1: % v0 draft 2003/10/30
2: % v1 revision 2005/2/28
3: %
4: %
5: %\documentclass[twocolumn,showpacs,preprintnumbers,amsmath,amssymb]{revtex4}
6: %\documentclass[preprint,eqsecnum,aps,floatfix,amsmath,amssymb,showpacs]{revtex4}
7: \documentclass[preprint,showkeys,eqsecnum,aps,floatfix,amsmath,amssymb]{revtex4}
8: %\documentclass[preprint]{tMPH2e}
9: \usepackage{graphicx}% figure files
10: \usepackage{dcolumn}% align table columns on decimal point
11: \usepackage{bm}% bold math
12: \textheight 9.0 in
13: \textwidth 6.5 in
14: \topmargin -0.5 in
15: \evensidemargin 0.0 in
16: \oddsidemargin 0.0 in
17:
18: \begin{document}
19: \newcommand\lsim{\alt}
20: \newcommand\stress[1]{{\it#1\/}}
21: \newcommand\whatis[1]{(\ref{#1})}
22: \newcommand\ie{{\it i.e.\/}}
23: \newcommand\etal{{\it et al.\/}}
24: \def\PD#1#2{\frac{\partial#1}{\partial#2}}
25:
26: \newcommand\Abar{\bar{A}}
27: \newcommand\Kbar{\bar{K}}
28: \newcommand\Sbr{\Sigma_{\beta|\gamma}}
29: \newcommand\Sab{\Sigma_{\alpha|\beta}}
30: \newcommand\Sabr{\Sigma_{\alpha|\beta\gamma}}
31: \newcommand\Sar{\Sigma_{\alpha|\gamma}}
32: \newcommand\mt{\widetilde m}
33: \newcommand\htilde{\tilde h}
34: \newcommand\Sm{S^\pm_M(\mt)}
35:
36: \newcommand\celc{\hbox{\thinspace$^\circ\hbox{\rm C}$}}
37: \newcommand\stu{\hbox{\hbox{erg}/\hbox{cm}$^2$}}
38: \newcommand\kelvin{\hbox{K}}
39: \newcommand\gcm{\hbox{g}/\hbox{cm}^3}
40: \newcommand\Ao{\thinspace{\buildrel \kern0.15em\scriptstyle\circ \over A}}
41: \newcommand\gl{\buildrel > \over {_<}}
42: %\newcommand\lsim{\buildrel < \over \sim}
43: %\newcommand\gsim{\buildrel > \over \sim}
44: %\newcommand\alt{\lesssim}
45: %\newcommand\agt{\gtrsim}
46: \newcommand\rrho{\check \rho}
47: \newcommand\rh{\check h}
48: \newcommand\rK{\check K}
49:
50: \newcommand\Pexpt{P_{\rm expt}}
51: \newcommand\Qexpt{Q_{\rm expt}}
52:
53: \newcommand\emspace{\hbox to 1 em {}}
54:
55: %\preprint{APS/123-QED}
56: \title{Interfacial tensions near critical endpoints:\\
57: experimental checks of EdGF theory}
58:
59: \author{SHUN-YONG ZINN and MICHAEL E. FISHER*}
60: \affiliation{%
61: Institute for Physical Science and Technology,\\
62: University of Maryland, College Park, MD 20742, USA}%
63: \date{22 March 2005}
64:
65: \begin{abstract}
66: Predictions of the extended de Gennes-Fisher local-functional
67: theory for the universal scaling functions of interfacial tensions near
68: critical endpoints are compared with
69: experimental data. Various observations of the
70: binary mixture isobutyric acid $+$ water are correlated to facilitate
71: an analysis of the experiments of Nagarajan,
72: Webb and Widom who observed the vapor-liquid interfacial tension
73: as a function of \stress{both} temperature and density.
74: Antonow's rule is confirmed and, with the aid of previously studied
75: {\it universal amplitude ratios\/}, the
76: crucial analytic ``background'' contribution to the surface tension
77: near the endpoint is estimated.
78: The residual singular behavior thus uncovered
79: is consistent with the theoretical scaling predictions and confirms the
80: expected lack of symmetry in $(T-T_c)$. A searching test of theory,
81: however, demands more precise and extensive
82: experiments; furthermore, the analysis highlights, a previously noted but
83: surprising, three-fold discrepancy in the magnitude of the surface
84: tension of isobutyric acid $+$ water relative to other systems.
85: \end{abstract}
86:
87: \keywords{interfacial tensions; critical endpoints; local-functional theories; near-critical binary fluids; universal critical amplitude ratios}
88:
89: \maketitle
90:
91: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
92: %
93: % SEC
94: %
95: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
96: \section{Introduction}
97: \noindent
98: On passing through a critical endpoint, a binary liquid
99: mixture of, say, species B and C, in
100: the presence of the vapor~$\alpha$, exhibits phase separation
101: below~$T_c$ into
102: two phases~$\beta$ and~$\gamma$ while above $T_c$ there is only
103: one homogeneous liquid
104: phase~$\beta\gamma$~\cite{Rowlinson, Widom77, Widom80, NWW}.
105: The~$\gamma$ phase, rich in C, may be supposed to have the
106: higher density so that in a gravitational field it lies below the~$\beta$
107: phase. In fact
108: there are binary mixtures which display lower consolute points so that
109: phase separation occurs
110: above~$T_c$ and mixing below~$T_c$; but since there are found to be
111: no basic differences in
112: criticality, only upper consolute points will be considered explicitly.
113:
114: Renormalization group theory has recently shown that the
115: critical behavior realized
116: at a critical endpoint lies in the same universality class
117: as on the critical
118: locus that results when the pressure is increased so that
119: the vapor phase $\alpha$ is fully suppressed~\cite{Diehl2000, Diehl2001}.
120: Nevertheless, near a critical
121: endpoint, new \stress{bulk} and \stress{interfacial} singularities arise
122: \cite{Fisher90a, Fisher90b}. These are of particular interest in the
123: interfacial tensions~$\Sbr$,
124: $\Sab$, $\Sar$, and $\Sabr$ corresponding to the various interfaces
125: indicated by the subscripts \cite{Widom77, Widom80, NWW, Fisher90a,
126: Fisher90b}. Indeed, except for the first, namely $\Sbr(T)$, all these tensions
127: are functions \stress{both} of the temperature, $T$, \stress{and} of the
128: overall density, $\rho$ (or, equally, of the overall composition).
129:
130: In terms of the standard notation, with
131: $t \equiv (T-T_c)/T_c$, the behavior of the \stress{critical}
132: \stress{surface}
133: \stress{tension} $\Sbr(T)$ that
134: \stress{vanishes} at the critical endpoint can be described by
135: \begin{equation}
136: \Sbr(T) \approx K |t|^\mu, \quad t \to 0-, \quad (h=0),
137: \label{eq5.1:sbr}
138: \end{equation}
139: where $K$ is a \textit{non}universal critical amplitude while
140: in $d \enspace (\le 4)$ spatial dimensions
141: the exponent satisfies the scaling relations
142: \begin{eqnarray}
143: \mu &=& 2 \beta + \gamma - \nu = 2 - \alpha - \nu, \nonumber\\
144: &=& \beta (\delta + 1) - \nu = (d - 1) \nu,
145: \label{eq5.1:exprel}
146: \end{eqnarray}
147: first derived, specifically for the surface tension, by Widom
148: \cite{Widom65, Widom72}. The symbol $h$ in parentheses in
149: \whatis{eq5.1:sbr} denotes the \stress{ordering} \stress{field}
150: [conjugate to the order parameter $\sim\negthinspace(\rho - \rho_c)$] that is
151: introduced in the scaling description \cite{Widom77, Widom80, Fisher90a,
152: Fisher90b, Zinn99, Zinn03}. It is an analytic combination of the
153: temperature $T$ and the chemical potentials, $\mu_B$ and $\mu_C$, or,
154: e.g., the chemical potential difference, $\Delta \mu = \mu_C - \mu_B$,
155: and the pressure, $p$, (all these being thermodynamic fields in the
156: standard sense); furthermore, $h$ is defined so that it vanishes \stress{at} the
157: \stress{critical} \stress{endpoint}, specified, say, by $(T_c, \rho_c,
158: p_c)$, and is identically zero on the whole $(\beta, \gamma)$
159: coexistence surface (or phase boundary) in the $(T, \mu_B, \mu_C)$
160: thermodynamic space: see, e.g., Fig.~1 of {\bf I} \cite{Zinn03}.
161:
162: The other surface tensions between the
163: noncritical vapor or \stress{spectator} \stress{phase} $\alpha$
164: and the liquid phases $\beta$, $\gamma$, and $\beta\gamma$ share a common
165: ``background'' contribution, $\Sigma_0(T, h)$, which
166: varies analytically and does \stress{not} vanish at
167: criticality. Thus we can write \cite{Fisher90a, Fisher90b}
168: \begin{eqnarray}
169: \Sabr(T) &\approx& K^+ |t|^\mu + \Sigma_0(T, 0), \quad t \to 0+,
170: \quad (h = 0),
171: \nonumber\\
172: \Sab(T) &\approx& K^- |t|^\mu + \Sigma_0(T, 0), \quad t \to 0-,
173: \quad (h = 0-),
174: \label{eq5.1:sabr}\label{eq5.1:sab}
175: \end{eqnarray}
176: where $K^+$ and $K^-$ are again nonuniversal amplitudes. When,
177: as generally expected \cite{Rowlinson}, the wetting temperature, $T_W$,
178: lies below $T_c$, the intermediate $\beta$ phase
179: spreads over the $\alpha | \gamma$ interface;
180: then Antonow's rule~\cite{Rowlinson, Zinn03} holds, that is,
181: \begin{equation}
182: \Sar(T) = \Sab(T) + \Sbr(T), \quad (h=0) \label{eq5.1:antonow}
183: \end{equation}
184: which by (\ref{eq5.1:sbr}) and (\ref{eq5.1:sab}) implies
185: \begin{equation}
186: \Sar(T) \approx (K + K^-) |t|^\mu + \Sigma_0(T, 0), \quad t\to0-, \quad h =
187: {0+}.
188: \label{eq5.1:sar}
189: \end{equation}
190:
191: More generally, on the whole thermodynamic surface bounding the $\alpha$ phase,
192: we have
193: \begin{equation}
194: \Sigma(T, h) = \Delta\Sigma(T, h) + \Sigma_0(T, h),
195: \label{eq5.1:sigma}
196: \end{equation}
197: where $\Delta\Sigma(T,h)$ and $\Sigma_0(T, h)$ are the singular
198: and regular (or analytic) parts of the surface tension, respectively.
199: Above $T_c$ one has $\Sigma = \Sabr$, while below $T_c$ one has
200: $\Sigma = \Sar$ when, by convention, $h > 0$
201: but $\Sigma = \Sab$ when $h < 0$.
202:
203: According to general scaling principles
204: the singular part can be written asymptotically as
205: \begin{equation}
206: \Delta\Sigma(T,h) \approx K|t|^\mu S_M^\pm[\mt(T, h)],
207: \quad\quad\quad\quad \mt\equiv M(T,h)/B|t|^\beta,
208: \label{eq5.1:Spm}
209: \end{equation}
210: where the superscripts $+$ and $-$ stand for $t \gl 0$, respectively:
211: see {\bf I}.
212: Because in experiments the density
213: is more readily accessible than the chemical potentials the argument
214: $\mt$ of the
215: scaling functions, $S_M^\pm(\bullet)$, has been chosen
216: proportional to the order
217: parameter which, in leading order, may be taken as $M \equiv (\rho -
218: \rho_c)$. The
219: coefficient $B$ represents the coexistence curve amplitude according
220: to $M_0 \approx B|t|^\beta$. The nonuniversal amplitudes~$B$ and~$K$ are
221: introduced in \whatis{eq5.1:Spm} to make the scaling functions
222: $S_M^\pm(\bullet)$ and their arguments
223: dimensionless. Then the $S_M^\pm(\bullet)$ are expected to be
224: \stress{universal} with,
225: by virtue of Antonow's rule~\whatis{eq5.1:antonow},
226: \begin{equation}
227: S_M^-(1) - S_M^-(-1) = 1. \label{eq5.1:SMpmnorm}
228: \end{equation}
229:
230: Owing to the technical and conceptual difficulties arising from the fluctuations
231: of capillary waves, the renormalization group approach has not so far
232: been successful
233: in calculating the $\Sm$. Instead, advances have been made with the aid
234: of phenomenological theories
235: based on local-functional concepts. The pioneering studies, both
236: theoretical and experimental, have been made by
237: Widom and his co-workers~\cite{Widom77,
238: Widom80, NWW}. However, as noted in \cite{Fisher90a,
239: Fisher90b}, their theory predicts a ``correction term'' varying as
240: $|t|^\gamma$ which becomes
241: more singular than the leading $|t|^\mu$ term when $d > 3 - \eta$
242: as is relevant in real three-dimensional systems. This
243: defect was remedied in the EdGF or ``extended de Gennes-Fisher''
244: theory proposed in~\cite{Fisher90a, Fisher90b} and implemented recently
245: in {\bf I} \cite{Zinn03} which forms the basis for the present report.
246: By using an accurate representation of the
247: equation of state near criticality~\cite{Zinn99},
248: the EdGF scaling functions~$\Sm$ have been calculated explicitly
249: and presented in {\bf I}.
250: On evaluating the scaling functions
251: in zero field, one obtains estimates for the Fisher-Upton
252: ratios \cite{Fisher90a, Fisher90b}, namely,
253: \begin{equation}
254: P \equiv (K^+ + K^-)/K = 0.137_5\pm0.002,
255: \quad Q \equiv K^+/K^- = -0.834\pm0.002. \label{eq5.1:PQ}
256: \end{equation}
257:
258: These ratios have been studied experimentally by
259: Woermann and coworkers~\cite{aKreuser, Mainzer} for the binary fluid
260: 2,6-dimethyl pyridine $+$ water: their
261: results are discussed in Sec.~II below.
262:
263: To go further and test the basic theoretical predictions for
264: the interfacial scaling functions $S_M^\pm(\mt)$ \stress{away}
265: from the $h = 0$ axis, the notable experimental data of
266: Nagarajan, Webb, and Widom ({\bf NWW})~\cite{NWW} for mixtures of
267: isobutyric acid and water offer, to our knowledge, the only
268: available opportunity.
269: For the same mixture, Howland,
270: Wong, and Knobler ({\bf HWK})~\cite{Knobler} measured the critical
271: surface tension
272: $\Sbr(T)$ while Greer~\cite{Greer} measured the densities on the coexistence curve.
273: These two experiments provide what prove to be valuable
274: consistency checks and calibrations for the {\bf NWW}
275: data as will be seen below.
276:
277: It may be noted that there
278: are interesting surface-tension experiments on other
279: quasi-binary and binary mixtures, such as $n$-octadecane and $n$-nonadecane
280: in ethane~\cite{Pegg85},
281: 2,5-lutidine in water~\cite{Privat}, and 2-butoxyethanol in
282: water~\cite{Woerman95}. However, our attention will focus on the
283: isobutyric acid $+$ water system since only in this case are there
284: sufficient measurements of the surface tensions
285: near the critical endpoint to warrant an attempt to extract the
286: interfacial scaling functions.
287:
288: Since the surface tension vanishes at the critical point
289: as $|t|^\mu$ with $\mu \simeq 1.26$, and hence faster than linearly,
290: the background contribution to the surface tensions
291: proves highly
292: significant as anticipated in {\bf I}. Thus, in analyzing the
293: experimental data, it is essential to determine the background
294: $\Sigma_0(T,h)$ carefully. By virtue of the analyticity of
295: the background, we may suppose it can be well represented
296: near the critical endpoint,
297: $(t,h)=(0,0)$, by an expansion of the form
298: \begin{equation}
299: \Sigma_0(T, h) = \Sigma_c + \Sigma_t t + \Sigma_h h +
300: \Sigma_{tt} t^2 + \Sigma_{th} t h + \Sigma_{hh} h^2 + \cdots.
301: \label{eq5.1:S0}
302: \end{equation}
303: Examination of the {\bf NWW} experimental data (see their Fig.~8)
304: reveals a
305: definite upward curvature in the plots of $\Sigma$ vs.\ $T$ at
306: fixed composition; but the theoretical results for the singular part
307: $\Delta \Sigma (T,0)$
308: \stress{alone} shows the opposite, downward curvature above $T_c$: see
309: {\bf I} Fig.~5(a) and the negative value of $Q$ in \whatis{eq5.1:PQ}.
310: To capture this dominant behavior, the
311: expansion~\whatis{eq5.1:S0} must contain at least quadratic terms in $t$.
312:
313: In the analysis reported by {\bf NWW} of their own data,
314: the background was assumed to be representable by the form
315: \begin{equation}
316: \Sigma_b(T, M) \simeq \Sigma_c + \Sigma_t t + \Sigma_{tM}\thinspace t M.
317: \label{eq5.1:NWWS0}
318: \end{equation}
319: However, this expression is \stress{not} smooth in $t$ and $h$, as expected
320: on general grounds,
321: since on the critical
322: isotherm it implies $(\partial \Sigma_b / \partial t) \simeq \Sigma_t +
323: \Sigma_{tM} M$ while $M$ varies as $\hbox{sgn}(h) |h|^{1/\delta}$
324: with $1/\delta \simeq 0.21$ which is highly singular. Furthermore, the
325: {\bf NWW} background assumption
326: seems to have led to the conclusion that $S_M^+(\mt)$ \stress{above} $T_c$
327: was indistinguishable from $S_M^-(\mt)$ \stress{below} $T_c$: see {\bf
328: NWW} Fig.~10.
329: However, when a suitable background of the form \whatis{eq5.1:S0} is used
330: in the analysis, one sees
331: that the data support a difference between $S_M^+(\mt)$ and $S_M^-(\mt)$
332: as, indeed, predicted theoretically in {\bf I} and expected quite
333: generally: see Sec.~V below.
334:
335: \begin{center}
336: {\bf Units and Exponents}
337: \end{center}
338:
339: \noindent
340: For convenience of reference we have set out in Table~I the exponent
341: values adopted in the analyses reported below. They correspond, of
342: course, to those for the three-dimensional Ising universality
343: class~\cite{Zinn98, Butera2000, Pelissetto2002, Guida1998,
344: Campostrini2002}.
345:
346: Since extensive experimental data will be discussed it is useful to
347: specify here the units employed. Thus: (i) mass densities will be
348: measured in $\gcm$ and denoted \stress{in} \stress{these} \stress{units}
349: as $\rrho \equiv \rho / (\gcm)$; (ii) for surface tensions,
350: $\Sigma$, and the corresponding amplitudes, $K$, etc., units of $\stu$
351: will be employed and indicated by $\rK$, etc.; (iii) since the reduced
352: free energy density will be taken in $\hbox{cm}^{-3}$ units, appropriate
353: units for the field $h$, conjugate to the order parameter $M = (\rho -
354: \rho_c)$, are $\hbox{g}^{-1}$ indicated correspondingly by $\rh$. The
355: units for various critical amplitudes, $B$, $C^\pm$, and the
356: coefficients $\Sigma_{th}$ in \whatis{eq5.1:S0}, etc. follow similarly.
357:
358: \begin{center}
359: {\bf Outline}
360: \end{center}
361:
362: \noindent
363: The balance of this article is set out as follows. In Sec.~II,
364: the theoretical predictions
365: \whatis{eq5.1:PQ} for the universal ratios $P$ and $Q$ are
366: discussed in the light of the experiments of
367: Woermann and
368: coworkers~\cite{aKreuser, Mainzer}.
369: The density data of Greer~\cite{Greer} along the coexistence curve
370: of isobutyric acid $+$ water
371: and the corresponding critical surface tension data of {\bf HWK}~\cite{Knobler}
372: are analyzed in Sec.~III to determine the critical amplitudes $B$ and
373: $K$. These results are used
374: to calibrate the {\bf NWW} data which are much more extensive but not as precise and,
375: thus, harder to analyze with confidence on their own.
376: However, the {\bf NWW} observations are found to be fully consistent
377: with the earlier measurements and, furthermore, confirm the validity of
378: Antonow's law in the temperature range studied: see Fig.~2 below.
379:
380: On this basis the analysis can be carried forward to determine the
381: crucial background term, $\Sigma_0(t,h)$. The results along the
382: coexistence curve and on the critical isochore above $T_c$ (\ie, for
383: zero field, $h \negthinspace = \negthinspace 0$) are reported in
384: Sec.~IV.A and displayed in Fig.~3. To estimate $\Sigma_0(t,h)$ on the
385: critical isotherm, $T = T_c$ $(t=0)$ for nonzero $h$, it is necessary to
386: re-express the {\bf NWW} observations at constant density in terms of
387: the field. For this purpose one needs information regarding the bulk
388: equation of state that goes beyond the value of $B$. In the absence of
389: direct bulk measurement on the critical isotherm, one can progress by
390: appealing to bulk {\it universal amplitude ratios\/} (known numerically
391: with appreciable reliability~\cite{Zinn98, Butera2000, Pelissetto2002}),
392: combined with the further universal critical ratio $S^+$ that relates
393: the surface tension amplitude $K$ to the correlation length
394: \cite{Zinn98}: see (A.1) in the Appendix where the relevant theory is
395: reviewed and its application explained.
396:
397: By this route one can estimate the amplitude $C^+$ for the ordering
398: susceptibility and thence, as explained in Secs.~IV.B, IV.C, and the
399: Appendix, the necessary conversions can be implemented. Thereby the
400: field-dependence of the background may be gauged. The singular part,
401: $\Delta \Sigma(t,h)$, then follows by subtraction although,
402: unfortunately, with rather limited precision that precludes, for
403: example, any direct determination of the surface tension amplitudes,
404: $K^c_>$ and $K^c_<$, \stress{on} the critical isotherms. Nevertheless,
405: in Sec.~V, the theoretically predicted universal scaling functions
406: are compared in suitable plots with the
407: experimental surface tension data of {\bf NWW}.
408: To the extent that consistency is well established,
409: the results are encouraging: however, significantly
410: more precise and extensive data close to criticality will be required to
411: enable sharper tests of the theory.
412:
413: The analysis, furthermore, encounters anew a puzzling dilemma concerning
414: the degree to which the dimensionless surface-tension ratio, $S^+$, for
415: the isobutyric acid $+$ water system actually conforms to the expected
416: value in light of independent observations of the critical scattering
417: and the values found for other systems. The issue is brought up in
418: Sec.~IV.B and pursued in some detail in the Appendix although without
419: resolution. The considerations presented highlight the need for further
420: experiments and, perhaps, for further theoretical developments.
421:
422: The article is summarized briefly in Sec.~VI.
423:
424: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
425: %
426: % SEC
427: %
428: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
429: \section{Lutidine and Water: Universal Ratios $\boldsymbol P$ and
430: $\boldsymbol Q$}
431: \noindent
432: Predictions for the surface tension ratios $P$ and $Q$
433: defined in \whatis{eq5.1:PQ} have
434: been tested experimentally by Mainzer-Althof and Woermann
435: ({\bf MW})~\cite{Mainzer}.
436: A binary liquid mixture of 2,6-dimethyl pyridine
437: [2,6-(CH$_3$)$_2$(C$_5$H$_3$N)], which is also known as 2,6-lutidine, and water
438: was
439: prepared at the critical composition and used in the experiment.
440: The surface tensions $\Sab(T)$, $\Sar(T)$, and $\Sabr(T)$
441: were measured, for the reduced temperature range $-0.028 \lsim t \lsim 0.019$, by
442: the inverted pendant drop (or rising bubble) method: analysis of
443: the contour line of the bubbles gives the surface tension. Also, the mass
444: densities of the
445: liquid phases $\rho_\beta(T)$, $\rho_\gamma(T)$, and $\rho_{\beta\gamma}(T)$
446: were measured. For the critical surface tension, {\bf MW} re-analyzed the
447: experimental data of Kreuser and Woermann~\cite{aKreuser} using the new
448: {\bf MW} density data.
449:
450: In the {\bf MW} analysis of the surface tension data, the background
451: \whatis{eq5.1:S0} was employed truncated at the linear term, $\Sigma_t
452: \thinspace t$. At first {\bf MW} fitted the data to \whatis{eq5.1:sabr} and
453: \whatis{eq5.1:sar} by
454: allowing separate background terms for $\Sab$, $\Sar$, and $\Sabr$ (\ie, different
455: sets of $\Sigma_c$ and $\Sigma_t$) and varying
456: $\Sigma_c$, $\Sigma_t$, $K^\pm$, and $(K+K^-)$.
457: For the exponent $\mu$ they assumed the value $1.26$.
458: The results did not agree
459: with the original Fisher-Upton ({\bf FU}) values
460: $P \simeq 0.1_2$ and $Q \simeq -0.83$ \cite{Fisher90a, Fisher90b}.
461: Furthermore, it appeared that Antonow's rule was \stress{not}
462: obeyed. However, when the analysis was performed with, as demanded
463: theoretically,
464: a \stress{common} background, $\Sigma_0(T)$,
465: Antonow's rule was found to be satisfied (as expected): this procedure yielded
466: \begin{equation}
467: \Pexpt \simeq 0.00 \pm 0.08, \quad
468: \Qexpt \simeq -1.0 \pm 0.2.
469: \label{eq5.2:fit1}
470: \end{equation}
471: The $P$ ratio is evidently not quite consistent with the {\bf FU} value
472: but, significantly, the $Q$ value is consistent, both in sign \stress{and}
473: magnitude.
474:
475: Other fitting procedures were also explored by {\bf MW} by imposing the
476: theoretical value of \stress{either} $K^+/K \enspace [= PQ/(1+Q)]$ \stress{or}
477: $K^-/K \enspace [= P/(1+Q)]$ and using a common background. Averaging the
478: central values of two different {\bf MW} fits and extending error estimates
479: to fully cover the fits yield
480: \begin{equation}
481: \Pexpt \simeq 0.11^{+0.05}_{-0.08}, \quad
482: \Qexpt \simeq {-0.835}^{+0.06}_{-0.12}.
483: \label{eq5.2:fit2}
484: \end{equation}
485: These values are consistent both with \whatis{eq5.2:fit1} and, now, with the
486: original {\bf FU} estimates. Furthermore, the more recent improved EdGF estimates
487: for $P$ and $Q$ from {\bf I}, quoted above in
488: \whatis{eq5.1:PQ}, agree remarkably well with these albeit biased fits to the
489: experimental data.
490:
491: It is appropriate to mention some earlier work.
492: Pegg, Goh, Scott, and Knobler~\cite{Pegg85} studied quasi-binary mixtures of
493: $n$-octadecane and $n$-nonadecane in ethane and measured the surface tensions
494: through and near both the \stress{upper} and the \stress{lower}
495: critical endpoints that occur in
496: the vicinity of the \stress{hidden} \stress{tricritical}
497: \stress{point} that arises in this system.
498: Their data, although limited, certainly support Antonow's rule and, more
499: importantly for us, suggest strongly that the ratio $K^+/K$ is negative and of
500: order unity in accordance with the EdGF predictions. However, the
501: upper and lower endpoints are less than $0.2 \thinspace \celc$ apart and,
502: mainly, for that reason, the 8 data points for the surface tensions $\Sbr$
503: (and the 3 for $\Sar$) are not sufficient to warrant quantitative study.
504:
505: The surface tension of the water and 2,5-lutidine system has been measured by
506: Privat and co-workers~\cite{Privat} at the critical composition of the lower
507: critical consolute point as well as off that critical composition. The authors
508: claim to verify Antonow's rule. However, their quoted fits for the amplitudes
509: $K$, $K^+$, and $K^-$ seem problematical relative to the graphical
510: presentation of their data. Thus, and especially in the light of the
511: {\bf MW}
512: experiments on the closely related 2,6-lutidine system, this work cannot be
513: regarded as yielding useful experimental estimates for the amplitude
514: ratios~$P$ and~$Q$.
515:
516: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
517: %
518: % SEC
519: %
520: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
521: \section{Isobutyric Acid and Water: Nonuniversal Amplitudes}
522: %\vskip-\bigskipamount
523: \subsection{Coexistence Curve Amplitude}
524: \noindent
525: The coexistence curve of isobutyric acid $+$ water has been studied
526: carefully by Greer~\cite{Greer}, who measured the coexisting mass
527: densities, $\rho_\beta(T)$ and $\rho_\gamma(T)$, with a precision
528: of 20 ppm from 3.5 K below the critical temperature at intervals
529: as small as 5 mK (with a precision of $\pm 1$ mK) for a sample
530: prepared at close to the critical composition. Two runs were made
531: and the data are recorded in Appendix A of Ref.~\onlinecite{Greer}.
532: Greer's analysis demonstrated that an optimal choice of order parameter
533: (as judged in relation to the number density of a pure fluid in a
534: lattice-gas representation) was the volume fraction of one component,
535: say $\phi_B$. The surface tension experiments of {\bf NWW}, which are
536: of principal concern to us, however, also used the mass density in the
537: vicinity of the critical endpoint as the primary controlled observable.
538: Accordingly, we have reanalyzed Greer's data using the mass difference,
539: $\rho - \rho_c$, as the order parameter. This choice, of course,
540: affects the real (and apparent) magnitudes of the various correction
541: terms; but these will play no more than an auxiliary role in our
542: analysis. Further discussion of fits for the volume fraction
543: difference, $\Delta \phi$, are given in \cite{Zinn1997}: one learns that
544: a range of reasonable fits to the data serve to determine the critical
545: point to within $\pm 10$ to $15$ mK while the leading critical amplitude
546: $B$ can be found with a precision of no better than $\pm 1$ \%.
547:
548: More specifically, we have fitted the data to
549: \begin{equation}
550: \hbox{$1\over2$} \Delta \rho
551: \equiv \hbox{$1\over2$} \left [ \rho_\gamma(T) - \rho_\beta(T) \right ]
552: = B |t|^\beta \thinspace
553: [ 1 + b_\theta |t|^\theta + b_1 t], \label{eq5.3:form}
554: \end{equation}
555: with $t = (T-T_c)/T_c$ as usual, but, as a useful crosscheck, also
556: in terms of the asymptotically equivalent variable,
557: \begin{equation}
558: t' = 1 - T_c/T = t / (1+t), \label{eq5.3:tp}
559: \end{equation}
560: as in a previous study \cite{Zinn}. Using primes to denote the
561: amplitudes fitted with $t'$ we find, from the first run \cite{Zinn1997},
562: \begin{eqnarray}
563: &&T_c \simeq 25.996_7\celc, \quad \check B \simeq \check B' \simeq
564: 0.03104, \nonumber\\
565: &&b_\theta \simeq b'_\theta \simeq -1.73, \quad
566: b_1 \simeq -9.783, \quad b'_1 \simeq -9.455.\label{eq5.3:Brho}
567: \end{eqnarray}
568: The difference $b'_1 - b_1 \simeq 0.328$ is close to $\beta$ as it
569: should be for consistency in light of \whatis{eq5.3:tp}.
570:
571: The exponent $\theta \simeq 0.5$ (See Table~I) in \whatis{eq5.3:form}
572: specifies the leading correction to scaling. General scaling
573: considerations \cite{Fisher2000, Kim1998} indicate that a further
574: correction term, $|t|^{2\beta}$, should also be present which, since
575: $2 \beta \simeq 0.65 < 1$, should in fact dominate the linear term. The
576: corresponding amplitude $b_{2\beta}$ may well be rather small:
577: nevertheless, the data cannot resolve two such terms so that the fitted
578: values of the coefficients $b_\theta$ and $b'_\theta$ must be regarded
579: as no more than \textit{effective} \textit{amplitudes}.
580:
581: On setting $b_\theta = b_1 = 0$ in \whatis{eq5.3:form} Greer found
582: $\check B = 0.0265 \pm 0.0005$; however, on allowing for $b_\theta \ne
583: 0$ the value $0.0315 \pm 0.0025$ resulted; this is fully consistent with
584: \whatis{eq5.3:Brho}.
585: Repeating the same procedure for the second experimental run yields
586: \begin{eqnarray}
587: &&T_c \simeq 25.969_6 \celc, \quad \check B \simeq \check B' \simeq
588: 0.03175, \nonumber\\
589: &&b_\theta \simeq b'_\theta \simeq -2.635, \quad
590: b_1 \simeq -14.563, \quad b'_1 \simeq -14.234.\label{eq5.3:Brho2}
591: \end{eqnarray}
592: The difference $b'_1 - b_1 \simeq 0.329$ is again close to $\beta$.
593: Clearly, the deviations of the fitted values here from those in
594: \whatis{eq5.3:Brho} provide a measure of the accuracy available in
595: estimating $T_c$ and $B$.
596:
597: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
598: %
599: % SUBSEC
600: %
601: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
602: \subsection{Comparison with NWW Coexistence Curve Data}
603: \noindent
604: We may now use the fits to \whatis{eq5.3:form} to calibrate the
605: experiments
606: on the same system by {\bf NWW} who prepared mixtures of
607: isobutyric acid $+$ water at various compositions and determined
608: the surface tensions
609: by optically measuring the wavelength of surface waves generated
610: by a transducer. In Table~1 of {\bf NWW}, about 140 $(T, \rho, \Sigma)$
611: data
612: points are presented: these are what we study here. We have
613: extrapolated some of the observations to the coexistence curve;
614: but, as indicated briefly in Sec.~IV.A, only very small
615: changes in the values of $\rho$ and $\Sigma$ are entailed.
616:
617: To proceed, we first consider the diameter of the coexistence curve in the
618: density variable, namely,
619: \begin{equation}
620: {\overline \rho}(T) = \hbox{$1\over2$} (\rho_\beta + \rho_\gamma),
621: \label{eq5.3:diameter}
622: \end{equation}
623: since, as mentioned, the density was directly observed by {\bf NWW}.
624: Owing to the mixing of thermodynamic fields near criticality in fluids,
625: one expects a dominant singular term $|t|^{1-\alpha}$ to appear in the
626: diameter~\cite{Fisher2000, Kim1998, Widom70, Rehr71, Fisher75}.
627: However, the {\bf NWW} data do not reveal a signature of any such singular term.
628: Rather, the coexistence curve appears almost symmetric in the
629: $(\rho, T)$ plane \cite{GreerDiameter}. Thus the diameter
630: may be well represented by a constant as
631: \begin{equation}
632: {\check {\overline \rho}}(T) \simeq \check \rho_c \simeq 0.9936.
633: \label{eq5.3:rhoc}
634: \end{equation}
635:
636: As discussed by Greer \cite{Greer}, the critical temperature of the
637: isobutyric acid $+$ water system is particularly sensitive to ionic
638: impurities \cite{Kim2001} and, indeed,
639: experience shows that the observed critical temperatures of
640: nominally the same binary
641: fluid mixtures typically vary from experiment to
642: experiment by amounts exceeding the stated errors. For
643: fitting the {\bf NWW} data, therefore, we have adopted
644: their value of $T_c = 26.310 \pm 0.001 \celc$,
645: even though it
646: lies outside the uncertainty limits implied by \whatis{eq5.3:Brho}
647: and \whatis{eq5.3:Brho2}.
648: As seen in Fig.~\ref{fig5:coex}, by using \whatis{eq5.3:form} and
649: \whatis{eq5.3:rhoc} with the amplitude estimate
650: \begin{equation}
651: \check B = 0.0314 \pm 0.0004,\label{eq5.3:amplB}
652: \end{equation}
653: the {\bf NWW} data can be represented
654: rather satisfactorily. This demonstrates the consistency of their
655: observations with other careful studies and establishes a reliable
656: estimate for the amplitude $B$ which will be employed below.
657:
658: It may be remarked that {\bf NWW} advocated the use in theoretical analysis
659: of a
660: (conventionally defined) volume fraction, $\phi$, in place of
661: the density, $\rho$ \cite{Phi}. Study of their data reveals, however, that the
662: dependence of $\phi$ on $\rho$ is both surprisingly nonlinear and
663: significantly temperature
664: dependent. Furthermore, contrary to the expectations of {\bf NWW}
665: (see p.~5781) and Greer, the coexistence curve appears
666: to be more symmetric and
667: regularly behaved in terms of $\rho$ than of $\phi$. Accordingly, we have
668: accepted the directly observed variable $\rho$
669: as the order parameter in all the following analysis.
670:
671: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
672: %
673: % SUBSEC
674: %
675: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
676: \subsection{Critical Surface-Tension Amplitude $\boldsymbol K$}
677: \noindent
678: In order to obtain the critical amplitude $K$ as defined for the surface
679: tension in \whatis{eq5.1:sbr},
680: the capillary-rise data obtained by {\bf HWK} \cite{Knobler}
681: for isobutyric acid $+$ water will be employed.
682: In their Fig.~8, one can clearly
683: see one point that deviates significantly from the general trends, namely,
684: that corresponding in {\bf HWK} Table~VII to $T_c - T = 1.088$ K or
685: $t \simeq -0.0036$, and $\check \Sigma_{\beta|\gamma} = 0.0191$.
686: This point has thus been omitted in the present analysis.
687:
688: Since only eight data points are available, the correction-to-scaling terms
689: are not expected to play a detectable role. First, therefore,
690: the data have been fitted to the form
691: \begin{equation}
692: \Sigma_{\beta|\gamma} = \Sigma_c + K |t|^\mu,
693: \label{5.3:sbr}
694: \end{equation}
695: where, of course, $\Sigma_c \equiv 0$ is anticipated.
696: The reported value of
697: $T_c$ is $299\thinspace\kelvin$ \cite{Knobler}; but we have not found
698: in the account of {\bf HWK} a satisfying description as to how this
699: value was determined. We therefore adopted
700: \begin{equation}
701: T_c = 299.00 \thinspace \kelvin = 25.85 \thinspace \celc,
702: \label{eq5.3:KnoblerTc}
703: \end{equation}
704: provisionally but have also examined how estimates for $K$ are affected by
705: changes in the assigned value of $T_c$.
706:
707: A weighted least-squares fit then yields
708: \begin{equation}
709: \check \Sigma_c = -0.002 \pm 0.008, \quad \check K = 34.6 \pm 5.7,
710: \label{eq5.3:roughK}
711: \end{equation}
712: where the uncertainties quoted here and below are twice the standard deviation.
713: By imposing $\Sigma_c = 0$, which is clearly consistent with the data, one finds
714: \begin{equation}
715: \check K = 32.8 \pm 1.7, \label{eq5.3:sc0}
716: \end{equation}
717: where the central estimate has shifted by 5\% while the uncertainty is reduced
718: by 30\% in comparison with \whatis{eq5.3:roughK}. Repeating the analysis
719: with $\Sigma_c = 0$ but using $t' = t/(1+t)$ as a variable \cite{Zinn}, we find
720: $\check K = 32.6 \pm 1.7$ in good agreement.
721:
722: As a further check one can calculate $\Sigma_{\beta|\gamma} /
723: |t|^\mu$ from each data point. This ratio approaches $K$ when $t\to 0$,
724: and an average yields $\check K \simeq 32.7$, suggesting that the corrections
725: to the pure power law are relatively small.
726:
727: To gauge the sensitivity to the imposed value of $T_c$, we changed the value
728: by $\pm 0.1$ K. Considering that the temperature was controlled to within
729: $0.01$ K \cite{Knobler}, this variation is fairly large. The results
730: (obtained by fixing $\Sigma_c = 0$) are $\check K = 30.6 \pm 1.7$ for
731: $T_c = 25.95 \thinspace \celc$ and $\check K = 35.3 \pm 1.8$ for
732: $T_c = 25.75 \thinspace \celc$.
733: Despite the changes of $\pm 7$ \%, the overlap of the
734: error bars allows consistency with the previous
735: estimates for $K$.
736:
737: Finally, the data were fitted to $\Sigma = K \thinspace |(T/T_c) - 1|^\mu$
738: treating both $T_c$ and $K$ as parameters. Minimizing $\chi^2$ yielded
739: $\check K \simeq 33.8_5$ and $T_c \simeq 25.78 \thinspace \celc$.
740: This value for $T_c$ is $0.07$ K lower than the asserted value
741: \whatis{eq5.3:KnoblerTc} but
742: the amplitude estimate again lies in the indicated range.
743: Repeating the procedure with $\Sigma = K' \thinspace |1 - (T_c/T)|^\mu$,
744: yields the same value for $T_c$ but $\check K' \simeq 33.5$,
745: about 1\% smaller. Overall we believe that
746: \begin{equation}
747: \check K = 33.7 \pm 2.5, \label{eq5.3:Kexp}
748: \end{equation}
749: summarizes the evidence conservatively \cite{Kprob}, although a more precise
750: knowledge of $T_c$ in the {\bf HWK} experiments could be valuable.
751:
752: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
753: %
754: % SUBSEC
755: %
756: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
757: \subsection{Comparison with NWW Critical Surface-Tension Data}
758: \noindent
759: It is informative to compare the experiments performed by {\bf HWK}
760: with those of {\bf NWW} to check, in particular, for
761: mutual consistency: see Fig.~\ref{fig5:chk}. For convenience
762: {\bf NWW} data points on
763: the coexistence curve have been numbered consecutively from~1 to~15.
764: Except for point~8, other points joined by parallelograms were measured at the
765: same temperature. Points~1--7 lie on the high density side of the
766: coexistence curve and hence represent $\Sar(T,
767: h\negthinspace=\negthinspace0+)$, while points~9--15 correspondingly lie on
768: the low density side and represent $\Sab(T, h\negthinspace=\negthinspace0-)$.
769:
770: To explain the significance of the fixed-$T$ parallelograms,
771: consider, as an example,
772: points~2 and~14. By Antonow's rule we have $\Sar = \Sab + \Sbr$
773: and so by using
774: $\Sbr(T) \approx K|t|^\mu$ with the estimate \whatis{eq5.3:Kexp}
775: and by varying $\Sab$ within the uncertainty limits, one can find
776: a range of $\Sar$ values that are consistent with Antonow's rule.
777: Evidently,
778: the central value of the data point~14 corresponds to the lower limit of the
779: data point~2, while the upper limit corresponds to the central value of data
780: point~2. These mutually consistent limiting pairs of
781: the data points~2 and~14 have been joined by parallel lines.
782: In other words, the vertical sides of the
783: parallelogram connecting data points~2 and~14 represent error limits on
784: the {\bf NWW} data that are consistent with
785: the {\bf HWK} experiment and with Antonow's rule. Thus we notice that
786: data point~3 represents a high estimate of $\Sab$ while
787: data point~13 is relatively low. Despite this (3,13) worst case,
788: it can be concluded that the two experiments are mutually consistent and that
789: the estimate~$K$ in~\whatis{eq5.3:Kexp} is reliable \cite{Kprob} .
790: Note that it would have
791: been very difficult to obtain a reliable value of $K$ from the {\bf NWW} data alone,
792: owing to the necessity of extrapolating and differencing their (not highly
793: precise) data to obtain $\Sbr$ via Antonow's rule.
794:
795: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
796: %
797: % SUBSEC
798: %
799: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
800: \subsection{Other Surface Tension Amplitudes}
801: \noindent
802: Having determined the two amplitudes $K$ and $B$, which serve as
803: metrical factors in the scaling formulation of the critical-endpoint
804: behavior of the surface tension [see \whatis{eq5.1:Spm}], we are
805: prepared to take the next step towards analyzing the {\bf NWW} data.
806: The first issue must be to establish the background contribution,
807: $\Sigma_0(T,h)$, by estimating the coefficients $\Sigma_c$, $\Sigma_t$,
808: etc., in \whatis{eq5.1:S0}. To that end it proves helpful to have
809: at hand values for the surface tension amplitudes $K^+$ and $K^-$ as
810: defined in \whatis{eq5.1:sabr}. Ideally, of course, these should be
811: unambiguously determined by the data themselves: but, as already seen in
812: Sec.~II, it is unrealistic at this stage in the development of
813: experimental techniques to expect to do more than
814: verify---optimistically at a fully convincing
815: level---\stress{consistency} of the theory with the observational data.
816: Accordingly, we report here the values
817: \begin{equation}
818: \check K^+ = -23.3 \pm 1.8, \quad \check K^- = 27.9 \pm 2.1,
819: \label{eq5.3:Kpm}
820: \end{equation}
821: that follow from \whatis{eq5.3:Kexp} on accepting the EdGF calculations for
822: the universal ratios $P$ and $Q$ given in \whatis{eq5.1:PQ}.
823:
824: To determine the field-dependent coefficients $\Sigma_h$ and
825: $\Sigma_{hh}$ in the background expansion \whatis{eq5.1:S0} we will need
826: to examine the data for $\Sigma(t,h)$ on the critical isotherm. The
827: expected behavior is
828: \begin{equation}
829: \Sigma_c(h) \approx K^c_{\gl} |h|^{\mu/\Delta} + \Sigma_c + \Sigma_h h +
830: \Sigma_{hh} h^2, \qquad (T = T_c)
831: \label{eq5.3:TcS0}
832: \end{equation}
833: where, again, values for the amplitudes $\check K^c_>$ and $\check
834: K^c_<$, for $h \gl 0$, respectively, will prove of importance. To
835: estimate these one may appeal to the EdGF theoretical values reported in
836: {\bf I} (6.9) for the universal ratios
837: \begin{equation}
838: S^c_{\gl} \equiv {K^c_{\gl} \over K} \left ( {B \over C^+} \right )
839: ^{\mu / \Delta}.
840: \label{eq5.3:Scgl}
841: \end{equation}
842: In the second factor in this expression, which enters because $K^c_{\gl}$
843: and $K$ have different dimensions, the nonuniversal amplitude $C^+$
844: specifies the magnitude above $T_c$ of the divergence of the basic
845: ordering susceptibility as $1/t^\gamma$. Note that the notation for
846: amplitudes used here and below accords with that set out in the Appendix
847: of \cite{Zinn98}.
848:
849: The appropriate value of $C^+$, and of other nonuniversal amplitudes for
850: isobutyric acid $+$ water, is taken up in Sec.~IV.B when the
851: field-dependence is considered explicitly. At that stage, we will make
852: contact with Moldover's extensive analysis \cite{Moldover85, Moldover86}
853: of surface tension, light scattering, and bulk thermodynamic experiments
854: for a range of single-component and binary fluid systems designed to
855: test the hyperscaling (or ``two-scale factor universality'') hypothesis
856: for criticality. Various details and some questions they raise are
857: presented in the Appendix.
858:
859: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
860: %
861: % SEC
862: %
863: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
864: \section{Determination of the Surface Tension Background}
865: \noindent
866: As discussed in Sec.~I, to analyze the {\bf NWW} surface
867: tension measurements appropriately,
868: the background~$\Sigma_0(T,h)$ must at least contain quadratic terms in $t$;
869: it then seems reasonable to include quadratic terms in $h$
870: also as in \whatis{eq5.1:S0}.
871: In order to determine the background
872: coefficients $\Sigma_c$, $\Sigma_t$, etc.,
873: we adopt the following strategy. First, the data on the
874: $h\negthinspace=\negthinspace0$
875: axis, \ie, for the mixture at the critical composition,
876: will be fitted to obtain
877: $\Sigma_c$, $\Sigma_t$, and $\Sigma_{tt}$.
878: From the data on the critical isotherm, $t=0$,
879: the coefficients $\Sigma_h$ and $\Sigma_{hh}$ can then be examined.
880: Finally, to obtain the cross-term
881: coefficient $\Sigma_{th}$, the derivative
882: $(\partial \Sigma / \partial t)_h$ may be studied
883: on the critical isotherm.
884:
885: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
886: %
887: % SUBSEC
888: %
889: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
890: \subsection{Along the Coexistence Curve}
891: \noindent
892: In zero-field, the surface tension above $T_c$ is,
893: following the Introduction, expected to behave as
894: \begin{equation}
895: \Sabr(T) \approx K^+ |t|^\mu + \Sigma_c + \Sigma_t t + \Sigma_{tt} t^2,
896: \qquad (T \ge T_c).
897: \label{eq5.4:sabr}
898: \end{equation}
899: Below $T_c$, there are two distinct vapor-liquid interfaces
900: and two surface tensions, $\Sab$ and $\Sar$, corresponding to the
901: coexisting phases $\beta$ and $\gamma$.
902: By Antonow's rule (which as shown is
903: satisfied by the {\bf NWW} data) the difference $(\Sar - \Sab)$
904: is the critical surface tension $\Sbr$ and thus it suffices to study
905: only $\Sab(T)$ below $T_c$.
906: The identical expression \whatis{eq5.4:sabr} then applies except that
907: $K^+$ is to be replaced by $K^-$.
908:
909: To proceed, one may first fit $\Sabr(T)$ using
910: the data above $T_c$ at the very close-to-critical composition
911: tabulated by {\bf NWW}, by regarding
912: $K^+$, $\Sigma_c$, $\Sigma_t$, and $\Sigma_{tt}$ as free parameters.
913: These fits suggest $K^+ > 0$ which, in accord with the generally observed
914: upwards curvature (compare with {\bf I} Fig.~5),
915: has the \textit{opposite} sign to the prediction for $K^+$ in
916: \whatis{eq5.3:Kpm}.
917: We conclude that the {\bf NWW} data above $T_c$ are not, on their own,
918: of adequate
919: precision to reliably distinguish the amplitudes of three terms varying as $t$,
920: $t^\mu$, and $t^2$; hence, we cannot independently estimate $K^+$.
921:
922: However,
923: as indicated in the {\bf MW} analysis \cite{Mainzer} discussed in
924: Sec.~II, the situation may
925: be improved by imposing the theoretical value of the ratio $K^+/K$;
926: but this is then
927: equivalent to imposing $K^+$ because, as seen in Sec.~III,
928: the critical amplitude $K$ has been already estimated reliably
929: by comparison with
930: the {\bf HWK} data. Since $\Sabr(T)$ and $\Sab(T)$ share the
931: common background $\Sigma_0(T,0)$, we likewise impose $K^-$ as given
932: in \whatis{eq5.3:Kpm}.
933: Of course, by adopting this procedure we will be primarily checking the
934: \stress{consistency} of our theory with the {\bf NWW} data which,
935: unfortunately, are not
936: adequate for estimating $K^+$ and $K^-$ directly as might originally
937: have been hoped.
938:
939: Accordingly,
940: by subtracting the singular term in \whatis{eq5.4:sabr} and likewise
941: below $T_c$,
942: we can generate experimental data for the zero-field background
943: alone: see Fig.~\ref{fig5:sigma0}.
944: Again we use the {\bf NWW} data taken at the (close-to) critical composition above
945: $T_c$ and the (close-to) low-density side of the coexistence curve.
946: A least-squares fit then yields
947: \begin{equation}
948: \check \Sigma_c \simeq 26.284, \quad
949: \check \Sigma_t \simeq -14.6, \quad
950: \check \Sigma_{tt} \simeq 381.
951: \label{eq5.4:h0fit}
952: \end{equation}
953: The corresponding plot is represented by the dashed curve
954: in Fig.~\ref{fig5:sigma0} and
955: seen to be very good. This demonstrates
956: that these coefficient estimates are most reasonable although, as will be
957: discussed, they prove to be less than fully satisfactory in the near
958: critical region below~$T_c$.
959:
960: To examine the fit further
961: consider the plots shown in Fig.~\ref{fig5:chk}.
962: The dashed curve represents the surface tensions $\Sab(T)$ and $\Sar(T)$
963: on the coexistence
964: curve as reproduced from the amplitudes $K$ and $K^-$ and the background
965: coefficients \whatis{eq5.4:h0fit}: to translate $t \propto T_c - T$ into
966: the density on the coexistence curve, the inner fit described in the
967: caption of Fig.~\ref{fig5:coex} is used. The agreement with the central
968: values of the {\bf NWW} data on the low-density side
969: is encouraging;
970: however, the fit reveals significant, systematic discrepancies on the
971: high density side where the plot lies outside the observational range for all
972: the points 1--7.
973:
974: To complement the overall or ``full-range''
975: approach just described and to understand the
976: discrepancies uncovered on the high-density side of the coexistence curve,
977: let us
978: return to Fig.~\ref{fig5:chk} and focus on the {\it data closest to
979: criticality\/}.
980: In particular, as explained in the construction of the parallelograms in this
981: figure, consistency with the {\bf HWK}
982: data for $\Sbr(T)$, together with Antonow's
983: rule~\cite{Fenzl} requires that the allowed observational uncertainties
984: quoted for the {\bf NWW}
985: data should be reduced to correspond to the vertical sides
986: of each parallelogram. Accordingly, for $T \lsim T_c$
987: consider the truncated forms
988: \begin{eqnarray}
989: \Sab(T) &\simeq& K^-|t|^\mu + \Sigma_c + \Sigma_t t, \label{eq5.4:linSab}\\
990: \Sar(T) &\simeq& (K^-+K) |t|^\mu + \Sigma_c + \Sigma_t t,\label{eq5.4:linS0}
991: \end{eqnarray}
992: in which the background has been restricted to a linear term.
993: By trial and error, one can discover those values of $\Sigma_c$ and
994: $\Sigma_t$ that give acceptable fits to those data points lying closest
995: to the critical point, specifically, for the pairs
996: (7,9), (6,10), (5,11), (4,12), and
997: (3,13). Slight shifts in the nominal value of $\rho_c$ from
998: \whatis{eq5.3:rhoc} may also be examined but are
999: found to be detrimental for changes
1000: exceeding $\Delta \check \rho_c = 10^{-4}$.
1001:
1002: The predominant conclusion is that $\Sigma_c$, the common value
1003: of the surface tensions $\Sab$, $\Sar$, and $\Sabr$ \stress{at} criticality,
1004: cannot actually be as low as suggested by the full-range
1005: fit \whatis{eq5.4:h0fit}.
1006: The acceptable ``critical fits'' yield, instead, the range
1007: \begin{equation}
1008: \check \Sigma_c = 26.294 \pm 0.004.
1009: \label{eq5.4:Sc}
1010: \end{equation}
1011: Indeed, {\bf NWW} actually quote the estimate
1012: $\check \Sigma_c \simeq 26.30$, which is
1013: much closer to this ``critical value'' than to \whatis{eq5.4:h0fit}. The
1014: corresponding
1015: preferred value for the background slope, is found to be
1016: $\check \Sigma_t = -17.5$; but reasonable fits could be obtained in the range
1017: $-16.5 > \check \Sigma_t > -18.5$. In all these fits,
1018: however, the fitting curve in Fig.~\ref{fig5:chk} became too broad at the larger values of $| \rho -
1019: \rho_c |$, failing to fit the data-point pairs (1,15) and (2,14). This last
1020: defect, however, was overcome by including in
1021: \whatis{eq5.4:linSab} and \whatis{eq5.4:linS0} the
1022: quadratic term $\Sigma_{tt} t^2$. As the assigned value $\Sigma_{tt}$ is
1023: increased, keeping $\Sigma_c$ fixed as proves essential, the optimal value for
1024: $\Sigma_t$ correspondingly falls. Thus $\check \Sigma_t = -16.0$ and
1025: $\check \Sigma_{tt} = 250$ give very good fits in
1026: Fig.~\ref{fig5:chk}. However, these particular
1027: values do not describe well the
1028: highest data points \textit{above} $T_c$ in Fig.~\ref{fig5:sigma0} (for $t >
1029: 0.015$). One could well disregard this range since the data extend
1030: below $T_c$ only down to $|t| \simeq 0.011$. However, one can find
1031: values that prove rather satisfactory in fitting \stress{all\/} the $h=0$
1032: data, \ie, on the coexistence curve and at the critical composition
1033: above $T_c$, namely,
1034: \begin{equation}
1035: \check \Sigma_c \simeq 26.294, \quad
1036: \check \Sigma_t \simeq -15.0, \quad
1037: \check \Sigma_{tt} \simeq 400.
1038: \label{eq5.4:h0fitb}
1039: \end{equation}
1040: These coefficients yield the solid curve shown in Fig.~\ref{fig5:sigma0}:
1041: evidently the fit displays a systematic displacement
1042: above the central values of the
1043: data points but by rather small amounts that remain \stress{within}
1044: all the uncertainty ranges.
1045: Furthermore, the agreement with the coexistence data
1046: is now much improved as seen by the solid curve in Fig.~\ref{fig5:chk}.
1047: Only very minor departures from the acceptable uncertainty ranges
1048: now arise. Accordingly we will
1049: retain the assignment \whatis{eq5.4:h0fitb} in all the following analysis.
1050:
1051: At this stage we may conclude that the {\bf NWW}
1052: data are, first, fully consistent
1053: with the best previous observations of the coexistence curve
1054: and the critical surface tension and, second,
1055: are quite consistent with the EdGF based estimates
1056: of the ratios $K^+/K$ and $K^-/K$ even though,
1057: unfortunately, the data cannot critically test these ratios.
1058:
1059: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1060: %
1061: % SUBSEC
1062: %
1063: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1064: \subsection{On the Critical Isotherm}
1065: \noindent
1066: It transpires, in retrospect, that {\bf NWW}
1067: did not make observations of the surface tensions
1068: very close to the critical
1069: isotherm $T=T_c = 26.310 \thinspace \celc$. We may, however, presume
1070: that the isotherms just
1071: above and below $T_c$, namely, at $T = 26.49 \thinspace \celc$ and
1072: $T = 26.05 \thinspace \celc$ constitute rough lower and upper bounds,
1073: respectively, on $\Sigma_c(h)$.
1074: On the critical isotherm, the surface tension
1075: should behave as in \whatis{eq5.3:TcS0} with $\mu / \Delta \simeq 0.806$.
1076: The optimal assignment of the critical point value $\Sigma_c$ has already
1077: been determined in the previous subsection. One might wish to determine
1078: the amplitudes $K^c_>$ and $K^c_<$ directly from the data; but,
1079: regrettably there is, once again, no hope of that! Rather, with the
1080: limited aim of testing only consistency with the theory, we will adopt
1081: theoretical estimates for these two amplitudes. At this point one
1082: encounters a rather sharp and surprising \stress{dilemma}: as explained
1083: further in the Appendix, two significantly different sets of EdGF-based
1084: predictions arise from \whatis{eq5.3:Scgl} and {\bf I} (6.9). The first
1085: set, derives \stress{entirely} from the two metrical factors $K$ and
1086: $B$, that we have determined for isobutyric acid $+$ water, combined
1087: with well-confirmed values for various {\it universal critical amplitude
1088: ratios\/} \cite{Zinn98}: as explained in the Appendix this yields
1089: \begin{equation}
1090: \rlap{\hskip -.1\displaywidth\hbox{{\bf A}:}}% \qquad\qquad
1091: \check K^c_> = (5.0 \pm 0.6) \negthinspace \times \negthinspace 10^{-18},
1092: \quad\quad\quad
1093: \check K^c_< = -(1.5_5 \pm 0.1_7) \negthinspace \times \negthinspace 10^{-18}.
1094: \label{eq5.4:KcglA}
1095: \end{equation}
1096:
1097: The second set also utilizes the estimate \whatis{eq5.3:Kexp} for $K$
1098: in \whatis{eq5.3:Scgl} together with \whatis{eq5.3:amplB} for $B$; but
1099: instead of using the purely theoretically derived value, namely,
1100: \begin{equation}
1101: \rlap{\hskip -.286\displaywidth\hbox{{\bf A}:}}% \qquad\qquad
1102: \check C^+ = (3.04 \pm 0.30) \negthinspace \times \negthinspace 10^{-26},
1103: \label{eq5.4:CpA}
1104: \end{equation}
1105: for the amplitude $C^+$ in \whatis{eq5.3:Scgl}, it employs the much larger
1106: estimate
1107: \begin{equation}
1108: \rlap{\hskip -.291\displaywidth\hbox{{\bf B}:}}% \qquad\qquad
1109: \check C^+ = (14.2 \pm 0.6) \negthinspace \times \negthinspace 10^{-26},
1110: \label{eq5.4:CpB}
1111: \end{equation}
1112: that follows, as explained in the Appendix, from {\it independent\/}
1113: light scattering and thermodynamic experimental data for isobutyric acid
1114: $+$ water \cite{Moldover85, Moldover86}. This then leads to the
1115: alternative estimates
1116: \begin{equation}
1117: \rlap{\hskip -.103\displaywidth\hbox{{\bf B}:}}% \qquad\qquad
1118: \check K^c_> = (17.5 \pm 1.5) \negthinspace \times \negthinspace 10^{-18},
1119: \quad\quad\quad
1120: \check K^c_< = -(5.4 \pm 0.5) \negthinspace \times \negthinspace 10^{-18},
1121: \label{eq5.4:KcglB}
1122: \end{equation}
1123: that, likewise, are much larger than in \whatis{eq5.4:KcglA}.
1124:
1125: Granted the values of $K^c_{\gl}$ in the critical point background
1126: \whatis{eq5.3:TcS0}, one is left with only
1127: the coefficients $\Sigma_h$ and $\Sigma_{hh}$ to be determined
1128: by fitting. However,
1129: since the {\bf NWW} measurements were made in terms of the density,
1130: it is necessary
1131: to convert the $(t, M \negthinspace = \negthinspace\rho \negthinspace -
1132: \negthinspace\rho_c)$ data to $(t, h)$ using an equation
1133: of state. To this end, it is natural to employ the
1134: scaled equation of state derived in \cite{Zinn99} on the basis of
1135: the extended sine model. The metric factor $B$ given in
1136: \whatis{eq5.3:amplB} and the appropriate values of $C^+$ [in
1137: \whatis{eq5.4:CpA} or \whatis{eq5.4:CpB}] may be used.
1138: The data spread over the range
1139: $-4.8 \times 10^{21} \lsim \check h \lsim 2.4 \times 10^{20}$
1140: [using \whatis{eq5.4:CpA}] and, hence, reach more widely
1141: on the negative $h$ (or low density)
1142: side; but for fitting we will focus on the available symmetric range.
1143:
1144: On the basis of the {\bf A} values above we thus
1145: find that the assignment
1146: \begin{equation}
1147: \check \Sigma_h \simeq 3.72 \negthinspace \times \negthinspace 10^{-23}
1148: \quad \quad \quad \hbox{and} \quad \quad \quad
1149: \check \Sigma_{hh} \simeq 6.65 \negthinspace \times \negthinspace 10^{-43},
1150: \label{eq5.4:Tcfits}
1151: \end{equation}
1152: provides a fit that lies well between the upper and lower limits
1153: for $|h|$ not too large: see Fig.~\ref{fig5:attc}.
1154: Comparable fits can be found using the {\bf B} values
1155: but with, of course, substantially different values of $\Sigma_h$ and
1156: $\Sigma_{hh}$ and a different accessible range of $h$. For the balance of
1157: this article, however, we will stay with the `purely theoretical' {\bf A}
1158: estimates \whatis{eq5.4:KcglA} and \whatis{eq5.4:CpA}.
1159:
1160: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1161: %
1162: % SUBSEC
1163: %
1164: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1165: \subsection{Cross Term}
1166: \noindent
1167: In order to estimate the cross coefficient $\Sigma_{th}$ in the
1168: background \whatis{eq5.1:S0}, one should examine
1169: the derivative $(\partial \Sigma / \partial
1170: t)_h$ at $T=T_c$. Since the measurements were made
1171: at constant $M$ (or density), it is
1172: appropriate to invoke the relation
1173: \begin{equation}
1174: \left( \PD{\Sigma}{t} \right)_{\negthinspace\negthinspace h} =
1175: \left(\PD{\Sigma}{t}\right)_{\negthinspace\negthinspace M}
1176: + \left( \PD{\Sigma}{M} \right)_{\negthinspace\negthinspace t}
1177: \left( \PD{M}{h} \right)_{\negthinspace\negthinspace t}.
1178: \label{eq5.4:xrel}
1179: \end{equation}
1180: However, owing to the measurement uncertainties, the estimates of
1181: $(\partial \Sigma / \partial t)_M$ obtainable from the
1182: {\bf NWW} data are extremely noisy.
1183: Hence, one cannot seriously attempt to estimate $\Sigma_{th}$.
1184: One may reasonably guess, nevertheless, that the geometric mean
1185: of $\Sigma_{tt}$ and $\Sigma_{hh}$, namely,
1186: using \whatis{eq5.4:h0fitb} and \whatis{eq5.4:Tcfits},
1187: \begin{equation}
1188: \left ( \check \Sigma_{tt} \check \Sigma_{hh} \right)^{1/2}
1189: = 1.6 \negthinspace \times \negthinspace 10^{-20},
1190: \label{eq5.4:S2x}
1191: \end{equation}
1192: will
1193: provide some indication of the possible order of magnitude of the
1194: cross-coefficient. However,
1195: we comment below on the sensitivity of the final
1196: comparisons with experiment to the value of
1197: $\Sigma_{th}$.
1198:
1199: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1200: %
1201: % SEC
1202: %
1203: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1204: \section{Scaling Functions for the Surface Tension}
1205: \noindent
1206: Having found, to the somewhat limited degree feasible,
1207: a reasonable description of the important surface-tension
1208: background, we may examine the experimental data for consistency with the
1209: predicted universal scaling functions $S_M^\pm(\mt)$
1210: introduced in \whatis{eq5.1:Spm}.
1211:
1212: In order to subtract the background $\Sigma_0(t,h)$, from the full surface
1213: tension data, it is necessary, as before,
1214: to convert the variables $(t, M \negthinspace = \negthinspace \rho
1215: \negthinspace - \negthinspace \rho_c)$ to $(t, h)$
1216: because the experiments employed the density as a primary variable.
1217: Adopting the {\bf NWW} critical
1218: temperature $T_c = 26.310 \thinspace \celc$ and using $B$ from
1219: \whatis{eq5.3:amplB}, the first step is to compute
1220: $\mt = \penalty-100000 (\rho - \rho_c)/B|t|^\beta$.
1221: The universal, scaled equation of state $\mt = Q_\pm(\tilde h)$ [for
1222: $t \gl 0$] may be used in the form of the extended sine model of
1223: \cite{Zinn99} to implement the conversion to $\tilde h$.
1224: The corresponding $h$ follows via
1225: $h = B \tilde h |t|^\Delta / C^+$ where, for the present discussion we
1226: use \whatis{eq5.4:CpA} for $C^+$ as in Fig.~\ref{fig5:attc}.
1227: Then, using the background with the coefficients given
1228: in \whatis{eq5.4:h0fitb} and \whatis{eq5.4:Tcfits}, an estimate
1229: for the singular part of the surface tension, $\Delta \Sigma (T, h)$,
1230: can be extracted from the {\bf NWW} data.
1231:
1232: Note that in the first instance,
1233: the cross term, $\Sigma_{th} t h$, may
1234: simply be dropped. The resulting scaling contribution, $\Delta \Sigma
1235: / K |t|^\mu$, is plotted versus $\mt$, the scaled density deviation,
1236: in Fig.~\ref{fig5:SMpm}.
1237:
1238: Since the field-dependence of the background $\Sigma_0(t,h)$ has been
1239: estimated only from data that lie on the two isotherms closest to
1240: criticality and fall within the restricted range
1241: $|\check h| \lsim 2.4 \negthinspace \times \negthinspace 10^{20}$,
1242: we have distinguished, by solid dots \stress{versus} crosses and plusses,
1243: between these observations, corresponding to the restricted range of $h$,
1244: and (selected) further data falling outside this range.
1245:
1246: Because the background contribution constitutes such a large part of the
1247: total surface tension, it is not surprising that the subtraction needed
1248: to isolate the singular piece, $\Delta \Sigma (t, h)$, is subject to
1249: relatively large uncertainties. This appears to be the main reason why
1250: many of the crosses and plusses, derived from the low-density data
1251: (beyond the fitted range of $h$), fall significantly below the scaling
1252: loci.
1253: For the data inside the range (the solid dots in Fig.~4),
1254: the agreement between the
1255: theoretical and experimental results is moderately encouraging considering the
1256: uncertainties of the original data
1257: and the inherent difficulties of the analysis. Certainly, the data of
1258: {\bf NWW}
1259: support the expected difference between $S_M^+(\mt)$
1260: and $S_M^-(\mt)$.
1261:
1262: It may be noted, nevertheless, that
1263: just as pointed out by {\bf NWW},
1264: the logarithmic singularity associated with the complete-wetting
1265: transition cannot be detected in these or other surface-tension plots.
1266: However, this is due not only
1267: to the limited precision of the experimental data but also
1268: to the fact that in the usual
1269: representations of the data the singularity is not really visible
1270: even in theoretical plots where it is known analytically to be present:
1271: compare with Figs.~5 and 6 of {\bf I}.
1272:
1273: If one also includes a \stress{positive} cross-coefficient
1274: $\Sigma_{th}$ in the background
1275: with the magnitude given in \whatis{eq5.4:S2x},
1276: one finds slight changes of less than 1\% in
1277: the data points plotted in Fig.~\ref{fig5:SMpm}. However, when the
1278: cross-term is inserted with a \stress{negative} sign, the majority
1279: of the data points display changes of $\sim\negthinspace5$\% and a few
1280: points prove still more sensitive.
1281: This situation is hardly satisfying but, in light of the experimental
1282: challenges and the resulting precision of the {\bf NWW} data,
1283: it is clear that our analysis can proceed little further.
1284:
1285: Finally, as pointed out theoretically by Ramos-G\'omez and Widom
1286: \cite{Widom80} and mentioned briefly in {\bf I},
1287: the surface tension isotherms when examined vs density
1288: ($\propto M$) above $T_c$ should, in general, exhibit
1289: \stress{crossings}.
1290: Although hints of such crossings have appeared in some binary fluid
1291: systems \cite{Campbell68, McLure79, LastNote}, they were not seen in the
1292: {\bf NWW} experiments. The primary reason seems to be that for the
1293: range of densities $\rho > \rho_c$ explored, the isotherms chosen
1294: for observation above
1295: $T_c$ were somewhat too widely spaced. Indeed, as illustrated in
1296: Fig.~\ref{fig6:isox}, the anticipated crossings for isobutyric acid $+$
1297: water can be exhibited on the basis of our fits to the {\bf NWW} data.
1298: All that is needed is to choose temperatures above $T_c$ spaced apart by
1299: intervals of no more than $1 \thinspace \celc$ and to extend the
1300: observations to densities slightly further above $\rho_c$. The crossings,
1301: while sensitive to the presence of the background term, should be
1302: visible even in the singular part of the surface tension, $\Delta
1303: \Sigma(T, \rho)$. Note, however, that since our fits were based only on
1304: the restricted (symmetric) range of $h$, the plots lying above the short
1305: horizontal lines in Fig.~\ref{fig6:isox} may be less reliable than at
1306: lower densities.
1307:
1308: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1309: %
1310: % SUMMARY
1311: %
1312: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1313: \section{Summary}
1314: \noindent
1315: In pioneering experiments to study critical endpoint behavior,
1316: Nagarajan, Webb, and Widom \cite{NWW} made extensive measurements
1317: of the vapor-liquid interfacial
1318: tension of isobutyric acid and water mixtures as a function of
1319: temperature at various compositions. Since
1320: the precision of the surface tension data was somewhat
1321: limited, we used the observations of the critical surface tension
1322: made by Howland \etal~\cite{Knobler} and the coexistence curve
1323: measurements of Greer~\cite{Greer} to cross-calibrate the {\bf NWW}
1324: data.
1325: We found full consistency between the various observations and verified
1326: agreement with Antonow's rule near the critical endpoint. In this
1327: way reliable estimates of the coexistence curve amplitude $B$ and the
1328: surface tension amplitude $K$ were obtained: see \whatis{eq5.3:amplB}
1329: and \whatis{eq5.3:Kexp}.
1330: Using known universal amplitude ratios \cite{Zinn98}
1331: and these two amplitudes, which can serve as the only required metric factors
1332: in a hyperscaling formulation,
1333: we estimated the nonuniversal susceptibility amplitude $C^+$.
1334: Thereby the experimental data as a function of density could be
1335: transformed in order to uncover the field-dependence of the crucial
1336: surface-tension background contribution in the form (1.10).
1337: Owing to the considerable uncertainties, however,
1338: we could not resolve the cross-coefficient
1339: $\Sigma_{th}$. Nevertheless, even without this term, the scaling
1340: plots for the singular part of the interfacial tension as derived from
1341: the experimental data proved consistent with the theoretical
1342: predictions based on the EdGF theory of {\bf I} \cite{Zinn03}:
1343: see Fig.~\ref{fig5:SMpm}.
1344:
1345: On the other hand, as discussed in the Appendix, significant
1346: discrepancies come to light when the Howland \etal~data are compared,
1347: via hyperscaling expectations, with evidence from other critical
1348: systems. Owing to conflicting experimental reports \cite{Kprob,
1349: Moldover85}, however, it seems impossible to resolve this issue and
1350: assess its significance without further experiments and, perhaps, new
1351: theoretical insights.
1352:
1353: More rewarding from a theoretical perspective and less open to questions
1354: of precision, the critical surface tension
1355: measurements of Mainzer-Althof and Woermann~\cite{Mainzer} on 2,6-dimethyl
1356: pyridine $+$ water at the critical composition, show rather satisfactory
1357: agreement
1358: between the experimentally determined universal ratios $P$ and $Q$
1359: [see (1.9)] and the EdGF predictions. However, as regards the full
1360: scaling aspects of the theory, more detailed, more precise and more
1361: reliable data are sorely needed to execute more searching tests!
1362:
1363: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1364: %
1365: % ACK
1366: %
1367: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1368: \begin{acknowledgments}
1369: \noindent
1370: We appreciate the keen interest of B.~Widom and comments from J.~Indekeu
1371: and M.~R.~Moldover.
1372: We are indebted to C.~M.~Knobler and I.~L.~Pegg for informative
1373: discussions regarding their surface tension measurements.
1374: The support of the National Science Foundation
1375: through grants CHE 99-81772 and CHE 03-01101 is gratefully acknowledged.
1376:
1377: Especially on this occasion, it is a pleasure for M.E.F. to acknowledge
1378: the inspiration provided over more than four decades by the limpidly
1379: insightful researches of Benjamin Widom.
1380: \end{acknowledgments}
1381:
1382: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1383: %
1384: % APPENDIX
1385: %
1386: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1387: \appendix*
1388: \section{Nonuniversal Critical Amplitudes}
1389: \noindent
1390: In Sec.~III the coexistence curve amplitude $B$ for isobutyric acid $+$
1391: water (IBA:W) was determined from the data of Greer \cite{Greer} and
1392: checked against the {\bf NWW} data yielding \whatis{eq5.3:amplB}. Likewise,
1393: analysis of the {\bf HWK} capillary-rise surface tension data yielded
1394: the critical amplitude $K$ reported in \whatis{eq5.3:Kexp} that also
1395: accords well with the {\bf NWW} data using forced capillary waves:
1396: indeed, no larger deviation in \whatis{eq5.3:Kexp}, than, say, $\pm 6.0$ (or
1397: $\pm 18$\%) could be reconciled with the {\bf NWW} observations
1398: \cite{ThankCK}. We stress this point because Moldover
1399: \cite{Moldover85}, in comparing these IBA:W data with results for many
1400: other fluid systems, uncovered an inconsistency in the value of the
1401: expected-to-be universal amplitude ratio
1402: \begin{equation}
1403: S^+ \equiv \Kbar (\xi^+_1)^2 = 0.377 \pm 0.011
1404: \quad\quad\quad \hbox{with} \quad\quad\quad
1405: \Kbar \equiv K/k_B T_c. \label{eqA:Kxi2}
1406: \end{equation}
1407: Here, in the notation of the Appendix and Table~I of \cite{Zinn98},
1408: the amplitude $\xi^+_1$ specifies the
1409: magnitude of the divergence of the second-moment
1410: correlation length above $T_c$ via $\xi_1(T) \approx \xi^+_1 /t^\nu$,
1411: while the value quoted for $S^+$ is that reported in \cite{Zinn98}.
1412:
1413: If we now accept this theoretical result
1414: we may proceed to predict the value of $\xi_1^+$
1415: for IBA:W. Thence, using the universal ratio \cite{Zinn98}
1416: \begin{equation}
1417: S_0 \equiv C^+ \Kbar / B^2 \xi^+_1 = 1.17 \pm 0.06,
1418: \label{eqA:S0}
1419: \end{equation}
1420: together with the estimates for $K$ and $B$, we can compute the
1421: susceptibility amplitude $C^+$ that is needed in Sec.~V to establish the
1422: $h$ scale in Fig.~4. This route, indeed, yields the {\bf A} value for
1423: $C^+$ reported in \whatis{eq5.4:CpA} and, with the further aid of
1424: \whatis{eq5.3:Scgl} and {\bf I} (6.9), the values for $K^c_>$ and $K^c_<$
1425: given in \whatis{eq5.4:KcglA}. These results, in turn, are incorporated in
1426: the plots in Fig.~5 that serve to test the scaling hypothesis.
1427:
1428: The correlation length amplitude predicted in this fashion is $\xi_{\rm
1429: 1, pre}^+ =
1430: 2.10 \pm 0.08 \thinspace \Ao$ [where the {\bf HWK} value
1431: \whatis{eq5.3:KnoblerTc}
1432: has been adopted for $T_c$ in \whatis{eqA:Kxi2}]. However, as reported by
1433: Moldover \cite{Moldover85},
1434: in light scattering experiments by Chu and co-workers~\cite{Chu73, Chu68},
1435: the second-moment
1436: correlation length above $T_c$ was evaluated for IBA:W.
1437: The fits presented led them to conclude
1438: $\nu \simeq 0.613$ and $\xi_1^+ = 3.57 \pm 0.07 \thinspace \Ao$.
1439: This correlation amplitude
1440: exceeds the predicted value, $\xi^+_{1, {\rm pre}}$, by 60\% or more!
1441: On re-analyzing their
1442: data with the higher value $\nu = 0.632$ imposed, we obtain
1443: $\xi_1^+ = 3.09 \pm 0.05 \Ao$. This is somewhat closer to
1444: $\xi_{1, \rm pre}^+$
1445: although the uncertainty limits fail to overlap by 36\%.
1446: From turbidity
1447: measurements, Beysens \etal~\cite{Beysens} concluded $\xi_1^+ = 3.62_5 \pm
1448: 0.06_5 \Ao$
1449: and $\nu \simeq 0.630$; this value of $\xi_1^+$ agrees with Chu
1450: \etal~and is some 17\% higher than that from
1451: our reanalysis.
1452:
1453: At a basic theoretical level the discrepancies uncovered here are of
1454: serious concern. If, for concreteness, one accepts
1455: $\xi^+_{\rm 1, expt} \simeq 3.6_0 \pm 0.07 \thinspace \Ao$
1456: as the observed value, in accord with these experiments,
1457: the universal ratio $S^+$ in \whatis{eqA:Kxi2}
1458: would yield a prediction for $\check K$ of only 12.0,
1459: about 36\% of that observed! Conversely, one might be tempted to assert
1460: that the
1461: experimentally determined value of the (supposedly universal) ratio
1462: for IBA:W is $S^+_{\rm expt} = 1.06 \pm 0.03$, in place of
1463: \whatis{eqA:Kxi2},
1464: the uncertainty indicated here taking into account various sources of
1465: experimental and theoretical imprecision.
1466:
1467: Similarly, the {\bf B} values for $C^+$ and $K^c_{\gl}$ reported in
1468: \whatis{eq5.4:CpB} and \whatis{eq5.4:KcglB} follow from the observed value of
1469: $B$ and the experimental value, $\xi_{1, \rm expt}^+$, by using
1470: \cite{Zinn98}
1471: \begin{equation}
1472: Q_c \equiv (\xi_1^+)^3B^2/C^+ = 0.323_6 \pm 0.006.
1473: \end{equation}
1474:
1475: In facing this discrepant situation let us focus first on the
1476: theoretical value \whatis{eqA:Kxi2} for $S^+$ (which, indeed, was
1477: incorporated into the numerical scaling calculations reported in {\bf I}
1478: and tested here). One might initially notice that 0.377 is
1479: significantly higher than the various theoretical, RG and Monte Carlo
1480: estimates, namely, $\simeq \negthinspace 0.21$, $0.24$, and $0.28$,
1481: available to Moldover in 1985 \cite{Moldover85}.
1482: In fact \whatis{eqA:Kxi2} rests on the analysis \cite{Zinn} of difficult but
1483: carefully performed Monte Carlo simulations for the nearest-neighbor
1484: $(d=3)$-dimensional Ising model by Hasenbush and Pinn \cite{Pinn94}
1485: (combined with the well established universal ratio $\xi_1^+ / \xi_1^- =
1486: 1.96 \pm 0.01$ \cite{Zinn, Liu}).
1487: The simulations \cite{Pinn94} could
1488: approach criticality no closer than $|t| > 0.015$. By contrast, the
1489: experimental data for IBA:W
1490: lie in the region $|t| = 0.0008$ to $0.010$. Although
1491: it seems unlikely, it is conceivable that
1492: exceptionally strong correction-to-scaling terms in the surface tension
1493: of the simple cubic Ising model could dominate very close to $T_c$
1494: so that \whatis{eqA:Kxi2} represents a significant under-estimate of $S^+$.
1495: Alternatively, finite-size effects in the simulations may be playing a
1496: far larger role than appreciated although no evidence to suggest this
1497: was observed and the issue was by no means neglected \cite{Gelfand90,
1498: Pinn94}.
1499:
1500: On the other hand, it might be that for reasons associated with the
1501: lattice structure and capillary-wave fluctuations, the standard simple
1502: cubic lattice gas/Ising model is inadequate for certain real fluid
1503: mixtures. Most certainly, the conventional near-neighbor
1504: lattice gases neglect the long-range power-law van der Waals
1505: interactions that lead to interphase potentials decaying, normal to
1506: an interface, only as $1/z^3$. And while these van
1507: der Waals forces play only a \textit{subdominant} role in \textit{bulk} critical
1508: behavior, they are known to be especially significant in surface and
1509: wetting phenomena \cite{Dietrich}. It is, thus, possible that
1510: they seriously distort the near critical interfacial tension or even
1511: destroy the expected universal character of the ratio $S^+$.
1512:
1513: At present, however, effective theoretical tools for addressing this
1514: interesting issue are not apparent. Furthermore, it is not unfair to
1515: remark that these various rather deep theoretical issues do not directly
1516: affect our
1517: basic \textit{scaling} \textit{analysis} of the {\bf NWW} observations
1518: of surface tensions near a critical endpoint. Indeed, in the first
1519: instance only the experimentally determined critical amplitudes $K$ and
1520: $B$ enter: universal or nonuniversal relations between these and other
1521: critical amplitude do not play a definitive role in the primary scaling
1522: issues.
1523:
1524: Finally, we must return to the experimental evidence. It is, indeed,
1525: striking that the value for $S^+$ in \whatis{eqA:Kxi2} accords remarkably
1526: well with the experimental values assembled by Moldover
1527: \cite{Moldover85} for Ar, Xe, ${\rm N}_2$, ${\rm O}_2$, ${\rm CO}_2$,
1528: ${\rm CH}_4$ and ${\rm SF}_6$ and for the
1529: binary fluids triethylamine $+$ water, nitroethane $+$ 3-methylpentane,
1530: cyclohexane $+$ aniline,
1531: methanol $+$ cyclohexane, and even for various polystyrene $+$ ${\rm
1532: C}_7{\rm H}_{14}$ solutions: see Fig.~1, Table I and also the {\it Note
1533: added\/} regarding Ref.~86 in \cite{Moldover85} and Table I of
1534: \cite{Moldover86}. Indeed, by and large, Moldover's analysis confirms
1535: rather generally the hyperscaling hypothesis, \stress{both} for the
1536: interfacial \stress{and} for the bulk thermodynamic properties. (See also
1537: in \cite{Zinn1997}.)
1538:
1539: These observations tend to reinforce confidence in the estimate
1540: \whatis{eqA:Kxi2} for the universal surface tension amplitude---at least as
1541: regards `simple' pure fluids and binary mixtures. But does the weak
1542: acid IBA:W
1543: constitute such a `simple' system? On the one hand, as an electrolyte,
1544: a population of positive and negative ions will be present in all the
1545: liquid phases. Furthermore, it is known that the critical point of
1546: IBA:W is
1547: extremely sensitive to small impurities and, especially, to
1548: ionic impurities \cite{Greer, Knobler05}. Ionic impurities, in
1549: particular, may be expected to have significant effects on the various
1550: liquid-liquid, liquid-vapor and liquid-solid interfaces; but even in the
1551: absence of impurities, an interphase Galvani potential difference will
1552: appear across the critical liquid-liquid interface together with an
1553: associated electrically charged double layer \cite{REF51, REF52, REF53,
1554: REF54}.
1555:
1556: On the other hand, Moldover \cite{Moldover85} in addressing
1557: the discrepancy issue
1558: notes, as a private communication from I. L. Pegg and C. M. Knobler,
1559: that further surface-tension measurements on IBA:W using a different
1560: technique (actually light scattering from thermally excited capillary
1561: waves \cite{Knobler05}) yielded the preliminary result $\check K = 11.6
1562: \pm 1.0$. This value is almost three times smaller than found in the
1563: {\bf HWK} experiments with which, as indicated above, it cannot be
1564: reconciled. Nevertheless, if this value for $\check K$ is
1565: used with the experimentally determined values,
1566: $\xi_{1, \rm expt}^+$, discussed above it yields a value for $S^+
1567: \propto \check K (\xi_1^+)^2$ close to that found experimentally for the
1568: other systems and, hence, in \textit{accord} with the expected universal
1569: value stated in (A.1) \cite{Moldover85, Moldover86}! On this basis one
1570: might (despite \cite{ThankCK}) be tempted to disregard the {\bf HWK} data
1571: and argue that our analysis of the {\bf NWW} observations should be
1572: seriously reconsidered.
1573:
1574: Unfortunately, the preliminary study of IBA:W by Pegg and Knobler has
1575: never been published \cite{Knobler05}; nor, to our knowledge, have any
1576: subsequent measurements been made by others on this system.
1577: Thus, while one can and,
1578: perhaps, should attribute the nearly
1579: three-fold difference in the critical
1580: surface tension to unidentified impurities or other agencies
1581: distinguishing the {\bf HWK}
1582: and {\bf NWW} samples of isobutyric acid from the later sample
1583: investigated by Pegg and Knobler (in 1984), the true causes
1584: of the discrepancies and their
1585: significance remain obscure. We must, instead, hope that more sensitive
1586: techniques will be developed and employed on the IBA:W and
1587: other binary systems in
1588: order to more stringently test the scaling theories for critical
1589: endpoints that have been developed.
1590:
1591: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1592: %
1593: % REF
1594: %
1595: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1596: \bibliography{ms}%
1597: \centerline{\vbox{\hrule width 3 in height 2 pt}}
1598:
1599: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1600: %
1601: % TABLE
1602: %
1603: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1604: % TABLE I
1605: \newpage
1606: \begin{table}
1607: \caption{Adopted values of fluid critical exponents. For convenience
1608: and consistency with the universal amplitude ratios assembled in
1609: \cite{Zinn98} we have employed the corresponding exponent values
1610: (satisfying the scaling relations). More recent studies
1611: \cite{Butera2000, Pelissetto2002, Guida1998, Campostrini2002}
1612: lead to our currently
1613: preferred exponent estimates shown here in parentheses. However, for
1614: the present analysis the differences are of negligible consequence.
1615: We may supplement the relations \whatis{eq5.1:exprel} with $\Delta =
1616: \beta \delta = \beta + \gamma$ and $\gamma = (2 - \eta) \nu$.}
1617: \label{tbl:exp}
1618: \begin{ruledtabular}
1619: \begin{tabular}{c c c c c}
1620: \emspace\emspace\emspace $\beta$ & $\gamma$ & $\nu$ & $\mu$ &
1621: $\theta$ \emspace\emspace\emspace\\
1622: \hline
1623: \emspace\emspace\emspace 0.3266 & 1.2392 & 0.6308 & 1.2616 &
1624: 0.54 \emspace\emspace\emspace\\
1625: \emspace\emspace\emspace (0.3260) & (1.2390) & (0.6303) & (1.2606) &
1626: (0.52) \emspace\emspace\emspace\\
1627: \end{tabular}
1628: \end{ruledtabular}
1629: \end{table}
1630:
1631: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1632: %
1633: % FIGURE CAPTIONS
1634: %
1635: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1636: \hbox to 1 in { }
1637: \newpage
1638: \noindent
1639: {\bf FIGURE CAPTIONS}
1640:
1641: \noindent
1642: FIG.~1.
1643: The coexistence curve of the isobutyric acid and water
1644: mixture. The dots represent the experimental data of {\bf NWW}.
1645: The inner curve is the fit \whatis{eq5.3:form} using the
1646: amplitude $\check B = 0.0310$ with $a_\theta$ and $a_1$
1647: from \whatis{eq5.3:Brho} while the outer curve employs $\check B = 0.0318$
1648: in accord with \whatis{eq5.3:Brho2}; the critical temperature has been taken
1649: as $T_c = 26.310 \celc$ in accord with the estimate of {\bf NWW}.
1650: \bigskip
1651:
1652: \noindent
1653: FIG.~2.
1654: Comparison of surface-tension measurements along the coexistence
1655: curve: the open circles with error bars represent the {\bf NWW}
1656: data. As explained in the text, the parallelograms have been drawn
1657: using $\Sbr = K|t|^\mu$ with the amplitude
1658: \whatis{eq5.3:Kexp} obtained from the {\bf HWK} data.
1659: The vertical sides of the parallelograms indicate the ranges in which
1660: consistency between the two experiments can be achieved
1661: in accord with Antonow's rule. The dashed curve,
1662: is based on the fit \whatis{eq5.4:h0fit} obtained below for the background
1663: $\Sigma_0(T,0)$,
1664: while the solid curve embodies the values \whatis{eq5.4:h0fitb}.
1665: See text in Sec.~IV.A for details.
1666: \bigskip
1667:
1668: \noindent
1669: FIG.~3.
1670: The fit to the surface tension background in zero-field. See text
1671: for details of the generation of the data points (black dots). The dashed
1672: curve is drawn using the estimates \whatis{eq5.4:h0fit}. The solid curve
1673: represents the fit \whatis{eq5.4:h0fitb}.
1674: \bigskip
1675:
1676: \noindent
1677: FIG.~4.
1678: A fit for the surface tension on the critical isotherm
1679: as a function of the ordering field. The dots
1680: joined by the dashed lines represent the {\bf NWW}
1681: isotherms nearest the critical
1682: temperature that serve as approximate
1683: upper and lower bounds. See text for details.
1684: \bigskip
1685:
1686: \noindent
1687: FIG.~5.
1688: Comparison of scaling plots
1689: for the vapor-liquid surface tension near the critical endpoint of
1690: isobutyric acid $+$ water. The solid curves represent the theoretically
1691: predicted scaling functions $S_M^\pm(\mt)$ using the ameliorated EdGF theory.
1692: The dots are derived from the {\bf NWW}
1693: data for $T>T_c$ and $T<T_c$ that fall within the symmetric
1694: range $|\check h| \lsim 2.4 \negthinspace \times \negthinspace 10^{20}$
1695: used in fitting the surface tension background.
1696: The other symbols ($\times$ for $T>T_c$ and $+$ for $T<T_c$)
1697: are some of the {\bf NWW} data
1698: falling outside the symmetric range of $h$ (and hence not covered by the
1699: background fit).
1700: Note that an offset of $\Delta S = 100$
1701: has been used for clarity in the vertical scales
1702: for $T \gl T_c$. See the text for further details.
1703: \bigskip
1704:
1705: \noindent
1706: FIG.~6.
1707: Illustration of predicted crossings of surface-tension isotherms of
1708: isobutyric acid $+$ water above $T_c$.
1709: The isotherms have been computed using the EdGF scaling functions
1710: $S^\pm_M(\mt)$ from {\bf I} with a background of the form~\whatis{eq5.1:S0}
1711: suitably matched to the experimental data.
1712: The dotted, solid, dashed, and dot-dashed curves represent isotherms
1713: $\Delta T \equiv T - T_c = 0$,
1714: $1.0$, $2.0$, and $3.0 \celc$, respectively.
1715: The narrow horizontal and vertical lines mark
1716: the critical values of $\Sigma$ and $\rho$.
1717: The short horizontal lines mark the ordering field values up to which
1718: the fitting was performed.
1719: See the text for further details.
1720:
1721: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1722: %
1723: % FIGURES
1724: %
1725: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1726: %FIG
1727: \begin{figure}
1728: \includegraphics{fig1_coex}
1729: \caption{\label{fig5:coex}
1730: }
1731: \end{figure}
1732:
1733: %FIG
1734: \begin{figure}
1735: \includegraphics[scale=0.95]{fig2_chk}
1736: \caption{\label{fig5:chk}
1737: }
1738: \end{figure}
1739:
1740: %FIG
1741: \begin{figure}
1742: \includegraphics{fig3_sigma0}
1743: \caption{\label{fig5:sigma0}
1744: }
1745: \end{figure}
1746:
1747: %FIG
1748: \begin{figure}
1749: \includegraphics{fig4_attc}
1750: \caption{\label{fig5:attc}
1751: }
1752: \end{figure}
1753:
1754: %FIG
1755: \begin{figure}
1756: \includegraphics[scale=0.85]{fig5_smpm}
1757: \caption{\label{fig5:SMpm}
1758: }
1759: \end{figure}
1760:
1761: %FIG
1762: \begin{figure}
1763: \includegraphics[scale=1.0]{fig6_isox}
1764: \caption{\label{fig6:isox}
1765: }
1766: \end{figure}
1767:
1768: \end{document}
1769: