cond-mat0401118/art.tex
1: \documentclass[pre,twocolumn,aps,floats]{revtex4}
2: %\documentclass[pre,aps,floats]{revtex4}
3: \usepackage{graphicx,epstopdf}
4: %\input isolatin1.sty
5: %\documentstyle[preprint,aps,epsfig]{revtex}
6: 
7: \newcommand{\csch}{\mbox{csch}}
8: \newcommand{\sech}{\mbox{sech}}
9: \newcommand{\re}{\mbox{Re\,}}
10: \newcommand{\Du}{\mbox{D\hspace{-0.5pt}u}}
11: \newcommand{\del}{\mbox{\boldmath{$\nabla$}}}
12: \newcommand{\nb}{\mbox{\bf n}}
13: \newcommand{\Fb}{\mbox{\bf F}}
14: \newcommand{\Jb}{\mbox{\bf J}}
15: \newcommand{\rb}{\mbox{\bf r}}
16: \newcommand{\xb}{\mbox{\bf x}}
17: \newcommand{\ub}{\mbox{\bf u}}
18: 
19: 
20: 
21: \begin{document}
22: %\draft
23: 
24: %\title{Intermediate time scale in the relaxation of
25: %a dilute electrochemical cell}
26: 
27: \title{ Diffuse-Charge Dynamics in Electrochemical Systems }
28: 
29: \author{Martin Z. Bazant$^{1,2}$, Katsuyo
30: Thornton$^3$, and  Armand Ajdari$^2$}
31: 
32: \affiliation{
33: $^1$ Department of Mathematics, Massachusetts Institute of Technology,
34: Cambridge, MA 02139 \\
35: $^2$ Laboratoire de Physico-Chimie Th{\'e}orique, UMR ESPCI-CNRS 7083,
36: 10 rue Vauquelin, F-75005 Paris, France\\
37: $^3$ Department of Materials Science and
38: Engineering, Northwestern University, Evanston, IL 60201}
39: 
40: % first calculations: June 1999, Bordeaux.
41: % Began paper: Paris, July 2000.
42: % First draft without much numerics: Jan 2003
43: % First COMPLETE draft: Jan 2004.
44: 
45: \date{\today}
46: 
47: \begin{abstract}
48: 
49: The response of a model micro-electrochemical system to a
50: time-dependent applied voltage is analyzed. The article begins with a
51: fresh historical review including electrochemistry, colloidal science,
52: and microfluidics. The model problem consists of a symmetric binary
53: electrolyte between parallel-plate, blocking electrodes which suddenly
54: apply a voltage.  Compact Stern layers on the electrodes are also
55: taken into account. The Nernst-Planck-Poisson equations are first
56: linearized and solved by Laplace transforms for small voltages, and
57: numerical solutions are obtained for large voltages.  The ``weakly
58: nonlinear'' limit of thin double layers is then analyzed by matched
59: asymptotic expansions in the small parameter $\epsilon = \lambda_D/L$,
60: where $\lambda_D$ is the screening length and $L$ the electrode
61: separation. At leading order, the system initially behaves like an RC
62: circuit with a response time of $\lambda_D L / D$ (not
63: $\lambda_D^2/D$), where $D$ is the ionic diffusivity, but nonlinearity
64: violates this common picture and introduce multiple time scales. The
65: charging process slows down, and neutral-salt adsorption by the
66: diffuse part of the double layer couples to bulk diffusion at the time
67: scale, $L^2/D$. In the ``strongly nonlinear'' regime (controlled by a
68: dimensionless parameter resembling the Dukhin number), this effect
69: produces bulk concentration gradients, and, at very large voltages,
70: transient space charge. The article concludes with an overview of more
71: general situations involving surface conduction, multi-component
72: electrolytes, and Faradaic processes.
73: 
74: \end{abstract}
75: 
76: 
77: \maketitle
78: 
79: 
80: %%%%%%%%%%%%%%%%%%%%%%%%%
81: \section{Introduction }
82: %%%%%%%%%%%%%%%%%%%%%%%%%
83: 
84: There is rapidly growing interest in micro-electrochemical or
85: biological systems subject to time-dependent applied voltages or
86: currents. For example, AC voltages applied at microelectrodes can be
87: used to pump liquid
88: electrolytes~\cite{ramos98,ramos99,green00a,gonzalez00,green02,ajdari00,brown01,studer02,mpholo03,ramos03,nadal02b},
89: to separate or self-assemble colloidal
90: particles~\cite{yeh97,trau97,faure98,green00b,nadal02a,marquet02,ristenpart03}, and
91: to manipulate biological cells and
92: vesicles~\cite{helfrich74,mitov93,pethig96}. Conversely, oscillating
93: pressure-driven flows can be used to produce frequency-dependent
94: streaming potentials to probe the structure of porous
95: media~\cite{journiaux97,reppert01,reppert02}.
96: 
97: A common feature of these diverse phenomena is the dynamics of diffuse
98: charge in microscopic systems.  Although the macroscopic theory of
99: neutral electrolytes with quasi-equilibrium double layers is very well
100: developed in electrochemistry~\cite{bard,newman} and colloidal
101: science~\cite{hunter,russel,lyklema}, microscopic double-layer
102: charging at subdiffusive time scales is not as well
103: understood. Although much progress has been made in various disjoint
104: communities, it is not so widely appreciated, and some open questions
105: remain, especially regarding nonlinear effects. The goals of this
106: paper are, therefore, (i) to review the relevant literature and (ii)
107: to analyze a basic model problem in considerable depth, highlighting
108: some new results and directions for further research.
109: 
110: \begin{figure}
111: \includegraphics[width=3in]{cartoon.eps}
112: \caption{ Sketch of the model problem. A voltage $2V$ is suddenly
113: applied to a dilute, symmetric, binary electrolyte between
114: parallel-plate, blocking electrodes separated by
115: $2L$. \label{fig:cartoon} }
116: \end{figure}
117: 
118: To illustrate the physics of diffuse-charge dynamics, consider the simplest
119: possible case sketched in Fig.~\ref{fig:cartoon}: a dilute $z$:$z$ electrolyte
120: suddenly subjected to a DC voltage, $2V$, by parallel-plate blocking
121: electrodes separated by $2L$. Naively, one might assume a uniform bulk electric
122: field, $E = V/L$, but the effect of the applied voltage is not so
123: trivial. Ions migrate in the bulk field and eventually screen it completely
124: (since ``blocking electrodes'' do not support a Faradaic current).
125: 
126: What is the characteristic time scale of this response? For charge relaxation,
127: one usually quotes the time, $\tau_D = \lambda_D^2 /D$, for diffusion with a
128: diffusivity $D$ across one Debye screening length,
129: \begin{equation}
130: \lambda_D = \sqrt{\frac{\varepsilon k T}{2z^2e^2C_b}} ,
131:  \label{eq:lambda}
132: \end{equation}
133: where $C_b$ is the average solute concentration, $k$ Boltzmann's
134: constant, $T$ the temperature, $e$ is the electronic charge, and $\varepsilon$ the permittivity of the
135: solvent~\cite{hunter,russel,lyklema}. The Debye time, $\tau_D$, is a material
136: property of the electrolyte, which for aqueous solutions ($\lambda_D
137: \approx 1-100$ nm, $D \approx 10^3$ $\mu$m$^2$/s) is rather small, in
138: the range of ns to $\mu$s. More generally, when Faradaic reactions occur (for
139: a non-blocking electrode), diffuse charge may also vary on the much
140: slower, geometry-dependent scale for bulk diffusion given by $\tau_L = 
141: L^2/D$, proportional to the square of the electrode separation.
142: 
143: These two relaxation times, $\tau_D$ for the charge density and
144: $\tau_L$ for the concentration, are usually presented as the only ones
145: controlling the evolution of the system, e.g.~ as in the recent
146: textbooks of Hunter~\cite{hunter} (Ch.\ 8) and Lyklema~\cite{lyklema}
147: (Chs.\ 4.6c). Dimensional analysis, however, allows for many other time
148: scales obtained by combining these two, such as the harmonic mean,
149: \begin{equation}
150: \tau_c = \sqrt{\tau_D \tau_L} = \frac{\lambda_D L}{D},  \label{eq:tmixed}
151: \end{equation}
152: proportional to the electrode separation (not squared). Below, we will
153: show that this is the primary time scale for diffuse-charge dynamics in
154: electrochemical cells, although $\tau_D$, $\tau_L$, and other time
155: scales involving surface properties also play important roles,
156: especially at large voltages (even without Faradaic processes). The
157: same applies to highly polarizable or conducting colloidal particles,
158: where $L$ is the particle size.
159: %Does a need to be introduced here? (Why not "where $L$ is given by the particle size"?) 
160: 
161: Although the basic charging time, $\tau_c$, is familiar in several scientific
162: communities~\cite{macdonald70,korn81,dukhin80,dukhin93}, it
163: is not as widely known as it should be. Recently, it
164: has been rediscovered as the (inverse) frequency of ``AC pumping'' at
165: patterned-surface micro-electrodes~\cite{ramos98,ajdari00}. As in the past, its
166: theoretical justification has sparked some controversy~\cite{scott01,ramos01}
167: related to the applicability of classical circuit
168: models~\cite{macdonald90,geddes97} in which $\tau_c$ arises as the ``RC time''
169: of a bulk resistor in series with a double-layer capacitor (see below).
170: 
171: Here, we attempt to unify and modestly extend a large body of prior
172: work on diffuse-charge dynamics in the context of our model problem,
173: paying special attention to effects which undermine the classical
174: circuit approximation. Going beyond most previous mathematical
175: studies, we allow for compact-layer capacitance, bulk concentration
176: polarization, and large voltages outside the linear regime. For the
177: nonlinear analysis, the method of matched asymptotic
178: expansions~\cite{bender,hinch,kevorkian} must be adapted for multiple time
179: scales at different orders of the expansion, so the problem also
180: presents an opportunity for mathematicians to develop time-dependent
181: boundary-layer theory.
182: 
183: We begin in section~\ref{sec:history} by reviewing some of the
184: relevant literature on electrochemical relaxation. In
185: section~\ref{sec:setup} we state the mathematical problem for a
186: suddenly applied DC voltage, and in section~\ref{sec:lin} we analyze
187: the linear response  using Laplace transforms. In
188: section~\ref{sec:nondim} we non-dimensionalize the problem and
189: describe numerical solutions, used to test our analytical
190: approximations. In section~\ref{sec:nonlin}, we derive uniformly valid
191: asymptotic expansions in the ``weakly nonlinear'' limit of thin double
192: layers and discuss the connection with circuit models. Apparently for
193: the first time (for this problem), in section~\ref{sec:higher} we 
194: analyze higher-order corrections, and in section~\ref{sec:strong} we
195: briefly discuss the ``strongly nonlinear'' regime at large voltages,
196: where the expansions are no longer valid.  In section~\ref{sec:concl},
197: we conclude by briefly discussing extensions to higher dimensions, general
198: electrolytes, and Faradaic processes and posing some open questions.
199: 
200: 
201: \section{ Historical Review }
202: \label{sec:history}
203: 
204: \subsection{ Electrical Circuit Models }
205: 
206: In electrochemistry, the most common theoretical approach is to
207: construct an equivalent electrical circuit, whose parameters are fit
208: to experimental impedance spectra or pulsed-voltage responses, as
209: recently reviewed by Macdonald~\cite{macdonald90} and
210: Geddes~\cite{geddes97}. The basic idea of an equivalent circuit is
211: apparently due to Kohlrausch~\cite{kohlrausch1873} in 1873, and the
212: first mathematical theory of Kohlrausch's ``polarization capacitance''
213: was given by Warburg at the end of the nineteenth
214: century~\cite{warburg1899,warburg1901}. Warburg argued that AC
215: electrochemical response is dominated by pure diffusion of the active
216: species and can be described a bulk resistance in series with a
217: frequency-dependent capacitance, which combine to form the ``Warburg
218: impedance''. 
219: 
220: Earlier, Helmholtz~\cite{helmholtz1853,helmholtz1879} had
221: suggested that the solid-electrolyte interface acts like a thin
222: capacitor, for which he apparently coined the term, ``double
223: layer''~\cite{bard}. In 1903 Kr\"uger~\cite{kruger1903} unified
224: Warburg's bulk impedance with Helmholtz' double-layer capacitor in the
225: first complete AC circuit model for an electrochemical cell, which
226: forms the basis for the modern ``Randles circuit''~\cite{randles47}.
227: In this context, the relaxation time for charging of the double layers
228: has been known to depend on the electrode separation, via the bulk
229: resistance, for at least a century.
230: 
231: The study of diffuse charge in the double layer was initiated in the
232: same year by Gouy~\cite{gouy1903}, who suggested that excess ionic
233: charge in solution near the electrode could be viewed as a
234: capacitance, $C_D = \varepsilon/\lambda_D$.  He was also the first to
235: derive Equation~(\ref{eq:lambda}) for $\lambda_D$ (obviously with
236: different notation) in his original theory of the diffuse double layer
237: in equilibrium~\cite{gouy1909,gouy1910}. With the availability of
238: Einstein's relation~\cite{einstein1905} for the mobility, $\mu =
239: D/kT$, at that time, the DC bulk resistance (per unit area) could have
240: been calculated as
241: \begin{equation}
242: R_b = \frac{V }{J} = \frac{L E_0}{\sigma_b E_0} = \frac{\lambda_D^2
243: L}{\varepsilon D}  \label{eq:Rb}
244: \end{equation}
245: (for a symmetric binary electrolyte of equal mobilities), where $J$ is
246: the current density and 
247: \begin{equation}
248: \sigma_b = \frac{\varepsilon D}{\lambda_D^2} = \frac{ 2(ze)^2 C_0
249:   D}{kT}  \label{eq:Kb}
250: \end{equation}
251: is the bulk conductivity. Therefore, basic time scale in
252: Eq.~(\ref{eq:tmixed}) has essentially been contained in circuit models
253: since roughly 1910 as the relaxation time,
254: \begin{equation}
255: \tau_c = R_b C_D = \frac{\lambda_D L}{D} = \frac{C_D L}{\sigma_b} =
256: \frac{\varepsilon L}{\lambda_D \sigma_b}
257: \end{equation}
258: although $\tau_c$ was not stated explicitly
259: as $\lambda_D L /D$ for perhaps another fifty years
260: ~\cite{macdonald70}.
261: 
262: Today, Gouy's screening length bears the name of Debye, who rederived
263: it in 1923 as part of his seminal work with
264: H\"uckel~\cite{debye23a,debye23b} on charge screening in bulk
265: electrolytes, using an equivalent formalism. Debye and H\"uckel solved
266: for the spherical screening cloud around an ion, and, due to the low
267: potentials involved, they linearized the transport equations, allowing
268: them to handle general electrolytes. When Gouy~\cite{gouy1910}
269: considered the identical problem of screening near a flat, blocking
270: electrode more than a decade earlier, he obtained exact solutions to
271: full nonlinear equations for the equilibrium potential profile in
272: several cases of binary electrolytes, $z_+/z_- = 1$, $2$, and
273: $\frac{1}{2}$, where $z_+$ and $z_-$ are the cation and anion charge
274: numbers, respectively. 
275: 
276: A few years later, Chapman~\cite{chapman1913} independently derived
277: Gouy's solution for a univalent electrolyte, $z_+ = z_- = 1$, the
278: special case of ``Gouy-Chapman theory'' for which they are both
279: primarily remembered today. Chapman also gave a simple form for the
280: charge-voltage relation of the diffuse-layer capacitor in this case,
281: which, upon differentiation, yields a simple formula for the nonlinear
282: differential capacitance of the diffuse layer,
283: \begin{equation}
284: C_D(\zeta) = \frac{\varepsilon}{\lambda_D} \,
285: \cosh\left(\frac{ze\zeta}{kT}\right), \label{eq:C_GC}
286: \end{equation}
287: where $\zeta$ is the voltage across the diffuse layer in thermal
288: equilibrium. (Here, we include the trivial extension to a general
289: $z$:$z$ electrolyte.)  Combining Eqs.~(\ref{eq:Rb}) and
290: (\ref{eq:C_GC}), we also obtain the basic relaxation time, $\tau_c$,
291: in Eq.~(\ref{eq:tmixed}) multiplied by a potential-dependent factor in
292: the usual case of nonzero equilibrium zeta potential (in the absence
293: of an applied voltage). This factor may be neglected in the
294: Debye-H\"uckel limit of small potentials, $\zeta \ll kT/ze$, but it
295: becomes important at large potentials and generally slows down the
296: final stages of double-layer charging.
297: 
298: More sophisticated models of the double layer were proposed by many
299: subsequent authors~\cite{delahay,bockris} and incorporated into AC
300: circuit models for electrochemical
301: cells~\cite{bard,sluyters70,parsons90}. Naturally, the original ideas of
302: Helmholtz and Gouy were eventually combined into a coherent
303: whole. In 1924, Stern~\cite{stern24} suggested   
304: decomposing the double layer into a ``compact'' (Helmholtz) part within
305: a molecular distance of the surface and a ``diffuse'' (Gouy) part
306: extending into the solution at the scale of the screening
307: length. Physically, the compact layer is intended to describe ions (at
308: the outer Helmholtz plane) whose solvation
309: molecules are in contact with the surface, although specifically
310: adsorbed ions (themselves in contact with the surface) may also be included~\cite{vorsina39}. Regardless of the precise microscopic
311: picture, however, Stern introduced the compact layer as an intrinsic surface
312: capacitance, which cuts off the divergent capacitance of
313: the diffuse layer, Eq.(\ref{eq:C_GC}), at large zeta potentials.
314: 
315: Using this model of two capacitors in series and neglecting
316: specific adsorption, Grahame~\cite{grahame47} applied Gouy-Chapman
317: theory for the diffuse part and inferred the nonlinear differential
318: capacitance of the compact part from his famous experiments on
319: electrified liquid-mercury drops. Macdonald~\cite{macdonald54} then
320: developed a mathematical model for double layers at metal electrodes
321: by viewing the compact layer as a parallel-plate capacitor, as we do
322: below, although he also allowed its thickness and capacitance to vary
323: due to electrostriction and dielectric saturation~\cite{grahame50}. 
324: The reader is referred to
325: various recent reviews~\cite{macdonald90,geddes97,parsons90}
326: to
327: learn how other effects neglected below, such as specific adsorption
328: and Faradaic processes for non-blocking electrodes, have been included
329: empirically in modern circuit models.
330: 
331: In spite of a century of research, open questions remain about the
332: applicability of circuit models~\cite{geddes97}, and even the most
333: sophisticated fits to experimental data still suffer from
334: ambiguities~\cite{macdonald90}.  One problem is the somewhat arbitrary
335: distinction between the diffuse layers and the bulk electrolyte, which
336: in fact comprise a single, continuous region. Even accepting this
337: partition, it is clear that the non-uniform evolution of ionic
338: concentrations in both regions cannot be fully captured by homogeneous
339: circuit elements~\cite{hollings03}. Another problem is the further
340: partitioning of the double layer into two (or more) poorly defined
341: regions at atomic lengths scales, where macroscopic continuum theories
342: (e.g. for dielectric response) are of questionable
343: validity~\cite{cooper78}.
344: 
345: \subsection{ Microscopic Transport Models }
346: \label{sec:hist2}
347: 
348: An alternative theoretical approach, pursued below, is to solve the
349: time-dependent Nernst-Planck
350: equations~\cite{nernst1888,nernst1889,planck1890} for ionic transport
351: across the entire cell (outside any molecular-scale compact layers)
352: without distinguishing between the diffuse-charge layers and the
353: quasi-neutral bulk.  Because this ``phenomenological''~\cite{korn81}
354: approach requires solving Poisson's equation for the mean-field
355: electrostatic potential (self-consistently generated by the continuum
356: charge density) down to microscopic (and sometimes atomic) length
357: scales, it lacks the thermodynamic justification of traditional
358: macroscopic theories based on bulk electroneutrality and
359: electrochemical potentials~\cite{newman}. Nevertheless, it addresses
360: time-dependent charge-relaxation phenomena, which do occur in real
361: systems, with fewer {\it ad hoc} assumptions than circuit models and
362: thus may be considered closer to first principles. The use of the
363: Nernst-Planck equations at scales smaller than the screening length
364: (but still larger than atomic dimensions) is also supported by the
365: success of Gouy-Chapman theory in predicting the diffuse-layer
366: capacitance in a number of experimental systems
367: (e.g. Refs.~\cite{grahame47,macdonald54}), because the theory is based
368: on the steady-state Nernst-Planck equations for thermal equilibrium.
369: The main difficulty in working with the Nernst-Planck equations, aside
370: from mathematical complexity, is perhaps in formulating appropriate
371: boundary conditions at the electrode surface, just outside any compact
372: layers.
373: 
374: Although the response to a suddenly applied DC voltage has been
375: considered by a few authors in the
376: linear~\cite{lemay53,buck69,korn77b} and
377: nonlinear~\cite{korn81,korn78} regimes, as we also do below, much more
378: analysis has been reported for the case of weak AC forcing, where the
379: equations are linearized and the time dependence is assumed to be
380: sinusoidal. These simplifications are made mainly for analytical
381: convenience, although they have direct relevance for the
382: interpretation of impedance spectra. An early analysis of this type
383: was due to Ferry~\cite{ferry48}, who considered the response of a
384: semi-infinite electrolyte to an oscillating charge density applied at a
385: electrode surface. Ferry's treatment is formally equivalent to the
386: classical theory of dielectric dispersion in bulk
387: electrolytes~\cite{debye28,falkenhagen34}.  Naturally, in both cases
388: the same time scale, $\tau_D = \lambda_D^2/D$, arises, and the
389: relaxation of the double layer has no dependence on the macroscopic
390: geometry.
391: 
392: Ferry's analysis of a single electrode is consistent with the common
393: intuition that double-layer charging should be a purely microscopic
394: process, but one might wonder how the electrode could draw charge
395: ``from infinity'' when an infinite electrolyte has infinite
396: resistance. Indeed, as emphasized by Buck~\cite{buck69} and
397: Macdonald~\cite{macdonald70} and confirmed by detailed comparisons
398: with experimental impedance spectra, Ferry's analysis is fundamentally
399: flawed, starting from the boundary conditions: It is not possible to
400: control the microscopic charge density at an electrode surface and
401: neglect its coupling to bulk transport processes; instead, one imposes
402: a voltage relative to another electrode and observes the resulting
403: current (or vice versa), while the surface charge density evolves
404: self-consistently.  
405: 
406: Buck~\cite{buck69} eventually corrected Ferry's analysis to account
407: for the missing ``IR drop'' across two electrodes which imposes the
408: initial surface charge density (or, alternatively, voltage). Nevertheless, the
409: physical picture of a double-layer responding locally to ``charge
410: injection'', independent of bulk transport processes, persists to the
411: present day.  For example, recent textbooks on colloidal science
412: (Hunter~\cite{hunter}, p.\ 408; Lyklema~\cite{lyklema}, p.\ 4.78)
413: present a slightly different version of Ferry's analysis (attributed
414: to O'Brien) as the canonical problem of ``double-layer relaxation'':
415: the response of a semi-infinite electrolyte to a suddenly imposed,
416: constant surface charge density. This gives some insight into high
417: frequency dielectric dispersion of non-polarizable colloids (the usual
418: case), but it is not relevant for polarizable particles and
419: electrodes. Three decades after Buck and Macdonald, it is worth
420: re-emphasizing the fundamental coupling of double-layer charging of
421: bulk transport in finite, polarizable systems.
422: 
423: The mathematical theory of AC response for a finite, two-electrode
424: system began with Jaff\'e's analysis for
425: semiconductors~\cite{jaffe33,jaffe52} and was extended to liquid
426: electrolytes by Chang and Jaff\'e~\cite{chang52}. A number of
427: restrictive assumptions in these studies, such as a uniform electric
428: field, were relaxed by Macdonald~\cite{macdonald53} for semiconductors
429: and electrolytes and independently by Friauf~\cite{friauf54} for ionic
430: crystals. These authors, who gave perhaps the first complete
431: mathematical solutions, also allowed for bulk generation/recombination
432: reactions, which are crucial for electrons and holes in
433: semiconductors. Subsequent authors mostly neglected bulk reactions in
434: studies of
435: liquid~\cite{macdonald70,buck69,baker68,macdonald74a,macdonald74b} and
436: solid~\cite{beaumont67,korn77b} electrolytes, while focusing on other
437: effects, such as arbitrary ionic valences and the compact layer.
438: 
439: Although it is implicit in earlier work, Macdonald~\cite{macdonald70}
440: first clearly identified the geometry-dependent time scale, $\tau_c =
441: \lambda_D L/D$, (in this form) as governing the relaxation of an
442: electrochemical cell. It was also derived independently by Kornyshev
443: and Vorontyntsev~\cite{korn77b,korn78} in the Russian literature on
444: solid electrolytes with one mobile ionic species~\cite{korn81}. With
445: Itskovich~\cite{korn77a}, these authors also modeled the compact-layer
446: capacitance via a mixed Dirichlet-Neumann condition on the
447: Nernst-Planck equations. This classical boundary
448: condition~\cite{bard}, also used below, introduces another length
449: scale, $\lambda_S$, the effective width of the Stern layer, which also
450: affects the time scales for electrochemical relaxation.
451: 
452: Other important surface properties have also been included in
453: mathematical analyses of AC response. For example, several recent
454: studies of blocking electrodes have included the effect of a nonzero
455: equilibrium zeta potential (away from the point of zero
456: charge)~\cite{hollings03,gunning95,scott00a,scott00b}, building on the
457: work of Delacey and White~\cite{delacey82}. A greater complication is
458: to include Faradaic processes at non-blocking electrodes through
459: boundary conditions of the Butler-Volmer
460: type~\cite{newman,delahay,bockris}, as suggested by
461: Levich~\cite{levich49} and Frumkin~\cite{frumkin55}. This approach has
462: been followed in various analyses of AC response around base states of
463: zero~\cite{korn81,korn77a,korn78,macdonald76a,macdonald76b,macdonald77}
464: and nonzero~\cite{bonnefont01,bonnefont_thesis} steady Faradaic
465: current. Numerical solutions of the time-dependent Nernst-Planck
466: equations have also been developed for AC response and more general
467: situations~\cite{delacey82,brumleve78,murphy92}, following the
468: original work of Cohen and Cooley~\cite{cohen65}.
469: % The effect of surface
470: % roughness has also been studied recently~\cite{pajkossy94}, although still
471: % relatively little mathematical analysis of electrochemical relaxation in two or
472: % three dimensions has been done.
473: 
474: 
475: \subsection{ Colloids and Microfluidic Systems }
476: \label{sec:colloids}
477: 
478: Diffuse-charge dynamics occurs not only near electrodes, but also
479: around colloidal particles and in microfluidic systems, where the
480: coupling with fluid flow results in time-dependent, nonlinear
481: electrokinetic phenomena. This review may be the first to unify some
482: of the fairly disjoint literatures on diffuse-charge dynamics in these
483: areas with the older literature in electrochemistry discussed
484: above. Compared to the latter, more sophisticated mathematical
485: analyses are often done in colloidal science and
486: electro-microfluidics, starting from the Nernst-Planck equations for
487: ion transport and the Navier-Stokes equations for fluid mechanics in
488: two or three dimensions. On the other hand, with the notable exception
489: of the Ukrainian school~\cite{dukhin93,lyklema,dukhin02,murtsovkin96}, less
490: attention is paid to surface properties, and simple boundary
491: conditions are usually assumed (constant zeta potential and complete
492: blocking of ions) which exclude diffuse-charge dynamics.
493: 
494: This might explain why the material time scale, $\tau_D$, is
495: emphasized as the primary one for double-layer relaxation around
496: colloidal particles~\cite{hunter,russel,lyklema}, although the mixed
497: time scale, $\tau_c = (L/\lambda_D)\tau_D$, has come to be recognized
498: as controlling bulk-field screening by
499: electrodes~\cite{hollings03,gunning95,scott00a,scott00b}. This
500: thinking can be traced back to the seminal work of Debye and
501: Falkenhagen~\cite{debye28,falkenhagen34} on dielectric dispersion in
502: bulk electrolytes, mentioned above. In that context, when a background
503: field $E_b$ is applied to an electrolyte, the relevant geometrical
504: length is the size of the screening cloud around an ion, $L =
505: \lambda_D$, over which a voltage, $E_b \lambda_D$, is effectively
506: applied. The relevant RC time for the polarization of the
507: screening cloud is then, $\tau_c = \lambda_D \lambda_D/D =
508: \tau_D$. The possible role of geometry is masked by the presence of
509: only one relevant length scale, $\lambda_D$.
510: 
511: For colloidal particles, which are usually much larger than the
512: double-layer thickness, the second time scale, $\tau_a = a^2/D$, for
513: bulk diffusion around a particle of radius, $a$, becomes important,
514: especially in strong fields. If there is significant surface conduction
515: or the particle is conducting, the ``RC'' time scale, $\tau_c =
516: \lambda_D a/D$, can also become important. In general, double-layer
517: relaxation is thus sensitive to the size and shape of the
518: particle. Although it is largely unknown (and rarely cited) in the
519: West, many effects involving non-uniform double-layer polarization
520: around colloidal particles have been studied 
521: under the name, ``non-equilibrium electric surface
522: phenomena''~\cite{dukhin80}, as recently reviewed by S.\ S.\
523: Dukhin~\cite{dukhin93,dukhin02}.
524: 
525: The colloidal analog of our model problem involving a blocking
526: electrochemical cell is that of an ideally polarizable, metal particle
527: in a suddenly applied background electric field. This situation has
528: received much less attention than the usual case of non-conducting
529: particles of fixed surface charge density, but it has an interesting
530: history.  The non-uniform polarization of the double layer for a metal
531: particle was perhaps first described by Levich~\cite{levich}, using
532: Helmholtz' capacitor model.  Simonov and
533: Shilov~\cite{simonov73a,simonov77} later considered diffuse charge and
534: showed that the metal particle acquires an induced dipole moment
535: opposite to the field over the time scale, $\tau_c = \lambda_D a /D$,
536: as bulk conduction transfers charge from the part of the double-layer
537: facing away from the field to the part facing toward the field. The
538: two hemispheres may be viewed as capacitors coupled through a
539: continuous bulk resistor~\cite{simonov77}, as in the RC circuit model
540: of DC electrochemical cells described above. The charging process
541: continues until the redistribution of diffuse-charge completely
542: eliminates the normal component of the electric field, responsible for
543: charging the double layer.
544: 
545: Diffuse-charge dynamics is important in the context of colloids
546: because it affects electrokinetic phenomena. In the metal-sphere
547: example, the remaining tangential component of the field interacts
548: with the non-uniform induced diffuse charge (and zeta potential) to
549: cause nonlinear electro-osmotic flows~\cite{murtsovkin96,gamayunov86},
550: which causes hydrodynamic interactions between colloidal
551: particles. Although these flows have little effect on the
552: electrophoresis of charged polarizable particles in uniform DC
553: fields~\cite{simonov73b,simonova76b}, they significantly affect
554: dielectrophoresis in nonuniform AC fields~\cite{shilov81,simonova01},
555: where the time-dependence of double-layer relaxation also plays an
556: important role. 
557: 
558: These developments followed from pioneering studies of S.\ S.\ Dukhin, B.\ V.\ Deryagin, and collaborators~\cite{dukhin80,dukhin93,deryagin80,dukhin74} 
559: on the effects of surface conduction and concentration
560: gradients on electrical polarization and electro-osmotic flows around
561: highly charged non-conducting particles, which was also extended to
562: polarizable particles~\cite{murtsovkin96}. (Similar ideas were also
563: pursued later in the Western literature, with some new
564: results~\cite{hunter,obrien83,hinch84,prieve84,anderson89}.)  Earlier still,
565: Bikerman~\cite{bikerman33,bikerman35,bikerman40} presented the
566: original theory of surface conduction in the double layer, and
567: Overbeek~\cite{overbeek43} first calculated in detail the effect
568: of non-equilibrium double-layer polarization on electrophoresis.
569: %  Some
570: % of these exotic NESP involving diffuse-charge dynamics may have
571: % relevance for recent experiments on colloidal self-assembly near
572: % electrodes in AC fields~\cite{yeh97,trau97}.
573: 
574: Diffuse-charge dynamics has begun to be exploited in microfluidic
575: devices, albeit without the benefit of the prior literature in
576: electrochemistry and colloidal science discussed above. In a series of
577: recent papers, Ramos and collaborators have predicted and observed
578: ``AC electro-osmosis'' at a pair of blocking
579: micro-electrodes~\cite{ramos98,ramos99,green00a,gonzalez00,green02}. Their
580: simple explanation of double-layer dynamics~\cite{ramos98,ramos99},
581: supported by a mathematical analysis of AC response in two
582: dimensions~\cite{gonzalez00}, is similar to that of Simonov and Shilov
583: for a metal particle in an AC field~\cite{simonov77}, and the
584: resulting electro-osmotic flow is of the type described by Gamanov et
585: al.\ for metal particles~\cite{gamayunov86}. An important difference,
586: however, is that AC electro-osmosis occurs at fixed micro-electrodes,
587: whose potentials are controlled externally, as opposed to free
588: colloidal particles.  Ajdari~\cite{ajdari00} has proposed a similar
589: means of pumping liquids using AC voltages applied at an array of
590: micro-electrodes, where broken symmetries in surface geometry or
591: chemistry generally lead to net pumping past the array, as observed in
592: subsequent experiments~\cite{brown01,studer02,mpholo03,ramos03}. These
593: are all examples of the general principle of ``induced-charge
594: electro-osmosis''~\cite{bazant03,squires03}, where diffuse-charge
595: dynamics at polarizable surfaces (not necessarily electrodes) is used
596: to drive micro-flows with AC or DC forcing. Clearly, the full range of
597: possible microfluidic applications of time-dependent nonlinear
598: electrokinetics has yet to be explored.
599: 
600: 
601: \subsection{ The Limit of Thin Double Layers}
602: 
603: All of the analytical studies cited above that go beyond linear
604: response (and most that do not) are based on the thin-double-layer
605: approximation, $\lambda_D \ll L$. In this limit, the bulk electrolyte
606: remains quasi-neutral, and the double layer remains in thermal
607: quasi-equilibrium, even with time dependent forcing (slower than
608: $\tau_D$)~\cite{newman,lyklema,hunter,russel,dukhin93}. The same limit
609: also justifies the general notion of circuit models for the diffuse
610: part of the double layer and, in the absence of concentration
611: gradients, the neutral bulk region.
612: 
613: As first shown by Grafov and Chernenko~\cite{grafov62,chernenko63} in
614: the Soviet Union and by Newman~\cite{newman66} and
615: Macgillivray~\cite{macgillivray68} in the United States, the thin
616: double-layer approximation for electrochemical cells can be given
617: ``firm'' (but not necessarily ``rigorous'') mathematical justification
618: by the method of matched asymptotic
619: expansions~\cite{bender,hinch,kevorkian} in the small parameter,
620: $\epsilon = \lambda_D/L$. For steady Faradaic conduction, the usual
621: leading-order approximation involves a neutral bulk with charged
622: boundary layers of $O(\epsilon)$ dimensionless width, which has since
623: been established rigorously in a number of studies by
624: mathematicians~\cite{rubinstein90,henry89,louro91,henry95,park97,barcilon97}.  The
625: standard asymptotic approximation breaks down, however, near Nernst's
626: diffusion-limited current, where the concentration at the cathode
627: vanishes. At the limiting current~\cite{smyrl67}, the boundary layer
628: expands to $O(\epsilon^{2/3})$ width, while at still larger
629: currents~\cite{rubinstein79}, a layer of ``space charge'' extends out
630: to $O(1)$ distances into the bulk region, although the effect of realistic
631: boundary conditions (Faradaic processes, compact layer, etc.) remains
632: to be studied in these exotic regimes. Matched asymptotic expansions
633: are also beginning to be used for time-dependent electrochemical
634: problems with Faradaic processes~\cite{bonnefont01,bonnefont_thesis}
635: below the limiting current.
636: 
637: Perhaps because it originated in fluid mechanics~\cite{kevorkian}, the
638: method of matched asymptotic expansions has been used extensively
639: in colloidal science and microfluidics
640: ~\cite{gonzalez00,dukhin93,dukhin74,obrien83,hinch84,prieve84,anderson89,dukhin69,dukhin70},
641: albeit with varying degrees of mathematical rigor. In any case, the
642: advantages of the technique are (i) to justify the assumption 
643: of equilibrium
644: structure for the double layers (at leading order), regardless of
645: transport processes in the neutral bulk, and (ii) to view the double
646: layers as infinitely thin at the bulk length scale, which is
647: particularly useful in multidimensional problems. For statics or
648: dynamics at the bulk diffusion time, it is usually possible to
649: construct uniformly valid approximations by adding the inner and
650: outer solutions and subtracting the overlap.
651: 
652: The thin double layer approximation is ``asymptotic'' as $\epsilon
653: \rightarrow 0$, which means that the ratio of the approximation to the
654: exact solution approaches unity for sufficiently small $\epsilon$,
655: with all other parameters held fixed. For any fixed $\epsilon > 0$ (no
656: matter how small), however, the approximation breaks down at
657: sufficiently large voltages. The general
658: criterion
659: \begin{equation}
660: \frac{\lambda_D}{a} \, \cosh\left(\frac{ze\zeta}{2kT}\right)\ll 1
661: \label{eq:valid}
662: \end{equation}
663: is often quoted for the validity of Smoluchowski's formula for the
664: electrophoretic mobility of a thin-double-layer
665: particle~\cite{hunter}, as justified by numerical
666: calculations~\cite{obrien78}. This is related to S. S. Duhkin's
667: seminal work on double-layer distortion around a spherical
668: particle~\cite{dukhin93,dukhin74,dukhin70}: In the case of highly charged
669: particles, $\zeta \gg kT/ze$, the ``Dukhin
670: number'' $\Du$ (which he called ``$Rel$'') controls corrections to
671: the thin-double-layer limit, $\Du=0$.
672: 
673: The Dukhin number is defined as the ratio of the double-layer surface
674: conductivity, $\sigma_s$, to the bulk conductivity, $\sigma_b$, in
675: Eq.~(\ref{eq:Kb}) per geometrical length, $a$: $\Du =
676: \sigma_s/\sigma_b a$.  Although its effect on electrophoresis was
677: first explored in detail by Dukhin, the same dimensionless group was
678: defined a few decades earlier by Bikerman~\cite{bikerman40}, who also
679: realized that it would play a fundamental role in electrokinetic
680: phenomena.  In a symmetric binary electrolyte with equal
681: diffusivities, the Dukhin number can be put in the simple
682: form,
683: \begin{eqnarray}
684: \Du &=& \frac{2\lambda_D(1+m)}{a} \left[ \cosh\left(
685: \frac{ze\zeta}{2kT}\right) - 1 \right] \nonumber \\
686: &  =&  \frac{4\lambda_D(1+m)}{a} \sinh^2\left(
687: \frac{ze\zeta}{4kT}\right).   \label{eq:Du}
688: \end{eqnarray}
689: where 
690: \begin{equation}
691: m = \left(\frac{kT}{ze}\right)^2 \frac{2\varepsilon}{\eta D} 
692: \end{equation}
693: is a dimensionless number giving the relative importance of
694: electro-osmosis compared with electro-migration and diffusion in surface
695: conduction, and $\eta$ is the viscosity. This form is due to Deryagin
696: and Dukhin~\cite{deryagin69}, who generalized Bikerman's original
697: results~\cite{bikerman33,bikerman35} to account for electro-osmotic
698: surface conductance ($m>0$). For $\Du \ll 1$ the
699: double-layer remains in its equilibrium state at constant zeta
700: potential, but for $\Du \gg 1$ it becomes distorted as surface
701: conduction draws current lines into the double layer.  For a detailed   
702: pedagogical discussion, we refer to Lyklema~\cite{lyklema}.
703: 
704: It is interesting to note that (at least at large zeta potentials) the
705: Dukhin number is similar to the ratio of the effective RC
706: time, $\tau_c(\zeta)$, away from the point of zero charge ($\zeta \neq
707: 0$) to the bulk diffusion time, $\tau_a$:
708: \begin{equation}
709: \frac{\tau_c(\zeta)}{\tau_a} = \frac{\lambda_D}{a} \,
710: \cosh\left(\frac{ze \zeta}{kT}\right)
711: \end{equation}
712: where we have used Eqs.~(\ref{eq:Rb})--(\ref{eq:C_GC}).  Moreover, the
713: usual condition (\ref{eq:valid}) for the validity of the thin
714: double-layer approximation in quasi-steady electrokinetic problems is
715: also a statement about time scales: $\tau_c(\zeta) \ll \tau_L$. When
716: this condition is violated, the usual RC charging dynamics is
717: slowed down so much by nonlinearity that bulk diffusion may complicate
718: the picture. Whether this does in fact occur depends on if the
719: nonlinearity is strong enough to cause significant concentration
720: depletion in the bulk for a given geometry and forcing. Understanding
721: this issue requires going beyond leading order in asymptotic analysis,
722: which is not trivial.
723: 
724: In spite of extensive work on the asymptotic theory of diffuse-charge
725: dynamics, difficult open questions remain.  The leading-order
726: thin-double-layer approximation is well understood in many cases, but
727: higher-order corrections have been calculated in only a few heroic
728: instances, such as the asymptotic analysis of diffusiophoresis by
729: Prieve et al.~\cite{prieve84}. Moreover, such detailed analysis has
730: mostly (if not exclusively) been done for quasi-steady problems. For
731: time-dependent problems of double-layer charging, it seems that
732: higher-order terms in uniformly-valid matched asymptotic expansions
733: have never been calculated.  
734: 
735: Even the leading-order behavior is poorly understood when the {\it
736: induced} zeta potential is large enough to violate the condition
737: (\ref{eq:valid}). In that case, the effective Dukhin number varies
738: with time, as the total zeta potential evolves in time and space. On
739: the other hand, the Russian literature on non-equilibrium
740: electro-surface phenomena at large $\Du$ mostly pertains to highly
741: charged particles in weak fields, where the constant {\it equilibrium}
742: zeta potential is large, but the time-dependent induced zeta potential
743: is small.
744: 
745: Below, we begin to explore these issues in the much simpler context of
746: a one-dimensional problem involving parallel-plate electrodes, which
747: excludes surface conduction and electro-osmotic flow.  We shall see
748: that this requires extending standard boundary-layer theory, which
749: deals with multiple length scales, to account for simultaneous
750: multiple time scales.  Before examining the nonlinear theory,
751: however, we state the mathematical model and study its exact solution
752: in the linear limit of small potentials.
753: 
754: 
755: 
756: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
757: \section{ The Basic Mathematical Problem }
758: 
759: \label{sec:setup}
760: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
761: 
762: 
763: As the simplest problem retaining the essential features of
764: diffuse-charge dynamics, we consider a dilute, completely dissociated
765: $z\!:\!z$ electrolyte, limited by two parallel, planar, blocking
766: electrodes at $X=\pm L$. We describe the concentrations of the charged
767: ions by continuum fields, $C_{\pm}(X,\tau)$, which satisfy the
768: Nernst-Planck equations,
769: \begin{equation}
770: \frac{\partial C_\pm}{\partial \tau} = - \frac{\partial}{\partial X} \left[
771: -D \frac{\partial C_\pm}{\partial X} \mp \mu ze C_\pm \frac{\partial
772:   \Phi}{\partial X}
773: \right]   \label{eq:eqdim}
774: \end{equation}
775: (without generation/recombination reactions), where $\Phi$ is the
776: electrostatic potential, which describes the Coulomb interaction in a
777: mean-field approximation.  For simplicity we assume that the diffusion
778: coefficients of the two ionic species are equal to the same constant,
779: $D$, and obtain the mobility, $\mu$, from the Einstein relation, $\mu
780: = D/k T$. The total ionic charge density, $\rho_e$, controls the
781: spatial variation of the potential, $\Phi$, through Poisson's 
782: equation,
783: \begin{equation}
784: - \varepsilon \frac{\partial^2 \Phi}{\partial X^2} = \rho_e =  ze (C_+ - C_-)
785: \label{Poi}
786: \end{equation}
787: where $\varepsilon$ is the dielectric permittivity of the solvent,
788: assumed to be a constant.
789: 
790: As described above, we focus on ``ideally polarizable'' or ``completely
791: blocking'' electrodes without Faradaic processes, so the ionic fluxes have to
792: vanish there:
793: \begin{equation}
794: F_{\pm} = -D \frac{\partial C_{\pm}}{\partial X} \mp \frac{ze D}{k_BT} C_{\pm}
795: \frac{\partial \Phi}{\partial X}
796:  =0, \ \ \mbox{for } X=\pm L.
797: \end{equation}
798: The Faradaic current density, $J = ze(F_+ - F_-)$, also vanishes at
799: the electrodes, although it can be nonzero elsewhere as diffuse charge
800: moves around inside the cell. We also take into account the intrinsic
801: capacitance of the electrode surface through a mixed boundary
802: condition for the potential~\cite{bard,korn77a,bonnefont01}. The surface
803: capacitance may represent a Stern layer of polarized solvent molecules~\cite{stern24}
804: and/or a dielectric coating on the electrode~\cite{macdonald54}. If $V_{\pm}(t)$ is the external
805: potential imposed by the external circuit on the electrode at $X=\pm
806: L$, then we assume
807: \begin{equation}
808: \Phi = V_{\pm} \mp \lambda_S \frac{\partial
809: \Phi}{\partial X} , \ \ \mbox{ at } X = \pm L,
810: \label{BCpot}
811: \end{equation}
812: where $\lambda_S$ is an effective thickness for the compact part of
813: the double layer. For a simple dielectric layer, this is equal to its
814: actual thickness times the ratio, $\varepsilon/\varepsilon_S$, of dielectric
815: constants of the solvent and the Stern layer, $\varepsilon_S$.
816: 
817: In order to study nonlinear effects and avoid imposing a time
818: scale, we consider the response to a step in voltage (a suddenly applied DC voltage),
819: rather than the usual case of weak AC forcing. For times $\tau <0$, no
820: voltage is applied, and we assume no spontaneous charge accumulation at the
821: electrodes.  The initial ionic
822: concentrations are uniform, $C_{\pm}(X,\tau <0)=C_0$. For $\tau> 0$, a
823: voltage difference $2V$ is applied between the two electrodes,
824: $V_{\pm}(\tau >0)=\pm V$, and we solve for the evolution of the
825: concentrations and the potential. As $\tau\rightarrow\infty$, the bulk
826: electric field at the center, $|E(0,\tau)| = \partial \Phi /
827: \partial X$, decays from its initial value, $V/L$, to zero, due to screening by
828: diffuse charge which is transferred from the right side of the cell
829: ($0 < X < 1$) to the left ($-1 < X < 0$). The relaxation is complete
830: when the Faradaic current decays to zero in steady state, from its
831: initial uniform value, $J(X,0) = J_0 = - \sigma_b V/L= -2 (ze)^2 C_0 D V/ kT L$.
832: 
833: 
834: 
835: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
836: \section{ Linear Dynamics }
837: \label{sec:lin}
838: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
839: 
840: \subsection{ Transform Solution for Arbitrary $\lambda_D$, $\lambda_S$,
841: $L$ }
842: 
843: For applied potentials much smaller than the thermal voltage, $V \ll
844: k_BT/ze$, the equations can be linearized, $C_{\pm}= C_0 +\delta
845: C_{\pm}$, so that the ionic charge density,
846: $\rho_e=ze(C_+-C_-)=ze(\delta C_+ -\delta C_-)$, obeys the
847: Debye-Falkenhagen equation~\cite{debye28},
848: \begin{equation}
849: \frac{1}{D} \frac{ \partial \rho_e}{\partial \tau}
850:  \approx \frac{\partial^2 \rho_e}{\partial X^2} - \kappa^2 \rho_e
851: \label{Deb}
852: \end{equation}
853: where $\kappa=\lambda_D^{-1}$ is the inverse screening length.
854: This equation can also be written as a conservation law,
855: \begin{equation}
856: \frac{\partial \rho_e}{\partial \tau} = - \frac{\partial J_e}{\partial X}
857: \end{equation}
858: in terms of the linearized total ionic electrical current,
859: \begin{equation}
860: J_e \approx - D\frac{\partial \rho_e}{\partial X} - D \kappa^2\varepsilon
861: \frac{\partial \Phi}{\partial X}
862: \end{equation}
863: which vanishes at the blocking electrodes, $X=\pm L$.
864: 
865: To solve the model problem, which involves a step-potential in time, it is
866: convenient to use Laplace transforms, defined by
867: \begin{equation}
868: {\hat f}(S) = \int_0^\infty d\tau\,\, e^{-S\tau} f(\tau).
869: \end{equation}
870: As $\rho_e(X)=0$ for $\tau<0$, the Laplace transforms of Eqs.~(\ref{Poi}) and
871: (\ref{Deb}) are
872: \begin{eqnarray}
873:  \frac{\partial^2 \hat \rho_e}{\partial X^2} &=&  k^2 {\hat \rho_e}
874: \label{hDeb} \\
875: -\varepsilon \frac{\partial^2 \hat\Phi}{\partial X^2} & = & {\hat \rho_e}
876: \label{hPoi}
877: \end{eqnarray}
878: where 
879: \begin{equation}
880: k(S)^2= \frac{S}{D} + \kappa^2.
881: \end{equation}
882: The general antisymmetric solution to Eq.~(\ref{hDeb}) is,
883: \begin{equation}
884: \hat{\rho}_e(X,S)= A \sinh(k X)
885: \end{equation}
886: for some constant $A(S)$, which, substituting into Eq.~(\ref{hPoi})
887: and integrating, yields,
888: \begin{equation}
889: - \epsilon_w \frac{\partial \hat{\Phi}}{\partial X}(X,S) =
890: \frac{A}{k}  \cosh(k X) + B,  \label{ES}
891: \end{equation}
892: where the constant $B(S)$, determined by $\hat{J_e}(\pm L,S) = 0$, is given by
893: \begin{equation}
894: B = A k \cosh(\kappa_S L) (\kappa^{-2} - k^{-2}) .
895: \end{equation}
896: Integrating Eq.~(\ref{ES}) again and enforcing antisymmetry yields the
897: Laplace transform of the potential,
898: \begin{equation}
899: {\hat \Phi}(X,S) = - A \frac{\cosh(k L)}{\varepsilon k^2}
900: \left( \frac{\sinh(k X)}{\cosh(k L)} + \frac{k S X}{\kappa^2 D} .
901: \right)
902: \end{equation}
903: The remaining constant,
904: \begin{equation}
905: A= \frac{-k^2 \varepsilon V S^{-1} \sech(k L) } {
906: \tanh(kL)+\lambda_S k +\frac{kSL}{\kappa^2 D}
907: \left(1+\frac{\lambda_S}{L}\right) }
908: %\right)^{-1}
909: \end{equation}
910: is determined by the Stern-layer boundary condition,
911: Eq.~(\ref{BCpot}).
912: 
913: \subsection{ Long-time Exponential Relaxation }
914: 
915: There is a great deal of information about transients in the Laplace
916: transform of exact solution to the linear problem. For times much
917: smaller than the Debye time, $\tau \ll \tau_D = \lambda_D^2/D$ (or $S
918: \gg \kappa^2 D$), there is no significant response, so we are mainly
919: interested in the response at longer times, $\tau \gg \tau_D$ (or $S
920: \ll \kappa^2 D$). There are many ways to see that this is generally an
921: exponential relaxation dominated by the mixed time scale discussed
922: above, $\tau_c = \lambda_D L/D$, although several other time scales
923: allowed by dimensional analysis also play a role.
924: 
925: \subsubsection{ Diffuse Charge Density at a Surface }
926: 
927: Let us focus on one quantity, for example, the Laplace transform of
928: the charge density at the anode, ${\hat \rho_e}(X=L,S)$.  The exact
929: formula is
930: \begin{equation}
931: {\hat \rho_e}(L,S) = A \sinh(k L) ,
932: %= \frac{-k^2 \epsilon_w V S^{-1} \tanh(k L) }
933: %{ \tanh(kL)+\lambda_S k +\frac{kSL}{\kappa^2 D} \left(1+\frac{\lambda_S}{L}\right)
934: %}.
935: \end{equation}
936: which is difficult to invert analytically. (Keep in mind that $k$
937: depends on $S$.) For times much longer than the Debye time, we
938: consider the limit, $S \ll \kappa^2 D$, in which the Laplace transform
939: takes the much simpler asymptotic form,
940: \begin{equation}
941: {\hat \rho_e}(L,S) \sim \frac{K_{\rho}\, S^{-1}}{1 + \tau_{\rho} S}
942: \end{equation}
943: where
944: \begin{equation}
945: K_{\rho} = -\frac{\varepsilon V \kappa^2}{1 + \kappa \lambda_S \coth(\kappa L) }
946: \end{equation}
947: and
948: \begin{equation}
949: \tau_{\rho} = \frac{L}{\kappa D} \, \left[ \frac{\coth(\kappa L)\left(1 +
950: \frac{3\lambda_S}{2L}\right) - \frac{1}{2} \kappa \lambda_S \csch^2(\kappa L)
951: - \frac{1}{\kappa L}}{1 + \kappa \lambda_S \coth(\kappa L) } \right]
952: \label{eq:taucrho}
953: \end{equation}
954: Since the Laplace transform of $1-\exp(-\tau/\tau_o)$ is
955: $S^{-1}/(1+S\tau_o)$, this result clearly shows that the buildup of
956: the charged screening layer occurs exponentially over a
957: characteristic response time given by Eq.~(\ref{eq:taucrho}),
958: %  In the limit of thin Debye
959: % layers $L \gg \kappa^{-1}$, this formula simplifies further,
960: % \begin{equation}
961: % {\hat \rho_e}(z=L,S\to 0) \simeq -\frac{\kappa^2 \epsilon_w
962: % V}{(1+\lambda_s\kappa)}\,\, \frac{S^{-1}} { 1+S \frac{L
963: % (1+\lambda_s/L)}{\kappa D(1+\lambda_s\kappa)}}
964: % \end{equation}
965: which is of order, $L/\kappa D = \lambda_D L /D = \tau_c$, for both thin and
966: thick double layers. Corrections introduce other mixed scales involving the
967: Stern length, such as $\lambda_S L/D$ and $\lambda_S \lambda_D/D$, as well as
968: the Debye time, $\lambda_D^2/D$.
969: 
970: 
971: Note that the same time scale can also be seen in the linear response to a weak
972: oscillatory potential, $V_{\pm}=\pm V \re (e^{i\omega \tau})$, which naturally
973: leads to
974: \begin{equation}
975: \rho_e(L,\tau) \sim K_\rho \, \re\left( \frac{e^{i\omega\tau}}{1 +
976:   i\omega\tau_\rho} \right)
977: \end{equation}
978: for frequencies well below the Debye frequency, $\omega \ll \omega_D =
979: D/\lambda_D^2$. Similar results for AC response near the point of zero charge
980: have been obtained by many authors, as cited above. The characteristic
981: frequency, $\omega_c = 1/\tau_c \approx D/\lambda_D L$, also arises the context
982: of AC electro-osmotic fluid pumping near
983: micro-electrodes~\cite{ramos98,ajdari00}, because diffuse-layer charging
984: controls the time-dependence of the effect.
985: 
986: 
987: \subsubsection{ Total Diffuse Charge in an Interface }
988: 
989: We now show that the same form of long-time exponential relaxation, with a
990: somewhat {\it different} characteristic time, also holds for other quantities,
991: such as the total diffuse charge near the cathode,
992: \begin{equation}
993: Q(t) = \int_{-L}^0 \rho(X,t) dX ,
994: \end{equation}
995: which plays a central role in the nonlinear analysis below. In the
996: limit of thin double layers, this is simply the total interfacial
997: charge (per unit area) of the diffuse part of the double layer. Here
998: we consider the total diffuse charge near a surface more generally,
999: even when the Debye screening length is much larger than the electrode
1000: separation. In the latter case, the concept of an ``interface'' is not
1001: well defined, since the two sides of the cell interact very strongly,
1002: but we can still study the overall separation of diffuse charge caused
1003: by the applied voltage.
1004: 
1005: Using Eqs.~(\ref{hPoi}) and (\ref{ES}), the Laplace transform of the
1006: total cathodic charge is,
1007: \begin{equation}
1008: \hat{Q}(S) = A k^{-1}\left[1 - \cosh(k L)\right] .
1009: \end{equation}
1010: Once again, this is difficult to invert analytically, so we focus on
1011: the long-time limit,
1012: \begin{equation}
1013: \hat{Q}(S) \sim \frac{K_Q S^{-1}}{1 + \tau_Q S} ,
1014: \end{equation}
1015: for $S \ll \kappa^2 D$, where
1016: \begin{equation}
1017: K_Q = \frac{ \varepsilon V \kappa \left[1 - \sech(k L)\right]
1018: }{\tanh(\kappa L)  + \kappa \lambda_S}
1019: \end{equation}
1020: and
1021: \begin{eqnarray}
1022: \tau_Q &=& \frac{L}{\kappa D} \left\{ \frac{1 + \frac{1}{2}\sech^2(\kappa
1023:       L) + \frac{3\lambda_S}{2L}}{\tanh(\kappa L) + \kappa \lambda_S}
1024:        \right. \nonumber \\
1025: & & \left. - \frac{\sech(\kappa L)\tanh(\kappa
1026:       L)}{2\left[1 - \sech(\kappa L)\right]} - \frac{1}{2\kappa L} \right\}   \label{eq:taucQ}
1027: \end{eqnarray}
1028: In the limit of thin double layers, the same basic time scale, $\tau_c
1029: = L/\kappa D = \lambda_D L /D$, arises as in the case of the surface
1030: charge density. A subtle observation is that the relaxation of the
1031: total interfacial charge, although still exponential, has a somewhat
1032: different time scale as a function of $\epsilon = \lambda_D/L$ and
1033: $\delta = \lambda_S/\lambda_D$. (See Fig.~\ref{fig:tc} below.) This
1034: apparently new result shows that charging dynamics has a nontrivial
1035: dependence on time and space, even for very weak potentials.
1036: 
1037: \begin{figure}
1038: \includegraphics[width=3in]{fig2.eps}
1039: \caption{ Analytical results for the exponential relaxation time from
1040:   the linear theory for weak applied potentials, ($V \ll
1041:   kT/ze$). The time scale for relaxation of the surface diffuse-charge
1042:   density, $t_\rho$, from Eq.~(\protect\ref{eq:tcrho}) is shown in (a)
1043:   and that of the total interfacial (half-cell) diffuse charge,
1044:   $t_Q$, from Eq.~(\protect\ref{eq:tcQ}) in (b). In each case, the
1045:   charging time, scaled to $\tau_c=L/\kappa D =\lambda_D L/D$, is
1046:   plotted versus the dimensionless diffuse-layer thickness, $\epsilon
1047:   = \lambda_D/L$, for different dimensionless Stern-layer thicknesses,
1048:   $\delta = \lambda_S/\lambda_D = 0, 0.1, 1, 10$ (solid, dot, dash, and dot-dash lines, respectively).
1049:   \label{fig:tc} }
1050: \end{figure}
1051: 
1052: 
1053: 
1054: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1055: \section{ Dimensionless Formulation and Numerical Solution }
1056: \label{sec:nondim}
1057: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1058: 
1059: \subsection{ Basic Equations }
1060: \label{sec:eqns}
1061: 
1062: In preparation for analysis of the full,
1063: nonlinear problem, we cast it in a dimensionless form using $L$ as the
1064: reference length scale and $\tau_c=\lambda_D L/D$ as the reference
1065: time scale, as motivated by the linear theory. Time and space are then
1066: represented by $t=\tau D/\lambda_D L$ and $x=X/L$, and the problem
1067: is reformulated through reduced variables: $c=(C_+ +C_-)/2C_0$ for the
1068: local salt concentration, $\rho= (C_+-C_-)/2C_0=\rho_e/(2C_0ze)$ for
1069: the charge density, and $\phi=ze\Phi/k_BT$ for the electrostatic
1070: potential. The solution is determined by only three dimensionless
1071: parameters: $v=zeV/k_BT$, the ratio of the applied voltage to the
1072: thermal voltage, $\epsilon=\lambda_D/L$, the ratio of the Debye length
1073: to the system size, and $\delta=\lambda_S/\lambda_D$, the ratio of the
1074: Stern length to the Debye length~\cite{bonnefont01}.
1075: 
1076: With these definitions, the dimensionless equations for $-1 < x < 1$ and
1077: $t>0$ are
1078: \begin{eqnarray}
1079: \frac{\partial c}{\partial t} &=& \epsilon\, \frac{\partial}{\partial x}\left(
1080: \frac{\partial c}{\partial x} + \rho \frac{\partial \phi}{\partial x}\right)
1081: \label{eq:c} \\ \frac{\partial \rho}{\partial t} &=& \epsilon\,
1082: \frac{\partial}{\partial x}\left( \frac{\partial \rho}{\partial x} + c
1083: \frac{\partial \phi}{\partial x}\right)  \label{eq:rho} \\ -\epsilon^2 \,
1084: \frac{\partial^2 \phi}{\partial x^2} &=& \rho  \label{eq:phi}
1085: \end{eqnarray}
1086: with boundary conditions at $x = \pm 1$,
1087: \begin{eqnarray}
1088: \frac{\partial c}{\partial x} + \rho \frac{\partial \phi}{\partial x} &=& 0
1089: \label{eq:bcc} \\ \frac{\partial \rho}{\partial x} + c \frac{\partial
1090: \phi}{\partial x} &=& 0 \label{eq:bcrho} \\ v - \delta\,\epsilon\,
1091: \frac{\partial \phi}{\partial x} &=& \pm\phi \label{eq:sternbc}
1092: \end{eqnarray}
1093: and initial conditions, $c(x,0)=1$, $\rho(x,0) = 0$, and $\phi(x,0) =
1094: v\, x$. Note that the limit of a negligible screening length, 
1095: $\epsilon\rightarrow 0$, is {\it singular} because it is impossible to
1096: satisfy all the boundary conditions when $\epsilon=0$. Physically,
1097: this corresponds to the limit of exact charge neutrality, $\rho=0$,
1098: which is always violated to some degree at electrochemical interfaces.
1099: 
1100: The total diffuse charge near the cathode is
1101: \begin{equation}
1102: q(t) = \int_{-1}^0 \rho(x,t) dx ,
1103: \end{equation}
1104: scaled to $2zeC_0 L$. The dimensionless Faradaic current density is,
1105: \begin{equation}
1106: j_F = \frac{\partial \rho}{\partial x} + c \frac{\partial \phi}{\partial x}
1107: ,
1108: \end{equation}
1109: scaled to $2zeC_0D/L$ (Nernst's diffusion-limited
1110: current~\cite{bonnefont01}).
1111: 
1112: \subsection{ Time Scales for Linear Response }
1113: 
1114: The time scale for exponential relaxation of the surface charge
1115: density in the linear theory above, Eq.~(\ref{eq:taucrho}), has the
1116: dimensionless form,
1117: \begin{equation}
1118: t_{\rho} = \frac{(1 + 3\delta\epsilon/2)\,\coth(\epsilon^{-1}) -
1119:     \delta\,\csch^2(\epsilon^{-1})/2 - \epsilon }{
1120:     1+\delta\,\coth(\epsilon^{-1}) } .  \label{eq:tcrho}
1121: \end{equation}
1122: As shown in Fig.~\ref{fig:tc}(a), this formula shows that for a wide
1123: range of diffuse and Stern layer thicknesses, the basic time scale is
1124: always roughly of order, $\lambda_D L/D$, since $t_c$ is of order $1$.  In the
1125: limit of a thin diffuse double layer, the dimensionless time scale
1126: has the  form,
1127: \begin{equation}
1128: t_{\rho} = \frac{1}{1+\delta} +   \left(\frac{3\delta -
1129:   2}{2(1+\delta)}\right) \, \epsilon  + O(e^{-\epsilon^{-1}}).  \label{eq:tcrhothin}
1130: \end{equation}
1131: with exponentially small errors.  In the limit of a thin Stern layer,
1132: the time scale becomes
1133: \begin{eqnarray}
1134: t_{\rho} &=& \coth(\epsilon^{-1}) - \epsilon + \nonumber \\
1135: & &  \left[ 5\epsilon \coth(\epsilon^{-1}) - 2
1136: \coth(\epsilon^{-1}) - \csch^2(\epsilon^{-1}) \right] \,
1137: \frac{\delta}{2} \nonumber\\ & &  + \,
1138: O(\delta^2) .
1139: \end{eqnarray}
1140: For simultaneously thin Stern and diffuse layers, we obtain the simple
1141: result,
1142: \begin{equation}
1143: t_{\rho} \sim 1 - \epsilon - \delta
1144: \end{equation}
1145: which, as in Fig.~\ref{fig:tc}(a), shows that increasing either $\epsilon
1146: = \lambda_D/L$ or $\delta = \lambda_S/\lambda_D$ tends to reduce the
1147: charging time in this limit, compared to the leading-order value,
1148: $\lambda_D L /D$. Putting the units back, this expression can be
1149: written as,
1150: \begin{equation}
1151: \tau_{\rho} \sim \frac{\lambda_D L}{D} - \frac{\lambda_D^2}{D} -
1152: \frac{\lambda_S L}{D}
1153: \end{equation}
1154: for $\lambda_S \ll \lambda_D \ll L$, which clearly shows the Debye
1155: time, $\lambda_D^2/D$, appearing only as a small perturbation of the
1156: intermediate time scale, $\lambda_D L/D$, for the relaxation of the cell.
1157: 
1158: \begin{figure*}
1159: \includegraphics[width=5in]{fig3.eps}
1160: \caption{ Profiles for $t=0$ (solid), $0.1$ (dot), $0.5$ (dash),
1161: $1$ (dot-dash), $2$ (dot-dot-dot-dash), $\infty$ (long dash) of the
1162: dimensionless charge density, $\rho(x,t)$, for dimensionless voltages
1163: (a) $v=0.1$ and (b) $v=1$, and of the dimensionless potential, $\phi$,
1164: for (c) $v=0.1$ and (d) $v=1$ ($\epsilon=0.05$,
1165: $\delta=0.1$).
1166: \label{fig:fields} }
1167: \end{figure*}
1168: 
1169: 
1170: Similar results hold for the relaxation time for the total half-cell
1171: charge, Eq.~(\ref{eq:taucQ}), which has the dimensionless form,
1172: \begin{equation}
1173: t_Q = \frac{1 + \frac{1}{2}\sech^2(\epsilon^{-1}) +
1174:   \frac{3\delta\epsilon}{2}}{\tanh(\epsilon^{-1}) + \delta} -
1175: \frac{\epsilon}{2} - \frac{\sech(\epsilon^{-1})\tanh(\epsilon^{-1})}{2\left[1 -
1176:     \sech(\epsilon^{-1})\right]}   \label{eq:tcQ}
1177: \end{equation}
1178: For thin double layers, we obtain the same leading-order behavior,
1179: \begin{equation}
1180: t_Q \sim \frac{1}{1+\delta} - \left[ \frac{1-2\delta}{2(1+\delta)}
1181:   \right]\,\epsilon +  O(e^{-\epsilon^{-1}}) \label{eq:tcQthin}
1182: \end{equation}
1183: although the correction term is somewhat different for thick diffuse
1184: layers. For simultaneously thin diffuse and Stern layers, the
1185: dimensionless relaxation time for the total charge becomes,
1186: \begin{equation}
1187: t_Q \sim 1 - \frac{\epsilon}{2} - \delta .
1188: \end{equation}
1189: For a detailed summary of how the two time scales, $t_{\rho}$ and $t_Q$,
1190: depend on the parameters, $\epsilon$ and $\delta$, see Figure \ref{fig:tc} (a)
1191: and (b), respectively.
1192: 
1193: 
1194: \subsection{ Numerical Solution}
1195: %\label{sec:num}
1196: 
1197: Our dimensionless model problem, stated in Section ~\ref{sec:eqns}, is
1198: straightforward to solve numerically using finite differences, at
1199: least if $\epsilon$ is not too small. (Ironically, as shown below,
1200: analytical progress is much easier in this singular limit.)  To
1201: resolve the boundary layer where the gradient is large, a variable
1202: size mesh is used, along with second-order-accurate differencing that
1203: accounts for the variable grid sizes.  The third-order Adams-Bashforth
1204: method is used in time.  The number of the grid points and the 
1205: ratio of the smallest to largest grid size are
1206: varied depending on the values of $\epsilon$ and $v$.  
1207: The numerical convergence is verified though multiple runs of
1208: different resolutions, and as a result, up to 1024 points are used
1209: in calculations for higher $v$.
1210: 
1211: To maximize the importance of diffuse charge,
1212: we first consider a rather larger value of $\epsilon$, even for a
1213: micro-electrochemical system, $\epsilon = 0.05$, say for $\lambda_D =
1214: 5$ nm and $L = 0.1 \mu$m. The Stern length is always of molecular
1215: dimensions, so we choose $\lambda_S = 5$ \AA, and thus $\delta =
1216: 0.1$. The time evolutions of the charge and potential are shown in
1217: Fig.~\ref{fig:fields} for $v = 0.1$ and $v=1$. At room temperature
1218: (and $z=1$), these voltages correspond to $V = 2.5$ mV and
1219: $V=25$ mV, respectively, which, when transferred to the diffuse layer
1220: after screening give maximum electric fields of order $10$ V/$\mu$m.
1221: 
1222: \begin{figure}
1223: \includegraphics[width=\linewidth]{fig4.eps}
1224: \caption{ (a) The dimensionless current density, $j(t)$ (in units of
1225: $2zeC_0 D/L$), and (b) the dimensionless total diffuse charge on the
1226: cathodic side of the cell, $q(t)$ (in units of $2 z e C_0 L$), scaled
1227: to $q_o = v/(1+\delta)$, versus dimensionless time, $t$ (in units of
1228: $\tau_c = \lambda_D L/ D$). Numerical results for dimensionless
1229: voltages, $v= 1$ (dot), $2$ (dash), $3$ (dot-dash), and $4$
1230: (dot-dot-dot-dash) are compared with linear dynamics in the thin
1231: double-layer limit: $q(t)/q_o \sim 1 - e^{-(1+\delta) t}$ and $j(t)/v
1232: \sim e^{-(1+\delta) t}$ (solid lines) as $v, \epsilon \rightarrow 0$.
1233: The breakdown of linear theory for $v \geq 1$ is highlighted in (c),
1234: where the data in (b) is replotted for longer times. \label{fig:q} }
1235: \end{figure}
1236: 
1237: 
1238: The current, $j$, and the total cathodic diffuse charge, $q$, are
1239: plotted versus time, $t$, in Fig.~\ref{fig:q} for applied voltages, $v
1240: = 1,2,3,$ and $4$. In all cases, the linearization is accurate at early
1241: times ($t<1$) since the dimensionless voltage across the diffuse layer
1242: remains small ($<1$). For $v=1$, the linear approximation is
1243: reasonable for all times, but for somewhat larger voltages, $v= 2, 3,$ and $4$,
1244: the relative error becomes unacceptable at long times, $t > 1$. 
1245: %I removed a paragraph change here --K 
1246: Not only is the limiting value of the total charge significantly
1247: underestimated, but the dynamics also continues for a longer time,
1248: with a qualitatively different charging profile.  The largest applied
1249: voltage, $v=4$, shows this effect most clearly, as there is a
1250: secondary relaxation at a much larger time scale of order $t =
1251: 1/\epsilon = 20$. Unlike the other cases, which display the expected
1252: steady increase in charge of an RC circuit, for $v\geq 4$ the total
1253: charge quickly reaches a maximum value, after the initial RC charging
1254: process, and then slowly decays toward its limiting value.
1255: 
1256: We are not aware of any previous theoretical prediction of such a
1257: non-monotonic charging profile, so it is a major focus of this work
1258: (in sections ~\ref{sec:higher} and \ref{sec:strong}). It is
1259: reminiscent of the Warburg impedance due to bulk diffusion of
1260: current-carrying ions at the time scale, $\tau_L = L^2/D$, or
1261: $1/\epsilon$ in our units, in (linear) response to Faradaic processes,
1262: which consume or produce them at an electrode. Here, however, there
1263: are no Faradaic processes, so any such bulk diffusion must be related
1264: to the adsorption or desorption of ions in the diffuse part of the
1265: double layer. Moreover, the over-relaxation of the charge density is
1266: part of the {\it non-linear} response to a large applied voltage, so
1267: it will require more sophisticated analytical methods.
1268:  
1269: 
1270: 
1271: 
1272: 
1273: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1274: \section{ Weakly Nonlinear Dynamics }
1275: \label{sec:nonlin}
1276: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1277: 
1278: \subsection{ Asymptotic Analysis for Thin Double Layers }
1279: 
1280: The remarkable robustness of the charging time well into the nonlinear
1281: regime (at least for the primary relaxation phase) can be predicted analytically using matched
1282: asymptotic expansions in the singular limit of thin double layers,
1283: $\epsilon = \lambda_D/L \ll 1$. Most (if not all) previous studies of
1284: time-dependent problems using asymptotic analysis have scaled time to
1285: the diffusion time, $\tau_L = L^2/D$. In this section, we will see how
1286: the correct charging time scale, $\tau_c = \lambda_D L/D$, arises
1287: systematically from asymptotic matching at leading order.  We also
1288: consider, apparently for the first time, the general case of arbitrary
1289: voltage, $v= ze V/k_BT$, and Stern-layer thickness, $\delta =
1290: \lambda_S/\lambda_D$, with a time-dependent zeta-potential (i.e. potential drop
1291: over the diffuse layer). We also
1292: study higher-order corrections, which involve some bulk diffusion at
1293: the time scale, $\tau_L$.
1294: 
1295: As usual, matched asymptotic expansions only produce a series of
1296: ``asymptotic'' approximations to the solution, in the sense that
1297: higher terms in the expansions vanish more quickly than the leading
1298: terms as $\epsilon \rightarrow 0$, with the other parameters, $v$ and
1299: $\delta$, held fixed at arbitrary values. For any fixed $\epsilon >0$
1300: (no matter how small), there could be $\epsilon$-dependent
1301: restrictions on $v$ and $\delta$ for various truncated expansions to
1302: produce accurate approximations. We refer to the regime where such
1303: conditions hold as ``weakly nonlinear'', as opposed to the ``strongly
1304: nonlinear'' regime where the asymptotic expansions break down
1305: (described below in section~\ref{sec:strong}).
1306: 
1307: 
1308: \subsection{ Outer and Inner Expansions }
1309: 
1310: 
1311: We begin by seeking regular asymptotic expansions (denoted by a bar
1312: accent) in the bulk ``outer'' region, e.g.
1313: \begin{equation}
1314: c(x,t) \sim \bar{c}(x,t) = \bar{c}_0 + \epsilon \, \bar{c}_1 +
1315: \epsilon^2\, \bar{c}_2 + \ldots .   \label{eq:cout}
1316: \end{equation}
1317: Substituting such expansions into Eqs.~(\ref{eq:c})--(\ref{eq:phi})
1318: and equating terms order by order yields a hierarchy of
1319: partial-differential equations. At leading order in $\epsilon$, we
1320: find that the bulk concentration does not vary in time, $\bar{c}_0 = 
1321: 1$, simply because the charging time scale, $\tau_c$, is much {\it
1322: smaller} than the bulk diffusion time scale, $\tau_L$. The
1323: leading-order potential is linear, 
1324: \begin{equation}
1325: \bar{\phi}_0 = \bar{j}_0(t)\,  x,  \label{eq:phi0}
1326: \end{equation}
1327: where $\bar{j}_0(0) = v$. Since the leading-order bulk concentration is
1328: uniform, $\bar{j}_0(t)$ is the leading-order current density. The
1329: leading-order charge density,
1330: \begin{equation}
1331: \bar{\rho}_2 = -\frac{\partial^2 \bar{\phi}_0}{\partial x^2}
1332: \end{equation}
1333: vanishes because the leading-order potential, Eq.\ (\ref{eq:phi0}), is harmonic, although
1334: at next order, $O(\epsilon^3)$, a nonzero
1335: bulk charge density, $\bar{\rho}_3$, arises due to concentration polarization. (See below.)
1336: These arguments justify the usual assumption of bulk electroneutrality
1337: to high accuracy, even during interfacial charging, as long as the
1338: dynamics are ``weakly nonlinear''.
1339: 
1340: The regular outer approximations must be matched with singular ``inner''
1341: approximations in the boundary layers.  The problem has the following
1342: symmetries about the origin,
1343: \begin{equation}
1344: \begin{array}{c}
1345: c(-x,t) = c(x,t) \\ 
1346: \rho(-x,t) = -\rho(x,t) \\
1347:  \phi(-x,t) = -\phi(x,t)
1348: \end{array}
1349:  \label{eq:sym}
1350: \end{equation}
1351: so we consider only the boundary layer at the cathode,
1352: $x=-1$, by transforming the equations to the inner coordinate, $y =
1353: (x+1)/\epsilon$:
1354: \begin{eqnarray}
1355: \epsilon \, \frac{\partial \tilde{c}}{\partial t} &=&
1356: \frac{\partial}{\partial y}\left( \frac{\partial \tilde{c}}{\partial y} +
1357: \tilde{\rho} \frac{\partial \tilde{\phi}}{\partial y}\right) \label{eq:cin}
1358: \\ \epsilon \, \frac{\partial \tilde{\rho}}{\partial t} &=&
1359: \frac{\partial}{\partial y}\left( \frac{\partial \tilde{\rho}}{\partial y}
1360: + \tilde{c} \frac{\partial \tilde{\phi}}{\partial y}\right)
1361: \label{eq:rhoin}
1362: \\ -\frac{\partial^2 \tilde{\phi}}{\partial y^2} &= &\tilde{\rho} \label{eq:phiin}
1363: \end{eqnarray}
1364: This scaling removes the singular perturbation in
1365: Poisson's equation, so we can seek regular asymptotic expansions for the
1366: inner approximations (denoted by tilde accents), e.g.
1367: \begin{equation}
1368: c(x,t) \sim \tilde{c}(y,t) = \tilde{c}_0 + \epsilon \, \tilde{c}_1 +
1369: \epsilon^2 \, \tilde{c}_2 + \ldots
1370: \end{equation}
1371: Matching with the bulk approximations {\it in
1372: space} involves the usual van Dyke conditions, e.g.
1373: \begin{equation}
1374: \lim_{y\rightarrow\infty} \tilde{c}(y,t) \sim \lim_{x\rightarrow -1}
1375: \bar{c}(x,t) ,
1376: \end{equation}
1377: which implies $\tilde{c}_0(\infty,t) = \bar{c}_0(-1,t)$,
1378: $\tilde{c}_1(\infty,t) = \bar{c}_1(-1,t)$, etc., but we will also have
1379: to make sure that the expansions are properly synchronized {\it
1380: in time}. In particular, we will have to worry about the appearance of
1381: multiple time scales at different orders.
1382: 
1383: Substituting the inner expansions into the rescaled 
1384: equations ~(\ref{eq:cin})--(\ref{eq:phiin}) causes the time-dependent terms
1385: to drop out at leading order.  Physically, this quasi-equilibrium
1386: occurs because the charging time, $\tau_c$, is much {\it larger} than
1387: the Debye time, $\tau_D$, characteristic of local dynamics in the
1388: boundary layer (at the scale of the Debye length, $\lambda_D$). As a
1389: result, we systematically arrive at classical Gouy-Chapman profiles
1390: for the equilibrium diffuse layer at leading order,
1391: \begin{equation}
1392: \tilde{c}_\pm \sim e^{\mp \tilde{\psi}}, \ \ \tilde{c}_0 = \cosh
1393: \tilde{\psi}_0 , \ \ \tilde{\rho}_0 = - \sinh \tilde{\psi}_0
1394: \end{equation}
1395: where the excess voltage relative to the bulk, 
1396: \begin{equation}
1397: \tilde{\psi}(y,t) =
1398: \tilde{\phi}(y,t) - \bar{\phi}(-1,t) \sim \tilde{\psi}_0 + \epsilon\, 
1399: \tilde{\psi}_1 + \ldots,
1400: \end{equation}
1401: satisfies the Poisson-Boltzmann equation at leading order,
1402: \begin{equation}
1403: \frac{\partial^2 \tilde{\psi}_0}{\partial y^2} = \sinh \tilde{\psi}_0 .
1404: \end{equation}
1405: Note that matching implies $\tilde{\psi}_0(\infty,t) =
1406: \tilde{\psi}_1(\infty,t)=\ldots = 0$. The dimensionless zeta
1407: potential, $\tilde{\zeta}(t) = \tilde{\psi}(0,t)$, varies as the
1408: diffuse layer charges.
1409: 
1410: After the first integration we apply matching to the electric
1411: field,
1412: \begin{equation}
1413: \frac{\partial\tilde{\phi}}{\partial y}(\infty,t) \sim \epsilon \, 
1414: \frac{\partial\bar{\phi}}{\partial x}(-1,t) \ \longrightarrow \ 
1415: \frac{\partial\tilde{\phi}_0}{\partial y}(\infty,t) = 0,
1416: \label{eq:Ematch}
1417: \end{equation}
1418: to obtain 
1419: \begin{equation}
1420: \frac{\partial \tilde{\psi}_0}{\partial y} = - 2 \, \sinh(\tilde{\psi}_0/2).
1421: \label{eq:psiy}
1422: \end{equation}
1423: After the second integration,
1424: \begin{equation}
1425: \tilde{\psi}_0(y,t) = - 4\, \tanh^{-1}(e^{-(y+K(t))}),   \label{eq:psiin0}
1426: \end{equation}
1427: we are left with a constant, 
1428: \begin{equation}
1429: K(t) = \log\coth (-\tilde{\zeta}_0(t)/4),   \label{eq:K}
1430: \end{equation}
1431: to be determined from $\tilde{\zeta}_0(t)$ (below) by the Stern
1432: boundary condition at the cathode surface, $y=0$, and the coupling to
1433: the bulk region.  The offset parameter, $K(t)$, which also appears in
1434: the concentration and charge density,
1435: \begin{eqnarray}
1436: \tilde{c}_0(y,t) &= & 1 + 2\, \csch^2(y + K(t)) \label{eq:cin0} \\ 
1437: \tilde{\rho}_0(y,t) &= & 2\, \csch(y+K(t))\coth(y+K(t)) \label{eq:rhoin0}
1438: \end{eqnarray}
1439: is quite sensitive to Faradaic reactions~\cite{bonnefont01}, but here
1440: we focus only on the effect of compact-layer capacitance.
1441: 
1442: 
1443: \subsection{ Time-dependent Matching }
1444: 
1445: It seems we have reached a paradox: Both the bulk and the boundary
1446: layers are in quasi-equilibrium at leading order, and yet there must
1447: be some dynamics, if we have chosen the proper time scale. The
1448: resolution lies in taking a closer look at asymptotic
1449: matching. Physically, we are motivated to consider the dynamics of the
1450: total diffuse charge, which has the scaling, $q(t) \sim \epsilon\, \tilde{q}(t)
1451: $, where
1452: \begin{equation}
1453: \tilde{q} = \int_0^\infty \tilde{\rho}(y,t)\, dy \sim \tilde{q}_0
1454: + \epsilon\, \tilde{q}_1 + \epsilon^2 \, \tilde{q}_2  + \ldots \label{eq:qt}
1455: \end{equation}
1456: Taking a time derivative using Eq.~(\ref{eq:rhoin}) and applying the no-flux
1457: boundary condition (\ref{eq:bcrho}), we find
1458: \begin{equation}
1459: \frac{d \tilde{q}}{dt} = \lim_{y \to \infty} \frac{1}{\epsilon}
1460: \left(\frac{\partial \tilde{\rho}}{\partial y} + \tilde{c} \frac{\partial
1461: \tilde{\phi}}{\partial y}\right)   
1462:  \sim \lim_{x\to -1} \left(\frac{\partial
1463: \bar{\rho}}{\partial x} + \bar{c} \frac{\partial \bar{\phi}}{\partial
1464: x}\right)
1465: \label{eq:qmatch}
1466: \end{equation}
1467: where we have applied matching to the {\it derivatives} (flux
1468: densities).  Substituting the inner and outer expansions yields a
1469: hierarchy of matching conditions. At leading order, we have
1470: \begin{equation}
1471: \frac{d\tilde{q}_0}{dt}(t) = \bar{j}_0(t),
1472:   \label{eq:dq}
1473: \end{equation}
1474: which shows that we have chosen the right time scale because this is a
1475: balance of O(1) quantities.  Moreover, it can be shown that any other
1476: choice of scaling would lead to a breakdown of asymptotic matching in
1477: the limit $\epsilon \rightarrow 0$. (For example, in the analogous
1478: Equations (42)--(43) of Ref.~\cite{gonzalez00} for small AC
1479: potentials, the time-dependent term vanishes in this limit, showing
1480: that the proper scaling was not used.) Therefore, the correct charging
1481: time scale, Eq.~(\ref{eq:tmixed}), in the weakly nonlinear regime
1482: follows systematically from time-dependent asymptotic matching at
1483: leading order.
1484: 
1485: The physical interpretation of Equation~(\ref{eq:dq}) is clear: At
1486: leading order, the boundary layer acts like a capacitor, whose
1487: total charge (per unit area), $\tilde{q}$, changes in response to the
1488: transient Faradaic current density, $\bar{j}(t)$, from the bulk.  The
1489: matching condition can also be understood physically as a statement of
1490: current continuity across the diffuse
1491: layer. Substituting Poisson's equation~(\ref{eq:phiin}) into
1492: Eq.~(\ref{eq:qt}), integrating, and matching the electric field using
1493: Eq.~(\ref{eq:Ematch}), we see that the left hand side of
1494: Eq.~(\ref{eq:dq}) is simply the leading-order (dimensionless) {\it
1495: displacement current
1496: density}~\cite{bonnefont01,bonnefont_thesis,brumleve78,cohen65} at the
1497: cathode surface,
1498: \begin{equation}
1499: \frac{d \tilde{q}_0}{dt} = \frac{\partial}{\partial t} \frac{\partial
1500: \tilde{\phi}_0}{\partial y}(0,t) = \tilde{j}_0(t) ,
1501: \end{equation}
1502: so the matching condition simply reads, $\tilde{j}_0(t) =
1503: \bar{j}_0(t)$. This transient displacement current exists in the
1504: external circuit, even if there is no Faradaic current.
1505: 
1506: 
1507: 
1508: \subsection{ Leading-order Dynamics }
1509: 
1510: Using Eqs.~(\ref{eq:phiin}), (\ref{eq:Ematch}) and (\ref{eq:psiy}),
1511: the integral in Eq.~(\ref{eq:qt}) can be performed at leading order to
1512: obtain the Chapman's formula for the total diffuse charge,
1513: \begin{equation}
1514: \tilde{q}_0 = - 2\, \sinh(\tilde{\zeta}_0/2) . \label{eq:qzeta}
1515: \end{equation}
1516: The Stern boundary condition, Eq.~(\ref{eq:sternbc}), then yields,
1517: \begin{equation}
1518: %\Delta\phi_D + 2\, \delta\, \sinh(\Delta\phi_D/2) = \Delta\phi_i ,
1519: \tilde{\zeta}_0 + 2\, \delta\, \sinh(\tilde{\zeta}_0/2) = \bar{j}_0(t)
1520: - v = \tilde{\Psi}_0 ,
1521: \label{eq:phid}
1522: \end{equation}
1523: where $\tilde{\Psi}(t) = - v - \bar{\phi}(-1,t) \sim \tilde{\Psi}_0 +
1524: \epsilon \tilde{\Psi}_1 + \ldots$ is the total voltage across the
1525: compact and diffuse layers.  Substituting into the matching condition,
1526: Eq.~(\ref{eq:dq}), we obtain an ordinary, initial-value problem,
1527: either for the leading-order double-layer voltage
1528: % ($ \Delta\phi_i(t)\rightarrow v$),
1529: \begin{equation}
1530: -\tilde{C}_0(\tilde{\Psi}_0)\, \frac{d\tilde{\Psi}_0}{dt} =
1531: \tilde{\Psi}_0 + v,\ \ \ \tilde{\Psi}_0(0)=0, \label{eq:PsiODE}
1532: \end{equation}
1533: or for the leading-order current density
1534: % ($j(t) \rightarrow 0$),
1535: \begin{equation}
1536: \tilde{C}_0(\bar{j}_0-v)\, \frac{d\bar{j}_0}{dt} = -\bar{j}_0 , \ \ \
1537: \bar{j}_0(0) = v,  \label{eq:jODE}
1538: \end{equation}
1539: where $\tilde{C}_0(\tilde{\Psi}_0) = d\tilde{q}_0/d\tilde{\Psi}_0$ is
1540: the differential capacitance for the double layer as a function of its
1541: total voltage, relative to the potential of zero charge. 
1542: 
1543: \begin{figure}
1544: \includegraphics[width=2.5in]{RC.eps}
1545: \caption{ Sketch of the equivalent RC circuit for the leading-order
1546: weakly nonlinear approximation: compact-layer and diffuse-layer
1547: capacitors in series with a bulk resistor. Although remarkably robust,
1548: the circuit approximation is violated by higher-order corrections,
1549: especially at large voltages.
1550: \label{fig:RC}}
1551: \end{figure}
1552: 
1553: The effective double-layer capacitance is given by
1554: \begin{equation}
1555: \tilde{C}_0 = \frac{1}{\mbox{sech}(\tilde{\zeta}_0/2) + \delta}   \label{eq:Ci}
1556: \end{equation}
1557: where $\tilde{\zeta}_0$ is related to $\tilde{\Psi}_0$ by
1558: Eq.~(\ref{eq:phid}). A similar formula arises in the classical
1559: circuit model of Macdonald~\cite{macdonald54}. Indeed, the
1560: leading-order charging dynamics from asymptotic analysis corresponds
1561: exactly to the nonlinear RC circuit shown in Fig. ~\ref{fig:RC}. We
1562: expect, however, that the {\it ad hoc} circuit approximation cannot
1563: describe higher-order asymptotic approximations, where the finite
1564: thickness of the double layer becomes important.
1565: 
1566: Linearizing for small voltages, $\tilde{C}_0 \sim
1567: 1/(1+\delta)$, we obtain the same results as before in the limit
1568: $\epsilon\rightarrow 0$, now by a completely different method,
1569: \begin{eqnarray}
1570: \bar{j}_0(t) &\sim& v\, e^{-(1+\delta) t} = v + \tilde{\Psi}_0(t) \\
1571: \tilde{q}_0(t) &\sim& \frac{v (1 - e^{-(1+\delta) t})}{1+\delta}  
1572: \end{eqnarray}
1573: As shown in Fig.~\ref{fig:q} for $\delta=0.1$, the linearization
1574: describes the charging dynamics fairly accurately, even for somewhat large
1575: voltages ($v \approx 1$), as long as $\delta$ is not too small.  One way to understand
1576: this is that the total differential capacitance satisfies the uniform
1577: bounds,
1578: \begin{equation}
1579: \frac{1}{1+\delta} = \tilde{C}_0(0) \leq \tilde{C}_0(\tilde{\Psi}_0) <
1580: \tilde{C}_0(\infty) = \frac{1}{\delta} , \label{eq:Cbound}
1581: \end{equation}
1582: in the linear and nonlinear regimes. Moreover, the linearization is
1583: always accurate at early times (up to $t \approx 1$ or $\tau \approx
1584: \tau_c$) for any applied voltage, as long as the initial
1585: zeta potential (or diffuse charge) is small. This is also clearly seen
1586: in Fig.~\ref{fig:q}.
1587: 
1588: 
1589: \begin{figure}
1590: \includegraphics[width=3in]{fig6.eps}
1591: \caption{ The comparison of the full numerical solution with the
1592: leading-order asymptotic results: (a) $j/v$, (b) $q/q_{lin}$ (early
1593: evolution), and (c) $q/q_{lin}$ (long time evolution).  The full
1594: numerical solution are shown with dot ($v=1$), dot-dash (with open squares in (c)) ($v=2$) and
1595: long dash with open triangles ($v=3$).  The leading-order asymptotic results are
1596: plotted with dash ($v=1$), dot-dot-dot-dash (with filled squares in (c)) ($v=2$) and long dash with filled triangles
1597: ($v=3$).  The curves for $v=3$ are omitted in (a) and (b) for clarity.
1598: The solid lines show the linear dynamics in the thin double-layer
1599: limit.  \label{fig:qa}}
1600: \end{figure}
1601: 
1602: 
1603: The dynamical equation (\ref{eq:PsiODE}) or (\ref{eq:jODE}) is
1604: first-order and separable, so its exact solution is easily expressed
1605: in integral form,
1606: \begin{equation}
1607: \tilde{\Psi}_0(t) = \bar{j}_0(t) - v = - F^{-1}(t)  \label{eq:jexact}
1608: \end{equation}
1609: where 
1610: \begin{equation}
1611: F(z)  = \int_0^z \frac{\tilde{C}_0(u)\,du}{u+v}
1612: \label{eq:F}
1613: \end{equation}
1614: The integral can be evaluated numerically and the total charge
1615: recovered from the Eqs.~(\ref{eq:qzeta}) and (\ref{eq:phid}). The
1616: results in Fig.~\ref{fig:qa} show that the leading-order dynamics
1617: compares fairly well with the numerical solution to the full nonlinear
1618: problem for $\epsilon=0.05$ and $\delta=0.1$, at least for the decay
1619: of the current density, especially at early times ($t \approx 1$). The
1620: limiting value of the total diffuse charge is also approximated much
1621: better than in the linear theory (Fig.~\ref{fig:q}), due to the
1622: nonlinear differential capacitance, Eq.~(\ref{eq:Ci}). For large
1623: voltages ($v > 1$), however, total charge shows some secondary
1624: dynamics at longer time scales ($t \gg 1$), which is not fully
1625: captured by the leading-order asymptotic approximation (or the
1626: corresponding circuit model). As we shall see below, this can only be
1627: understood by considering higher-order terms which violate the circuit
1628: approximation.
1629: 
1630: 
1631: % CONVERGENCE TEST
1632: %\begin{figure}
1633: %\includegraphics[width=3in]{figconv.eps}
1634: %\caption{
1635: %\label{fig:figconv} }
1636: %\end{figure}
1637: 
1638: 
1639: For moderately large voltages ($v \approx 1$), we can expand around
1640: $u=v$ in the integrand of Eq.~(\ref{eq:F}) and obtain a
1641: long-time exponential decay,
1642: \begin{equation}
1643: \bar{j}_0(t)  = v + \tilde{\Psi}_0(t) \propto e^{- t/\tilde{C}_0(v)} 
1644: \label{eq:jlate}
1645: \end{equation}
1646: as $t \rightarrow \infty$. This reveals a (dimensionless)
1647: characteristic time, $t_c = \tilde{C}_0(v)$, which is larger than that
1648: of the linear regime, $t_c = \tilde{C}_0(0) = 1/(1+\delta)$, by at most
1649: a factor of $1+1/\delta$ ($=11$ in our numerical examples). Although
1650: this factor is non-negligible, the characteristic time, $\tau_c$, is
1651: still the basic time scale, rather than $\tau_L$ and $\tau_D$ which differ
1652: from $\tau_c$ by factors of $\epsilon$, i.e. usually two or more
1653: orders of magnitude. As the voltage is increased, however,
1654: nonlinearity always becomes important, and one of its generic effects
1655: is to slow down the relaxation process.
1656: 
1657: 
1658: In order to simplify the response function, $F(z)$, and other
1659: quantities, it is useful to consider the regular limit of thin Stern
1660: layers, $\delta\rightarrow 0$ (taken after the singular limit of thin
1661: diffuse layers, $\epsilon \rightarrow 0$). In this common physical
1662: regime, where $\lambda_S \ll \lambda_D \ll L$, the following
1663: asymptotic expansions can be derived by iteration~\cite{hinch} from
1664: Eqs.~(\ref{eq:qzeta}), (\ref{eq:phid}), and (\ref{eq:Ci}):
1665: \begin{eqnarray}
1666: \tilde{\zeta}_0 &\sim& \tilde{\Psi}_0 - 2\delta \sinh\frac{\tilde{\Psi}_0}{2}
1667: + \delta^2 \sinh \tilde{\Psi}_0 + \ldots \label{eq:phida} \\ 
1668: \tilde{q}_0 &\sim& - 2 \sinh\frac{\tilde{\Psi}_0}{2} 
1669: + \delta \sinh \tilde{\Psi}_0 \nonumber\\ &
1670: & - \delta^2 \left( \sinh \tilde{\Psi}_0 \cosh\frac{\tilde{\Psi}_0}{2} +
1671: \sinh^3\frac{\tilde{\Psi}_0}{2} \right)
1672: \label{eq:qa} \\ 
1673: \tilde{C}_0 &\sim& \cosh\frac{\tilde{\Psi}_0}{2} 
1674: - \delta \cosh \tilde{\Psi}_0  \nonumber \\ & & +
1675: \delta^2 \left( \cosh \tilde{\Psi}_0\, \cosh\frac{\tilde{\Psi}_0}{2} +
1676: \frac{1}{2}\, \sinh \tilde{\Psi}_0 \, \sinh\frac{\tilde{\Psi}_0}{2}
1677: \right. \nonumber \\
1678: & & \left. + \frac{3}{2}\, \sinh^2\frac{\tilde{\Psi}_0}{2}\,
1679: \cosh\frac{\tilde{\Psi}_0}{2} \right)  \label{eq:Ca} 
1680: \end{eqnarray}
1681: The response function can then be expanded in somewhat simpler (but
1682: still nontrivial) integrals,
1683: \begin{equation}
1684: F(z) \sim \int_0^z \frac{\cosh(u/2)\, du}{u+v} - \delta \int_0^z
1685: \frac{\cosh(u)\, dx}{u+v} + \ldots
1686: \end{equation}
1687: in the limit $\delta \rightarrow 0$.
1688: 
1689: 
1690: %\begin{figure}
1691: %\includegraphics[width=3in]{fig5.eps}
1692: %\caption{ For $\delta=0.1$, plots of (a) the diffuse-layer voltage,
1693: %$\Delta\phi_D$, and (b) the differential total capacitance, each
1694: %versus the dimensionless interfacial voltage, $-\tilde{\Psi}_0$ (dotted
1695: %lines).  (All quantities are dimensionless.) The corresponding points
1696: %test the asymptotic approximations, Eqs.~(\protect\ref{eq:phida}) and
1697: %(\protect\ref{eq:Ca}), to first (cross) and second order (diamond) in
1698: %$\delta$, respectively.\label{fig:C} }
1699: %\end{figure}
1700: 
1701: 
1702: 
1703: \begin{figure*}
1704: \includegraphics[width=6in]{fig7.eps}
1705: \caption{ The potential (a), charge density (b) and concentration (c)
1706: at $t=1$ for a large dimensionless voltage, $v=1$, with
1707: $\epsilon=0.05$ and $\delta=0.1$. The full numerical solution (dashed 
1708: lines) is compared with the leading-order uniformly valid
1709: approximation (dotted lines),
1710: Eqs.~(\ref{eq:uphi})--(\ref{eq:urho}).  The analytical 
1711: approximations are almost indistinguishable from the numerical
1712: solutions for $\phi$ and $\rho$, but not for $c$, which shows
1713: errors of a few percent ($< \epsilon$) just outside the double layers.
1714: \label{fig:compare} }
1715: \end{figure*}
1716: 
1717: 
1718: \subsection{ Uniformly Valid Approximations }
1719: 
1720: Asymptotic analysis tells us not only the behavior of integrated
1721: quantities like total charge and voltage, but also the complete
1722: spatio-temporal profiles of the charge density and potential. As
1723: usual, uniformly valid approximations (in space) are constructed by
1724: adding the outer and inner approximations and subtracting the
1725: overlaps. Taking advantage of the symmetries in Eq.~(\ref{eq:sym}), we
1726: obtain the following leading-order approximations:
1727: \begin{eqnarray}
1728: \phi(x,t) &\sim& \overline{j}(t)\, x +
1729: \tilde{\psi}_0\left(\frac{1+x}{\epsilon},t\right) - 
1730: \tilde{\psi}_0\left(\frac{1-x}{\epsilon},t\right) \label{eq:uphi}  \\ c(x,t)
1731: &\sim& \tilde{c}_0\left(\frac{1+x}{\epsilon},t\right) +
1732: \tilde{c}_0\left(\frac{1-x}{\epsilon},t\right) - 1 \label{eq:uc} \\
1733: \rho(x,t) &\sim& \tilde{\rho}_0\left(\frac{1+x}{\epsilon},t\right) -
1734: \tilde{\rho}_0\left(\frac{1-x}{\epsilon},t\right) \label{eq:urho}
1735: \end{eqnarray}
1736: where the boundary-layer contributions are given by
1737: Eqs.~(\ref{eq:psiin0})--(\ref{eq:rhoin0}) and Eq.~(\ref{eq:phid}),
1738: which express the effect of the compact layer. The time dependence of
1739: the leading-order approximations is entirely determined by the bulk
1740: current density, $\bar{j}_0(t)$, or the double-layer voltage,
1741: $\tilde{\Psi}_0(t)$, via Eqs.~(\ref{eq:jexact}) and (\ref{eq:F}).
1742: 
1743: As shown in Fig.~\ref{fig:compare}, the time-dependent approximations
1744: for $\phi$ and $\rho$ are in excellent agreement with our numerical
1745: results well into the nonlinear regime ($v=1$), even for a fairly
1746: large boundary-layer thickness, $\epsilon = 0.05$. The charge density
1747: clearly shows the expected separation into three regions: a neutral
1748: bulk with two charged boundary layers of $O(\epsilon)$ width.  On the
1749: other hand, for the same parameters, the leading-order approximation
1750: of $c$ is not nearly as good. As expected, the concentration exhibits
1751: a homogeneous bulk region and two inhomogeneous boundary layers of
1752: $O(\epsilon)$ width, which are fairly well described, but there are
1753: also intermediate regions of depleted concentration extending far into
1754: the bulk, which are not captured at leading order.
1755: 
1756: 
1757: 
1758: \section{ Higher-Order Effects }
1759: \label{sec:higher}
1760: 
1761: \subsection{ Neutral-Salt Adsorption by the Double Layer }
1762:  
1763: We have seen that each diffuse-charge layer acquires an excess salt 
1764: concentration relative to the outer region. At leading order, however,
1765: there is no sign of how the extra ions got there.  This paradox, which
1766: also applies to circuit models, is apparent from symmetry alone,
1767: Eq.~(\ref{eq:sym}) --- Diffuse charge near the cathode grows by
1768: bulk electromigration, which creates equal and opposite diffuse charge
1769: near the anode. In contrast, the excess concentration is the same in
1770: both double layers, so it can only arrive there by {\it diffusion} of
1771: neutral electrolyte from the bulk, which is excluded at leading order.
1772: 
1773: The key to understanding higher-order terms, therefore, is the total
1774: excess concentration per unit surface area in (say) the cathodic
1775: diffuse layer, $w(t) = \epsilon\, \tilde{w}(t)$, where
1776: \begin{equation}
1777: \tilde{w}(t) = \int_0^\infty \left[\tilde{c}(y,t)-\bar{c}_0(-1,t)\right]
1778: dy = \tilde{w}_0(t) + \epsilon\, \tilde{w}_1(t) + \ldots   \label{eq:w}
1779: \end{equation}
1780: is analogous to the scaled total diffuse charge, $\tilde{q}(t)$. (Note
1781: that $\bar{c}_0(-1,t)=1$ in our model problem, but Equation
1782: ~(\ref{eq:w}) is more general.)  We proceed with matching in the same
1783: manner as above.  Taking a time derivative using Eq.~(\ref{eq:cin})
1784: and applying the no-flux boundary condition (\ref{eq:bcc}), we find
1785: \begin{equation}
1786: \frac{d \tilde{w}}{dt} = \lim_{y \to \infty} \frac{1}{\epsilon}
1787: \left(\frac{\partial \tilde{c}}{\partial y} + \tilde{\rho} \frac{\partial
1788: \tilde{\phi}}{\partial y}\right)   
1789:  \sim \lim_{x\to -1} \left(\frac{\partial
1790: \bar{c}}{\partial x} + \bar{\rho} \frac{\partial \bar{\phi}}{\partial
1791: x}\right)
1792: \label{eq:wmatch-full}
1793: \end{equation}
1794: Substituting the inner and outer expansions yields another 
1795: hierarchy of matching conditions. At leading order, we have
1796: \begin{equation}
1797: \frac{d\tilde{w}_0}{d\bar{t}}(t)
1798: = \frac{1}{\epsilon}\, \frac{d\tilde{w}_0}{dt}(t) = \frac{\partial
1799:   \bar{c}_1}{\partial x}(-1,t) ,
1800:   \label{eq:wmatch}
1801: \end{equation}
1802: which, unlike Eq.~(\ref{eq:dq}), involves a {\it new time variable},
1803: \begin{equation}
1804: \bar{t} ={\epsilon} t = \frac{\epsilon\,\tau}{\tau_c} =
1805: \frac{\tau}{\tau_L} ,
1806: %%%% QQQ Martin, please make sure my changes are OK above
1807: % What did you do?
1808: \end{equation}
1809: scaled
1810: to the bulk diffusion time, $ \tau_L=L^2/D$. Physically, this matching
1811: condition simply expresses mass conservation: The (zeroth order)
1812: excess concentration in the diffuse layer varies in response to the
1813: (first order) diffusive flux from the bulk.
1814: 
1815: In Eq.~(\ref{eq:w}), the left-hand side is given by the leading-order
1816: inner approximation calculated above.  Substituting
1817: Eq.~(\ref{eq:cin0}) into Eq.~(\ref{eq:w}), integrating, and using
1818: Eq.~(\ref{eq:K}) yields,
1819: \begin{eqnarray}
1820: \frac{d\tilde{w}_0}{dt}(t) &=& 2 \frac{d}{dt} \coth K(t) \nonumber \\
1821: &=& 2 \frac{d}{dt} \cosh \frac{\tilde{\zeta}_0(t)}{2}  \label{eq:dw0dt} \\
1822: &=& - \frac{\tilde{q}_0(t)}{2} \,  \frac{d\tilde{\zeta}_0(t)}{dt} \nonumber
1823: \end{eqnarray}
1824: where we have used the identity, $\cosh 2z = -\coth\log\tanh z$.
1825: Recall that the leading-order zeta potential, $\tilde{\zeta}_0(t)$, is
1826: related via Eq.~(\ref{eq:phid}) to the leading-order bulk current
1827: density, $\bar{j}_0(t)$, or interfacial voltage, $\tilde{\Psi}_0(t)$,
1828: given by Eqs.~(\ref{eq:jexact}) and (\ref{eq:F}). Integrating
1829: Eq.~(\ref{eq:dw0dt}) and requiring $\tilde{w}_0=0$ for
1830: $\tilde{\zeta}_0=0$, we also obtain a simple expression for the excess
1831: concentration,
1832: \begin{equation}
1833: \tilde{w}_0 = 4\, \sinh^2 \frac{\tilde{\zeta}_0}{4} ,
1834: \label{eq:w0}
1835: \end{equation}
1836: which also holds for the static Gouy-Chapman solution.  Of course,
1837: this is another sign that (at leading order) a thin double layer stays
1838: in quasi-equilibrium, even while charging.
1839: 
1840: 
1841: \subsection{ The Sign of the Donnan Effect }
1842: 
1843: Before proceeding to calculate the bulk dynamics, we comment on the
1844: sign of the excess concentration in the diffuse part of the double
1845: layer, which corresponds to a {\it positive} adsorption of neutral
1846: salt. In contrast, it is commonly believed that double-layer salt
1847: adsorption is always negative, resulting in an excess neutral
1848: concentration in the nearby bulk electrolyte. Lyklema calls this the
1849: ``Donnan effect'' with reference to Donnan's general papers on
1850: membrane equilibria~\cite{donnan11a,donnan11b,donnan24} and describes
1851: how it is used to infer surface areas from experimental measurements
1852: of concentration variations upon charging (``salt
1853: sieving")~\cite{lyklema,lyklema1}. In the present case of
1854: diffuse-layer adsorption, the following argument is given: Since the
1855: equilibrium co-ion concentration in the double layer near a charged
1856: surface is reduced compared to the bulk, there must be an excess of
1857: co-ions, and hence an excess of neutral concentration (and
1858: counter-ions) in the nearby bulk.
1859: 
1860: How can this belief be reconciled with our analytical and numerical
1861: results, which clearly demonstrate the opposite effect in a model
1862: problem? The difference is that we consider a {\it finite} system with
1863: global ion conservation, while Lyklema considers an open system into
1864: which ions are apparently injected. We also explicitly calculate the
1865: time-dependent response to interfacial charging, while he describes the
1866: steady state after new ions somehow arrive ``from infinity''.
1867: 
1868: Our conclusion is therefore the opposite: Since the equilibrium
1869: counter-ion concentration is enhanced in the double layer, there must
1870: be a depletion of counter-ions, and hence a reduction in neutral
1871: concentration (and co-ions) in the nearby bulk. Gouy's nonlinear
1872: theory shows that at large voltages the excess of counter ions exceeds
1873: the reduction in co-ions in the diffuse layer, so this result is quite
1874: intuitive.
1875: 
1876: 
1877: In real systems, it may be that compact-layer effects, such as the
1878: specific adsorption of ions on the surface, can lead to overall
1879: negative adsorption by the double layer. According to the
1880: Nernst-Planck-Poisson equations, however, the average concentration of
1881: all ions (regardless of species) is always increased in the diffuse
1882: part of the double layer relative to the bulk in any finite system.
1883: The associated local depletion of neutral salt has important
1884: implications for time-dependent electrokinetic phenomena at
1885: polarizable surfaces (section ~\ref{sec:colloids}) since bulk
1886: concentration gradients alter electric fields and produce
1887: diffusio-osmotic slip.
1888: 
1889: 
1890: 
1891: \subsection{ Bulk Diffusion at Two Time Scales }
1892: 
1893: We now proceed to calculate how the bulk concentration is
1894: depleted in time and space during double-layer charging in our model
1895: problem.  The matching condition, Eq.~(\ref{eq:wmatch}), seems to contradict
1896: the analysis above, since it introduces a new time variable,
1897: $\bar{t}$. However, this is the same time scale for the first-order
1898: (diffusive) dynamics in the bulk,
1899: \begin{equation}
1900: \frac{\partial \bar{c}_1}{\partial \bar{t}}
1901: = \frac{1}{\epsilon} \,  \frac{\partial \bar{c}_1}{\partial t}
1902: = \frac{\partial^2 \bar{c}_1}{\partial x^2} .   \label{eq:c1eq}
1903: \end{equation}
1904: We must solve this equation starting from $\bar{c}_1(x,0)=0$ with a
1905: time-dependent prescribed flux at $x=-1$ given by
1906: Eqs.~(\ref{eq:wmatch}) and (\ref{eq:dw0dt}). We also enforce symmetry
1907: about the origin, Eq.~(\ref{eq:sym}). 
1908: 
1909: The Laplace transform of the solution is:
1910: \begin{eqnarray}
1911: \hat{\bar{c}}_1(x,s)
1912: &=& - \frac{\sqrt{s}\,
1913:     \cosh(x\sqrt{s})}{\sinh(\sqrt{s})}\, \int_0^\infty   
1914:     e^{-s\bar{t}} \tilde{w}_0(\bar{t}/\epsilon)
1915:     d\bar{t}   \nonumber \\
1916: &=& \int_0^\infty e^{-s\bar{t}} \bar{c}_1(x,\bar{t}) d\bar{t}   
1917:     \label{eq:c1lap} 
1918: \end{eqnarray}
1919: where $\tilde{w}_0(t)$ is determined by $\tilde{\zeta}_0(t)$ from
1920: Eq.~(\ref{eq:w0}). The prefactor,
1921: \begin{equation}
1922: \hat{G}(s) = \frac{\cosh(x\sqrt{s})}
1923: 	{\sqrt{s}\, \sinh(\sqrt{s})} ,  \label{eq:c1t}
1924: \end{equation}
1925: is the Laplace transform of $G(\bar{t})$, the Green function for
1926: the diffusion equation, Eq. ~(\ref{eq:c1eq}), for a sudden unit flux
1927: of ions at time $\bar{t} = 0^+$ injected at the boundary:
1928: \begin{equation}
1929: G(x,0)=0, \ \ \frac{\partial G}{\partial x}(-1,\bar{t}) =
1930: \delta^+(\bar{t}) .  \label{eq:c1bc1}
1931: \end{equation}
1932: The same Green function also arises in the equivalent problem of an
1933: initial unit source adjacent to a reflecting wall,
1934: \begin{equation}
1935: G(x,0) = \delta(x+1^+), \ \ \frac{\partial G}{\partial
1936: x}(-1,\bar{t}) = 0 . \label{eq:c1bc2}
1937: \end{equation}
1938: In this form, the Green function can be obtained by inspection,
1939: \begin{equation}
1940: G(x,\bar{t}) = \frac{1}{\sqrt{\pi
1941: \bar{t}}} \, \sum_{m=-\infty}^\infty e^{-(x - 2m+1)^2/4\bar{t}} ,
1942: \label{eq:c1profile}
1943: \end{equation}
1944: using the method of images.
1945: 
1946: 
1947: Since $\hat{\bar{c}}_1(x,s)$ is expressed as a product of two
1948: Laplace transforms, Eq.~(\ref{eq:c1lap}), the inverse is equal to 
1949: the convolution of the two original functions:
1950: \begin{equation}
1951: \bar{c}_1(x,\bar{t}) = -\int_0^{\bar{t}} d\bar{t}^\prime \, 
1952: G(x,\bar{t}^\prime-\bar{t}) \, \frac{\partial \tilde{w}_0}{\partial \bar{t}}
1953: \left(\frac{\bar{t}^\prime}{\epsilon}\right). \label{eq:c1int1} 
1954: \end{equation}
1955: This form clearly demonstrates that the boundary forcing occurs over the
1956: fast, charging time, $t = \bar{t}/\epsilon$, while the response
1957: described by the Green-function kernel occurs over the slow, diffusion
1958: time, $\bar{t}$. The separation of time scales is apparent in
1959: the equivalent expression,
1960: \begin{equation}
1961: \bar{c}_1(x,t) = - \int_0^t dt^\prime
1962: G\left(x,\epsilon(t^\prime-t)\right)\,
1963: \tanh\left(\frac{\tilde{\zeta}_0(t^\prime)}{2}\right)\,
1964: \bar{j}_0(t^\prime) \label{eq:c1j}
1965: \end{equation}
1966: which can be derived from Eq.~(\ref{eq:c1int1}) using
1967: Eq.~(\ref{eq:dw0dt}). This form shows explicitly how solution for the
1968: current at leading order, Eq.~(\ref{eq:jexact}), fully determines the
1969: bulk concentration at first order.
1970: 
1971: 
1972: \begin{figure}
1973: \includegraphics[width=\linewidth]{fig8.eps}
1974: \caption{ Weakly nonlinear dynamics for $v=1$, $\epsilon=0.05$,
1975: $\delta=0.1$, showing the effect of bulk diffusion. The concentration
1976: from the full numerical solution is shown in the half cell (a) and in
1977: the diffuse layer (b) for $t=0.5$ (solid), $t=1$ (dot), $2$ (dash),
1978: $4$ (dot-dash), $8$ (dot-dot-dot-dash), and $20$ (long
1979: dash). \label{fig:weak} }
1980: \end{figure}
1981: 
1982: Before further analysis of the exact solution for $c_1(x,t)$, we
1983: describe its physical significance. The bulk concentration at
1984: first-order exhibits diffusive relaxation at two different time
1985: scales, $t=O(1)$ and $\bar{t}=O(1)$, or with units, $\tau =
1986: O(\lambda_DL/D)$ and $\tau = O(L^2/D)$, respectively.  For $t=O(1)$
1987: and $\bar{t} = O(\epsilon)$, the initial double-layer charging process
1988: proceeds without any significant changes in concentration at the bulk
1989: length scale, $x = O(1)$. During this phase, each diffuse-charge layer
1990: acquires an $O(\epsilon)$ amount of excess concentration, given by
1991: Eq.~(\ref{eq:w0v}).  This excess concentration has been acquired by a
1992: diffusive process, which at this time scale corresponds to a bulk
1993: diffusion layer of $O(\sqrt{\epsilon})$ width near each electrode.
1994: This implies an $O(\sqrt{\epsilon})$ depletion of the neutral salt
1995: concentration in the bulk diffusion layers. These scaling arguments
1996: are confirmed by Fig.~\ref{fig:compare}(c) for $v=1$ and
1997: $\epsilon=0.05$, where at time $t = 1$ (or $\bar{t} = \epsilon$) the
1998: diffusion layers are roughly of width $\sqrt{2\bar{t}} =
1999: \sqrt{2\epsilon} \approx 0.3$. The formation and spreading of the
2000: diffusion layers is also shown in more detail in Fig.~\ref{fig:weak}.
2001: 
2002: 
2003: \subsection{ Evolution of the Diffusion Layers }
2004: 
2005: In the previous section, we derived the time-dependent outer
2006: approximation, Eq.~(\ref{eq:cout}), to first order,
2007: \begin{equation}
2008: \bar{c}(x,t) \sim 1 + \epsilon \, \bar{c}_1(x,t),  \label{eq:cc1}
2009: \end{equation}
2010: which displays dynamics at both the RC time and the bulk diffusion
2011: time. The result, Eqs.~(\ref{eq:c1profile})--(\ref{eq:c1j}), is fairly
2012: complicated, so in this section we try to gain some simple analytical
2013: insight. In the limit $\epsilon \rightarrow 0$, the initial charging
2014: process at the time scale $\bar{t} = O(\epsilon)$ is instantaneous,
2015: and we are left with only the slow relaxation of the bulk diffusion
2016: layers. Explicitly taking this limit in Eq.~(\ref{eq:c1lap}) with
2017: $\bar{t} = \epsilon t$ fixed, 
2018: \begin{equation}
2019: %\lim_{\epsilon\rightarrow 0} \hat{\bar{c}}_1(x,s) 
2020: %&=& -\tilde{w}_0(\infty) \hat{G}(s) \nonumber \\ 
2021: \lim_{\epsilon\rightarrow 0}
2022: \bar{c}_1(x,\bar{t}) = -\tilde{w}_0(\infty) G(x,\bar{t}) ,   \label{eq:c1approx}
2023: \end{equation}
2024: we see that the slow-scale evolution of the diffusion layers is 
2025: given by the Green function, $G(x,\bar{t})$, with a source of
2026: strength, $-\tilde{w}_0(\infty)$, equal to the leading-order total salt
2027: adsorption. According to Eqs.(\ref{eq:phid}) and (\ref{eq:w0}) with
2028: $\bar{j}_0(\infty)=0$, this is given by
2029: \begin{equation}
2030: \tilde{w}_0(\infty) = 4\, \sinh^2 \frac{f^{-1}(v)}{4}, 
2031: \label{eq:w0v}
2032: \end{equation}
2033: where
2034: \begin{equation}
2035: f(\zeta) = \zeta + 2\delta \sinh(\zeta/2) \label{eq:fzeta}
2036: \end{equation}
2037: which reduces to 
2038: \begin{equation}
2039: \tilde{w}_0(\infty) = 4\, \sinh^2 \frac{v}{4},  \label{eq:wtd0}
2040: \end{equation}
2041: in the absence of any compact layers ($\delta=0$).
2042: 
2043: 
2044: 
2045: \begin{figure}
2046: \includegraphics[width=\linewidth]{fig9.eps}
2047: \caption{ \label{fig:c1} Simple approximations of the bulk diffusion
2048: layers for weakly nonlinear charging dynamics with $v=1$,
2049: $\epsilon=0.05$, $\delta=0.1$. The full numerical solution (solid) is
2050: compared with the approximate first-order expansion at the diffusion
2051: time scale, given by Eqs.~(\ref{eq:c1approx}), (\ref{eq:c1profile}),
2052: and (\ref{eq:cc1}). Also, shown is the latter plus the zeroth-order
2053: inner approximation, Eq.~(\ref{eq:cin0}), for the diffuse layers
2054: (dashed). }
2055: \end{figure}
2056: 
2057: 
2058: This simple approximation describes two diffusion layers created at
2059: the electrodes slowly invading the entire cell. At first, they have
2060: simple Gaussian profiles,
2061: \begin{equation}
2062: \bar{c}(x,\bar{t}) \sim 1 - \frac{\epsilon
2063: \tilde{w}_0(\infty)}{\sqrt{\pi \bar{t}}} \left[ e^{-(x +
2064: 1)^2/4\bar{t}} + e^{-(x -1)^2/4\bar{t}} \right]  \label{eq:cinit}
2065: \end{equation}
2066: for $\bar{t} \ll 1$, which is qualitatively consistent with the
2067: numerical results in Fig.~\ref{fig:compare}(c). To attempt a
2068: quantitative comparison, we also need $t \gg 1$ to use
2069: Eqs.~(\ref{eq:c1approx}) and (\ref{eq:c1profile}). As shown in
2070: Fig.~(\ref{fig:c1}), the approximation is reasonable for $t=3$ with an
2071: error of roughly $\epsilon^2 = 0.0025$. The two diffusion layers
2072: eventually collide, and the concentration slowly approaches a
2073: (reduced) constant value,
2074: \begin{equation}
2075: \bar{c}(x,\bar{t}) \sim 1 - \epsilon \tilde{w}_0(\infty)   \label{eq:cinf}
2076: \end{equation}
2077: for $\bar{t} \gg 1$, as expected from the steady-state excess
2078: concentration in the double layers. (This result may be checked by
2079: replacing the sum in Eq.~(\ref{eq:c1profile}) with an integral in the
2080: limit $\bar{t} \rightarrow \infty$.)
2081: 
2082: 
2083: \subsection{ Bulk Concentration Polarization }
2084: 
2085: As mentioned above, the bulk charge density remains very small,
2086: $\bar{\rho} = O(\epsilon^3)$, even during double-layer relaxation, but
2087: changes in neutral bulk concentration affect the potential at first
2088: order. Substituting the outer expansions into Eq.~(\ref{eq:rho}) and
2089: collecting terms at $O(\epsilon)$, we have
2090: \begin{equation}
2091: 0 = \frac{\partial}{\partial x}\left[ \bar{c}_0 
2092:   \frac{\partial\bar{\phi}_1}{\partial x} + \bar{c}_1 
2093:   \frac{\partial\bar{\phi}_0}{\partial x} \right].
2094: \end{equation}
2095: This is easily integrated using $\bar{c}_0 = 1$ to obtain the
2096: first-order contribution to the bulk electric field,
2097: \begin{equation}
2098: \frac{\partial\bar{\phi}_1}{\partial x} = \bar{j}_1(t) - \bar{j}_0(t)
2099: \, \bar{c}_1(x,t),
2100: \end{equation}
2101: where the second term describes concentration polarization, i.e. the
2102: departure from a harmonic potential, which would be predicted by Ohm's
2103: law.  The first term is a uniform bulk field (or current) determined
2104: by first-order perturbation in double-layer charge. This follows from
2105: the matching condition, Eq.~(\ref{eq:qmatch}), at first order:
2106: \begin{equation}
2107: \frac{d\tilde{q}_1}{dt} = \bar{j}_1(t).
2108: \end{equation}
2109: where $\tilde{q}_1(t)$ is obtained by solving the inner problem at
2110: first order.
2111: 
2112: 
2113: \subsection{ Perturbations in Double-layer Structure }
2114: 
2115: Unfortunately, the first-order inner problem is difficult to solve
2116: analytically because the perturbed concentration profiles are no
2117: longer in thermal equilibrium during the initial charging phase. To
2118: see this, note that the time derivatives in Eqs.~(\ref{eq:cin}) and
2119: (\ref{eq:rhoin}) contribute nonzero (but known) terms at first order,
2120: \begin{eqnarray}
2121: \frac{\partial \tilde{c}_0}{\partial t} &=& 
2122: \frac{\partial}{\partial y}\left( \frac{\partial \tilde{c}_1}{\partial y} +
2123: \tilde{\rho}_0 \frac{\partial \tilde{\phi}_1}{\partial y} +
2124: \tilde{\rho}_1 \frac{\partial \tilde{\phi}_0}{\partial y}\right)
2125: \label{eq:c1in} \\
2126: \frac{\partial \tilde{\rho}_0}{\partial t} &=& 
2127: \frac{\partial}{\partial y}\left( \frac{\partial \tilde{\rho}_1}{\partial y} +
2128: \tilde{c}_0 \frac{\partial \tilde{\phi}_1}{\partial y} +
2129: \tilde{c}_1 \frac{\partial \tilde{\phi}_0}{\partial y}\right)
2130: \label{eq:rho1in} \\
2131:  -\frac{\partial^2 \tilde{\phi}_1}{\partial y^2} &= &\tilde{\rho}_1
2132: \label{eq:phi1in} 
2133: \end{eqnarray}
2134: although one still solves a system of linear ordinary differential
2135: equations in $y$ at each $t$, since $\tilde{c}_0(y,t)$,
2136: $\tilde{\rho}_0(y,t)$, and $\tilde{\phi}_0(y,t)$ are known.
2137: 
2138: The general problem seems daunting, but some progress can
2139: be made at the scale of bulk diffusion, $\bar{t} = O(1)$ or $t =
2140: O(\epsilon^{-1})$, where the leading-order concentration profiles
2141: remain in thermal equilibrium, without any explicit time
2142: dependence. This will give us some insight into secondary charge
2143: relaxation at the time scale of bulk diffusion. In this limit, the
2144: Equations~(\ref{eq:c1in}) and (\ref{eq:rho1in}) can be integrated to
2145: obtain: \begin{eqnarray} - \frac{\partial \tilde{c}_1}{\partial y}
2146: &\sim& \tilde{\rho}_0 (y,\infty)\frac{\partial
2147: \tilde{\phi}_1}{\partial y} + \tilde{\rho}_1 \frac{\partial
2148: \tilde{\phi}_0}{\partial
2149:  y}(y,\infty) 
2150: \label{eq:c1inb} \\
2151: - \frac{\partial \tilde{\rho}_1}{\partial y} &\sim&
2152: \tilde{c}_0 (y,\infty)\frac{\partial \tilde{\phi}_1}{\partial y} +
2153: \tilde{c}_1 \frac{\partial \tilde{\phi}_0}{\partial
2154:  y}(y,\infty) 
2155: \label{eq:rho1inb} 
2156: \end{eqnarray}
2157: after applying the usual van Dyke matching conditions.
2158: Substituting from Poisson's equation, $\tilde{\rho}_n = -\partial^2
2159: \tilde{\phi}_n/\partial y^2$, at orders $n=0,1$ into
2160: Eq.~(\ref{eq:c1inb}), integrating, and applying matching again, we
2161: obtain:
2162: \begin{equation}
2163: \tilde{c}_1(y,\bar{t}) \sim \frac{\partial \tilde{\phi}_0}{\partial
2164:  y}(y,\infty) \,  \frac{\partial \tilde{\phi}_1}{\partial
2165:  y}(y,\bar{t}) + \bar{c}_1(-1,\bar{t})
2166: \end{equation}
2167: for $\bar{t} > 0$ and $t = \bar{t}/\epsilon \gg 1$.  From the previous
2168: section, we also have the leading-order inner concentration,
2169: \begin{equation}
2170: \tilde{c}_0(y,\infty) = \bar{c}_0(-1,\bar{t}) + \frac{1}{2}
2171: \tilde{E}_0(y,\infty),
2172: \end{equation}
2173: where $\bar{c}_0(-1,\bar{t})=1$ is the leading-order outer
2174: concentration, and 
2175: \begin{equation}
2176: \tilde{E}_0(y,\infty) = -\frac{\partial \tilde{\phi}_0}{\partial
2177: y}(y,\infty) = 2 \sinh \frac{\tilde{\psi}_0(y,\infty)}{2},
2178: \end{equation}
2179: is the leading-order inner electric field in steady-state.  Finally,
2180: we substitute these expressions into Eq.~(\ref{eq:rho1inb}) and use
2181: Eq.~(\ref{eq:rho1in}) to obtain a master equation for the first-order
2182: inner electric field, $\tilde{E}_1(y,\bar{t})= -\frac{\partial
2183: \tilde{\phi}_1}{\partial y}(y,\bar{t})$, at the bulk-diffusion time
2184: scale:
2185: \begin{equation}
2186: \frac{\partial^2 \tilde{E}_1}{\partial y^2}= \left(1 + \frac{3}{2}
2187: \tilde{E}_0^2\right) \tilde{E}_1 +  \bar{c}_1 \tilde{E}_0
2188: \label{eq:E1master} .
2189: \end{equation}
2190: This linear equation with a non-constant coefficient must be solved
2191: subject to the boundary conditions, $\tilde{E}_1(\infty,\bar{t}) = 0$
2192: and $\tilde{E}_1(0,\bar{t}) = -\tilde{q}_1(\bar{t})$. The perturbation
2193: of the total charge, $\tilde{q}_1(\bar{t})$ is obtained by another
2194: integration of the field to get the first-order inner potential, while
2195: applying the Stern boundary condition.
2196: 
2197: For our purposes here, it suffices to point out that the spatial
2198: profile of the first-order inner electric field in
2199: Eq.~(\ref{eq:E1master}) varies with the outer concentration,
2200: $\bar{c}_1(-1,\bar{t})$, at the slow time scale of bulk
2201: diffusion. Notably, this can lead to a secondary relaxation of the
2202: total diffuse charge, in response to the evolution of the diffusion
2203: layers. We observe this slow relaxation phase in our numerical
2204: solutions of the full equations, especially at large voltages. In
2205: particular, it is presumably associated with the non-monotonic
2206: charging profile for $v = 4$ shown in Fig.~\ref{fig:q}(c). A detailed
2207: analysis of this interesting effect from the setup above would require
2208: solving the first-order inner problem numerically, so we leave it for
2209: future work.
2210: 
2211: 
2212: \section {Strongly Nonlinear Dynamics }
2213: \label{sec:strong}
2214: 
2215: \subsection{ Steady State and the Dukhin Number }
2216: 
2217: We stress again that the asymptotic expansions derived above are valid
2218: in the limit of thin double layers, $\epsilon \rightarrow 0$, with the
2219: other two dimensionless parameters, $v$ (applied voltage) and $\delta$
2220: (relative compact-layer capacitance), held constant. For any fixed
2221: $\epsilon > 0$, there is no guarantee that the approximation remains
2222: accurate as the other parameters are varied. Having just calculated
2223: the bulk concentration to first order in the regular
2224: expansion, Eq.~(\ref{eq:cout}), we can now check {\it a posteriori}
2225: under what conditions it remains a good approximation.
2226: 
2227: A simple check involves the constant bulk concentration,
2228: Eq.~(\ref{eq:cinf}), after the charging process is completed.  The
2229: assumption that the first correction is much smaller than the leading
2230: term requires, $\alpha_s \equiv \epsilon \, \tilde{w}(\infty) \ll 1$.
2231: Linearizing Eq.~(\ref{eq:fzeta}) for $\delta \ll 1$, we can write this
2232: condition in a closed form:
2233: \begin{equation}
2234: 4 \, \epsilon \, \sinh^2 \left( \frac{v}{4(1+\delta)} \right) \ll 1
2235: \end{equation}
2236: Putting the units back, we have  
2237: \begin{equation}
2238: \alpha_s(\zeta_0) = \frac{4\lambda_D}{L} \, \sinh^2 \left( \frac{ze
2239: \zeta_0}{4 kT} \right) \ll 1 \label{eq:valid-steady}
2240: \end{equation}
2241: where $\zeta_0 \approx V/(1+\delta)$ is the steady-state zeta
2242: potential, long after the DC voltage is applied. 
2243: 
2244: The condition, $\alpha_s(\zeta_0) \ll 1$, for the validity of the {\it
2245: steady-state} asymptotic expansion is identical to that of small
2246: Dukhin number, $\Du(\zeta_0) \ll 1$, from Eq.~(\ref{eq:Du}) in the
2247: limit of no electro-osmosis ($m=0$), which may seem surprising since
2248: there is no surface conduction in our one-dimensional model
2249: problem. (Hence, we use the symbol $\alpha_s$ rather than $\Du$.)  The
2250: reason is that in both cases --- Dukhin's problem of electrophoresis
2251: of highly charged particles in weak applied fields and ours of electrode
2252: screening in strong applied fields --- the double layer absorbs a significant
2253: amount of neutral salt from the bulk (reverse Donnan effect). 
2254: 
2255: Net charge adsorption relative to the point of zero charge is measured by the
2256: {\it total} zeta potential,
2257: \begin{equation}
2258: \zeta_{tot} = \zeta_{eq} + \zeta_{ind}
2259: \end{equation}
2260: where $\zeta_{eq}$ is the uniform equilibrium zeta potential
2261: (reflecting the initial surface charge) and $\zeta_{ind}$ is the
2262: non-uniform induced zeta potential (resulting from diffuse-charge
2263: dynamics). In Dukhin's problem, the former may be large,
2264: $\Du(\zeta_{eq}) > 1$, but the latter is always small, $\zeta_{ind}
2265: \ll kT/ze$, so that the charging dynamics is linearized (or
2266: ignored). In our model problem, the situation is reversed: We assume
2267: $\zeta_{eq}=0$ (for simplicity), but we allow for a large applied
2268: voltage, $\zeta_{ind} \approx v/(1+\delta) > kT/ze$, in which case the
2269: dynamics is nonlinear. In both cases, the steady state is well
2270: described by weakly nonlinear asymptotics as long as
2271: $\alpha_s(\zeta_{tot}) = \Du(\zeta_{tot}) \ll 1$. When this condition is
2272: violated, double-layer charging and surface conduction may cause
2273: significant changes in the steady-state bulk concentration.
2274: 
2275: 
2276: \subsection{ Breakdown of Weakly Nonlinear Asymptotics }
2277: 
2278: In general, weakly nonlinear {\it dynamics} breaks down at
2279: somewhat smaller voltages, where $\zeta_{tot} > kT/e$ but
2280: $\alpha_s(\zeta_{tot}) = \Du(\zeta_{tot}) \ll 1$, because neutral-salt
2281: adsorption causes a {\it temporary, local depletion} of bulk
2282: concentration exceeding that of the steady state, after diffusional
2283: relaxation. In our model problem, the maximum change in bulk
2284: concentration occurs just outside the diffuse layers at $x = \pm 1$,
2285: just after the initial charging process finishes at time scale, $t =
2286: 1$ or $\bar{t} = \epsilon$. From 
2287: Eq.~(\ref{eq:cinit}), we have the first two terms of the weakly
2288: nonlinear asymptotic expansion there:
2289: \begin{equation}
2290: \bar{c}(\pm 1,\epsilon) \sim 1 - \sqrt{\frac{\epsilon}{\pi}} \,
2291:   \tilde{w}_0(\infty).
2292: \end{equation}
2293: At that time, the newly created diffusion layers have spread to
2294: $O(\sqrt{\epsilon})$ width, so the concentration is depleted locally
2295: by $O(\epsilon/\sqrt{\epsilon})=O(\sqrt{\epsilon})$, which is much
2296: more than the uniform $O(\epsilon)$ depletion remaining after bulk
2297: diffusion.
2298:  
2299: Therefore, in order for the time-dependent correction term to be
2300: uniformly smaller than the leading term, we need
2301: \begin{equation}
2302: \alpha_d \equiv  \sqrt{\frac{\epsilon}{\pi}} \, \tilde{w}_0(\infty) =
2303: \frac{\alpha_s}{\sqrt{\pi \epsilon}} \ll 1 .
2304: \end{equation}
2305: The relevant dimensionless parameter,  
2306: \begin{equation}
2307: \alpha_d(\zeta_{tot}) = 4 \sqrt{\frac{\lambda_D}{\pi L}} \, \sinh^2 \left( \frac{ze
2308: \zeta_{tot}}{4 kT} \right)  
2309: \end{equation}
2310: is larger than $\alpha_s(\zeta_{tot})$ (and the Dukhin number) by a factor
2311: of $\sqrt{L/\pi\lambda_D}$ in the limit of thin double layers.  For
2312: weakly nonlinear dynamics to hold, the applied voltage
2313: cannot greatly exceed the thermal voltage,
2314: \begin{equation}
2315: \zeta_{tot} \approx \frac{V}{1+\delta} < \frac{kT}{ze} \, \log \frac{
2316:   L}{\lambda_D} , \label{eq:valid-dyn}
2317: \end{equation}
2318: even for very thin double layers, $\lambda_D \ll L$, due to the
2319: logarithm.  In comparison, the applied voltage can be twice as large
2320: before the steady-state bulk concentration is significantly affected
2321: (and surface conduction becomes important in higher dimensions).
2322: This
2323: could have interesting consequences for induced-charge
2324: electrokinetic phenomena~\cite{bazant03} at moderate applied voltages
2325: where $\alpha_d > 1$ but $\alpha_s = \Du < 1$.
2326: 
2327: \subsection{ Strongly Nonlinear Asymptotics }
2328: 
2329: When condition (\ref{eq:valid-dyn}) is violated, electrochemical
2330: relaxation becomes much more complicated because double-layer charging
2331: is coupled to bulk diffusion. As long as $\alpha_d<1$, 
2332: however, the
2333: bulk remains quasi-neutral at all times. This regime of strongly
2334: nonlinear dynamics is demonstrated by the numerical solution in
2335: Fig.~\ref{fig:strong}, for $v = 4$, $\epsilon=0.05$, and $\delta=0.1$,
2336: in which case $\alpha_d = 0.545$.
2337: %
2338: % gnuplot> print 4*sqrt(0.05/pi)*(sinh(1.0/1.1)**2)
2339: % 0.545371019206225
2340: %
2341: In spite of the substantial $O(1)$
2342: amount of charge transfered from one diffuse layer to the other, each
2343: retains almost exactly the same $O(\epsilon)$ width as at lower 
2344: voltages, and bulk electro-neutrality remains an excellent
2345: approximation for all times. The initial charging process up
2346: to $t \approx 1$ creates a diffusion layer of neutral salt which
2347: relaxes into the bulk at the scale $\bar{t} \approx 1$ (or $t =
2348: \bar{t}/\epsilon \approx 20$).
2349: 
2350: \begin{figure*}
2351: \includegraphics[width=6in]{fig10.eps}
2352: \caption{ Strongly nonlinear charging dynamics for $v=4$ with
2353: $\epsilon=0.05$ and $\delta=0.1$. The potential (a), charge density
2354: (b), and concentration are shown in the half cell (top) and in the
2355: diffuse layer (bottom) for $t=0.5$ (solid), $1$ (dot), $2$ (dash), $4$
2356: (dot-dash), $8$ (dot-dot-dot-dash), and $20$ (long
2357: dash). \label{fig:strong} }
2358: \end{figure*}
2359: 
2360: In the strongly nonlinear regime, if $\alpha_s$ is not too small,
2361: double-layer charging is slowed down so much by nonlinearity that
2362: it continues to occur as the bulk diffusion layers evolve. One way to
2363: see this is that the effective RC time for the late stages of
2364: charging in Eq.(\ref{eq:jlate}) is
2365: \begin{equation}
2366: t_c(v) = C_i(v) \approx \cosh \frac{v}{2} \approx 2 \sinh^2 \frac{v}{4}
2367: \approx \frac{\alpha_s}{2\epsilon}  
2368: \end{equation}
2369: where we use the leading-order approximation of the differential
2370: capacitance, Eq. (\ref{eq:Ca}), for $\delta \ll 1$. In units of the
2371: bulk diffusion time, the nonlinear relaxation time is $\bar{t}_c =
2372: \epsilon t_c = \alpha_s/2$. 
2373: 
2374: To make analytical progress, one would consider the joint limits
2375: \begin{equation}
2376: \epsilon \rightarrow 0 \ \ \mbox{ and } \ \ v \rightarrow \infty \ \
2377: \mbox{with } \ \ \alpha_d(v)>0 \mbox{ fixed}
2378: \end{equation}
2379: and expect the approximations to remain acceptable at somewhat
2380: larger voltages, as long $\alpha_s(v) < 1$.  Such analysis is beyond
2381: the scope of this article, but at least we indicate how the leading
2382: order approximation would be calculated. (Going beyond leading order
2383: seems highly nontrivial.) 
2384: 
2385: At leading order in the bulk, we have the
2386: usual equations for a neutral binary electrolyte (with equal ionic diffusivites),
2387: \begin{equation}
2388: \frac{\partial \bar{c}_0}{\partial \bar{t}} = \frac{\partial^2
2389: \bar{c}_0}{\partial x^2} \ \ \ \mbox{and} \ \ \ \frac{\partial
2390: }{\partial x} \left(\bar{c}_0 \, \frac{\partial
2391: \bar{\phi}_0}{\partial x}\right) = 0
2392: \end{equation}
2393: with $\bar{\rho} = O(\epsilon^2)$. Integrating the second equation, we
2394: obtain a constant, uniform current density, $\bar{j}_0(t)$,  as before,
2395: but the electric field is modified by concentration polarization,
2396: \begin{equation}
2397: \frac{\partial \bar{\phi}_0}{\partial x} =
2398: \frac{\bar{j}_0(t)}{\bar{c}_0(x,\bar{t})}.
2399: \end{equation}
2400: The effective boundary conditions come from asymptotic matching with
2401: the diffuse layers as before, 
2402: \begin{equation}
2403: \epsilon \, \frac{d\tilde{q}_0}{d\bar{t}} = \bar{j}_0(t) \ \ \ \mbox{and
2404: } \ \ \ \epsilon \, \frac{d\tilde{w}_0}{d\bar{t}} =  \frac{\partial
2405: \bar{c}_0}{\partial x}(-1,\bar{t})
2406: \end{equation}
2407: only now the diffusive flux entering the diffuse layers (second equation)
2408: appears at leading order.  The ionic concentrations retain 
2409: Gouy-Chapman equilibrium profiles modified
2410: quasi-statically by the evolving nearby bulk concentration:
2411: \begin{eqnarray}
2412: \tilde{q}_0(\bar{t}) &=& - 2 \sqrt{\bar{c}_0(-1,\bar{t})}
2413: \sinh\left(\frac{\tilde{\zeta}_0(\bar{t})}{2}\right)   \\
2414: \tilde{w}_0(\bar{t}) &=& 4 \sqrt{\bar{c}_0(-1,\bar{t})}
2415: \sinh^2\left(\frac{\tilde{\zeta}_0(\bar{t})}{4}\right)  
2416: \end{eqnarray}
2417: where
2418: \begin{equation} 
2419: \tilde{\zeta}_0(\bar{t}) - \tilde{q}_0(\bar{t})\,  \delta =
2420: \tilde{\Psi}_0(\bar{t})  = - v -
2421: \bar{\psi}_0(-1,\bar{t}) .
2422: \end{equation}
2423: It seems exact solutions are not possible in terms of elementary
2424: functions.
2425: The equations are ``stiff'', since they involve a short time
2426: scale, $\bar{t} = \epsilon$, for the initial phases of charging, but
2427: at least the spatial boundary layers have been ``integrated out'',
2428: which is convenient for numerical solutions.
2429: 
2430: \subsection{ Space Charge at Very Large Voltages }
2431: 
2432: We close this section by noting some intriguing, new possibilities, 
2433: further into the strongly nonlinear regime.  At large voltages, such
2434: that $\alpha_d > 1$, it seems a {\it transient space charge layer}
2435: should form since the bulk concentration would be depleted almost
2436: completely near the diffuse layers by the initial charging process.
2437: In steady-state problems of Faradaic conduction, it is well known that
2438: double-layer structure is altered from its Gouy-Chapman equilibrium
2439: profile at a limiting current~\cite{smyrl67} and may turn into an
2440: extended space charge layer above a limiting
2441: current~\cite{rubinstein79}, but here we see that similar effects may
2442: also occur temporarily with large time-dependent voltages, in the
2443: absence of any Faradaic processes (at blocking electrodes).  At still
2444: larger voltages, such that $\alpha_s > 1$, double layer charging
2445: consumes most of the bulk concentration, presumably leaving the entire
2446: bulk region in a state of ``space charge''.  
2447: 
2448: Such situations may seem quite exotic in macroscopic systems, where
2449: $\epsilon = \lambda_D/L$ is extremely small, but in microsystems
2450: perhaps they could occur. The mathematical model neglects bulk
2451: reactions (e.g. leading to hydrogen bubble formation), nonlinear
2452: dielectric properties, electro-convection, or other effects which may
2453: hinder the formation of space charge in real systems. Nevertheless,
2454: the rich nonlinear behavior of the model merits further mathematical
2455: study, as a challenging problem in time-dependent boundary-layer
2456: theory.
2457: 
2458: 
2459: 
2460: \section{ Beyond the Model Problem }
2461: \label{sec:concl}
2462: 
2463: We conclude by discussing more general situations, which contain some
2464: new physics, absent in our simple model problem.  For thin double layers,
2465: the same methods of asymptotic analysis could be applied to derive
2466: effective equations in which the double layers are incorporated into
2467: boundary conditions, better suited for analytical or numerical
2468: work. Here, we simply sketch the results and suggest some other model
2469: problems for further study. 
2470: 
2471: \subsection{ Two or More Dimensions }
2472:  
2473: In the weakly nonlinear regime, where $\alpha_d < 1$ for all times
2474: over all double layers, our analysis  extends trivially to higher
2475: dimensions, as long as the surface curvature does not introduce another
2476: length scale much
2477: smaller than $L$. In that case, the double layers are
2478: locally ``flat'', and the boundary-layer calculations remain
2479: unchanged. Following the same procedure, we find that the bulk concentration is
2480: uniform at leading order, $\bar{c}_0 = 1$, and the bulk potential,
2481: $\bar{\phi}(\rb,t)$, is a harmonic function,
2482: \begin{equation}
2483: \nabla^2 \bar{\phi}_0 = 0 ,  \label{eq:laplace}
2484: \end{equation}
2485: subject to a (dimensionless)  RC boundary condition at each
2486: electrode surface,
2487: \begin{equation}
2488: \frac{\partial \tilde{q}_0}{\partial t} = \tilde{C}(\bar{\phi}_0 -
2489: \phi_e) \, \frac{\partial (\bar{\phi}_0-\phi_e)}{\partial t} =
2490: \nb\cdot\del \bar{\phi}_0 ,   \label{eq:RCBC}
2491: \end{equation}
2492: where $\nb$ is the unit normal pointing into the electrolyte and  
2493: $\phi_e(\rb,t)$ is the local electrode potential relative to the
2494: solution. The latter is equal to the local applied voltage plus the
2495: equilibrium zeta potential:
2496: \begin{equation}
2497: \phi_e(\rb,t) = V(\rb,t) + \zeta_{eq}(\rb)
2498: \end{equation}
2499: which accounts for any pre-existing double-layer charge (neglected in
2500: our calculations above).  A Neumann boundary condition,
2501: $\nb\cdot\del\bar{\phi}_0$, is imposed at any inert, non-polarizable
2502: %%%% QQQ Missing right hand side in the above boundary condition????
2503: % NO. The prior eq is the boundary condition (it is an ODE).
2504: surface, such as a channel side wall. 
2505: 
2506: Another complication in two or more dimensions is the possibility of
2507: electro-osmotic flow. The fluid velocity in the bulk usually satisfies
2508: the Stokes equations, which may be unsteady for high-frequency
2509: forcing. In the weakly nonlinear regime, the classical
2510: Helmholtz-Smoluchowski formula gives the fluid slip in terms of the
2511: local zeta potential and tangential bulk electric
2512: field~\cite{hunter,russel,lyklema}.
2513: 
2514: Equations ~(\ref{eq:laplace}) and (\ref{eq:RCBC}) model the
2515: electrolyte as a bulk Ohmic resistor with a capacitor skin at
2516: electrode interfaces.  The linearized version of these equations (with
2517: $\tilde{C}$ = constant) has been studied extensively, e.g.\ in the
2518: context of metallic colloids~\cite{murtsovkin96,simonov77}, AC
2519: electro-osmosis~\cite{ramos99,green00a,gonzalez00}, AC
2520: pumping~\cite{ajdari00}, and other phenomena of induced-charge
2521: electro-osmosis~\cite{bazant03,squires03}. The nonlinear version,
2522: however, has apparently not been analyzed, even though it may have
2523: relevance for experiments, in which the condition, $v \ll 1$ ($V \ll
2524: kT/ze$), is routinely violated.
2525: 
2526: More significant modifications arise at leading order in the strongly
2527: nonlinear regime (or at higher order in the weakly nonlinear regime). 
2528: Ohm's law breaks down due to
2529: concentration gradients, as the double layers absorb a significant
2530: amount of neutral salt from the bulk. In two or more dimensions, the
2531: dimensionless leading-order equations for $\bar{c}_0(\rb,t)$ and
2532: $\bar{\phi}_0(\rb,t)$ in section~\ref{sec:strong} take the form, 
2533: \begin{equation}
2534: \frac{\partial \bar{c}_0}{\partial \bar{t}} = \nabla^2\bar{c}_0,
2535: \ \ \ \del\cdot(\bar{c}_0\del\bar{\phi}) = 0 ,
2536: \end{equation}
2537: where we scale time to the bulk diffusion time. This assumes $\alpha_d
2538: < 1$ so that no transient space charge layers form. 
2539: 
2540: The effective boundary
2541: conditions still involve the small parameter, $\epsilon$, as in one dimension, since
2542: the natural scale is the RC charging time, but there are some new
2543: terms in higher dimensions:
2544: \begin{eqnarray}
2545: \epsilon\, \frac{\partial \tilde{q}_0}{\partial \bar{t}} & = & 
2546: \nb\cdot(\bar{c}_0 \del \bar{\phi}_0) -   
2547: \Du \, \del_s \cdot \tilde{J}_s   \label{eq:Js} \\
2548: \epsilon\, \frac{\partial \tilde{w}_0}{\partial \bar{t}} & = & 
2549: \nb\cdot \del \bar{c}_0 -   
2550: \Du \, \del_s \cdot (\tilde{D}_s \del_s \tilde{w}_0)  \label{eq:Ds}
2551: \end{eqnarray}
2552: The last term in Eq.~(\ref{eq:Js}) is the surface divergence of the
2553: (leading-order) dimensionless 
2554: tangential current, $\tilde{J}_s$, in the diffuse layer; the size of this term compared
2555: to the normal current is governed by a Dukhin number, based on the
2556: largest expected total zeta potential. Similarly, the last term in
2557: Eq.~(\ref{eq:Ds}) is the surface divergence of the tangential diffusive
2558: flux in the diffuse layer, where $\tilde{D}_s$ is a dimensionless surface
2559: diffusivity; again, this term is of order $\Du$ smaller than the
2560: normal diffusive flux.  
2561: 
2562: Formulae for $\tilde{J}_s$ and $\tilde{D}_s$ can be derived
2563: systematically using the matched asymptotic expansions, which is
2564: beyond the scope of this paper. The classical results of
2565: Bikerman~\cite{bikerman33,bikerman35} and Deryagin and
2566: Dukhin~\cite{deryagin69} are available for the case of weak applied
2567: voltages ($\zeta_{ind}\ll kT/e$) and large equilibrium surface charges
2568: ($\zeta_{eq} > kT/e$, $\Du(\zeta_{eq})\approx 1$), and many Russian
2569: authors have studied electrokinetic phenomena in this
2570: regime~\cite{dukhin80,dukhin93}.  The case of strongly nonlinear
2571: dynamics ($\zeta_{ind} > kT/e$,  $\alpha_d(\zeta_{ind}) \approx 1$),
2572: however, should be revisited in more detail to see if any changes
2573: arise for strong, time-dependent applied voltages. We suggest as a
2574: basic open question analyzing the electrochemical response of a metal
2575: cylinder or sphere in a strong, suddenly applied, uniform background
2576: DC field.
2577: 
2578: Another interesting issue is the stability of our one-dimensional
2579: solution. One should consider small space-dependent perturbations of
2580: the solution at various large voltages, in both the weakly nonlinear
2581: and strongly nonlinear regimes. The general transient analysis in two
2582: or more dimensions with the same equations and boundary conditions
2583: presents an interesting challenge.
2584: 
2585: 
2586: \subsection{ General Electrolytes and Faradaic Reactions }
2587: 
2588: Even in one dimension, it would be interesting to extend our analysis
2589: to more general situations involving asymmetric or multicomponent
2590: electrolytes, which undergo Faradaic processes at electrode surfaces. 
2591: Restoring dimensions, the bulk electrolyte is described by the $N$ ionic
2592: concentrations, $C_i$, $i=1,2,\ldots,N$, satisfying mass conservation,
2593: \begin{equation}
2594: \frac{\partial C_i}{\partial \tau} = - \del\cdot\Fb_i
2595: \end{equation}
2596: where $\Fb_i$ is the flux density due to diffusion and
2597: electromigration,
2598: \begin{equation}
2599: \Fb_i = - D_i \del C_i - \mu_i z_i e C_i \del \Phi  
2600: \end{equation}
2601:  as in Eq.~(\ref{eq:eqdim}). For thin double layers, at leading order 
2602: the bulk remains neutral (as long as $\alpha_d < 1$ to avoid space 
2603: charge formation), so the potential is determined implicitly by the condition of
2604: electroneutrality, 
2605: \begin{equation}
2606: \rho_e = \sum_{i=1}^N z_i e C_i = 0 .
2607: \end{equation}
2608: These are the standard equations of bulk
2609: electrochemistry~\cite{newman}, but interesting physical effects are
2610: contained in the effective boundary conditions. 
2611: 
2612: Generalizing the total surface charge density $q$ and excess surface
2613: concentration $w$, we define $\Gamma_i$ to be the surface
2614: concentration of species $i$ absorbed in the diffuse layer. To be precise, it
2615: is the integral of the leading-order excess concentration relative to
2616: the bulk over the inner coordinate, as in Eqs.~(\ref{eq:qt}) and
2617: (\ref{eq:w}). For example, $q = ze(\Gamma_+ - \Gamma_-)/2$ and $w = (\Gamma_++
2618: \Gamma_-)/2$ for a symmetric binary electrolyte. 
2619: 
2620: Following the procedure above, the boundary
2621: conditions on the leading-order bulk approximation are of the form:
2622: \begin{equation}
2623: -\frac{\partial \Gamma_i}{\partial \tau} = \nb\cdot \Fb_i +
2624: \del_s\cdot\Fb_{si} + R_i
2625: \end{equation}
2626: where $\Fb_{si}(C_i,\Phi)$ is the surface flux density of species $i$ in the
2627: double layer~\cite{deryagin69} and $R_i(\{C_i\},\Phi)$ is the reaction-rate density
2628: for any Faradaic processes consuming (or producing) species $i$ at the
2629: surface. The usual assumption for $R_i$ involves Arrhenius kinetics,
2630: as in the Butler-Volmer equation, but the Frumkin correction for
2631: concentration variations across the diffuse layer must be taken into
2632: account~\cite{newman,bard}. 
2633: 
2634: The general system of nonlinear equations is challenging to solve, even
2635: numerically, due to multiple length and time scales. Boundary-layer
2636: theory provides only a partial simplification by integrating out the
2637: smallest length scale. As described in
2638: section~\ref{sec:history}, various special cases of the effective
2639: equations have been considered in the literature,
2640: but much remains to be done, especially for strongly nonlinear
2641: dynamics in large applied voltages. In micro-electrochemical or
2642: biological systems, this regime is easily reached, so it merits
2643: additional mathematical study and comparison with experimental data,
2644: in part to test the applicability of the Nernst-Planck equations in
2645: micro-systems. Another interesting aspect 
2646: is the coupling of electrochemical dynamics to fluid flow, which is
2647: finding new applications in microfluidic devices.
2648: 
2649: \acknowledgments
2650: 
2651: This work was supported in part by the MRSEC Program  of the National
2652: Science Foundation under award number DMR 02-13282 (MZB), by ESPCI
2653: through the Paris Sciences Chair (MZB), and by an ACI from the
2654: Minist\`ere de la Recherche (AA). We are grateful to K. Chu and the
2655: MIT librarians for help in obtaining articles from the ``ancient''
2656: and Russian literatures.
2657: 
2658: %\references
2659: \begin{thebibliography}{}
2660: 
2661: \include{art-bib}
2662: 
2663: \end{thebibliography}
2664: \end{document}
2665: