physics0205060/ms.tex
1: % Template article for preprint document class `elsart'
2: % SP 2001/01/05
3: 
4: % Use the option doublespacing or reviewcopy to obtain double line spacing
5: %\documentclass[doublespacing]{elsart}
6: \documentclass{elsart}
7: 
8: % The amssymb package provides various useful mathematical symbols
9: %\usepackage{amssymb}
10: \usepackage{amsmath}
11: \usepackage{graphicx}
12: 
13: \graphicspath{{img/}}
14: 
15: \makeatletter
16: \def\@date{22 May 2002}
17: \makeatother
18: 
19: \renewcommand{\l}{\ell}
20: \renewcommand{\L}{L}%\mathcal{L}}
21: 
22: \begin{document}
23: 
24: \begin{frontmatter}
25: \title{Laser controlled molecular switches and transistors}
26: \author{J\"org Lehmann},
27: \author{S\'ebastien Camalet},
28: \author{Sigmund Kohler}, and
29: \author{Peter H\"anggi\corauthref{cor1}}
30: \corauth[cor1]{Corresponding author. Tel.: +49-821-598-3250; 
31: fax: +49-821-598-3222; e-mail: Peter.Hanggi@physik.uni-augsburg.de}
32: \address{Institut f\"ur Physik, Universit\"at Augsburg, 
33:          Universit\"atsstra{\ss}e~1, D-86135 Augsburg, Germany}
34: 
35: \begin{abstract}
36: We investigate the possibility of optical current control through single molecules
37: which are weakly coupled to leads.  A master equation approach for the transport
38: through a molecule is combined with a Floquet theory for the time-dependent
39: molecule.  This yields an efficient numerical approach to the evaluation of
40: the current through time-dependent nano-structures in the presence of a finite
41: external voltage.  We propose tunable optical current switching in two- and
42: three-terminal molecular electronic devices driven by properly adjusted laser
43: fields, \textit{i.e.} a novel class of molecular transistors.
44: 
45: \end{abstract}
46: 
47: \begin{keyword}
48:   molecular electronics \sep quantum control 
49:   \PACS 85.65.+h \sep 33.80.-b \sep 73.63.-b \sep 05.60.Gg
50: \end{keyword}
51: \end{frontmatter}
52: 
53: \section{Introduction}
54: 
55: Spurred by the ongoing experimental progress in the field of molecular
56: electronics \cite{Reed1997a,Cui2001a,Joachim2000a,Schon2001a,Reichert2002a},
57: the theoretical interest in transport properties of molecules has revived
58: \cite{Nitzan2001a}.  Tight-binding models for the wire have been used to
59: compute current-voltage characteristics, within a scattering approach
60: \cite{Mujica1994a} and from electron transfer theory \cite{Nitzan2001a}.  Both
61: approaches bear the same essentials \cite{Nitzan2001b}.  For high temperatures,
62: the wire electrons loose their quantum coherence and the transport is dominated
63: by incoherent hopping between neighbouring sites \cite{Petrov2001a}.  Recently,
64: the current-voltage characteristics has been obtained from a quantum-chemical
65: \textit{ab initio} description of the molecule \cite{Heurich2001a}.  The
66: results were in good agreement with recent experiments \cite{Reichert2002a}.
67: 
68: Typical electronic excitation energies in molecules are in the range up to an
69: eV and, thus, correspond to light quanta from the optical and the infrared
70: spectral regime where most of today's lasers work.  It is therefore natural to
71: use such coherent light sources to excite molecules and to study their
72: influence on the transport properties aiming to find a way of manipulating
73: currents.  One particularly prominent example of quantum control is the
74: so-called coherent destruction of tunnelling (CDT), \textit{i.e.} the
75: suppression of the tunnelling dynamics in an \textit{isolated} bistable
76: potential by the purely coherent influence of an oscillating bias
77: \cite{Grossmann1991a,Grossmann1992a,Morillo1993a,Goychuk1996a,Grifoni1998a}.
78: The crucial point there is that the long-time dynamics in a periodically driven
79: quantum system is no longer dominated by the energies, but rather by the
80: so-called quasienergies \cite{Grifoni1998a,Shirley1965a,Sambe1973a}.  The
81: latter may be degenerate for properly chosen driving parameters yielding a
82: divergent time-scale.  Inspired by these results, we address in this Letter the
83: question of controlling by use of properly tailored laser fields the transport
84: through time-dependent \textit{open} systems, \textit{i.e.} systems that allow
85: for a particle exchange with external leads.
86: 
87: For the computation of electrical currents through wires exposed to strong
88: laser fields, we put forward Floquet approach \cite{Lehmann2002b}.  The central
89: idea of this method lies in a non-perturbative solution of the Schr\"odinger
90: equation of the isolated time-dependent wire plus laser field, while the
91: wire-lead coupling is treated perturbatively.  The resulting density operator
92: equation is decomposed into a time-dependent Floquet basis permitting a
93: numerically efficient treatment.  We generalise here this method to the
94: analysis of networks with an arbitrary number of contacts to leads.
95: Subsequently we apply the so obtained formalism to the investigation of optical
96: current switching in two- and three-terminal devices as prototypical examples
97: of a new class of molecular transistors.
98: 
99: \section{Model for wire and leads}
100: 
101: %%%% Figure 1 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
102: \begin{figure}
103:   \begin{center}
104:     \includegraphics[width=6truecm]{fig1.eps}
105:   \end{center}
106:   \caption{Molecular circuit consisting of $N=6$ sites of which the
107:     sites $1,\ldots,L$ are coupled to $\L=4$ leads.}
108:   \label{fig:system}
109: \end{figure}
110: %%%% Figure 1 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
111: 
112: We embark by specifying the model Hamiltonian of the entire system as
113: sketched in Fig.~\ref{fig:system}.  It consists of the molecule in the
114: laser field, ideal leads, and the molecule-leads coupling Hamiltonian,
115: \begin{equation}
116: H(t)=H_{\text{molecule}}(t) + H_{\text{leads}} +
117: H_{\text{molecule-leads}}\ .
118: \end{equation}
119: The irradiated molecule is modelled by a tight-binding description taking into
120: account $N$ molecular orbitals $|n\rangle$, which are relevant for the
121: transport.  Disregarding the electron-electron interaction, the most general
122: form of the Hamiltonian reads
123: \begin{equation}
124: \label{eq:Hmolecule} H_{\text{molecule}}(t)=\sum_{n,n'} H_{nn'}(t)\,
125: c_n^\dagger c_{n'},
126: \end{equation}
127: where the fermionic operators $c_n$ and $c_n^\dagger$ destroy and create,
128: respectively, an electron in the molecular orbital $|n\rangle$. The sums
129: extend over all tight-binding orbitals.  The $\mathcal{T}$-periodic
130: time-dependence of the single-particle Hamiltonian
131: $H_{nn'}(t)=H_{nn'}(t+\mathcal{T})$, reflects the influence of the laser field
132: with frequency $\Omega=2\pi/\mathcal{T}$.
133: %
134: The $L$ ideal leads are described by grand-canonical ensembles of electrons
135: at temperature $T$ and electro-chemical potential $\mu_\l$, $\l=1,\dots,L$.
136: Thus, the lead Hamiltonian reads
137: $H_\mathrm{leads} = \sum_{q\l} \epsilon_{q\l} \, c_{q\l}^\dagger c_{q\l}$,
138: where $c_{q\l}$ destroys an electron in state $q$ in lead $\l$.
139: All expectation values of lead operators can be traced back to
140: $\langle c_{q\l}^\dagger c_{q'\l'} \rangle = \delta_{qq'}
141: \delta_{\l\l'} f(\epsilon_{q\l}-\mu_\l)$, where $f(\epsilon)
142: =(1+\e^{\epsilon/k_B T})^{-1}$ denotes the Fermi function.
143: The model is completed by the molecule-leads tunnelling Hamiltonian
144: \begin{equation}
145: H_{\text{molecule-leads}}
146: =\sum_{q\l} V_{q\l} \, c_{q\l}^\dagger \, c_{\l} + \mathrm{h.c.}\ ,
147: \end{equation}
148: that connects each lead directly to one of the suitably labelled molecular
149: orbitals.  Since we are not interested here in the effects that arise from the
150: microscopic details of the molecule-lead coupling, we restrict our analysis in
151: the following to energy-independent couplings, \textit{i.e.} $\Gamma_{\l} = (2\pi/\hbar)
152: \sum_q |V_{q\l}|^2 \, \delta(\epsilon - \epsilon_{q\l})$.
153: 
154: 
155: \section{Perturbative description and Floquet ansatz}
156: 
157: Let us assume that the dynamics of the driven wire is dominated by the
158: time-dependent wire Hamiltonian so that the coupling to the leads can be taken
159: into account as a perturbation.  This allows to derive by use of standard
160: methods the approximate equation of motion for the total density operator
161: $\varrho(t)$, 
162: \begin{align}
163: \label{eq:master}
164: \dot\varrho(t) =
165: & -\frac{i}{\hbar}[H_{\rm molecule}(t)+H_{\rm leads},\varrho(t)] \\
166: & -\frac{1}{\hbar^2}\int\limits_0^\infty \!\mathrm{d}\tau
167:   [H_\mathrm{molecule-leads},[\widetilde{H}_\mathrm{molecule-leads}(t-\tau,t),
168:   \varrho(t)]] . \nonumber
169: \end{align}
170: We have omitted a transient term that depends purely on the initial preparation.
171: The tilde denotes operators in the interaction picture with respect to the
172: molecule plus the lead Hamiltonian, $\widetilde{O}(t,t')=U_0^\dagger(t,t') O(t)
173: U_0(t,t')$, where $U_0(t,t')$ is the time-evolution operator without the
174: coupling.  The first term describes the coherent dynamics of the electrons on
175: the wire, while the second term represents incoherent hopping of electrons
176: between the leads and the wire.
177: 
178: The net (incoming minus outgoing) electrical current that flows from lead $\l$
179: into the molecule is then given by the rate at which the electron number in the
180: corresponding lead decreases multiplied by the electron charge $-e$,
181: \begin{equation}
182:   \label{eq:current}
183:   I_\l(t) = e\, \frac{\mathrm{d}}{\mathrm{d}t}\langle N_\l\rangle\ .
184: \end{equation}
185: Note that this expectation value is time-dependent through the non-equilibrium
186: density operator $\varrho(t)$.  To evaluate the right-hand side of
187: Eq.~(\ref{eq:current}), we employ Eq.~\eqref{eq:master} to derive after some
188: algebra the result
189: \begin{equation}
190: \label{eq:current_general}
191: \begin{split}
192: I_\l(t)  = 
193:  -e   \Gamma_{\l} \bigg\{ &\mathop{\rm Re} 
194: \int\limits_0^\infty \frac{\mathrm{d}\tau}{\hbar} \! 
195: \int  \!\frac{\mathrm{d}\epsilon}{\pi}\,
196: e^{i\epsilon\tau/\hbar}          
197: f(\epsilon-\mu_\l)\big\langle[c_\l,
198: \tilde c_{\l}^\dagger(t-\tau,t)]_+\big\rangle \\
199:  &- \big\langle c_\l^\dagger
200:  c_{\l}\big\rangle\bigg\}\ .
201: \end{split}
202: \end{equation}
203: This expression still contains the yet unknown expectation values of solely
204: those wire operators with a direct connection to lead $\l$.  It depends in
205: particular on the Heisenberg operators $\tilde c_\l^\dagger(t-\tau,t)$ and thus
206: implicitly on the dynamics of the driven wire.  Let us therefore focus on the
207: single-particle dynamics of the periodically time-dependent wire Hamiltonian
208: $H_{nn'}(t)$.  An established procedure for the solution of the corresponding
209: Schr\"odinger equation is to employ a Floquet ansatz amounting to a
210: non-perturbative treatment of the external driving.  There one uses the fact
211: that a complete set of solutions is of the form
212: $|\Psi_\alpha(t)\rangle=\exp(-i\epsilon_\alpha t/\hbar)|\Phi_\alpha(t) \rangle$.
213: The so-called quasienergies $\epsilon_\alpha$ take over the role of the
214: energy eigenvalues in static systems and govern the long-time dynamics.  The Floquet
215: modes $|\Phi_\alpha(t)\rangle$ obey the time-periodicity of the driving field
216: which allows to express them as a Fourier series, $|\Phi_\alpha(t)\rangle =
217: \sum_{k=-\infty}^{\infty} \exp(-ik\Omega t) |\Phi_{\alpha,k}\rangle$.  They can
218: be obtained from the eigenvalue equation
219: \cite{Shirley1965a,Sambe1973a,Grifoni1998a}
220: \begin{equation}
221: \label{floquet_hamiltonian}
222: \Big(\sum_{n,n'}|n\rangle H_{nn'}(t) \langle n'|
223: -i\hbar\frac{\mathrm{d}}{\mathrm{d}t}\Big)
224: |\Phi_{\alpha}(t)\rangle = \epsilon_\alpha |\Phi_{\alpha}(t)\rangle
225: \ .
226: \end{equation}
227: Moreover, the Floquet modes define the complete set of operators
228: \begin{equation}
229: \label{eq:c_alpha}  
230: c_\alpha(t) = \sum_n \langle\Phi_\alpha(t)|n\rangle\, c_n  ,
231: \end{equation}
232: whose time-evolution assumes the convenient form $\tilde{c}_\alpha(t-\tau,t) =
233: \exp(i\epsilon_\alpha\tau/\hbar) c_\alpha(t)$.  The orthogonality of the Floquet
234: states at equal times \cite{Shirley1965a,Sambe1973a,Grifoni1998a} yields the
235: back-transformation $c_n=\sum_\alpha\langle n|\Phi_\alpha(t)\rangle$ and thus
236: results in the required spectral decomposition.
237: Using (\ref{eq:c_alpha}) and performing the energy and the $\tau$-integration
238: in Eq.~(\ref{eq:current_general}), we obtain for the time-averaged current
239: the main result
240: \begin{equation}
241: \label{eq:meancurrent}
242: \begin{split}
243: \bar{I}_\l = 
244: -\frac{e\Gamma_\l}{\hbar}\sum_{\alpha k}
245: \Big[&
246: \langle \Phi_{\alpha, k}|\l\rangle \langle \l|\Phi_{\alpha, k}\rangle 
247:   f(\epsilon_\alpha+k\hbar\Omega-\mu_\l) 
248: \\ &
249: -\sum_{\beta k'}
250:  \langle \Phi_{\alpha, k'+k} |\l\rangle\langle \l|\Phi_{\beta k'} \rangle
251:  R_{\alpha\beta,k}
252: \Big] .
253: \end{split}
254: \end{equation}
255: Here, we have introduced the expectation values $R_{\alpha\beta}(t)= \langle
256: c^\dagger_\alpha(t)c_{\beta}(t)\rangle$, which assume in the long-time limit the
257: time-periodicity of the driven system and thus can be decomposed into the
258: Fourier series $R_{\alpha\beta}(t)= \sum_k \exp(-ik\Omega t) R_{\alpha\beta,k}$.
259: It is straightforward to derive for the $R_{\alpha\beta,k}$ from the
260: density operator equation~(\ref{eq:master}) the following set of inhomogeneous linear
261: equations
262: \begin{align}
263: \label{eq:mastereq_fourier}
264: \lefteqn{\frac{i}{\hbar}(\epsilon_\alpha-\epsilon_\beta+k \hbar \Omega)R_{\alpha\beta,k}}
265: \\ && =
266:  \frac{1}{2}\sum_{\l k'} \Gamma_\l
267:     \Big\{ &
268:     \sum_{\beta'k''}
269:                 \langle\Phi_{\beta,k''+k'}|\l\rangle
270:                 \langle \l|\Phi_{\beta',k''+k}\rangle
271:                 R_{\alpha\beta',k'}
272: \nonumber \\&& {}+{} &
273:     \sum_{\alpha'k''}
274:                 \langle\Phi_{\alpha',k''+k'}|\l\rangle
275:                 \langle \l|\Phi_{\alpha,k''+k}\rangle
276:                 R_{\alpha'\beta,k'}
277: \nonumber \\ && {}-{} &
278:     \langle\Phi_{\beta,k'-k}|\l\rangle
279:     \langle \l|\Phi_{\alpha,k'}\rangle
280:     f(\epsilon_\alpha+k'\Omega-\mu_\l)
281: \nonumber \\ && {}-{} &
282:     \langle\Phi_{\beta,k'}|\l\rangle
283:     \langle \l|\Phi_{\alpha,k'+k}\rangle
284:     f(\epsilon_\beta+k'\Omega-\mu_\l)
285:    \Big\}
286: \nonumber\ ,
287: \end{align}
288: which will be solved numerically.  We have found that even in the case of
289: strong driving where the Floquet states comprise many sidebands, a few Fourier
290: coefficients $R_{\alpha\beta,k}$ are in fact sufficient to obtain numerical
291: convergence.  This justifies \textit{a posteriori} the use
292: of the Floquet states as a basis set.
293: 
294: To conclude the technical part of this work, we note that our approach goes
295: beyond a linear response treatment of the driving and additionally does not use
296: a so-called rotating wave approximation (RWA) \cite{Blumel1991a,Bruder1994a},
297: which neglects the oscillatory contributions to $R_{\alpha\beta}(t)$ by the
298: ansatz $R_{\alpha\beta,k}=P_\alpha\,\delta_{\alpha,\beta}\,\delta_{k,0}$.  In
299: fact, we found that a RWA solution delivers inaccurate results in the vicinity
300: of quasienergy degeneracies.
301: 
302: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
303: \section{Optical current gate}
304: 
305: As a first setup that may be suitable as a current control device, we
306: investigate the transport through a two-level system, \textit{i.e.} a wire that consists
307: of $N=2$ sites --- one of them is coupled to the left lead and the other to
308: the right lead.  Then the time-dependent wire Hamiltonian reads in the basis of
309: the molecular orbitals
310: \begin{equation}
311:   \label{eq:twosites}
312:     H_\mathrm{molecule}(t) = 
313:     \begin{pmatrix}
314:       E_\mathrm{L} && -\Delta\\
315:       -\Delta      && E_\mathrm{R}
316:     \end{pmatrix}
317:     + A 
318:     \begin{pmatrix}
319:       1 && 0\\
320:       0 && -1
321:     \end{pmatrix}\cos(\Omega t)\ ,
322: \end{equation}
323: where $\Delta$ denotes the tunnel matrix element between the two sites and
324: $E_\mathrm{L}$ and $E_\mathrm{R}$ are the corresponding on-site energies.  The
325: laser field contributes to the Hamiltonian (\ref{eq:twosites}) a time-dependent
326: bias with amplitude $A=-e\mathcal{E}d$, \textit{i.e.} charge times electrical
327: field strength times the site-to-site distance.  Note that the electrical field
328: may be drastically enhanced due to the presence of the metallic tips
329: \cite{Demming1998a}.  The effective coupling to each lead is assumed to be
330: equal, $\Gamma_{\l} = \Gamma$, and an external voltage $V$ is taken into account
331: by a difference in the electro-chemical potentials,
332: $\mu_\mathrm{L}-\mu_\mathrm{R}=-eV$.
333: 
334: We use in all numerical calculations the tunnel matrix element $\Delta$ as the
335: energy unit and assume that the effective couplings to the leads are by one order of
336: magnitude smaller, $\hbar\Gamma=0.1\Delta$. This corresponds to
337: a large contact resistance and ensures the applicability of a perturbational
338: approach.  A realistic value is $\Delta=0.1\mathrm{eV}$, resulting in a current
339: unit $e\Gamma=0.256\,\mathrm{\mu A}$.  For a site-to-site distance of $10${\AA}
340: and a laser frequency $\Omega=10\Delta/\hbar$, the driving amplitudes
341: considered below correspond to an electrical field amplitude of
342: $10^6\,\mathrm{V}/\mathrm{cm}$ at $1\,\mu\mathrm{m}$ wavelength.
343: 
344: %%%% Figure 2 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
345: \begin{figure}
346:   \begin{center}
347:     \includegraphics[width=7.5truecm]{fig2.eps}
348:   \end{center}
349:   \caption{Average current and quasienergy spectrum versus driving amplitude
350:     for a wire which consists of two sites between two electrodes (cf.\ inset)
351:     for unbiased ($E_\mathrm{R}=E_\mathrm{L}=0$, solid lines) and
352:     biased ($E_\mathrm{R}=-E_\mathrm{L}=0.1\Delta$, dashed lines)
353:     on-site energies.
354:     The leads' chemical potentials are $\mu_\mathrm{R}=-\mu_\mathrm{L}=10 \Delta$;
355:     the other parameters read $\hbar\Omega=10\Delta$, $k_B T=0.25\Delta$, 
356:     $\hbar\Gamma=0.1\Delta$.
357:     }
358:   \label{fig:twolevel}
359: \end{figure}%
360: %%%% Figure 2 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
361: The time-averaged current $\bar{I}=\bar{I}_\mathrm{L}=-\bar{I}_\mathrm{R}$
362: through the molecule in a case where both on-site energies are equal is
363: depicted in Fig.~\ref{fig:twolevel}.  As a striking feature, we find that at
364: certain values of the driving amplitude, the current collapses to less than 1\%
365: of its maximal value reached in the absence of the driving.  Closer inspection
366: (not shown) reveals that the width of this depression is proportional to the
367: molecule-lead coupling $\Gamma$.  Comparison with the quasienergy spectrum in
368: the lower panel demonstrates that the current break downs occur at quasienergy
369: crossings.  This relates the present phenomenon to the CDT, \textit{i.e.} the
370: standstill of the tunnel dynamics in a driven bistable potential at quasienergy
371: crossings \cite{Grossmann1991a}.  For the \textit{isolated} two-level system
372: (\ref{eq:twosites}) with $\Delta\ll\Omega,A$, the CDT condition reads
373: $\mathrm{J}_0(2A/\hbar\Omega)=0$ \cite{Grossmann1992a}, \textit{i.e.} the suppression
374: of the tunnelling dynamics is related to the zeros of the Bessel function
375: $\mathrm{J}_0$.  As our analysis reveals, the same condition results in a
376: suppression of the transport through the \textit{open} system.
377: 
378: An external voltage may be of peculiar influence to the on-site energies of a
379: molecular wire \cite{Mujica2000a} and cause an effective bias $E_\mathrm{L}\neq
380: E_\mathrm{R}$ in originally symmetric molecules.  Thus, a crucial question is
381: whether the above current suppression is stable against such a modification.
382: The broken lines in Fig.~\ref{fig:twolevel} demonstrate that this is indeed the
383: case.  Although the quasienergies now form an avoided crossing, the current
384: breakdowns do survive; they are even more pronounced, but slightly shifted
385: towards larger driving amplitudes.  This robustness of CDT based current
386: control combined with the huge on/off ratio suggests the presented setup as a
387: promising alternative to structural chemistry-based switching devices
388: \cite{Chen1999a,Collier2000a}.
389: 
390: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
391: \section{Optical current router}
392: 
393: An experimentally more ambitious configuration consists in a planar molecule
394: with $N=4$ sites, three of which are coupled to a central site and are
395: directly connected to leads (cf.\ inset of Fig.~\ref{fig:switch1}).  We borrow
396: from electrical engineering the designation $\mathrm{E}$, $\mathrm{C}_1$, and
397: $\mathrm{C}_2$.  Here, an external voltage is always applied such that
398: $\mathrm{C}_1$ and $\mathrm{C}_2$ have equal electro-chemical potential,
399: \textit{i.e.} $\mu_{\mathrm{C}_1}=\mu_{\mathrm{C}_2}\neq\mu_\mathrm{E}$.  In a perfectly
400: symmetric molecule, where all on-site energies equal each other, reflection
401: symmetry at the horizontal axis ensures that any current which enters at
402: $\mathrm{E}$, is equally distributed among $\mathrm{C}_{1,2}$, thus
403: $I_{\mathrm{C}_1}=I_{\mathrm{C}_2}=-I_\mathrm{E}/2$.
404: %
405: Since this structure in Fig.~\ref{fig:switch1} is essentially two-dimensional,
406: we have to take also the polarisation of the laser field into account.  We
407: assume it to be linear with an polarisation angle $\phi$ as sketched in the
408: inset of Fig.~\ref{fig:switch1}.  The effective driving amplitudes of the
409: orbitals which are attached to the leads acquire now a geometric factor which
410: is only the same for both orbitals $\mathrm{C}_1$ and $\mathrm{C}_2$ when
411: $\phi=0$.  For any other polarisation angle, the mentioned symmetry is broken
412: and the outgoing currents may be different from each other.  The difference may
413: be huge, as depicted in Fig.~\ref{fig:switch1}.  Their ratio varies from unity
414: for $\phi=0$ up to the order of 100 for $\phi=60^\circ$.  Thus, adapting the
415: polarisation angle enables one to route the current towards the one or the other
416: drain.
417: %%%% Figure 3 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
418: \begin{figure}
419:   \begin{center}
420:     \includegraphics[width=7.5truecm]{fig3.eps}
421:   \end{center}
422:   \caption{Average currents through contacts $\mathrm{C}_1$ (solid)
423:     and $\mathrm{C}_2$ (broken) as a function of the polarisation angle $\phi$
424:     for the three-terminal device depicted in the inset.  The chemical
425:     potentials are $\mu_\mathrm{E}=-\mu_{\mathrm{C}_1} =-\mu_{\mathrm{C}_2}=50
426:     \Delta$; the on-site energies $E_n=0$.  The driving field is specified by
427:     the strength $A=25\Delta$ and the angular frequency $\Omega=10\Delta/\hbar$;
428:     the effective coupling is $\hbar\Gamma=0.1\Delta$ and the temperature $k_B
429:     T=0.25\Delta$.  The maximal value of the current ratio
430:     $I_{\mathrm{C}_1}/I_{\mathrm{C}_2}\approx 100$ is assumed at $\phi=60^\circ$. }
431:   \label{fig:switch1}
432: \end{figure}%
433: %%%% Figure 3 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
434: 
435: Alternatively, one can keep the polarisation angle at $\phi=0$ and break the 
436: reflection symmetry by using an intrinsically asymmetric molecule, as
437: sketched in the inset of Fig.~\ref{fig:switch2}. This allows to 
438: control sensitively the ratio of the outgoing currents by the strength $A$ of the
439: external field, cf.\ Fig.~\ref{fig:switch2}.  The switching range comprises
440: up to four orders of magnitude with an exponential sensitivity.
441: %%%% Figure 4 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
442: \begin{figure}
443:   \begin{center}
444:     \includegraphics[width=7.5truecm]{fig4.eps}
445:   \end{center}
446:   \caption{Ratio of the outgoing average currents versus driving strength $A$
447:     for the three-terminal device at a polarisation angle $\phi=0$.
448:     The filled circle in the inset depicts a site with an on-site energy
449:     $E_{\mathrm{C}_1}$ that differs from the others.
450:     All other on-site energies and parameters as in Fig.~\ref{fig:switch1}.
451:     }
452:   \label{fig:switch2}
453: \end{figure}
454: %%%% Figure 4 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
455: 
456: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
457: \section{Concluding remarks}
458: 
459: We have presented a method for the efficient numerical computation of currents
460: through periodically time-dependent networks with two or more contacts to
461: external leads.  The application to two types of setups substantiated that
462: external fields bear a wealth of possibilities for the manipulation of
463: electrical currents: in a molecular wire the current can be suppressed by
464: proper time-dependent fields.  In a three terminal device, it is possible to
465: route by tailored optical fields the current that enters from a source towards
466: the one or the other drain.
467: 
468: The authors hope that their proposals will motivate experimentalists to accept
469: the challenge of implementing the proposed molecular wire schemes in the
470: laboratory.  The two-terminal current gate can possibly be realized using break
471: junctions exposed to a laser field.  Alternatively, one could use a
472: self-assembled, laser-irradiated maze-like layer of sparsely distributed
473: conducting molecules on a conducting surface. Then by positioning a scanning
474: tunnelling microscope tip directly over one such molecule, it should be
475: possible to measure the features of the predicted gating behaviour.
476: Experimentally more ambitious is the realization of the arrangement in
477: Fig.~\ref{fig:switch1} with a planar molecule contacted to three electrodes.
478: Here again, laser-irradiated self-assemblies of molecules such as carbon
479: nanotube complexes or of biomolecules like metalised DNA \cite{Braun1998a}, or
480: the use of cationic lipid-DNA complexes \cite{Raedler1998a} as DNA-nanocables,
481: with the centre-molecule covalently bound to such planar structures, might make
482: the experiment feasible.
483: 
484: A completely different realization of our findings should be possible in
485: semiconductor heterostructures. There, instead of a molecule, coherently
486: coupled quantum dots \cite{Blick1996a} form the central system. Furthermore,
487: owing to the lower level spacings, the suitable frequency of the coherent
488: radiation source is then in the microwave spectral range.
489: 
490: 
491: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
492: \ack
493: We appreciate helpful discussions with Igor Goychuk and Gert-Ludwig Ingold.
494: This work has been supported by the Deutsche Forschungsgemeinschaft
495: through SFB 486  and by the Volkswagenstiftung under grant No.\ I/77 217.
496: 
497: 
498: \begin{thebibliography}{10}
499: \expandafter\ifx\csname url\endcsname\relax
500:   \def\url#1{\texttt{#1}}\fi
501: \expandafter\ifx\csname urlprefix\endcsname\relax\def\urlprefix{URL }\fi
502: 
503: \bibitem{Reed1997a}
504: M.~A. Reed, C.~Zhou, C.~J. Muller, T.~P. Burgin, J.~M. Tour, Conductance of a
505:   molecular junction, Science 278 (1997) 252.
506: 
507: \bibitem{Cui2001a}
508: X.~D. Cui, A.~Primak, X.~Zarate, J.~Tomfohr, O.~F. Sankey, A.~L. Moore, T.~A.
509:   Moore, D.~Gust, G.~Harris, S.~M. Lindsay, Reproducible measurement of
510:   single-molecule conductivity, Science 294 (2001) 571.
511: 
512: \bibitem{Joachim2000a}
513: C.~Joachim, J.~K. Gimzewski, A.~Aviram, Electronics using hybrid-molecular and
514:   mono-layer devices, Nature 408 (2000) 541.
515: 
516: \bibitem{Schon2001a}
517: J.~H. Sch\"on, H.~Meng, Z.~Bao, Self-assembled monolayer organic field-effect
518:   transistors, Nature 413 (2001) 713.
519: 
520: \bibitem{Reichert2002a}
521: J.~Reichert, R.~Ochs, D.~Beckmann, H.~Weber, M.~Mayor, H.~v.~Loehneysen,
522:   Driving current through single organic molecules, Phys. Rev. Lett. 88 (2002)
523:   176804.
524: 
525: \bibitem{Nitzan2001a}
526: A.~Nitzan, Electron transmission through molecules and molecular interfaces,
527:   Ann. Rev. Phys. Chem. 52 (2001) 681.
528: 
529: \bibitem{Mujica1994a}
530: V.~Mujica, M.~Kemp, M.~A. Ratner, Electron conduction in molecular wires. i. a
531:   scattering formalism, J. Chem. Phys. 101 (1994) 6849.
532: 
533: \bibitem{Nitzan2001b}
534: A.~Nitzan, A relationship between electron-transfer rates and molecular
535:   conduction, J. Phys. Chem. A 105 (2001) 2677.
536: 
537: \bibitem{Petrov2001a}
538: E.~G. Petrov, P.~H\"anggi, Nonlinear electron current through a short molecular
539:   wire, Phys. Rev. Lett. 86 (2001) 2862.
540: 
541: \bibitem{Heurich2001a}
542: J.~Heurich, J.~C. Cuevas, W.~Wenzel, G.~Sch\"on, Electrical transport through
543:   single-molecule junctions: from molecular orbitals to conduction channels,
544:   cond-mat/  (2001) 0110147.
545: 
546: \bibitem{Grossmann1991a}
547: F.~Grossmann, T.~Dittrich, P.~Jung, P.~H\"anggi, Coherent destruction of
548:   tunneling, Phys. Rev. Lett. 67 (1991) 516.
549: 
550: \bibitem{Grossmann1992a}
551: F.~Gro{\ss}mann, P.~H\"anggi, Localization in a driven two-level dynamics,
552:   Europhys. Lett. 18 (1992) 571.
553: 
554: \bibitem{Morillo1993a}
555: M.~Morillo, R.~I. Cukier, Control of proton transfer reactions with external
556:   fields, J. Chem. Phys. 98 (1993) 4548.
557: 
558: \bibitem{Goychuk1996a}
559: I.~A. Goychuk, E.~G. Petrov, V.~May, Control of the dynamics of a dissipative
560:   two-level system by a strong periodic field, Chem. Phys. Lett. 253 (1996)
561:   428.
562: 
563: \bibitem{Grifoni1998a}
564: M.~Grifoni, P.~H\"anggi, Driven quantum tunneling, Phys. Rep. 304 (1998) 229.
565: 
566: \bibitem{Shirley1965a}
567: J.~H. Shirley, Solution of the {S}chr\"odinger equation with a {H}amiltonian
568:   periodic in time, Phys. Rev. 138 (1965) B979.
569: 
570: \bibitem{Sambe1973a}
571: H.~Sambe, Steady states and quasienergies of a quantum-mechanical system in an
572:   oscillating field, Phys. Rev. A 7 (1973) 2203.
573: 
574: \bibitem{Lehmann2002b}
575: J.~Lehmann, S.~Kohler, P.~H\"anggi, A.~Nitzan, Molecular wires acting as
576:   quantum ratchets, Phys. Rev. Lett. 88 (2002) 228305.
577: 
578: \bibitem{Blumel1991a}
579: R.~Bl\"umel, A.~Buchleitner, R.~Graham, L.~Sirko, U.~Smilansky, H.~Walter,
580:   Dynamical localisation in the microwave interaction of {R}ydberg atoms: The
581:   influence of noise, Phys. Rev. A 44 (1991) 4521.
582: 
583: \bibitem{Bruder1994a}
584: C.~Bruder, H.~Schoeller, Charging effects in ultrasmall quantum dots in the
585:   presence of time-varying fields, Phys. Rev. Lett. 72 (1994) 1076.
586: 
587: \bibitem{Demming1998a}
588: F.~Demming, J.~Jersch, K.~Dickmann, P.~I. Geshev, Calculation of the field
589:   enhancement on laser-illuminated scanning probe tips by the boundary element
590:   method, Appl. Phys. B 66 (1998) 593.
591: 
592: \bibitem{Mujica2000a}
593: V.~Mujica, A.~E. Roitberg, M.~Ratner, Molecular wire conductance: Electrostatic
594:   potential spatial profile, J. Chem. Phys. 112 (2000) 6834.
595: 
596: \bibitem{Chen1999a}
597: J.~Chen, M.~A. Reed, A.~M. Rawlett, J.~M. Tour, Large on-off ratios and
598:   negative differential resistance in a molecular electronic device, Science
599:   286 (1999) 1550.
600: 
601: \bibitem{Collier2000a}
602: C.~P. Collier, G.~Mattersteig, E.~W. Wong, Y.~Luo, K.~Beverly, J.~Sampaio,
603:   F.~M. Raymo, J.~F. Stoddart, J.~R. Heath, A [2]catenane-based solid state
604:   electronically reconfigurable switch, Science 289 (2000) 1172.
605: 
606: \bibitem{Braun1998a}
607: E.~Braun, Y.~Eichen, U.~Sivan, G.~Ben-Yoseph, {DNA}-templated assembly and
608:   electrode attachment of a conducting silver wire, Nature 391 (1998) 775.
609: 
610: \bibitem{Raedler1998a}
611: O.~J. R\"adler, I.~Koltover, A.~Jamieson, T.~Salditt, C.~R. Safinya, Structure
612:   and interfacial aspects of self-assembled cationic lipid-{DNA} gene carrier
613:   complexes, Langmuir 14 (1998) 4272.
614: 
615: \bibitem{Blick1996a}
616: R.~H. Blick, R.~J. Haug, J.~Weis, D.~Pfannkuche, K.~v. Klitzing, K.~Eberl,
617:   Single-electron tunneling through a double quantum dot: The artificial
618:   molecule, Phys. Rev. B 53 (1996) 7899.
619: 
620: \end{thebibliography}
621: 
622: 
623: \end{document}
624: