1: %\documentclass[aps, prb, hyperref, floatfix, preprint]{revtex4}
2: %\documentclass[aps, prb, preprint]{revtex4}
3: \documentclass[aps, prb, twocolumn, floatfix]{revtex4}
4:
5: \usepackage{graphics}
6: \usepackage{epsfig}
7: \usepackage{dcolumn} % Align table columns on decimal point
8: \usepackage{subfigure}
9: \usepackage{float}
10: \usepackage{morefloats}
11: \usepackage{flafter}
12: \usepackage{afterpage}
13:
14: \newcommand{\rfc}[1]{{ \bf{(Request for comment: #1)}}}
15: \newcommand{\fixme}[1]{{ \bf{(FIX: #1 )}}}
16: \newcommand{\sinf}{$ S({\infty}) $}
17: \newcommand{\note}[1]{{\bf{#1} }}
18: %\newcommand{\fig}{../FIGS/}
19: \newcommand{\fig}{./}
20: \newcommand{\dd}[1]{{$\Delta \Delta${#1}}}
21:
22: \newcommand{\dns }{$\Delta n_s$ }
23: \newcommand{\jprime }{$\log(j)$ }
24: \newcommand{\phii }{{$\phi_{i}$ }}
25: \newcommand{\phiinosp}{{$\phi_{i}$}}
26: \newcommand{\Phy }{$\Phi$ }
27: \newcommand{\jmin }{$j_{min}$ }
28: \newcommand{\cd }{${\it j_i }$ }
29: \newcommand{\cdnosp}{${\it j_i}$ }
30:
31: \newcommand{\nsprime }{$n'_s$}
32: \newcommand{\localexp }{$\beta_{local}$ \hspace{1pt}}
33: \newcommand{\localexpnosp }{$\beta_{local}$}
34: \newcommand{\nep }{$n_{ep}$}
35: \newcommand{\goodgap}{
36: \hspace{\subfigtopskip}
37: \hspace{\subfigbottomskip}
38: \vspace{\subfigtopskip}
39: \vspace{\subfigbottomskip}
40: }
41:
42: \newcommand{\threshvolt }{$V_T$ \hspace{0.75pt}}
43: \newcommand{\threshvoltnosp}{$V_T$}
44: \newcommand{\vt }{$V_T$}
45: \newcommand{\csigma }{$C_{\Sigma }$ \hspace{0.75pt}}
46: \newcommand{\csigmanosp }{$C_{\Sigma }$}
47: \newcommand{\cinter }{$C_{I}$ \hspace{0.75pt}}
48: \newcommand{\cinternosp }{$C_{I}$}
49: \newcommand{\esigma }{$E_{c_{\Sigma }}$\hspace{0.05pt}}
50: \newcommand{\esigmanosp }{$E_{c_{\Sigma}}$}
51: \newcommand{\eovercsigma}{$\frac{e}{C_{\Sigma}}$}
52: \newcommand{\covercsigma}{$\frac{C}{C_{\Sigma}}$}
53: \newcommand{\crossvolt }{{$\Delta V_{X}$} \hspace{0.75pt}}
54: \newcommand{\crossvoltmax }{{$\Delta V_{X}^{max}$} \hspace{0.75pt}}
55: \newcommand{\mtr}{$\overline{\Gamma}$}
56: \newcommand{\ltth }{L$^{\frac{2}{3}} $~}
57: \newcommand{\ltthnosp}{L$^{\frac{2}{3}}$}
58: \newcommand{\lmtth }{L$^{-\frac{2}{3}} $~}
59: \newcommand{\lmoth }{L$^{-\frac{1}{3}} $~}
60: \newcommand{\lotw }{L$^{\frac{1}{2}} $~}
61: \newcommand{\lmotw }{L$^{-\frac{1}{2}} $~}
62: \newcommand{\lmotwnosp }{L$^{-\frac{1}{2}}$}
63: \newcommand{\lotwnosp }{L$^{\frac{1}{2}}$}
64: \newcommand{\loth }{L$^{\frac{1}{3}} $~}
65: \newcommand{\lothnosp }{L$^{\frac{1}{3}} $}
66: \newcommand{\lfth }{L$^{\frac{5}{3}} $~}
67: \newcommand{\leth }{L$^{\frac{8}{3}} $~}
68: \newcommand{\lfthnosp }{L$^{\frac{5}{3}}$}
69: \newcommand{\ftth }{$\frac{2}{3}$}
70: \newcommand{\ftthwsp }{$\frac{2}{3} $~}
71: \newcommand{\fttw }{$\frac{3}{2}$}
72: \newcommand{\fttwsp }{$\frac{3}{2} $~}
73: \newcommand{\ffth }{$\frac{5}{3}$}
74: \newcommand{\ffthwsp }{$\frac{5}{3} $~}
75: \newcommand{\feth }{$\frac{8}{3}$}
76: \newcommand{\ffthsp }{$\frac{5}{3} $~}
77: \newcommand{\fotw }{$\frac{1}{2}$}
78: \newcommand{\fotwsp }{$\frac{1}{2} $~}
79: \newcommand{\foth }{$\frac{1}{3}$}
80: \newcommand{\ecsm }{$\langle \frac{e}{C_\sigma} \rangle$}
81: \newcommand{\ecs }{$\frac{e}{C_{Sigma}}$}
82:
83: \newcommand{\CVSID}[1]{{ %
84: \cfoot[ \mbox{#1} ] { \mbox{#1} } %
85: }}
86:
87: \begin{document}
88: \title{Effects of disorder on electron transport
89: in \\ arrays of quantum dots}
90:
91: \author{Shantenu Jha} \author{A. Alan Middleton}
92: \affiliation{Department of Physics, Syracuse University, Syracuse,
93: NY, 13244, USA} \date{\today}
94:
95: \begin{abstract}
96: Using analytical and numerical methods, we investigate the
97: zero-temperature transport of electrons in a model of quantum dot
98: arrays with a disordered background potential, where the electrons
99: incoherently tunnel between the dots. One effect of the disorder is
100: that conduction through the array is possible only for voltages
101: across the array that exceed a critical voltage $V_T$. We
102: investigate the behavior of arrays in three voltage regimes: below
103: the critical voltage, above, but arbitrarily close to, the critical
104: voltage, and further above the critical voltage. For voltages less
105: than $V_T$, we find that the features of the invasion of charge onto
106: the array depend on whether the dots have uniform or varying
107: capacitances. We compute the first conduction path at voltages just
108: above $V_T$ using a transfer-matrix style algorithm. Though only
109: the first path can be studied using this technique, it can be used
110: to elucidate the important energy and length scales. We find that
111: the geometrical structure of the first conducting path is
112: essentially unaffected by the addition of capacitive or tunneling
113: resistance disorder. We also investigate the effects of this added
114: disorder to transport further above the threshold. We find that
115: qualitative behavior is dominated by the presence of the background
116: potential, rather than capacitive or tunneling disorder, at least as
117: long as these additional disorders do not have an extremely broad
118: distribution. We use finite size scaling analysis to explore the
119: nonlinear current-voltage relationship near $V_T$. The scaling of
120: the current $I$ near $V_T$, $I\sim(V-V_T)^{\beta}$, gives similar
121: values for the effective exponent $\beta$ for all varieties of
122: tunneling and capacitive disorder, when the current is computed for
123: voltages within a few percent of threshold. We do note that the
124: value of $\beta$ near the transition is not
125: % the current $I$ near $V_T$, $I\sim(V-V_T)^{\beta_{\rm eff}}$, gives
126: % similar values for the effective exponent $\beta_{\rm eff}$ for all
127: % varieties of tunneling and capacitive disorder, when the current is
128: % computed for voltages within a few percent of threshold. We do note
129: % that the value of $\beta_{\rm eff}$ near the transition is not
130: converged at this distance from threshold and difficulties in
131: obtaining its value in the $V\searrow V_T$ limit.
132: \end{abstract}
133:
134: \pacs{} \keywords{} \maketitle
135:
136: \section{Introduction}
137: It is now possible to engineer arrays of nanoparticles (1-2 nm in
138: diameter) in various geometrical configurations~\cite{wybourne02} and
139: to lithographically fabricate arrays of low capacitance islands
140: separated by tunnel junctions with reasonable control over array
141: parameters~\cite{kurdak00_1}. Such quantum dot arrays (QDA) have been
142: the subject of intense investigation recently~\cite{sohn,
143: wingreen+sohn}. In spite of the relatively well controlled
144: properties of these arrays however, there are limitations on the
145: homogeneity of such systems. Disorder at the sub-micron length scales
146: arises due to a variety of reasons, is inevitable and significantly
147: influences the properties of these otherwise well ordered arrays. For
148: ligand coated nanoparticles the variation in coating properties and
149: separation result in different resistances to electron tunneling. As
150: a consequence of the poly-dispersion in the sizes of metallic
151: nanoparticles, the charging energies of the individual islands differ.
152: Similarly for lithographically fabricated tunnel junctions, islands
153: with variable charging energies arise due to a dispersion of island
154: sizes or due to fluctuating capacitative coupling between dots and the
155: underlying gate. Given the pervasiveness of random background
156: charges, nonuniform charging energy and fluctuations in tunneling
157: resistance across the array, an important focus of the current work is
158: to study the effect of these on transport properties of electrons.
159: Due to limitations of the fabrication process, it is difficult to
160: control the different types of disorder independently, whereas this
161: can can be relatively easily addressed by computer simulations.
162:
163: There are many dynamical systems in which strongly interacting
164: particles exhibit collective transport in a random
165: environment~\cite{dfisher98}. Even though the underlying microscopic
166: details are different, systems like the vortex glass in type-II
167: superconductors and charge density waves~\cite{dfisher87}, share some
168: general features in the long wavelength limit, e.g., they are both
169: characterized by the presence of a well defined threshold force below
170: which the system is essentially static and above which the system has
171: a non-linear response. Electron transport in disordered QDA also
172: provide a useful system to study problems of qualitative similarity.
173: An advantage of QDA is that the primary interactions and the
174: fundamental physics are relatively better understood and arguably
175: under greater experimental control.
176:
177: Using a combination of analytic and numerical techniques, we
178: investigate electron transport at zero temperature in arrays of
179: disordered small capacitance islands which are capacitatively
180: uncoupled to their neighbors. We use this system both as a model for
181: collective transport of discrete charges in a random environment and
182: for better understanding the role of disorder. By studying similar
183: systems, but for different values of parameters, different regimes of
184: the collective transport problem can be addressed. These regimes may
185: be characterized by the relative strengths of disorder, tunneling
186: rates and electron-electron interaction. These regimes are accessible
187: experimentally too, as arrays can be fabricated with varying degree of
188: tunability of the coupling between the array elements~\cite{marcus95}.
189: For example, the model in Refs.~[\onlinecite{Enomoto99},
190: \onlinecite{Kosterlitz98}] is similar to the model we study in this
191: paper -- in that offset charge disorder is included, although Gaussian
192: distributed as opposed to uniformly distributed -- but the screening
193: length is assumed infinite. (Transport in this regime is believed to
194: belong to the same universality class of two dimensional magnetic
195: vortex model in disordered superconducting films.)
196: %Also, for 2DEG at semiconductor heterointerfaces the self capacitance
197: %of the dots tends to be large due to the metallic gates and hence cuts
198: %off the long range Coulomb interactions.
199: Ref.~[\onlinecite{hirasawa98}] which investigates 2DEG at
200: semiconductor heterointerfaces designed to keep the self-capacitance
201: and disorder low -- the I-V characteristic is better explained by
202: charge soliton injection as no threshold voltage is observed. Such
203: systems can be used to study two dimensional Coulomb gases which
204: undergo charge Kosterlitz-Thouless (KT) transitions.
205:
206: \subsection{Outline \label{sec:outline}}
207:
208: There have been many experimental and simulation papers
209: ~\cite{jaeger01, jaegerprl04, berk98, berk95, ancona02} (to cite just
210: a few) which have used the theory developed in
211: Ref.~[\onlinecite{mw93}] by Middleton and Wingreen (MW). This paper
212: expands on the MW discussion of electron transport in disordered
213: arrays. The original model and its extension to include other forms
214: of disorder are described in the remainder of this section.
215: One-dimensional arrays -- both below and above threshold -- are
216: discussed in detail in section~\ref{sec:onedim}. In
217: section~\ref{sec:subthresh} results for 2D arrays below the threshold
218: voltage are presented including several results not discussed
219: previously. As the original MW paper sketched only briefly the
220: connection between the independent conducting paths and the properties
221: of a directed polymer in random media (DPRM), a major aim of
222: section~\ref{sec:paths} and~\ref{sec:subthresh} is to establish the
223: connection on a more rigorous basis. In section~\ref{sec:paths} we
224: discuss the morphology and current carrying properties of the first
225: conducting path at threshold for QDA. It also provides some of the
226: details required to understand the non-linear scaling of current (I)
227: with voltage (V), which is addressed in section~\ref{sec:2d_dynamics}.
228:
229: There have been several papers that have used numerical approaches to
230: investigate transport in arrays (both 1D and 2D) in the presence of
231: random background charges as well as other types of
232: disorder~\cite{likharev_prb03, melsen97, johansson, ancona02}. Some
233: approaches, have used discrete event simulation techniques to model
234: the individual tunneling events, whereas some have explicitly used
235: computed transition rates in a master-equation approach. The common
236: aim is to compute the general I-V characteristics, which as a
237: consequence of the collective behavior of electron tunneling is
238: non-trivially dependent on the individual rates. We focus on a
239: statistical physics approach to the problem, thus laying a theoretical
240: basis for the scaling exponents observed experimentally and
241: numerically.
242:
243: \subsection{The Model \label{sec:model}}
244: The three main energy scales of QD are the charging energy
245: (\esigmanosp), the electron in-a-box energy levels ($\Delta$) and the
246: thermal energy ($kT$). As a consequence of the small size of these
247: islands and tunnel junctions the capacitance involved are in the femto
248: to atto-Farad range, thus the charging energy -- which is the increase
249: in energy due to the addition of a single electron is given by $e^2/
250: C_{\Sigma}$ -- of these islands is large. A characteristic feature of
251: QD is the clear separation of internal energy scales $\Delta$ and
252: \esigma. An external energy scale ($kT$) is set by the temperature of
253: interest, which determines the levels that are resolved and
254: participate in transport. When \esigma $\gg$ $kT$ the role of thermal
255: fluctuations can be ignored. Depending upon the temperatures of
256: interest, $\Delta$ maybe comparable to $kT$ or different; for $kT \gg
257: \Delta$, the discrete energy level spectrum of the QD do not play a
258: role during transport. Metallic dots are different from
259: semi-conducting dots by the fact that typically the level spacings for
260: metallic dots are much smaller compared to other energies. At
261: sufficiently low temperatures, the scale of which is set by, \esigma
262: $\gg$ $kT$, the addition of a single extra electron to an dot
263: increases the dot energy; in spite of the increased energy, the dot is
264: stable to thermal energy fluctuations, which in turn makes it
265: unfavorable for more electrons to tunnel onto the same dot, resulting
266: in its blocking other electrons onto the dot. This is called the {\it
267: Coulomb blockade} regime.
268:
269: The parameters required to characterize QDA can vary over a large
270: range of values and consequently so do the properties of QDA. Thus,
271: it is instructive to understand the parameter space of QDA in order to
272: appreciate the details of the model. The main parameters used to
273: characterize QDA, as opposed to individual quantum dots (QD) are: the
274: tunneling resistance ($R_T$) which to a first approximation is a
275: measure of how well confined the electrons are on a dot, the inter-dot
276: capacitance (\cinternosp) and the dot capacitance (\csigmanosp) which
277: is a function of the junction, gate and self-capacitances. The
278: relative values of \csigma and \cinter are important as it determines
279: the extent of electrostatic coupling between dots in the array. The
280: exact value of \csigma depends on the system under consideration. For
281: example, typically the self-capacitance of lithographically prepared
282: arrays is negligible compared to the other capacitances, thus \csigma
283: is a function of the dot-gate and tunnel junction capacitances (for
284: example in Ref.~[\onlinecite{berk98}] \csigma = C$_g$ + 4C). For
285: nanoparticles with diameters of a few {\it nm}, the self-capacitance
286: becomes important and should possibly be considered in the computation
287: of \csigmanosp \cite{wybourne01}. Either way, \csigma still sets the
288: scale for the charging energy. Independent of the actual experimental
289: setup considered, as long as \csigma $\gg$ \cinternosp, the dots are
290: considered to be capacitatively uncoupled to each other and the
291: electrostatic energy is determined by on site interactions only.
292: However if \csigma is comparable or less than \cinter the dots are
293: capacitatively coupled. A screening length ($\lambda$) can be thought
294: of the distance (in units of dots) upto which the charge on a dot can
295: be felt electrostatically, i.e., distance that an excess charge placed
296: on a dot will effect neighboring dots by polarization. The
297: polarization decreases exponentially with $\lambda$, which in turn
298: decreases with the ratio of $\frac{C_I}{C_{\Sigma}}$; this is
299: consistent with understanding that there is a stronger screening of
300: the electrons on a dot from electrons on adjacent dots as the
301: capacitative coupling between a dot and the back gate increases. For
302: \csigma $\gg$ \cinter, $\lambda \approx
303: {(\frac{C_I}{C_{\Sigma}})}^{1/2}$.
304:
305: The main modification to the original MW model is the introduction of
306: nonuniform dot capacitance \csigma and tunneling resistances $R_T$.
307: The effects of underlying charge impurities trapped at the interfaces
308: and in the substrate are captured in a random background charge on
309: each dot. The effect of background charges is modeled as offset
310: charges on each dot ($q_i$). The offset charge at any site is
311: considered to be [0,1[, as any value outside this range will be
312: compensated by electron hopping. Arrays with only offset charge
313: disorder are referred to as UC (uniform capacitance) systems. The
314: area and capacitative coupling of an island to the underlying electron
315: gas varies from dot-to-dot in an array. These fluctuations in the
316: dot-gate and self-capacitance of dots, along with stray capacitances
317: are incorporated by assuming a varying dot capacitance \csigmanosp.
318: As \csigma controls the charging energy of the dot, a non-uniform
319: \csigma results in different charging energies \esigma of dots.
320: Arrays with both offset charge disorder and a varying \csigmanosp, are
321: referred to as DC (disordered capacitance) systems. Fluctuations in
322: the tunneling resistance -- either due to varying distance between
323: metallic dots or the varying material properties of the tunnel
324: junction separating the metallic islands between dots -- is captured
325: by assuming a log-normal distribution of tunneling resistances. Arrays
326: that incorporate variation in tunneling resistance as well as offset
327: charge disorder, but with a fixed value of \csigmanosp, are referred
328: to as RT (resistance disorder) systems.
329:
330: We assume small metallic islands are separated from each other by
331: tunnel junctions of resistance $R_T$ but capacitatively coupled to
332: neighboring dots (\cinternosp). We assume a constant capacitance
333: \cinter between neighboring dots and between the left and right leads
334: and dots adjacent to them. The dots are assumed to be separated from
335: an underlying back gate by an insulating layer. Each dot is
336: capacitatively coupled to the back gate with a capacitance
337: \csigmanosp. The leads and back gate are assumed to have infinite self
338: capacitance. As a consequence of the proximity of the back gate to the
339: dots \csigma $\gg$ \cinternosp, the screening-length is taken to be
340: less than one lattice spacing. Consequently the capacitative coupling
341: between dots is neglected.
342:
343: We will consider arrays where the single-energy levels of the dots are
344: essentially a continuum at the Fermi level in the strongly Coulomb
345: blockaded (\esigma $\gg kT$) regime. Thus tunneling is between levels
346: determined by \esigma. Where a spread in values of $R_T$ is
347: considered, we assume tunneling resistance between any two dots is
348: still sufficiently large to consider electrons localized on a site
349: ($R_T \gg h/{e^2}$). This is the regime of the ``orthodox theory'' of
350: single electron tunneling and is applicable for both the micron sized
351: lithographically defined SET (e.g., metal islands embedded in a
352: substrate \cite{berk95} and separated by tunnel junctions or
353: semi-conductor islands separated by barriers \cite{marcus95}) as well
354: as the 3D metallic grains. According to the ``orthodox
355: theory''~\cite{orthodox} of a tunneling event across a tunnel
356: junction, tunneling rates (transition probability per unit time)
357: associated with an event are given by,
358: \begin{equation}
359: \label{eq:rate}
360: \Gamma = {\Delta E \over e^2R_T}{ 1 \over [1-\exp(-\frac{\Delta E}{k T})]}
361: \end{equation}
362:
363: where $\Delta E $ is the difference in the free energy of the system
364: before and after the tunneling event, $R_T$ is the tunneling
365: resistance of the junction involved in the tunneling event, T the
366: temperature and $k$ is the usual Boltzmann constant. The kinetic
367: energy gained by the tunneling electron is assumed to be dissipated.
368: The value of $R_T$ is assumed to be much greater than $h/e^2$. This
369: essentially implies that the wavefunction of electrons are localized
370: to a single dot which permits the number of electrons on any single
371: dot to be treated as a classical variable. It should be pointed out
372: that the orthodox theory is still valid for arrays in the limit $C_i$
373: $\gg$ \csigmanosp, but not for dots in the other limits of $R_T \ll
374: R_Q$ and \esigma $\ll kT$.
375:
376: In this limit the energy is all electrostatic and is determined by a
377: capacitance matrix $C_{ij}$ and is represented as:
378: \begin{equation}
379: \label{eq:energyexpression}
380: E = V_L Q_L + V_R Q_R + \frac{1}{2} \sum_{ij}{(Q_{i} +
381: q_{i})}{C^{-ij}}{(Q_{j}+q_{j})},
382: \end{equation}
383: where $Q_L$ ($Q_R$) are the charges on the left (right) leads, which
384: are at voltages $V_L$ ($V_R$) and $C^{-ij}$ is the inverse of the
385: matrix of capacitances between dots $i$ and $j$. The diagonal
386: elements of $C_{ij}$ are the sum of all capacitances associated with a
387: dot and the off-diagonal elements are the negative of the inter dot
388: capacitances. Hence for a N$\times$N array in the limit of
389: $\frac{C_I}{C_{\Sigma}}$ $\rightarrow$ 0, the capacitance matrix is a
390: N$\times$N diagonal matrix.
391:
392: In the limit of small screening length (less than 1 dot spacing) and
393: the presence of offset charge disorder the voltage on dot i is given
394: by $V_i$ is $(Q_i + q_i)/C_{\Sigma} $.
395:
396: At zero temperatures the expression~(\ref{eq:rate}) for tunneling
397: rates reduces to
398: \begin{equation}
399: \label{eq:zeroT_rate}
400: \Gamma = {\Delta E \over e^2R_T}{\Theta(\Delta E)}
401: \end{equation}
402: hence a charge may tunnel from dot i to j, only if such an event
403: lowers the free energy of the array i.e.
404: \begin{equation}
405: \label{eq:condition}
406: V_i > V_j + e/C_{\Sigma}
407: \end{equation}
408:
409: %\input{oneD}
410: %\input{subthresh}
411: %\input{paths}
412: %\input{dynamics}
413: %\input{otherresults}
414: %\input{conclusions}
415: %\appendix*
416:
417: %\renewcommand{\fig}{../FIGS/1D/}
418:
419: \section{1D arrays \label{sec:onedim}}
420:
421: Before attempting to understand the detailed properties of two
422: dimensional arrays, we begin by an attempt to understand the
423: relatively simpler case of a linear chain of quantum dots, as they
424: facilitate an understanding of some of the ideas required later.
425: There have been several experiments aimed at understanding the
426: conduction properties of essentially one dimensional arrays of
427: nanoparticles \cite{tinkham99, wybourne01}. The ability of metallic
428: nanoparticles to be patterned using polymers templates makes them
429: attractive candidates for potential future self-assembling electronic
430: devices.
431:
432: \subsection{Uniform \csigmanosp: Insulating State
433: \label{subsec:onedim_uniform}}
434:
435: We start by exploring the tunneling of electrons onto the array from
436: the emitter lead. In the zero temperature limit, as the capacitance of
437: the leads is assumed to be infinite, electrons can flow onto the array
438: when the voltage of the emitter lead ($V_L$) is equal to or greater
439: than the voltage of the leftmost dot as given by
440: Eqn.~\ref{eq:condition}. At this applied voltage, an electron cannot
441: tunnel from the leftmost dot to the next dot, say represented by the
442: index $i$, if dot $i$ has an offset charge impurity $q_i$ greater than
443: the offset charge impurity $q$ of the leftmost dot. Electrons tunnel
444: onto the array only if it is possible to do so without an increase in
445: the free energy of the system. This is no longer possible for the
446: configuration in Fig.~\ref{firstincr}. In this configuration the
447: electron residing on the leftmost dot is considered to be {\it
448: pinned}.
449:
450: %\renewcommand{\fig}{../FIGS/GEN/}
451:
452: \begin{figure}
453: \subfigure[]{\epsfig{figure= \fig spunif_schm.0.eps, scale=0.5}
454: \label{firstincr}}
455: \subfigure[]{\epsfig{figure= \fig spunif_schm.1.eps, scale=0.5}
456: \label{secondincr}}
457: \subfigure[]{\epsfig{figure= \fig spunif_schm.2.eps, scale=0.5}
458: \label{thirdincr}}
459: \caption{A schematic illustrating the build of charges in a 1D array
460: as the emitter voltage is progressively increased to threshold.}
461: \end{figure}
462:
463: %\renewcommand{\fig}{../FIGS/1D/}
464:
465: As shown in Figs.~\ref{secondincr} and~\ref{thirdincr} the emitter
466: lead voltage has to be increased by at least one unit in order that
467: the electrons can overcome the barrier. As the value of the emitter
468: lead voltage is successively increased, there will be a cascade of
469: electrons tunneling onto the array from the lead, until they get
470: progressively pinned and it is no longer energetically possible for
471: electrons to penetrate further into the array. The flow of charges
472: onto the array at a given emitter lead voltage until they are all
473: pinned due to the disorder and thus no further electrons can tunnel
474: onto the array constitutes an {\it avalanche}.
475:
476: %\renewcommand{\fig}{../FIGS/1D/}
477:
478: \begin{figure}
479: \subfigure[]{\epsfig{figure= \fig 1D.thresh+fluct.1.0.eps, scale=0.6}
480: \label{1D.vthresh.uniform}}
481: \subfigure[]{\epsfig{figure= \fig 1D.thresh+fluct.2.0.eps, scale=0.6}
482: \label{1D.vthresh.nonuniform}}
483: \caption{In Fig.\ref{1D.vthresh.uniform} a plot of the scaling of $V_T$ and
484: root mean square fluctuations of $V_T$ with system size for a 1D
485: array when the \csigma is uniform. In
486: Fig.\ref{1D.vthresh.nonuniform} scaling of $V_T$ and root mean
487: square fluctuations in $V_T$ with system size for a 1D array but
488: with non-uniform distribution of \csigmanosp.}
489: \end{figure}
490:
491: There exists a unique value of the emitter lead voltage -- which
492: depends upon the underlying disorder profile -- at which electrons
493: will be able to reach the collector lead for the first time
494: (Fig.~\ref{thirdincr}). This well-defined voltage value is referred
495: to as the threshold voltage ($V_T$). \threshvolt separates the
496: conducting phase from an insulating phase. Typically in order to
497: reach the collector lead end of an array L dots long, an electron will
498: have to overcome $\frac{L}{2}$ upward steps. These steps can be can
499: be understood as the average number of steps a random walk in 1D makes
500: in a given direction, thus the mean threshold voltage should be given
501: by
502: \begin{equation}
503: \label{eq:threshold}
504: {\overline{V_T}} = (\frac{L}{2})(\frac{e}{C_{\Sigma}})
505: \end{equation}
506: where the over-bar represents an averaging over disorder realizations.
507: Sample-to-sample fluctuations in the $V_T$ can be thought of as the
508: root-mean-square fluctuations of a random walk in 1D which scales as
509: $N^{1/2}$, where N is the number of steps of the random walk. Hence
510: fluctuations in $V_T$ should scale with system size as
511: \begin{equation}
512: \label{eq:threshold_fluct}
513: {\sigma(V_T)} \sim L^{1/2}
514: \end{equation}
515: The scaling of both $V_T$ and $\sigma(V_T)$ with system size, as shown
516: in Fig.~\ref{1D.vthresh.uniform} are consistent with the above
517: explanation.
518:
519: \subsection{Uniform \csigmanosp: Conducting State}
520: The threshold voltage represents the lowest voltage at which electrons
521: can tunnel across the array, hence for $V_L > V_T$, current flows
522: through the array. For a given disorder realization, \threshvolt
523: depends on the number of up-steps encountered due to the offset charge
524: impurities.
525:
526: If $V_L$ is marginally greater than $V_T$, so that $\nu \equiv (V_L -
527: V_T)/V_T \ll 1$, then the discreteness of charge and offset charge
528: impurities play a crucial role in determining the current. At these
529: voltages the current is determined by the slowest tunneling rate
530: ($\Gamma_{slow}$) between any two neighboring dots in the array
531: (analogous to the net flow of traffic being determined by the
532: bottleneck in the path of flow), which on the average is given by
533: $\frac{V_L - V_T}{eRL}$, where L is the number of dots in the 1D
534: array.
535:
536: This can be understood as follows: $(V_L - V_T)$ represents the
537: voltage increment over $V_T$. In principle the voltage drop can be
538: anywhere between between 0 and $V_L - V_T$ for a given pair of dots.
539: For an array with L dots there are L+1 ($\sim$ L when L is large)
540: voltage drops (tunneling rates), hence the minimum voltage drop across
541: any two dots is on the {\it average} $\frac{V_L - V_T}{L}$, which
542: results in a tunneling rate given by Eqn.~(\ref{eq:zeroT_rate}) to be
543: $\frac{V_L - V_T}{eRL}$. Using $V_T = \frac{eL}{2C_{\Sigma}}$ we get
544: $\Gamma_{slow}$ = $\frac{V_L - V_T}{2RC_{\Sigma} V_T}$. As $I =
545: e\Gamma_{slow}$ we have
546: \begin{equation}
547: \label{eq:current_1D}
548: I = (\frac{e}{2RC_{\Sigma}})\nu.
549: \end{equation}
550:
551: As can be seen from simulation results in
552: Fig.~\ref{all_run_curr.1.0.1D.eps}, in the limit of low $\nu$ and for
553: large system sizes, the local value of the exponent is consistent with
554: 1. Note that the for smaller system sizes (L $\le$ 1000) the effective
555: exponent is quite far away from 1. The fact that chains at least
556: larger than 1000 are required is an important observation. We will
557: revisit its implication later in the chapter.
558:
559: In the opposite regime of a high applied voltage, $\nu \gg 1$, the
560: current is determined by the average tunneling rate across a pair of
561: dots. This is given by the $\frac{1}{eR}$ of the average voltage drop
562: across a pair of dots, $\frac{V_L - V_T}{L}$, i.e.,
563: ${\overline{\Gamma}} = \frac{V_L - V_T}{eRL}$, which gives the same
564: scaling expression for the current with $\nu$ as
565: Eqn.~(\ref{eq:current_1D}). For values of $\nu \sim 1$, a crossover
566: from slow point dominated current linear scaling to high applied
567: voltage linear scaling is observed.
568:
569: \begin{figure}
570: \subfigure[]{\epsfig{figure=\fig curr.1.0.eps,
571: scale=0.45}\label{1D_curr_unif.eps}}
572: \subfigure[]{\epsfig{figure=\fig rs-4.all.1.0.eps,
573: scale=0.45}\label{eff_exp.1.0.eps}}
574: \subfigure[]{\epsfig{figure=\fig all_curr.1.0.1D.eps, scale=0.45}
575: \label{all_curr.1.0.1D.eps}}
576: \subfigure[]{\epsfig{figure=\fig all_run_curr.1.0.1D.eps, scale=0.45}
577: \label{all_run_curr.1.0.1D.eps}}
578: \caption{\label{curr.1D.1.0}
579: I-V curves for 1D array with uniform capacitances drawn on a
580: log-log plot. Fig.~\ref{eff_exp.1.0.eps} is the plot of the local
581: slope of the current versus $\nu$ derived from the data in plot
582: Fig.~\ref{1D_curr_unif.eps}. The value of the local exponent is
583: computed using the values of the current at neighboring values of
584: $\nu$, and assigned a value equal to the geometric mean of the two
585: $\nu$. For clarity small systems have been separated from the
586: largest three 1D chains simulated. In general at low voltages the
587: current scales linearly with reduced voltage ($\nu$). Flat regions
588: for smaller system sizes arise because the corresponding increase
589: in voltage is less than 1 unit, so although the rate to tunnel
590: onto the first dot increases, the rate at the ``slow point'' does
591: not change. Consequently there is no change in the amount of
592: current flowing through the array. Theoretically the current is
593: expected to scale linearly at low $\nu$ and again at high $\nu$
594: with a crossover region in between. As shown in
595: Fig.~\ref{eff_exp.1.0.eps}, the observed effective exponent is
596: approximately 0.85 for smaller system sizes, however it is close
597: to 1.0 for the largest 1D arrays at the smallest $\nu$ values as
598: shown in Fig.~\ref{all_curr.1.0.1D.eps} and
599: Fig.~\ref{all_run_curr.1.0.1D.eps}. }
600: \end{figure}
601:
602: \subsection{Non-Uniform \csigmanosp: Insulating State}
603:
604: %\renewcommand{\fig}{../FIGS/1D/}
605:
606: \begin{figure}
607: \subfigure[]{\epsfig{figure= \fig spunif_schm.eps, scale=0.6}
608: \label{spunif_schm}}
609: \subfigure[]{\epsfig{figure= \fig spdis_schm.eps, scale=0.6}
610: \label{spdis_schm}}
611: \caption{Fig.~\ref{spunif_schm} is a schematic of dot voltages in a 1D array
612: of dots with uniform \csigmanosp. The block heights indicate the
613: increase in potential on the addition of an electron. The block
614: heights are the same for different dots as \csigma is the same.
615: The broken line boxes at the top indicate the potential if another
616: electron were added. The left arrow indicate the difference in
617: potential that will determine the rate of tunneling when the
618: occupation number of the leftmost dot is 4 and its right neighbor
619: is 2. Similarly the right arrow indicates the rate of tunneling
620: when the occupation number of the last dot is 1 and the preceeding
621: dot to its left is 2. Fig.~\ref{spdis_schm} is a similar
622: schematic when \csigma is non-uniform, with \csigma for the
623: second, fourth and fifth dots different.}
624: \end{figure}
625:
626: As mentioned, the introduction of dots with non-uniform \csigma leads
627: to an array with dots of different charging voltages. For our
628: simulations, we assume \csigma to be uniformly distributed between 1.0
629: and a maximum fluctuation of 2.0. If we attribute the variation in
630: \csigma by a factor of 2 to fluctuations in the size of the dots, it
631: corresponds to a change in a variation in the linear dimension of dots
632: by a factor of $\sqrt2$. The determination of the \threshvolt gets
633: complicated by the presence of both offset charge disorder and varying
634: charging energies. This is illustrated in Fig.~\ref{spdis_schm} where
635: the spacings between voltages are different for different dots; this
636: is in contrast to dots with uniform \csigma as shown in
637: Fig.~\ref{spunif_schm}. The increment in voltage required in order to
638: tunnel between a pair of dots is due to two independent random
639: variables -- $\frac{1}{C_\Sigma}$ and ${\beta}_i$, where
640: $1/{C_{\Sigma}}$ is the charging energy of dot i and ${\beta}_i$ is
641: between 0 and 1 which represents the required increment due to the
642: offset charges. $V_T$ can be written as the
643: ${\Sigma}_i({\beta}_i/C_{\Sigma})$, where the summation runs over the
644: number of dots in the array, L. $V_T$ can therefore be written as
645: $L.\langle {\beta}_i \rangle.\langle \frac{1}{C_{\Sigma}} \rangle$,
646: where $\langle\rangle$ represents the average values. Hence
647: \begin{equation}
648: \label{eq:vthresh.dis}
649: {\overline{V_T}} = \langle 1/{C_{\Sigma}} \rangle L/2,
650: \end{equation}
651: where $\frac{1}{C_{\Sigma}}$ is $\log(2.0)$ for the assumed maximum
652: value of 2.0. Although $V_T$ scales as L, similar to the UC arrays,
653: $\sigma(V_T)$ behavior is more complicated. An exact analytical
654: expression for $\sigma(V_T)$ -- which can be derived using the
655: expectation values of $1/{C_{\Sigma}}$ and $1/{C_{\Sigma}}^2$ gives,
656: \begin{equation}
657: \label{eq:sigma_dis_vthresh.}
658: \sigma(V_T) = \sqrt{\frac{L}{6} - \frac{L}{4} (\log(2))^2}.
659: \end{equation}
660: Thus, up to leading order $\sigma(V_T)$ scales as $L^{1/2}$. This is
661: consistent with our results as can be seen in
662: Fig.~\ref{1D.vthresh.nonuniform}.
663:
664: An often used technique to explore the disorder energy scale is to
665: study the response of the system on changing the boundary
666: condition~\cite{braymoore87}. For disordered 1D QDA, we change the
667: boundary condition at the right lead and study the change in threshold
668: voltage. We define ${\Delta}V_T({\delta}V_R)$ as the difference in
669: $V_T$ on changing the value of $V_R$ by ${\delta}V_R$. Recall that
670: for UC arrays $V_T$ was completely determinable by the number of
671: up-steps in the offset charge impurities. For the uniform capacitance
672: case the response is trivial: a shift in the $V_R$ by ${\delta}V_R$ --
673: where ${\delta}V_R$ is $\frac{e}{C_{\Sigma}}$ or a
674: multiple thereof -- changes %$\langle V_T \rangle$
675: \threshvolt by the same amount. This is a consequence of the response
676: being periodic in voltage. For the 1D QDA with disordered
677: capacitances, however, a shift in $V_R$ by ${\delta}V_R$ guarantees a
678: change in $V_T$ by ${\delta}V_R$ {\it only on the average}, due to a
679: consequence of the threshold voltage being invariant to the zero-level
680: of the lead voltages. Specific values of ${\Delta}V_T$ depend upon
681: the specific disorder configuration. As a consequence of the
682: invariance just mentioned,
683: \begin{eqnarray}
684: \label{eq:invariance1.}
685: \langle {\Delta}V_T\rangle =
686: P({\Delta}V_T \neq 0)\langle{\Delta}V_T\rangle_{{\Delta}V_T \neq0}+ \nonumber \\ P({\Delta}V_T=0)\langle{\Delta}V_T\rangle_{{\Delta}V_T=0}
687: \end{eqnarray}
688: where $P({\Delta}V_T \neq 0)$ ( $P({\Delta}V_T=0)$ ) is the
689: probability that the threshold voltage changes (does not change),
690: $\langle{\Delta}V_T\rangle_{{\Delta}V_T \neq 0}$ the average of the
691: non-zero values of ${\Delta}V_T=0$. % the difference in $V_T$ and
692: Also $\langle{\Delta}V_T\rangle_{{\Delta}V_T=0}$ is 0.
693:
694: %\renewcommand{\fig}{../FIGS/1D/}
695: \begin{figure}%[htb]
696: \subfigure[]{\epsfig{figure=\fig
697: Prob_nonzeroVt.VR1.0.2.0.eps,scale=0.6} \label{prob_nonzeroVt}}
698: \subfigure[]{\epsfig{figure=\fig Mean_nonzeroVt.VR1.0.2.0.eps,
699: scale=0.6} \label{mean_nonzeroVt}}
700: \caption{
701: Fig.~\ref{prob_nonzeroVt} plots the probability for 1D systems
702: with non-uniform \csigma that that \threshvolt changes when the
703: right lead voltage is increased by one unit. The probability
704: decreases as L$^{-\frac{1}{2}}$. For 1D systems with non-uniform
705: \csigmanosp, the mean value of non-zero ${\Delta}V_T$ scales as
706: the square root of the system size. As a consequence of
707: invariance, the mean value of all ${\Delta}V_T$ is equal to the
708: value by which the right lead voltage is incremented.}
709: \end{figure}
710: The response to a change in the right lead voltage can be formulated
711: in terms of a 1D random walk problem. Given the initial and final
712: points of a random walk, one can ask what is the probability that a
713: random walk starting a distance {\it a} from the initial point of the
714: original walk, intercepts the original walk before a distance $l$? If
715: we assume that interception with the original walk results in
716: annihilation, we can ask of the surviving walks -- what is the typical
717: separation of the end-point of surviving walks from the end-point of
718: the original random walk? It is known \cite{boltzmann_award}, that
719: the probability of ``survival'' decreases as $l^{1/2}$ and the typical
720: separation scales as $l^{1/2}$ (the square root of the mean standard
721: deviation of a $l$ step random walk). The mapping to the random walk
722: problem is carried out by considering $V_R$ as the origin of the walk,
723: the potential of each dot at threshold (minus the gradient) to be the
724: positions of the original random walk and finally ${\delta}V_R$ as the
725: distance $a$ of the initial point of the second random walk from the
726: original random walk. Thus, we expect that the probability of
727: ${\Delta}V_T \neq 0$ (i.e.,
728: survival) %non-zero change in the threshold voltage
729: and the mean of the non-zero ${\Delta}V_T$
730: ($\langle{\Delta}V_T\rangle_{{\Delta}V_T \neq 0}$) should scale as
731: \lmotw and \lotwnosp respectively.
732: %should scale as \lotwnosp.
733: This is consistent with our numerical results as shown in
734: Fig.~\ref{prob_nonzeroVt} , although there are significant deviations
735: at smaller system sizes.
736:
737: \subsection{Non-uniform \csigmanosp: Conducting State}
738:
739: We discussed earlier how for UC arrays in the regime of low $\nu$, the
740: value of the current is determined by the presence of dynamically
741: important slow points. An important distinction that arises in DC
742: arrays is that the location and value of slow points is less regular.
743: For UC arrays, the value of smallest voltage drop -- and hence the
744: minimal tunneling rate -- was bound to increase every time the emitter
745: lead voltage was incremented by one unit (\eovercsigma).
746: \begin{figure}%[t]
747: \center \includegraphics[scale=0.6]{\fig single_sample_plat.eps}
748: \caption{\label{plateau}
749: I-V characteristic for a single sample 1D QDA with disordered
750: \csigmanosp. Unlike UC arrays, at small reduced voltages an the
751: current value remains constant over a range of $\nu$ values -- hence
752: the observed plateau(s). This happens when the smallest tunneling
753: rate does not change in spite of increasing $\nu$. When the rate at
754: the slow point does change, however, the value of the current jumps
755: resulting in the step like features.}
756: \end{figure}
757: Unlike UC arrays, the amount by which the emitter lead voltage must be
758: increased in order to overcome the slow point does not have a
759: well-defined lower bound and varies significantly from
760: sample-to-sample and with the value of reduced voltage. As can be
761: seen from the presence of plateaus in Fig.~\ref{plateau}, the I-V at
762: low $\nu$ for a single DC array is qualitatively different to a
763: sample-averaged I-V.
764:
765: At a given value of $\nu$, the mean tunneling rate (\mtr) is
766: proportional to the average potential gradient, and is given by,
767: \begin{equation}
768: \label{eq:relate_rate_nu}
769: \overline{\Gamma} = \frac{\nu}{2}{\langle \frac{e}{C_{\Sigma}} \rangle}.
770: \end{equation}
771: Based upon the relative values of the \mtr ~and the typical maximum
772: fluctuations from \mtr, we can categorize the applied voltage into
773: three regimes. These regimes are: (i) when the maximum fluctuation
774: are larger then the mean tunneling rate; (ii) when the maximum
775: fluctuations are of the same value as the average gradient, and (iii)
776: when the maximum fluctuations are much less then the average gradient.
777:
778: In regime (iii) the fluctuations about the mean gradient can be
779: ignored; they no longer influence the current value and consequently
780: the current is given by the Eqn.~(\ref{eq:current_1D}). The average
781: potential profile at any given dot can be computed using the average
782: potential gradient and the distance of the dot from the boundary. By
783: subtracting the dot potentials from the averaged potential profile at
784: each site, we can calculate the fluctuations, and thus the roughness
785: of the voltage surface. In regime (i) the roughness of the voltage
786: surface scales as \lotwnosp. Assuming Gaussian distribution of the
787: fluctuations, the mean value of the maximum of N variables from a
788: distribution with mean $\mu$ and standard deviation $\sigma$ is given
789: by $\mu$ + $\sigma$$\sqrt{a \log N}$~\cite{kinnison}. The deviation
790: from the mean tunneling rate is maximum at the slow point. Given that
791: the slow point can occur anywhere between any two dots (i.e., 0 and
792: L), the typical value of the maximum fluctuation scales as $\mu$ +
793: $\sigma \sqrt{a \log (\frac{L}{2})}$. The increment in voltage
794: $\Delta{\nu}$ should thus be greater than the maximum barrier
795: (fluctuation) in order that the slow point be overcome, i.e., an {\it
796: additional} electron flows over the slow point which in turn will
797: result in an increase in current. Therefore the probability that a
798: change by ${\Delta{\nu}}$ will overcome a slow point is given by the
799: probability that the typical maximum fluctuation is less than
800: ${\Delta{\nu}}$. As ${\Delta{\nu}} \sim L$, P(${\Delta{\nu}} >$
801: typical maximum) $\sim \frac{L}{L^{1/2} \sqrt{a \log(L)}}$, which
802: approaches 1 as L gets larger.
803:
804: If step like features persist in the I-V for single samples, then
805: given the sample-to-sample fluctuations in the location of the
806: plateaus, the sample averaged I-V curve will be more or less flat. As
807: seen in Fig.~\ref{all_curr_1D.2.0}, there is a voltage upto which the
808: averaged current is more or less static. The value of this voltage
809: decreases with increasing system size -- consistent with the arguments
810: that the same increase in $\nu$ is more likely to result in a slow
811: point being overcome as L gets larger. In spite of the irregular
812: change in the value of the minimum rate, the average value of the
813: minimum voltage drop across any two dots remains $\frac{V_L -
814: V_T}{L}$. Hence the average value of the minimum rate remains as
815: before -- $\frac{V_L - V_T}{eRL}$, which implies that once the
816: ``static current regime'' is overcome the current should scale
817: linearly with voltage. Thus in spite of the introduction of variable
818: \csigmanosp, current scales linearly with $\nu$ in regimes (i) and
819: (iii), similar to UC arrays. The crossover from linear scaling in
820: regime (i) to (iii) -- corresponds to regime (ii) and is more
821: complicated to understand analytically.
822:
823: \begin{figure}
824: \subfigure[]{\epsfig{figure=\fig all_curr.2.0.1D.eps, scale=0.6}
825: \label{all_curr_1D.2.0}}
826: \subfigure[]{\epsfig{figure=\fig all_run_curr.2.0.1D.eps, scale=0.6}
827: \label{all_run_curr_1D.2.0}}
828: \caption{
829: Analogous to Fig.~\ref{curr.1D.1.0}, the I-V curves and \localexp
830: for 1D arrays with disordered \csigma. The major difference is that
831: the value of the effective exponent, even for the largest 1D arrays,
832: does not appear to plateau at 1.0, but seem to flatten out at 0.85.}
833: \end{figure}
834:
835: In Fig.~\ref{all_curr_1D.2.0} and Fig.~\ref{all_run_curr_1D.2.0} we
836: plot the current and local exponent values for the largest systems (L
837: $\geq$ 500). As shown in the plot of effective exponents
838: (Fig.~\ref{all_curr_1D.2.0}), numerically we find $\beta$= 0.85 $\pm$
839: 0.02 in the low voltage regime (0.05 $< \nu <$0.5) and 1.2 $\pm 0.05$
840: in the high voltage regime ($\nu >$ 1.0). Similar exponent values are
841: found for system sizes less than L=500 (not shown).
842:
843: %\renewcommand{\fig}{../FIGS/2D/SUB/}
844: \section{2D Arrays: Insulating State}\label{sec:subthresh}
845:
846: We saw in the previous section, how for one-dimensional arrays, charge
847: flowed onto the array from the emitter lead till it was energetically
848: favorable. In this section we will attempt to develop an
849: understanding of the progressive build up of charge in two-dimensional
850: arrays, as the emitter lead voltage ($V_L$) is increased; the
851: tunneling of charge is still governed by Eqn.~\ref{eq:condition}. The
852: flow of charge onto the array can thus be viewed as lowering the
853: energy. Such {\it relaxation} of charges so as to lower the system
854: energy, is analogous to several different systems where the system
855: reaches a lower energy via a series of avalanches~\cite{dfisher98,
856: plastic2}.
857:
858: The threshold voltage is the minimal emitter lead voltage possible
859: such that when electrons tunneling onto the array from the emitter
860: lead have sufficient potential to overcome the disorder barriers and
861: reach the collector lead. Given a disorder configuration it is not
862: trivial to determine the {\it minimal} voltage for 2D arrays. A naive
863: approach might be to think of the L$ \times $W array as W, 1D arrays
864: of L dots each; trivially compute the ``threshold'' voltage for each
865: of the 1D arrays and then find the minimum. The computed minimum
866: would still probably be overestimating the true threshold voltage.
867: Determining the threshold voltage can be formulated as an optimization
868: problem, but motivated by the aim of understanding the physical
869: buildup of charge in QDA, we take a different approach. For a given
870: emitter voltage, we add charges till a meta-stable insulating state is
871: reached; then the emitter lead voltage is progressively increased,
872: building up charges until an insulating state no longer exists. The
873: value of the emitter lead voltage at which electrons first tunnel onto
874: the collector lead is our computed threshold voltage.
875:
876: \subsection{Avalanches}
877: As briefly mentioned earlier that an avalanche at a given voltage is
878: the flow of charge onto the array until the flow is arrested by
879: disorder. Avalanches in QDA are qualitatively similar to those found
880: in other systems with collective elastic transport. Some well studied
881: examples are vortex flow in disordered superconductors~\cite{plastic4}
882: and the avalanches when an interface like a CDW moves in quenched
883: disordered systems~\cite{narayan94, aam93}. For 2D arrays, the
884: location where charges tunnel in a given avalanche, helps develop the
885: notion of connected elastic domains~-- {\it basins}.
886:
887: We define $q_{i}(V_{L}^{-})$ as the charge of site $i$ before the
888: emitter lead voltage is incremented to $V_L$ and $q_{i}(V_{L}^{+})$ as
889: the charge of site $i$ after the emitter lead voltage has been
890: incremented to $V_L$. The physical size $A$ of an avalanche is the
891: number of sites where $q_{i}(V_{L}^{-}) \neq q_{i}(V_{L}^{+})$, and
892: the volume is $\sum_{i} {q_{i}(V_{L}^{-}) - q_{i}(V_{L}^{+})}$. If we
893: set $n(A,V_L)$ to be the number of avalanches between size $A$ and
894: $A+dA$, at an emitter lead voltage of $V_L$ and define $N(A)$ =
895: $\int_{0}^{V_T} n(A,V) dV$ , then $N(A)$ can be thought of as the
896: number of such avalanches that occur in going from a $V_L$= 0 to $V_L$
897: = $V_T$.
898:
899: %\renewcommand{\fig}{../FIGS/2D/SUB/}
900: \begin{figure}
901: \subfigure[]{\epsfig{figure=\fig symm.NA_A.1.0.eps, scale=0.6}
902: \label{symm.NA_A.1.0.eps}} \goodgap
903: \subfigure[]{\epsfig{figure=\fig symm.l_vs_A.1.0.eps, scale=0.6}
904: \label{symm.l_vs_A.1.0.eps}} \\
905: \caption{
906: Scaling collapse for the distribution of avalanche sizes N(A), for
907: symmetrical systems with uniform \csigma is plotted in
908: Fig.~\ref{symm.NA_A.1.0.eps}. The exponent {\it a}, gives the
909: typical size of the largest avalanches as L$^{a}$ ; avalanches of
910: sizes greater than L$^{a}$ become increasingly improbable.
911: Fig.~\ref{symm.l_vs_A.1.0.eps} shows the collapse of data for the
912: mean linear size of avalanches with size between A and A+dA for
913: systems with uniform \csigmanosp. From the scaling collapse we
914: estimate $d_f$ to have a value of (= $\rho/\sigma$) = 1.5.}
915: \end{figure}
916:
917:
918: We explore the distribution of avalanche sizes for square samples
919: (L$\times$L). The size of an avalanche $A$ can also be thought of as
920: the ``surface area'' -- which is equal to the number of dots that
921: electrons tunnel onto during an avalanche at a given $V_L$. As the
922: size of avalanches vary over several orders of magnitude -- starting
923: with avalanches of size 1 to system spanning avalanches -- we use
924: logarithmic bin sizes. Logarithmic binning is natural for exploring
925: power laws and reduces fluctuations in plots.
926:
927: \begin{figure}
928: \subfigure[]{\epsfig{figure=\fig symm.NA_A.2.0.eps, scale=0.6}
929: \label{symm.NA_A.2.0.eps}} \goodgap
930: \subfigure[]{\epsfig{figure=\fig symm.l_vs_A.2.0.eps, scale=0.6}
931: \label{symm.l_vs_A.2.0.eps}} \\
932: \caption{
933: The data collapse for distribution of avalanche sizes for
934: symmetrical arrays with non-uniform \csigma is plotted in
935: Fig.~\ref{symm.NA_A.2.0.eps}, whilst the collapse of data for the
936: mean linear size of avalanches with size between A and A+dA for
937: symmetrical systems with non-uniform \csigma is plotted in
938: Fig.~\ref{symm.l_vs_A.2.0.eps}. From the range of exponents for
939: which the scaling collapse is acceptable we get the $d_f$ (=
940: $\rho/\sigma$) = 1.5 $\pm$ for avalanches.}
941: \end{figure}
942:
943: Using standard finite size scaling, we conjecture a scaling form for
944: $N(A)$ to be of the type:
945: \begin{equation}
946: \label{eq:aval_scal_form}
947: N(A) = L^{b}{\hat N}({A/L^{a}}),
948: \end{equation}
949: where ${\hat N}(x)$ scales as $x^c$ in the limit of $x \ll 1$ and
950: ${\hat N}(x)$ approaches a constant in the limit of $x \approx 1$,
951: where $L$ is the length of the system. The exponents $a$ and $b$ are
952: determined to be those exponents for which a scaling plot of
953: $N(A)/L^{b}$ vs $A/L^{a}$ yields a single scaling function ${\hat
954: N}({A/L^{a}})$. The two exponents are not independent and can be
955: shown to be related by the relation $a + b = 3$.~\footnote{The sum of
956: the product $N(A)A$, for all avalanches upto threshold, scales as
957: $L^{3.0}$ (for systems of size L$\times$L), from which the relation
958: $a+b=3.0$ can be derived by using the scaling ansatz in the
959: integral, $\int N(A) dA \sim L^{3.0}$. For logarithmic binning as $n
960: dA = N(A) d(ln A)$, $N(A)dA = n(A)A dA$ where n(A) represents the
961: number of avalanches in the linear bin [$A, A+dA$].} In addition to
962: the two exponents $a$ and $b$, a third exponent $c$, can be used to
963: make the curve flat in the regime where $A < L^a$, which for square
964: systems of length and width L is related to the other two exponents by
965: $b - ac=2$.~\footnote{for avalanches of size $A < L^a$, N(A) scales as
966: $L^{2.0}$. Also for $A < L^a$, $N(A)/L^b \sim {(A/L^a)}^c$ where c
967: is $<$ 0, which for a given A leads to $L^{b - ac} \sim L^2$, i.e.,
968: $b - ac = 2.0$.} As there are two constraints for the three scaling
969: exponents, we get only one independent exponent from the scaling
970: collapse of the distribution of the sizes of avalanches. As shown in
971: Fig.~\ref{symm.NA_A.1.0.eps} the collapse of data to a single scaling
972: function is satisfactory, which indicates that the dimension of the
973: avalanches is \ffth. The typical size of the largest avalanches is
974: given by \lfthnosp. To study further the morphology of the
975: avalanches, we compute the the mean of the maximum length of
976: avalanches with sizes between A and A+dA. We collapse the data as
977: shown in Fig.~\ref{symm.l_vs_A.1.0.eps} onto a single curve and
978: determine that exponents $\sigma$ and $\rho$, defined in the scaling
979: function:
980: \begin{equation}
981: \label{eq:linear_size_scal_form}
982: l(A) = L^{\sigma}{\hat L}({A/L^{\rho}}),
983: \end{equation}
984: to have values consistent with \ftth and 1 respectively. We get a
985: collapse to $l/L^{\sigma} \sim ({A/L^{\rho}})^{\kappa}$ for all system
986: sizes L, thus $l \sim A^{\kappa}$ and the relation constraining the
987: exponents is therefore ${\sigma} = {\rho} {\kappa}$ or ${\kappa} =
988: {\sigma}/{\rho}$. We know that the $A \sim {l^{d_f}}$, where $d_f$ is
989: defined to be the fractal dimension, from which we get ${\kappa} =
990: 1/{d_f}$ thus $d_f = {\rho}/{\sigma}$. From the computed values of
991: ${\rho}$ and ${\sigma}$, $d_f$ works out to be 1.5. This is consistent
992: with the conclusions from the distribution of avalanche sizes.
993:
994: Finally, we have also investigated the avalanche structure using the
995: radius of gyration~($R_g$) of avalanches, which is defined in the
996: usual way as:
997: \begin{equation}
998: \label{eq:rg_exp}
999: {R_g}^2 = {{\sum(r_i - {\overline r})^2}\over N},
1000: \end{equation}
1001: and study the scaling of the mean $R_g$ for avalanches of sizes
1002: between $A$ and $A+dA$. Numerical evidence~\cite{jhathesis} indicates
1003: that the scaling of the area with $R_g$ is similar to the fractal
1004: dimension of the avalanches, which implies that the avalanche
1005: morphology is compact, i.e., does not have any significant holes.
1006:
1007: We have investigated avalanche structure using three ways and the
1008: results of all three are consistent with the hypothesis that typical
1009: avalanches are compact with dimension of \ffth. For systems with
1010: uniform \csigmanosp, the sequence of dots at which avalanches
1011: originate is periodic in left lead voltage (with periodicity
1012: \eovercsigma). We can thus think of ``basins'' of dimension
1013: $\frac{5}{3}$ evolving as charge flows into the array, with some
1014: basins growing at the expense of others. In general the basin
1015: structure is not isotropic, as they have a preferred growth direction
1016: and the linear size in the direction transverse to this preferred
1017: direction grows only as $l^{\frac{2}{3}}$ where {\it l} is the linear
1018: extent in the preferred direction. Thus in a square samples of length
1019: and width L, there are approximately $N_b(l) \sim L/l^{\frac{2}{3}}$
1020: independent regions of activity, where $N_b(l)$ is defined as the
1021: number of basins at a distance $l$ from the left lead. Hence to
1022: increase the chances of having the large basins that scale with L, we
1023: simulated systems of length L and width a multiple of
1024: L$^{\frac{2}{3}}$ (width = N$_b$ \ltthnosp).
1025:
1026: \begin{figure}
1027: \subfigure[]{\epsfig{figure=\fig NA_A.1.0.eps, scale=0.6}
1028: \label{NA_A.1.0.eps}} \goodgap
1029: \subfigure[]{\epsfig{figure=\fig l_vs_A.1.0.eps, scale=0.6}
1030: \label{l_vs_A.1.0.eps}} \\
1031: \caption{
1032: Scaling collapse for the distribution of avalanche size for UC
1033: arrays of size L$\times$ L$^{2/3}$. The exponent {\it a} and {\it
1034: c} are the same for symmetrical systems to within errors, while
1035: {\it b} differs. Fig.~\ref{l_vs_A.1.0.eps} plots the data collapse
1036: for the mean linear size of avalanches with sizes between A and A+dA
1037: for systems of size L $\times$L$^{2/3}$ with uniform \csigmanosp.
1038: Exponents are the same as those for symmetrical systems.}
1039: \end{figure}
1040:
1041: Similar to square samples, exponents in the scaling collapse for the
1042: distribution of avalanche sizes for systems of size L$\times$\ltth
1043: (Fig.~\ref{NA_A.1.0.eps}), are not independent but constrained by two
1044: relations.
1045:
1046: Given that the total number of dots is \lfth the sum of the product
1047: $N(A)A$ scales as \leth rather than L$^{3.0}$; hence $a + b=$ \feth.
1048: Also the number of avalanches in the bin [A, A+dA] scales as \lfth, so
1049: $b - ac=$ \ffth~in this case.
1050:
1051: An important difference in the avalanche structures between the
1052: uniform and disordered \csigma systems is the lack of periodicity
1053: (irregular) in emitter lead voltage and that the basins no longer
1054: evolve by quenching other basins (they overlap).
1055:
1056: Avalanches in DC arrays are not periodic in emitter lead voltage and
1057: basins don't typically evolve by quenching other basins -- they tend
1058: to overlap. This behavior is different to UC arrays. By using the
1059: three methods discussed earlier we find that capacitance disorder does
1060: not affect the structure of the avalanches.
1061: Fig.~\ref{symm.NA_A.2.0.eps} shows the scaling collapse for the
1062: distribution of avalanche sizes N(A), for square arrays with
1063: nonuniform \csigmanosp. The constraining equations in this case are
1064: now $a + b = 2.8$ (as the sum of the product of N(A)A for all
1065: avalanches $\sim L^{2.8}$) and $b - ac = 2.0$. In spite of the
1066: presence of capacitance disorder, the value for the exponent $a$ is
1067: the same to within errors for the value for uniform \csigmanosp.
1068: Exponents characterizing the scaling of mean linear size and mean
1069: $R_g$ with area for DC arrays also agree to within errors with
1070: exponent values from UC arrays. Thus avalanches remain essentially
1071: compact with a dimensionality of \ffth. For L$\times$\ltth DC arrays
1072: there is no change in the values of the exponent {\it a}, though the
1073: constraining equations change to $a + b = 2.47$ and $b- ac =
1074: 1.67$~\cite{jhathesis}.
1075:
1076: In this subsection, we have used finite-size analysis and been able to
1077: successfully relate several finite-size exponents via scaling
1078: relations. Table~\ref{table:symbol_table} provides a quick summary of
1079: the values and the context in which they are used. Taken along with
1080: the fact that these exponents and scaling relations help characterize
1081: the transition from an insulating to a conducting state (the
1082: conducting state is yet to be discussed), it is reasonable to view the
1083: transition as a {\it critical transition} with associated critical
1084: exponents and behavior.
1085:
1086: \subsection{Interface motion}
1087: The maximum advance of charge into the system at a given $V_L$ can be
1088: used to define an interface. Properties of the interface can be used
1089: to understand other properties of the system like fluctuations in
1090: $V_T$. Some details are required about the way we define the
1091: interface. At a given $V_L$, there will be some dots onto which
1092: electrons have not tunneled yet, defined relative to the original
1093: stable configuration reached by relaxing an original configuration
1094: with $0 < V_i < e/{C_\Sigma}$. We refer to such dots as zero excess
1095: dots. We can define the interface as either of the following: (i)
1096: contour of leftmost sites along each row which has not had an electron
1097: tunnel onto it, or (ii) the contour of last sites along each row which
1098: has had an electron tunnel onto it. The two although seemingly similar
1099: are different in the sense that the second definition considers the
1100: case where there may be ``bubbles'' of zero electron dots enclosed
1101: behind the interface. The difference, however, is not significant as
1102: the long wavelength properties of the interface (e.g., roughness) do
1103: not seem to depend upon which definition is used. As $V_L$ is
1104: increased, electrons tunnel onto arrays, via avalanches and if
1105: electrons tunnel onto a zero excess dots, the interface advances. The
1106: motion of the interface in response to a driving force, can be
1107: described in the language of an elastic medium driven through a random
1108: potential. We will argue that some quantitative correspondences exist
1109: in fact. The dynamics of such elastic interfaces through quenched
1110: disorder has been extensively studied in recent years \cite{barabasi},
1111: e.g., CDW, flux lines in type II superconductors etc, fluid flow
1112: through a porous medium to name some, flux front in thin films of type
1113: II superconductors \cite{surdeanu99}, combustion of paper
1114: \cite{alava1997, alava2000}.
1115:
1116: \begin{figure}
1117: \subfigure[]{\epsfig{figure= \fig width_vs_time.moa-1.1.0.eps,
1118: scale=0.6}
1119: \label{width_vs_time.moa=1.1.0.eps}}
1120: \subfigure[]{\epsfig{figure= \fig width_vs_time.moa-1.2.0.eps,scale=0.6}
1121: \label{width_vs_time.moa=1.2.0.eps}}
1122: \caption{
1123: Collapse of the roughness of the interface for $L \times L^{2/3}$
1124: systems using the scaling form given by
1125: Eqn.~\ref{eq:family-vicsek}. Here W is the system width and $t$ is
1126: measured by the distance of the mean location of the interface
1127: from the emitter lead. Fig.~\ref{width_vs_time.moa=1.1.0.eps} is
1128: for systems with uniform \csigmanosp, where the values of the
1129: exponents used in the collapse are $\alpha$= 0.5 and z=1.5.
1130: Fig.~\ref{width_vs_time.moa=1.2.0.eps} is for systems with
1131: non-uniform \csigma and the values of the exponents used in the
1132: collapse are $\alpha$= 0.5 and z=1.5 as well.}
1133: \end{figure}
1134:
1135: We define the roughness (width) of the interface as the square root of
1136: the mean of the square of the fluctuation from the mean position. On
1137: increasing $V_L$ the interface advances further into the system and
1138: gets rougher. As charge builds up behind the interface, the advance
1139: of the interface is analogous to the growth of a surface due to
1140: deposition of a material. It is well known, that such surfaces become
1141: increasingly rough with time, gradually reaching a saturation width.
1142: For QDA as the advance of the interface is governed by $V_L$; it plays
1143: the role of time, which upto a constant factor is the same as the mean
1144: position of the interface. Using the well known scaling
1145: form~\cite{familyvicsek}:
1146: \begin{equation}
1147: \label{eq:family-vicsek}
1148: w(L, t) = L^{\alpha}{\hat W}(t/L^z),
1149: \end{equation}
1150: we were able to collapse data on the width of the interface with time
1151: onto a single scaling curve Fig.~\ref{width_vs_time.moa=1.1.0.eps}. We
1152: initially used symmetrical L$\times$L systems to study the properties
1153: of the interface. Due to the large values of the dynamic exponent z
1154: (1.5), we were able to study only small system sizes with interfaces
1155: with saturated width. Consequently in order to study steady state
1156: properties of larger interface lengths, it is prudent to study
1157: non-square systems like L$\times$\ltthnosp, thereby permitting a more
1158: accurate determination of the exponents and hence the universality
1159: class the interface growth process belongs to.
1160:
1161: From the the collapse in Fig.~\ref{width_vs_time.moa=1.1.0.eps}, we
1162: find values of the roughness exponent $\alpha$ = 0.5 and dynamic
1163: exponent z = 1.5 -- therefore the growth exponent $\beta_g$ = 0.33.
1164: This is consistent with the roughening of the interface being in the
1165: KPZ universality class~\cite{barabasi}, where $\alpha$= \fotw and $z$=
1166: \fttw. The KPZ universality class is consistent with the symmetries
1167: of the system, viz., rotation is a symmetry on large scales
1168: \cite{mw93}, interactions are short range and the speed of the
1169: interface advance lacks large fluctuations. In light of the {\it
1170: assumed} lack of spatial correlation of the underlying charge
1171: disorder for dot arrays (statistically Galilean invariant) and the
1172: fact that the interface {\it will} move forward when the emitter lead
1173: voltage is increased by one unit (\eovercsigma), a description of the
1174: interface advance in terms of thermal KPZ equation seems valid.
1175:
1176: Some avalanches involve electrons hoping onto a zero excess dot -- a
1177: new-site. When an avalanche involves new-sites, the interface is
1178: reconfigured; the distribution of the avalanches that involve
1179: new-sites provides information on the reconfiguration (advance) of the
1180: interface.
1181: \begin{figure}
1182: \subfigure[]{\epsfig{figure=\fig NS_S.1.0.eps, scale=0.6}
1183: \label{NS_S.1.0.eps}} \goodgap
1184: \subfigure[]{\epsfig{figure=\fig un_cross_iface.eps, scale=0.6}
1185: \label{un_cross_iface.eps}} \\
1186: \caption{
1187: Scaling collapse for the distribution of the number of avalanches
1188: involving between s and s+ds new-sites for uniform \csigmanosp. The
1189: exponents are related by $\mu + \delta = 1.67$ and $\delta - \mu
1190: \upsilon = 0.67$. The value of $\mu = 0.67$ indicates that when the
1191: interface moves it typically advances by 1 dot spacing.
1192: Fig.~\ref{NS_S.1.0.eps} shows an avalanche crossing the interface
1193: for uniform \csigmanosp. Notice that the amount by which the
1194: avalanche overshoots the interface is of the order one.}
1195: \end{figure}
1196: The scaling collapse for the distribution of avalanches that have
1197: between s and s+ds new-sites for UC arrays is shown in
1198: Fig.~(\ref{NS_S.1.0.eps}). We define $N(s)$ analogous to $N(A)$,
1199: where $s$ is the number of new-sites visited in an avalanche. For
1200: L$\times$\ltth samples the sum of the product $N(s)s$ for all
1201: avalanches, scales as the number of dots in the array (\lfth). Thus
1202: the constraint on the exponents $\mu$ and $\delta$ in the scaling
1203: ansatz:
1204: \begin{equation}
1205: \label{eq:new-site_scal_form}
1206: N(s) = L^{\delta}{\hat \eta}(s/L^{\mu}),
1207: \end{equation}
1208: is given by $\mu + \delta= 1.67$. Another constraint is determined by
1209: the scaling of the number distribution of avalanches with the number
1210: of new-sites for a given bin ([s, s+ds]), with system length as \ltth,
1211: which results in only one independent exponent in the scaling ansatz.
1212: Hence $\delta - \mu \upsilon = 0.67$. We find that the exponent values
1213: from the collapse consistent with these constraints. We interpret the
1214: value of the exponent $\mu$ = 0.67 as giving the typical number of new
1215: sites involved in an avalanche of linear length {\it l} as l$^{0.67}$.
1216: We know that the width of the an avalanche of linear length {\it l},
1217: is also l$^{\frac{2}{3}}$, which indicates that the avalanche
1218: typically involves one new dot for each dot along the width. Thus the
1219: interface advances smoothly on the average by 1 dot along the width of
1220: the basin of activity. Fig.~\ref{un_cross_iface.eps} shows the
1221: configuration of the interface at a given $V_L$ and an avalanche that
1222: crosses the interface with the portion to the right of the interface
1223: being the new-sites involved in the avalanche. These new-sites will
1224: determine the new configuration of the interface after the avalanche.
1225: \begin{figure}
1226: \subfigure[]{\epsfig{figure=\fig NS_S.2.0.eps, scale=0.6}
1227: \label{NS_S.2.0.eps}} \goodgap
1228: \subfigure[]{\epsfig{figure=\fig dis_cross_iface.eps, scale=0.6}
1229: \label{dis_cross_iface.eps}} \\
1230: \caption{
1231: Scaling collapse for the distribution of the number of avalanches
1232: involving between s and s+ds new-sites for L $ \times L^{2/3}$
1233: systems with non-uniform \csigmanosp. The exponents are related by
1234: $\mu + \delta = 1.67$ and $\delta - \mu \upsilon = 0.67$. The value
1235: of $\mu = 1.0$ indicates that when a segment of the interface moves
1236: it typically advances by $l^{1/3}$. Fig.~\ref{dis_cross_iface.eps}
1237: shows an avalanche crossing the interface for non-uniform \csigma
1238: systems. Typically the overshoot of the $l^{2/3}$ portion of the
1239: interface is $l^{1/3}$.}
1240: \end{figure}
1241:
1242: Further information on the movement of the interface can be obtained
1243: by studying the voltages (V$_L$) at which an avalanche that involves
1244: new-sites occurs, or equivalently when the interface advances. We can
1245: define \crossvolt as the difference in $V_L$ between two avalanches
1246: that manage to cross the interface (there may be several avalanches
1247: that do not cross the interface between two interface crossing
1248: avalanches). Based upon the assumption that the advance within basins
1249: should be independent, it can be shown~\cite{jhathesis} that
1250: \crossvolt scales as W$/$l$^{\frac{2}{3}}$.
1251:
1252: We've seen how the structure of the avalanches is the same
1253: irrespective of the presence or absence of disorder in the
1254: capacitance, even though there are changes of major significance in
1255: the motion of the interface. If however, as shown in
1256: Fig.~\ref{NS_S.2.0.eps} we attempt a scaling collapse for the number
1257: of new sites covered in an avalanche the exponent values are different
1258: from the uniform \csigma exponent values. The exponent $\mu$ has a
1259: value 1.0 to within errors. This value can be interpreted as follows:
1260: the dimensionality of avalanches is \ffth, which means for linear size
1261: $l$ the width is typically $l^{\frac{2}{3}}$. When an interface
1262: crossing avalanche occurs, the average amount by which it overshoots
1263: the interface scales as $l^{\frac{1}{3}}$, hence covering
1264: $l^{\frac{2}{3}}$ x $l^{\frac{1}{3}}$ new sites.
1265:
1266: This can be seen in Fig.~\ref{dis_cross_iface.eps}, which represents a
1267: typical interface crossing avalanche in a sample with disordered
1268: \csigmanosp, where the avalanche overshoots the interface by a
1269: significant amount compared to the uniform \csigmanosp (where the
1270: overshoot was of order one spacing). For DC arrays avalanches do not
1271: occur with any fixed regularity -- either spatial or temporal -- so a
1272: large number of avalanches may occur which do not reconfigure the
1273: interface, followed by an avalanche that rearranges the interface by a
1274: large amount. Compared to the smooth motion of the interface in arrays
1275: with uniform \csigmanosp, the motion of the interface for DC arrays is
1276: rather jerky. It is important to mention that the motion appears
1277: jerky locally, but at a coarse-grained scale and {\it on average} the
1278: velocity of the interface is well defined and smooth till it reaches
1279: the collector lead.
1280:
1281: \subsection{Threshold voltages revisited}\label{sec:revisited}
1282: Similar to one-dimensional systems the \threshvolt for two-dimensional
1283: systems scales linearly with the system length. For two-dimensions
1284: however, there is an additional dependence on the width of the
1285: samples, which can be understood using the concepts of basins and
1286: interface advance from earlier subsections. It also helps explain
1287: voltage fluctuations.
1288:
1289: \begin{figure}
1290: \subfigure[]{\epsfig{figure=\fig V_t-by-L.1.0.eps, scale=0.6}
1291: \label{V_t-by-L.1.0.eps}} \goodgap
1292: \subfigure[]{\epsfig{figure=\fig V_t-by-L.2.0.eps, scale=0.6}
1293: \label{V_t-by-L.2.0.eps}} \goodgap
1294: \caption{
1295: The dependence of $V_T$ on the number of basins, N$_b$(L) for
1296: uniform \csigma systems is plotted in Fig.~\ref{V_t-by-L.1.0.eps}.
1297: The dependence of $V_T$ on the number of basins (N$_b$(L)) for
1298: non-uniform \csigma systems is plotted in
1299: Fig.~\ref{V_t-by-L.2.0.eps}.}
1300: \end{figure}
1301: With increasing $V_L$ the interface advances further along into the
1302: system until finally electrons reach the collector lead at $V_T$.
1303: Fig.~\ref{V_t-by-L.1.0.eps} shows how $V_T$ normalized by system
1304: length (L) depends upon the width of the system studied. When $N_b(L)
1305: >$ 1, in addition to fluctuations within a single basin, $V_T$ is
1306: determined by the basin that moves the interface to the right lead the
1307: earliest. With increasing $N_b(L)$, the expectation value of the
1308: maximum advance of the interface at a given $V_L$ increases,
1309: explaining the decreasing value of $\frac{V_T}{L}$. The
1310: sample-to-sample fluctuations in $V_T$ can be attributed to the
1311: roughness of the advancing interface. We saw that the roughness
1312: exponent $\alpha$ for the interface was \fotw. Assuming a value of z=
1313: \fttw, for L $\times$ W samples, where W = \ltth the interface reaches
1314: its saturation width given by W$^{\frac{1}{2}}$, which is \loth. As
1315: shown in Table \ref{table:sigmavt}, fluctuations in $V_T$ agree with
1316: this picture. Local values of the threshold fluctuation exponents are
1317: plotted in Fig.~\ref{eff_exp.threshfluct.uc} and
1318: \ref{eff_exp.threshfluct.dc} and they are consistent with the scenario
1319: depicted.
1320:
1321: %\renewcommand{\fig}{../FIGS/2D/PATHS/}
1322:
1323: \begin{figure}
1324: \subfigure[]{\epsfig{figure= \fig eff_exp.threshfluct.1.0.0.0.eps,
1325: scale=0.6} \label{eff_exp.threshfluct.uc}}
1326: \subfigure[]{\epsfig{figure= \fig eff_exp.threshfluct.2.0.0.0.eps,
1327: scale=0.6} \label{eff_exp.threshfluct.dc}}
1328: \caption{Effective exponents for the fluctuation of threshold voltage
1329: for arrays with uniform capacitance and arrays with disordered
1330: capacitance are plotted in Fig.~\ref{eff_exp.threshfluct.uc} and
1331: Fig.~\ref{eff_exp.threshfluct.dc} respectively. Sample-to-sample
1332: fluctuations in $V_T$ for nonuniform \csigma arrays are similar to
1333: the uniform \csigma and scale as $L^{0.33}$ to within errors for
1334: all values of $N_b(L)$.}
1335: \end{figure}
1336:
1337: %\renewcommand{\fig}{../FIGS/2D/SUB/}
1338:
1339: A finite-size scaling length can be defined in terms of the
1340: characteristic fluctuations in $V_T$. ${\sigma (V_T)}/V_T$ can be
1341: thought of as defining the scale characterized by $L^{-1/{\nu}_T}$,
1342: where ${\nu}_T$ is the exponent characterizing the finite-size effects
1343: on the transition and for both uniform and disordered \csigma systems
1344: has a value of \fttw.
1345:
1346: Analogous to the 1D systems, we investigated the response to changed
1347: boundary conditions -- measured by the change in right lead voltage
1348: $V_r$ -- by measuring difference in $V_T$, as a method of probing the
1349: disorder energy scale. It is shown~\cite{jhathesis} that
1350: $\langle{\Delta}V_T\rangle_{{\Delta}V_T \neq 0}$ scales as
1351: L$^\frac{1}{3}$, i.e., for 2D when \threshvolt changes, it changes on
1352: average by a value given by L$^{\frac{1}{3}}$. We have discussed the
1353: mapping between Eden growth (which is in the KPZ universality class)
1354: and DPRM~\cite{eden}. Using the analogy, the maximum point of advance
1355: of the interface in our systems can be mapped to the ground state of a
1356: DPRM. It is known that the free energy fluctuations of the ground
1357: states, both for sample-to-sample fluctuations and higher order
1358: excitations scale as L$^{\frac{1}{3}}$. For systems whose ground
1359: state (maximum advance of interface) is unable to overcome the
1360: increased voltage of the right lead (a change in boundary conditions)
1361: the next lowest energy state (interface position) needs to be enough
1362: to overcome the changed boundary condition; the energy of which is
1363: typically L$^{\frac{1}{3}}$ higher than the ground state. The \loth
1364: behavior can be understood without invoking the mapping between Eden
1365: growth and DPRM. We saw in the subsection on interfaces, that the
1366: mean voltage increment to move the interface so as to have a new
1367: maximum position scaled as \loth. When the right boundary condition is
1368: changed, either the last avalanche is able to overcome the increased
1369: right lead voltage (in which case $\Delta V_L = 0$) or requires an
1370: increase in $V_L$, in order to surmount the barrier at the right lead.
1371:
1372: \begin{table}
1373: \caption{Symbols used and comparison of values
1374: for uniform and non-uniform \csigma samples.}
1375: \begin{tabular}{cccc}
1376: Symbol & Used in & uniform \csigma & non-uniform \csigma\\
1377: \hline
1378: a, b, c & N(A) vs A & 1.7, 1.3, -0.43 & 1.7, 1.1, -0.55 \\
1379: $\rho$, $\sigma$, $\kappa$ & l vs A & 1.0, 0.63, 0.58 & 1.0, 0.67, 0.67\\
1380: $\alpha$, $\beta_g$, z & Family-Vicsek scaling & 0.5, 0.33, 1.5 & 0.5, 0.33, 1.5 \\
1381: $\mu$, $\delta$, $\upsilon$ & N(s) vs s & 0.67, 1.0, -0.3 & 1.0, 0.67, -0.05\\
1382: $\theta$ & fluctuations in $V_T$ & 0.33 & 0.33 \\
1383: $\tau$ & mean of nonzero ${\Delta}V_T$ & 0.0 & 0.3 \\
1384: \end{tabular}
1385: \label{table:symbol_table}
1386: \end{table}
1387:
1388: \begin{table}
1389: \caption{
1390: Fit values for $\theta$ for uniform and non-uniform \csigma samples,
1391: with different number of basins ($N_b$).
1392: For $\sigma({V_T})= AL^{\theta}$
1393: both A and $\theta$ were fit parameters.}
1394: \begin{tabular}{ccc}
1395: \hline
1396: $N_b(L)$ & uniform $C_{\Sigma}$ & non-uniform $C_{\Sigma}$ \\
1397: \hline
1398: 1 & 0.33 $\pm$ 0.01 & 0.33 $\pm$ 0.01 \\
1399: 2 & 0.34 $\pm$ 0.01 & 0.34 $\pm$ 0.01 \\
1400: 4 & 0.35 $\pm$ 0.01 & 0.34 $\pm$ 0.01 \\
1401: 8 & 0.36 $\pm$ 0.02 & 0.35 $\pm$ 0.01 \\
1402: \end{tabular}
1403: \label{table:sigmavt}
1404: \end{table}
1405:
1406: In addition to a well defined critical point, a true continuous phase
1407: transition is characterized by the fact that there aren't any
1408: characteristic length scales in the system, i.e., fluctuations take
1409: place at {\it all} length scales. Obviously this is not true for
1410: finite size systems -- where possibly all length scales upto the {\it
1411: system size}, but no larger can be present. This system-size
1412: dependent {\it cut-off} explains why we see finite-size deviation from
1413: the true (infinite-size) values. There is a systematic dependence of
1414: these deviations with the system sizes studied. By examining these
1415: systematic dependences on the scaling exponents, we try to extrapolate
1416: to the infinite-size value of the exponents. The fact that a
1417: system-size independent behavior is possible over a range of system
1418: sizes (e.g., the collapse of several system sizes onto a single curve)
1419: is the crucial indication that finite-size scaling approach is
1420: successful, which in turn is an indication of a phase transition.
1421: Hence the claim that \threshvolt can be viewed as the critical point
1422: of a phase transition.
1423:
1424: \section{2D Arrays at Threshold}{\label{sec:paths}}
1425:
1426: $V_T$ is defined as the lowest lead voltage at which there exists at
1427: least one dot in the column adjacent to the emitter lead, onto which
1428: electrons can tunnel and subsequently find a way onto the collector
1429: lead. Having established the existence of a threshold voltage in the
1430: previous section, our aim in this section is to understand the current
1431: conduction in the array at {\it exactly} the threshold voltage. A few
1432: samples of the first conducting path
1433: at \threshvolt %ground state paths
1434: are shown in Fig.~\ref{sample_gpaths}, from which it can be seen that
1435: there are frequent splittings and possible recombination of paths,
1436: leading to an overall complicated geometry and topology. We
1437: investigate the structure and the current carrying capacity of the
1438: ground state paths. We will find immediate application of our
1439: understanding in the next section, when we investigate the I-V
1440: characteristics of 2D arrays. In the next subsection we present the
1441: details of how the first conducting path at \threshvolt is determined.
1442: We then present results on the transverse deviations (meandering) of
1443: the path -- characterized by a wandering exponent ($\zeta$) --
1444: followed by a discussion of the main structural features of paths.
1445: Finally, we discuss the current density profiles at the end-points and
1446: establish a connection between structure and current density profiles
1447: and compare the ground state path for UC arrays and DC arrays.
1448:
1449: %\renewcommand{\fig}{../FIGS/2D/PATHS/} %\renewcommand{\thesubfigure}{}
1450:
1451: \begin{figure}
1452: \subfigure[a]{\epsfig{figure= \fig gpath1.eps, scale=0.4} \label{a}}
1453: \subfigure[b]{\epsfig{figure= \fig gpath2.eps, scale=0.4} \label{b}}\\
1454: \subfigure[c]{\epsfig{figure= \fig gpath3.eps, scale=0.4} \label{c}}
1455: \caption{A few ground state path samples illustrating the merges and splits
1456: along a single path. Some splits are ``effective'' in that paths
1457: do not recombine, for example, the middle split in figure a. Some
1458: paths have multiple splits but none go onto persist till the end
1459: (figure c). Then there are paths that split early on and then wrap
1460: around -- due to periodic boundary conditions, and merge with the
1461: original path (figure b). This results in a mouth width equal to
1462: the system width. The gray-scale is an encoding for the current
1463: density at a given dot.}\label{sample_gpaths}
1464: \end{figure}
1465:
1466: %\renewcommand{\thesubfigure}{(\alph{subfigure}) \space} % revert to default
1467:
1468: \subsection{Computing the ground state path}
1469: \label{sec:gory_details}
1470:
1471: %\renewcommand{\fig}{../FIGS/2D/V_T/}
1472:
1473: To describe the ground state path, in addition to determining the
1474: location of where electrons flow, we are interested in determining the
1475: relative probability of an electron tunneling through a given site. We
1476: will use relative probabilities as measured by current densities ($j$)
1477: [to be defined in Eqn.~(\ref{eq:curr_dncty})] and not macroscopic
1478: currents (I). We use a transfer-matrix approach to determine the
1479: relative probability of electrons tunneling through dots. This
1480: involves computing the probability of being in state {\it i}, using
1481: known probabilities of being in all possible previous states {\it j}
1482: and the transition probability of going from the states {\it j} to
1483: state {\it i}. Due to the numerical uniqueness of the random
1484: potential at each site, there is in practice only a single dot onto
1485: which electrons can tunnel from the emitter lead at $V_T$. We assign
1486: this dot, which is at the same potential as the emitter lead, a
1487: current density of 1.0, as all current flowing onto the array passes
1488: through this {\it head} dot. It can be shown that an electron cannot
1489: tunnel onto any other dot in the left most column other than the root
1490: of the spanning avalanche~\cite{jhathesis} -- the head dot. Thus all
1491: other dots in the leftmost column are assigned a probability of 0.0
1492: (the boundary condition).
1493:
1494: As electrons can tunnel onto a dot only from dots that have a higher
1495: electro-static potential. Hence in order to compute the current
1496: density of a dot (probability of an electron flowing onto a dot), the
1497: current density of all neighboring dots which have a higher potential
1498: should be known. The current density of any dot i (\cdnosp), is
1499: computed as the product of the current densities of dots in the
1500: immediate vicinity of $i$ with the probability of tunneling from the
1501: neighboring dot onto dot $i$, summed over all dots:
1502: \begin{equation}
1503: \label{eq:curr_dncty}
1504: j_i = \sum_{V_{j} > V_i} j_{j} p_{j -> i},
1505: \end{equation}
1506: where $j_{j}$ is the current density of dot j and $p_{j->i}$ is the
1507: probability of tunneling from a dot $j$ to $i$. $p_{j->i}$ is
1508: computed as,
1509: \begin{equation}
1510: \label{eq:prob_tnlng}
1511: p_{j \rightarrow i} = {\Gamma_{j \rightarrow i}}/{\Gamma_{out}^{j}} ,
1512: \end{equation}
1513: where ${\Gamma_{j \rightarrow i}}/{\Gamma_{out}^{j}}$ is the ratio of
1514: the tunneling rate from dot {\it j} onto dot {\it i} over the sum of
1515: all outgoing rates from dot {\it j}. As the probability of tunneling
1516: onto a dot from a dot at lower potential is zero. Thus starting with
1517: the head dot with a current density of 1.0 and sorting all dots in
1518: decreasing order of potential, the current density is computed for the
1519: dot with next highest potential. As the current densities and the
1520: probabilities of tunneling are known for all dots from which electrons
1521: can tunnel onto it, the current density for the new dot can be
1522: determined using Eqn.(\ref{eq:curr_dncty}).
1523:
1524: A special case of \cd is $j_L(i)$, which is defined as the current
1525: density from the $i$th dot in the last column onto the collector lead.
1526: It is useful to note that the sum of the $j's$ along any column can be
1527: greater than 1.0 (e.g. when there is intra-column hopping), but the
1528: sum of all current densities between adjacent two columns must be
1529: equal to 1.0 (essentially a current continuity equation). Thus the sum
1530: of all $j_L(i)$ will be 1.0.
1531:
1532: There isn't a simple connection between the current densities computed
1533: using our approach and the actual macroscopic currents that can be
1534: carried by a path. The current densities approach taken here, provides
1535: information on the relative proportion of the current that would
1536: tunnel off the dots onto the collector lead, i.e., be carried along
1537: different paths, but says nothing about the exact values corresponding
1538: to a given path. It is possible for example, that at threshold, a
1539: simple non-splitting path conduct more current than a highly complex
1540: path with many splittings and recombinations.
1541:
1542: %\renewcommand{\fig}{../FIGS/2D/PATHS/}
1543: \subsection{Ground state path properties} \label{sec:phys_prop}
1544: \subsubsection{Path meandering, widths and energy fluctuations}\label{sec:paths_wander}
1545:
1546: The number of end-points ($n_{ep}$) is defined as the number of dots
1547: in the last column -- adjacent to the collector lead -- which have a
1548: nonzero value of \cd (strictly speaking, a non-zero value of
1549: $j_L(i)$). As shown in Fig.~\ref{prob_nep.1000.100.1.0.0.0}
1550: \begin{figure}%[t]
1551: \includegraphics[scale=0.60]{\fig prob_nep.1000.100.1.0.0.0.eps}
1552: \caption{
1553: The probability distribution of the $n_{ep}$ taken
1554: determined using approximately 75000 disorder realizations.
1555: There is a good fit to an exponential line, implying
1556: that there is an exponentially decreasing probability of
1557: a path having higher number of end-points.
1558: Other system sizes have a similar distribution.\label{prob_nep.1000.100.1.0.0.0} }
1559: \end{figure}
1560: it becomes exponentially less probable to find paths with an an
1561: increasing number of end-points. It is relatively simple to implement
1562: a tracking algorithm that by working downwards from the head node
1563: computes the trajectory of the electron hopping and determines the
1564: number of transverse (inter-row) deviations en-route to the end nodes.
1565: The deviation of end-points is of interest and for the {\it i}$^{th}$
1566: dot is given by $y_L(i)$. The current density weighted transverse
1567: deviation can be defined as:
1568: \begin{equation}
1569: \label{eq:phii_defn}
1570: \phi_i = {j_L}{y_L}(i)
1571: \end{equation}
1572: and the current density weighted mean transverse deviation as:
1573: \begin{equation}
1574: \label{eq:Phi_defn}
1575: \Phi = {\sum \phi_i}
1576: \end{equation}
1577:
1578: \begin{figure}
1579: \includegraphics[scale=0.60]{\fig eff_exp.wandering.1.0.0.0.eps}
1580: \caption{\label{wandering.uc} Plot of the value of
1581: the effective value of the fluctuation of $\Phi$ with L which
1582: gives the wandering exponent. The wandering exponent approaches a
1583: value consistent with $\frac{2}{3}$.}
1584: \end{figure}
1585:
1586: As there are often more than a single end-point, thus the weighted
1587: mean \Phy determines the location of the effective end-point of the
1588: path and thereby the deviation of the path from the head node. The
1589: value of \Phy averaged over many samples will be zero, as there is an
1590: equal probability of the effective end-point being on either side. A
1591: look at the values of \Phy over many disorder realizations, reveals
1592: essentially a Gaussian distribution about the mean value and the
1593: standard deviation of the distribution grows as L$^{\zeta}$, where
1594: $\zeta$ is the wandering exponent and is found to have a value of
1595: $\frac{2}{3}$. This sets the scale for typical sample-to-sample
1596: fluctuations of the effective end-point. Fig.~\ref{wandering.uc}
1597: plots the local value of the exponents, which in general is a useful
1598: way to get a handle on the finite-size dependencies of the exponents.
1599: \footnote{It can be argued that if we are simulating systems of sizes
1600: L $\times $ \ltth, then the wandering exponent can be no larger than
1601: \ftth. To establish that $\zeta$ is not system width limited, we
1602: simulated systems~\cite{jhathesis} with multiple basins, i.e., with
1603: widths greater than \ltth. We find from a plot of the effective
1604: local exponent values for $\zeta$, that the value of $\zeta$ is
1605: still consistent with \ftth}
1606:
1607: \begin{figure}
1608: \subfigure[]{\epsfig{figure= \fig mouth_wdth+effexp.1.0.0.0.eps, scale=0.6}\label{mouth_width_uc}}
1609: \subfigure[]{\epsfig{figure= \fig mouth_wdth_flct+effexp.1.0.0.0.eps, scale=0.6}\label{mouth_width_flct.uc}}
1610: \caption{The width of the mouth of the path is defined as the $[max(y_L(i))
1611: - min(y_L(i))]$. In Fig.~\ref{troubledfig}(a) the left y-axis
1612: gives a plot of the width as a function of system size; using the
1613: right y-axis gives the local slope of the width with system size.
1614: As can be seen the value of the exponent settles to around 0.37
1615: $\pm 0.03$. Fig.\ref{mouth_width_flct.uc} shows the results for
1616: the fluctuation of the width of the mouth of the path. The growth
1617: in the fluctuations of the width are consistent with \lotwnosp.
1618: The left-hand y-axis represents the sample-to-sample fluctuations
1619: of the width, whereas the right-hand y-axis gives the value of the
1620: local slope.}\label{troubledfig}
1621: \end{figure}
1622:
1623: Computing the transverse deviation associated with dots helps
1624: calculate the width of the ``mouth'' of the path in the last column,
1625: which is defined as the difference of the transverse coordinates of
1626: the extreme end-points. It is of interest to understand how the width
1627: of the mouth grows with system size. The wandering exponent provides
1628: insight into the typical fluctuations of the location of current
1629: density weighted end-point but does the width of the mouth grow with
1630: the same exponent? The mean value of mouth-widths for UC arrays is
1631: shown in Fig.~\ref{troubledfig}(a). The growth in the width is
1632: consistent with \loth -- definitely different from \ltthnosp. This is
1633: indicative of a situation where the location of the effective
1634: end-point fluctuates as \ltth but the distribution of the extreme
1635: end-points around the effective end-point gets ``smaller'' relative to
1636: the effective end-point fluctuations. The increase in the mean width
1637: of the mouth tells us that paths that require the same potential
1638: difference as the ground-state path to within O(1), are to be found
1639: upto \loth around the effective end-point. The fluctuations in the
1640: %width of the mouth is plotted in Fig. \ref{mouth_width_flct.uc}. The
1641: width of the mouth is plotted in Fig.~\ref{troubledfig}(b). The
1642: fact that the fluctuation in the width scales as \lotwnosp, indicates
1643: that the extremities of the mouth are determined by randomly juggling
1644: end points. We've discussed the fluctuations in the threshold voltage
1645: in the section~\ref{sec:revisited}; we remind the reader,
1646: that as shown in Fig.~\ref{eff_exp.threshfluct.uc} and
1647: Fig.~\ref{eff_exp.threshfluct.dc}, the fluctuations in the threshold
1648: voltage scales as L$^{\frac{1}{3}}$.
1649:
1650: As a quick consistency check that the wandering exponent is not
1651: dependent on the width, we compared the wandering exponent for systems
1652: with $N_b = 4$ (L$ \times N_b$L$^{2/3}$) with those of systems with
1653: $N_b$ = 1. Although there are significant differences at smaller
1654: system sizes, for larger system sizes the boundary effects become less
1655: significant. For systems with $N_b=$4, the convergence to the
1656: $L^{2/3}$ is sooner than for $N_b=$1, indicative that boundary effects
1657: dominate at small system lengths.
1658:
1659: \subsection{Path geometry and topology: Gap sizes, merge lengths and
1660: typical splitting distances}{\label{sec:paths_structure}}
1661:
1662: So far our understanding of the structure of the ground state path is
1663: that there are possibly many branchpoints leading to multiple
1664: end-points. The current density weighted transverse deviation leads to
1665: an effective-end-point, with sample-to-sample fluctuations of \ltth
1666: and mouth-width which scales as \lothnosp. As a consequence of the many
1667: interspersed end-points between the extremal end-points of the mouth,
1668: the mouth has a fine structure not accessible by studying only the
1669: transverse fluctuations of the effective end-point and widths. We
1670: would like to understand the details of fine structure of the mouth,
1671: viz., to understand if any two physically contiguous end-points are
1672: part of the same path segment or if they belong to two different path
1673: segments. In general, if two end-points belong to different segments,
1674: typically how far back did those segments separate? Answers to these
1675: questions, will help understand several important length scales
1676: characterizing paths at threshold. It will also provide
1677: additional metrics to compare the ground-state paths of systems with
1678: different disorders (UC, DC and RT).
1679:
1680: \begin{figure}%[htb]
1681: \subfigure[]{\epsfig{figure= \fig comp_alldelta_l.1.0.0.0.eps, scale=0.60}
1682: \label{comp_alldelta_l.1.0.0.0}}
1683: \caption{
1684: Fig.~\ref{comp_alldelta_l.1.0.0.0} is a plot showing the mean
1685: lateral length for all gaps of size $\Delta$. The mean lateral
1686: length of a gap of size $\Delta$, can be thought of as the typical
1687: distance at which the path to the head-node from two end-points a
1688: distance $\Delta$ apart will have merged. The fit indicates that
1689: the lateral length scales as $3/2$ of the gap size. The plot also
1690: shows a comparison of the mean lateral lengths for the maximum
1691: and minimum gaps for an effective branchpoint. }
1692: \end{figure}
1693:
1694: In order to compute the size of gaps separating the end-points and to
1695: compute how frequently and over what length scales paths split we need
1696: to look beyond the transverse deviations of each point. We need to
1697: track the complete trajectory electrons may take before tunneling onto
1698: the end-points. This involves knowing the location of all
1699: branchpoints along the path , where a branchpoint maybe defined as a
1700: location at which a split occurs, i.e., where there is more than one
1701: neighboring dot onto which electrons can tunnel. There are many
1702: splits along the path; the majority of the splits along the path do
1703: not survive and merge a short distance after splitting. In the
1704: thermodynamic limit, not all branchpoints are of interest, but only
1705: those branchpoints that go on to produce end-points -- do not merge
1706: after splitting. We find the last possible location that is common to
1707: the trajectories associated with the two end-points of interest. This
1708: point is referred to as the effective branchpoint corresponding to the
1709: two end-points. Having thus determined the location of the effective
1710: branchpoints for all pairs we can compute the position at which paths
1711: to the end-points last overlap. Equivalently, this location can be
1712: used to establish a lateral distance from the collector lead that
1713: paths from end-points a transverse distance $\Delta y_L$ apart will
1714: most likely merge by.
1715:
1716: We then compute the mean value of the lateral length of gaps of size
1717: $\Delta$. Given that the wandering exponent has a value of
1718: $\frac{2}{3}$, one would expect that the paths to two end-points
1719: separated by $\Delta {y_L}$ would be typically joined upto a distance
1720: ${\Delta {y_L}}^{1/{\zeta}}$ from the collector leads. This is
1721: analogous to the typical separation of the optimal paths to two ends
1722: of a DPRM that are $\Delta y_{L}$ apart, viz., ${\Delta
1723: {y_L}}^{1/{\zeta}}$). As shown in
1724: Fig.~\ref{comp_alldelta_l.1.0.0.0} our findings are in good agreement
1725: with this expectation; the mean lateral length for all the $\frac{n
1726: (n-1)}{2}$ gaps for the $n$ end-points is used in the fit. The fact
1727: that the path structure of QDA is similar to the scale-invariant tree
1728: structure of DPRM, is further indication that ground state paths of
1729: QDA are in the same universality class of DPRM. The significance of
1730: this conclusion will be explored later. Also plotted are the mean
1731: lateral lengths of the maximal and minimal sized gaps of an effective
1732: branchpoint, which scale like $\frac{5}{3}$ and $\frac{4}{3}$
1733: respectively with the transverse size of the gaps.
1734:
1735: \begin{figure}
1736: \subfigure[]{\epsfig{figure= \fig
1737: fit.histo.delta.1000.100.1.0.0.0.eps, scale=0.6}
1738: \label{fit.histo.delta.1000.100.1.0.0.0}
1739: } \subfigure[]{\epsfig{figure= \fig
1740: fit.histo.l.1000.100.1.0.0.0.eps, scale=0.6}
1741: \label{fit.histo.l.1000.100.1.0.0.0}
1742: }
1743: \caption{
1744: Fits to the probability distributions of the gap sizes
1745: (Fig.~\ref{fit.histo}(a)) and the lateral length of the gaps
1746: (Fig.~\ref{fit.histo}(b)) for a single system size. Although not
1747: shown, there is not a system size dependence at which probability
1748: decreases (though there is a system size dependent cut-off). Fits
1749: indicate the probability of occurrence of a gap of size $\Delta$
1750: decreases as ${\Delta}^{-1}$. A similar dependence is observed
1751: for $l$, which is consistent with expected probability
1752: distribution for $l$ computed using a variable transformation from
1753: $\Delta$ to $l$ (where $\Delta \sim l^{2/3}$). Given the
1754: distribution of $l$, the chance that paths that split, will
1755: survive as independent paths all the way to the end gets smaller
1756: the earlier they split.}\label{fit.histo}
1757: \end{figure}
1758:
1759: The gap between two end-points is defined as the number of the
1760: intermediate dots separating them. Thus for two physically adjacent
1761: end-points (irrespective of whether they belong to the same path
1762: segment or not), the gap is defined to be of zero size. We computed
1763: the effective branchpoints for all pairs of end-points (there are
1764: $\frac{n (n-1)}{2}$ pairs for n end-points) and computed the lateral
1765: and transverse sizes of the gaps. The results of the probability
1766: distribution for a fixed system size are shown in
1767: Fig.~\ref{fit.histo.delta.1000.100.1.0.0.0} and
1768: Fig.~\ref{fit.histo.l.1000.100.1.0.0.0}. The probability
1769: distributions represent the simple fact that large gaps resulting from
1770: earlier permanent splittings of the path at the threshold become less
1771: probable. As the path at threshold and the end-points have become
1772: reachable within the last $O(1)$ increase in potential, all path
1773: segments at threshold must be equal to each other to within voltage of
1774: $O(1)$. Consequently they will overlap for the most part. We have
1775: seen that the sample to sample fluctuation in the threshold voltages
1776: scales as \loth which should set the scale for the typical difference
1777: between non-overlapping paths, thus at threshold we expect
1778: typically $O(1) \over L^{\frac{1}{3}}$ %$\frac{O(1)}{L^{\frac{-1}{3}}}$
1779: fraction of paths segments to not overlap.
1780: Consequently if for a given system size, we were to plot the mean of
1781: all lateral sizes of the gaps as a fraction of the system length, we'd
1782: expect to see a $L^{-\frac{1}{3}}$ dependence.
1783:
1784: By studying the distribution of the location of the effective
1785: branchpoints, and found that it became increasingly improbable that
1786: they would be located closer to the emitter lead~\cite{jhathesis}. It
1787: is also useful to compute the number of effective branchpoints (depth)
1788: encountered on a path to an end-point and how the depth varies for the
1789: different end-points in a given sample. This tells us if the paths to
1790: the end-points are typically similar, as well as permitting us to
1791: determine a correlation between physical proximity of end-points and
1792: difference in depths. Thus we determine the average number of
1793: end-points for a given mean depth of end-points.
1794: \begin{figure}
1795: \subfigure[]{\epsfig{figure= \fig nep_dcg.1000.100.1.0.0.0.eps, scale=0.60}
1796: \label{nep_dcg}}
1797: \caption{
1798: In Fig.~\ref{nep_dcg_all} the mean value of the $d_{cg}$ of the
1799: end-points is plotted as a function of $n_{ep}$. As the number of
1800: end-points increases the mean depth of the end-points increases
1801: logarithmically. This is indicative of an essentially balanced
1802: tree. For a perfectly balanced tree the number of end-points
1803: increases exponentially with the depth, whilst for an hierarchically
1804: split tree (in the limit of a perfectly unbalanced tree) the mean
1805: value of the depth increases linearly with the number of
1806: end-points.}\label{nep_dcg_all}
1807: \end{figure}
1808: If the tree structure was perfectly random the number of end-points
1809: would grow like the square root of the depth. On the other hand for
1810: an essentially unbalanced trees (where all the splittings take place
1811: on the path to one particular end-point), the number of end-points,
1812: $n_{ep}$ grows linearly with the mean depth. For an essentially
1813: balanced tree each path splits essentially with equal probability in
1814: which case the number of end-points grows as some number to the power
1815: of the mean depth (for a perfectly balanced tree this would be the
1816: number would be 2). As shown in Fig.~\ref{nep_dcg_all}, we find that the
1817: mean depth grows logarithmically with the number of end-points,
1818: characteristic of an essentially balanced tree.
1819:
1820:
1821: \begin{figure}
1822: \includegraphics[scale=0.60]{\fig DyL_cgdepth.1000.100.1.0.0.0.eps}
1823: \caption{\label{DyL_cgdepth.1000.100.1.0.0.0}
1824: $\Delta y_L$ is the transverse separation of two end-points on the
1825: path and $\Delta d_{cg}$ is the difference in the $d_{cg}$ between the
1826: nodes. Given the logarithmic increase in the value of
1827: $\Delta d_{cg}$ with transverse separation, the paths to the nodes
1828: typically have a net difference of one additional branchpoint with
1829: every scale of two increase in the separation between nodes.
1830: }
1831: \end{figure}
1832:
1833: Finally we'd like to determine if the trees (representing the
1834: ground-state paths) are spatially homogenous and if the path segments
1835: from the different end-points to the head node, are essentially
1836: similar in the number of effective branchpoints encountered. To do
1837: so, we investigate the difference in the depth (on the path to the
1838: end-point), represented as $\Delta d_{cg}$, of end-points separated by
1839: a transverse distance $\Delta y_L$. As shown in
1840: Fig.~\ref{DyL_cgdepth.1000.100.1.0.0.0}, the difference in depths
1841: increases logarithmically with transverse separation of end-points.
1842: Given the logarithmic dependence of $\Delta d_{cg}$ on transverse
1843: separation of end-points, the paths to the nodes typically have a net
1844: difference of one additional branchpoint with every scale of two
1845: increase in the separation between nodes.
1846:
1847: It is fair to assume that path segments that are not immediate
1848: descendents of the same parent are essentially independent; every pair
1849: of end-points with a $\Delta d_{cg} > 1$ can be considered independent
1850: and thus the plot in Fig.~\ref{DyL_cgdepth.1000.100.1.0.0.0} provides
1851: a measure of the number of independent paths segments (channels) that
1852: reach the collector lead. This could possibly be experimentally
1853: verified by studying the spectrum of the discrete current at the
1854: collector lead. Using this definition of 'independent' paths, we find
1855: that the number of independent channels increases logarithmically in
1856: the transverse direction upto the width of the mouth.
1857:
1858:
1859: \subsection{Current densities}{\label{sec:currentdensities}
1860:
1861: We have found that the structure and the topology of the first
1862: conducting path to have several interesting features and
1863: characteristic length scales. An important question is what is the
1864: profile of the current densities at the end-points?
1865: Also, what does the fluctuations of the current densities within the
1866: mouth tell us about the overall structure of the paths?
1867:
1868: To address what the current density values for a pair of
1869: end-points tell us, we take $j^>$ and $j^<$ as the values of the
1870: larger and smaller current density (for the two end-points in
1871: consideration) respectively, and as a measure of the difference in the
1872: number of splittings encountered for two end-points define $\Delta
1873: n_s$ as:
1874: \begin{equation}
1875: \label{eq:delta_ns_defn}
1876: \Delta n_s = \log({j^{>}\over j^{<}})
1877: \end{equation}
1878:
1879: \begin{figure}
1880: \subfigure[]{\epsfig{figure=\fig
1881: bda.bin_delta_yL_delta_ns.1.0.0.0.eps, scale=0.60}
1882: \label{delta_yL_delta_ns.uc}} \goodgap
1883: \caption{
1884: $\Delta n_s$ is a measure of the difference in the number of
1885: splits in the paths traversed to get to two end-points. In
1886: Fig.~\ref{delta_yL_delta_ns.uc} the dependence of the mean value
1887: of $\Delta n_s$ on the separation $\Delta y_L$ is plotted. As can
1888: be seen initially there is a logarithmic dependence of $\Delta
1889: n_s$ on $\Delta y_L$ (the fit shown here is for the largest system
1890: size) before gradually crossing over to a $\Delta y_L$ independent
1891: value. Data for widely differing system sizes (from L=343 to
1892: L=3375). To within errors, the plateau value appears to be
1893: independent of the system size.}\label{delta_yL_delta_ns.uc.all}
1894: \end{figure}
1895: In Fig.~\ref{delta_yL_delta_ns.uc.all}, we plot the value of $\Delta n_s$
1896: as the transverse separation between end-points increases. The best
1897: fit to the largest system size considered (L=3375, W=225) is
1898: consistent with a logarithmic dependence over the range 1 to about 20.
1899: From the logarithmic increase in $\Delta n_s$ with transverse distance
1900: ($\Delta y_L$), it follows that $j^{>} = \Delta y_L j^{<}$, i.e., with
1901: increasing distance between the two points considered, the larger
1902: current density ($j^{>}$) tends to get larger relative to the smaller
1903: ($j^{<}$) -- increasing linearly in $\Delta y_L$. We know from insight
1904: gained from the structure of the paths, that as the transverse
1905: distance separating two end-points increases, the paths taken to the
1906: two end-points separate earlier. If the paths to the end-points after
1907: separation typically undergo the same number of current splittings,
1908: then on average there would not be any variation in $\Delta n_s$ with
1909: distance; but given the slow but definite distance dependence, it is
1910: consistent to conclude that one path undergoes more current splits
1911: than the other, and that for end-points separated by a greater
1912: transverse distance, the correlation in current densities will be less
1913: than for those end-points which have greater overlap in their
1914: paths.
1915:
1916: It is useful to point out the similarity between the logarithmic
1917: dependence of $\Delta n_s$ on $\Delta y_L$ as in
1918: Fig.~\ref{delta_yL_delta_ns.uc} and the logarithmic dependence between
1919: the $d_{cg}$ on $\Delta y_L$ as shown
1920: Fig.~\ref{DyL_cgdepth.1000.100.1.0.0.0}. In general, given the
1921: similarity in the properties of the current densities at the
1922: end-points and the structure of the path, it appears to be the case
1923: that the effective branchpoints not only determine the structure of
1924: the paths but also play a role in determining the current profiles of
1925: the end-points.
1926:
1927: \subsection{QDAs with capacitance disorder} {\label{sec:path_ucrd}}
1928:
1929: In this subsection, we study the properties of the first path for DC
1930: arrays and begin by investigating the transverse deviations of the
1931: path and the structure of the mouth for ground-state paths.
1932: \begin{figure}%[tb]
1933: \subfigure[]{\epsfig{figure= \fig eff_exp.wandering.2.0.0.0.eps,
1934: scale=0.6} \label{wandering.dc}}
1935: \subfigure[]{\epsfig{figure=
1936: \fig comp.nep_dcg.UC_DC.eps, scale=0.6}
1937: \label{UC_DC.nep_dcg}}
1938: \caption{The values of the local slopes for the wandering exponent
1939: computed for arrays with disordered capacitance are plotted in
1940: Fig.~\ref{wandering.dc}. From Fig.~\ref{UC_DC.nep_dcg} $d_{cg}$
1941: increases logarithmically with $n_{ep}$. Both are essentially
1942: indistinguishable from UC arrays.}\label{path_dc_all}
1943: \end{figure}
1944: As shown in Fig.~\ref{path_dc_all}(a), the wandering exponent gradually
1945: approaches the value of $\zeta=$ \ftthwsp for larger systems, which is
1946: similar to UC arrays. In Fig.~\ref{UC_DC.nep_dcg}, the relationship
1947: between $d_{cg}$ and $n_{ep}$ is shown to be logarithmic (recall that
1948: the probability of occurrence decreases exponentially as $n_{ep}$
1949: increases).
1950:
1951: In Figs.~\ref{UC_DC.delta} we plot the distribution of the gaps and the mean
1952: lateral distance of splitting for gaps of size $\Delta$. Both the distribution of the gaps sizes (and thereby
1953: lateral size of the gaps) and the mean lateral distance dependence on
1954: gap sizes are similar to UC arrays.
1955:
1956: % \begin{figure}%[htb]
1957: \begin{figure}%[h]
1958: \subfigure[]{\epsfig{figure= \fig comp.bin.alldelta_l.UC_DC.eps,
1959: scale=0.60} \label{UC_DC.alldelta_l}}
1960: \subfigure[]{\epsfig{figure= \fig comp.histo.alldelta.UC_DC.eps,
1961: scale=0.6} \label{UC_DC.alldelta}}
1962: \caption{Comparing the gap distribution and spacing for UC and DC in
1963: Fig.~\ref{UC_DC.alldelta_l} and Fig.~\ref{UC_DC.delta}(b),
1964: indicates that they are essentially indistinguishable.}
1965: \label{UC_DC.delta}
1966: \end{figure}
1967:
1968: In addition to the structure of the path, current flow properties are
1969: also indistinguishable to UC systems as shown by the sample averaged
1970: fluctuations of the current-density weighted transverse
1971: locations~\cite{jhathesis}.
1972: % Fig.~\ref{sigma_phi.dc}.
1973: % and the scatter plots of the $n_s$ and $n'_s$ Fig.~\ref{ns_nprime.dc}.
1974: From the data as presented in this section, ground state paths are
1975: effectively indistinguishable from the UC. It is highly unlikely that
1976: any further investigations will indicate any significant differences
1977: between the ground path structures for the UC and DC systems.
1978:
1979:
1980: %\renewcommand{\fig}{../FIGS/2D/PATHS/}
1981:
1982:
1983: DPRM is controlled by a zero fixed point thus the ground state (lowest
1984: energy) strongly determines the properties of the system. Given the
1985: fact that the ground-state conducting path is in the same universality
1986: class as the DPRM, one would expect excited conducting paths -- those
1987: with energy higher than the ground state at higher voltages as well as
1988: ground-states at non-zero temperatures -- to be strongly influenced by
1989: the structure and energetics of the first conducting path. Thus the
1990: connection between QDA and DPRM in addition to providing an idea of the
1991: structure of the ground state path at zero temperature, indicates that
1992: an extension of the approach used here to study the ground state path
1993: might possibly be used in determining sensitivity to boundary
1994: conditions and temperature changes of the ground state paths. The
1995: latter is of significant practical importance. Given the putative
1996: similarity between QDA and DPRM we can use results obtained in the
1997: DPRM case to predict a temperature sensitivity: namely that the ground
1998: state configuration is sensitive to temperature changes and will most
1999: likely rearrange. As to whether this is sufficient to change any
2000: scaling properties will require explicit numerical and analytic work.
2001:
2002:
2003: \section{Conduction in 2D Arrays \label{sec:2d_dynamics}}
2004: % \section{Conducting State}\label{sec:dynamics}
2005:
2006: In the previous sections, we saw how the threshold voltage can be
2007: viewed as the critical point of a continuous phase transition and
2008: explored the associated critical phenomenon {\it at and below the
2009: critical point}. This sets the stage to address the next, and
2010: arguably most important question in our investigation of disordered
2011: QDA -- the nature of the critical phenomenon for voltages {\it above
2012: the critical point}. Based based on the strength of the driving
2013: force ($V$) relative to the strength of internal interactions and
2014: disorder strength, roughly three distinct regimes can be identified.
2015: The first regime can be thought of as when the scales of disorder,
2016: interaction and driving force are all similar. In this regime the
2017: role of disorder is generally crucial and the interactions between the
2018: many degrees-of-freedom result in strong deviations from a mean-field
2019: behavior. This regime typically occurs when $V$ is very close to the
2020: threshold voltage. A second regime lies at the other end of the
2021: spectrum, where the driving force is extremely strong compared to the
2022: strength of disorder and interactions between the degrees of freedom;
2023: in this regime the disorder and interactions become irrelevant and the
2024: system is driven into a linear response mode. Our {\it primary} focus
2025: will be on the investigation of the critical behavior and dynamical
2026: response close to the transition -- corresponding to the first regime.
2027: We will study the dynamic response by computing the I-V
2028: characteristics for a range of different systems sizes. Details of
2029: theory and implementation of our numerical simulations can be found in
2030: Ref.~[\onlinecite{jhathesis}].
2031:
2032: The relative strength of the interactions and disorder in turn has
2033: been used to broadly classify two widely differing types of collective
2034: transport: weak disorder relative to the strength of interactions most
2035: likely leads to an elastic structure without breaking up; an example
2036: of which are CDW. In general when the disorder is strong relative to
2037: the interactions, the elastic structure breaks and transport is far
2038: more inhomogeneous and plastic like. Examples of transport in such a
2039: plastic regime include, the flow of a non-wetting fluid in porous
2040: medium~\cite{porous_fluid}, the transport of strongly pinned
2041: two-dimensional Abrikosov flux array~\cite{plastic3}, driven
2042: collective transport of neutral carriers in randomly varying
2043: traps~\cite{plastic2, plastic6} and the flow of a fluid with no
2044: elastic interactions flowing down a rough inclined
2045: plane~\cite{plastic5} (the dirty windshield problem). We will find
2046: that conduction in the low $\nu$ regime is plastic-like, i.e., along
2047: well defined narrow channels.
2048:
2049: Recall that below $V_T$, the concept of an advancing elastic interface
2050: -- defined as the contour of maximum advance of charge along a given
2051: row was useful. As a consequence of our definition, this elastic
2052: interface is no longer well defined at driving voltages above
2053: threshold, and thus not the interface that tears and results in
2054: plastic flow. This leads to an interesting situation where as a
2055: consequence of asymmetry around threshold, the variables and
2056: description of the system on the opposite sides of the critical point
2057: are different; consequently the same exponents are not valid both
2058: above and below the transition point. This is unlike many continuous
2059: phase transitions, especially equilibrium (e.g., two-dimensional Ising
2060: magnets in the absence of an external field) but even non-equilibrium
2061: phase transitions (e.g., CDW) where the same exponents with possibly
2062: the same values characterize the critical regimes on either side of
2063: the critical point.
2064:
2065: It is instructive to review the MW scaling hypothesis originally
2066: presented in Ref.~[\onlinecite{mw93}] to understand current flow in a
2067: two-dimensional arrays before discussing the numerical results.
2068: Similar to one-dimensional arrays, any current-carrying channel at a
2069: given emitter lead voltage $V_L$, there will be $\frac{V_L - V_T}
2070: {(\frac{e}{C_{\Sigma}})}$ extra charges on average. The locations of
2071: these extra charges can be viewed as charge steps relative to the
2072: threshold configuration. Where exactly these extra charges are
2073: located on the channel depends on the underlying disorder; typically
2074: charge down-steps are where the tunneling rates are sufficiently
2075: smaller than the mean tunneling rates. The location of these steps
2076: give the {\it most likely } locations of a split in the path; and thus
2077: can be used to define a correlation length $\xi_\parallel$, where
2078: $\xi_\parallel = \frac{eL}{(V_L - V_T)C_{\Sigma}}$, where $L$ is the
2079: linear dimension between the emitter and collector leads. We have seen
2080: that the transverse deviation ($\xi_\perp$) of a path segment of
2081: length $\xi_\parallel$ is given by $\xi_\parallel^{2/3}$. Also $V_T
2082: \sim L$, thus $\xi_\parallel \sim {\nu}^{-1}$ and therefore $\xi_\perp
2083: \sim v^{-2/3}$. $\xi_\perp$ sets the scale for separation between
2084: channels {\it before splitting}. The number of channels ($N_{ch}$) at
2085: the collector lead will thus be given by $\frac{W}{\xi_\perp}$, where
2086: W is the width of the array. Under the assumption that each channel
2087: reaching the collector acts as an independent one-dimensional
2088: current-carrying chain, the current in a channel is $I_{ch} \sim \nu$.
2089: Thus the total current carried by the array will be given as:
2090: \begin{equation}{\label{eqn:mw_hypo}}
2091: I \sim {N_{ch} \times I_{ch}} \hspace{0.2in} \sim {\nu}^{5/3}
2092: \end{equation}
2093:
2094: It is important to discuss some of the assumptions that the MW
2095: hypothesis depends critically on.
2096:
2097: Firstly, that each channel behaves like a one-dimensional array and
2098: the current in the 1D array grows linearly with $\nu$. Secondly, that
2099: number of channels grows like $\sim {\nu}^{2/3}$, which in turn is
2100: dependent upon two assumptions. The first is that the transverse
2101: deviations grows like $l^{2/3}$, where $l$ is a linear dimension of
2102: the path. We have extensively verified this to be true {\it at} $V_T$;
2103: it is fair to assume that it is above $V_T$ too. The second
2104: assumption is that the most likely effective splits -- splits that do
2105: not result in a recombination -- take place at the charge down-steps.
2106: This has been harder to verify rigorously; at best we find for arrays
2107: at $V_T$ that the sample-averaged probability of effective splits
2108: decreased as $\sim \frac{1}{l}$. We find that the fluctuations in the
2109: current carrying capacity of one-dimensional arrays decreases as $\sim
2110: {\nu}^{-1/2}$
2111:
2112: It was originally predicted \cite{mw93} that to observe the true
2113: exponent arrays larger than $400 \times 400$ would be required. We
2114: will go onto show that the linear dimension required before the ``true
2115: exponent'' might be observed is probably an order of magnitude larger
2116: than initially estimated. A significant portion of the remainder of
2117: this section will be devoted in support of this statement.
2118:
2119: \subsection{Simulation results, analysis and discussion}
2120: \label{sec:dynamics}
2121: %\renewcommand{\fig}{../FIGS/2D/DYN/}
2122:
2123: From our analysis of the avalanches and path meandering, we developed
2124: an understanding that on average, a L$ \times$L$^{\frac{2}{3}}$ sized
2125: system contains one independent basin and thus one channel. Before we
2126: can understand the current carrying capacity of several interacting
2127: channels, it is important to determine how the current carrying
2128: properties of a single channel changes with driving voltage.
2129: Consequently, we will most often investigate the I-V curves of
2130: asymmetrical systems of length L and width L$^{\frac{2}{3}}$. Due to
2131: finite size effects and computational cost, this also happens to be a
2132: practical approach.
2133:
2134: We are investigating the scaling properties of a hypothesized power
2135: law between I and $\nu$, thus rather than perform a coarse grained fit
2136: to the entire I-V and generate a single value for the scaling
2137: exponent, we compute ``local'' values of the exponent $\beta$ as a
2138: function of $\nu$. We find this a more useful and meaningful
2139: representation for determining how the scaling relation between the I
2140: and $\nu$ changes. The procedure although helpful, is not
2141: sufficiently sophisticated to be a complete replacement for a rigorous
2142: fitting, as error bounds with confidence intervals are not easily
2143: determinable from this approach.
2144:
2145: \begin{figure}
2146: \subfigure[]{\epsfig{figure= \fig allcurr.1.0.eps, scale=0.6}
2147: \label{allcurr.1.0}} \goodgap
2148: \subfigure[]{\epsfig{figure= \fig all_run_slope.1.0.eps, scale=0.6}
2149: \label{all_run_slope.1.0}}
2150: \caption{
2151: I-V curves for 2D arrays with offset charge disorder for a range of
2152: system sizes. The \localexp for the I-V curves in
2153: Fig.~\ref{all_curr_uc}(a) is plotted in Fig.~\ref{all_curr_uc}(b).
2154: For a true power law scaling, the value of \localexp for different
2155: system sizes should overlap. As can be seen this does not happen for
2156: values less than $\nu = 0.1$. Following the MW scaling hypothesis,
2157: we expect a plateau at values less than $\nu = 0.1$. The inability
2158: to see clearly a definitive plateau is primarily due to large finite
2159: size effects.}\label{all_curr_uc}
2160: \end{figure}
2161:
2162: The local exponents for UC arrays are plotted in
2163: Fig.~\ref{all_curr_uc}(b), from which there is a clear dependence
2164: on system size for the local exponents. There isn't a range of $\nu$,
2165: however small, where the local exponent curves for all the different
2166: system sizes lie on a single curve, as would be expected for a valid
2167: scaling regime. Thus it is difficult to claim that there is a {\it
2168: unique single} value of the local exponent for all sizes, even over
2169: the smallest regime of $\nu$. At lower values of $\nu$ the
2170: statistical noise starts to dominate and the true value of the local
2171: exponent becomes unclear.
2172:
2173: The aim of rigorously verifying the MW scaling hypothesis numerically
2174: does not appear to be easily attainable with available computational
2175: resources at the present moment; thus it remains open, as to
2176: whether the MW scaling hypothesis is valid. If we assume that MW is
2177: the correct hypothesis, we can at best determine the constraints on
2178: system sizes and values of $\nu$ to establish a regime for the
2179: validity of the hypothesis. This is somewhat analogous to determining
2180: an upper bound of the reduced variable upto which critical behavior
2181: can be observed:
2182:
2183: It is known that for CDW one has to be within $f \leq 0.01$
2184: \cite{thorne_pt} (where $f = {(F- F_c)\over F_c}$) of the critical
2185: point in order to observe associated critical phenomenon. Similarly
2186: for high-T$_{c}$ superconductors (copper oxide) in three-dimensions,
2187: by some estimates~\cite{goldenfeld} the critical region exists for $t=
2188: 10^{-4}$ where $t = {(T- T_c)\over T_c}$ The quoted estimates are from
2189: analytical calculations and supported by numerical data. With the
2190: caveat that it is much harder to estimate correctly using numerical
2191: data alone, we hazard an estimate of the critical region for QDA
2192: solely on numerical data. From the plot of the local slopes for the
2193: largest two-dimensional QDA simulated ($L=2744$ and $W=196$ in
2194: Fig.~\ref{all_curr_uc}(b)) and the plot of \localexp for largest
2195: one-dimensional systems (L$=2000$) a similar upper bound would be
2196: somewhere between $\nu = 0.01 - 0.1$ for L$\times$L$^{\frac{2}{3}}$
2197: arrays. As can be seen from Fig.~\ref{all_curr_uc}(b), there are
2198: strong indications of a plateau in \localexpnosp, albeit over a small
2199: region -- for values of $\nu$ around 0.1 -- and only for the largest
2200: arrays. In addition, from the very brief flattening out of \localexp
2201: for $1000 \times 100$ arrays around $\nu =0.1$ before dipping, it is
2202: conceivable that for values of $\nu < 0.1$, the ``true exponent''
2203: value lie somewhere in between 1.5-2.0; this is consistent with the
2204: hypothesized value of \localexp $=5/3$.
2205: %plateau at \localexp $=5/3$ for $\nu < 0.1$ for $2744 \times 196$ systems.
2206: If at all, this will be the critical region and the likely value of
2207: the critical exponent.
2208:
2209: It is clear that simulations of even larger system sizes will be
2210: required to observe a plateau for at least a decade in $\nu$ -- which
2211: is the really the minimum range over which a power law should be
2212: observed before definitive claims of scaling are valid. As mentioned
2213: simulations of systems large than $2744 \times 196$ are currently
2214: computationally not feasible.
2215:
2216: For values of $\nu > 0.1$ the values of \localexp are influenced by a
2217: crossover to a peak value of approximately $2.0$, before being driven
2218: into the linear regime. This bump in the values of the \localexp
2219: corresponds to a regime outside of the putative MW regime, when new
2220: splits in the current carrying channels are taking place at all length
2221: scales and thus there are rapidly increasing new outlets giving rise
2222: to the value of $2.0$ for the \localexpnosp. New channels open, but
2223: are not all independent; for L$ \times$L$^{\frac{2}{3}}$ arrays these
2224: newly opened channels will typically merge with the ground state path.
2225: The effective value of \localexp at 2.0 appears to be a coincidence, a
2226: malicious one for several experiments seem to encounter this value too.
2227: The effective exponent value at a given $\nu$ is sensitive to the
2228: ratio of the length to widths, albeit in a complex fashion.
2229:
2230: As a consequence of finite-sizes, a crossover region over which the
2231: effective exponent is different from the ``true exponent'' arises.
2232: The crossover region gets larger for smaller system sizes.
2233: \begin{figure}
2234: \subfigure[]{\epsfig{figure=\fig scale_local_slope-8.1.0.eps,
2235: scale=0.6}
2236: \label{scale_local_slope-8.1.0}}\goodgap
2237: \subfigure[]{\epsfig{figure=\fig symm_scale_local_slope-8.1.0.eps,
2238: scale=0.6}
2239: \label{symm_scale_local_slope-8.1.0}} \\
2240: \caption{
2241: The collapse of the local slopes for uniform capacitance systems
2242: is plotted in Fig.~\ref{scale_local_slope_uc_all}(a). Collapse of
2243: the local slopes for symmetric systems with uniform gate
2244: capacitance is plotted in Fig.~\ref{scale_local_slope_uc_all}(b).
2245: Note that the value of ${\nu}_T$ that gives good data collapse is
2246: different from the value of ${\nu}_T$ that gives similar collapse
2247: for systems of size L x
2248: L$^{2/3}$.}\label{scale_local_slope_uc_all}
2249: \end{figure}
2250: Somewhat analogous to the finite-size scaling exponent $\nu_{T}$
2251: characterizing fluctuations in the threshold voltage, we attempt to
2252: define a finite-size exponent $\nu_{l}$, which helps characterize this
2253: crossover region over which values of the \localexp for arrays
2254: L$\times$ L$^{\frac{2}{3}}$ deviate from the true exponent. From the
2255: plot in Fig.~\ref{scale_local_slope-8.1.0} we find that the best
2256: estimate is given by $\nu_{l} = 1.5$. Although the quality of the
2257: collapse is by no means satisfactory, a couple of trends are
2258: noticeable: there appears to be a a region over which the \localexp
2259: appears to lie on a single curve (roughly over $\nu L^{1/{\nu_{l}}}$
2260: values from 1 to 10) and one notices that larger systems appear to
2261: stay on the collapsed curve upto smaller values of $\nu
2262: L^{1/{\nu_{l}}}$. It is interesting to note that {\it it appears}
2263: that $\nu_{l}$ is similar to $\nu_{T}$, which if true would imply the
2264: existence of single finite-size length scale. Also, for small systems
2265: at large $\nu$, there are strong signs from the values of \localexp
2266: that the transition towards linear behavior has begun.
2267:
2268: There are at least a couple of factors possibly preventing the
2269: observance of a true scaling region. Firstly, there are very strong
2270: finite size effects. Systems of a sufficient size are required before
2271: the putative scaling behavior can be discerned. For example, 1D
2272: channel sizes need to be long before the linear dependence of I on
2273: $\nu$ can be observed. We estimate from our analysis in
2274: Sec(\ref{sec:onedim})that they should be at least longer than 1000
2275: dots. So although using brute force computational power we have
2276: reduced significantly the statistical noise for smaller systems (e.g.,
2277: $343 \times 49$, $216 \times 36$) , these systems sizes are
2278: insufficient to actually observe the putative scaling and we observe
2279: an effective exponent not in agreement with the theoretically expected
2280: scaling values. Secondly, statistical noise needs to be reduced
2281: significantly further for larger systems. The reduction of
2282: statistical noise for large system sizes, especially at lower $\nu$,
2283: is strongly dependent on the cost of determining {\it correctly} the
2284: value of the current. We will elaborate on this further.
2285:
2286: Additional complications arise from the fact that there are large
2287: fluctuations and large timescales associated with channel formation.
2288: It is the complexity associated with both determining correctly the
2289: channel structure as well as the converged current value that makes
2290: the exact and proper simulation of electron flow in arrays such a
2291: difficult task. The two issues are in someways aspects of the same
2292: problem -- the timescales required to form a steady state current
2293: pattern are long and broadly distributed between samples. This
2294: phenomenon is common to several other dynamical systems involving
2295: collective transport and disorder~\cite{plastic3, plastic2}. The
2296: timescale required for current patterns to reach a steady state
2297: appears to be different from the timescale required for the current
2298: values to reach steady state. For a particular sample considered, the
2299: difference in current after the last channel formed was only $2 \%$,
2300: but while investigating systems of the same size and at similar
2301: $\nu$'s (to within a factor of 2), we noticed that when channels
2302: formed, there were concomitant changes in the current by more than $20
2303: \%$. This wide variation is part of the problem -- for it is
2304: difficult to estimate how much, a well formed channel will contribute
2305: to the overall current. Any adaptive algorithm based upon channel
2306: formation and activity isn't easy. As L gets larger the problem gets
2307: more acute. A somewhat similar problem, is the long time scale
2308: required to form a channel, even if there is just a single channel
2309: involved in conduction.
2310:
2311: As a consequence of the above features, it can be difficult to
2312: determine the current value correctly. At best, we can strive to
2313: minimize the probability of getting an incorrect current value. As
2314: with other simulation schemas, it soon becomes a problem of optimizing
2315: a finite amount of resources -- the reduction of statistical noise has
2316: to be traded off with systematic errors. Naively one would expect
2317: that the channels that conduct most of the current would form early
2318: on, and thus with simulations of {\it sufficient duration} the major
2319: current carrying channels will have reached a steady-state, both in
2320: terms of current carried and formation. This is not necessarily the
2321: case and even if it were, given the broad range of times for this to
2322: happen between samples of a given size, it would require setting all
2323: simulation runs to be sufficiently long to accommodate the longest
2324: time to steady-state. In addition to being difficult to estimate {\it
2325: a priori}, it would be computationally no more efficient than using
2326: an algorithm that determines dynamically whether the current channels
2327: have reached a steady-state.
2328:
2329: After accounting for initial transient effects, we set a bin size to
2330: be 10000 and compute the current in the first two bins, based upon
2331: which we use a convergence criteria (to be described later) to
2332: determine if the current has reached a steady state. If the current
2333: hasn't converged, the bin-sizes are doubled, i.e., the number of
2334: electrons that we wait for to tunnel off are doubled, after which
2335: similar checks to determine the steady state is done on the next two
2336: bins. This process disregards the history and values of previous
2337: bins. One of the reasons this is done, is because it can take very
2338: long for the steady-state distribution to be free of initial
2339: transients and biases. It is difficult to use two successive bins
2340: {\it from the same initial configuration for convergence} to determine
2341: in a definitive way whether we have reached a steady state. Our
2342: approach to correctly determining steady-state current, is to use two
2343: different initial conditions and to simulate until they each reach
2344: values that: (i) individually converge, and (ii) converge with respect
2345: to each other. (This is somewhat analogous to the situation for
2346: simulations of glassy systems where at least two different starting
2347: configurations are adopted as a measure to check against getting stuck
2348: in a local minima while exploring state space). We refer to the two
2349: starting states as the ``hot'' and ``cold'' configurations
2350: respectively. The classification of hot and cold states reflects the
2351: fact that the cold initial configuration has been prepared by the
2352: addition of electrons so as to have a smooth spatial gradient of
2353: electron potential from the emitter lead to the collector lead for the
2354: given value of $\nu$, while the hot initial configuration has a smooth
2355: spatial gradient of electron potential but corresponding some value
2356: greater $\nu'$ than the required value of $\nu$.
2357:
2358: The rate at which the local values of current change can be very
2359: different for hot and cold states; it is also typically very different
2360: for different samples. It is possible that a simple measure of
2361: convergence like setting an acceptable upper limit on the percent
2362: difference between the values of local current in two bins before
2363: considering the current to have converged, mistake the slow change to
2364: be an incorrect convergence. Any measure of convergence whether hybrid
2365: or for a single state should take into account the fluctuations in the
2366: value of the local current. Our method for determining convergence
2367: can be summarized as follows: After reaching threshold, we initialize
2368: two different states -- hot and cold. We start by simulating the hot
2369: configuration until it converges after which we switch to the cold
2370: configuration. We check if hot and cold states satisfy the
2371: convergence criteria so as to distinguish it from the convergence test
2372: of two successive bins from the same starting convergence (single
2373: convergence). Every time the cold configuration is checked for single
2374: convergence. If the cold configuration is singly converged, but the
2375: hybrid convergence criteria isn't satisfied then we switch to the
2376: previously saved hot configuration. In general, we check for hybrid
2377: convergence every time a single convergence check is performed and
2378: toggle between hot and cold states every time either one of them
2379: satisfies single convergence but the test for hybrid convergence isn't
2380: satisfied. This is an attempt to keep the dynamical evolution of the
2381: hot and cold simulations somewhat in phase. By ensuring that the
2382: individual configurations have separately converged -- once the
2383: test of hybrid convergence is passed, it is fair to assume that we
2384: have determined the steady-state current value and pattern. For the
2385: same net computational resource, the hybrid convergence method
2386: provides higher quality data~\cite{jhathesis}.
2387:
2388: Related to the ongoing analysis of statistical versus systematic
2389: errors, we present some final remarks about simulations of larger
2390: system versus more disorder averaging: just because the ability to
2391: simulate larger systems may exist, does not make it necessarily an
2392: efficient use of computational resources. Bigger may not always be
2393: better -- for it maybe possible to get more accurate results by
2394: simulating larger number of samples (disorder realizations) of smaller
2395: system sizes than smaller number of samples of larger system sizes.
2396:
2397: A system is said to be self-averaging \cite{binder+landau} for a
2398: variable A, if the error in $n$ statistically independent measurements
2399: of A ($\Delta$A) tends to go to zero as $L \rightarrow \infty$, i.e.,
2400: \begin{equation}
2401: \Delta A(n,L) = \sqrt{(<A^2> - {<A>}^2)/n}, \hspace{0.2in} n \gg 1
2402: \end{equation}
2403: If it goes to an L-independent value the system lacks self-averaging.
2404: We computed $\Delta V_T(n, L)$ for $n=10000$ samples and we find that
2405: one-dimensional arrays are {\it strongly} self-averaging, i.e.,
2406: $\Delta V_T(n, L) \sim {(nL^d)}^{-1/2}$, where $d=1$. Two-dimensional
2407: arrays are weakly self-averaging: in this case $\Delta V_T(n, L) \sim
2408: {(nL^{-2/3})}$. Generally if a system is self-averaging than
2409: simulations of larger system sizes is not counter-productive.
2410: Recapitulate that from plots of \localexp in one-dimensions, we saw
2411: that linear chains of lengths greater than a 1000 are required to
2412: observe linear scaling; this in turn in a way sets a lower bound on
2413: the sizes of two-dimensional arrays.
2414:
2415: \subsection{Disordered capacitance}
2416:
2417: %\renewcommand{\fig}{../FIGS/2D/DYN/}
2418: \begin{figure}%[hbt]
2419: \subfigure[]{\epsfig{figure=\fig allcurr.2.0.eps,
2420: scale=0.6}\label{all_curr.2.0}}
2421: \subfigure[]{\epsfig{figure=\fig all_run_slope.2.0.eps,
2422: scale=0.6}\label{all_run_slope.2.0}}
2423: \caption{I-V curves for systems with disordered capacitance are plotted in
2424: Fig.~\ref{all_curr_dc}(a). Local slopes of the I-V curves are
2425: plotted in Fig.~\ref{all_curr_dc}(b).}\label{all_curr_dc}
2426: \end{figure}
2427:
2428: We plot the I-V curves and the $\beta_{local}$ for DC arrays in
2429: Fig.~\ref{all_curr_dc}(a). As can be seen in Fig.~\ref{all_curr_dc}(b)
2430: there does not exist a regime where a definite single value of the
2431: exponent describes the scaling of I with $\nu$ for all system sizes.
2432: Unlike UC systems, we have not simulated arrays of size 2744 $\times$
2433: 196 but only upto 1000$\times$100 -- which goes to substantiate the
2434: dependence of the putative scaling exponent on system sizes.
2435:
2436: \section{Other Results}\label{sec:or}
2437:
2438: %\renewcommand{\fig}{../FIGS/1D/}
2439:
2440: \begin{figure}%[htb]
2441: \center \includegraphics[scale=0.6]{\fig
2442: all_run_curr.1.0.1.0.1D.eps}
2443: \caption{
2444: \label{all_run_curr.1.0.1.0.1D} The \localexp for
2445: 1D arrays with offset charge disorder and tunneling resistance
2446: disorder. The effective exponents are quantitatively similar to 1D
2447: arrays without tunneling disorder. For larger system sizes the
2448: value of \localexp approaches 1 at smaller $\nu$. Although the
2449: slow points are a consequence of a combination of tunneling
2450: resistance fluctuations and small voltage differences, the basic
2451: mechanism of overcoming slow points with increasing voltage remains
2452: and thus the linear dependence on increasing $\nu$.}
2453: \end{figure}
2454:
2455: We briefly discuss results for for one-dimensional arrays with
2456: tunneling disorder. Along with the understanding of paths at threshold
2457: from section~\ref{sec:paths}, we can use it to infer some basic
2458: features of the ground-state path of RD arrays, even though we have
2459: not explicitly simulated 2D systems with tunneling disorder. The
2460: \localexp for 1D arrays with tunneling disorder is shown in
2461: Fig.~\ref{all_run_curr.1.0.1.0.1D}. The \localexp are qualitatively
2462: and even quantitatively similar to 1D arrays without tunneling
2463: disorder. We note that the value of \localexp approaches 1 at smaller
2464: $\nu$ for larger system sizes. The system sizes and values of $\nu$ at
2465: which they approach 1 are essentially similar to 1D arrays without
2466: tunneling disorder. So although the slow points now are a consequence
2467: of a combination of tunneling resistance fluctuations and local
2468: minimums in the potential gradient (rate differences), the basic
2469: mechanism as outlined earlier for 1D arrays of overcoming slow points
2470: with increasing voltage remains valid and thus the linear dependence
2471: on increasing $\nu$. This combined with results of the transverse
2472: deviation of paths at threshold, where we found that $\zeta$ scales as
2473: L$^{\frac{2}{3}}$, irrespective of the type of disorder, indicates
2474: that {\it a priori} there is no reason to expect that splittings will
2475: occur any differently (pre-factors may change) and thus the
2476: probability is very small that I-V scaling on introducing tunneling
2477: disorder will be any different in the thermodynamic limit.
2478:
2479: %\renewcommand{\fig}{../FIGS/2D/PATHS/}
2480:
2481: \begin{figure}%[htb]
2482: \subfigure[]{\epsfig{figure=\fig eff_exp.wandering.1.0.1.0.eps,
2483: scale=0.60} \label{wandering.rd} }
2484: \subfigure[]{\epsfig{figure=\fig comp.nep_dcg.UC_RD.eps, scale=0.60}
2485: \label{UC_RD.nep_dcg} }
2486: \caption{Fig.~\ref{wandering.rd} plots the dependence of the standard
2487: deviation of $\Phi$ with L for arrays with both offset charge
2488: disorder and random tunneling resistances is plotted. Note that
2489: the in spite of the introduction of resistance disorder the
2490: scaling exponent of the transverse meanderings is not different
2491: from the meandering of the first path for UC arrays.
2492: Fig.~\ref{UC_RD.nep_dcg} shows a comparison of the relationship
2493: between $n_{ep}$ and $d_{cg}$ for UC and RD arrays. A logarithmic
2494: dependence ($d_{cg} \approx \log(n_{ep})$) holds for both.}\label{otherresults.1}
2495: \end{figure}
2496:
2497: Similar to the comparisons of $\zeta$ between DC and UC arrays, we
2498: compute the wandering exponent for RD arrays. We find that the value
2499: of the wandering exponent $\zeta$, as shown in
2500: Fig.~\ref{otherresults.1}(a), approaches the value of $\frac{2}{3}$ as the
2501: size of systems simulated gets larger. Also, as shown in
2502: Fig.~\ref{otherresults.1}(b), the structural properties of the
2503: ground-state path as measured by the relationship between the depth
2504: and the number of end-points is similar to that of UC. The
2505: probability distribution of gaps and the mean lateral length of
2506: separation for gaps of size $\Delta$ are plotted in
2507: Fig.~\ref{UC_RD.alldelta} and Fig.~\ref{UC_RD.alldelta_l}
2508: respectively. Differences with UC if any are not significant. From
2509: the comparisons between UC and RD arrays as well as UC and DC arrays,
2510: it can be confidently said that the main features of the ground-state
2511: path -- meandering, structure and geometry -- are invariant to the
2512: type of underlying disorder.
2513:
2514: \begin{figure}%[th]
2515: \subfigure[]{\epsfig{figure= \fig comp.histo.alldelta.UC_RD.eps,
2516: scale=0.6}
2517: \label{UC_RD.alldelta} }
2518: \subfigure[]{\epsfig{figure= \fig comp.bin.alldelta_l.UC_RD.eps,
2519: scale=0.6} \label{UC_RD.alldelta_l} }
2520: \caption{Plots comparing the probability distribution of gap sizes
2521: and the dependence of mean lateral length with gap sizes for UC
2522: and RD arrays. The mean lateral length scales as
2523: ${\Delta}^{\frac{3}{2}}$ for RD arrays as shown in
2524: Fig.~\ref{UC_RD.alldelta_l} whereas, Fig.~\ref{UC_RD.alldelta}
2525: compares the probability distribution of the gap sizes for UC and
2526: RD arrays.}
2527: \end{figure}
2528:
2529: There have been suggestions based on experiments
2530: Ref.~[\onlinecite{jaeger01}] that the presence of the tunneling
2531: disorder could lead to greater transverse fluctuations because of the
2532: introduction of additional possible bottlenecks arising due to the
2533: large fluctuations in the tunneling resistances. Based upon numerical
2534: simulations, we do not notice any changes from the properties of UC
2535: arrays in the transverse meandering or the structure of the paths.
2536: %\renewcommand{\fig}{../FIGS/2D/V_T/}
2537: \begin{figure}
2538: \subfigure[]{\epsfig{figure= \fig
2539: view_thresh_prob.v1.0.100.100.1.0.0.0.0.0000.400.eps,scale=0.25}
2540: \label{compare_gs_path.uc}} \subfigure[]{\epsfig{figure= \fig
2541: view_thresh_prob.v1.0.100.100.1.0.1.0.0.0000.400.eps,scale=0.25}
2542: \label{compare_gs_path.rt}}
2543: \caption{Comparing the ground state path at $\nu = 0.0$.
2544: Fig.~\ref{compare_gs_path}(b) is an array with exactly similar
2545: charge disorder as the array in Fig.~\ref{compare_gs_path}(a), but
2546: with tunneling resistance disorder included. Although the current
2547: densities at various locations are different, the overall structure
2548: is similar.}\label{compare_gs_path}
2549: \end{figure}
2550: To validate further the claim we carried out the following numerical
2551: experiment: we first computed the path in a few UC arrays. Keeping
2552: everything else the same, we introduced disorder in the tunneling
2553: resistances in the otherwise similar arrays and computed the paths.
2554: For the four different arrays we experimented with, we did not find
2555: any significant changes in the structure of path (although actual
2556: values of the current densities will be different).
2557: Fig.~\ref{compare_gs_path} shows the results for one of them.
2558: Although not conclusive, this is indicative that resistance disorder
2559: at most changes the current density distribution for a given sample
2560: and that change is indistinguishable when averaged over many samples.
2561: It is important to mention that the ground state paths in
2562: Fig.~\ref{compare_gs_path} and similar experiments, were not computed
2563: using the transfer-matrix approach, but was dynamically determined at
2564: $\nu = 0.0$. It is possible, however, that due to greater dynamical
2565: freedom in selecting current flow paths at higher values of $\nu$,
2566: there still be differences in the properties of current carrying
2567: paths.
2568:
2569: \section{Summary and Conclusions}\label{sec:sac}
2570:
2571: Using computer simulations we can easily control the presence of
2572: different disorder types and thus discern the individual and
2573: collective effects. In doing so, we find that the presence of
2574: background charge disorder is the dominant type of disorder, and
2575: although there are some minor changes for arrays with variable
2576: capacitance and tunneling disorder, the main scaling arguments and
2577: exponents characterizing the arrays at \threshvolt and in the
2578: conducting regime close to \threshvolt remain unchanged. A study of
2579: the interface properties in section \ref{sec:subthresh} indicated that
2580: the ground-state path for two-dimensional QDA should belong to the
2581: same universality class as the DPRM. By looking at the structure and
2582: the transverse deviations of the ground-state path we were able to
2583: establish the connection conclusively. We also saw in
2584: sections~\ref{sec:onedim} and~\ref{sec:paths} that the introduction of
2585: disordered \csigma does not change the current-scaling exponents for
2586: one-dimensional arrays nor of the ground-state paths. From
2587: section~\ref{sec:2d_dynamics}, it appears that the scaling exponent
2588: $\zeta$ for 2D arrays does not depend upon the types of disorder
2589: simulated either.
2590:
2591: The dominance of charge disorder is probably due to the fact that the
2592: disorder energy scale is set by the presence of the background charge
2593: impurities, is the crucial energy scale of the system. This in part
2594: is due to the fact that the fluctuation between the charging energy of
2595: dots as a consequence of the particular parameter values we choose
2596: ($C_{\Sigma}^{max}$ = 2.0), is less than the fluctuation in
2597: electrostatic energy due to offset charges being chosen randomly
2598: between [0,1[. The presence of tunneling disorder does not change the
2599: energetics of the arrays, i.e., $V_T$ and fluctuations in $V_T$. It is
2600: important to remark, however, that if the offset-charge impurities
2601: were disregarded and only a non-uniform \csigma considered, the arrays
2602: would still exhibit a threshold voltage, separating the insulating and
2603: conducting phases and most properties would still be similar to the
2604: situation where there was only offset-charge impurities. If the
2605: energy fluctuations due to of non-uniform \csigmanosp, was greater
2606: than the background charges the claim would be that the dominant form
2607: of disorder was the \csigma {\it although the properties of the array
2608: would be essentially insensitive to which was the dominant disorder}.
2609:
2610: %non-uniform \esigma and the addition of background offset-charge
2611: %impurities had a minor contribution,
2612:
2613: From our discussion in section~[\ref{sec:2d_dynamics}], we make the
2614: important conclusion that it is most likely that the MW hypothesis is
2615: correct and valid for disordered QDA irrespective of the actual
2616: relative strengths of the disorder. It is important to appreciate
2617: that one needs {\it to get sufficiently close to threshold to observe
2618: the scaling and that too only for large systems}.
2619:
2620: From our experience, a naive approach to determining the steady state
2621: current consistently underestimated the current values, which tends to
2622: get more acute at lower values of $\nu$. As a consequence, a higher
2623: putative value of $\beta$ would be observed. It also follows that
2624: there is a need for careful simulations, for with slightly less
2625: diligence it would have been tempting to predict an exponent range of
2626: 2.0-2.25. This is intertwined with the issue of high computational
2627: cost, which arises from a combination of the need to compute the
2628: converged current accurately for a {\it single} sample and the need to
2629: simulate large systems as a consequence of strong finite-size effects.
2630: It is not easy to formulate an elegant algorithmic solution to this
2631: problem. Although, parallelization is a well defined and often used
2632: approach to reduce time-to-solution of a problem, our problem does not
2633: appear to be a suitable candidate, for as mentioned, one of the
2634: primary bottlenecks in our simulations is the extremely long times
2635: required to reach a steady-state configuration. It is
2636: physically-meaningless to run a simulation at a time $t_2$ without
2637: state information at time $t_1$, where $t_2$ is a time later than time
2638: $t_1$; thus there is a fundamental limitation on speed-up that can be
2639: achieved via parallelization. But parallelization possibly along the
2640: lines of Ref.~[\onlinecite{korniss00}] may be a possible route forward.
2641:
2642: We have focused on QDA in the extreme limit where the screening-length
2643: is less than the spacing between dots. The opposite regime of
2644: essentially infinite screening-length has been well studied, both
2645: numerically and theoretically for non-disordered arrays
2646: \cite{bakhalov89} and recently for arrays with a random background
2647: potential~\cite{likharev_prb03} -- although neither of these studies,
2648: nor others that we are aware of, use the {\it statistical physics}
2649: approach that we have used. Surprisingly, there has been little
2650: activity in the regime representing the middle ground, viz., a
2651: screening length from a few upto a dozen dot spacings. With a
2652: screening-length more than a single dot spacing, the on-site
2653: interaction model that we have used in this work is not valid and
2654: computational approaches will require fundamental reworking.
2655: Ironically this regime is important (and interesting), as most
2656: nanoparticle arrays as a consequence of the absence of an underlying
2657: gate {\it most probably} have an electrostatic screening-length of a
2658: few dot spacings.
2659:
2660: In summary, we have investigated the effect of disorder on the
2661: transport of electrons in arrays of mesoscopic sized metallic islands,
2662: at, below and above a critical voltage $V_T$. In contrast to
2663: experiments, using computer simulations we can easily control the
2664: effects of different disorder types. We find that the presence of
2665: background charge disorder is the dominant type of disorder and
2666: although there are some minor changes with the addition of variable
2667: capacitance and tunneling disorder, the main scaling arguments and
2668: exponents characterizing the arrays at threshold and in the conducting
2669: regime remain unchanged. Our numerical results indicate a value for
2670: the exponent $\beta$ to be in the range 1.5-2.0.
2671:
2672: \begin{acknowledgments}
2673: One of us (SJ) would like to thank Dave McNamara for helpful
2674: discussions during the initial stages of the work. This work was
2675: supported in part by the National Science Foundation
2676: DMR-0109164. %(DMR-9702242)
2677:
2678: \end{acknowledgments}
2679: \bibliographystyle{apsrev}
2680: \bibliography{qdots}
2681: \end{document}
2682: