1: \documentclass[twocolumn,showpacs,preprintnumbers,amsmath,amssymb]{revtex4}
2: %\documentclass[preprint,showpacs,preprintnumbers,amsmath,amssymb]{revtex4}
3:
4: % Some other (several out of many) possibilities
5: %\documentclass[preprint,aps]{revtex4}
6: %\documentclass[preprint,aps,draft]{revtex4}
7: %\documentclass[prb]{revtex4}% Physical Review B
8: \usepackage{graphicx}% Include figure files
9: \usepackage{dcolumn}% Align table columns on decimal point
10: \usepackage{bm}% bold math
11:
12:
13: \begin{document}
14:
15:
16: \title{The Low-Frequency Character of the Thermal Correction to the Casimir Force between
17: Metallic Films\\
18: LA-UR 02-4696}
19:
20: \author{J.R. Torgerson and S.K. Lamoreaux}
21:
22: \affiliation{University of California,Los Alamos National
23: Laboratory,Physics Division P-23, M.S. H803, Los Alamos, NM 87545}
24:
25: \date{today}
26:
27:
28: \begin{abstract}
29:
30: The frequency spectrum of the finite temperature correction to the
31: Casimir force can be determined by use of the Lifshitz formalism
32: for metallic plates of finite conductivity. We show that the
33: correction for the $TE$ electromagnetic modes is dominated by
34: frequencies so low that the plates cannot be modelled as ideal
35: dielectrics. We also address issues relating to the behavior of
36: electromagnetic fields at the surfaces and within metallic
37: conductors, and calculate the surface modes using appropriate
38: low-frequency metallic boundary conditions. Our result brings the
39: thermal correction into agreement with experimental results that
40: were previously obtained. We suggest a series of measurements
41: that will test the veracity of our analysis.
42:
43: \end{abstract}
44: \pacs{12. 20. Ds, 41. 20. -q, 12. 20. Fv}
45: \maketitle
46:
47: \section{Introduction}
48:
49: A recent paper \cite{1}, in which finite conductivity and
50: temperature corrections to the Casimir force between metal plates
51: are simulatenously considered, suggests a large thermal correction
52: to the force at distances greater than about 1 $\mu$m. This
53: correction deviates significantly from experimental results
54: \cite{2,decca1} and previous theoretical work, and has attracted
55: considerable interest. The principal conclusion in \cite{1}
56: leading to this discrepancy is that the $TE$ electromagnetic mode
57: (${\bf E}$ parallel to the surface) does not contribute to the
58: force at finite temperature. Arguments against the analysis given
59: in \cite{1} have been numerous \cite{3,4,5,6} but the arguments
60: have not been universally accepted \cite{7,8}.
61:
62: A careful numerical analysis of the problem leads us and others to
63: conclude that the results presented in \cite{1} are mathematically
64: correct. As we show here, this analysis does not accurately
65: represent the experimental arrangement used in \cite{2}. The
66: aspect of the problem that has not been considered in detail is
67: the appropriateness of a dielectric model of the metallic plates
68: at low frequencies, which, as we will show, are most relevant for
69: the thermal correction. The first purpose of this note is to
70: expand on our previous work \cite{9} and to point out that the
71: proper boundary conditions for conductors have not yet been
72: directly applied to this problem, and to show that the
73: experimental result \cite{2} can be fully explained by this
74: application.
75:
76: The second purpose of this note is to contrast the points of view
77: put forward in \cite{1} and \cite{15}. Use of the surface
78: impedance to calculate the waveguide modes, as was done in
79: \cite{15} allows description of the Casimir force by a single
80: analytic function in the complex $\omega$ plane \cite{12}.
81: Treating metals as dielectrics, as was done in \cite{1}, leads to
82: the requirement that different boundary conditions must be used
83: when the skin depth of the electromagnetic field is smaller than
84: the electron mean free path in the metal. Therefore, with the
85: dielectric treatment, the Casimir force cannot be described by a
86: single analytic function so the techniques used in \cite{1} are
87: not applicable to the problem.
88:
89: Finally, we suggest that the analysis in \cite{1} is applicable to
90: insulating dielectrics, and possibly to materials such as
91: intrinsic or lightly dope Ge or Si where the skin depth is longer
92: than the electron mean free path. Measurements with a dielectric
93: such as Diamond would provide an excellent test of the theory and
94: allow the possibility to discharge the surface by use of
95: ultraviolet light. The ultimate purposes of this note are to call
96: for further theoretical studies and experimental measurements as
97: suggested here.
98:
99: \section{Spectrum of the $TE$ Mode Thermal Correction of the Casimir Force}
100:
101:
102: Following Ford \cite{10}, the spectrum of the Casimir force is
103: given by Eqs. (2.3) and (2.4) of Lifshitz' seminal paper
104: \cite{11}. We note that
105: \begin{equation}
106: {1\over 2}\coth {\hbar\omega\over 2kT}={1\over 2} + {1\over
107: \exp(\hbar\omega/ kT)-1}={1\over 2} + g(\omega)
108: \end{equation}
109: and we only include the second term on the right-hand side in the
110: determination of the spectrum of the thermal correction. From Eq.
111: (2.4) of \cite{11}, the spectrum of the $TE$ mode excitation
112: between parallel plates can be described by
113: \begin{eqnarray}\label{lif}
114: \left[{\hbar\over \pi^2 c^3}\right]F_\omega&=&\left[{\hbar\over
115: \pi^2 c^3}\right]\omega^3 g(\omega)\cr &\times &{\rm Re}\int_C
116: p^2dp
117: \left[{(s+p)^2\over(s-p)^2}e^{-2ip\omega a/c}-1\right]^{-1},
118: \end{eqnarray}
119: \begin{equation}\label{s}
120: s=\sqrt{\epsilon(\omega)-1+p^2}
121: \end{equation}
122: where $a$ is the plate separation, and we have assumed that the
123: plates are made of the same material with vacuum between them. The
124: integration path $C$ can be separated into $C_1$ for $p=1$ to 0,
125: which describes the effect of plane waves, and $C_2$ with pure
126: imaginary values $p=i0$ to $i\infty$ for exponentially damped
127: (evanescent) waves.
128:
129: In anticipation that the effect is a low-frequency phenomenon, we
130: use the parameters for Au in \cite{1} for ${\rm Im}\
131: \epsilon=\epsilon_2$ and employ the Kramers-Kronig relations to
132: determine ${\rm Re}\ \epsilon=\epsilon_1$. We find for frequencies
133: $\omega<10^{14}$ s$^{-1}$ that, to good approximation,
134: \begin{equation}\label{epsilon}
135: \epsilon_1={-1.48\times 10^{4}\over 1+(\omega/\omega_0)^2};\ \
136: \epsilon_2={1.8\times 10^{18}\over \omega(1+(\omega/\omega_0)^2)}
137: \end{equation}
138: with $\omega_0=3.3\times 10^{13}\ {\rm s}^{-1}$.
139:
140: In \cite{1}, a net deviation from the zero-temperature value of
141: the Casimir force is predicted to be about 25\% for a plate
142: separation of $1\ \mu$m at 300 K. The experimental results
143: reported in \cite{2} had their greatest sensitivity around 1
144: $\mu$m, and disagree significantly with the results in \cite{1}.
145: As a comparison, we numerically integrate Eq. {\ref{lif}) for
146: $a=1\ \mu$m and $T=300$ K, using Eq. (\ref{epsilon}) for the
147: permittivity. The results are shown in Fig. 1, where we have
148: separated the results from the two integration paths. In Fig.
149: 1a, it can be seen that there is no significant deviation from the
150: perfectly conducting case. On the other hand, the contribution
151: from evanescent waves, shown in Fig. 1b, is large and the
152: integrated value is in good agreement with the result given in
153: \cite{1}.
154:
155: We see immediately that the main contributions of the $TE$-mode
156: finite conductivity correction are around $\omega=10^{10}-10^{13}$
157: s$^{-1}$. This behavior is due to an approximately quadratic
158: increase with $\omega$ of the $C_2$ integral and a suppression
159: beginning at $\omega=kT/\hbar=4\times 10^{13}$ s$^{-1}$ due to
160: $g(\omega)$. This is a low frequency range and we can question
161: certain assumptions in \cite{1} and in the Lifshitz calculation,
162: among others, in regard to theoretical predictions relevant to the
163: experimental arrangement in \cite{2}.
164:
165:
166: \section{Low Frequency Limit and Field Behavior in Metallic Materials}
167:
168:
169: When the depth of penetration of the electromagnetic field into a
170: metal,
171: \begin{equation}
172: \delta=c/\sqrt{2\pi\mu\sigma\omega}
173: \end{equation}
174: where $\sigma$ is the conductivity and $\mu$ is the permeability
175: (for Au and Cu, $\sigma\approx 3\times 10^{17}\ {\rm s}^{-1}$,
176: $\mu=1$), becomes of the same order as the mean free path of the
177: conduction electrons, it is no longer possible to describe the
178: field in terms of a dielectric permeability \cite{12,13}. This
179: occurs for optical frequencies $\omega\approx 5\times 10^{13}$
180: s$^{-1}$ for metals such as Au and Cu where the mean free path, at
181: 300 K, is about $3\times 10^{-6}$ cm \cite{14} (p. 259). At
182: frequencies above $10^{14}$ s$^{-1}$ the permeability description
183: again becomes valid because on absorbing a photon, a conduction
184: electron acquires a large kinetic energy and has a shortened mean
185: free path. However, in the interaction of a field with a material
186: surface, $\bf E$ and $\bf H$ can always be related linearly
187: through the surface impedance (which relates the electric field at
188: the surface to a surface current hence magnetic field); this
189: approach has been used in calculation of the Casimir force
190: \cite{15}. A related correction arises from the the plasmon
191: interaction with the surface which becomes significant near the
192: plasma frequency of the metal, and has been estimated as nearly
193: 10\%\space \cite{16} for sub-$\mu$m plate separations.
194:
195: The proper boundary conditions for a conducting plane have been
196: discussed by Boyer \cite{17}. He points out that when (using here
197: the notation of \cite{1}) $\omega\ll\eta^2\rho/4\pi$, where $\rho$
198: is the resistivity and $\eta$ is the dissipation, the usual
199: dielectric boundary conditions are not applicable. For Au, using
200: the parameters in \cite{1}, this limit is met for
201: $\omega\ll4\times 10^{14}\ {\rm s}^{-1}$. This corresponds to an
202: optical wavelength of 5\,$\mu$m, which implies that for plate
203: separations significantly larger than this, and of course for
204: $\omega\rightarrow 0$, the plates must be treated as good
205: conductors.
206:
207: The boundary conditions for a conducting surface are discussed in
208: \cite{18} (Sec. 8.1). At low frequencies (e.g., where the
209: displacement current can be neglected), a tangential electric
210: field at the surface of a conductor will induce a current ${\bf
211: j}_\|=\sigma {\bf E}_\|$, where $\sigma$ is the conductivity. The
212: presence of the surface current leads to a discontinuity in the
213: normal derivative of ${\bf H}_\|$, hence a discontinuity in the
214: normal derivative of ${\bf E}_\|$, at the boundary of a conducting
215: surface. These boundary conditions are quite different from the
216: dielectric case where the fields and their derivatives are assumed
217: continuous.
218:
219: These boundary conditions are applicable when the skin depth of
220: the electromagnetic field is much smaller that the characteristic
221: wavelength of the field. The wavelengths that contribute most to
222: the Casimir force correspond to wavevector $k\approx 1/4a$, {\it
223: independent of frequency}, by numerical determination. When
224: $k<\sqrt{2}/\delta$, the boundary conditions are applicable. This
225: is well-satisfied over the entire frequency range of the finite
226: temperature effect for the conditions of the experiment \cite{2};
227: when $\omega>10^{11}\ {\rm s^{-1}}$ in which case $\delta< 0.7\
228: \mu$m and the relationship, for $a=1\ \mu$m
229: \begin{equation}
230: {1\over 4 a}= 2.5\times 10^3< {\sqrt{2}\over \delta}=1.4\times
231: 10^4\ {\rm cm}
232: \end{equation}
233: at the lowest frequency of interest. In this frequency range,
234: specifying the $k$ vector in the material as a boundary condition
235: is not warranted.
236:
237: This can also be understood by noting that the propagation of
238: electromagnetic field in a conductive material is described by the
239: diffusion equation. If we imaging a spatially periodic varying
240: field on the surface of the material as $\exp(ikz)$, the
241: variations propagate into the material, damped exponentially as
242: $\exp(-(k^2+2/\delta^2)z)$ into the material, where
243: $\sqrt{2}/\delta$ is interpreted as the diffusion length. We
244: therefore see that over the frequency range of interest, the
245: conducting boundary conditions are appropriate. In this limit,
246: the displacement current is small compared to the real current,
247: ${\bf j}=\sigma {\bf E}$, for good conductors of interest here.
248:
249: \section{Electromagnetic Modes between Metallic Plates}
250:
251: We are interested in modes between two conducting plates separated
252: by a distance $a$. In the limit that the plates are thin films of
253: thickness $\delta$, the skin depth, we can assume that the plates
254: are infinitely thick and the problem is considerably simplified.
255: This is valid for the experiment \cite{2} where the Cu/Au metallic
256: film was 1 $\mu$m. Essentially all of the $TE$ mode thermal
257: correction comes in the $10^{11}$ and $10^{13}$ s$^{-1}$ range as
258: shown in Fig. 1, so $0.07<\delta<0.7\ \mu$m.
259:
260: Taking the $\hat z$ axis as perpendicular to the plates, and the
261: mode propagation direction along $\hat x$, for the case of $TE$
262: modes (also referred to as $H$ or magnetic modes), $E_x=0$. The
263: plates surfaces are located at $z=0$ and $z=a$. For a perfect
264: conductor, $\partial H_z/\partial z=0$ at the conducting surfaces.
265: A finite conductivity makes this derivative non-zero, and can be
266: estimated from the small electric field $E_y$ that exists at the
267: surface of the plate, (see \cite{18}, Sec. 8.1 and Eq. (8.6)),
268: %
269: \begin{equation}
270: \vec E_\|=\hat y E_y=\sqrt{\omega\over 8\pi \sigma}(1-i)\hat
271: n\times\vec H_\|.
272: \end{equation}
273: %
274: where $\vec H_\|= \hat x H_x$ and it is assumed that the
275: displacement current in the metal plate can be neglected
276: ($\sigma>>\omega$), and that the inverse of the mode wavenumber is
277: less than $\delta$. $E_y$ and $H_x$ are related through Maxwell's
278: equation $\vec\nabla\times\vec H=\partial \vec E/c\partial t$.
279: Assuming a time dependence of $e^{-i\omega t}$, and vacuum between
280: the plates,
281: %
282: \begin{equation}\label{diffeq}
283: {\partial H_x\over \partial z}=\pm {i\omega\over c} E_y.
284: \end{equation}
285: where $\pm$ indicates sign of $\hat n$ at $z=0$ and $z=a$
286: respectively. The boundary conditions at the surfaces are thus
287: \begin{equation}
288: {\partial\over\partial z}H_x=\pm i\sqrt{\omega\over 8 \pi
289: \sigma}\left({\omega\over c}\right) (1-i)H_x\equiv \pm \alpha H_x.
290: \end{equation}
291: %
292: %
293: Solutions of the form $H_x(z)=Ae^{Kz}+Be^{-Kz}$, where
294: $K^2=k^2-\omega^2/c^2$ and $k$ is the transverse wavenumber, can
295: be constructed for the space between the conducting plates. The
296: eigenvalues $K$ can be determined by the requirement that Eq.
297: (\ref{diffeq}) be satisfied at $z=0$ and $z=a$. With the usual
298: substitution $\omega=i\xi$, the eigenvalues $K$ are then given by
299: (see \cite{19}, Sec. 7.2)
300: %
301: \begin{equation}\label{gte}
302: G_{TE}(\xi)\equiv{(\alpha + K)^2\over (\alpha-K)^2}e^{2Ka}-1=0
303: \end{equation}
304: %
305: and the force can be calculated by the techniques outlined in
306: \cite{19}, Sec. 7.3.
307:
308: This result can be recast in the notation of the Lifshitz
309: formalism, and the spectrum of the thermal correction can be
310: calculated as before. Noting that $K=i\omega p/c$,
311: \begin{equation}
312: F_\omega=\omega^3g(\omega)\int_C p^2dp\left[{(\alpha+i\omega
313: p/c)^2\over(\alpha-i\omega p/c)^2}e^{-2i\omega p/c}-1\right]^{-1}.
314: \end{equation}
315: Results of a numerical integration are shown in Fig. 2, where it
316: can be seen by comparison with Fig. 1 that the metallic plate
317: boundary condition does not show a significant contribution from
318: the $C_2$ integral of the $TE$ mode thermal correction and is
319: therefore similar to that for the ``perfect conductor'' boundary
320: condition. This reconciles the discrepancy between the prediction
321: in \cite {1} and the experimental results reported in \cite{2}.
322: Note that the function Eq. (\ref{gte}) in only applicable where
323: the skin depth is small compared to the mode wavelength. Our
324: result is in agreement, in its range of applicability, with the
325: analysis presented in \cite{15} which is valid for all
326: frequencies. In this note, we essentially determined the surface
327: impedance from the bulk properties, which is possible in over the
328: frequency range of interest.
329:
330: \section{Conclusion}
331:
332: The problem of calculating the $TE$ mode contribution to the
333: Casimir force has been previously treated with the ``Schwinger
334: prescription'' \cite{20} of setting the dielectric constant to
335: infinity before setting $\omega=0$. This prescription has become
336: controversial \cite{21}, a term that can be used to describe the
337: entire history of the theory of the temperature correction.
338: However, there is no doubt that the issues brought up in \cite{1}
339: are important.
340:
341: The purpose of our calculation is to take a different approach and
342: to study the low-frequency behavior of the correction in order to
343: understand its character. We have shown that the finite
344: temperature correction in \cite{1} is a low-frequency phenomenon.
345: The frequency is sufficiently low so that treating the plates as
346: bulk dielectrics is not valid. By use of a more realistic
347: description of the field interaction with the plates we show that
348: the modes between metallic plates of finite conductivity produce a
349: finite temperature correction in close agreement with the
350: perfectly conducting case. The principal difference between our
351: result and the previous work is that we have allowed for the
352: possibility that the derivatives of the fields at the conducting
353: boundary are discontinuous. This possibility exists because the
354: fields produce currents in the conducting plates that are
355: discontinuous across the boundary between the vacuum and the
356: conductor. Although it is tempting to model the finite
357: conductivity as a modification to the dielectric permittivity,
358: such a model fails when the mean free path of the conduction
359: electrons exceeds the penetration depth of the electromagnetic
360: field, and thus fails for frequencies of interest for the thermal
361: correction to the $TE$ electromagnetic mode.
362:
363: We have shown that the conducting boundary conditions that are
364: applicable for frequencies where the $TE$ mode thermal correction
365: has its significant contribution lead to a net increase of the
366: $TE$ mode force, and is of the same magnitude as the perfectly
367: conducting case. This result is in agreement with the
368: experimental results reported in \cite{2}. However, additional
369: and improved experiments with large plate separations (greater
370: than 2 $\mu$m) with both conducting and dielectric plates would
371: provide the definitive test. A particularly tempting dielectric
372: would be diamond which offers both theoretical and experimental
373: benefits: its dielectric properties can be calculated from
374: first-principles, and stray surface charges can be eliminated by
375: exposing it to ultraviolet light, making it photoconductive. A
376: semiconductor such as lightly-doped Germanium would also provide a
377: useful test of this theory. Ge with resistivity $40\ \Omega$cm is
378: readily available and would have a skin depth about 1,000 times
379: that of Cu or Au. Additional high-accuracy measurements at
380: long-range with Au or Cu are also important for testing the
381: theory. We are presently constructing a new torsion pendulum
382: system that will be able to measure the Casimir force with 1\%
383: accuracy at distances greater than 2 $\mu$m, at a fixed
384: temperature of 300 K. We hope that this note will spur further
385: theoretical work on the questions and basic analysis presented
386: here.
387:
388: \bibliography{L2}
389: \bibliographystyle{plain}
390:
391:
392:
393: \begin{figure*}
394: \includegraphics{fig1L.eps}
395: \caption{The net finite-temperature contribution to the Casimir
396: force is determined by $F=(\hbar/\pi c^3)\int_0^\infty F_\omega
397: d\omega$ and is attractive when $F>0$. a: The two curves represent
398: the $C_1$ path for perfectly conducting plates (dashed curve) and
399: for plates with permittivity given by Eq. (\ref{epsilon})(solid
400: curve). The net force force for the latter is $0.95$ times the
401: perfectly conducting case. b: For a perfect conductor, the $C_2$
402: integral is zero. The net contribution from the $C_2$ path is
403: $-169$ times the perfectly conducting contribution from the $C_1$
404: path, and its addition to the $TE$ mode zero-point contribution
405: reduces the net $TE$ mode force to nearly zero, which is the
406: result obtained in \cite{1}. All are for $a=1\ \mu$m, $T=300$ K.}
407: \end{figure*}
408:
409: \begin{figure*}
410: \includegraphics{fig2L.eps}
411: \caption{Numerical results for $F_\omega$ using the finite
412: conductivity boundary conditions. The integrated force for the
413: $C_2$ path contribution is $1.47$ times greater than the $C_1$
414: integration, and the total net force for both paths is $1.75$
415: times greater than the perfectly conducting case. Treatment of the
416: plates as conducting metals fails above $\omega =10^{14}\ {\rm
417: s}^{-1}$. All are for $a=1\ \mu$m, $T=300$ K.}
418: \end{figure*}
419:
420: \end{document}
421: