1: \documentclass[aps,floats,prb,showpacs,twocolumn]{revtex4}
2:
3: \newcommand{\beq}{\begin{eqnarray}} \newcommand{\eeq}{\end{eqnarray}}
4:
5: \def\eq{&=&}
6:
7: \usepackage{amsfonts,amsmath} \usepackage{bm} \usepackage{dcolumn}
8: \usepackage{epsfig} \usepackage{latexsym}
9:
10: \begin{document}
11:
12: \title{Electron transport in two--dimensional arrays}
13:
14: \author{Julia S. Meyer$^1$, Alex Kamenev$^1$, and Leonid I.
15: Glazman$^{1,2}$}
16:
17: \affiliation{$^1$Department of Physics, and $^2$W.I. Fine Theoretical
18: Physics Institute,\\
19: University of Minnesota, Minneapolis, MN55455, USA}
20:
21: \date{\today}
22:
23: \pacs{73.23.-b, 73.23.Hk, 71.45.Lr, 71.30.+h }
24:
25: \begin{abstract}
26: We study charge transport in a granular array with high inter--grain
27: conductances. We show that the system exhibits a
28: Berezinskii--Kosterlitz--Thouless crossover from the
29: high--temperature conducting state into a low--temperature
30: insulating state. The crossover takes place at a critical
31: temperature $T_{\rm BKT}\propto E_c\exp\{-g\}$, where $E_c$ is the
32: charging energy of a grain and $g\gg 1$ is the dimensionless
33: inter--grain conductance. A uniformly applied gate voltage drives
34: the insulator into a conducting charge liquid state followed by an
35: insulating lattice--pinned Wigner crystal state at larger values of
36: the gate voltage. Technically, we establish correspondence between
37: the charge and phase representations, employing the instanton gas
38: summation in the framework of the phase model.
39: \end{abstract}
40:
41: \maketitle
42:
43: \section{Introduction}
44:
45: Granular arrays have recently attracted much attention as analytically
46: tractable systems to study the interplay of interactions and
47: scattering~\cite{Gerber97,general,Beloborodov01,tschersich,AGK}. The
48: advantage of granular systems is the possibility to separate the
49: scattering--induced quantum interference phenomena from
50: electron--electron interaction effects. Quantum coherence is relevant
51: as long as the typical dwell--time in a grain $\tau_{\rm dwell}\sim
52: \hbar/(g\delta)$ is longer than the dephasing time $\tau_\varphi$.
53: Here $g$ is the dimensionless inter--grain conductance measured in
54: units of $e^2/(2\pi\hbar)$, and $\delta$ is the mean level spacing of
55: the grains. For two--dimensional disordered interacting systems the
56: dephasing time is known to be~\cite{AAK82} $\tau_\varphi\approx \hbar
57: g/T$. The quantum interference is thus suppressed if $\tau_\varphi <
58: \tau_{\rm dwell}$, which is the case at $T>g^2\delta$. Hereafter we
59: assume this condition to be satisfied and consider the temperature
60: range $g^2\delta<T<E_c$ (where $E_c=e^2/(2C)$ is the charging energy
61: of the grains). This allows us to focus on the interaction--induced
62: phenomena, while omitting the interference (incoherent regime).
63:
64: Earlier studies of incoherent two--dimensional arrays led to
65: conflicting theoretical results~\cite{tschersich,fs2,fs}. In
66: Refs.~[\onlinecite{fs2,fs}] the low--temperature insulating state was
67: found for sufficiently small inter--grain conductance, $g<1$. Upon
68: elevating the temperature, the array was shown to undergo a
69: Berezinskii--Kosterlitz--Thouless (BKT) transition into the conducting
70: state. The transition temperature $T_{\rm BKT}=T_{\rm BKT}(g)$ was
71: predicted to go to zero at some critical value of conductance
72: $g_c\approx 1.8$. At larger conductances, $g> g_c$, the metallic state
73: was claimed to persist down to zero temperature. In
74: Ref.~[\onlinecite{tschersich}] the same model as in
75: Refs.~[\onlinecite{fs2,fs}] was studied in the regime of large
76: conductances, $g\gg 1$. Using a perturbative renormalization group
77: (RG) analysis, the renormalized inter--grain conductance was shown to
78: behave as $g\Rightarrow g(T)=g-{2\over d}\ln(E_cg/T)$ (where $2d$ is
79: the coordination number of the lattice). This correction is
80: essentially similar (and at $T\approx g\delta$ crosses
81: over~\cite{vinokur}) to the interaction--induced
82: Altshuler--Aronov~\cite{altshuler-aronov} conductivity corrections
83: known for homogeneous disordered systems. At $T\sim E_c
84: g\exp\{-dg/2\}$, the conductance is renormalized down to $g(T)\sim
85: {\cal O}(1)$. Thus, one may expect that the system approaches an
86: insulating state at low enough temperature even for large bare
87: inter--grain conductance. Whether such a ``high--$g$'' insulator
88: indeed exists, and -- if so -- its nature and relation (if any) to the
89: BKT transition, found~\cite{fs2,fs} for ``low--$g$'' systems, was not
90: clarified.
91:
92: In the present work we show how to reconcile these findings. In
93: particular, we show that there exists a finite $T_{\rm BKT}\propto
94: \exp\{-g\}$ at $g\gg1$. To this end one needs to go beyond the
95: perturbative analysis~\cite{tschersich} of the model considered in
96: Refs.~[\onlinecite{tschersich,fs2,fs}] and include non--perturbative
97: -- instanton -- field configurations. A similar program was recently
98: carried out for one--dimensional incoherent arrays~\cite{AGK}. The
99: conductivity of one--dimensional arrays was found to display activated
100: (insulating) behavior with the charge gap $\sim E_c\exp\{-g/4\}$,
101: which is parametrically larger than the temperature where the
102: perturbative corrections become large. It was also shown that the
103: proper low--temperature representation of incoherent arrays is that of
104: pinned {\em charge}--density wave fluctuations. [This should be
105: contrasted with the fluctuating phase (or voltage) picture employed in
106: Refs.~[\onlinecite{tschersich,fs2,fs}].] The activation gap
107: corresponds to the energy needed to create a long unit--charge
108: soliton.
109:
110: In the present paper the charge representation is derived and analyzed
111: for the two--dimensional setup. We find that the charge excitations
112: are localized unit--charge two--dimensional solitons. At $g\gg 1$ the
113: solitons interact logarithmically over a large range of distances.
114: This leads to a sharp BKT crossover~\cite{BKT-b,BKT-kt} between a
115: low--temperature insulating phase with bound charge--anti-charge pairs
116: (and an exponentially small number of free charges) and a
117: high--temperature conducting phase, where the pairs are unbound. The
118: BKT temperature $T_{\rm BKT}(g)$ remains finite (though exponentially
119: small), $T_{\rm BKT}(g)\sim E_c g \exp\{-g\}$, for an {\em arbitrarily
120: high} bare conductance $g$. The zero--temperature quantum phase
121: transition at $g=g_c$, found in Refs.~[\onlinecite{fs2,fs}], thus,
122: does {\em not} exist. Instead, there is a fast but continuous drop of
123: the transition temperature $T_{\rm BKT}(g)$ in the vicinity of $g\sim
124: 1$.
125:
126: The issue of whether a classical phase transition or a crossover
127: occurs as the temperature is lowered, depends sensitively on the
128: details of the model. The true BKT transition takes place only if the
129: interactions between the charged solitons are logarithmic at
130: arbitrarily large distances. This is the case for arrays with
131: inter--grain capacitances only (in the absence of the grain's
132: self--capacitance, no electric field lines can leave the 2d plane).
133: In the presence of the self--capacitance, the interaction is
134: logarithmic in a wide, but finite range of distances:
135: $1<l\lesssim\exp\{g/2\}$ (hereafter distances are measured in units of
136: the array lattice spacing). In the latter case, below the BKT
137: temperature the array's conductivity is not zero (in contrast to the
138: former case), but rather exhibits the activation behavior,
139: \begin{equation}
140: \label{eq-res}
141: \sigma\simeq g\exp\left\{-\frac{\Delta}T\;\right\},
142: \end{equation}
143: where the activation gap is given by $\Delta \simeq g T_{\rm BKT}\gg
144: T_{\rm BKT}$. Upon raising the temperature, above the BKT temperature
145: the conductivity sharply increases as
146: \begin{equation}
147: \label{eq-above}
148: \sigma=gK\exp\left\{-2b\sqrt{\frac{T_{\rm BKT}}{T-T_{\rm BKT}}}\right\},
149: \end{equation}
150: where $K$ and $b$ are non-universal constants of order
151: unity~\cite{mooij}. Finally, above the transition region, the
152: conductivity crosses over to the perturbative prediction
153: $\sigma=g-\ln(gE_c/T)$. We thus have a generally consistent picture
154: based on the BKT physics at any value of bare conductance, $g$.
155:
156: A gate voltage induces a uniform background charge $q\in [0,1]$.
157: However, for small gate voltages the array remains in the
158: particle-hole symmetric state with an integer number of electrons per
159: dot. The transition into a non--uniform state (with a non--integer
160: average number of electrons per dot) takes place at a critical
161: dimensionless charge density $q^*=\Delta/(2E_c)$. Its physics is
162: similar to the transition from the Meissner to the Abrikosov state in
163: type II superconductors upon increasing an external magnetic field. In
164: our case the role of the magnetic field is played by the gate voltage,
165: $q$, with $q^*$ corresponding to the lower critical field $H_{c1}$.
166: The Abrikosov lattice in turn corresponds to the 2d Wigner crystal
167: formed by the unit--charge solitons. Such a crystal is easily pinned
168: by the lattice. Consequently, at low temperatures, the array is in
169: the insulating phase with a residual activation conductivity,
170: associated with the thermal creation of defects. The Wigner crystal
171: melts at a temperature~\cite{Tm} of the order of $T_{\rm BKT}$,
172: leading to a sharp crossover into the conducting phase.
173:
174: Methodologically, interacting systems may be modeled in two
175: alternative ways: in terms of either phase or charge degrees of
176: freedom. The two are canonically conjugated and the choice between
177: them is a matter of convenience. The phase representation is easier to
178: derive microscopically starting from the fermionic tunneling
179: Hamiltonian. By this reason it was used in the vast majority of works
180: on granular systems and quantum
181: dots~\cite{general,Beloborodov01,tschersich,fs,fs2}. We found it more
182: convenient, however, to work in the charge representation, which is
183: the natural language to describe the insulating phase. In the quantum
184: dot context, the charge description was introduced in
185: Refs.~[\onlinecite{flensberg,matveev}]. Here we employ its
186: generalization to 2d granular arrays. We introduce the model in
187: Sec.~\ref{sec-charge} and show that it exhibits the BKT crossover in
188: Sec.~\ref{sec-BKT}. Finite gate voltages and the pinned Wigner crystal
189: phase are discussed in Sec.~\ref{sec-gate}. To facilitate comparison
190: with the body of work on the phase representation, we include the
191: proof of equivalence of the two models in Appendix~\ref{app-phase}.
192:
193: \section{Charge representation}
194: \label{sec-charge}
195:
196: In this section, we introduce the charge representation for incoherent
197: ($\delta\to 0$) interacting arrays. To keep the presentation compact,
198: we assume all contacts to be single--channel, characterized by a
199: reflection amplitude $r<1$. The generalization to the multichannel
200: case and, in particular, to $g\gg1$ is discussed at the end of the
201: section and, in more detail, in Appendix~\ref{app-gamma}, while the
202: proof of the equivalence of the resulting charge model to the more
203: widely used phase model is outlined in Appendix~\ref{app-phase}.
204:
205: \subsection{Single contact}
206:
207: Consider a point contact between a quantum dot and a metallic
208: reservoir. Such a point contact allows for a small number of
209: propagating transverse modes which may be thought of as
210: one--dimensional electron liquids (with the contact situated at the
211: origin, $z=0$). Here we consider the case of a single propagating
212: mode, deferring the consideration of multi--mode contacts to
213: Appendix~\ref{app-gamma}. The corresponding one--dimensional electron
214: liquid may be bosonized in the conventional
215: way~\cite{flensberg,matveev} and described in terms of the bosonic
216: field $\theta(\tau,z)$. Its gradient $\partial_z \theta(\tau,z)$ has
217: the meaning of a local electron density. As a result, the electron
218: number on the dot may be written as $N=\int_0^\infty\! dz\, \partial_z
219: \theta(\tau,z)=-\theta(\tau,0)$ and, thus, the Coulomb energy takes
220: the form $(eN)^2/(2C)=E_c\theta^2(\tau,0)$. Finally the
221: imaginary--time action of the bosonic field reads
222: \begin{eqnarray}
223: \label{eq-dot}
224: S[\theta(\tau,z)] &= &\int\limits_0^\beta d\tau
225: \left\{ \int\limits_{-\infty}^{\infty}dz
226: \left[(\partial_\tau\theta)^2+(\partial_z\theta)^2\right]\right.
227: \nonumber \\
228: && + \left. E_c\theta^2(\tau,0) -
229: \frac{Dr}\pi\cos[2\pi\theta(\tau,0)]\right\}.
230: \end{eqnarray}
231: The last term in this expression describes backscattering at the point
232: contact with the reflection amplitude $r$, while $D$ is the electronic
233: bandwidth.
234:
235: One may integrate out all degrees of freedom with $z\neq 0$, retaining
236: the field $\theta(\tau)\equiv \theta(\tau,0)$ only. The corresponding
237: action reduces to
238: \begin{equation}
239: \label{eq-dot1}
240: S[\theta] = \frac1T\sum_m \left( \pi|\omega_m|\theta_{m}^{2}+
241: E_{\rm c}\theta_{m}^2 \right) - \frac{Dr}\pi\!\int\limits_0^\beta
242: \!d\tau\,\cos(2\pi\theta(\tau)),
243: \end{equation}
244: where $\omega_m=2\pi Tm$, and we have introduced the Matsubara
245: representation though the transformation $\theta_m=\int_0^\beta
246: d\tau\, \theta(\tau)e^{-i\omega_m\tau}$. The dissipative term,
247: $\pi|\omega_m|\theta_{m}^{2}$, is generated as a result of integrating
248: out the continuum spectrum of the degrees of freedom on the dot. Its
249: appearance is a consequence of the assumption that the mean level
250: spacing is the smallest energy scale in the model, $\delta\to 0$.
251:
252: \subsection{2d array}
253: \label{subsec-array}
254:
255: We now generalize the single--contact action, Eq.~(\ref{eq-dot1}), to
256: the 2d array geometry. To this end we introduce the vector index ${\bf
257: l}$ to label the grains. We also introduce two fields
258: $\theta_{x,{\bf l}}(\tau)$ and $\theta_{y,{\bf l}}(\tau)$ which
259: describe charge transport from grain ${\bf l}$ in the positive $x$ and
260: $y$ directions, respectively. In these notations, the instantaneous
261: electron density on the grain ${\bf l}$ is given by the lattice
262: divergence $\nabla\cdot\vec\theta_{\bf l} \equiv \theta_{x,{\bf
263: l+e}_x}-\theta_{x,{\bf l}}+\theta_{y,{\bf l+e}_y}-\theta_{y,{\bf
264: l}}$ (cf.~Fig.~\ref{fig1}). With the backscattering in the
265: contact, characterized by the reflection amplitude $r$, the action
266: reads
267: \begin{widetext}
268: \begin{equation}
269: \label{eq-matveev} S\!\left[\,\vec\theta\,\right] = \sum_{\bf l}
270: \left\{\frac1T\sum_m \left( \pi|\omega_m|\vec\theta_{{\bf
271: l},m}^{\;2} + E_{\rm c}(\nabla\cdot\vec\theta_{{\bf l},m})^2
272: \right) - \frac{Dr}\pi\sum_{i=x,y}\int\limits_0^\beta d\tau\;
273: \cos(2\pi\theta_{i,{\bf l}}(\tau))\right\},
274: \end{equation}
275: \end{widetext}
276: where $D$ is again the bandwidth. As in Eq.~(\ref{eq-dot1}), the first
277: term in the action~(\ref{eq-matveev}) describes the dissipative
278: dynamics originating from integrating out degrees of freedom within
279: the grains, the second term is responsible for the charging, and the
280: third one describes backscattering in the contacts.
281:
282: \begin{figure}[h]
283: \centerline{\epsfxsize=3in\epsfbox{Fig1.eps}}
284: \caption{Granular array with a square lattice. The
285: massless modes $\eta$, explained in the text, correspond to
286: circular currents around a plaquette and, therefore, do not
287: contribute to charge transport.}
288: \label{fig1}
289: \end{figure}
290:
291: For a square lattice of linear size $M$, the array contains $M^2$
292: grains and $2M^2$ contacts between them. Consequently the model is
293: written in terms of $2M^2$ bosonic degrees of freedom, $\theta_{i,{\bf
294: l}}(\tau)$. In the limit $r\to 0$, the masses of these modes are
295: provided only by the $M^2$ charging terms $E_{\rm
296: c}(\nabla\cdot\vec\theta_{{\bf l},m})^2$. Therefore, if the
297: backscattering is neglected, only half of the degrees of freedom of
298: the model are massive. To see this explicitly, one may rewrite the 2d
299: vector field $\vec \theta$ through two scalar fields
300: $\vec\theta=\nabla{\chi}+\nabla\times\eta$ and notice that the
301: charging term contains only the field $\chi$ while the curl--field
302: $\eta$ fully decouples from it.
303:
304: In order to find an effective low-energy theory, we shall proceed with
305: the renormalization group (RG) scheme based on the integration of the
306: high frequency Matsubara modes, $D'<|\omega_m|<D$, accompanied by the
307: appropriate change of the backscattering amplitude $r$. As long as the
308: coefficient in front of the cosine--term in Eq.~(\ref{eq-matveev}) is
309: less than the running bandwidth $D'$, one may treat the fields as
310: Gaussian, governed by the first two terms in Eq.~(\ref{eq-matveev}).
311: As a result, the backscattering amplitude renormalizes as
312: \begin{equation}
313: \label{eq-renormalization}
314: Dr\enspace\Rightarrow\enspace Dr \exp\left\{-{(2\pi)^2\over 2}\langle
315: \theta^{2}_i\rangle\right\}\, ,
316: \end{equation}
317: where the averaging in $\langle \theta^{2}_i\rangle$ is performed over
318: high--frequency fluctuations. Passing to the momentum representation
319: and taking into account both $\chi$ and $\eta$ components of the
320: fluctuations, one finds
321: \begin{eqnarray}
322: \langle\theta^{2}_i\rangle\! &=& \!\frac
323: T{4M^2}\!\!\!\sum\limits_{|\omega_m|=D'}^{D}\sum\limits_{
324: q_x,q_y=1}^{M}\left(\frac1{E_{\bf
325: q}\!+\!\pi|\omega_m|}\!+\!\frac1{\pi|\omega_m|}\right)\!
326: \nonumber \\
327: &\simeq&
328: \!\frac1{4\pi^2}\left(\ln\frac{ D}{E_c}+\ln\frac{D}{D'}
329: \right)=\frac1{2\pi^2}\ln\frac D{\sqrt{E_c D'}}\, ,
330: \label{eq-theta2}
331: \end{eqnarray}
332: where $E_{\bf q}=4E_c\sum_{i=x,y}\sin^2(\pi q_i/(2M))$ is the mass
333: spectrum of the $\chi$ modes. In the second line we have assumed that
334: $D'<E_c$ (in the opposite case, $D'>E_c$, one should substitute
335: $\sqrt{E_c D'}$ by $D'$). Notice that the presence of the lower limit,
336: $D'$, in this expression is due to the massless rotational modes of
337: the field $\eta$. [Note also the difference with the 1d--system, where
338: all modes are massive and, therefore, the result corresponding to
339: Eq.~(\ref{eq-theta2}) is independent on the lower limit~\cite{AGK}.]
340: Combining Eqs.~(\ref{eq-renormalization}) and (\ref{eq-theta2}), one
341: finds that upon integrating out the high--frequency modes, the
342: coefficient of the cosine potential renormalizes as $Dr \Rightarrow
343: \sqrt{E_cD'}\,r$. As was discussed above, this procedure works as long
344: as $\sqrt{E_cD'}\,r<D'$, that is $D'>T_0$, where $T_0= E_c r^2$ is the
345: ``freezing'' temperature. For smaller bandwidths, the cosine--term
346: itself provides a mass for the rotational modes $\eta$. As a result,
347: all modes acquire a mass and, thus, the renormalized backscattering
348: amplitude looses its sensitivity to the lower limit $D'$. Therefore,
349: we arrive at the conclusion that for $D'< T_0$ the cosine amplitude
350: saturates at a value about $T_0$. Integrating in this way all
351: Matsubara components, except the static one, $m=0$ (it is obvious from
352: Eq.~(\ref{eq-theta2}) that the $m=0$ component can not be handled in
353: the same way), one obtains an effective classical model with the
354: action
355: \begin{equation}
356: \label{eq-classical}
357: S_{\rm cl}[\vec\theta] = \frac{E_c}T \sum_{\bf l}
358: \Big\{(\nabla\cdot\vec\theta_{\bf l})^2 -
359: \frac{\gamma(T)}{2\pi^2}\sum_{i=x,y}\cos(2\pi\theta_{i,{\bf
360: l}})\Big\},
361: \end{equation}
362: where $\gamma(T)=2\pi \sqrt{T/E_c}\,\,r$ for $T>T_0$ and
363: $\gamma(T)\simeq 2\pi r^2$ for $T<T_0$.
364:
365: So far we have formulated the model for an array with single--channel
366: contacts. Generalization to the $N\geq 2$ channels is achieved by
367: introducing bosonic modes $\vec\theta_{{\bf l},\alpha}(\tau)$ for
368: every channel $\alpha=1\dots N$. One may then integrate out the
369: antisymmetric modes (including the spin modes) for every contact,
370: retaining only the symmetric (charge) mode $\vec\theta_{{\bf
371: l}}=\sum_{\alpha=1}^N \vec\theta_{{\bf l},\alpha}$, see
372: Appendix~\ref{app-gamma} for details. The main result of such a
373: procedure may be summarized by the redefinition of the effective
374: backscattering amplitude~\cite{nazarov,ABG} $r\Rightarrow c_N
375: \prod_{\alpha=1}^N r_\alpha$ in the action of the charge mode ($c_N$
376: is a numerical coefficient with a finite limit $c_\infty$).
377: Consequently the characteristic freezing temperature changes to
378: $T_0\simeq E_c \prod_{\alpha=1}^N r^2_\alpha$. Then, the charge mode
379: may be described by the same effective classical model
380: Eq.~(\ref{eq-classical}) with $\gamma(T)\simeq \sqrt{T/E_c}\,
381: \prod_{\alpha=1}^N r_\alpha$ for $T>T_0$ and $\gamma(T)\simeq
382: \prod_{\alpha=1}^N r^2_\alpha$ for $T<T_0$.
383:
384: A model which adequately describes an array of metallic grains assumes
385: that the contacts between grains consist of a large number $N$ of
386: weakly transmitting
387: channels~\cite{general,Beloborodov01,tschersich,fs,fs2}. For
388: sufficiently large $N$, the total conductances of the junctions may
389: still be high, $g=\sum_{\alpha=1}^N t_\alpha^2 >1$, where
390: $t_\alpha^2=1-r_\alpha^2\ll 1$ is the transmission probability in
391: channel $\alpha$. In this case, one finds that $\prod_\alpha
392: r_\alpha=\exp\left\{\frac12\sum_\alpha\ln(1-t_\alpha^2)\right\}\approx
393: \exp\{-g/2\}$. Employing the expressions derived above, one obtains
394: that the freezing temperature is of the order $T_0\simeq g E_ce^{-g}$,
395: while the effective amplitude of the cosine potential in
396: Eq.~(\ref{eq-classical}) is given by
397: \begin{eqnarray}
398: \label{eq-gamma}
399: \gamma(T)\simeq
400: \begin{cases}
401: \sqrt g\,\, e^{-g/2}\sqrt{\frac T{E_c}}\,; \,\,\,\, & T>T_0,\\
402: g \,e^{-g}\,; & T<T_0\, .
403: \end{cases}
404: \end{eqnarray}
405: In Appendix~\ref{app-phase}, we show how Eqs.~(\ref{eq-classical}) and
406: (\ref{eq-gamma}) follow from the phase
407: model~\cite{general,Beloborodov01,tschersich,fs,fs2}, demonstrating
408: that the two models based on the charge and phase representations,
409: respectively, are reduced to the same effective classical system.
410: Notice that, since the charge model was derived for $r_\alpha\ll 1$,
411: while $t_\alpha\ll 1$ is assumed in the phase model, the coincidence
412: of $\gamma(T)$ may be expected at best with exponential accuracy. The
413: algebraic pre--exponential function of $g$ is a result of the
414: evaluation in the framework of the phase model, see
415: Appendix~\ref{app-phase}.
416:
417: \section{BKT transition}
418: \label{sec-BKT}
419:
420: In this Section we analyze the physics of the classical charge model
421: specified by Eqs.~(\ref{eq-classical}) and (\ref{eq-gamma}). Two
422: issues are discussed: (i) the spectrum of its charged excitations and
423: their interactions, and (ii) the low--frequency charge dynamics and
424: the dc conductivity of the array. We finally put our findings in
425: perspective by comparing them with the results of previous studies.
426:
427: \subsection{Charge spectrum}
428:
429: The lowest energy configuration of the action (\ref{eq-classical}) is
430: given by $\vec\theta=0\;(\mbox{mod}\;1)$ everywhere. Localized
431: excitations must have integer $\vec\theta$ far away from the core to
432: minimize the cosine potential. The total charge of such localized
433: excitation is $e\!\int(d^2l)\,\nabla\cdot\vec\theta=e\!\int d\vec
434: s\cdot\vec \theta$, where the line integral on the r.h.s. is
435: calculated over a distant contour enclosing the excitation. It is
436: clear therefore that the charge of the excitation is quantized in
437: integer numbers of $e$. The simplest (and only stable) charged
438: excitations have charge $\pm e$. They consist of a large (i.e.,
439: spread out over $\sim 1/\gamma\gg 1$ grains) localized 2d soliton of
440: unit charge, connected to a 1d string of links with $\theta_i=1$. The
441: other end of the string may either go to the system boundary or be
442: terminated by an anti-soliton with charge $-e$. The soliton solution
443: centered at ${\bf l}=0$ can be written in the form $\vec\theta_{\bf
444: l}=1-\vec\vartheta({\bf l})$ for the links along the string and
445: $\vec\theta_{\bf l}=\vec\vartheta({\bf l})$ everywhere else, where
446: $|\vec\vartheta(|{\bf l}|\to\infty)|\to0$. Minimizing the action,
447: Eq.~(\ref{eq-classical}), with respect to $\vec\vartheta$, one finds
448: the saddle point equation for the soliton solution,
449: \begin{equation}
450: \nabla(\nabla\cdot\vec\vartheta)-{\gamma\over 2\pi} \sum_{i=x,y}\sin
451: (2\pi\vartheta_i){\bf e}_i=0.
452: \label{saddlepoint}
453: \end{equation}
454: Except for a domain consisting of ${\cal O}(1)$ links closest to the
455: core of the soliton, $\vartheta$ is small, justifying an expansion of
456: the sine--term in the saddle point equation. As a result,
457: Eq.~(\ref{saddlepoint}) takes the form
458: $\nabla(\nabla\cdot\vec\vartheta)-\gamma\vec\vartheta=0$. Its
459: unit--charge solution is:
460: \begin{eqnarray}
461: \label{eq-soliton}
462: \vec\vartheta({\bf l}) \eq -\frac{\sqrt\gamma}{2\pi}\,
463: K_1\!\left(\frac l{\xi_{\rm
464: s}}\right)\,{\bf e}_l,
465: \end{eqnarray}
466: where $K_1$ is a modified Bessel function, and $\xi_{\rm
467: s}=1/\sqrt\gamma \gg1$, justifying the continuum approximation;
468: finally ${\bf e}_l\equiv {\bf l}/l$. The solution is normalized as
469: $\int(d^2l)\,\nabla\cdot\vec\theta=1$ to obey the charge quantization.
470:
471: Substituting this solution back into the action,
472: Eq.~(\ref{eq-classical}), one finds that the soliton energy originates
473: primarily from the cosine potential part of the action and is given by
474: $\Delta\simeq E_c[\gamma(T)/2\pi]\ln \xi_{\rm s}$. The large
475: logarithmic factor $\ln \xi_{\rm s}=-{1\over 2} \ln\gamma\gg 1$ is due
476: to the $\propto 1/l^2$ behavior of the charge density in the wide
477: range of distances $1<l<\xi_{\rm s}$. At larger distances, $l>\xi_{\rm
478: s}$, the charge density decays exponentially. As a result, the
479: solitons interact logarithmically up to a distance $\xi_{\rm s}$
480: beyond which the interaction is exponentially screened. Since the
481: density of thermally--excited solitons is $n_{\rm s}\approx
482: \exp\{-\Delta/T\}$, the mean distance between them is $l_{\rm
483: s}=n_{\rm s}^{-1/2}\approx \exp\{\Delta/(2T)\}$. It becomes
484: comparable to $\xi_{\rm s}$ at $T\approx \Delta/(2\ln \xi_{\rm s})=
485: E_c\gamma(T)/(4\pi)$. This condition is satisfied at temperatures
486: about the ``freezing'' temperature, $T\sim T_0$. Thus, at $T < T_0$,
487: the thermally-excited charges are essentially non--interacting, while,
488: at $T>T_0$, there is a neutral (in average) gas of logarithmically
489: interacting solitons and anti-solitons.
490:
491: In the latter regime the partition function of the charged degrees of
492: freedom may be written therefore as
493: \begin{equation}
494: \label{eq-partition}
495: Z=\sum\limits_{n=0}^{\infty}\frac{f^{n}}{n!} \int (d^2l_1)\ldots
496: (d^2l_n)\,e^{\pm\frac{E_c\gamma(T)}{2\pi T}\sum\limits_{k,k'}^n
497: \ln|{\bf l}_k-{\bf l}_{k'}|}\, ,
498: \end{equation}
499: where $\ln f\simeq E_c\gamma(T)/T$ is the fugacity of the logarithmic
500: gas, originating from the solitons core energy. The plus/minus signs
501: in the exponent correspond to soliton--soliton and
502: soliton--anti-soliton interactions, respectively.
503:
504: It is well known that the Coulomb gas in 2d described by
505: Eq.~(\ref{eq-partition}) undergoes the BKT
506: transition~\cite{BKT-b,BKT-kt} at a critical temperature $T_{\rm
507: BKT}\approx E_c\gamma/(4\pi)$. For $T<T_{\rm BKT}$, the charges are
508: bound in charge--anti-charge pairs. The residual density of free
509: charges is exponentially small and given by $n_{\rm s}\approx
510: \exp\{-\Delta/T\}$, where $\Delta=T_{\rm BKT}\ln \xi_{\rm s}^{\;2}$.
511: The value of $\Delta$ is finite but large, as long as the solitons
512: interact with each other logarithmically over a broad range of
513: distances $\xi_{\rm s}\gg 1$. Notice that the Coulomb interactions in
514: our model are strictly on--site (only the self--capacitance, $C$, is
515: included). The long range of the soliton--soliton interactions is due
516: to the fact that in a strongly coupled array, $g\gg 1$, the charge is
517: spread over a large distance $\xi_{\rm s}\sim\exp\{g/2\}$.
518:
519: One can modify the model to include mutual capacitances $C'$ between
520: neighboring grains (and thus to include long--range Coulomb
521: interactions). It is straightforward to show that such modification
522: alters the range of logarithmic interactions as $\xi_{\rm s} \to
523: \xi_{\rm s}\sqrt{1+C'/C}$, while the charging energy now reads
524: $E_c=e^2/(2(C+C'))$. In the limit $C\to 0$, while $C'$ remains
525: finite, the interaction range diverges, $\xi_{\rm s}\to \infty$. In
526: fact, this was to be anticipated: since without the self--capacitance
527: no electric field lines can leave the system, one deals with the true
528: 2d Coulomb interaction, which is logarithmic. In this case $\Delta\to
529: \infty$ and the density of free charges below $T_{\rm BKT}$ is
530: strictly zero. This is the case of the genuine BKT phase transition.
531: For non--vanishing self--capacitance, $C>0$, the interactions are
532: screened at distances exceeding $\xi_{\rm s}$. Therefore, the density
533: of free charges is finite at any temperature and the phase transition
534: is smeared into a sharp crossover. Above the transition/crossover
535: temperature the density of free charges rapidly increases
536: as~\cite{BKT-kt} $n_{\rm s}\sim\exp\{-2b\sqrt{T_{\rm BKT}/(T-T_{\rm
537: BKT})}\}$, where $b$ is a constant of order unity, driving the
538: array into the conducting phase.
539:
540: \subsection{dc conductivity}
541:
542: In order to discuss the dc conductivity of the array, one needs to
543: restore the low frequency, $\omega\ll T$, dynamics of the classical
544: charge model, Eq.~(\ref{eq-classical}). This may be done formally by
545: keeping the dissipative dynamical term in the action. Notice that in
546: the multichannel case, see Appendix~\ref{app-gamma}, the coefficient
547: in front of $|\omega_m|$ acquires a factor $N^{-1}$, where $N$ is the
548: number of channels. In the presence of strong backscattering, it
549: actually reads $\pi g^{-1} |\omega_m|\vec\theta_{{\bf l},m}^{\;2}$ and
550: corresponds to the conventional Ohmic dissipation. Since we focus on
551: the low frequencies, it is convenient to pass to the Keldysh
552: representation (to avoid dealing with the analytical continuation) and
553: consider its semi--classical limit. The latter is known to be
554: equivalent to a certain Langevin dynamics~\cite{Kamenev01}.
555:
556: Here we prefer to take a more phenomenological route, leading to the
557: same conclusions. Let us consider the static equations of motion
558: following from Eq.~(\ref{eq-classical}): $\partial_i\left({e\over
559: C}\nabla\cdot\vec\theta_{\bf l}\right)-{e\gamma\over 2\pi C}\sin
560: (2\pi\theta_{i,{\bf l}})=0$. Since $\frac eC\nabla\cdot\vec\theta_{\bf
561: l}\equiv V_{\bf l}$ is the voltage on grain ${\bf l}$, the equation
562: simply expresses the fact that in the absence of charge quantization,
563: $\gamma\to 0$, all grains are equipotential: $\nabla V_{\bf l}=V_{{\bf
564: l+e}_i}-V_{\bf l}=0$. Once currents are allowed to flow in the
565: array this condition should be substituted by the Kirchhoff law,
566: $V_{{\bf l+e}_i}-V_{\bf l}=RI_{i,{\bf l}}$, where $R=2\pi \hbar /(e^2
567: g)$ is the contact resistance, and $I_{i,{\bf l}}=e\partial_t
568: \theta_{i,{\bf l}}$ is the current flowing between grains ${\bf l}$
569: and ${\bf l+e}_i$. Restoring also the $\gamma$--term in the equation
570: of motion, one thus finds
571: \begin{eqnarray}
572: \label{eq-langevin}
573: &&\frac{\pi}{g}\, \partial_t\vec\theta -
574: E_c\Big[\nabla(\nabla\!\cdot\!\vec\theta) - \frac{\gamma}{2\pi}
575: \sum_i {\bf e}_i\sin(2\pi\theta_i)\Big] \nonumber\\
576: &=& -\frac{e}{2}\, {\bf E} +\vec\xi(t)\,.
577: \end{eqnarray}
578: On the right hand side we have included an external electric field
579: ${\bf E}$, as well as the Gaussian noise, $\vec\xi(t)$, with the
580: correlator
581: \begin{equation}
582: \label{eq-noise}
583: \langle \xi_{i,{\bf l}}(t)\xi_{i',{\bf l'}}(t')\rangle = \frac{2\pi
584: T}{g}\, \delta(t-t')\delta_{{\bf l},{\bf l}'}\delta_{i,i'}\, ,
585: \end{equation}
586: to satisfy the fluctuation--dissipation theorem.
587:
588: Our goal is to calculate the current, $I$, in presence of a weak
589: uniform field, $E$. To this end we employ Drude--type arguments,
590: saying that $I=en_{\rm s} v$, where $n_{\rm s}$ is the carrier
591: concentration and $v$ is their drift velocity. The only mobile
592: carriers in the system are the solitons, Eq.~(\ref{eq-soliton}), whose
593: concentration, $n_{\rm s}$, we have discussed in detail above. Now we
594: concentrate on the drift velocity, $v$. We look for a solution of
595: Eq.~(\ref{eq-langevin}) (without the noise) in the form
596: $\vec\theta({\bf l},t)=\vec\theta_0({\bf l}\!-\!{\bf
597: v}t)+\vec\theta_1({\bf l}\!-\!{\bf v}t)+\vec\alpha$. Here
598: $\vec\theta_0({\bf l})$ is the static soliton solution in the absence
599: of the external field, whereas $\vec\theta_1\sim {\bf E}$ is a small
600: modification of the soliton's shape due to the presence of the
601: external field. Finally the constant vector $\vec\alpha$ is
602: determined by the shift of the minimum of the periodic potential in
603: the field: $E_c\gamma\sum_i{\bf e}_i\sin(2\pi\alpha_i)=\pi{\bf E}$.
604: Choosing ${\bf E}=E{\bf e}_x$ and ${\bf v}=v{\bf e}_x$, and
605: linearizing Eq.~(\ref{eq-langevin}), one finds that $\vec\theta_1$
606: satisfies the equation
607: \begin{equation}
608: \label{eq-linearized}
609: E_c \hat {\cal F}_{\{\vec\theta_0\}}\vec\theta_1=\frac
610: v{g}\partial_x\vec\theta_0-\frac
611: E{\pi}\sin^2(\pi\theta_{0,x}){\bf e}_x,
612: \end{equation}
613: where $\hat {\cal F}_{\{\vec\theta_0\}}\vec\theta_1 \equiv
614: \nabla(\nabla\!\cdot\!\vec\theta_1)-\gamma \sum_i {\bf
615: e}_i\cos(2\pi\theta_{0,i})\theta_{1,i}$. The velocity, $v$, is
616: determined by the condition that the r.h.s. of
617: Eq.~(\ref{eq-linearized}) is orthogonal to the translational
618: zero--mode of the operator $\hat {\cal F}_{\{\vec\theta_0\}}$, given
619: by $\partial_x\vec\theta_0$. This requirement leads to $v\sim gE$.
620: Finally, the dc conductivity is given by $\sigma \simeq g\,n_{\rm
621: s}(T)$.
622:
623: As a result, all the conclusions, drawn above, regarding the BKT
624: transition/crossover in the soliton density may be directly translated
625: to the array's conductivity. In particular, for the self--capacitance
626: model at $T<T_{\rm BKT}$ (employing that at low temperature $\ln\gamma
627: \simeq g$) we find Eq.~(\ref{eq-res}), i.e. $\sigma\simeq
628: g\exp\{-\Delta/T\}$. Above $T_{\rm BKT}$, the conductivity behaves as
629: $\sigma \simeq g\exp\{-2b\sqrt{T_{\rm BKT}/(T-T_{\rm BKT})}\}$, see
630: Eq.~(\ref{eq-above}). At even higher temperatures, this behavior
631: crosses over to the result~\cite{tschersich} of the perturbative
632: calculation, $\sigma=g-\ln(gE_c/T)$.
633:
634: \subsection{Phase diagram}
635:
636: Using the results of the previous sections, we are now in a position
637: to discuss the phase diagram of the array. An array having
638: inter--grain capacitances $C'$ only exhibits a BKT phase transition
639: between the low--temperature insulating and the high--temperature
640: conducting phases. Its phase diagram on the plane temperature vs. bare
641: inter--grain conductance, $g$, is shown in Fig.~\ref{fig2}. Unlike
642: previous works~\cite{fs2,fs} that predicted a zero--temperature metal
643: for $g>g_c\simeq 1$, we find that the low--temperature phase is an
644: insulator for arbitrarily large $g$. The critical temperature, $T_{\rm
645: BKT}(g)$, however, drops sharply at $g\simeq 1$ and, at large $g\gg
646: 1$, behaves as $T_{\rm BKT}\sim E_cg \exp\{-g\}$. As shown in
647: Appendix~\ref{app-phase}, the disagreement is not a consequence of the
648: different model we use, but can be traced back to the disregard of the
649: quantum fluctuations of phase in the earlier works. By contrast,
650: Ref.~[\onlinecite{tschersich}] uses a perturbative renormalization
651: scheme that neglects instanton configurations. However, it is
652: precisely these instanton configurations that reflect the discreetness
653: of charge which is the key point in identifying the transition.
654:
655: \begin{figure}[h]
656: \centerline{\epsfxsize=3.5in\epsfbox{Fig2.eps}}
657: \caption{BKT temperature as a function of $g$. For $g\gg1$, $T_{\rm
658: BKT}$ is exponentially small, but remains finite. a) $C=0$: the
659: true transition exists for any value of $g$. Conductance is zero
660: below the transition temperature. b) $C>0$: crossover takes places.
661: It is sharp only if $g\gg1$ and/or $C'\gg C$. In the regime
662: $T<T_{\rm BKT}$, the system shows activation behavior with the gap
663: $\Delta=T_{\rm BKT}\ln(\xi_{\rm s}^{\;2})$.}
664: \label{fig2}
665: \end{figure}
666:
667: In the presence of the self--capacitance, $C>0$, the screening length
668: $\xi_{\rm s}\simeq \sqrt{1+C'/C}\, \exp\{g/2\}$ is finite, and the
669: transition is smeared into a crossover. The crossover is sharp as long
670: as $\xi_{\rm s}\gg 1$. Regardless of the ratio $C/C'$, this is the
671: case for $g\gg 1$. In this regime the charge gap is parametrically
672: larger than the crossover temperature, $\Delta\simeq gT_{\rm BKT}$,
673: and therefore the residual conductivity below the crossover, though
674: finite, is exponentially small, $\sigma\lesssim g\exp\{-g\}$. As a
675: result, quantitatively, there is little difference between the models
676: with and without self--capacitance. This is not the case for
677: $g\lesssim 1$: unless $C'\gg C$, the crossover is gone, and the
678: conductivity follows the simple activation law $\sigma\simeq
679: g\exp\{-\Delta/T\}$ with $\Delta \simeq E_c$.
680:
681: \section{Finite gate voltage}
682: \label{sec-gate}
683:
684: So far we have restricted ourselves to the case of zero gate voltage
685: only. A finite gate voltage induces a continuous background charge
686: $q\propto V_{\rm gate}$ on the grains. In this case the charging term
687: in the action Eq.~(\ref{eq-classical}) has to be replaced with $S_{\rm
688: cl}^{\rm (c)}[\vec\theta;q]=E_c/T \sum_{\bf
689: l}(\nabla\cdot\vec\theta_{\bf l}-q)^2$. Alternatively one may shift
690: the $\vec\theta$ field by $ q{\bf l}$ to move the $q$--dependence into
691: the pinning term $\frac{\gamma(T)}{2\pi^2}\cos(2\pi(\theta_i+ql_i))$.
692: Since grain coordinates $l_i$ take only integer values, the model is
693: periodic in the $q$--space with unit period. In this work we restrict
694: ourselves to a uniform gate voltage $q({\bf l})=q=const$, leaving
695: considerations of a random background charge for future studies.
696:
697: For small $q$, the system is in a particle--hole symmetric ``neutral''
698: state: the ground state is still (as for $q=0$) characterized by
699: $\vec\theta_{\bf l}=0$. At some finite value $q=q^*$, a transition
700: takes place, where the ground state becomes charged (with a
701: non--integer average number of electrons per dot) and spatially
702: non--uniform. To find $q^*$, let us compute the soliton energy in the
703: presence of $q$. Since the $q$--dependence of the Hamiltonian is a
704: pure boundary effect, one immediately finds the soliton energy
705: $\Delta(q)=\Delta(0)-2qE_c$. At $q^*=\Delta(0)/(2E_c)$ the soliton
706: energy $\Delta(q)$ vanishes. This marks the transition into the
707: charged state: for $q>q^*$, solitons are created at no cost.
708:
709: In the charged phase, the density of solitons in the array is finite
710: even at zero temperature. In order to find the soliton density at
711: $q>q^*$, one has to take into account interaction between the
712: solitons. At small densities, $n_{\rm s}<\xi_{\rm s}^{-2}=\gamma$, the
713: interaction between solitons is exponentially weak. The energy cost
714: associated with a soliton density $n_{\rm s}$ reads
715: \begin{eqnarray}
716: E_<(n_{\rm s})=n_{\rm
717: s}\Delta(q)+\frac2\pi E_c \gamma \,n_{\rm s}
718: K_0\left(\sqrt{\frac\gamma{n_{\rm s}}}\right),
719: \end{eqnarray}
720: where the second contribution is given by the interaction energy of a
721: pair of solitons separated by the distance $1/\sqrt n_{\rm s}$,
722: multiplied by the soliton density. Since $\sqrt{\gamma/n_{\rm
723: s}}\gg1$, we can use the asymptotic expression for the Bessel
724: function, $K_0(x)\sim x^{-1/2}\exp\{-x\}$ for $x\to\infty$. The
725: optimal density is determined by the minimum of $E_<(n_{\rm s})$.
726: Minimization of $E_<(n_{\rm s})$ with respect to $n_{\rm s}$ yields
727: $n_{\rm s}(q)\sim\gamma/\ln^2[\gamma/(q-q^*)]$, where we used that
728: $\Delta(q)=2E_c(q^*-q)$. Thus, at $q\gtrsim q^*$, the soliton density
729: rises rapidly until at $q-q^*\simeq\gamma$, the distance between
730: solitons reaches $\xi_{\rm s}$. For $n_{\rm s}>\xi_{\rm s}^{-2}$, the
731: solitons start to interact logarithmically. Consequently, the
732: expression for the energy has to be modified as
733: \begin{eqnarray}
734: E_>(n_{\rm s})\eq n_{\rm
735: s}\Delta(q)+\frac1{2\pi} E_c\gamma \,n_{\rm
736: s}^2\!\!\int\limits_{1/\sqrt{n_{\rm s}}}^\infty \!\!l\,dl\;
737: K_0\left(\sqrt\gamma \,l\right)\nonumber\\
738: &\simeq& n_{\rm
739: s}\Delta(q)+ \frac1{2\pi}E_c n_{\rm s} \left(\!n_{\rm s} - \gamma
740: \,\ln\!\sqrt{\frac{n_{\rm s}}\gamma}\right)\!,
741: \end{eqnarray}
742: where the second contribution describes the interaction energy of the
743: solitons with density $n_{\rm s}$; in the volume $\xi_{\rm
744: s}^2=1/\gamma$, the interaction is logarithmic [$K_0(x)\simeq -\ln
745: x$ for $x\ll1$]. In this regime, the minimization yields $n_{\rm
746: s}(q)\sim 2\pi(q-q^*)+ (\gamma/4)\ln[( q-q^*)/\gamma]$ for
747: $q-q^*\gg\gamma$.
748:
749: Naively, one would expect the system to be no longer insulating once
750: the density of solitons becomes finite at $q>q^*$ -- which would be
751: the case if the solitons were mobile. However, even though the
752: soliton density in the system is finite, $n_{\rm s}>0$, it turns out
753: that -- except for a narrow region $q-q^*<\gamma$, where the
754: interaction between solitons is exponentially weak -- the solitons
755: form a Wigner crystal which is pinned due to the underlying lattice
756: structure. Thus, transport is still activated.
757:
758: To understand this fact, we use the analogy with the formation of
759: vortices in a type II superconducting film~\cite{deGennes}. The field
760: $\vec\theta$ may be viewed as ${\bf A}\times{\bf n}_z$, where ${\bf
761: A}$ is the vector potential and ${\bf n}_z$ is a unit vector normal
762: to the film. Since the local magnetic field is given as ${\bf
763: h}=h\,{\bf n}_z=\nabla\times{\bf A}$, the correspondence goes as
764: $\nabla\!\cdot\!\theta=h$ and the charge quantization in the array is
765: equivalent of the flux quantization in the superconductor,
766: $\int(d^2l)\,h=k$ ($k\in\Bbb{Z}$), where $h$ is measured in units of
767: the flux quantum $\phi_0$. In this analogy, the gate voltage
768: translates to the external magnetic field, $H$, and the gate voltage
769: $q^*$ corresponds to the critical magnetic field $H_{c1}$, where it
770: becomes energetically favorable to create vortices. The
771: correspondences are summarized in the following ``dictionary'':\\[-0.4cm]
772: \begin{center}
773: \begin{tabular}{cc|cc}
774: array &&& superconducting film\\[0.1cm]
775: \hline
776: &&&\\[-0.3cm]
777: $\vec\theta$ &&& ${\bf A}\times{\bf n}_z
778: = -{\lambda^2}\nabla h$\\
779: charge $\nabla\cdot\vec\theta$ &&& local magnetic
780: field $h$\\
781: $\xi_{\rm s}=1/\sqrt{\gamma}$ &&& penetration depth $\lambda$ \\
782: background charge $q$ &&& external magnetic field $H$\\
783: $q^*$ &&& $H_{c1}$\\[-0.1cm]
784: \end{tabular}
785: \end{center}
786: Above $H_{c1}$ there is a finite density of vortices in the system
787: which at low enough temperatures form an Abrikosov
788: lattice~\cite{abrikosov}. In a clean film, the vortex lattice is free
789: to move, but it is easily pinned by the system boundaries, the
790: underlying lattice structure (as in Josephson junction arrays) or any
791: sort of disorder. Upon increasing the temperature the vortex lattice
792: eventually melts~\cite{BKT-kt,Tm}, and above the melting temperature
793: $T_{\rm m}$ most of the vortices are free to move. The melting
794: temperature at finite $H>H_{c1}$ is smaller than, but parametrically
795: the same as the Berezinskii--Kosterlitz--Thouless temperature at zero
796: magnetic field~\cite{Tm}. Thus, at $T<T_{\rm m}$ the system is
797: superconducting while at $T>T_{\rm m}$ the moving vortices lead to
798: dissipation.
799:
800: \begin{figure}[h]
801: \centerline{\epsfxsize=3.25in\epsfbox{Fig3.eps}}
802: \caption{System properties as a function of an external gate
803: voltage. The system becomes charged at $q>q^*$. If the soliton
804: density is larger than $1/\xi_{\rm s}^{\;2}$, charges arrange into
805: a (pinned) Wigner crystal. a) Phase diagram. b) Conductivity at
806: $T<T_{\rm m}$ as a function of $q$.}
807: \label{fig3}
808: \end{figure}
809:
810: Translating back to our problem this means that at $q>q^*$ the
811: solitons form a Wigner crystal once their density is sufficiently
812: large such that the interaction is logarithmic. Only in the narrow
813: interval $q^*< q< q^* + \xi_{\rm s}^{-2}$ the system is in the
814: conducting charge liquid state. Upon increasing the gate voltage, the
815: Wigner crystal forms and, due to lattice pinning, the system is an
816: insulator at temperatures smaller than the melting temperature. The
817: latter is of the order of $T_{\rm BKT}$. Note that while for $q<q^*$
818: charge is carried by individual (thermally-activated) solitons, for
819: $q>q^*$ the mobile charges are lattice defects, whose core energy is
820: proportional to the logarithm of the lattice constant of the Wigner
821: crystal.
822:
823: \section{Conclusions}
824: \label{sec-conclusions}
825:
826: The conductivity of a granular material with small ($g\ll 1$)
827: inter--grain conductances is controlled by the Coulomb blockade effect
828: in separate grains~\cite{fs}. Charge transport in such an array occurs
829: by electron hops between single grains. Because of the Coulomb
830: blockade, the granular array behaves as an insulator at low
831: temperatures. The characteristic energy determining the activation of
832: the charge transport, is associated with the single--grain charging
833: energy $E_c$.
834:
835: In this paper, we investigated the properties of a granular array at
836: high inter--grain conductance, $g\gg 1$. We concentrated on the
837: simplest model, neglecting the spacing between the discrete electron
838: levels in the grains ($\delta\to 0$), and in the main part of the
839: paper we also assumed the ideal limit of zero background charge
840: ($q=0$). We found that the granular array at sufficiently low
841: temperatures remains an insulator even in the case of $g\gg 1$. The
842: large inter--grain conductance, however, does affect the nature of the
843: charge carrier. Instead of an integer (in the units of $e$)
844: uncompensated charge sitting on a single grain, it is rather a
845: charge--$e$ soliton involving many grains. There is a sharp crossover
846: to low conductance at temperature the $T_{\rm BKT}\propto
847: E_c\exp\{-g\}$. Below the crossover, the electron transport is
848: associated with the activation of solitons with charges $\pm e$; the
849: corresponding activation energy is relatively high, $\Delta\simeq
850: gT_{\rm BKT}$. The approach to the crossover region from the
851: high--temperature side can be described in terms of the correlation
852: radius for the Berezinskii--Kosterlitz--Thouless transition, see
853: Eq.~(\ref{eq-above}). Comparison of Eqs.~(\ref{eq-res}) and
854: (\ref{eq-above}) shows that the crossover width is $\delta T\sim
855: T_{\rm BKT}/g^2\ll T_{\rm BKT}$.
856:
857: The effect of a gate voltage $q$ applied uniformly throughout the
858: array is to much extent similar to that of an external magnetic field
859: applied to a type II superconductor. Until $q$ reaches a certain
860: critical value $q^*$ the system's behavior does not change
861: qualitatively (that is, it exhibits the BKT crossover from insulator
862: to metal). The charge gap, $\Delta$, and the crossover temperature,
863: $T_{\rm BKT}$, decrease with increasing $q$. At $q=q^*$ the gap
864: vanishes (while the crossover temperature remains finite), and for
865: even larger gate voltages there is a certain ground--state density,
866: $n_{\rm s}=n_{\rm s}(q)$, of charge solitons. As long as this density
867: is small $n_{\rm s}(q)<\xi_{\rm s}^{-2}$ the solitons are in a liquid
868: state and the array conducts. At larger density, $n_{\rm
869: s}(q)>\xi_{\rm s}^{-2}$, the solitons form a Wigner crystal pinned
870: by the lattice. As a result, the array is again in the insulating
871: state with the charge gap determined by the cost of a defect in the
872: Wigner crystal.
873:
874: Finally, let us mention related issues that are {\em not} addressed in
875: the present paper. The first one is the role of disorder. The most
876: relevant is charge disorder equivalent to a grain--dependent gate
877: voltage $q_{\bf l}$. In the extreme scenario one may assume that
878: $q_{\bf l}$ are independent random variables, uniformly distributed in
879: $[0,1]$. One would then like to solve the classical statistical
880: problem formulated by Eqs.~(\ref{eq-langevin}) and (\ref{eq-noise})
881: with random $q_{\rm l}$ in the argument of $\sin(2\pi (\theta_{i,{\bf
882: l}} + q_{\bf l}))$. Despite many similarities with the vortex
883: physics, one can not simply transfer the known results from the pinned
884: vortex lattice literature~\cite{blatter}. The reason is that random
885: charge, $q_{\bf l}$, translates into a strong (of order of $H_{c2}$)
886: fluctuating magnetic field, rather a than fluctuating pinning
887: potential.
888:
889: Another unaddressed issue is the role of quantum coherence, which
890: enters the problem through the mean level spacing $\delta$. Our
891: results are valid as long as $g^2\delta\lesssim T_{\rm BKT}$. In the
892: opposite limit, the quantum coherence effects (most notably Anderson
893: localization) start to interfere with the effects of
894: electron--electron interactions, considered here. One may expect that
895: both effects drive the system towards the insulating ground--state.
896: (It is worth mentioning that in both cases the characteristic length
897: happens to be exponentially large in the bare conductance, $g$.) The
898: structure of such insulator is not known currently.
899:
900: \begin{acknowledgments}
901: We are grateful to A. Altland for valuable discussions. This work
902: was supported by NSF grants DMR01-20702, DMR02-37296, and
903: EIA02-10736. Furthermore, JSM was supported by a Feodor Lynen
904: fellowship of the Humboldt Foundation. AK and JSM acknowledge the
905: hospitality of KITP (UCSB), where part of this research was done
906: under NSF grant PHY99-07949.
907: \end{acknowledgments}
908:
909: \begin{appendix}
910:
911: \section{Multi--channel contacts}
912: \label{app-gamma}
913:
914: In the main text (Sec.~\ref{subsec-array}), we derived a classical
915: model for the single--channel case. Here we discuss its generalization
916: to $N\geq2$ channels. For every channel $\alpha=1,\ldots,N$ of a
917: multi--channel contact, one introduces a field $\theta_\alpha(\tau)$.
918: Consider an $M\times M$ array with $N$ channels in each of the $2M^2$
919: contacts. The quadratic part of the action reads
920: \begin{eqnarray}
921: S_2 = \frac1T\sum_{{\bf l},m}\left(\pi|\omega_m|\sum_\alpha
922: \vec\theta_{{\bf l},\alpha}^{\;2}+E_c\Big(\sum_\alpha
923: \nabla\cdot\vec\theta_{{\bf l},\alpha}\Big)^2\right)\!,
924: \end{eqnarray}
925: while the backscattering is described by
926: \begin{eqnarray}
927: S_r = -\frac{E_c}\pi\sum_{{\bf l},\alpha} \sum_{i=x,y}r_{i{\bf
928: l}\alpha}\! \int\limits_0^\beta\! d\tau\,
929: \cos(2\pi\theta_{i,{\bf l},\alpha}),
930: \end{eqnarray}
931: where the high--energy modes $E_c<|\omega_m|<D$ have already been
932: integrated out. [At energies larger than $E_c$, all modes are
933: decoupled and, thus, can be integrated out for each channel
934: separately.] Here, in order to clarify the following evaluation
935: scheme, the reflection coefficients $r$ have been given indices
936: specifying the direction, contact, and channel.
937:
938: Only the $2M^2$ symmetric modes $\theta_{\bf
939: l}=\sum_\alpha\theta_{{\bf l},\alpha}$ couple to external
940: parameters, such as gate voltages. We want, thus, to find an
941: effective action for $\theta_{\bf l}$ by integrating out $2M^2(N-1)$
942: asymmetric modes. To this end let us change variables from
943: $\theta_{{\bf l},\alpha}$ ($\alpha=1\dots N$) to $\theta_{\bf l}$ and
944: $\tilde \theta_{{\bf l},\alpha}=\theta_{{\bf l},\alpha}-(\theta_{\bf
945: l}-\sum_{\alpha'>\alpha}\theta_{{\bf l},\alpha'})/(\alpha+1)$
946: ($\alpha,\alpha'=1\dots N-1$). While the symmetric fields $\theta_{\bf
947: l}$ are massive due to the charging term, all the asymmetric fields
948: $\tilde \theta_{{\bf l},\alpha}$ are massless. As a result, the
949: perturbation theory in powers of $r_{i{\bf l}\alpha}$ contains only
950: the terms that do not have massless fields $\tilde \theta_{{\bf
951: l},\alpha}$ in the exponents (cosines). Rewriting the
952: backscattering action in terms of the new fields, one can see that the
953: lowest order non--vanishing terms are of the order
954: $\prod_{\alpha=1}^Nr_{i{\bf l}\alpha}$, where the product runs over
955: {\em all} channels of a given contact:
956: \begin{widetext}
957: \begin{eqnarray}
958: Z_N\sim E_c^N\prod_{\alpha=1}^Nr_{i{\bf l}\alpha}\int
959: d\tau_\alpha\,\cos\left(\frac{2\pi}N\sum_\alpha \theta_{i,{\bf
960: l}}(\tau_\alpha)\right) \prod_{\alpha=1}^N \left\langle
961: \exp\left\{2\pi
962: i\left(\tilde\theta_{i,{\bf l}\alpha}(\tau_\alpha) -
963: \frac1\alpha\left(\tilde\theta_{i,{\bf l},\alpha}(\tau_N) +
964: \sum_{\alpha'<\alpha}\tilde\theta_{i,{\bf
965: l},\alpha}(\tau_{\alpha'}) \right)\right)\right\}\right
966: \rangle_{\tilde\theta_{{\bf l},\alpha}}\!\!\!.\nonumber
967: \end{eqnarray}
968: \end{widetext}
969: Taking the averages $\langle\dots\rangle_{\tilde\theta_{{\bf
970: l},\alpha}}$ with the actions
971: \begin{eqnarray}
972: S[\theta_{i,{\bf
973: l},\alpha}] = \frac1T \sum_m \frac{\alpha+1}\alpha\pi
974: |\omega_m|\tilde\theta_{i,{\bf
975: l},\alpha}^2\, ,
976: \end{eqnarray}
977: yields $E_c^{1-N}\prod_{\alpha=1}^N \prod_{\alpha'>\alpha}
978: (\tau_\alpha-\tau_{\alpha'})^{-2/N}$ for the product of correlators
979: $\prod_\alpha\langle\dots\rangle_{\tilde\theta_{{\bf l},\alpha}}$.
980:
981: Thus, the effective action for $\vec\theta_{\bf l}$ reads
982: \begin{widetext}
983: \begin{eqnarray}
984: S[\vec\theta]\!=\!\sum_{\bf l}\left\{\!\frac1T\sum_m\left( \frac\pi N
985: |\omega_m| \vec\theta_{{\bf l}}^{\;2}+E_c\Big(
986: \nabla\!\cdot\!\vec\theta_{\bf l}\Big)^2\right) -
987: \frac{E_c}\pi\! \sum_{i=x,y}
988: \prod_{\alpha=1}^N r_{i{\bf l}\alpha}\!\! \int\! d\tau_\alpha
989: \prod_{\alpha'>\alpha}
990: \frac1{(\tau_\alpha\!-\!\tau_{\alpha'})^{2/N}}
991: \cos\Big(\frac{2\pi}N \sum_\alpha\theta_{i,{\bf
992: l}}(\tau_\alpha)\Big) \!\right\}.
993: \end{eqnarray}
994: \end{widetext}
995: Note that it is important to keep the non--local in time structure of
996: the cosine--term.
997:
998: At this stage, we can proceed to integrate out all the remaining modes
999: except the static one, $\theta_{m=0}$ -- as in the single--channel
1000: case. The prefactor of the cosine--term $V_0=({E_c}/\pi) \prod_\alpha
1001: r_{i{\bf l}\alpha}$ is renormalized according to
1002: \begin{eqnarray}
1003: V_0&\to&V(T)=V_0\,\exp\!\Big\{-\frac{2\pi^2}{N^2}
1004: \sum_{\alpha,\alpha'} \langle
1005: \theta(\tau_\alpha)\theta(\tau_{\alpha'})
1006: \rangle_{\theta_{m\neq0}}\Big\}\nonumber\\
1007: &=& V_0\,\exp\!\Big\{-\!\sum_{m\neq0} f(\omega_m)\big(1 \!+\!
1008: \frac2N\sum_{\alpha,\alpha'>\alpha}
1009: \cos\omega_m\tau_{\alpha\alpha'}\big)\Big\},\nonumber
1010: \end{eqnarray}
1011: where $f(\omega_m)=T/(4M^2)\sum_{\bf q}\{(E_{\bf
1012: q}+\pi|\omega_m|)^{-1}+(\pi|\omega_m|)^{-1}\}$; see
1013: section~\ref{sec-charge}. Since typical time differences
1014: $\tau_{\alpha\alpha'}=\tau_\alpha-\tau_{\alpha'}$ are of the order
1015: $1/T$ (the time integrals are dominated by the upper limit of
1016: integration), the last cosine--term inside the exponent may be
1017: disregarded. As a result we find $V(T)=V_0\sqrt{T/E_c}$.
1018:
1019: Finally, one may perform the multiple time integrations in the
1020: prefactor of the cosine. The integral over the center--of--mass time
1021: $\tau=\sum_\alpha\tau_\alpha/N$ contributes a factor $1/T$, while the
1022: integration over $N-1$ independent time differences
1023: $\tau_{\alpha}-\tau$ yields a constant $c_N$ multiplied by the
1024: logarithmic factor~\cite{ABG} $\ln E_c/T$. The latter follows simply
1025: from power counting. The same logarithmic factor appears in the
1026: framework of the phase model, Appendix \ref{app-phase}, as a result of
1027: zero--mode integration. Since all our evaluations of $\gamma$ are done
1028: up to a numerical factor, we shall not keep this logarithm explicitly.
1029: We, thus, reproduce Eq.~(\ref{eq-classical}) with $\gamma(T)\sim
1030: \sqrt{T/E_c}\, \prod_{\alpha=1}^N r_\alpha$. Continuation to $T<T_0$
1031: follows the same way as discussed in the main text for the
1032: single--channel case.
1033:
1034: \section{Phase model}
1035: \label{app-phase}
1036:
1037: In this Appendix, we establish correspondence between the charge
1038: representation, employed in the paper, and the more commonly used
1039: phase model. The latter may be straightforwardly derived starting from
1040: the fermionic tunneling Hamiltonian. Integration over the fermionic
1041: degrees of freedom (under the assumption of vanishing level spacing in
1042: every grain, $\delta\to 0$), leads to a model formulated in terms of
1043: the dynamic phase variable~\cite{AES}, $\phi_{\bf l}(\tau)$. Its time
1044: derivative, $\dot\phi_{\bf l}(\tau)$, has the meaning of a fluctuating
1045: instantaneous voltage on grain ${\bf l}$. The resulting action is a
1046: straightforward generalization of the Ambegaokar-Eckern-Sch\"on (AES)
1047: action~\cite{AES} to the array geometry. It consists of the charging
1048: term,
1049: \begin{equation}
1050: \label{AEScharging}
1051: S_{\rm c}[\phi]=\int\limits_0^{\beta}\!\! d\tau \sum_{\bf
1052: l}\left[ {\dot\phi_{\bf l}^{\;2}\over 4E_c} - i q
1053: \dot\phi_{\bf l}\right]\, ,
1054: \end{equation}
1055: and the dissipative term
1056: \begin{equation}
1057: \label{eq-AES}
1058: S_{\rm d}[\phi]=\frac {g
1059: T^2}4\int\!\!\!\int\limits_0^{\beta}\!\! d\tau d\tau'
1060: \sum_{\langle{\bf l},{\bf l'}\rangle}
1061: \frac{\sin^2(\phi_{\bf ll'}(\tau)-\phi_{\bf ll'}(\tau'))}{\sin^2(\pi
1062: T(\tau-\tau'))}\, ,
1063: \end{equation}
1064: describing tunneling between nearest neighbor grains $\langle{\bf
1065: l},{\bf l'}\rangle$. Here, $\phi_{\bf ll'}=(\phi_{\bf l}-\phi_{\bf
1066: l'})/2$.
1067:
1068: The phase field $\phi_{\bf l}(\tau)$ obeys the boundary condition
1069: $\phi_{\bf l}(\beta)-\phi_{\bf l}(0)=2\pi W_{\bf l}$, where $W_{\bf
1070: l}\in \Bbb{Z}$ is an integer called winding number. In addition to
1071: the trivial configuration $\phi_{\bf l}=0$, the stationary
1072: configurations of the dissipative action, $S_{\rm d}$, are given by
1073: Korshunov instantons~\cite{korshunov}
1074: \begin{equation}
1075: e^{i\phi_{\bf l}^{(W_{\bf l})}}=\prod_{a=1}^{|W_{\bf
1076: l}|}\frac{e^{2\pi i\tau
1077: T}-z_{a}}{1-\bar z_{a}e^{2\pi i\tau T}}
1078: \end{equation}
1079: characterized by the spatially--dependent winding number $W_{\bf l}$
1080: and a set of complex parameters $|z_{a}|<1$. In the regime $T\ll E_c$,
1081: these configurations are a good approximation to the saddle points of
1082: the total action. In the same approximation the $z_{a}$ are zero--mode
1083: coordinates: the instanton action is almost $z$-independent (safe for
1084: the charging terms that weakly depends on $z$). Neglecting this
1085: dependence, the action for a certain winding number configuration,
1086: $\{W_{\bf l}\}$, reads
1087: \begin{eqnarray}
1088: \label{eq-sw}
1089: S_W \!\simeq\! \frac{\pi^2T}{E_c}\sum_{\bf
1090: l} W_{\bf l}^2
1091: +\frac g4\sum_{\langle{\bf l},{\bf l'}\rangle}|W_{\bf
1092: ll'}|,
1093: \end{eqnarray}
1094: where $W_{\bf ll'}\equiv W_{\bf l}-W_{\bf l'}$. In the regime, we are
1095: interested in, $T\ll E_c$ and $g\gg 1$, the dominant contribution
1096: comes from the second (tunneling) term in this expression. Since the
1097: latter depends only on the differences of winding numbers on
1098: neighboring grains, it favors configurations with spatially extended
1099: regions with a fixed constant winding number, e.g. $W_{\bf l}=\pm 1$
1100: for a closed set of grains ${\bf l}$. We shall refer to such sets of
1101: grains with a fixed non--zero winding number as ``islands''. A typical
1102: phase--field configuration contains, therefore, a number of
1103: ``islands'' (with fixed non--zero windings) embedded in the sea of
1104: $W=0$ grains. An example of such a configuration is shown in
1105: Fig.~\ref{fig4}. According to Eq.~(\ref{eq-sw}), the action cost of
1106: one such island with the winding number $W$ is $S= A W^2(\pi^2 T/E_c)
1107: + L|W|(g/2)$, where $A$ is the area of the island (number of grains
1108: inside) and $L$ is its circumference (number of contacts with a
1109: winding number jump across them).
1110:
1111: \begin{figure}[h]
1112: \centerline{\epsfxsize=2.5in\epsfbox{Fig4.eps}}
1113: \caption{A typical island configuration is shown. The
1114: numbers correspond to winding numbers in the phase model.}
1115: \label{fig4}
1116: \end{figure}
1117:
1118: The same picture was employed in previous studies of 2d granular
1119: arrays~\cite{fs2,fs}, where the island structure was mapped onto the,
1120: so called, solid--on--solid model. The latter is known to exhibit a
1121: BKT transition. [A true transition takes place if there is no cost for
1122: the island's area, as is the case for the model with mutual
1123: capacitances only. Indeed, in such a model both charging and tunneling
1124: terms provide a cost proportional to the island's circumference, $L$.
1125: In presence of the self--capacitance (and thus an area--proportional
1126: cost) the transition is smeared into a crossover.] What was missed in
1127: the previous studies is an account of fluctuations on top of the
1128: stationary island--like configurations.
1129:
1130: We provide such an account here. Consider a stationary configuration
1131: consisting of a single island with a fixed winding number $W$.
1132: Expanding to the second order in deviations $\phi_{\bf l}=\phi_{\bf
1133: l}^{(W_{\bf l})}+\varphi_{\bf l}$, one finds for the fluctuating
1134: part of the action
1135: \begin{equation}
1136: \label{eq-fluctuations}
1137: \delta S = \sum\limits_m \sum\limits_{\bf l,l'} \bar\varphi_{{\bf
1138: l},m} M_{{\bf l,l'},m}^{(W)}\, \varphi_{{\bf l'},m}\, ,
1139: \end{equation}
1140: where $m$ is a Matsubara index, and $M_{{\bf l,l'},m}^{(W)}\equiv{\bf
1141: M}_m^{(W)}$ is the fluctuation matrix. The fluctuation factor
1142: associated with this configuration is given by
1143: \begin{equation}
1144: \label{eq-determinant}
1145: \prod\limits_{m>1}\frac{\det {\bf M}_m^{(0)}}{\det {\bf
1146: M}_m^{(W)}}=\exp\left\{-\sum\limits_m \mbox{Tr}\ln \left({\bf
1147: M}_m^{(0)}\right)^{-1}{\bf M}_m^{(W)} \right\},
1148: \end{equation}
1149: where ${\bf M}_m^{(0)}$ is the corresponding fluctuation matrix for
1150: the flat ($W=0$) stationary configuration. In the regime $g\gg 1$, the
1151: dominant fluctuation contribution comes from the expansion of the
1152: tunneling term, $S_{\rm d}$. This leads to $M_{{\bf l,l'},m}^{(0)}=
1153: -g|\omega_m|$ for nearest neighbors $\langle {\bf l,l'}\rangle $,
1154: while $M_{{\bf l,l},m}^{(0)}=-\sum_{{\bf l'}\neq {\bf l}} M_{{\bf
1155: l,l'},m}^{(0)}$ and $M_{{\bf l,l'},m}^{(0)}=0$ otherwise. In
1156: presence of the island, the off-diagonal elements of the fluctuation
1157: matrix are changed to $M_{{\bf l,l'},m}^{(W)}=-g|\omega_{m-|W|}|$ (and
1158: the diagonal elements accordingly), only if ${\bf l}$ and ${\bf l'}$
1159: are nearest neighbors laying across the island's boundary. As a
1160: result, one may write ${\bf M}_{m}^{(W)}={\bf M}_{m}^{(0)} - |W| {\bf
1161: \delta M}$, where the matrix ${\bf \delta M}$ has entries $\pm 2\pi
1162: Tg$ along the island's boundary and zeros everywhere else. Returning
1163: to the calculation of the fluctuation factor,
1164: Eq.~(\ref{eq-determinant}), one finds $\mbox{Tr}\ln \left({\bf
1165: M}_m^{(0)}\right)^{-1}{\bf M}_m^{(W)} = \mbox{Tr}\ln\!\left[1-
1166: |W|\left({\bf M}_m^{(0)}\right)^{-1}\!\!\!{\bf \delta M}\!\right]
1167: \!\approx\! -|W|\mbox{Tr}\left({\bf M}_m^{(0)}\right)^{-1}\!\!\!{\bf
1168: \delta M}$. Higher order terms in the expansion of the logarithm are
1169: rapidly convergent upon Matsubara summation and, therefore, may be
1170: safely neglected. Since ${\bf M}_m^{(0)} \sim g|\omega_m|$, summation
1171: over the Matsubara index in $\sum_m |W|\mbox{Tr}\left({\bf
1172: M}_m^{(0)}\right)^{-1}\!\!{\bf \delta M}$ leads to the logarithmic
1173: divergence \cite{Grabert96}. It is cut off by the charging part of the
1174: action at $m\approx gE_c/T\gg 1$. The summation (trace) over spatial
1175: indices results in a factor proportional to $L$, the island's
1176: circumference, as it counts the number of non--zero entries in ${\bf
1177: \delta M}$. Finally, a careful evaluation of the numerical
1178: coefficient~\cite{unpub} leads to the fluctuation factor,
1179: Eq.~(\ref{eq-determinant}), equal to $(gE_c/T)^{L|W|/2}$.
1180:
1181: As a result, an island of winding number $W$ with area $A$ and
1182: circumference $L$ contributes to the partition function of the model
1183: with the relative factor
1184: \begin{equation}
1185: \label{eq-factor}
1186: P_W(A,L)\!=\! \left(e^{-\pi^2T/E_c}\right)^{A W^2}
1187: \!\!\left(\sqrt{gE_c\over T}\, e^{-g/2} \right)^{L|W|}\!.
1188: \end{equation}
1189: (Actually, the statistical weight of an island contains also a factor
1190: $(\ln E_c/T)^{|W|}$, coming from the zero--mode, $z_a$, integrations
1191: \cite{Grabert96}. This factor has its exact analog in the charge
1192: model, mentioned at the end of Appendix \ref{app-gamma}. Hereafter we
1193: omit it for brevity.)
1194:
1195:
1196: We shall show now that the perturbative expansion in powers of
1197: $\gamma(T)$ of the classical model, Eq.~(\ref{eq-classical}), leads to
1198: the same island picture. In this case, every island carries the
1199: relative factor $\tilde P_W(A,L)\!=\! \left(e^{-\pi^2T/E_c}\right)^{A
1200: W^2} \!\left(E_c\gamma(T)/ (2\pi^2 T)\right)^{L |W|}$. We can, thus,
1201: identify the two models provided $\gamma(T) \simeq \sqrt{gT/E_c}\,
1202: e^{-g/2}$. Notice that $\gamma(T)\propto \sqrt{T/E_c}$ is exactly what
1203: one expects for the high--temperature, $T>T_0$, charge model. At lower
1204: temperature, non--linear fluctuation corrections in the phase model
1205: diverge~\cite{tschersich} and the above treatment runs out of
1206: validity. However, having establish the equivalence of the phase and
1207: charge models at $T>T_0$, one may proceed with the analysis of the
1208: latter even at smaller temperatures.
1209:
1210: To complete the proof, we elucidate now the island structure of the
1211: perturbative expansion of the charge model, Eq.~(\ref{eq-classical}).
1212: Consider the expansion of the partition function ${\cal Z}=\int
1213: D\vec\theta\;\exp\left\{-S[\vec\theta]\right\}$, with the action
1214: $S[\vec\theta]$ given by Eq.~(\ref{eq-classical}), in powers of the
1215: small parameter $\gamma$. The partition function can be written as
1216: ${\cal Z}=\sum_{n=0}^\infty Z_n (E_c\gamma/T)^{n}$, where $Z_n$ is a
1217: product of $n$ cosine terms averaged with the action $S_c[\vec \theta]
1218: =E_c\sum_{\bf l}(\nabla\cdot\vec \theta_{\bf l})^2/T$. There are two
1219: types of contributions: a) terms with higher powers of the cosine
1220: taken at the same link and b) terms involving different links. The
1221: first class of terms describes perturbative corrections to the
1222: conductance of a single contact, and may be shown to be equivalent to
1223: those of Ref.~[\onlinecite{tschersich}] in the framework of the phase
1224: model. The second class corresponds to the instanton terms and is the
1225: subject of our focus. These terms exhibit ``island formation''. To
1226: illustrate this, let us label the coefficients $\gamma_{i,{\bf l}}$
1227: (even though we assume them all to be equal), where $i=x,y$. Terms of
1228: the form $\prod_{\underline{\alpha}}\gamma_{\underline{\alpha}}$ are
1229: non--zero only if the lines crossing all contacts
1230: ${\underline{\alpha}}=(i,{\bf l})$ form closed loops, see
1231: Fig.~\ref{fig4}. I.e. the lowest-order non-local term is proportional
1232: to $\gamma^4=\gamma_{x,{\bf l}}\gamma_{x,{\bf l+e}_x}\gamma_{y,{\bf
1233: l}}\gamma_{y,{\bf l+e}_y}$ -- involving all the four links
1234: surrounding grain ${\bf l}$. This property of the model is due to the
1235: presence of massless modes, as is explained below.
1236:
1237: Rewriting $\vec\theta=\nabla{\chi}+\nabla\times\eta$, one finds that
1238: the charging action takes the form $S_c[\chi,\eta] =E_c\sum_{\bf
1239: l}(\nabla \chi_{\bf l})^2/T$ and is thus $\eta$--independent. The
1240: rotational field $\eta$ is therefore strictly massless. As a result,
1241: as long as an argument of the cosine (exponential) function contains
1242: the $\eta$--field, it averages to zero. Indeed, to obtain $Z_n$ one
1243: has to average expressions of the form $\exp\{2\pi
1244: i(\sum_{j_x=1}^s(\pm)\theta_{x,{\bf
1245: l}_{j_x}}+\sum_{j_y=s+1}^n(\pm)\theta_{y,{\bf l}_{j_y}}\}$, where
1246: $j_i$ labels contacts in $i$--direction. The terms containing $\eta$
1247: in the argument of this exponent, vanish. Therefore non--vanishing are
1248: only those terms that have $\sum_{j_x}\pm(\eta_{\,{\bf
1249: l}_{j_x}}-\eta_{\,{\bf l}_{j_x}\!\!-{\bf e}_y})
1250: +\sum_{j_y}\pm(\eta_{\,{\bf l}_{j_y}}-\eta_{\,{\bf l}_{j_y}\!\!-{\bf
1251: e}_x})=0$. It can be seen that this condition corresponds to the
1252: island structure. As a result, every island brings a factor
1253: $(E_c\gamma/T)^L$, where $L$ is its circumference, that simply
1254: reflects the order of the perturbation theory needed to create the
1255: island. For a proper (i.e. island--like) term of the perturbation
1256: theory, the averaging over the massive $\chi$--fields results in the
1257: factor $\exp\{-\pi^2TA/E_c\}$, where $A$ is the area. Finally, the
1258: integer index $|W|$ corresponds to the possibility of having a
1259: non--zero term of the perturbation theory, where links surrounding an
1260: island are included $|W|$ times each. We have shown, thus, that the
1261: perturbation theory in the charge model, Eq.~(\ref{eq-classical}),
1262: produces the same island structure as the instanton expansion of the
1263: phase model -- with the same relative factors, Eq.~(\ref{eq-factor}).
1264: This completes the proof of the equivalence of the two models and
1265: provides the value of $\gamma(T)$, Eq.~(\ref{eq-gamma}), for $g\gg 1$.
1266:
1267: \end{appendix}
1268:
1269: \begin{thebibliography}{}
1270:
1271: \bibitem{Gerber97} A. Gerber, A. Milner, G. Deutscher, M. Karpovsky,
1272: and A. Gladkikh, Phys. Rev. Lett. {\bf 78}, 4277 (1997).
1273:
1274: \bibitem{general} D.S. Golubev and A.D. Zaikin, Phys. Lett. {\bf
1275: A169}, 475 (1992); G. Goeppert and H. Grabert, Euro. Phys. J. B
1276: {\bf 16}, 687 (2000).
1277:
1278: \bibitem{Beloborodov01} I.S. Beloborodov, K.B. Efetov, A. Altland, and
1279: F.W.J. Hekking, Phys. Rev. B {\bf 63}, 115109 (2001).
1280:
1281: \bibitem{tschersich} K.B. Efetov and A. Tschersich, Europhys. Lett.
1282: {\bf 59}, 114 (2002).
1283:
1284: \bibitem{AGK} A. Altland, L.I. Glazman, and A. Kamenev, Phys. Rev.
1285: Lett. {\bf 92}, 026801 (2004).
1286:
1287: \bibitem{AAK82} B.L. Altshuler, A.G. Aronov, and D.E. Khmelnitskii, J.
1288: Phys. C {\bf 15}, 7367 (1982).
1289:
1290: \bibitem{fs2} J.E. Mooij, B.J. van Wees, L.J. Geerligs, M. Peters, R.
1291: Fazio, and G. Sch\"on, Phys. Rev. Lett. {\bf 65}, 645 (1990).
1292:
1293: \bibitem{fs} R. Fazio and G. Sch\"on, Phys. Rev. B {\bf 43}, 5307
1294: (1991).
1295:
1296: \bibitem{vinokur} I.S. Beloborodov, K.B. Efetov, A.V. Lopatin, and
1297: V.M. Vinokur, Phys. Rev. Lett. {\bf 91}, 246801 (2003).
1298:
1299: \bibitem{altshuler-aronov} B.L. Altshuler and A.G. Aronov, in {\em
1300: Electron-Electron Interaction In Disordered Systems}, eds. A.J.
1301: Efros and M. Pollak, pp. 1-153, Elsevier, North-Holland, 1985.
1302:
1303: \bibitem{BKT-kt} J.M. Kosterlitz and D.J. Thouless, J. Phys. C {\bf
1304: 6}, 1181 (1973).
1305:
1306: \bibitem{BKT-b} V.L. Berezinskii, Zh. Eksp. Teor. Fiz. {\bf 59}, 907
1307: (1970) [Sov. Phys. JETP {\bf 32}, 493 (1971)].
1308:
1309: \bibitem{mooij} J.E. Mooij, in {\em Percolation, Localization, and
1310: Superconductivity}, NATO ASI Series B, Vol. 109, eds. A.M. Goldman
1311: and S.A. Wolf, pp. 325-370, Plenum Press, 1983.
1312:
1313: \bibitem{Tm} B.A. Huberman and S. Doniach, Phys. Rev. Lett. {\bf 43},
1314: 950 (1979); D.S. Fisher, Phys. Rev. B {\bf 22}, 1190 (1980).
1315:
1316: \bibitem{flensberg} K. Flensberg, Phys. Rev. B {\bf 48}, 11156 (1993).
1317:
1318: \bibitem{matveev} K.A. Matveev, Phys. Rev. B {\bf 51}, 1743 (1995); A.
1319: Furusaki and K.A. Matveev, Phys. Rev. B {\bf 52}, 16676 (1995).
1320:
1321: \bibitem{nazarov} Yu.V. Nazarov, Phys. Rev. Lett. {\bf 82}, 1245
1322: (1999).
1323:
1324: \bibitem{ABG} I.L. Aleiner, P.W. Brouwer, and L.I. Glazman, Phys. Rep.
1325: {\bf 358}, 309 (2002).
1326:
1327: \bibitem{Kamenev01} A. Kamenev, in Nato Science Series II, Vol. 72,
1328: eds. I.V. Lerner {\em et al.}, pp. 313-340, Kluwer Academic
1329: Publishers, Dordrecht, 2002.
1330:
1331: \bibitem{deGennes} P.G. de Gennes, {\em Superconductivity of Metals
1332: and Alloys}, W.A. Benjamin, New York, 1966.
1333:
1334: \bibitem{abrikosov} A.A. Abrikosov, Zh. Eksp. Teor. Fiz. {\bf 32},
1335: 1442 (1957) [Sov. Phys. JETP {\bf 5}, 1174 (1957)].
1336:
1337: \bibitem{blatter} G. Blatter, M.V. Feigel'man, V.B. Geshkenbein, A.I.
1338: Larkin, and V.M. Vinokur, Rev. Mod. Phys. {\bf 66}, 1125 (1994).
1339:
1340: \bibitem{AES} V. Ambegaokar, U. Eckern, and G. Sch{\"o}n, Phys. Rev.
1341: Lett. {\bf 48}, 1745 (1982); G. Sch{\"o}n and A.D. Zaikin, Phys.
1342: Rep. {\bf 198}, 237 (1990).
1343:
1344: \bibitem{korshunov} S.E. Korshunov, Pis'ma Zh. Eksp. Teor. Fiz. {\bf
1345: 45}, 342 (1987) [JETP Lett. {\bf 45}, 434 (1987)].
1346:
1347: \bibitem{Grabert96} X. Wang and H. Grabert, Phys. Rev. B {\bf 53},
1348: 12621 (1996); M.V. Feigel'man, A. Kamenev, A.I. Larkin, and M.A.
1349: Skvortsov, Phys. Rev. B {\bf 66}, 054502 (2002).
1350:
1351: \bibitem{unpub} A. Altland, L.I. Glazman, A. Kamenev, and J.S. Meyer,
1352: unpublished.
1353:
1354: \end{thebibliography}
1355:
1356: \end{document}
1357: