quant-ph0512183/si4.tex
1: \documentclass[twocolumn,prb,showpacs,amsmath,amssymb]{revtex4}
2: %\documentclass[preprint,showpacs,preprintnumbers,amsmath,amssymb]{revtex4}
3: 
4: % Some other (several out of many) possibilities
5: %\documentclass[preprint,aps]{revtex4}
6: %\documentclass[preprint,aps,draft]{revtex4}
7: %\documentclass[prb]{revtex4}% Physical Review B
8: 
9: \usepackage{graphicx}% Include figure files
10: %\usepackage{dcolumn}% Align table columns on decimal point
11: \usepackage{bm}% bold math
12: \usepackage{epsf}
13: 
14: %\nofiles
15: 
16: \newcommand{\hpl}[1]{[{\bf hpl: #1}]}
17: \newcommand{\sr}[1]{[{\bf sr: #1}]}
18: \newcommand{\mk}[1]{[{\bf mk: #1}]}
19: 
20: \newcommand{\beq}{\begin{equation}}
21: \newcommand{\eeq}{\end{equation}}
22: \newcommand{\beqa}{\begin{eqnarray}}
23: \newcommand{\eeqa}{\end{eqnarray}}
24: \newcommand{\beqan}{\begin{eqnarray*}}
25: \newcommand{\eeqan}{\end{eqnarray*}}
26: \newcommand{\ben}{\begin{enumerate}}
27: \newcommand{\een}{\end{enumerate}}
28: \newcommand{\bit}{\begin{itemize}}
29: \newcommand{\eit}{\end{itemize}}
30: 
31: \newcommand{\refeq}[1]{(\ref{#1})}
32: \newcommand{\code}[1]{{\smaller\texttt{#1}}}
33: \newcommand{\ep}{\thinspace . }
34: 
35: 
36: \newcommand{\mathbfx}[1]{{\mbox{\boldmath $#1$}}}
37: \newcommand{\mathbfxg}[1]{{\mbox{\boldmath $#1$}}}
38: \newcommand{\modten}[1]{\mathbfx{#1}}
39: \newcommand{\modvec}[1]{\mathbfx{#1}}
40: \newcommand{\phivec}{\mathbfx{\phi}}
41: \newcommand{\norm}[1]{|| #1 ||}
42: \newcommand{\Norm}[2]{|| #1 ||_{#2}}
43: \newcommand{\w}{\mathbfx{w}}
44: \newcommand{\x}{\mathbfx{x}}
45: \newcommand{\y}{\mathbfx{y}}
46: \renewcommand{\k}{\mathbfx{k}}
47: \newcommand{\kk}{\mathbfx{k}}
48: \newcommand{\ii}{\mathbfx{i}}
49: \newcommand{\jj}{\mathbfx{j}}
50: \renewcommand{\u}{\mathbfx{u}}
51: \renewcommand{\b}{\mathbfx{b}}
52: \newcommand{\A}{\mathbfx{A}}
53: \renewcommand{\a}{\mathbfx{a}}
54: \newcommand{\C}{\mathbfx{C}}
55: \newcommand{\F}{\mathbfx{F}}
56: \newcommand{\G}{\mathbfx{G}}
57: \newcommand{\Id}{\mathbfx{I}}
58: \newcommand{\J}{\mathbfx{J}}
59: \newcommand{\K}{\mathbfx{K}}
60: \newcommand{\U}{\mathbfx{U}}
61: \renewcommand{\L}{\mathbfx{L}}
62: \newcommand{\Null}{\mathbfx{0}}
63: \newcommand{\D}{\mathbfx{D}}
64: \newcommand{\M}{\mathbfx{M}}
65: \newcommand{\N}{\mathbfx{N}}
66: \newcommand{\T}{\mathbfx{T}}
67: \newcommand{\bft}{\mathbfx{t}}
68: \newcommand{\B}{\mathbfx{B}}
69: \newcommand{\Q}{\mathbfx{Q}}
70: \renewcommand{\P}{\mathbfx{P}}
71: \def\R{\mathbfx{R}}
72: \renewcommand{\c}{\mathbfx{c}}
73: \newcommand{\bfr}{\mathbfx{r}}
74: \newcommand{\p}{\mathbfx{p}}
75: \newcommand{\s}{\mathbfx{s}}
76: \newcommand{\z}{\mathbfx{z}}
77: \newcommand{\n}{\mathbfx{n}}
78: \newcommand{\f}{\mathbfx{f}}
79: \newcommand{\q}{\mathbfx{q}}
80: \renewcommand{\v}{\mathbfx{v}}
81: \newcommand{\e}{\mathbfx{e}}
82: \newcommand{\tde}{\modten{\varepsilon}}
83: \newcommand{\tds}{\modten{\sigma}}
84: 
85: \begin{document}
86: 
87: %\preprint{APS/123-QED}
88: 
89: \title{A finite element analysis of a silicon based double quantum
90: dot structure}
91: % Force line breaks with \\
92: \author{S. Rahman$^1$}
93: \email{S.Rahman.00@cantab.net}
94: % \altaffiliation[Also at ]{Physics Department, Cambridge University.}%Lines break automatically or can be forced with \\
95: \author{J. Gorman$^1$}%
96: \author{C. H. W. Barnes$^1$}%
97: \author{D. A. Williams$^2$}%
98: \author{H. P. Langtangen$^3$}%
99: \affiliation{$^1$Cavendish Laboratory, Cambridge
100: University, J J Thomson Avenue, Cambridge, CB3 OHE, UK}%
101: \affiliation{$^2$Hitachi Cambridge Laboratory, J J Thomson Avenue, Cambridge, CB3 OHE, UK}%
102: \affiliation{$^3$Simula Research Laboratory, Martin Linges v 17,
103: Fornebu P.O.Box 134, 1325 Lysaker, Norway}%
104: 
105: \date{\today}% It is always \today, today,
106:              %  but any date may be explicitly specified
107: 
108: \begin{abstract}
109: 
110: We present finite-element solutions of the Laplace equation for
111: the silicon-based trench-isolated double quantum-dot and the
112: capacitively-coupled single-electron transistor device
113: architecture. This system is a candidate for charge and spin-based
114: quantum computation in the solid state, as demonstrated by recent
115: coherent-charge oscillation experiments. Our key findings
116: demonstrate control of the electric potential and electric field
117: in the vicinity of the double quantum-dot by the electric
118: potential applied to the in-plane gates. This constitutes a useful
119: theoretical analysis of the silicon-based architecture for quantum
120: information processing applications.
121: 
122: \end{abstract}
123: 
124: \pacs{85.35.Be, 03.67.Lx}% PACS,  the Physics and Astronomy
125: %Classification Scheme.
126: 
127: \maketitle
128: 
129: Recent experiments conducted on trench-isolated double quantum-dot
130: (IDQD) structures have successfully demonstrated detection of
131: single-electron polarization,\cite{Emiroglu:03} and
132: coherent-charge oscillation.\cite{Gorman:05} This highlights the
133: possibility of constructing charge-based quantum computer circuits
134: in Si, with coherence times of the order
135: $100$~ns.\cite{NielsonBook,Loss:98} The architecture for a single
136: qubit device is a complex, three-dimensional structure consisting
137: of a single-electron transistor (SET), an IDQD, and gate
138: electrodes. This makes it difficult to determine theoretically the
139: system evolution by means of a complete and self-consistent
140: Schr\"{o}dinger-Poisson analysis. This is particularly the case if
141: the analysis were to fully take into account the device geometry
142: and all interactions while performing quantum manipulations within
143: the coherence time of the qubit.\cite{ft1}
144: 
145: In this paper, we present a significant contribution to such an
146: analysis by a finite-element solution of the Laplace equation with
147: the three-dimensional device geometry and material composition
148: taken into account. The aim of this work is to demonstrate the
149: electrostatic effect on the IDQD structure when voltages are
150: applied to the in-plane control gates of the device.
151: 
152: Figures~\ref{fig:schematic}(a) and \ref{fig:schematic}(b) show
153: device schematics in the $x$-$y$ and $x$-$z$-planes, respectively.
154: The trench isolation, illustrated in Fig.~\ref{fig:schematic}(b),
155: is formed by high-resolution electron-beam lithography and
156: reactive-ion etching. Each trench is approximately $150$~nm deep
157: and runs into the buried-oxide (BOX) layer of the
158: silicon-on-insulator (SOI) wafer. The active regions of the device
159: elements are P doped Si, which are electrically isolated from
160: other device elements, as seen in Fig.~\ref{fig:schematic}. In
161: this work, we use rectangular approximations to the device
162: elements and the etched profiles to simplify the
163: analysis.\cite{ft2}
164: 
165: \begin{figure}[!h]
166: \begin{center}
167: \rotatebox{0}{\scalebox{0.65}[0.65]{\includegraphics*[0.0in,4.25in][6.0in,8.25in]{f1.eps}}}
168: \end{center}
169: \caption{A schematic representation of the Si IDQD device used in
170: the numerical simulations. Rectangular approximations are made to
171: the device elements. (a) A view of the $x$-$y$ plane at the level
172: of the dashed line in (b). (b) A view of the $x$-$z$ plane at the
173: level of the dashed line in (a). The trench-like structures are
174: connected to the in-plane metal gates through a column of air.
175: Each trench extends about $150~$nm above the oxide base.}
176: \label{fig:schematic}
177: \end{figure}
178: 
179: The small dimensions of the quantum dots and the $20$~nm
180: constriction between them, which is fully depleted and acts as a
181: tunable inter-dot tunnel barrier, result in a significant
182: double-well type confinement potential for the electrons that
183: occupy discrete quantum states on each quantum
184: dot.\cite{Fujisawa:98,vanDerWiel:02} The tunable inter-dot
185: coupling causes the wave functions in the two dots to overlap and
186: hybridize so that they may be thought of as pseudo-molecular
187: states of an artificial two atom molecule. The IDQD is an
188: electrically-isolated component that is coupled only capacitively
189: to the rest of the circuit, including the SET for read-out, and
190: the in-plane control gates (G1 to G3) for manipulation. Voltages
191: applied to the gates G1 to G3 are used to tune the electric field
192: in the vicinity of the IDQD, and thus, the confinement potential
193: asymmetry and inter-dot tunnelling. Hence, an electron initially
194: localized on one quantum dot may be allowed to tunnel to the
195: opposite quantum dot by such manipulation.
196: 
197: An electric field is induced on the SET as a result of this
198: polarization process. This modulates the chemical potential of the
199: SET, and, therefore, the conductance through the source and drain
200: leads under a finite SET bias condition. To maximize the change in
201: conductance, the SET is initially tuned to the charge-sensitive
202: regime by an appropriate voltage bias at G4, and under a small
203: source-drain bias to ensure operation in the linear transport
204: regime. Such manipulation of the device has only been shown
205: through experiment so far. Therefore, a thorough theoretical
206: analysis is necessary to complement the recent experimental
207: findings, and build a more comprehensive understanding of the
208: physical mechanisms involved.
209: 
210: Analytic methods exist for calculating the electrostatic potential
211: in two-dimensional electron gases generated by patterned surface
212: gates on GaAs/AlGaAs heterostructures.\cite{Davies:94,Davies:95}
213: While these analytic methods yield useful results for such
214: devices, they are unsuitable for trench-isolated Si structures,
215: where the geometry is much more sophisticated. Therefore,
216: numerical methods have to be implemented. The finite-element
217: method is a well-suited means for simulation of
218: geometrically-complicated domains,\cite{Cook,Zienkiewicz} and is
219: commonly used to solve Poisson-type
220: equations.\cite{Iserles,DpBook2}
221: 
222: In order to determine the electric field throughout the modelled
223: device regions, the numerical solution to the Laplace equation in
224: three-dimensions is performed:
225: 
226: \begin{eqnarray}\label{eqn:poisson}
227:      \nabla \cdot [ \epsilon (\mathbf{x}) \nabla \phi (\mathbf{x}) ] &=& 0, \qquad \mathbf{x} \in \Omega \in \mathbb{R}^3, \label{eqn:lap1} \\
228:     \phi(\mathbf{x}) &=& D_i, \qquad \mathbf{x} \in \partial \Omega_{D_i}, \label{eqn:lap2}\\
229:      - \epsilon (\mathbf{x}) \frac{\partial \phi }{\partial n} &=&
230:      q,\qquad \mathbf{x} \in \partial \Omega_{N}, \label{eqn:l3}
231: \end{eqnarray}
232: 
233: \noindent where $\phi (\mathbf{x})$ is the electrostatic potential
234: and $\epsilon (\mathbf{x})$ is the material dielectric parameter.
235: The dielectric parameter varies discontinuously on moving through
236: the different materials - air ($\epsilon_0 = 8.85 \cdot 10^{-12}
237: ~$Fm$^{-1}$), Si ($11.0 \epsilon_0$) and SiO$_2$ ($4.5 \epsilon_0
238: $). We apply Dirichlet boundary conditions to the surfaces of the
239: metal gates and to the grounded base of the device, and the
240: Neumann boundary condition to the exposed surfaces. (We apply
241: Neumann boundary conditions with $q=0$, but for generality, we
242: include in our discussion the possibility of non-zero $q$).
243: 
244: The finite element solution of the Laplace equation is well
245: covered in the literature.\cite{Iserles,DpBook2,Mohan} The basic
246: idea of the finite element method is to approximate the unknown
247: fields, for example $\phi$ in the Laplace equation above, by
248: $\widetilde{\phi}$ which is a combination of linearly independent
249: basis functions $N_j$
250: 
251: \begin{equation}\label{eqn:fem4}
252: \widetilde{\phi}(\x) = \sum_{j=1}^M \phi_j N_j(\x) ,
253: \end{equation}
254: 
255: \noindent where $M$ is the number of basis functions and $\phi_j$
256: are the expansion coefficients to be determined. In the finite
257: element method, the computational domain $\Omega$ is divided into
258: a number of elements, and the $N_j$ are chosen to be piecewise
259: polynomials such that they are non-zero only in a `few' adjacent
260: elements.\cite{ft3,Iserles} The method then requires the
261: substitution of $\widetilde{\phi}$ into the Laplace equation, and
262: the residual $R = \nabla \cdot [\epsilon \nabla \widetilde{\phi}
263: ]$ to be orthogonal to the space spanned by a linearly independent
264: set $\{ W_1,\ldots,W_n\}$. In our calculations, we implement the
265: orthogonality through the weighted residual statement:
266: 
267: \begin{eqnarray}
268: \int_{\Omega} R W_i d\Omega = 0 , \qquad i=1,\hdots,M ,
269: \end{eqnarray}
270: 
271: \noindent and Galerkin's method i.e. set $W_i$ equal to the basis
272: functions $N_i$ to obtain
273: 
274: \begin{eqnarray}\label{eqn:fem1}
275:     \int_{\Omega} N_i [
276:     \frac{\partial}{\partial x} \epsilon \frac{\partial \widetilde{\phi}}{\partial
277:     x}
278:     +
279:     \frac{\partial}{\partial y} \epsilon \frac{\partial \widetilde{\phi}}{\partial
280:     y}
281:     +
282:     \frac{\partial}{\partial z} \epsilon \frac{\partial \widetilde{\phi}}{\partial
283:     z}]
284:     d\Omega  = 0 \ep
285: \end{eqnarray}
286: 
287: 
288: %Using Green's Theorem:
289: 
290: %\begin{eqnarray}\label{eqn:green}
291: %- \int_{\Omega}\nabla \cdot [k \nabla u] W_i d \Omega \\
292: %\nonumber = \int_{\Omega}k\nabla u \cdot \nabla W_i d \Omega -
293: %\int_{\partial \Omega} W_i k \frac{\partial u}{\partial n} d\Gamma
294: %,
295: %\end{eqnarray}
296: 
297: Using integration by parts, we reduce the order of derivatives in
298: Eq.~(\ref{eqn:fem1});
299: 
300: \begin{eqnarray}\label{fem2}
301:    -  \int_{\Omega}
302:      \epsilon \frac{\partial \widetilde{\phi}}{\partial x} \frac{\partial N_i}{\partial x}
303:     +
304:      \epsilon \frac{\partial \widetilde{\phi}}{\partial y} \frac{\partial N_i}{\partial y}
305:     +
306:     \epsilon \frac{\partial \widetilde{\phi}}{\partial z} \frac{\partial N_i}{\partial z}
307:     d\Omega
308:     \nonumber \\ \nonumber\\
309:     +\int_{\partial \Omega} N_i \epsilon \frac{\partial \widetilde{\phi}}{ \partial n} d\Gamma
310:     =0 \ep
311: \end{eqnarray}
312: 
313: 
314: The weighted residual method and integration by parts leads to a
315: natural mechanism for the incorporation of derivative boundary
316: condition given by Eq.~(\ref{eqn:l3}) for the Laplace operator
317: $\nabla \cdot [\epsilon \nabla \widetilde{\phi}]$. Hence, we
318: obtain after expanding the approximation for $\widetilde{\phi}$
319: 
320: %{\setlength\arraycolsep{2pt}}
321: \begin{eqnarray}\label{eqn:fem5}
322:       \int_{\Omega}
323:      \epsilon \frac{\partial N_i}{\partial x} \frac{\partial N_j}{\partial x} \phi_j
324:     +
325:      \epsilon \frac{\partial N_i}{\partial y} \frac{\partial N_j}{\partial y} \phi_j
326:     +
327:      \epsilon \frac{\partial N_i}{\partial z} \frac{\partial N_j}{\partial z} \phi_j
328:      d\Omega
329:      \nonumber\\ \nonumber\\
330:     +  \int_{\partial \Omega} N_i q d\Gamma
331:     =0 ,
332: \end{eqnarray}
333: 
334: \noindent where summation is implied over repeated indices. The
335: problem has now been reduced to one of matrix inversion;
336: 
337: \begin{equation}\label{eqn:femMain}
338:     K \Phi = F ,
339: \end{equation}
340: 
341: \noindent where the `stiffness' matrix $K$ is given by
342: 
343: \begin{equation}\label{eqn:femK}
344:     K_{ij} = \int_{\Omega}
345:      \epsilon \frac{\partial N_i}{\partial x} \frac{\partial N_j}{\partial x}
346:     +
347:      \epsilon \frac{\partial N_i}{\partial y} \frac{\partial N_j}{\partial y}
348:     +
349:      \epsilon \frac{\partial N_i}{\partial z} \frac{\partial N_j}{\partial z}
350:      d\Omega ,
351: \end{equation}
352: 
353: \noindent and the RHS vector is given by
354: 
355: \begin{equation}\label{eqn:femF}
356:     F_i=-\int_{\partial \Omega} N_i q d\Gamma ,
357: \end{equation}
358: 
359: \noindent and $\Phi$ is simply the vector of unknowns $\phi_j$.
360: Dirichlet boundary conditions are implemented by forcing
361: prescribed values of $\phi_j$.
362: 
363: %It is often convenient to decompose the integrals given by
364: %Eqs.~(\ref{eqn:femK})-(\ref{eqn:femF}) involving the entire domain
365: %$\Omega$ into integrals over the elements $\Omega_e$. The benefits
366: %of this decomposition are most evident for more complex partial
367: %differential equations and therefore we avoid further elaboration
368: %on this point.
369: 
370: In the formulation and solution of Eq.~(\ref{eqn:femMain}) we have
371: chosen linear basis functions corresponding to 8-noded brick
372: elements. This method therefore has a convergence rate for error
373: of 2.0.\cite{ft4}
374: 
375: We have employed Gaussian quadrature in three-dimensions for the
376: volume integrals and two-dimensions for surface integrals, and a
377: conjugate gradient method with Incomplete Lower and Upper (ILU)
378: factorization preconditioning to solve the resulting linear system
379: of equations. Sufficient resolution was obtained by a mesh with
380: $107\times 77\times 9$ nodes in the $x$, $y$ and $z$ directions,
381: respectively. We implemented adaptive mesh refinement along $z$
382: axis to improve accuracy in the vicinity of the active region.
383: 
384: 
385: \begin{figure}[!h]
386: \begin{center}
387: \rotatebox{0}{\scalebox{0.65}[0.65]{\includegraphics*[1.5in,0.1in][6.0in,6.2in]{f2.eps}}}
388: \end{center}
389: \caption{Two-dimensional slices taken from the full
390: three-dimensional solution of the Laplace equation. (a) A slice
391: from the $x$-$y$ plane at $z=420$~nm, through the center of the
392: gates and IDQD. (b) A slice from the $x$-$z$ plane at $y=190$~nm,
393: through the center of gate G2. The gate G4, is set to $-4.8$~V.
394: The gates G1, G2 and G3 are set to $1$~V, $2$~V and $-1$~V
395: respectively. The position of the IDQD is outlined.}
396: \label{fig:2dsolutions}
397: \end{figure}
398: 
399: 
400: Figures~\ref{fig:2dsolutions}(a) and \ref{fig:2dsolutions}(b) show
401: cross-sectional slices along orthogonal planes of the full
402: three-dimensional solution for the simulated electric potential.
403: The simulation was performed with the following parameters: the
404: gate potentials of G1, G2 and G3 are set to +$1$~V, $2$~V and
405: -$1$~V, respectively; G4 is set to $-4.8$~V. The voltage chosen
406: for G4 is approximately equal to that used for this gate in the
407: experimental demonstrations in Ref.~\onlinecite{Emiroglu:03} where
408: single electron polarization of the IDQD was obtained for such a
409: device. The choices of G1, G2 and G3, are also similar to those in
410: the experiments but for this simulation, the exact values are
411: chosen so that three-dimensional illustrations are as clear as
412: possible.
413: 
414: Figures~\ref{fig:2dsolutions}(a) and \ref{fig:2dsolutions}(b)
415: clearly demonstrate the effect of the applied gate voltages on the
416: potential landscape of the IDQD and the device as a whole; the
417: result of applying a voltage on the in-plane gates is that a
418: significant fraction of the applied voltage is induced on the
419: IDQD, despite the etched trench gap. The abrupt change in the
420: effective permittivity from the metallic gates to the voids, from
421: the voids to the SiO$_2$, and from Si to SiO$_2$, causes some
422: definition of the gates and the IDQD in the plots. The difference
423: between the relative permittivity of air, Si and SiO$_2$ leads to
424: a potential gradient, such that the absolute value of the
425: potential is prone to vanish more rapidly in air, compared with Si
426: and SiO$_2$. However, Fig.~\ref{fig:2dsolutions}(b) clearly
427: demonstrates that for this particular pillar height, which matches
428: the device used in experiment, the potential at the IDQD is due
429: mainly to the electric field vectors that are on a direct path
430: through the trench isolation, and not the underlying substrate.
431: This is consistent with experimental observations and is the
432: preferred mechanism of device operation, since it is relatively
433: easier in design and theoretical analysis, compared with the case
434: where the majority of the electric field is through the
435: semiconductor base and the field lines arrive at the IDQD from
436: several different paths.
437: 
438: \begin{figure}
439: 
440: \epsfxsize=9.0cm
441: 
442: \centerline{\epsffile{f3.eps}}
443: 
444: \caption{Cross-sectional curves through the IDQD with $x=320$~nm
445: and $z=420$~nm (the dots of each curve correspond to nodal points
446: of the computational grid). The gate G4 is set to -$4.8$~V. (a)
447: The electrostatic potential of the voltage on gate G2 is changed
448: from -$5$~V to +$5$~V. (b) The voltage applied to gate G1 is
449: varied from $0$~V to -$5$~V. The voltage applied to gate G3 is set
450: to the negative of that applied to gate G1 (in order to maximize
451: the electric field). } \label{fig:xscurves1}
452: 
453: \end{figure}
454: 
455: For a more quantitative analysis, we determine the electric
456: potential along the active region of the IDQD as a function of the
457: applied gate voltages. This is shown in
458: Figs.~\ref{fig:xscurves1}(a) and \ref{fig:xscurves1}(b), where
459: different gates are used to apply the in-plane electric field.
460: Figure~\ref{fig:xscurves1}(a) shows that the effect of varying the
461: voltage applied to gate G2, from -$5$~V to +$5$~V, is to induce a
462: voltage at the IDQD from -$2$~V to +$0.8$~V, respectively. Note
463: that G4 had lowered the overall potential by approximately
464: $0.55$~V in this case.
465: 
466: The data in Fig.~\ref{fig:xscurves1}(a) shows a maximum change of
467: $0.3$~V of the electrostatic potential at the IDQD, when the G2
468: gate voltage is raised or lowered by $1$~V. This field coupling
469: factor of $\sim$$30~\%$ is approximately one order of magnitude
470: greater than what was observed in experiment.\cite{Emiroglu:03}
471: However, the measured quantity in experiments is the SET current,
472: and the IDQD coupling terms are inferred from such measurements.
473: The task of calculating such coefficients exactly as measured in
474: experiment is beyond the scope of a purely electrostatic model,
475: since, with the SET present, the global system that must be
476: treated consists of interacting sub-systems of quantum
477: mechanically bound electrons. Therefore, we project that a
478: self-consistent Schr\"{o}dinger-Poisson analysis of the system
479: would yield results for the coupling coefficients that are closer
480: to actual values observed in experiment.
481: 
482: Our results are also consistent with the experimental
483: demonstrations of Ref.~\onlinecite{Emiroglu:03} where the voltage
484: on gate G2 is swept continuously from $-5$~V to $+5$~V, with G4
485: set to $\sim$$-4.8$~V, in order to demonstrate conductance
486: resonances of the SET currents but also resonances due to the
487: single electron polarization of the IDQD. (The device in
488: Ref.~\onlinecite{Emiroglu:03} had a slight asymmetry in the
489: alignment of the IDQD relative to G2, hence the ability of G2 to
490: polarize the IDQD.)
491: 
492: The abrupt changes in the potential at $115$~nm and $265$~nm are
493: due to the change of relative permittivity at the air-SiO$_2$
494: interface. The difference between the potential gradient in the
495: air and in the semiconductor regions is more evident in these
496: figures. Figure~\ref{fig:xscurves1}(b) shows the effect of
497: applying voltages of opposite sign to gates G1 and G3. This
498: results in a potential gradient across the IDQD, which has a
499: maximum value of $\sim$$0.007$~Vnm$^{-1}$ in these simulations.
500: This clearly demonstrates an effective mechanism for externally
501: tuning the internal potential asymmetry of the IDQD electronic
502: states.
503: 
504: In the experimental demonstrations of Ref.~\onlinecite{Gorman:05},
505: a voltage bias is pulsed across in-plane metallic gates, which are
506: placed perpendicularly to an IDQD as in our case, and this was
507: shown to result in the coherent oscillation of a single electron
508: charge present in an IDQD. This is again consistent with our
509: results which suggests a strong electric field is induced at the
510: IDQD due to the electric field at the in-plane gates.
511: 
512: In summary, we have successfully demonstrated, by means of
513: finite-element solutions to the Laplace equation, that the
514: electric potential and potential gradient across the confining
515: region of the IDQD in trench-isolated Si devices may be
516: manipulated effectively by the voltages applied to
517: capacitively-coupled in-plane gates. Our calculations show good
518: correlation with recent experimental demonstrations, where the
519: IDQD electron states are manipulated by such methods.
520: 
521: We thank S. Pfaendler and R. Schumann for comments and useful
522: discussions. SR, JG, and CB acknowledge the support of the EPSRC
523: through the QIP IRC. SR acknowledges the Cambridge-MIT Institute
524: for financial support.
525: 
526: \begin{references}
527: 
528: \bibitem{Emiroglu:03}   E. G. Emiroglu, D. G. Hasko
529: and D. A. Williams, Appl.\ Phys.\ Lett.\ {\bf 83}, 3942-3044
530: (2003).
531: 
532: \bibitem{Gorman:05}   J. Gorman, D. G. Hasko
533: and D. A. Williams,  Phys.\ Rev.\ Lett.\ {\bf 95}, 090502 (2005).
534: 
535: \bibitem{Loss:98} D. Loss and D. P. DiVincenzo, Phys.\
536: Rev.\ A {\bf 57}, 120-126 (1998).
537: 
538: \bibitem{NielsonBook} M. A. Nielsen and I. L. Chuang, Quantum Computation and Quantum Information,  {\it Cambridge University Press},
539: (2003).
540: 
541: \bibitem{ft1} The most significant interaction is that between the SET and the IDQD
542: because it directly influences the read-out operation and the
543: degree of back-action.
544: 
545: \bibitem{ft2} The device elements are in fact roughly rectangular, so this
546: approximation introduces only a small error to the results
547: presented.
548: 
549: \bibitem{vanDerWiel:02} W. G. van der Wiel, S. De
550: Francheshi, J. M. Elzerman, T. Fujisawa, S. Tarucha and L. P.
551: Kouwenhoven, Reviews of Modern Physics, {\bf 75}, 1-22, (2003).
552: 
553: \bibitem{Fujisawa:98} T. Fujisawa, T. H. Oosterkamp, W. G. van der Wiel, B. W. Broer, R. Aguado, S. Tarucha and L. P. Kouwenhoven, Science, {\bf 282},
554: 932-935, (1998).
555: 
556: \bibitem{Davies:95}    J. H. Davies, I. A. Larkin and E.
557: V. Sukhorukov J.\ Appl.\ Phys. {\bf 77}, 4504-4512 (1995).
558: 
559: \bibitem{Davies:94} J. H. Davies and I. A. Larkin, Phys.\ Rev.\ B {\bf 49(7)},
560: 4800-4809 (1994).
561: 
562: %\bibitem{Laux:88}  S. E. Laux and D. J. Frank and F.
563: %Stern, Surface Science {\bf 196}, 101-106 (1988).
564: 
565: \bibitem{Cook} R. D. Cook, D. S. Malkus and M. E. Plesha, {\it Concepts and Applications of Finite Element Analysis}, Wiley,
566: (1989).
567: 
568: \bibitem{Zienkiewicz} O. Zienkiewicz, {\it The Finite Element
569: Method}, McGraw-Hill, 4th edition (1994).
570: 
571: %\bibitem{Bathe} K. J. Bathe, {\it Finite Element Procedures}, Prentice Hall,
572: %(1996).
573: 
574: \bibitem{Iserles}A. Iserles, {\it A first course in the numerical analysis of differential equations}, Cambridge University Press,
575: (1995).
576: 
577: \bibitem{DpBook2} H. P. Langtangen, {\it Computational Partial
578: Differential Equations - Numerical methods and Diffpack
579: Programming}, Springer, 2nd edition (2003).
580: 
581: \bibitem{Mohan} L. Ramdas Ram-Mohan, {\it Finite Element and Boundary Element Applications in Quantum Mechanics}, Oxford University
582: Press, 1st edition, (2002).
583: 
584: \bibitem{ft3} The choice of piecewise polynomials results in a speed-up over spectral methods with little increase in error.
585: 
586: \bibitem{ft4} Higher-order basis functions were also implemented but had
587: not resulted in any significant improvement in the accuracy of the
588: results presented.
589: 
590: 
591: 
592: \end{references}
593: 
594: \end{document}
595: 
596: \begin{thebibliography}{99}
597: %\bibliographystyle{plain}
598: