir stress in an inhomogeneous medium. Ann. Phys. (N.Y.) 325, 579 (2010)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  22. C. Xiong, T.W. Kelsey, S.A. Linton, U. Leonhardt, Casimir forces for inhomogeneous planar media. J. Phys: Conf. Ser. 410, 012165 (2013)

    MATH  Google Scholar 

  23. W.M.R. Simpson, S.A.R. Horsley, U. Leonhardt, Divergence of Casimir stress in inhomogeneous media. Phys. Rev. A 87, 043806 (2013)

    Article  ADS  Google Scholar 

  24. K.A. Milton, Local and global Casimir energies: divergences, renormalization, and the coupling to gravity, in Casimir Physics, ed. by D. Dalvit, P. Milonni, D. Roberts, F. da Rosa (Springer, New York, 2011), pp. 39–95

    Chapter  MATH  Google Scholar 

  25. M. Bordag, D. Hennigt, D. Robaschik, Vacuum energy in quantum field theory with external potentials concentrated on planes, 1. Phys. A: Math. Gen. 25, 4483 (1992)

    Article  ADS  MathSciNet  Google Scholar 

  26. P. Candelas, Vacuum energy in the presence of dielectric and conducting surfaces. Ann. Phys. 143, 241 (1982)

    Article  ADS  MathSciNet  Google Scholar 

  27. A.A. Actor, Scalar quantum fields confined by rectangular boundaries. Fortschr. Phys. 43, 141 (1995)

    Article  MathSciNet  MATH  Google Scholar 

  28. S.A. Fulling, T.E. Settlemyre, K.A. Milton, Renormalization for a scalar field in an external scalar potential. Symmetry 10, 54 (2018)

    Article  MATH  Google Scholar 

  29. P. Parashar, K.A. Milton, Y. Li, H. Day, X. Guo, S.A. Fulling et al., Quantum electromagnetic stress tensor in an inhomogeneous medium. Phys. Rev. D 97, 125009 (2018)

    Article  ADS  Google Scholar 

  30. D.V. Vassilevich, Heat kernel expansion: user’s manual. Phys. Rep. 388, 279 (2003)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  31. M. Bordag, K. Kirsten, D.V. Vassilevich, Path-integral quantization of electrodynamics in dielectric media. J. Phys. A: Math. Gen. 31, 2381 (1998)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  32. S.K. Blau, M. Visser, A. Wipf, Zeta functions and the Casimir energy. Nucl. Phys. B 310, 163 (1988)

    Article  ADS  MathSciNet  Google Scholar 

  33. G. ’t Hooft, M.J.G. Veltman, Regularization and renormalization of gauge fields. Nucl. Phys. B 44, 189 (1972)

    Article  ADS  MathSciNet  Google Scholar 

  34. F. Bao, J.S. Evans, M. Fang, S. He, Inhomogeneity-related cutoff dependence of the Casimir energy and stress. Phys. Rev. A 93, 013824 (2016)

    Article  ADS  Google Scholar 

  35. Y. Li, K.A. Milton, X. Guo, H.L. Dodge, G. Kennedy, S.A. Fulling, Casimir forces in inhomogeneous media: Renormalization and the principle of virtual work. Phys. Rev. D 99, 125004 (2019)

    Article  ADS  MathSciNet  Google Scholar 

  36. S.A. Fulling, K.A. Milton, M. Bordag, Vacuum energy density and pressure near boundaries. Int. J. Mod. Phys. A 25, 2364 (2010)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  37. S.A. Fulling, K.A. Milton, J. Wagner, Energy density and pressure in power-wall models. Int. J. Mod. Phys. A 27, 1260009 (2012)

    Article  ADS  MATH  Google Scholar 

  38. K. Symanzik, Schrödinger representation and Casimir effect in renormalizable quantum field theory. Nucl. Phys. B 190, 1 (1981)

    Article  ADS  MathSciNet  Google Scholar 

  39. D.V. Vassilevich, Divergences in the vacuum energy for frequency-dependent interactions. Phys. Rev. D 79, 065016 (2009)

    Article  ADS  MathSciNet  Google Scholar 

  40. K.A. Milton, Hard and soft walls. Phys. Rev. D 84, 065028 (2011)

    Article  ADS  Google Scholar 

  41. J.D. Bouas, S.A. Fulling, F.D. Mera, K. Thapa, C.S. Trendafilova, J. Wagner, Investigating the spectral geometry of a soft wall. Proc. Symp. Pure Math. 84, 139 (2012)

    Article  MathSciNet  MATH  Google Scholar 

  42. S.W. Murray, C.M. Whisler, S.A. Fulling, J. Wagner, H.B. Carter, D. Lujan et al., Vacuum energy density and pressure near a soft wall. Phys. Rev. D 93, 105010 (2016)

    Article  ADS  Google Scholar 

  43. N. Graham, R.L. Jaffe, H. Weigel, Casimir effects in renormalizable quantum field theories. Int. J. Mod. Phys. A 17, 846 (2002)

    Article  ADS  MATH  Google Scholar 

  44. N. Graham, R.L. Jaffe, V. Khemani, M. Quandt, M. Scandurra, H. Weigel, Calculating vacuum energies in renormalizable quantum field theories: a new approach to the Casimir problem. Nucl. Phys. B 645, 49 (2002)

    Article  ADS  MATH  Google Scholar 

  45. N. Graham, R.L. Jaffe, V. Khemani, M. Quandt, M. Scandurra, H. Weigel, Casimir energies in light of quantum field theory. Phys. Lett. B 572, 196 (2003)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  46. N. Graham, R. Jaffe, V. Khemani, M. Quandt, O. Schroeder, H. Weigel, The Dirichlet Casimir problem. Nucl. Phys. B 677, 379 (2004)

    Article  ADS  MATH  Google Scholar 

  47. K.A. Milton, Casimir energies and pressures for delta-function potentials. J. Phys. A 37, 6391 (2004)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  48. J.S. Høye, I. Brevik, Casimir force and its relation to surface tension. Phys. Rev. A 95, 052127 (2017)

    Article  ADS  Google Scholar 

  49. T.-P. Hack, V. Moretti, On the stress-energy tensor of quantum fields in curved spacetimes-comparison of different regularization schemes and symmetry of the Hadamard/Seeley-DeWitt coefficients. J. Phys. A: Math. Theor. 45, 374019 (2012)

    Article  MathSciNet  MATH  Google Scholar 

  50. S. Hollands, R.M. Wald, Conservation of the stress tensor in perturbative interacting quantum field theory in curved spacetimes. Rev. Math. Phys. 17, 227 (2005)

    Article  MathSciNet  MATH  Google Scholar 

  51. B.S. DeWitt, Dynamical Theory of Groups and Fields (Gordon & Breach, New York, 1965)

    MATH  Google Scholar 

  52. C.W. Misner, K.S. Thorne, J.A. Wheeler, Gravitation (Freeman, San Francisco, 1973)

    Google Scholar 

  53. Y.S. Barash, V.L. Ginzburg, Electromagnetic fluctuations in matter and molecular (Van-der-Waals) forces between them. Sov. Phys. Uspekhi 18, 305 (1975)

    Article  ADS  Google Scholar 

  54. E.M. Lifshitz, L.P. Pitaevskii, Statistical Physics, Part 2: Theory of the Condensed State (Pergamon, New York, 1980)

    Google Scholar 

  55. D. Fursaev, D. Vassilevich, Operators, Geometry, and Quanta (Springer, New York, 2011)

    Book  MATH  Google Scholar 

  56. J. Schwinger, J.L.L. Deraad, K.A. Milton, Casimir effect in dielectrics. Ann. Phys. (N.Y.) 115, 1 (1978)

    Article  ADS  MathSciNet  Google Scholar 

  57. R.M. Wald, On the Euclidean approach to quantum field theory in curved spacetime. Commun. Math. Phys. 70, 221 (1979)

    Article  ADS  MathSciNet  Google Scholar 

  58. S.A. Fulling, S.N.M. Ruijsenaars, Temperature, periodicity and horizons. Phys. Rep. 152, 135 (1987)

    Article  ADS  MathSciNet  Google Scholar 

  59. Y.S. Barash, V.L. Ginzburg, Some problems in the theory of van der Waals forces. Sov. Phys. Uspekhi 27, 467 (1984)

    Article  ADS  Google Scholar 

  60. J.A. Stratton, Electromagnetic Theory (Wiley, New York, 2007)

    MATH  Google Scholar 

  61. J. Schwinger, On gauge invariance and Vacuuwn polarization. Phys. Rev. 82, 664 (1951)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  62. L.D. Landau, L.P. Pitaevskii, E.M. Lifshitz, Electrodynamics of Continuous Media (Butterworth-Heinemann, Oxford, 1984)

    Google Scholar 

  63. G.W. Gibbons, S.W. Hawking, Action integrals and partition functions in quantum gravity. Phys. Rev. D 15, 2752 (1977)

    Article  ADS  Google Scholar 

  64. P.B. Gilkey, Invariance Theory, the Heat Equation and the Atyah-Singer Index Theorem (Chemical Rubber Company Press, Boca Raton, 1994)

    Google Scholar 

  65. K.M. Apfeldorf, H.E. Camblong, C.R. Ordóñezc, Field redefinition invariance in Quantum Field Theory. Mod. Phys. Lett. A 16, 103 (2001)

    Article  ADS  MathSciNet  Google Scholar 

  66. F. Ravndal, Problems with the Casimir vacuum energy, in Problems with Vacuum Energy, 8 (2000). arXiv:hep-ph/0009208

  67. J.A. Rackers, J.W. Ponder, Classical Pauli repulsion: an anisotropic, atomic multipole model. J. Chem. Phys. 150, 084104 (2019)

    Article  ADS  Google Scholar 

  68. F.J. Dyson, A. Lenard, Stability of matter. I. J. Math. Phys. 8, 423 (1967)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  69. A. Lenard, F.J. Dyson, Stability of matter. II. J. Math. Phys. 9, 698 (1968)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  70. G. Volovik, The Universe in a Helium Droplet, vol. 117 (Clarendon Press, Oxford, 2006)

    Google Scholar 

  71. K.G. Wilson, The renormalization group and critical phenomena. Rev. Mod. Phys. 55, 583 (1983)

    Article  ADS  MathSciNet  Google Scholar 

  72. J. Mahanty, B.W. Ninham, Dispersion Forces (Academic Press, Boca Raton, 1976)

    Google Scholar 

  73. J.D. Jackson, Classical Electrodynamics (Wiley, New York, 1998)

    MATH  Google Scholar 

  74. S.L. Adler, Qunatum theory of the dielectric constant in real solids. Phys. Rev. 126, 413 (1962)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  75. M. Bordag, K. Kirsten, Heat kernel coefficients and divergencies of the Casimir energy for the dispersive sphere. Int. J. Mod. Phys. A 17, 813 (2002)

    Article  ADS  MathSciNet  Google Scholar 

  76. V. Moretti, One loop stress tensor renormalization in curved background: the relation between zeta function and point splitting approaches, and an improved point splitting procedure. J. Math. Phys. 40, 3843 (1999)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  77. F. Scarf, New soluble energy band problem. Phys. Rev. 112, 1137 (1958)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  78. J. Dereziński, M. Wrochna, Exactly solvable Schrödinger operators. Ann. Henri Poincaré 12, 397 (2011)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  79. M. Abramowitz, I.A. Stegun, Handbook of Mathematical Functions: With Formulas, Graphs, and Mathematical Tables (Dover, New York, 1965)

    MATH  Google Scholar 

  80. J.M. Douillard, Experimental approach of the relation between surface tension and interfacial thickness of simple liquids. J. Colloid Interface Sci. 337, 307 (2009)

    Article  ADS  Google Scholar 

  81. K. Penanen, M. Fukuto, R.K. Heilmann, I.F. Silvera, P.S. Pershan, \(^{4}\)He liquid-vapor interface below \(1 \, {{\rm K}}\) studied using x-ray reflectivity. Phys. Rev. B 62, 9621 (2000)

    Article  ADS  Google Scholar 

  82. W.H. Keesom, The second virial coefficient for rigid spherical molecules whose mutual attraction is equivalent to that of a quadruplet placed at its center. Proc. R. Nether. Acad. Arts Sci. 18, 636 (1915)

    Google Scholar 

  83. G. Luo, S. Malkova, S.V. Pingali, D.G. Schultz, B. Lin, M. Meron et al., Structure of the interface between two polar liquids: Nitrobenzene and water. J. Phys. Chem. Lett. B 110, 4527 (2006)

    Article  Google Scholar 

  84. R.S. Taylor, L.X. Dang, B.C. Garrett, Molecular dynamics simulations of the liquid/vapor interface of SPC/E water. J. Phys. Chem. 100, 11720 (1996)

    Article  Google Scholar 

  85. D. Beaglehole, Ellipsometric study of the surface of simple liquids. Physica B 100, 163 (1980)

    Article  Google Scholar 

  86. E.S. Wu, W.W. Webb, The liquid-vapor interface of Sulfur-Hexafluoride near the critical-point. J. Phys. C1(33), 149 (1972)

    Google Scholar 

  87. F. Buff, R. Lovett, F. Stillinger, Interfacial density profile for fluids in the critical region. Phys. Rev. Lett. 15, 621 (1965)

    Article  ADS  Google Scholar 

  88. J. Pressing, J.E. Mayer, Surface tension and interfacial density. J. Chem. Phys. 59, 2711 (1973)

    Article  ADS  Google Scholar 

  89. D. Beysens, M. Robert, Thickness of fluid interfaces near the critical point from optical reflectivity measurements. J. Chem. Phys. 87, 3056 (1987)

    Article  ADS  Google Scholar 

  90. J.D. van der Waals, Thermodynamische Theorie der Kapillarität unter Voraussetzung stetiger Dichteänderung. Z. Phys. Chemie 13, 657 (1894)

    Article  Google Scholar 

  91. J.W. Cahn, J.E. Hilliard, Free energy of a nonuniform system. I. Interfacial free energy. J. Chem. Phys. 28, 258 (1958)

    Article  ADS  MATH  Google Scholar 

  92. B. Lautrup, Physics of Continuous Matter: Exotic and Everyday Phenomena in the Macroscopic World (CRC Press, Boca Raton, 2011)

    Book  MATH  Google Scholar 

  93. J.E. Lennard-Jones, B.M. Dent, The change in lattice spacing at a crystal boundary. Proc. R. Soc. Lond. A 121, 246 (1928)

    ADS  Google Scholar 

  94. G. Jaffé, A statistical theory of liquids. I. Phys. Rev. 62, 463 (1942)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  95. J.G. Kirkwood, F.P. Buff, The statistical mechanical theory of surface tension. J. Chem. Phys. 17, 338 (1949)

    Article  ADS  Google Scholar 

  96. P.D. Shoemaker, G.W. Paul, L.E. Marc de Chazal, Surface tension of simple liquids from the radial distribution function. J. Chem. Phys. 52, 491 (1970)

    Article  ADS  Google Scholar 

  97. M.V. Berry, R.F. Durrans, R. Evans, The calculation of surface tension for simple liquids. J. Phys. A: Gen. Phys. 5, 166 (1972)

    Article  ADS  MATH  Google Scholar 

  98. J. Walton, D.J. Tildesley, J.S. Rowlinson, J.R. Henderson, The pressure tensor at the planar surface of a liquid. Mol. Phys. Int. J. Interface Between Chem. Phys. 48, 1357 (1983)

    Google Scholar 

  99. R. Laghaei, A.E. Nasrabad, B.C. Eu, Statistical-mechanical theory of rheology: Lennard-Jones fluids. J. Chem. Phys. 123, 234507 (2005)

    Article  ADS  Google Scholar 

  100. T. Yamaguchi, Stress-structure coupling and nonlinear rheology of Lennard-Jones liquid. J. Chem. Phys. 148, 243507 (2018)

    Article  Google Scholar 

  101. A. Jabbarzadeh, J.D. Atkinson, R.I. Tanner, Rheological properties of thin liquid films by molecular dynamics simulations. J. Non-Newtonian Fluid Mech. 69, 169 (1997)

    Article  Google Scholar 

  102. A. Trokhymchuk, J. Alejandre, Computer simulations of liquid/vapor interface in lennard-jones fluids: Some questions and answers. J. Chem. Phys. 111, 8510 (1999)

    Article  ADS  Google Scholar 

  103. J. Baacke, G. Krüsemann, Perturbative analysis of the divergent contributions to the Casimir energy. Z. Phys. C Particles Fields 30, 413 (1986)

    Article  ADS  Google Scholar 

  104. I.H. Brevik, B. Jensen, K. Milton, Comment on Casimir energy for spherical boundaries. Phys. Rev. D 64, 088701 (2001)

    Article  ADS  MathSciNet  Google Scholar 

  105. C. Lutken, F. Ravndal, Energy level shifts in atoms between metallic planes. Phys. Rev. A 31, 2082 (1985)

    Article  ADS  Google Scholar 

  106. R. Shankar, Effective field theory in condensed matter physics, in Conceptual Foundations of Quantum Field Theory, ed. by T.Y. Cao (Cambridge University Press, Cambridge, 1999), pp. 47–55

    Chapter  Google Scholar 

  107. P.A. Lagerstrom, Matched Asymptotic Expansions (Springer, New York, 1988)

    Book  MATH  Google Scholar 

  108. K.A. Milton, Julian Schwinger and the Casimir effect: the reality of zero point energy. Conf. Proc. C 9809142, 20 (1999)

    MathSciNet  MATH  Google Scholar 

  109. L. Bergström, Hamaker constants of inorganic materials. Adv. Colloid Interface Sci. 70, 125 (1997)

    Article  Google Scholar 

  110. E. Orowan, Surface energy and surface tension in solids and liquids. Proc. R. Soc. Lond. A 316, 473 (1970)

    Article  ADS  Google Scholar 

  111. S.M. Christensen, Vacuum expectation value of the stress tensor in an arbitrary curved background: the covariant point-separation method. Phys. Rev. D 14, 2490 (1976)

    Article  ADS  MathSciNet  Google Scholar 

Download references

Acknowledgements

This work was partially supported by the Natural Sciences and Engineering Research Council of Canada (NSERC). A.Z. also acknowledges financial support by the Killam Trust.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Andrei Zelnikov.

Heat kernel calculations

Heat kernel calculations

Let us define the new field

$$\begin{aligned} \begin{aligned} \phi =\mathrm {g}^{-1/4}\varPhi =\varepsilon ^{3/4}\varPhi \end{aligned}\end{aligned}$$
(78)

and the corresponding Green’s function

$$\begin{aligned} \begin{aligned} {\mathbb {G}}(\zeta ;{\varvec{x}},{\varvec{x'}})=\langle \upphi ({\varvec{x}})\upphi ({\varvec{x'}})\rangle =\mathrm {g}^{-1/4}({\varvec{x}})\,\mathrm {g}^{-1/4}({\varvec{x}}')\,{\widehat{G}}(\zeta ;{\varvec{x}},{\varvec{x'}}), \end{aligned}\end{aligned}$$
(79)

where \(\upphi ({\varvec{x}})\) is the operator corresponding to the field \(\phi ({\varvec{x}})\). Then, the inner product for this new field takes the covariant form \(\int \mathrm {d}{\varvec{x}}\, \varPhi _1 \varPhi _2=\int \mathrm {d}{\varvec{x}}\,\varepsilon ^{-3/2}\, \phi _1 \phi _2=\int \mathrm {d}{\varvec{x}}\sqrt{\mathrm {g}}\, \phi _1 \phi _2\). Equation (3) transforms to

$$\begin{aligned} \begin{aligned} {\mathbb {O}} \, {\mathbb {G}}(\zeta ;{\varvec{x}},{\varvec{x'}})=\frac{\delta ({\varvec{x}}-{\varvec{x'}})}{\sqrt{\mathrm {g}}}, \ \text {where} \ {\mathbb {O}}=\mathrm {g}^{-1/4}({\varvec{x}})\,{\widehat{O}}\,\mathrm {g}^{1/4}({\varvec{x}}). \end{aligned}\end{aligned}$$
(80)

Taking into account that

$$\begin{aligned} \begin{aligned} \partial _j \varepsilon ^{-3/4}&=\varepsilon ^{-3/4}\Big (\partial _j-\frac{3}{4}\eta _j \Big )\\ \partial _i\partial _j \varepsilon ^{-3/4}&=\varepsilon ^{-3/4} \Big (\partial _i\partial _j-\frac{3}{4}\eta _i\partial _j -\frac{3}{4}\eta _j\partial _i-\frac{3}{4}(\partial _i\eta _j)+\frac{9}{16}\eta _i\eta _j \Big ), \end{aligned}\end{aligned}$$
(81)

the new operator \({\mathbb {O}}\) becomes

$$\begin{aligned} \begin{aligned} {\mathbb {O}}&=\varepsilon \delta ^{ij}\,\Big (\partial _i\partial _j-\frac{1}{2}\eta _i\partial _j\Big ) -\frac{3}{4}\varepsilon \delta ^{ij}\,\Big (\partial _i\eta _j+\frac{1}{4}\eta _i\eta _j \Big ). \end{aligned}\end{aligned}$$
(82)

Note that

$$\begin{aligned} \begin{aligned} \nabla _k\eta ^k=\varepsilon \delta ^{ij}\Big (\partial _i\eta _j-\frac{1}{2}\eta _i\eta _j\Big )=\mathrm {g}^{ij}\Big (\partial _i\eta _j-\frac{1}{2}\eta _i\eta _j\Big ) \end{aligned}\end{aligned}$$
(83)

and the covariant Laplacian in metric \(\mathrm {g}_{ij}\) reads

$$\begin{aligned} \begin{aligned} \bigtriangleup&=\mathrm {g}^{ij}\nabla _i\nabla _j=\frac{1}{\sqrt{\mathrm {g}}}\partial _i\Big (\mathrm {g}^{ij}\sqrt{\mathrm {g}}\,\partial _j\Big ) =\mathrm {g}^{ij}\,\Big (\partial _i\partial _j-\frac{1}{2}\eta _i\partial _j \Big ). \end{aligned}\end{aligned}$$
(84)

Therefore, we obtain the covariant form for the operator in (80)

$$\begin{aligned} \begin{aligned} {\mathbb {O}}=\bigtriangleup -{\mathbb {V}}, \ \text {where} \ {\mathbb {V}}=\frac{3}{4}\Big (\nabla ^k\eta _k+\frac{3}{4}\eta ^k\eta _k\Big ) =\frac{3}{4}\varepsilon \delta ^{ij}\,\Big (\partial _i\eta _j+\frac{1}{4}\eta _i\eta _j \Big ), \end{aligned}\end{aligned}$$
(85)

which is Hermitian with the covariant measure \(\sqrt{\mathrm {g}}\). The potential \({\mathbb {V}}\) can also be obtained following the computations in [31, 64], which require the knowledge of the Christoffel symbols

$$\begin{aligned} \varGamma ^i_{jk}=-\frac{1}{2}\big [\delta _{ij}\eta _k+\delta _{ik}\eta _j -\delta _{jk}\eta _i \big ], \ \mathrm {g}^{jk}\varGamma ^i_{jk}&=\frac{1}{2} \varepsilon \,\eta _i, \end{aligned}$$
(86)

and the connection

$$\begin{aligned} \begin{aligned} \omega _i=\frac{1}{2}\big [\eta _i+\mathrm {g}_{ik}\mathrm {g}^{mn}\varGamma ^k_{mn}\big ]=\frac{3}{4}\eta _i. \end{aligned}\end{aligned}$$
(87)

Heat kernel corresponding to operator \({\widehat{O}}\) is \({\widehat{K}}(s|{\varvec{x}},{\varvec{x'}})=\mathrm {g}^{1/4}({\varvec{x}})\mathrm {g}^{1/4}({\varvec{x}}'){\mathbb {K}}(s|{\varvec{x}},{\varvec{x'}})\), where

$$\begin{aligned} \begin{aligned} {\mathbb {K}}(s|{\varvec{x}},{\varvec{x}}')=-e^{s{\mathbb {O}}}=-\frac{\varDelta ^{1/2}({\varvec{x}},{\varvec{x}}')}{(4\pi s)^{3/2}}e^{-\frac{\sigma ({\varvec{x}},{\varvec{x}}')}{2s}}\Big (a_0({\varvec{x}},{\varvec{x}}')+s\,a_1({\varvec{x}},{\varvec{x}}')+\dots \Big ); \end{aligned}\end{aligned}$$
(88)

here s is the proper time, \(\sigma ({\varvec{x}},{\varvec{x}}')\) a world function defined as one half of the square of geodesic distance between points in the metric \(\mathrm {g}_{ij}\) and \(\varDelta ^{1/2}({\varvec{x}},{\varvec{x}}') \equiv \mathrm {g}^{-1/2}({\varvec{x}}) \mathfrak {D}({\varvec{x}},{\varvec{x}}') \mathrm {g}^{-1/2}({\varvec{x}}')\), where \(\mathfrak {D}({\varvec{x}},{\varvec{x}}') = -\det \left( -\sigma _{;\mu \nu '}\right) \) is the van Vleck–Morette determinant. For the computation of the heat kernel of this operator, we need to know \(\varDelta ^{1/2}({\varvec{x}},{\varvec{x}}')\), \(a_0({\varvec{x}},{\varvec{x}}')\), and \(a_1({\varvec{x}},{\varvec{x}}')\) and their derivatives. In particular they include the curvature terms:

$$\begin{aligned} R_{ij}=\frac{1}{2}\partial _j\eta _i+\frac{1}{4}\eta _i \eta _j + \frac{1}{4} \delta _{ij}\delta ^{km}\big (2\partial _k\eta _m - \eta _k\eta _m\big ), \quad R = 2\varepsilon \,\delta ^{km}\Big (\partial _k\eta _m-\frac{1}{4}\eta _k\eta _m\Big ); \end{aligned}$$
(89)

here we used the relations

$$\begin{aligned}&\eta _{i;j}=\partial _j\eta _i+\eta _i\eta _j-\frac{1}{2}\delta _{ij}\delta ^{km}\eta _{k}\eta _m, \end{aligned}$$
(90a)
$$\begin{aligned} &\eta ^k{}_{;k}=\varepsilon \delta ^{ij} \big (\partial _j\eta _i-\frac{1}{2}\eta _{i}\eta _j), \partial _i (\mathrm {g}^{1/4})=-\frac{3}{4}\varepsilon ^{-3/4}\eta _i. \end{aligned}$$
(90b)

The regularized Green’s function (79), in the original notation, assumes the form

$$\begin{aligned} \begin{aligned} {\widehat{G}}_{\lambda _{0}}(\zeta ;{\varvec{x}},{\varvec{x'}})=\int _{\lambda _{0}^{2}}^{\infty } \mathrm {d}s\, {\widehat{K}}(s|{\varvec{x}},{\varvec{x}}') =\mathrm {g}^{1/4}({\varvec{x}})\mathrm {g}^{1/4}({\varvec{x}}')\int _{\lambda _{0}^{2}}^{\infty } \mathrm {d}s\,{\mathbb {K}}(s|{\varvec{x}},{\varvec{x'}}), \end{aligned}\end{aligned}$$
(91)

where \(\lambda _{0}\) is the cut-off parameter, and can be expanded as

$$\begin{aligned} \begin{aligned} {\widehat{G}}_{\lambda _{0}}(\zeta ;{\varvec{x}},{\varvec{x'}})=-\int _{\lambda _{0}^{2}}^{\infty } \mathrm {d}s\, \frac{\mathfrak {D}^{1/2}({\varvec{x}},{\varvec{x}}')}{(4\pi s)^{3/2}}e^{-\frac{\sigma ({\varvec{x}},{\varvec{x}}')}{2s}}\Big (a_0({\varvec{x}},{\varvec{x}}')+s\,a_1({\varvec{x}},{\varvec{x}}')+\dots \Big ) \end{aligned}\end{aligned}$$
(92)

Derivation of the regularized stress tensor then reduces to the following steps:

  1. i)

    Compute \(\varDelta ^{1/2}({\varvec{x}},{\varvec{x}}')\), \(a_0({\varvec{x}},{\varvec{x}}')\), and \(a_1({\varvec{x}},{\varvec{x}}')\);

  2. ii)

    Compute the second partial derivative of the Green’s function \(\partial _{{\varvec{x}}}\partial _{{\varvec{x}}'}{\widehat{G}}_{\lambda _{0}}(\zeta ;{\varvec{x}},{\varvec{x'}})\);

  3. iii)

    Take the limit \({\varvec{x}}={\varvec{x'}}\);

  4. iv)

    Integrate over the proper time s and then expand the result in the cut-off \(\lambda _{0}\);

  5. v)

    Compute \({\widehat{\sigma }}_{ij}(\zeta ;{\varvec{x}})\) by substituting the result to (11).

For the quantities in the limit of coincident points, we introduce notation \([\dots ]\equiv (\dots )|_{{\varvec{x}}={\varvec{x'}}}\). The expressions that we need to compute divergent parts (33) of the Green’s function \({\widehat{G}}_{\lambda _{0}}(\zeta ;{\varvec{x}},{\varvec{x'}})\) are [111]

$$\begin{aligned} {[}\sigma ]&=0, \quad [\sigma _{;i}]=0, \quad [\sigma _{;i'}]=0, \quad [\sigma _{;ij}]=\mathrm {g}_{ij}, [\sigma _{;ij'}]=-\mathrm {g}_{ij}, \end{aligned}$$
(93a)
$$\begin{aligned} {[}a_0]&=1, \quad [a_0{}_{;i}]=0, ~ [a_0{}_{;ij'}]=0, ~ [a_1]=\frac{1}{6}R-{\mathbb {V}}, ~ [a_1{}_{;i}]=\frac{1}{2}\big (\frac{1}{6}R_{;i}-{\mathbb {V}}_{;i}\big ), \end{aligned}$$
(93b)
$$\begin{aligned} {[}\varDelta ^{1/2}]&=1, \quad [(\varDelta ^{1/2})_{;i}]=0, \quad [(\varDelta ^{1/2})_{;ij}]=\frac{1}{6}R_{ij}, \quad [(\varDelta ^{1/2})_{;ij'}]=-\frac{1}{6}R_{ij}, \end{aligned}$$
(93c)
$$\begin{aligned}&[\partial _i\partial _{j'}\big (\mathrm {g}^{1/4}({\varvec{x}})\mathrm {g}^{1/4}({\varvec{x'}})\varDelta ^{1/2}\big )] =\varepsilon ^{-3/2}\left( \frac{9}{16}\eta _i\eta _j-\frac{1}{6}R_{ij}\right) . \end{aligned}$$
(93d)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Zelnikov, A., Krechetnikov, R. Scarf for Lifshitz. Eur. Phys. J. Plus 136, 755 (2021). https://doi.org/10.1140/epjp/s13360-021-01714-3

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1140/epjp/s13360-021-01714-3

Navigation