cond-mat0101157/wp.tex
1: %&LaTeX
2: \documentstyle[preprint,prb,aps,epsfig]{revtex} 
3: \begin{document}
4: \tighten
5: \title{An Electromechanical Which--Path Interferometer}
6: \author{A. D. Armour\cite{auth}} 
7: \address{The Blackett Laboratory, Imperial College of Science,
8: Technology and Medicine, London SW7~2BZ, United Kingdom}
9: \author{M. P. Blencowe\cite{auth2}}
10: \address{Department of Physics and Astronomy, Dartmouth College,
11: Hanover, New Hampshire 03755}
12: \date{\today}
13: \maketitle
14: \begin{abstract}
15: We investigate the possibility of an electromechanical which--path 
16: interferometer, in 
17: which electrons travelling through an 
18: Aharonov--Bohm  
19: ring incorporating a quantum dot in one of the arms are dephased by an interaction 
20: with the fundamental flexural mode of a 
21: radio frequency cantilever. The cantilever is positioned so that its tip
22:  lies just above the dot and a bias is applied so that an electric field
23:   exists between the dot and the tip. This electric field is modified when an 
24:   additional electron hops onto the dot, coupling the 
25:    flexural mode of the cantilever and the microscopic electronic 
26:   degrees of freedom. We analyze the transmission properties of this system 
27:   and the dependence of interference fringe visibility on the cantilever--dot 
28:   coupling and on the mechanical properties of the cantilever. 
29:   The fringes are progressively 
30:   destroyed as the interaction with the cantilever is turned up, in 
31:   part due to dephasing arising  from the entanglement of the electron and 
32:   cantilever states and also due to the thermal smearing that results from 
33:   fluctuations in the state of the cantilever. When the dwell time of the 
34:   electron on the dot is comparable to or longer than the cantilever period,
35:   we find coherent 
36:   features in the transmission amplitude. These features are washed out 
37:   when the cantilever is decohered by its coupling to the environment. 
38:   \end{abstract}
39: \pacs{PACS numbers: 85.85.+j, 85.35.Ds, 73.63.Kv, 73.23.Hk, 03.65.Yz }
40: 
41: \section{Introduction}
42: 
43: Which--path devices such as the canonical
44:  two--slit interference experiment, where a  measurement is made of the
45: path a particle takes,  
46:  have a long 
47: history going back as far as the early debates about 
48: complementarity.\cite{hist,Fey} 
49: Recent interest in their investigation has been stimulated
50: partly by advances in experimental
51: techniques, which have lead to the realization of several different
52: varieties of which--path systems in the laboratory,\cite{Buks,op} and
53: partly by accompanying developments in the theory of quantum
54: measurement.\cite{SS,Scully,Walls} However, it is  the
55: realization that a which--path experiment provides a very convenient
56: model system for developing and testing fundamental ideas about
57: decoherence in mesoscopic systems,\cite{Stern,hacr} that has  increased the
58: level of interest within the solid state physics community in particular.
59: 
60: 
61: Which--path experiments in solid state systems were recently pioneered
62: by Buks {\emph{et al.}}\cite{Buks} The solid state analog
63:  of the two--slit interference experiment is the measurement of the oscillations
64:   in the current passing  through an Aharonov--Bohm (AB) ring as a function 
65:   of the applied magnetic field. The path taken by an electron may be probed by 
66:   placing a measuring device close to one of the arms. Buks {\emph{et al.}}
67:    incorporated a quantum dot in one of the arms in order to slow the electrons 
68:    down, with a neighboring quantum point contact (QPC) serving as a 
69:    which--path detector.
70: An electron travelling around the arm of the ring containing the dot 
71:  dwells on the dot for a finite amount of time before 
72: moving on. The presence of the dot alone does not destroy 
73: the coherence of the electron transport through the ring, so long as the dwell time 
74: of the electrons is sufficiently short,\cite{Yac} but it does provide time for the 
75: electron to interact with an external measuring device. A QPC adjacent to the dot 
76: functions as a measuring device since it can be biased so that its 
77: conductance is very sensitive to changes in the occupancy of the dot: the 
78: passage of an electron via the path including the dot leaves behind which--path 
79: information in the QPC device, although actual knowledge of which path 
80: an electron took, so--called `true which--path information', can only be obtained 
81: via further measurement.\cite{Hack} In accord with theoretical
82: predictions,\cite{Le,Al,Gurvitz} the experiment demonstrated that
83:  the interference fringes are degraded when the interaction 
84: between the electrons and the measuring device is sufficiently 
85: sensitive
86: for information to be obtained that would help, even in principle,
87:  to determine which of the two possible paths an electron
88: took.    
89: 
90: The experiment of Buks {\emph{et al.}} has close parallels with 
91: the well--known thought experiment (see, e.g., Sec. 1-6 of Ref.\ 
92: \onlinecite{Fey}), in which 
93: a light source
94: is used to detect through which slit an electron passes in a  two--slit
95: interference experiment. In both cases dephasing is effected by
96: scatterers, each of which interact once, and once only, with the interfering
97: particle and whose interactions with each other may safely be ignored.
98: For the electron--light scattering scheme,\cite{Fey} the scatterers are photons
99: which probe the electron's state directly, whereas in the
100: solid--state experiment the scatterers are electrons in the QPC
101: which interact with the electron on the dot via electrostatic coupling. 
102: However, because the electron has a finite dwell time on the
103: dot, it has time to interact with  more than one scatterer in the QPC, 
104: so that in this case dephasing can be achieved via a series of 
105: very weak interactions rather than a single strong event, as is the
106: case in the photon thought experiment.
107: 
108: 
109: The present work is concerned with a variation on the system 
110: considered by Buks 
111: {\emph{et al.}}, 
112: in which a radio frequency mechanical cantilever is used, rather than a
113: QPC, to determine which path an electron takes. A coupling  between electrons 
114: residing on the dot (which is again on one of the AB interferometer arms) and 
115:  the
116: cantilever, whose tip is suspended over the dot, can be set up by developing a 
117: uniform electric field between the tip of the cantilever and the base of the dot. 
118: Electrons on the dot couple approximately linearly to the cantilever position, thus leading 
119: to a coupling between the flexural phonon modes of the cantilever and the
120: occupancy of the dot. Furthermore, because the coupling strength decays rapidly 
121: with increasing frequency  
122: and  also the flexural mode spectrum of a cantilever 
123: is quite sparse, it turns out that for micron scale  cantilevers  only the fundamental
124: flexural mode is relevant. Therefore, at low 
125: temperatures the cantilever can be
126: treated as a single quantum mechanical oscillator.  
127: 
128: 
129: As a consequence of the cantilever having effectively only 
130: one degree of freedom,  the electromechanical which--path 
131: interferometer exhibits qualitatively different behavior from the QPC 
132: which--path device. In particular, we 
133: find that the dephasing behavior of 
134: the electron due to the cantilever depends on the 
135: relative magnitudes of the electron dwell time on the dot and the cantilever 
136: period. When the  dwell time is short compared to the cantilever 
137: period, the dephasing occurs in a way analogous to Einstein's recoiling slit
138: thought--experiment.\cite{hist,SS} In contrast, when the dwell time of
139: the electron on the dot is comparable to or longer than the period of the 
140: cantilever,
141: a description of the dephasing in terms of the entanglement of the
142: cantilever and electron states becomes more appropriate than a
143: semiclassical picture of momentum transfer.
144: The effectively harmonic nature of  the cantilever motion
145: means that the degree of entanglement between the cantilever and 
146: electronic states
147: must be periodic, so long as the cantilever interacts only with 
148: the electron on the dot. If the dwell time
149:  of the electrons on the dot could be tuned to the cantilever period, 
150:  then we would 
151:    be able to erase all which--path information held in the state of the 
152:    cantilever. 
153:    This would give us direct 
154:     evidence that the 
155:    coupled cantilever and dot were behaving as a single coherent quantum system. 
156:  However, in practice electron dwell times on a quantum dot have a distribution
157:   of values and so we can only obtain indirect evidence for the quantum 
158:  coherence of the  cantilever--dot system, such as the coherent exchange of 
159:   energy quanta between the electron and the cantilever which give rise to 
160:  side resonances in the elastic transmission amplitude of the device. 
161:  
162:  The environment of the cantilever influences its behavior  in two important ways: 
163:  over short timescales it destroys the cantilever's phase coherence, whilst
164:  over longer timescales it damps the cantilever's motion.
165:  The electromechanical which--path system we propose acts as a probe of 
166:  the cantilever's decoherence due to it's interaction
167:   with the environment. We show that the cantilever's decoherence 
168:   inhibits the coherent
169:   exchange of energy between the cantilever and the dot and hence manifests 
170:   itself by washing out the side resonances in the elastic transmission 
171:   amplitude.  
172:      
173: 
174: 
175: The outline of this paper is as follows. In Sec. II we give our basic 
176: model for the electromechanical
177: which--path device and show that the interference properties of the AB
178: ring, modified to include the dot and cantilever, can be obtained by calculating
179: the elastic transmission amplitude of the arm containing the dot. In 
180: Sec. III 
181: we describe a simple tight--binding model for an electron on a  dot which is linearly
182: coupled to the cantilever (treated as a quantum oscillator) and to
183:  propagating states in the leads. We show that recent work on
184: inelastic resonant tunneling can be adapted to obtain the elastic
185: transmission amplitude. 
186: 
187: We present results for our model in Sec. IV, where we examine
188: dephasing as a function of the dot--cantilever coupling in both
189: the regime where the electron dwell time is short compared to the
190: cantilever period and  where they are comparable. 
191: 
192: In Sec. V we examine the influence of the cantilever's
193: environment on the  dephasing produced by the cantilever when the
194: dwell time is comparable to the period of the cantilever. We obtain a modified
195: expression for the transmission through the arm with the dot when the 
196: cantilever is coupled to an environment which consists of a bath of oscillators.  
197: 
198: 
199: In Sec. VI we outline the principal practical constraints on an
200: electromechanical which--path device and discuss its feasibility 
201: using currently available technology. We draw our conclusions in Sec. 
202: VII and discuss ways in which the present work could be extended.
203: Appendix A contains a detailed, classical analysis of the cantilever--dot 
204: coupling and the flexural mode spectrum of the cantilever which underpin 
205: the simplified model employed in the main text. Finally, appendix B contains 
206: details of a general calculation of the effect of the cantilever environment on
207: the transmission amplitude of the dot. 
208: 
209: \section{Theoretical Model}
210: 
211: The ultimate goal of our analysis of the electromechanical which--path device
212: is to obtain an expression for how
213: the magnetoresistance of the AB ring varies as a function of the coupling
214: between the cantilever and an electron on the dot. In particular, we
215: want to determine how  the amplitude and phase of the
216: oscillation in the current through the device as the magnetic field is
217: varied (i.e., the interference fringes) depend on the voltage between the
218: cantilever and the dot. 
219: In order to simplify the analysis, we will consider only the case
220: where the magnetoresistance is measured as an average over a 
221: time  much longer than any other timescale in the
222: problem.    
223:    
224: Our model builds  on the theoretical analysis of the
225: QPC which--path experiment,  carried out by a number of
226: groups.\cite{Le,Al,Gurvitz}  However,
227: there are also close parallels between the system we
228: consider and a quantum optical system analyzed recently,\cite{BJK,Bose} in which 
229: radiation is used to drive an oscillator--mounted mirror into non--classical states. 
230: 
231: Since we are interested in the behavior of the system averaged over
232: time, it is sufficient to work within the Landauer framework, so that
233: our task of obtaining the average magnetoresistance is equivalent to
234: calculating the transmission characteristics of the device.\cite{Datta}
235: The simplest way of thinking about the transmission probability through  
236:  the AB ring is in terms of a two--slit experiment: the total 
237: transmission probability, $\mathcal{T}$,  is given by the coherent sum of 
238: amplitudes for 
239: transmission through the left and right arms, $t_l$ and $t_r$,
240: \begin{equation}
241: {\mathcal{T}}=|t_l+t_r|^2.\label{sm}
242: \end{equation}
243: This approach was employed by Yacoby {\emph{et al.}}\cite{Yac} in their 
244: analysis of the first experiments on AB rings with a quantum dot 
245: included in one arm. However, in that case this simple 
246: formalism proved
247: inadequate as it neglects the contribution of electrons which have
248: undergone multiple reflections. The contribution to the transmission
249: of multiply reflected electrons leads to the enforcement of the
250: Onsager--type relation for the conductance from source (S) to 
251: drain (D): $G_{DS}(B)=G_{DS}(-B)$, which must apply for a
252: two--terminal device.\cite{But1,But2,Imry,Yac2}
253: If the AB ring is modified to absorb electrons which are scattered 
254: backwards in the device, then the Onsager--type condition is no longer satisfied 
255: and the simple--minded formula (\ref{sm}) is valid. Buks 
256: {\emph{et al.}} incorporated  just such a modification in 
257: their AB ring and we shall limit our analysis to this case.    
258:  We write 
259: $\tilde{t}_{\mathrm QD}$ for the transmission through the arm containing the dot and 
260: $t_0$ for the other amplitude. The presence of a magnetic flux, $\Phi$, induces 
261: an additional  relative phase shift between the two paths and so the 
262: transmission probability in the absence of the cantilever can be written as
263: \begin{equation}
264: {\mathcal{T}}=\left|t_0+\tilde{t}_{\mathrm QD}{\mathrm{e}}^{i 
265: 2\pi\Phi/\Phi_0}\right|^2
266: ={\mathcal{T}}^{(0)}+2{\mathrm{Re}}\left[ t^*_0\tilde{t}_{\mathrm QD}{\mathrm{e}}^{i 
267: 2\pi\Phi/\Phi_0}\right], \label{T}
268: \end{equation}
269: where $\Phi_0$ is the flux quantum and ${\mathcal{T}}^{(0)}$ is the 
270: flux independent term of the transmission probability. In practice, 
271: the transmission through the arm with the dot is much
272: less than that through the arm without the dot,  
273: $|\tilde{t}_{\mathrm QD}|\ll|t_0|$, and so ${\mathcal{T}}_0\approx |t_0|^2$.
274: 
275: When there is a non--zero interaction between the cantilever and the
276: electron on the dot, we must  explicitly include the fact that the
277: transmission depends on the initial  state of the
278: cantilever. Since the cantilever contains a macroscopic number 
279: of atoms we assume that
280: its initial state can always be described as a thermal mixture. Such a
281: procedure will necessarily lead us to neglect short time correlations
282: between the cantilever states, but this will be unimportant if the
283: magnetoresistance is averaged over a time which is long compared with
284: the characteristic timescale for the thermalization of the
285: cantilever's motion.\cite{steady} With the initial state of the 
286: cantilever assumed to be thermal we have,
287: \begin{equation}
288: {\mathcal{T}}=\sum_i\sum_f \rho_i\left[ \langle i|t_0 
289: +\hat{t}_{\mathrm QD}{\mathrm{e}}^{i 2\pi\Phi/\Phi_0} | f\rangle \left(\langle
290:   i|t_0 +\hat{t}_{\mathrm QD}{\mathrm{e}}^{i
291: 2\pi\Phi/\Phi_0} | f\rangle\right)^*\right], \label{T2} 
292: \end{equation}
293: where $\hat{t}_{\mathrm QD}$ is an operator on cantilever states only, $\rho_i$ is
294: the usual thermal weight ($\rho_i={\mathrm{e}}^{-\beta
295:   \epsilon_i}/\sum_j{\mathrm{e}}^{-\beta \epsilon_j}$, with
296: $\beta=1/k_{\mathrm{B}}T$ and $\epsilon_i$ the energy of the state),
297:  and we have made no  assumption about the final state
298:   of the cantilever mode. Because only the dot arm interacts with the cantilever
299: the interference term is diagonal in the cantilever modes,
300: \begin{equation}
301: 2{\mathrm{Re}}\left[ \sum_i \rho_i t^*_0\langle i|\hat{t}_{\mathrm QD}|i 
302: \rangle {\mathrm{e}}^{i 2\pi\Phi/\Phi_0}\right]=
303: 2{\mathrm{Re}}\left[t^*_0\langle\tilde{t}_{\mathrm QD}\rangle 
304:  {\mathrm{e}}^{i 2\pi\Phi/\Phi_0}\right],
305: \end{equation}  
306: where $\langle \tilde{t}_{\mathrm QD}\rangle$ is averaged over the cantilever thermal
307: state. Thus, only elastic scattering processes contribute to the 
308: interference term. 
309: 
310: At finite temperatures we also have to average the transmission amplitude for 
311: transport through the arm containing the dot over the Fermi 
312: distribution:\cite{Datta}
313: \begin{equation}
314: \langle\tilde{t}_{\mathrm QD}\rangle
315: =\int_{0}^{\infty} {d}\epsilon\left( -\frac{\partial f}{\partial 
316: \epsilon}\right) \langle t_{\mathrm QD}(\epsilon)\rangle. \label{therm}
317: \end{equation}
318: 
319: 
320: 
321: The knowledge of the transmission amplitude allows us to obtain the
322: amplitude of the periodic oscillations in the current, or in the
323: language of a generic two--slit experiment, the visibility of the
324: interference fringes. The visibility, $v$, in a two--slit experiment is 
325:  defined in terms of the maximum and minimum signal (in this case 
326: current), measured at the peaks and valleys of the fringes,
327:  respectively:\cite{vis}
328: \begin{equation}
329: v=\frac{{\mathrm{max.}}-{\mathrm{min.}}}{{\mathrm{max.}}+{\mathrm{min.}}}.
330: \end{equation}
331: 
332: In the case of the electromechanical which--path device considered here, the 
333: current is proportional to the transmission probability, $\mathcal{T}$, 
334: given in Eq. (\ref{T2}) above. We can write the transmission amplitudes through 
335: the two arms in modulus--argument form, 
336: $\langle\tilde{t}_{\mathrm QD}\rangle=|\langle\tilde{t}_{\mathrm QD}
337: \rangle|{\mathrm{e}}^{i\alpha}$ and 
338: ${t}_0^*=|t_{0}|{\mathrm{e}}^{-i\beta}$, so that the interference 
339: part of the transmission probability takes the form
340: \begin{equation}
341: 2{\mathrm{Re}}\left[t_0^*
342: \langle\tilde{t}_{\mathrm QD}\rangle{\mathrm{e}}^{i2\pi\Phi/\Phi_0}\right]
343: =2|{t}_0||\langle\tilde{t}_{\mathrm QD}\rangle|\cos\left( 2\pi\Phi/\Phi_0+ \alpha-\beta 
344: \right).
345: \end{equation}
346: Hence in this case, the visibility of the fringes is
347: \begin{equation}
348: v=\frac{2|{t}_0||\langle\tilde{t}_{\mathrm QD}\rangle|}{{\mathcal{T}}^{(0)}}
349: \simeq \frac{2|\langle\tilde{t}_{\mathrm QD}\rangle|}{|{t}_0|}. 
350: \end{equation}
351: The phase of the
352: transmission amplitude  determines the phase of the fringe
353: pattern as a function of the magnetic field. Any change in the phase
354: of the transmission amplitude should therefore be detectable as a change in the
355: phase of the whole interference pattern. 
356: 
357: In the absence of the dot, and for the ideal case of a ring in which both arms 
358: are identical, the transmission amplitudes for both arms are 
359: the same so that ${\mathcal{T}}_0=2 |{t}_0|^2$ and the visibility is unity. 
360: This demonstrates the utility of $v$ as a measure:  it does not depend 
361: on the value of the total transmission probability through the device,
362: but just on its interferometric properties.
363: 
364: In practice, a fringe visibility close to unity cannot be
365: achieved. Apart from the obvious difficulty in constructing a ring in which both 
366: arms are identical, there are two further reasons why the `intrinsic' 
367: visibility is less than unity.\cite{Yac} Firstly, there is more than one conduction 
368: channel in each arm and so the transport is never really a `one electron' problem. 
369: Secondly, thermal smearing can play an important r\^{o}le, since the thermal 
370: smearing length is typically comparable to the size of the ring at temperatures
371:  of around 100mK.\cite{Yac} However, for our device 
372:  the reduction in fringe visibility arising from sources 
373: other than the cantilever is irrelevant: all we 
374: require is that there should be a measurable change in the visibility of the 
375: fringes as the electric field coupling the dot and cantilever is turned on.
376: 
377: 
378: \section{Transmission Amplitude}
379: 
380: In order to obtain the transmission amplitude for the arm of the AB
381:  ring containing the dot, we employ a standard tight--binding model for the dot 
382: and the leads to which it is coupled. In
383: the Coulomb blockade regime, the quantum dot is modelled by a single localised 
384: state, at energy $\epsilon_0$, which is coupled to sets of non--interacting,
385:  propagating, states in
386:  the leads. The cantilever is treated as a single quantum oscillator mode of frequency
387:   $\omega_0$, although the generalisation to include a spectrum of modes is 
388:   straightforward, as we discuss below.   
389: The total Hamiltonian of the interferometer arm can be written as the
390: sum of two parts 
391: \begin{equation}
392: {\mathcal{H}}={\mathcal{H}}_0+{\mathcal{H}}_1 \label{Ham1},
393: \end{equation}
394: with non--interacting part
395: \begin{equation}
396: {\mathcal{H}}_0=\epsilon_0\hat{c}^{\dagger}\hat{c}+\sum_k\left( 
397: \epsilon_{kL}\hat{c}^{\dagger}_{kL}\hat{c}_{kL}+\epsilon_{kR}\hat{c}^{\dagger}_
398: {kR}\hat{c}_{kR}\right)+\hbar \omega_0 \hat{a}^{\dagger}\hat{a},
399: \end{equation}
400: and interacting part
401: \begin{equation}
402: {\mathcal{H}}_1=-\lambda \hat{c}^{\dagger}\hat{c}(\hat{a}+\hat{a}^{\dagger})
403: +\sum_k V_{kL}\left( 
404: \hat{c}^{\dagger}_{kL}\hat{c}+\hat{c}^{\dagger}\hat{c}_{kL}\right)
405: +\sum_k V_{kR}\left( 
406: \hat{c}^{\dagger}_{kR}\hat{c}+\hat{c}^{\dagger}\hat{c}_{kR}\right) \label{coup},
407: \end{equation}
408: where $\hat{c}$ and $\hat{a}$ operate on the states of the electron on the 
409: dot and the cantilever, respectively. The states in the left--hand(right--hand) 
410: leads have energies $\epsilon_{kL}$($\epsilon_{kR}$) and are operated on by 
411: $\hat{c}_{kL}$($\hat{c}_{kR}$), where the index $k$ runs over propagating 
412: states in the leads. The matrix elements for hopping onto(off) the dot 
413: are 
414: given by $V_{kL}$($V_{kR}$). The interaction between the electron on
415: the dot and the cantilever is modelled as a linear coupling between the
416: displacement of the flexural mode and the occupation of the dot, since
417: when the electron is not on the dot there is no displacement of the
418: cantilever.\cite{justify} We
419: analyze the interaction between the cantilever and the dot in some detail in 
420: appendix A, where the form of the interaction is derived in terms 
421: of the normal modes of the cantilever. We find that the coupling constant, 
422: $\lambda$, is given by the relation $\lambda=\xi 
423: eE\sqrt{\hbar/2m\omega_0}$ [see Eq.\ (\ref{coupling})], 
424: where $\xi$ is a geometrical factor of order one, $E$ is the electric field 
425: experienced by the additional electron on the dot, $m$ is the cantilever mass 
426: and $e (>0)$ the electronic charge.
427: 
428: 
429: The energy-dependent amplitude, $t_{\mathrm QD}(\epsilon)$, is calculated using
430: the usual  
431: $S$--matrix formalism\cite{W89} employed in transport theory. 
432: The amplitude for transmission from a state of energy $\epsilon$ to one of 
433: energy $\epsilon'$ is given by the element of the 
434: $S$--matrix\cite{W89} linking propagating states in the left and right--hand leads
435: \begin{equation}
436: \langle \epsilon', R| S|\epsilon, L \rangle =t_{\mathrm QD}(\epsilon,\epsilon'),
437: \end{equation} 
438: where $\langle  \epsilon', R|$($|  \epsilon, L\rangle$) is the state in the 
439: right(left)--hand lead with energy $\epsilon'$($\epsilon$). The total 
440: transmission amplitude at energy $\epsilon$ is then obtained by 
441: integrating over the final state energies
442: \begin{equation}
443: t_{\mathrm QD}(\epsilon)=\int_0^{\infty}t_{\mathrm QD}(\epsilon,\epsilon')d 
444: \epsilon'.\label{fermi}
445: \end{equation} 
446: 
447: The $S$--matrix element for a single electron tunneling from left to right 
448: through 
449: a single, localised, state which is coupled to a cantilever mode can be 
450: calculated using either the methods described by Wingreen {\emph{et al.}} 
451: \cite{W89} or those of Glazman and Shekhter.\cite{GS}
452:  For initial and final states of the cantilever
453: given by $\alpha_i$ 
454: and $\alpha_f$ respectively, the relevant matrix element is
455: \begin{eqnarray}
456: \langle \epsilon', \alpha_f, R|S|\epsilon, \alpha_i, L \rangle=&-&i\int\int 
457: \frac{dt_1dt_2}{\hbar^2}{\mathrm{e}}^{-\eta(|t_1|+|t_2|)}\cr
458:  &\times& \langle \epsilon', \alpha_f, R|
459: {\mathrm{e}}^{i{\mathcal{H}}_0t_2/\hbar}{\mathcal{H}}_1 
460: \hat{G}_R(t_2-t_1){\mathcal{H}}_1{\mathrm{e}}^{-i{\mathcal{H}}_0t_1/\hbar}
461: |\epsilon, \alpha_i, L \rangle,
462: \end{eqnarray}
463: where $\hat{G}_R(t)=-i\Theta(t){\mathrm{e}}^{-i{\mathcal{H}}t/\hbar}$
464: and $\eta$ is the usual small positive real number inserted to ensure 
465: convergence. 
466: 
467: We want to calculate the visibility of the fringes and so we need only calculate 
468: the coherent part of the transmission probability, and hence the elastic 
469: transmission amplitude. To evaluate the elastic transmission amplitude we  consider 
470: processes in which the state of the cantilever remains unchanged and so 
471: we write $\alpha_i=\alpha_f=\alpha$ and calculate an average over an ensemble 
472: of 
473: states of the cantilever (see the discussion in Sec. II above). Thus
474: \begin{equation}
475: \langle S \rangle=-V_R(\epsilon)V_L(\epsilon')\int\int 
476: \frac{{d}t_1dt_2}{\hbar^2}{\mathrm{e}}^{-\eta(|t_1|+|t_2|)}
477: \times {\mathrm{e}}^{i(\epsilon' t_2-\epsilon t_1)/\hbar}\Theta(t_2-t_1)\langle 
478: 0,\alpha| \hat{c}(t_2)\hat{c}^{\dagger}(t_1)|0, \alpha \rangle 
479: \end{equation}
480: with 
481: $|V_{L(R)}(\epsilon)|^2=\sum_k|V_{kL(R)}|^2\delta(\epsilon-\epsilon_{kL(kR)})$ 
482: and where  $|0,\alpha\rangle$ is the state with no electrons on the 
483: localised level and  the cantilever in state $\alpha$.
484: 
485: We assume that the coupling to the leads is independent of energy over
486: the range of interest and that it is symmetric so that  we can write
487: \begin{equation}
488: \Gamma=\Gamma_{L(R)}=2\pi|V_{L(R)}|^2.
489: \end{equation}
490: Since we are considering only the elastic part of the transmission
491: amplitude the Green function will be invariant with respect to time.
492:  Hence, we change variables to 
493: $\tau=t_2-t_1$ and $t_0=t_1$, so that the Green function takes the form
494: \begin{equation}
495: G^R(\tau,t_0)=-i\Theta(\tau)\langle 
496: \hat{c}(\tau+t_0)\hat{c}^{\dagger}(t_0)\rangle
497: \end{equation}
498: and the  invariance with respect to translations in time is 
499: equivalent to  the statement that the value of the Green function is 
500: independent of the choice of $t_0$. Averaging over $t_0$, and taking 
501: the limit $\eta \rightarrow 0^+$, 
502: the overall expression for the $S$--matrix element is then
503: \begin{equation}
504: \langle \epsilon', R|S |\epsilon , L\rangle=
505: -i\frac{\Gamma}{2 
506: \pi}\int_0^{\infty} \frac{d\tau}{\hbar} 
507: {\mathrm{e}}^{i\epsilon\tau/\hbar}G^R(\tau) 
508: \times 2\pi\delta(\epsilon-\epsilon').
509: \end{equation}
510: 
511: Thus, the final expression for the transmission amplitude of the dot is 
512: \begin{equation}
513: \langle t_{\mathrm QD}(\epsilon)\rangle=-i\Gamma
514: \int_0^{\infty}\frac{d\tau}{\hbar}
515: {\mathrm{e}}^{i\epsilon\tau/\hbar}G^R(\tau), \label{tqd}
516: \end{equation}
517: where the transmission amplitude is understood to be time--averaged in
518: the sense described in Sec. II, and the 
519: relevant Green function is 
520: \begin{equation}
521: G^R(\tau)=-i\Theta(\tau)\langle\hat{c}(\tau)\hat{c}^{\dagger}(0)\rangle, \label{tqd22} 
522: \end{equation}
523: where the expectation value is over a thermal distribution state of the 
524: cantilever  with no electrons present. This result for the 
525: transmission amplitude  is  very similar to that
526: studied by Aleiner {\emph{et al.}},\cite{Al} in their analysis of the
527: QPC which--path device.
528: 
529: The Green function can be evaluated using either operator algebra
530: techniques,\cite{GS} or many body perturbation theory,\cite{W89} and 
531: one obtains
532: \begin{equation}
533: G^R(\tau)=G^R_0(\tau){\mathrm{e}}^{-\phi(\tau)}, \label{tqd2}
534: \end{equation}
535: where 
536: \begin{equation}
537: G^R_0(\tau)=-i\Theta(\tau){\mathrm{e}}^{(-i\epsilon_0-\Gamma/2)\tau/\hbar},
538: \end{equation}
539: and it has been assumed that the renormalisation of the dot energy,
540: $\epsilon_0$, due to coupling to the leads can be ignored.
541: The factor due to coupling to the cantilever, $\phi(\tau)$, can be obtained by
542: calculating the contribution for the cantilever beginning and ending in a
543: particular coherent state, $|\nu\rangle$, and then carrying out a thermal
544: average over all coherent states with appropriate weightings,
545: \begin{equation}
546: {\mathrm{e}}^{-\phi(\tau)}=
547: {\mathrm{e}}^{i(\lambda/\hbar\omega_0)^2[\omega_0\tau-\sin(\omega_0 \tau)]}
548: {\mathrm{e}}^{-(\lambda/\hbar\omega_0)^2[1-\cos(\omega_0\tau)]}
549: \int\frac{{d}^2\nu}{\pi\overline{n}}{\mathrm{e}}^
550: {(\lambda/\hbar\omega_0)(\nu^*\mu-\nu\mu^*)}{\mathrm{e}}
551: ^{-|\nu|^2/\overline{n}},\label{ephi}
552: \end{equation}
553: where $\mu={\mathrm{e}}^{i\omega_0 t}-1$. 
554: Evaluating the thermal average, we find
555: \begin{equation}
556: \phi(\tau)
557: =\left(\frac{\lambda}{\hbar \omega_0}\right)^2\left\{i[\sin(\omega_0 
558: \tau)-\omega_0 \tau]+[1-\cos(\omega_0 \tau)](1+2\overline{n})\right\}, 
559: \label{ex2}
560: \end{equation}
561: with $\overline{n}$ the thermal occupation of the cantilever mode. 
562: 
563: It is
564: clear that the thermal average leads to a much more rapidly decaying
565: term, due to the extra factor of $2\overline{n}$. If the
566: cantilever remained in a coherent state throughout then it would be far less
567: effective, compared to the thermal state, at reducing the visibility of the 
568: fringes. This is because each coherent state affects the transmission amplitude 
569: in two different ways: the magnitude of the transmission is reduced by an amount 
570: which is independent of which state the cantilever is in and also
571:  a phase shift is induced whose size depends 
572: sensitively on the cantilever state.   
573: When we carry out the thermal average we are in effect averaging over a range
574: of different phase shifts. Such a procedure effectively destroys the
575: interference fringes whenever the thermal state includes contributions
576: in which the phase differs by $\sim 2\pi$, irrespective of the magnitude of the 
577: transmission amplitude for each of the coherent states which constitute the 
578: thermal mixture. The thermal averaging timescale is given by the damping time 
579: of the cantilever which will typically be much 
580: larger than the dwell time of the electron on the dot. In this case, the loss of 
581: fringe visibility arising from thermal smearing is  not due 
582: to which--path detection of individual electrons: a measurement of the  current 
583: averaged over times shorter than the damping time but longer than the 
584: dwell time would resolve  AB fringes with phase fluctuating in time.  In 
585: contrast, the state-independent reduction in the transmission 
586: amplitude would give rise to a reduction in the fringe visibility even for 
587: the time-resolved measurement, signifying which--path detection.   
588: However, since we are
589:  working in a steady state regime in which we consider measurements made over 
590: very long times we will not be able to make an explicit distinction between 
591: which--path detection and thermal smearing in this work.  
592: 
593: 
594: We can repeat this calculation for the more general case of coupling
595: to a whole series of non-interacting cantilever modes. The result has the
596: same basic structure as before, but the cantilever factor, $\phi(\tau)$, 
597: is modified and now takes the form of a sum of the contributions from each
598:  mode,\cite{W89,GS}  
599: \begin{equation}
600: \phi(\tau)=\sum_i \left(\frac{\lambda_i}{\hbar \omega_i}\right)^2
601: \left[i(\sin(\omega_i 
602: \tau)-\omega_i \tau)+(1-\cos(\omega_i 
603: \tau))(1+2\overline{n}_i)\right],\label{phinoenv}
604: \end{equation}
605: where $\lambda_i$, $\omega_i$, and $\overline{n}_i$ are the coupling
606: constant, frequency, and thermal occupation number of the $i$th cantilever
607: mode, respectively. However, we  find that for the purposes of
608: demonstrating the electromechanical which--path device we need only
609: consider a single mode, as is discussed in appendix A.
610: 
611: \section{Results for Isolated Cantilever} 
612: 
613: The important question which we need to answer is the following:
614: When does the cantilever destroy the AB fringes? There are two
615: very different regimes which we can explore by varying the
616: relative sizes of characteristic dwell time of the electron on the dot, 
617: $\tau_d=\hbar/\Gamma$, and the fundamental frequency of the cantilever: 
618: \begin{enumerate}
619: \item $\omega_0\tau_d\ll1$: in this case the  
620: periodic behavior of  the cantilever will not be relevant and the entanglement
621:  built up between the cantilever and electron states during the dwell time will
622:  be irreversible, leading to dephasing. 
623:   
624: \item $\omega_0\tau_d\sim 1$: in this regime the periodicity of the
625:   cantilever means that the 
626:  entanglement of the cantilever and dot states caused by their interaction 
627:  may be partially undone, or erased,
628:  although  the distribution of electron dwell times will make the effect 
629:  impossible to observe directly. 
630: \end{enumerate}
631: 
632: The first regime is easier to analyze theoretically, as the
633: calculations can be done analytically and invite direct comparisons
634: with the sliding--slit  thought experiment. The case where the dwell
635: time of the electron on the dot is comparable to the period of the
636: cantilever has to be analyzed numerically. It is also complicated by
637: the presence of the cantilever's environment which makes the entanglement of the
638: cantilever and dot states irreversible by breaking the overall periodicity in the 
639: expression (\ref{tqd}) for the transmission amplitude. In this
640: section, we calculate the magnitude and phase of the transmission
641: amplitude in both regimes, neglecting the effect of the
642: environment which we address in detail in Sec. V. This approach enables
643: a clear picture to be built up of exactly how the environment affects the 
644: behavior of the cantilever--dot system.    
645: 
646: 
647: \subsection{Regime Where $\omega_0\tau_d\ll1$}  
648: 
649: Since $\omega_0\tau_d\ll1$, we can simplify Eq.\
650: (\ref{tqd}) for
651: the transmission amplitude through the arm with the dot 
652: by expanding the harmonic functions to quadratic order in $\omega_0\tau_d$.
653: We obtain
654: \begin{equation}
655:  \langle t_{\mathrm QD}(\epsilon)\rangle \simeq-\Gamma 
656: \int_0^{\infty}\frac{{d}\tau}{\hbar}
657: {\mathrm{e}}^{[i(\epsilon-\epsilon_0)-\Gamma/2]\tau/\hbar
658: -(eE\Delta x_{\mathrm th}\tau)^2/2\hbar^2}, \label{exp}
659: \end{equation}
660: where $\Delta x_{\mathrm th}=\sqrt{(2\overline{n}+1)(\hbar/2m\omega_0)}$ is the
661: thermal position uncertainty of the cantilever and we have taken the 
662: geometrical factor in the coupling constant (\ref{coupling}) to be unity for simplicity.
663: The integral on the righthand side  can now be evaluated to obtain
664: \begin{equation}
665:  \langle t_{\mathrm QD}(\epsilon)\rangle \simeq -\sqrt{\frac{\pi}{2}}
666: \left(\frac{\Gamma}{eE\Delta x_{\mathrm th}}\right)
667: \exp\left(\frac{[\Gamma/2-i(\epsilon-\epsilon_0)]^2}
668: {2(eE\Delta {x_{\mathrm th}})^2}\right){\mathrm{Erfc}}
669: \left(\frac{\Gamma/2-i(\epsilon-\epsilon_0)}{\sqrt{2} eE\Delta x_{\mathrm th} }\right).
670: \end{equation}
671: 
672: If we ignore, for the moment, the thermal width of the electron energy
673: distribution in the leads, then we can deduce from this expression a 
674: rough criterion for dephasing 
675: in the region close to the electronic  resonance:
676: \begin{equation}
677: eE\Delta x_{\mathrm th}> \Gamma,
678: \label{dephasecond1}
679: \end{equation}
680: or written in another way,
681: \begin{equation}
682: eE\tau_{d}>\frac{\hbar}{\Delta x_{\mathrm th}}.
683: \label{dephasecond2}
684: \end{equation}
685: Practical considerations (see Sec. VI) restrict the dwell time to be 
686: a few ns or less and place an upper bound on the electric field: $E \sim 10^5$V/m. 
687: Thus, the destruction of the 
688: interference fringes requires
689:  $\Delta x_{\mathrm th}>10^{-2}$\AA, a value which is  certainly  achievable.
690: 
691: As previously discussed, 
692: the loss of fringe visibility is not only associated with which--path 
693: detection, but with
694: thermal smearing as well.  
695: We can obtain the condition for which--path detection 
696: by setting the cantilever temperature to zero in Eq. 
697: (\ref{dephasecond2}):
698: \begin{equation}
699: eE \tau_d >\frac{\hbar}{\Delta x_{\mathrm zp}}=2\Delta p_{\mathrm zp},
700: \label{whichpathcond}
701: \end{equation}
702: where $\Delta x_{\mathrm zp}=\sqrt{\hbar/2m\omega_{0}}$ denotes the 
703: zero-point position uncertainty. Because the reduction in the 
704: transmission amplitude due to each 
705: different coherent state is the same [c.f., Eq. (\ref{ephi})], this condition holds 
706: independently of which coherent state the cantilever is in.
707: Thus, we find that for which--path detection to occur the
708: classical impulse delivered to the cantilever during the dwell time must exceed
709: twice the zero--point momentum uncertainty of the cantilever.
710: 
711: 
712: 
713: Our result is  equivalent to that obtained by Bohr in his
714: famous discussion of Einstein's
715: sliding--slit {\emph{gedanken}} experiment.\cite{hist} In that case Bohr argued
716: that in order to
717: detect the passage of an electron through a given slit, the momentum 
718: uncertainty in the
719: slit must be less than the impulse transferred by the passing
720: electron, thus necessitating a corresponding 
721: latitude  in position of the slits (via the uncertainty principle)
722: which in turn  washes out the fringes by causing large phase shifts 
723: (i.e., $\sim 2\pi$) for
724: successive electrons. However, these phase shifts are not associated with
725: any kind of thermal fluctuation, instead they arise from the position uncertainty
726: of the quantum state of the slits.  
727: 
728: 
729:   
730: 
731: At finite temperatures we must take into account
732: not only the thermal state of the cantilever, but also the fact that
733:  the electron energies are  spread over a range $\sim 4k_{\mathrm{B}}T$
734: around the Fermi energy. We must therefore average
735: the transmission amplitude over energy, weighted by the derivative of the
736: Fermi distribution function [see Eq. (\ref{therm})]:
737: \begin{equation}
738: \langle \tilde{t}_{\mathrm QD} \rangle =  \int_{0}^{\infty}{d}\epsilon
739: \frac{\langle t_{\mathrm QD}(\epsilon) \rangle}{4k_{\mathrm{B}}T}
740: {\mathrm{sech}}^2\left( \frac{\epsilon-\epsilon_0}
741: {2k_{\mathrm{B}}T}\right),
742: \end{equation}
743: where we have assumed that the bias across the device is small enough for a 
744: linear response approach to be valid and that the average Fermi energy in the 
745: leads is tuned to the resonance $\epsilon_0$.
746: The effect of this procedure is to reduce both the coherent
747: transmission amplitude itself and the influence the cantilever has on
748: it. The explanation for this unexpected feature can be found by comparing the
749:  transmission amplitude, $\langle t_{\mathrm QD}(\epsilon)\rangle$, at and away 
750:  from the 
751: resonance. Whilst the transmission amplitude at, or close to,  resonance 
752: decays rapidly with increasing 
753: coupling to the cantilever, the situation is reversed when the electron energy
754:  is far from resonance, where the transmission amplitude is {\emph{enhanced}} by
755: the interaction with the cantilever. The reduction in  dephasing efficiency 
756: due to the thermal averaging over the electron 
757: energy distribution makes an interesting contrast to the effect of the 
758: thermal average over cantilever states, which increases the dephasing rate 
759: substantially.
760:       
761: 
762: Fig.~1 illustrates the effect of increasing the dimensionless coupling constant 
763: $\kappa=\lambda/\hbar\omega_0$ on the magnitude of the resonant transmission 
764: amplitude $|\langle t_{\mathrm QD}(\epsilon_0)\rangle|$, with
765: and without averaging over the thermal width of the electron energy
766: distribution in the leads. For this example, we have taken
767: $\omega_0=140$~MHz and $T=$20~mK (giving a thermal occupation 
768: $\overline{n}=18$), $m=8\times10^{-20}$~kg, and assumed 
769:  $\kappa$ to have a maximum value of about 3, consistent
770:  with the analysis of the cantilever--dot coupling in appendix A. 
771:  The effect of the thermal broadening of electron energies in reducing 
772:  the efficiency of the cantilever to cause dephasing is clear.
773:  
774:  It is important to note  that our  
775: treatment is entirely restricted to a one--electron picture of transport. 
776: Such an approach is valid so long as the electron gas in the leads and the dot
777: itself is non-degenerate. When the electron temperature drops towards 
778: zero the interactions between electrons can no longer be ignored and so our
779: model and the results which follow from it become inapplicable. At 
780: $T=0$ and
781: in the weak bias limit, 
782: inelastic transport through the dot would become impossible: with the absence of 
783: unoccupied states below the Fermi level and  the cantilever in 
784: the ground state an electron is unable to exchange energy quanta with 
785: the cantilever.
786:  
787: 
788: \subsection{Regime Where $\omega_0\tau_d\sim 1$}           
789: 
790: In the regime where the dwell time of the electron on the dot
791: approaches the period of the cantilever, the system takes on
792: a rather different character, as the dephasing interaction between the
793: cantilever and the electron on the dot is no longer a brief scattering
794: event, but a sustained interaction. In this regime we would 
795: expect to see evidence of the periodic behavior of the cantilever,
796: such as side resonances in the elastic transmission amplitude, but
797: in practice these would have to compete with thermal effects which tend to wash out
798: fine structure in the transmission characteristics.
799: 
800: In order to calculate the transmission amplitude, 
801: it is convenient to recast our earlier expression (\ref{tqd}) 
802: in the form
803: \begin{eqnarray}
804: \langle t_{\mathrm QD}(\epsilon) \rangle&=& -\frac{\Gamma}{\hbar\omega_0}
805: \frac{{\mathrm{e}}^{-(\lambda/\hbar\omega_0)^2(2\overline{n}+1)}}
806: {1-{\mathrm{e}}^{-(2\pi/\omega_0)[i(\epsilon_0-(\lambda^2/\hbar\omega_0)
807: -\epsilon)/\hbar)+\Gamma/2\hbar]}}
808: \nonumber \\ 
809: &&\times \int_0^{2\pi}{d}\tau'
810: \exp\Biggl\{\frac{-\tau'}{\omega_0}\left[\frac{i}{\hbar}
811: \left(\epsilon_0-\frac{\lambda^2}{\hbar\omega_0}-\epsilon \right)
812: +\frac{\Gamma}{2\hbar}  \right]\nonumber \\       
813: &&-i\left(\frac{\lambda}{\hbar\omega_0}\right)^2 \sin (\tau')
814: +(2\overline{n}+1)\left(\frac{\lambda}{\hbar\omega_0}\right)^2 
815: \cos (\tau')\Biggr\},\label{calc}
816: \end{eqnarray}
817: where $\tau'=\omega_0 \tau$. Of course, to model a practical experiment, we also 
818: need to average over the energy, $\epsilon$, to take into
819: account the effect of electron temperature. However, we
820:  look first at the dependence of the transmission amplitude on the incident energy. 
821:  Although this  would require `monochromatic' 
822: electrons,  important insight is gained into the behavior of the system
823: which in practice may be obscured by thermal effects.
824: 
825: In Fig.~2, the magnitude of the resonant transmission amplitude is plotted
826: against the coupling $\kappa$ for $\Gamma/\hbar\omega_0$=2, 1, and
827:  0.5, and with $\overline{n}=18$. The behavior is similar in all three cases for 
828:  small $\kappa$, with a rapid decay in the magnitude of the transmission amplitude,
829:   due to dephasing caused by a combination of  which--path detection and
830:  thermal fluctuations in the state of the cantilever.
831:   Notice, however, that
832:  for $\Gamma/\hbar\omega_0$=1,  oscillations eventually develop in the 
833: transmission, and that these become even more pronounced 
834: for $\Gamma/\hbar\omega_0$=0.5. 
835: 
836: 
837: 
838: 
839: 
840: In order to clarify the origin of the oscillations in
841:  $|\langle t_{\mathrm QD}(\epsilon)\rangle|$,
842:  Fig.~3 plots the evolution of the amplitude and phase of the
843: transmission amplitude as functions of both the coupling constant,
844: $\kappa$, and the detuning energy $\epsilon-\epsilon_0$, for 
845: $\Gamma/\hbar \omega_0=0.5$ and $\overline{n}=18$.  
846: It is now clear that there are side resonances at
847: $\epsilon-\epsilon_{0}+\lambda^2/\hbar\omega_0=
848: \pm\hbar\omega_0,\pm2\hbar\omega_0,...$, and that
849: these resonances drift in energy as the value of $\kappa$
850: is increased.\cite{W89,JW} 
851: It is this drift  which is responsible for the oscillations
852:  in $|\langle t_{\mathrm QD}(\epsilon)\rangle|$ as a function of the coupling, $\kappa$, 
853: for given fixed $\epsilon$. 
854: 
855: Note that if the cantilever temperature is set equal to 
856: zero, i.e., it is in 
857: its ground state, 
858: $\overline{n}=0$, then side resonances are still observed, but only on one 
859: side of the main peak, at $\epsilon-\epsilon_{0}+\lambda^2/\hbar\omega_0
860: =+\hbar\omega_0,
861: +2\hbar\omega_0,...$.This is because in its ground state the cantilever  
862: can only absorb energy quanta.
863: 
864: 
865:  
866: Under the conditions $k_{\mathrm B}T\gg\hbar\omega_{0}$, 
867: $\kappa\sim 1$, and near resonance, Eq. 
868: (\ref{calc}) has the following asymptotic approximation:
869: \begin{equation}
870:     \langle t_{\mathrm QD}(\epsilon) \rangle 
871:     \sim 
872:     -\frac{\Gamma}{2\lambda}\sqrt{\frac{\pi\hbar\omega_{0}}{k_{\mathrm 
873:     B}T}}
874:     \left\{\frac
875: {1+{\mathrm{e}}^{-(2\pi/\hbar\omega_0)[i(\epsilon_0-\lambda^2/\hbar\omega_0
876: -\epsilon)+\Gamma/2\hbar]}}{1-{\mathrm{e}}^{-(2\pi/\hbar\omega_0)
877: [i(\epsilon_0-\lambda^2/\hbar\omega_0
878: -\epsilon)+\Gamma/2\hbar]}}\right\}.\label{asympt}
879: \end{equation} 
880: The various above-discussed features in the transmission 
881: amplitude dependence on coupling and energy can be clearly seen in 
882: this simplifying approximation.   
883: 
884: 
885: Side resonances are a familiar feature in tunneling problems, but unlike those
886: found here they are usually 
887: associated with inelastic processes. However, as was  first
888: shown by Jauho and Wingreen,\cite{JW} who considered the classical--oscillator
889: version of the problem we address, such resonances can also occur in the
890: elastic transmission amplitude.  Side resonances in the elastic transmission 
891: indicate
892: coherent or virtual exchange of energy quanta between the electron on the dot 
893: and the cantilever, with no net energy interchange over the dwell 
894: time of the electron.  
895: In a fully quantum mechanical system, these processes must rely on the coupled 
896: cantilever--electron system maintaining its phase coherency and so we may expect 
897: that the influence on the cantilever of its environment should be detectable, 
898: at least in principle,  via its effect on these resonances. 
899: 
900: The side resonances have a separation in energy of $\hbar\omega_0$, but  
901: for any practically realizable system the thermal energy scale will be 
902: much larger, as we discuss in detail in Sec. VI. However, in  
903: the regime where $\hbar\omega_0\ll k_{\mathrm{B}} T$,
904: averaging over the thermal distribution of electron energies will wash 
905: out the side resonances.
906:  This means that it would not be
907: possible to find any trace of the coherent exchange of energy quanta between
908: the cantilever and the electron on the dot in the transmission 
909: characteristics. Only if the
910: electron energy width could be lowered or the frequency raised to the
911: point where $\hbar\omega_0\sim k_{\mathrm{B}}T$ would we then expect
912: such features to be visible. Note, in the case of photon assisted tunneling the photon
913: energy can be made comparable with $k_{\mathrm{B}} T$  
914: without difficulty.\cite{JW}
915: 
916: 
917: \section{Influence of Environment on Cantilever}
918: Thus far, we have treated the cantilever--dot system as isolated,
919: thereby ignoring the effect of the environment on their degrees of
920: freedom. This approximation is valid when the dwell time of the electron on 
921: the dot is  short compared to the decoherence time of the 
922: cantilever, 
923: as will most certainly be the case for $\omega_0\tau_d\ll1$. However, there is no 
924: reason  to assume
925: that this will also be the case when $\omega_0\tau_d\sim1$.  Indeed, we
926: would like to know how the environment of the cantilever affects the coherent 
927: oscillations which occur in the transmission amplitude as a function of the 
928: coupling strength  (for monochromatic electrons). Intuitively, we expect that the 
929: decoherence arising from the cantilever's environment  should wash out the side 
930: resonances in the transmission amplitude, but a detailed calculation is required 
931: to determine  the cantilever quality ($Q$) factor range
932:  for this to 
933: occur.  
934: 
935: 
936: In general,
937: both the interaction between the dot electron and its local environment (other
938: electrons, phonons, photons, etc. in the dot region) and the interaction
939: between the cantilever and its environment (the other collective vibrational modes,
940: as well as internal electronic degrees of freedom, external photons, 
941: gas molecules, etc.) will
942: contribute to the decoherence. However, since we can always measure the
943: properties of the fringes with the dot--cantilever interaction turned
944: off for an arbitrary dwell time ($\tau_d=\hbar/\Gamma$), we need only consider
945: the additional effect of the cantilever's environment, as the electron's environment
946: can be included via a renormalisation of the zero electric field
947: transmission amplitude.
948: 
949: 
950: 
951: \subsection{Estimate of Cantilever Decoherence Time}
952: 
953: We can obtain a rough estimate of the time--scale over which superpositions of 
954: cantilever states, resulting from the coupling to the dot electron, 
955: are likely to be decohered by the environment by using
956: a simple, heuristic approach which models the environment as an infinite set of 
957: harmonic oscillators.\cite{UZ,Bose} This 
958: approach leads to the prediction that for
959: a system with  a classical damping rate, $\gamma_c$, a linear superposition of
960:  two different coherent
961: states whose centers are a distance $\Delta x$ apart, where $\Delta x$
962: is greater than the thermal
963: de Broglie wavelength, $\lambda_{\mathrm th}=\hbar/\sqrt{2m k_{\mathrm{B}}T}$,
964: will decohere at a rate $\gamma_d$, given by
965: \begin{equation}
966: \gamma_d\sim\frac{2m\gamma_c k_{\mathrm{B}}T(\Delta x)^2}{\hbar^2}.\label{rot}
967: \end{equation}
968: 
969: However, we cannot apply this heuristic rule directly to the coupled 
970: cantilever--dot system as the displacement of the cantilever is a continuously 
971: varying function. An estimate for the decoherence rate of the cantilever can 
972: only be obtained using this method if the further approximation of 
973: averaging over the cantilever displacement is made.\cite{Bose} 
974: If the cantilever starts in any given 
975: coherent state when the electron arrives on the dot, then the
976: state of the cantilever at time $t$ later will be a different coherent state
977: centered a distance $\Delta x$ apart from the point on which the initial 
978: state was centred, where 
979: \begin{equation}
980: \Delta x(t)=\sqrt{\frac{2\hbar}{m\omega_0}}\ 
981: \kappa[1-\cos(\omega_0 t)].
982: \end{equation}
983: Thus we can obtain an estimate for the decoherence rate $\gamma_{d}$ of the 
984: cantilever due to its
985: environment for the case where $\tau_d\omega_0=1$ by averaging this displacement 
986: over the cantilever period:\cite{Bose}
987: \begin{equation}
988: \gamma_d \sim\frac{1}{(2\pi/\omega_0)}\frac{4\kappa^2\gamma_c k_{\mathrm{B}}T}
989: {\hbar\omega_0}\int_0^{2\pi/\omega_0}{d}t[1-\cos(\omega_0 
990: t)]^2
991: =6\kappa^2\gamma_c\frac{k_{\mathrm{B}}T}{\hbar 
992: \omega_0}.\label{decoherence}
993: \end{equation} 
994: For a nanotube cantilever of frequency $\sim140$MHz (see appendix A), 
995: the $Q$--factor\cite{vib}
996: can be of order $500$ and so the classical damping rate, $\gamma_c$, will be 
997: of order $3\times 10^5$~$\mathrm{s}^{-1}$. Thus, at a temperature of 
998: 20~mK, our
999: heuristic expression for the decoherence rate of the nanotube cantilever  gives 
1000: $1/\gamma_d\sim 3\times 10^{-9}$~s for $\kappa\simeq 3$. This result 
1001: signals that  the decoherence of the cantilever due to its environment has an effect
1002: whenever the dwell time approaches a magnitude of order $1$ns.
1003: 
1004: Whilst the heuristic model we have outlined is very useful for
1005: estimating whether or not the cantilever's environment is relevant for
1006: the calculation of the transmission amplitude, it is not expected to
1007: be very accurate, as it is an approximation even of the rule--of--thumb given 
1008: by   Eq. (\ref{rot}).
1009: In order to improve on this estimate, 
1010: we must  enlarge our simple model by adding to the system 
1011: Hamiltonian (\ref{Ham1}) the cantilever's 
1012: environment, modelled as an infinite bath of harmonic oscillators with 
1013: linear coupling to the cantilever.  
1014: Modelling the environment in this way is of course itself a fairly
1015: serious approximation. However, this approximation is ubiquitous in
1016: one form or another throughout the theory of open quantum 
1017: systems.\cite{CL,UZ,O,JZ} By extending our calculation to include this
1018:  model of the cantilever's environment, we can obtain predictions
1019: for how the transmission properties of our which--path device depend
1020: on the decoherence rate of the cantilever. Thus, we can use our
1021: theory to predict how changing the $Q$--factor of the cantilever
1022: affects the interference fringes.         
1023: 
1024: \subsection{Transport Properties with Environmental Coupling}
1025: 
1026: The standard model of the environment which we use is  an infinite bath of harmonic 
1027: oscillators which interact linearly with the cantilever, but do not interact 
1028: with each other. We have to assume an infinite bath of oscillators in order to 
1029: have a reservoir which remains in thermal equilibrium despite contact with the 
1030: cantilever. The infinite oscillator--bath model of the environment is also 
1031: equivalent to a quantum Langevin formalism in which the cantilever operators 
1032: have an equation of motion containing a damping term and a thermal noise 
1033: operator arising from the reservoir. 
1034: 
1035: We begin by considering just the interaction between the cantilever and its
1036: environment, before going on to include the effect of an electron on the dot and
1037: hence obtaining the transmission amplitude including the environment.
1038: The cantilever-environment Hamiltonian can be written as\cite{MW,Lou}
1039: \begin{equation}
1040: {\mathcal{H}}^{c}=\hbar \omega_0\hat{a}^{\dagger}\hat{a}
1041: +\sum_{\omega} \hbar \omega 
1042: \hat{A}^{\dagger}(\omega)\hat{A}(\omega) 
1043: +\sum_{\omega}[g(\omega)\hat{a}^{\dagger}\hat{A}(\omega) + 
1044: g^*(\omega)\hat{A}^{\dagger}(\omega)\hat{a}], \label{envham}
1045: \end{equation}
1046: where the cantilever states are again operated on by 
1047: $\hat{a}$, the bath oscillator of frequency $\omega$ is operated on by
1048:   $\hat{A}(\omega)$ and the coupling constants $g(\omega)$ depend on
1049:   the bath oscillator frequency.
1050: 
1051:  Using the Hamiltonian ${\mathcal{H}}^c$, the  equation 
1052: of motion for $\hat{a}$ in the interaction picture is 
1053: readily obtained,\cite{MW}
1054: \begin{equation}
1055: \dot{\hat{a}}(t)=(-i \omega-\gamma)\hat{a}(t)-\hat{F}(t),
1056: \end{equation}
1057: where 
1058: \begin{equation}
1059: \hat{F}(t)=i\sum_\omega g(\omega)\hat{A}(\omega,0){\mathrm{e}}^{-i \omega t}
1060: \end{equation}
1061: and the damping coefficient, $\gamma$, is given by
1062: \begin{equation}
1063: \gamma=\pi \eta_{b}(\omega_0)|g(\omega_0)|^2,
1064: \end{equation}
1065: where $\eta_{b}(\omega)\delta \omega$ is the number of bath oscillators in the 
1066: spectral range $\delta\omega$. The coefficient $\gamma$ provides the bridge 
1067: between the model and experiment, as it is simply related to the rate at which the 
1068: system loses energy after being excited (i.e., the classical damping rate),\cite{Lou}
1069: $\gamma_c=2\gamma$.
1070: 
1071: 
1072: Integration of the equation of motion for $\hat{a}(t)$, and a similar one for 
1073: $\hat{a}^{\dagger}(t)$, leads to explicit expressions for 
1074: $\hat{a}(t)$ and $\hat{a}^{\dagger}(t)$:
1075: \begin{equation}
1076: \hat{a}(t)=\hat{a}(0){\mathrm{e}}^{(-i\omega_0-\gamma)t}-{\mathrm{e}}
1077: ^{(-i\omega_0- 
1078: \gamma)t}\int_0^t dt' \hat{F}(t'){\mathrm{e}}^{(i\omega_0+\gamma)t'}, 
1079: \label{plaina}
1080: \end{equation} 
1081: \begin{equation}
1082: \hat{a}^{\dagger}(t)=\hat{a}^{\dagger}(0){\mathrm{e}}^{(i\omega_0-\gamma)t}-
1083: {\mathrm{e}}^{(i 
1084: \omega_0-\gamma)t}\int_0^t dt' 
1085: \hat{F}^{\dagger}(t'){\mathrm{e}}^{(-i\omega_0+\gamma)t'}. \label{adag}
1086: \end{equation}  
1087: 
1088: The time $t=0$ holds a special place in the theory since it is the time  at 
1089: which the interactions between the cantilever and the bath are turned on and 
1090: so the cantilever and the bath are apparently quite independent at this instant.
1091:  However, this  need not be the case and certainly would not be appropriate 
1092:  for the system  we are considering. In
1093: order to specify the model completely we need to give the expectation values 
1094: at $t=0$ for both the cantilever and the bath. If we set the initial 
1095: expectation values of the cantilever to be those of a thermal state at the
1096: same temperature as the bath, then we can describe the case in which the
1097: cantilever has been interacting with the bath for a time much longer than the 
1098: damping time, $1/\gamma_c$, {\emph{before}} $t=0$ and so is in equilibrium with
1099: the bath. Our only assumption is that
1100: quantum correlations between the cantilever and the bath can be neglected
1101: at $t=0$, the usual assumption made in calculating decoherence rates. 
1102: 
1103: 
1104:  
1105:  The expectation values are over  
1106: products of  cantilever and environment states.  We eventually want to calculate the 
1107: transmission amplitude for an electron interacting with
1108: a cantilever which is initially in a thermal state
1109: and so we choose to define the $t=0$ expectation values as:
1110: \begin{equation}
1111: \langle...\rangle=\frac{{\mathrm{Tr}}[...{\mathrm{e}}
1112: ^{-{\mathcal{H}}^c_0/k_{\mathrm{B}}T}]}
1113: {{\mathrm{Tr}}[{\mathrm{e}}^{-{{\mathcal{H}}^c_0}/k_{\mathrm{B}}T}]},
1114: \end{equation}
1115: where the Hamiltonian without interactions is 
1116: \begin{equation}
1117: {\mathcal{H}}^c_0=\hbar \omega_0\hat{a}^{\dagger}\hat{a}
1118: +\sum_{\omega} \hbar \omega 
1119: \hat{A}^{\dagger}(\omega)\hat{A}(\omega),
1120: \end{equation}
1121: and $T$ defines the fixed temperature of the environment.
1122: Using this definition we find
1123: \begin{eqnarray}
1124: \langle \hat{A}(\omega,0)\rangle&=&0 \nonumber \\
1125: \langle \hat{A}(\omega,0)\hat{A}(\omega',0)\rangle&=&0 \\
1126: \langle \hat{A}^{\dagger}(\omega,0)\hat{A}(\omega',0)\rangle&=&\delta_{\omega 
1127: \omega'}N(\omega),\nonumber 
1128: \end{eqnarray}   
1129: with
1130: \begin{equation}
1131: N(\omega)=\frac{1}{{{\mathrm{e}}}^{\hbar \omega/k_{\mathrm{B}}T}-1}.
1132: \end{equation}
1133: 
1134: For the cantilever itself the values of the zero-time correlation functions 
1135: represent the initial conditions of the problem. In 
1136: thermal equilibrium with the environment at temperature $T$,   
1137: \begin{eqnarray}
1138: \langle \hat{a}(0)\rangle&=&0\nonumber \\
1139: \langle \hat{a}(0)\hat{a}(0)\rangle&=&0 \\
1140: \langle \hat{a}^{\dagger}(0)\hat{a}(0)\rangle&=&\overline{n},\nonumber
1141: \end{eqnarray}   
1142: with
1143: \begin{equation}
1144: \overline{n}\equiv N(\omega_0)=\frac{1}{{\mathrm{e}}^{\hbar 
1145: \omega_0/k_{\mathrm{B}}T}-1}.
1146: \end{equation}
1147: 
1148: The purpose of extending the analysis to include the interactions between the
1149: cantilever and its environment is to see how they modify the 
1150: interaction between the cantilever and the electron {\emph{whilst it is on the dot}}. 
1151: Therefore we can choose our
1152: origin of  time, and hence the definition of the initial cantilever state, to 
1153: be the time when the electron jumps onto the dot. There is no need to explicitly
1154: include the interaction with the bath before the electron hops onto the dot as it is
1155: already implicitly included by assuming the cantilever is in a thermal state 
1156: at temperature $T$. However, when the electron is on the dot it drives 
1157: the cantilever away from equilibrium and we now need to include the 
1158: interaction with the environment explicitly to model the behavior 
1159: of the cantilever--dot system as accurately as possible. 
1160: 
1161: 
1162: 
1163: We obtain the transmission amplitude for the dot, including the 
1164: effect of the cantilever's environment by applying the $S$--matrix we used in
1165:  Sec. III to a generalization of the Hamiltonian (\ref{Ham1})
1166:  which includes the coupling of the cantilever to 
1167:  the bath of oscillators,
1168: \begin{equation}
1169: {\mathcal{H}}=\epsilon_0\hat{c}^{\dagger}\hat{c}+{\mathcal{H}}^c
1170: +{\mathcal{H}}_1+\sum_k\left( 
1171: \epsilon_{kL}\hat{c}^{\dagger}_{kL}\hat{c}_{kL}+\epsilon_{kR}\hat{c}^{\dagger}_
1172: {kR}\hat{c}_{kR}\right), 
1173: \end{equation}
1174: where ${\mathcal{H}}_1$, given by eqn (\ref{coup}), describes the 
1175: electron--cantilever coupling and the cantilever--environment part, ${\mathcal{H}}^c$, 
1176: is given by Eq.\ (\ref{envham}). 
1177:  
1178: 
1179: Using the methods of Glazman and Shekhter,\cite{GS} we can again separate out 
1180: the electronic 
1181: part of the transmission amplitude from the average over cantilever states, 
1182: and so we find
1183: \begin{eqnarray}
1184: \langle t_{\mathrm QD}(\epsilon) \rangle &=& 
1185: -\Gamma\int_0^{\infty}\frac{dt}{\hbar}{\mathrm{e}}^{[i(\epsilon-\epsilon_0)
1186: -\Gamma/2]t/\hbar}
1187: \langle{\mathrm{T}}_t 
1188: {\mathrm{e}}^{-i \int_0^{t}W_I(t')dt'/\hbar} \rangle.
1189: \end{eqnarray}
1190: The term in angled brackets on the right--hand side is known as the influence
1191:  functional, 
1192: ${\mathrm{T}}_t$ is the time ordering operator, and $W_I(t)$  is the
1193:  electron--cantilever coupling defined in the interaction picture 
1194: \begin{equation}
1195: W_I(t)=-{\mathrm{e}}^{i{\mathcal{H}}^c t/\hbar}
1196: \lambda(\hat{a}^{\dagger} +\hat{a})
1197: {\mathrm{e}}^{-i{\mathcal{H}}^c t/\hbar}
1198: =-\lambda[ \hat{a}^{\dagger}(t) +\hat{a}(t)],
1199: \end{equation}
1200: where $\hat{a}(t)$ and $\hat{a}^{\dagger}(t)$ are given by Eqs. (\ref{plaina}) and 
1201: (\ref{adag}) above.
1202: 
1203: 
1204: Here, we shall
1205: carry out a much simpler, approximate calculation of the influence 
1206: functional which involves performing a second order expansion in 
1207: $\lambda$ and then re-exponentiating. In appendix B, it is shown that 
1208: this approximate calculation in fact coincides with the exact result 
1209: obtained from a full calculation. 
1210: 
1211: The expansion to second order gives 
1212: \begin{equation}
1213: \langle{\mathrm{T}}_t {\mathrm{e}}^{-i \int_0^{t}W_I(t')dt'/\hbar} \rangle=
1214: 1-i\int_0^t \frac{dt'}{\hbar}\langle W(t')\rangle 
1215: +(-i)^2\int_0^t \frac{dt'}{\hbar}\int_0^{t'}\frac{dt''}{\hbar}\langle 
1216: W(t')W(t'')\rangle. \label{if}
1217: \end{equation}  
1218: The next step is to evaluate the correlation 
1219: functions which arise in the first and second order terms in the 
1220: expansion of the influence functional. At first order
1221: \begin{equation}
1222: \langle W(t') \rangle=-\lambda\langle  [\hat{a}(t')+\hat{a}^{\dagger}(t')]\rangle,
1223: \end{equation}
1224:  and at second order
1225: \begin{equation}
1226: \langle W(t')W(t'')\rangle = \lambda^2\{ \langle 
1227: \hat{a}(t') \hat{a}(t'')\rangle+ \langle \hat{a}(t') 
1228: \hat{a}^{\dagger}(t'')\rangle   
1229: +\langle \hat{a}^{\dagger}(t') \hat{a}(t'')\rangle + \langle 
1230: \hat{a}^{\dagger}(t') \hat{a}^{\dagger}(t'')\rangle \},
1231: \end{equation} 
1232: where $t'\geq t''$.
1233: 
1234: We can calculate all the correlation functions 
1235:  using the initial conditions defined above and the standard results of 
1236: quantum Langevin theory.\cite{MW,Lou}
1237: We  start by observing that three of the five correlation functions are 
1238: zero:
1239: \begin{equation}
1240: \langle \hat{a}(t')\hat{a}(t'')\rangle=\langle 
1241: \hat{a}^{\dagger}(t')\hat{a}^{\dagger}(t'')\rangle=0 \label{a}
1242: \end{equation}
1243: and
1244: \begin{equation}
1245: \langle [\hat{a}(t')+\hat{a}^{\dagger}(t')]\rangle=0.
1246: \end{equation}
1247: These results follow from the definitions above, since correlators of the type 
1248: $\langle \hat{a}(0)\hat{F}(t) \rangle$  decouple into products of one-time 
1249: correlation functions and the definitions imply that $\langle \hat{F}(t) 
1250: \rangle=\langle \hat{F}(t')\hat{F}(t'') \rangle=0$. 
1251: For the other two correlation functions, we have
1252: \begin{eqnarray}
1253: \langle \hat{a}(t')\hat{a}^{\dagger}(t'')\rangle&=&  (\overline{n}+1)
1254: {\mathrm{e}}^{-i\omega_0 (t'-t'')} {\mathrm{e}}^{-\gamma (t'-t'')} 
1255: \label{b} \\
1256: \langle \hat{a}^{\dagger}(t')\hat{a}(t'')\rangle&=&  \overline{n}
1257: {\mathrm{e}}^{i\omega_0 (t'-t'')} {\mathrm{e}}^{-\gamma (t'-t'')} \label{c}.
1258: \end{eqnarray}
1259: Thus, we again find
1260: \begin{equation}
1261: \langle t_{\mathrm QD}(\epsilon) \rangle=-\Gamma\int_0^{\infty}\frac{dt}{\hbar}
1262: {\mathrm{e}}^{[i(\epsilon-\epsilon_0)
1263: -\Gamma/2]t/\hbar}{\mathrm{e}}^{-\phi(t)}, 
1264: \label{exact}
1265: \end{equation} 
1266: but now with 
1267: \begin{equation}
1268: \phi(t)=\frac{\lambda^2}{\hbar^2} \int_0^t 
1269: {dt'}\int^{t'}_{0}{dt''}{{\mathrm{e}}}^{-\gamma (t'-t'')}\left[ (\overline{n}+1)
1270: {\mathrm{e}}^{-i\omega_0 (t'-t'')}+\overline{n}
1271: {\mathrm{e}}^{i\omega_0 (t'-t'')}\right], 
1272: \label{phi}
1273: \end{equation}
1274: where we have re-exponentiated the expansion in $\lambda$.
1275: 
1276: 
1277: 
1278: Carrying out a change of variables to $\tau=t'-t''$, the integrals are readily 
1279: evaluated to give
1280: \begin{equation}
1281: \phi(t)=\left(\frac{\lambda}{\hbar}\right)^2\left[ 
1282: \frac{(\overline{n}+1)}{(\gamma+i\omega_0)}\left(t+\frac{{\mathrm{e}}^{-(\gamma+i
1283: \omega_0)t}-1}{\gamma+i\omega_0}\right)+\frac{\overline{n}}{(\gamma-i\omega_0)}
1284: \left(t+\frac{{\mathrm{e}}^{-(\gamma-i\omega_0)t}-1}{\gamma-i\omega_0}\right)
1285: \right]. \label{phix}
1286: \end{equation}
1287: Setting $\gamma=0$, one may verify that (\ref{phix}) indeed reduces 
1288: to expression (\ref{ex2}) in the absence of the environment.
1289: 
1290: Under the conditions $Q\gg 1$, $k_{\mathrm B}T\gg\hbar\omega_{0}$, 
1291: $\kappa\sim 1$, and near resonance, the transmission amplitude 
1292: (\ref{exact}) has the following asymptotic approximation:
1293: \begin{equation}
1294:     \langle t_{\mathrm QD}(\epsilon) \rangle 
1295:     \sim 
1296:     -\frac{\Gamma}{2\lambda}\sqrt{\frac{\pi\hbar\omega_{0}}{k_{\mathrm 
1297:     B}T}}
1298:     \left\{\frac
1299: {1+{\mathrm{e}}^{-(2\pi/\hbar\omega_0)[i(\epsilon_0-\lambda^2/\hbar\omega_0
1300: -\epsilon)+\Gamma/2\hbar+(\kappa^{2}\gamma_{c}k_{\mathrm 
1301: B}T/\hbar\omega_{0})]}}{1-{\mathrm{e}}^{-(2\pi/\hbar\omega_0)
1302: [i(\epsilon_0-\lambda^2/\hbar\omega_0
1303: -\epsilon)+\Gamma/2\hbar+(\kappa^{2}\gamma_{c}k_{\mathrm 
1304: B}T/\hbar\omega_{0})]}}\right\}.\label{asymptenv}
1305: \end{equation} 
1306: Note that this approximation differs from the earlier-derived one for 
1307: the transmission amplitude 
1308: in the absence of the environment, Eq. (\ref{asympt}), merely by the 
1309: replacement of $\Gamma/2\hbar$ with $\Gamma/2\hbar+(\kappa^{2}\gamma_{c}k_{\mathrm 
1310: B}T/\hbar\omega_{0})$. Thus, a decoherence rate $\gamma_{d}$ can be identified 
1311: as the term $\kappa^{2}\gamma_{c}k_{\mathrm B}T/\hbar\omega_{0}$, which 
1312: agrees with the earlier-derived estimate 
1313: (\ref{decoherence}) up to an overall numerical factor. 
1314: Approximation (\ref{asymptenv}) clearly shows the washing out of
1315: the coherent, oscillatory behavior by the 
1316: environment when the decoherence time $1/\gamma_{d}$ is shorter than 
1317: the cantilever period. When the decoherence time exceeds the dwell 
1318: time $\hbar/\Gamma$, then the former has a 
1319: negligible effect on the transmission properties.      
1320: 
1321: 
1322: 
1323: 
1324: \subsection{Results for Cantilever Coupled to Environment}
1325: The expression for the transmission amplitude including the coupling to
1326: the cantilever's environment [Eq.\ (\ref{exact})] can be integrated
1327: numerically, thereby allowing us to explore the effect on the
1328: transmission characteristics of varying the cantilever's
1329: $Q$--factor under more general 
1330: conditions than those for which approximation (\ref{asymptenv}) is 
1331: justified. Fig.~4 shows the magnitude 
1332: and phase of the
1333: transmission  amplitude through the dot for $\Gamma/\hbar \omega_0=0.5$,
1334: $\overline{n}=18$ and $Q=\omega_0/\gamma_{c}=50$. The diagram 
1335: should be compared with
1336: Fig.~3 showing the behavior of the same system without
1337: environmental coupling. Fig.~5 shows the behavior at resonance of
1338: $|\langle t_{\mathrm QD}(\epsilon)\rangle|$ for 
1339: $Q=50$ and 500, as well as the $\gamma_{c}/\omega_0 =0$  (no environment)
1340: case for comparison.
1341: 
1342: It is clear from the figures that the coupling to the environment tends to
1343: destroy the side resonances in the transmission amplitude, as well as
1344: the associated features in the phase. This is because the environment acts to 
1345: degrade the coherent 
1346: superposition of cantilever states into which the interaction with the electron tries
1347: to drive the cantilever. Furthermore, the figures show that these 
1348:  environmental effects 
1349: become increasingly important as the cantilever--electron coupling,
1350: $\kappa$, is increased. This is because the larger $\kappa$ is, the greater 
1351: is the 
1352:  difference between the 
1353: states in the superposition into which the cantilever is driven, and consequently 
1354: the faster is the rate at which the superposition decoheres.        
1355: 
1356:  
1357: 
1358: \section{Practical Considerations}
1359: 
1360: 
1361: The model parameter ranges actually allowed in an experiment  are limited by 
1362: practical constraints. Up until now we 
1363: have only referred to these very loosely and have concentrated instead on the
1364: range of behavior which can occur in the which--path system under fundamental 
1365: constraints alone.
1366: 
1367: Probably
1368: the most important practical constraint affects the upper range of the electric field
1369: which can be developed between the cantilever and the dot. 
1370: The maximum allowable field is typically $\sim 10^5$~V/m before 
1371: breakdown of the two--dimensional electron gas occurs due to  
1372: deconfinement,\cite{Buksp} and so we
1373: must take this as the largest possible value in considering the
1374: practicality of the system. 
1375: 
1376: The temperature of
1377:  the system is limited by the difficulty of cooling conduction electrons  
1378:  to ultralow temperatures. It becomes extremely difficult to reduce the 
1379:  electron temperature  below about 20~mK,
1380:  because the electrons become practically decoupled from acoustic 
1381:  phonons. 
1382:  Therefore we take 20~mK to be 
1383:  the temperature minimum. 
1384: 
1385: The frequency range of the cantilever is crucial 
1386: for observing quantum coherent behavior such as  side resonances, as 
1387: well as their destruction 
1388: due to 
1389: decoherence. The lower limit 
1390: on the frequency range is set by the requirement that the cantilever 
1391: period should be comparable to the dwell time $\tau_d=\hbar/\Gamma$ of the electron on the quantum dot,  
1392: while the 
1393: upper limit is set
1394:  by the requirement that the interaction between the cantilever and
1395: the electron on the dot be sufficiently strong, given the limits on
1396: the electric field,  to lead to decoherence,
1397: again on a time--scale  comparable to the dwell time.
1398: The interaction between
1399: the electron on the dot and the cantilever is discussed in more
1400: detail in appendix A, along with the geometrical factors which arise in
1401: the calculation of the effect of the electric field on the flexural cantilever
1402: modes.
1403: The dwell time  is
1404: limited by the rate at which processes other than the interaction 
1405: with the cantilever cause decoherence, as well as how small a current can be measured 
1406: through the dot. Of course, the decoherence rate
1407: of all the background processes is very difficult to estimate
1408: theoretically; in experiment it can be done by
1409: observing the visibility of the fringes as a function of dwell time
1410: with the cantilever interaction switched off. This then provides a
1411: baseline with which to compare all later measurements where we are
1412: interested in the effect of the cantilever. Previous experiments 
1413: carried out at $T=0.1$~K by
1414: Yacoby {\emph{et al.}},\cite{Yac} which were designed to measure the phase of
1415: the transmission amplitude through a dot in the absence of any
1416: external probe, show that dwell times as long 10~ns and currents as low as 
1417: $\sim10^{-11}$~A lead to fringes which are still detectable.
1418: 
1419: It turns out that the best compromise between the two 
1420: competing frequency limits is achieved by using say a carbon nanotube 
1421: cantilever with a frequency of 100--200~MHz, giving a maximum 
1422: coupling $\kappa=\lambda/\hbar\omega_{0}\sim 3$ (appendix A). The   
1423: problem is not so much in finding a cantilever with a high enough frequency, but rather
1424: in obtaining a large enough coupling from the electric field to cause a 
1425: detectable amount of dephasing.  
1426: 
1427: 
1428: 
1429: In the light of these practical constraints it is clear that only some 
1430:  of the theoretical results obtained in our
1431: analysis would be observable in an experiment using current
1432:  technology. The overall destruction of the
1433:  AB interference fringes as the coupling between the cantilever and the
1434:  electron on the dot is increased should be detectable, both in the
1435:  regime where $\Gamma/\hbar\omega_0\gg 1$ and $\Gamma/\hbar\omega_0\sim1$.
1436: However, the restrictions on the temperature and cantilever
1437:  frequency imply that an experiment would have to be performed in the regime
1438:  where $k_{\mathrm{B}}T\gg \hbar\omega_0$. This means that the thermal
1439:  width of electron energies will wash out the side resonances 
1440:  in the transmission which were found to occur when 
1441: $\Gamma/\hbar\omega_0\sim1$. Under these circumstances, the effect of
1442:  the cantilever's environment on the transmission characteristics
1443:  would be obscured. However, we emphasise again that these
1444:  limitations are not fundamental: if the width of the electron
1445:  energies could be reduced or if the restriction on the maximum
1446:  cantilever frequency could be relaxed (by finding a way to increase
1447:  the coupling between the electron on the dot and the cantilever, for
1448:  example) then the coherent, quantum electromechanical features which our
1449:  analysis predicts should be observable. 
1450:  Whilst it is not clear how
1451:  the coupling constant between the cantilever and the dot could be
1452:  increased, it is possible to reduce the thermal width of
1453:  electron energies by using a double quantum dot system 
1454:  rather than a single one as we have considered here.\cite{Buks2}  
1455:    
1456: It is also  possible to conduct experiments in which  an ac bias is applied 
1457: to the AB ring. Such an experiment would provide an alternative  way of 
1458: investigating
1459: the interaction between the cantilever and its environment: if the ac frequency 
1460: is higher than the rate $\gamma_{c}$ at which the cantilever state changes due to 
1461: thermal fluctuations then the thermally-induced fluctuations in the phase of the 
1462: interference fringes should  be detectable, and distinguishable from 
1463: the destruction of the interference fringes due to which-path 
1464: detection. However, a detailed analysis of such
1465: an experiment requires extending the theory developed here to include
1466: time dependence, and so goes well beyond the present analysis.
1467: 
1468: 
1469: \section{Conclusions and Discussion}
1470: 
1471: We have carried out a theoretical analysis of a possible solid state
1472: which--path interferometer, in which electronic and mechanical degrees of freedom 
1473: become coupled. The visibility and phase of the interference
1474: fringes in the system depend strongly on the coupling between the
1475: electron on the dot and the cantilever. The reduction in visibility 
1476: with increasing coupling is 
1477: due in part to
1478: the cantilever measuring the path taken by the electron and also due 
1479: to thermal
1480: fluctuations in the state of the cantilever.
1481: 
1482: When the dwell time of the electron on the dot is short compared to
1483: the period of the cantilever, the system behaves in a way which is
1484: analogous to Einstein's celebrated recoiling slit experiment. In
1485: contrast, when the dwell time is comparable to the cantilever's period,
1486:  the cantilever and the electron on the
1487: dot   show signs of behaving  as a single coherent quantum system,
1488:  so long as the electrons incident on the dot have a
1489: sufficiently narrow energy width.  The coherency of the
1490: cantilever--dot system is inferred from the appearance of side
1491: resonances in the transmission characteristics. Including  the cantilever's
1492:  environment in the analysis, we find that the side
1493:  resonances are washed out   for small enough  oscillator 
1494:  $Q$--factor, while the average decrease in the fringe visibility with
1495:  increasing coupling to the cantilever is not affected.
1496: 
1497: The basic feature of the reduction of fringe visibility as a function
1498: of the coupling between the electron on the dot and the cantilever
1499: should be observable in an experiment using currently available technology. 
1500: However, the 
1501: more delicate features   such as the
1502: side resonances in the transmission amplitude for long dwell times 
1503: and the effect of the 
1504:  cantilever's environment on these resonances will usually be
1505:  obscured by the  energy thermal width of electrons incident on the dot.
1506: 
1507: There are two important ways in which our analysis could be
1508: extended. Most straightforwardly, we could examine what effect using a
1509: double dot, rather than a single one, would have on the behavior of the
1510: system. In particular, it would be interesting to see to what extent the
1511: electron energy width could be reduced, whilst maintaing
1512: an overall, measurable current through the device. 
1513: The second way our
1514: analysis could be extended would be to go beyond our steady state
1515: treatment to obtain the time dependence of the transmission
1516: amplitude. A time dependent analysis would allow us to make predictions
1517: about the way in
1518: which the thermal fluctuations in the state of the cantilever cause
1519: fluctuations in the phase of the interference fringes. Such
1520: fluctuations may prove to be observable in ac experiments and may also
1521: provide us with another way of inferring information about how the
1522: cantilever interacts with its environment.
1523:   
1524: \acknowledgements
1525: M.P.B. would especially like to thank Eyal Buks for many helpful 
1526: conversations and invaluable suggestions. Very helpful conversations 
1527: with Sougato Bose are also gratefully acknowledged.
1528: A.D.A. thanks Angus MacKinnon for a series of very useful discussions.  Funding was provided by the 
1529: EPSRC under Grant No. GR/M42909/01.
1530: 
1531: \appendix
1532: \section{Cantilever}
1533: 
1534: The cantilever in the which--path experiment we propose must fulfil a
1535: number of quite stringent properties: it must have a fundamental 
1536: frequency in the range 100--200~MHz and it must also  
1537: be a conductor. These requirements can be satisfied conveniently
1538:  by using a carbon nanotube as a cantilever, rather than a device which has
1539: been fabricated via some kind of etching process from a much larger substrate. 
1540: 
1541: Carbon nanotubes have a number of remarkable physical properties which
1542: are beginning to exploited for practical purposes. Recent experiments
1543: have seen them employed as hyper--sensitive tips in AFM experiments,
1544:  with the nanotube attached to the end of a conventional AFM tip to
1545:  extend the effective range of resolution of the device.\cite{AFM} A similar
1546:  apparatus could be used to bring a nanotube into position to act as
1547:  the cantilever in our which--path experiment, either in the
1548:  geometrical configuration explicitly considered above or some
1549:  variation of it  which would lead to slightly different geometrical
1550:  factors, but not change the underlying linear form of the
1551:  cantilever--dot interaction.   
1552:  
1553: In this appendix we carry out an analysis of the coupling between the
1554:  cantilever
1555: and the electron on the dot, and the normal modes of a nanotube cantilever.  
1556: We verify the
1557:   linear dependence of the cantilever energy on the electric field assumed 
1558:  in the text and derive the form of the coupling constant, $\lambda$.
1559:   By examining the mode spectrum of a nanotube cantilever, we 
1560:  confirm the possibility of obtaining a cantilever with a fundamental flexural
1561:  mode of order $10^8$~Hz, and justify our assumption that for any given cantilever it 
1562:  is sufficient to consider just the lowest mode.  
1563: 
1564:  
1565: \subsection{Cantilever--Dot Coupling}
1566: 
1567: We begin by determining the relation between the
1568: maximum electric field at the surface of the dot
1569: and the voltage applied to the conducting cantilever, before analysing the 
1570: details of the effect of the electron on the cantilever energy.   
1571: 
1572: We will consider a cantilever of length $L\sim1\ \mu$m, positioned so
1573: that its tip lies over the centre of the dot and at a height $d\sim
1574: 0.1\ \mu$m.\cite{justify} When an extra electron is added to the dot it
1575: will cause a (classical) vertical displacement of the tip $z\ll d$. We treat the
1576: cantilever as a conducting needle, since it's length will be far greater than 
1577: either of the other two dimensions. An applied voltage induces a line
1578: charge density, $\sigma$, on the cantilever, but because the cantilever is 
1579: necessarily finite in length, the charge density is not entirely uniform and so
1580: for an exact treatment we should write the line charge density as $\sigma(x)$ 
1581: where $x$
1582: runs along the length of the cantilever from the tip ($x=0$). However,
1583: for a cantilever with a large enough aspect ratio, the charge density can be 
1584: approximated as constant with only a small error [the
1585: error is of order $\delta$ with $\delta^{-1}=2\ln(L/r)$, where $r$ is
1586: the radius, for a rod shaped cantilever\cite{jack}]. For the cantilever we are
1587: considering the radius may be as small as $1.5$nm, as we discuss
1588: below, and so the error in assuming a uniform charge density will be
1589: less than 10\%, which is acceptable as our aim is to obtain an order
1590: of magnitude estimate for the  interaction strength.    
1591: 
1592: The charge induced on the cantilever leads to an electric field at the
1593: surface of the dot, 
1594: \begin{equation}
1595: {\mathbf{E}}=\int_0^{L}\frac{\sigma(-x\hat{\mathbf{i}}+d\hat{\mathbf{j}})
1596: {d}x}{4\pi\epsilon_0(x^2+d^2)^{3/2}},
1597: \end{equation} 
1598: taking $\sigma=cV$ where $c$ is an unimportant  constant and $V$
1599: is the voltage applied to the cantilever. The unit vectors are defined
1600: so that $\hat{\mathbf{i}}$ runs along the cantilever away from its tip and
1601: $\hat{\mathbf{j}}$ points down from the tip towards the dot.
1602: Since $L\gg d$, we find
1603: \begin{equation}
1604: |{\mathbf{E}}|\simeq \frac{2^{1/2}cV}{4\pi\epsilon_0 d}.
1605: \end{equation}
1606: 
1607: We also need to know how the potential energy of the cantilever varies
1608: for small vertical displacements, $z$, about the equilibrium position,
1609: $d_0$, where $d=d_0-z$. For a section of the cantilever of
1610: length $\Delta x$ centered at $x$, the potential energy due to the
1611: displacement is
1612: \begin{equation}
1613: \Phi(x)=\frac{qcV\Delta x}{4\pi\epsilon_0\sqrt{(d_0-z)^2+x^2}},
1614: \end{equation}
1615: where $q=-e$ is the excess charge on the dot. Since $d_0\gg z$, we obtain 
1616: the $z$--dependent part of the potential energy as
1617: \begin{equation}
1618: \Phi_z(x)=z\frac{qcVd_0\Delta x}{4\pi\epsilon_0[d_0^2+x^2]^{3/2}}.
1619: \end{equation} 
1620: This is of course linear in $z$ as we anticipated in our model. However,
1621: in order to obtain the effective coupling between the electric field
1622: and each of the cantilever modes we need to rewrite the $z$--dependent
1623: part of the potential energy in terms of the normal modes of the
1624: cantilever and so we turn now to the mode spectrum of the cantilever.
1625: 
1626: 
1627:  
1628: \subsection{Cantilever Modes}
1629: 
1630: 
1631: A single--walled nanotube cantilever  can be obtained with lengths $\sim
1632: 1\ \mu$m, diameters $\sim 3$~nm, and a Young's modulus $\sim 1$~TPa.\cite{nano} 
1633: Recent experiments,\cite{vib,modes2} have shown that the
1634: flexural modes of such nanotubes have frequencies which lie in the MHz--GHz
1635: regime. We can treat a nanotube as a rigid hollow rod and so obtain
1636: the frequencies of the flexural modes,\cite{modes}  
1637: \begin{equation}
1638: \omega_i=\frac{\beta^2_i}{2L^2}\sqrt{ \frac{Y (a^2+b^2)}{\rho}},
1639: \end{equation}   
1640: where $a$ and $b$ are the outer and inner  diameters, $L$ the length,
1641: $Y$ the Young's modulus and $\rho$ the density of the tube. The factors
1642: $\beta_i$ arise from geometrical considerations and are the solutions
1643: of the equation $\cos(\beta_i)\cosh(\beta_i)=-1$. 
1644: 
1645: The energy of the cantilever can be written in terms of its classical normal
1646: modes as 
1647: \begin{equation}
1648: H=\sum_i\left(\frac{1}{2}m\omega^2_i q^2_i 
1649: +\frac{p^2_i}{2m}\right)I_i,
1650: \end{equation}
1651: where 
1652: \begin{equation}
1653: I_i=\frac{1}{L}\int_0^L\Gamma_i^2{d}x
1654: \end{equation}
1655: and
1656:  \begin{eqnarray}
1657: \Gamma_i(L-x)=&&[\cos(\beta_i)+\cosh(\beta_i)][\sin(\beta_i
1658: x/L)-\sinh(\beta_i x/L)]\nonumber\\
1659: &&-[\sin(\beta_i)+\sinh(\beta_i)][\cos(\beta_i x/L)-\cosh(\beta_i x/L)]
1660: \end{eqnarray}
1661: are the vibrational mode
1662: eigenfunctions.\cite{modes} If we modify the definitions of the
1663: canonical variables slightly to define $q_i'=q_iI_i^{1/2}$ and 
1664: $p_i'=p_iI_i^{1/2}$, then the Hamiltonian takes the conventional form 
1665: \begin{equation}
1666: H=\sum_i\left(\frac{1}{2}m\omega^2_i (q'_i)^2 +\frac{(p'_i)^2}{2m}\right).
1667: \end{equation}
1668: We can add in the additional potential energy due to small amplitude
1669: deflections in the electric field by expanding the displacement $z$ in
1670: terms of the normal modes,
1671: \begin{equation}
1672: V(z)=-\sum_i\frac{ e cV d_0}{4\pi\epsilon_0 I_i^{1/2}}
1673: \left(\int_0^L  \frac{\Gamma_i(x){d}x}{[(x-L)^2+d_0^2]^{3/2}}
1674: \right)q'_i, 
1675: \end{equation}
1676: where the origin of $x$ has been shifted.
1677: 
1678: We can  quantise the full Hamiltonian for the cantilever in the
1679: usual way, and so we are able to associate a position operator
1680: of the form  
1681: \begin{equation}
1682: \hat{q}'_i=\left(\frac{\hbar}{2m\omega_i}\right)^{1/2}(\hat{a}^{\dagger}_i
1683: +\hat{a}_i)
1684: \end{equation}
1685: with each mode.
1686: Thus, it follows that the interaction between the electron on the dot and
1687: the cantilever can be written in the form proposed [Eq.\ (\ref{coup})]
1688: with a  coupling constant between the cantilever and the
1689: dot which depends on the mode we are considering:
1690: \begin{equation}
1691: \lambda_i= e 
1692: E\xi_i\sqrt{\frac{\hbar}{2m\omega_i}},\label{coupling}
1693: \end{equation}
1694: where $\xi_i$ is a dimensionless constant of order unity defined
1695: by the relation
1696: \begin{equation}
1697: \xi_i=\frac{d_0^2}{(2I_i)^{1/2}}
1698: \int_0^L  \frac{\Gamma_i(x){d}x}{[(x-L)^2+d_0^2]^{3/2}}.
1699: \end{equation}
1700: 
1701: We can see from Eqs. (\ref{tqd}) and (\ref{ex2}) that the
1702:  effect  of a particular cantilever mode on the electron
1703: interference in the which--path device depends on
1704: $(\lambda_i/\hbar\omega_i)^2$, rather than on $\lambda_i$ alone. Thus, the
1705: ratio of the coupling between the fundamental and the $i$th excited mode goes
1706: as $(\omega_0/\omega_i)^{3}=(\beta_0/\beta_i)^{6}$. Because the electric field
1707: is strongly limited, we will always be working in the regime where the
1708: coupling constant is just sufficient to cause detectable effects. 
1709: Therefore, since $(\beta_0/\beta_1)^6\sim0.005$, our assumption that only
1710: the fundamental mode is relevant is  justified.   
1711: 
1712: As a concrete example, we consider a nanotube cantilever of length 
1713: $1.4$~$\mu$m and 
1714: outer radius $3.3$~nm.\cite{nano}
1715: In this case, the fundamental frequency is $140$~MHz and the mass is of order 
1716: $8\times 10^{-20}$~kg. If the tip--dot distance is set at $\sim 
1717: 0.1$~$\mu$m, then 
1718: $\xi_0\sim 1.3$ and for a maximum electric field of $10^5$~V/m, the corresponding  
1719: maximum coupling constant $\kappa=\lambda/\hbar\omega_0\sim 3$. 
1720: 
1721: 
1722: 
1723: \section{Non--Perturbative Calculation of Environmental Coupling}
1724: 
1725: In Sec. V, we calculated the transmission amplitude through the dot, 
1726: including the effect of the cantilever's environment, by expanding to 
1727: second order in the interaction between the cantilever and 
1728: the electron on the dot, $\lambda$, and then re-exponentiating.
1729: In  this appendix, we justify this result with a full calculation.
1730: The fact that a second order expansion in $\lambda$ leads to the exact 
1731: result is at first somewhat surprising. 
1732: However, the important step is the re-exponentiation of the truncated series 
1733: expansion (\ref{if}): we implicitly equate the 
1734: influence functional with not just the first few terms in an expansion, but 
1735: with an infinite series of terms which are themselves 
1736: composed of products of the second order terms we evaluated. This procedure is 
1737: generally acceptable as an approximation in the limit of small $\lambda$, but
1738:  in this particular  case the series generated in fact coincides with the exact form 
1739:  obtained from a linked cluster expansion.\cite{Mahan}
1740:  
1741:  
1742: We begin by adopting the notation
1743: \begin{eqnarray*}
1744: \hat{O}(t)&=&-{\mathrm{e}}^{(-i\omega_0- \gamma)t}\int_0^t dt' 
1745: \hat{F}(t'){\mathrm{e}}^{(i\omega_0+\gamma)t'} \\
1746: \hat{O}^{\dagger}(t)&=&-
1747: {\mathrm{e}}^{(i 
1748: \omega_0-\gamma)t}\int_0^t dt' 
1749: \hat{F}^{\dagger}(t'){\mathrm{e}}^{(-i\omega_0+\gamma)t'},
1750: \end{eqnarray*}
1751: so that 
1752: \begin{eqnarray*}
1753: \hat{a}(t)&=&\hat{a}(0){\mathrm{e}}^{(-i\omega_0-\gamma)t}+\hat{O}(t) \\
1754: \hat{a}^{\dagger}(t)&=&\hat{a}^{\dagger}(0){\mathrm{e}}^{(+i\omega_0-\gamma)t}+
1755: \hat{O}^{\dagger}(t).
1756: \end{eqnarray*}
1757: The operators $\hat{O}(t)$ and $\hat{O}^{\dagger}(t)$ operate only on the 
1758: variables of the oscillator bath and they both commute with $\hat{a}(0)$ and 
1759: $\hat{a}^{\dagger}(0)$ which operate on the states of the cantilever alone. 
1760: 
1761: The object which we wish to evaluate is the influence functional
1762: \begin{equation}
1763: \left\langle{\mathrm{T}}_t {\mathrm{e}}^{-i \int_0^{t}W_I(t')dt'/\hbar} 
1764: \right\rangle
1765: =\left\langle {\mathrm{T}}_t 
1766: {\mathrm{e}}^{i\lambda 
1767: \int_0^{t}\left[\hat{a}(0){\mathrm{e}}^{(-i\omega_0-\gamma)t'}+\hat{a}
1768: ^{\dagger}(0){\mathrm{e}}^{(+i\omega_0-\gamma)t'}\right]dt'/\hbar}\right\rangle
1769: \left\langle {\mathrm{T}}_t 
1770: {\mathrm{e}}^{i\lambda
1771:  \int_0^{t}\left[\hat{O}(t')+\hat{O}^{\dagger}(t')\right]dt'/
1772:  \hbar}\right\rangle, \label{c2}
1773: \end{equation}
1774: where we have exploited the commutation properties of the operators to 
1775: factor the expression into two terms, which we can now write as $C_1\times C_2$.
1776: 
1777: 
1778: The method we use to evaluate $C_1$ and $C_2$ is based on 
1779: the application of Wick's theorem, as described in the books by
1780:  Mahan\cite{Mahan}  
1781: and Louisell.\cite{Lou} We can write the first term as
1782: \begin{equation}
1783: C_1=\left\langle {\mathrm{T}}_t 
1784: {\mathrm{e}}^{i\lambda 
1785: \left[\int_0^{t}\hat{a}(0){\mathrm{e}}^{(-i\omega_0-\gamma)t'}+\hat{a}
1786: ^{\dagger}(0){\mathrm{e}}^{(+i\omega_0-\gamma)t'}\right]dt'/\hbar}\right\rangle=
1787: \sum_{n=0}^{\infty}i^n U_n(t),
1788: \end{equation}
1789: where 
1790: \begin{eqnarray*}
1791: U_n(t)&=&\frac{(\lambda/\hbar)^n}{n!}\int_0^t dt_1 \ldots\int_0^t
1792:  dt_n 
1793: \left\langle {\mathrm{T}}_t \left[
1794: \hat{a}(0){\mathrm{e}}^{(-i\omega_0-\gamma)t_1}+\hat{a}
1795: ^{\dagger}(0){\mathrm{e}}^{(+i\omega_0-\gamma)t_1}\right]\ldots\right. \\
1796: &&\times\left.\left[\hat{a}(0){\mathrm{e}}^{(-i\omega_0-\gamma)t_n}+\hat{a}
1797: ^{\dagger}(0){\mathrm{e}}^{(+i\omega_0-\gamma)t_n}\right]\right\rangle. 
1798: \end{eqnarray*}
1799: An obvious simplification arises from the fact that the averages over odd 
1800: numbers of operators will always vanish and so we can replace $n$ by the even 
1801: index $2m$. 
1802: 
1803: We now apply Wick's theorem which allows us to write the average of products 
1804: of pairs of operators as products of the averages of pairs of operators. Thus 
1805: for $U_{2m}$, we have
1806: \begin{equation}
1807: U_{2m}=\frac{(\lambda/\hbar)^{2m}i^{2m}}{(2m)!}\int_0^t dt_1 \ldots
1808: \int_0^tdt_{2m} \sum_C\left\{ D_0(t_1-t_i)...D_0(t_j-t_{2m}) \right\}, 
1809: \end{equation}
1810: where the summation is over all possible pairing combinations of the $2m$ time 
1811: labels and 
1812: \begin{equation}
1813: iD_0(t_1-t_2)=\left\langle {\mathrm{T}}_t 
1814: \left[\hat{a}(0){\mathrm{e}}^{(-i\omega_0-\gamma)t_1}+\hat{a}
1815: ^{\dagger}(0){\mathrm{e}}^{(+i\omega_0-\gamma)t_1}\right] 
1816: \left[\hat{a}(0){\mathrm{e}}^{(-i\omega_0-\gamma)t_2}+\hat{a}
1817: ^{\dagger}(0){\mathrm{e}}^{(+i\omega_0-\gamma)t_2}\right]\right\rangle.
1818: \end{equation}
1819: The summation over all possible combinations for each term allows us to 
1820: re-exponentiate so that $
1821: C_1={\mathrm{e}}^{-\phi_0(t)}$, 
1822: where
1823: \begin{equation}
1824: \phi_0(t)=\frac{i}{2}\left(\frac{\lambda}{\hbar}\right)^2\int_0^tdt_1
1825: \int_0^tdt_2 D_0(t_1-t_2).
1826: \end{equation}
1827: It is important to notice that the exact form of the function of time and 
1828: frequency multiplying the operators $\hat{a}(0)$ and $\hat{a}^{\dagger}(0)$ is 
1829: unimportant. These functions give the operators their individual time labels, 
1830: but because they are just algebraic functions they do not affect the validity of 
1831: Wick's theorem. 
1832: 
1833: 
1834: Now we must consider the second term,
1835: \begin{equation}
1836: C_2=\left\langle{\mathrm{e}}^{i\lambda\int_0^t\left[\hat{O}(t)+\hat{O}^{\dagger}
1837: (t)\right]{d}t'/\hbar}\right\rangle. 
1838: \end{equation}
1839: We can make progress by separating out the underlying operators of the 
1840: oscillators in the bath:
1841: \begin{eqnarray}
1842: \hat{O}(t)&=&-{\mathrm{e}}^{(-i\omega_0-\gamma)t}\int_0^tdt'\sum_{\omega}
1843: \hat{A}(\omega,0)g(\omega){\mathrm{e}}^{-i\omega t'}
1844: {\mathrm{e}}^{(i\omega_0 +\gamma)t} \\
1845: &=&\sum_{\omega}\hat{A}(\omega,0)g(\omega)\left[-{\mathrm{e}}^{(-i\omega_0
1846: -\gamma)t} \int_0^tdt'{\mathrm{e}}^{-i\omega t'}
1847: {\mathrm{e}}^{(i\omega_0 +\gamma)t'}\right] \\
1848: &=&\sum_{\omega}\hat{A}(\omega,0)g(\omega)f(\omega,\omega_0,\gamma,t)
1849: \end{eqnarray} 
1850: and similarly,
1851: \begin{equation}
1852: \hat{O}^{\dagger}(t)=\sum_{\omega}\hat{A}^{\dagger}
1853: (\omega,0)g^*(\omega)f^*(\omega,\omega_0,\gamma,t).
1854: \end{equation}
1855: We do not need to calculate $f$ explicitly since it is just an algebraic
1856: function and so always commutes. Because the oscillators in the heat bath 
1857: are all non--interacting, almost all the $\hat{A}(\omega,0)$ and 
1858: $\hat{A}^{\dagger}(\omega,0)$ operators commute. The only exceptions are 
1859: annihilation and creation operators of the same frequency. This means that we 
1860: can write the expectation value as a product over all the frequencies,
1861: \begin{equation}
1862: C_2=\prod_{\omega}\left\langle 
1863: {\mathrm{e}}^{i\lambda\int_0^t\left[\hat{A}(\omega,0)g(\omega)f+\hat{A}^{\dagger}
1864: (\omega,0)g^*(\omega)f^*\right]dt'/\hbar}\right\rangle. 
1865: \end{equation}
1866: 
1867: The advantage of decoupling $C_2$ into a product of terms is that each of these 
1868: terms can be handled in the same way as $C_1$. Because $\hat{A}(\omega,0)$ and 
1869: $\hat{A}^{\dagger}(\omega,0)$ are boson operators, Wick's theorem can again be 
1870: applied so that eventually we obtain
1871: \begin{equation}
1872: C_2=\prod_{\omega}{\mathrm{e}}^{-\phi_{\omega}(t)},
1873: \end{equation}
1874: where
1875: \begin{equation}
1876: \phi_{\omega}(t)=\frac{i\lambda^2}{2\hbar^2}\int_0^tdt_1
1877: \int_0^tdt_2 D_{\omega}(t_1-t_2),
1878: \end{equation}
1879: with 
1880: \begin{eqnarray}
1881: iD_{\omega}(t_1-t_2)&=&\left\langle {\mathrm{T}}_t 
1882: \left[\hat{A}(\omega,0)g(\omega)f(\omega,\omega_0,\gamma,t_1)+\hat{A}
1883: ^{\dagger}(\omega,0)g^*(\omega)f^*(\omega,\omega_0,\gamma,t_1)\right]\right. \\ \nonumber
1884: &&\times\left. 
1885: \left[\hat{A}(\omega,0)g(\omega)f(\omega,\omega_0,\gamma,t_2)+\hat{A}
1886: ^{\dagger}(\omega,0)g^*(\omega)f^*(\omega,\omega_0,\gamma,t_2)\right]\right\rangle.
1887: \end{eqnarray}
1888: 
1889: The overall expression for the influence functional can now be written in a 
1890: simplified form
1891: \begin{equation}
1892: \left\langle{\mathrm{T}}_t {\mathrm{e}}^{-i 
1893: \int_0^{t}W_I(t')dt'/\hbar}\right\rangle=
1894: {\mathrm{e}}^{-\left[\phi_0(t)+\sum_{\omega}\phi_{\omega}(t)\right]}
1895: ={\mathrm{e}}^{-\phi(t)},
1896: \end{equation}
1897: where
1898: \begin{equation}
1899: \phi(t)=\frac{i\lambda^2}{\hbar^2}\int_0^tdt_1
1900: \int_0^{t_1}dt_2 \left\{D_0(t_1-t_2)+\sum_{\omega}D_{\omega}
1901: (t_1-t_2)\right\}.
1902: \end{equation}
1903: However, because the operators in the bath are independent, only the averages 
1904: including  
1905: pairs of operators from the same oscillator are non--zero. Thus, we can simplify 
1906: the sum over $D_{\omega}(t)$ functions: 
1907: \begin{eqnarray}
1908: \sum_{\omega}D_{\omega}(t_1-t_2)&=&\sum_{\omega}\left\langle 
1909: \left[\hat{A}(\omega,0)g(\omega)f(\omega,\omega_0,\gamma,t_1)+\hat{A}^{\dagger}
1910: (\omega,0)g^*(\omega)f^*(\omega,\omega_0,\gamma,t_1)\right]\right.\\ \nonumber
1911: &&\left.\times\left[\hat{A}(\omega,0)g(\omega)
1912: f(\omega,\omega_0,\gamma,t_2)+\hat{A}^{\dagger}
1913: (\omega,0)g^*(\omega)f^*(\omega,\omega_0,\gamma,t_2)\right]
1914: \right\rangle\\
1915: &=&\left\langle  
1916: \left[\hat{O}(t_1)+\hat{O}^{\dagger}(t_1)\right]
1917: \left[\hat{O}(t_2)+\hat{O}^{\dagger}(t_2)\right] 
1918: \right\rangle.
1919: \end{eqnarray}
1920: 
1921: 
1922: Since the $\hat{O}(t)$ and $\hat{O}^{\dagger}(t)$ operators commute with 
1923: $\hat{a}(0)$ and $\hat{a}^{\dagger}(0)$ we can complete the process of 
1924: recombination to obtain
1925: \begin{equation}
1926: i\left\{D_0(t_1-t_2)+\sum_{\omega}D_{\omega}(t_1-t_2)\right\}
1927: =\left\langle\left[\hat{a}(t_1)+\hat{a}^{\dagger}(t_1)\right]
1928: \left[\hat{a}(t_2)+\hat{a}^{\dagger}(t_2)\right]\right\rangle. \label{dv}
1929: \end{equation}
1930: 
1931: We can now use our previous results in Eqs. (\ref{a}), (\ref{b}), and (\ref{c}) 
1932: to evaluate the averages in Eq. (\ref{dv}):
1933: \begin{equation}
1934: \left\langle\left[\hat{a}(t_1)+\hat{a}^{\dagger}(t_1)\right]
1935: \left[\hat{a}(t_2)+\hat{a}^{\dagger}(t_2)\right]\right\rangle
1936: ={\mathrm{e}}^{-\gamma (t_1-t_2)}\left[ (\overline{n}+1)
1937: {\mathrm{e}}^{-i\omega_0 (t_1-t_2)}+\overline{n}
1938: {\mathrm{e}}^{i\omega_0 (t_1-t_2)}\right].
1939: \end{equation}
1940: Thus, our final expression for the influence functional  is
1941: \begin{equation}
1942: \left\langle{\mathrm{T}}_t {\mathrm{e}}^{-i \int_0^{t}W_I(t')dt'/\hbar} 
1943: \right\rangle
1944: ={\mathrm{e}}^{-\phi(t)},
1945: \end{equation}
1946: with
1947: \begin{equation}
1948: \phi(t)=\frac{\lambda^2}{\hbar^2}\int_0^tdt_1
1949: \int_0^{t_1}dt_2 {{\mathrm{e}}}^{-\gamma (t_1-t_2)}\left[ 
1950: (\overline{n}+1)
1951: {\mathrm{e}}^{-i\omega_0 (t_1-t_2)}+\overline{n}
1952: {\mathrm{e}}^{i\omega_0 (t_1-t_2)}\right]. \label{phic}
1953: \end{equation}
1954: Comparing Eq. (\ref{phic}) with Eq. (\ref{phi}), it is clear that the expression
1955:  we obtained for the influence functional in Sec. V is exact.
1956: 
1957: 
1958: 
1959: 
1960: \begin{references}
1961: \bibitem[*]{auth}Electronic address: a.armour@ic.ac.uk
1962: \bibitem[\dag]{auth2}Electronic address: miles.p.blencowe@dartmouth.edu
1963: 
1964: \bibitem{hist}N. Bohr, in {\emph{Albert Einstein: Philosopher
1965:       Scientist}}, edited by P.A. Schilpp, (Library of Living
1966:   Philosophers, Evanston, Ill., 1949), vol. 7, p. 200.
1967: \bibitem{Fey} R. Feynman, R. Leighton and M. Sands, {\emph{The Feynman
1968:       Lectures on Physics}} (Addison Wesley, Mass.,  1965), vol. III.
1969: \bibitem{Buks} E. Buks, R.  Schuster, M. Heiblum, D. Mahalu, and
1970:       V. Umansky, {{Nature}} {\bf{391}}, 871 (1998). 
1971: \bibitem{op} S. D\"{u}rr, T. Nonn, and G. Rempe,   {{Nature}} 
1972: {\bf{395}}, 33 (1998). 
1973: \bibitem{SS}W. K. Wooters and W. H. Zurek, {{Phys. Rev. D}} {\bf{19}}, 473 
1974: (1979).
1975: \bibitem{Scully} M. O. Scully, B.--G. Englert, and H. Walther,  {{Nature}} 
1976: {\bf{351}}, 111 (1991). 
1977: \bibitem{Walls} P. Storey, S. Tan, M. Collett, and D. Walls,  {{Nature}} 
1978: {\bf{367}}, 626 (1994). 
1979: \bibitem{Stern} A. Stern, Y. Aharanov, and Y. Imry, {{Phys. Rev. A}} 
1980: {\bf{41}}, 3436 (1990).
1981: \bibitem{hacr} For a review of previous work on phase coherent transport
1982: and controlled dephasing see G. Hackenbroich, /cond-mat/0006361
1983: \bibitem{Yac}A. Yacoby, M. Heiblum, D. Mahalu, and H. Shtrikman 
1984: {{Phys. Rev. Lett.}} {\bf{74}}, 4047 (1995). 
1985: \bibitem{Hack} G. Hackenbroich, B. Rosenow, and H. A. Weidenm\"{u}ller,
1986:   {{Europhys. Lett.}} {\bf{44}}, 693 (1998).
1987: \bibitem{Le} Y. Levinson, {{Europhys. Lett.}} {\bf{39}}, 299 (1997).
1988: \bibitem{Al} I. L. Aleiner, N. S. Wingreen, and Y. Meir,
1989:  {{Phys. Rev. Lett.}} {\bf{79}}, 3740 (1997).
1990: \bibitem{Gurvitz} S. A. Gurvitz, {{Phys. Rev. B}} {\bf{56}}, 15215 (1997).
1991: \bibitem{BJK} S. Mancini, V. I. Man'ko, and P. Tombesi,  {{Phys. Rev. A}} 
1992: {\bf{55}}, 3042 (1997);
1993: S. Bose, K. Jacobs, and P. L. Knight , {{Phys. Rev. A}} 
1994: {\bf{56}}, 4175 (1997).
1995: \bibitem{Bose}S. Bose, K. Jacobs, and P. L. Knight,
1996: {{Phys. Rev. A}}  {\bf{59}}, 3204 (1999).
1997: \bibitem{Datta} S. Datta, {\emph{Electronic Transport in Mesoscopic
1998:       Structures}}, (Cambridge University Press, Cambridge, U.K., 1995).
1999: \bibitem{But1} M. B\"{u}ttiker, IBM J. Res. Develop. {\bf{32}}, 317 (1988).
2000: \bibitem{But2} A. Levy Yeyati and   M. B\"{u}ttiker,
2001:     {{Phys. Rev. B}} {\bf{52}}, R14360 (1995).
2002: \bibitem{Imry} Y. Imry, {\emph{Introduction to Mesoscopic Physics}}, 
2003:   (Oxford University Press, Oxford, U.K., 1997). 
2004: \bibitem{Yac2}A. Yacoby, R. Schuster, and M. Heiblum,  
2005: {{Phys. Rev. B}} {\bf{53}}, 9583 (1996).
2006: \bibitem{steady}Our assumption is that the which--path device is 
2007: operated in a steady--state mode, whereby measurements are only made 
2008: after any transitory behavior has decayed so that the probability 
2009: distribution of (initial) cantilever states is stationary. 
2010: This steady--state distribution 
2011: is further assumed to be the usual thermal distribution. These assumptions will 
2012: only
2013: be strictly correct in the limit as the current goes to zero, but can be expected 
2014: to provide  predictions which are at least qualitatively correct for the average
2015:  behavior over long times  even with 
2016: finite currents.
2017: \bibitem{vis}P. D. D. Schwindt, P. G. Kwiat, and B.--G. Englert, 
2018:  {{Phys. Rev. A}} {\bf{60}}, 4285 (1999).
2019: \bibitem{justify} We assume that all positions are measured from the
2020:   equilibrium height of the cantilever above the dot, including the
2021:   effect of the residual electrons on the dot which do not participate 
2022:  in the transport. 
2023: \bibitem{W89}N. S. Wingreen, K. W. Jacobsen, and J. W. Wilkins, 
2024: {{Phys. Rev. B}} {\bf{40}}, 11834 (1989).
2025: \bibitem{GS}L. I. Glazman and R. I. Shekhter, Zh. Eksp. Teor. Fiz. 
2026: {\bf 94}, 292 (1987) 
2027: [{{Sov. Phys.---JETP}} {\bf{67}}, 163 (1988)].
2028: \bibitem{JW}A.--P. Jauho and N. S. Wingreen, 
2029:   {{Phys. Rev. B}} {\bf{58}}, 9619 (1998).
2030: \bibitem{UZ}W. G. Unruh and W. H. Zurek, 
2031: {{Phys. Rev. D}} {\bf{40}}, 1071 (1989).
2032: \bibitem{vib}P. Poncharal, Z. L. Wang, D. Ugarte, and W. A. de Heer, 
2033:  {{Science}} {\bf{283}}, 1513 (1999).
2034: \bibitem{CL} A. O. Caldeira and A. J. Leggett, 
2035: {{Ann. Phys. (N.Y.)}} {\bf{149}}, 374 (1983).
2036: \bibitem{O}R. Omn\`{e}s, {\emph{The Interpretation of Quantum Mechanics}},
2037: (Princeton University Press, Princeton, NJ, 1994).  
2038: \bibitem{JZ}E. Joos and H. D. Zeh, {{Z. Phys. B}} {\bf{59}} 223 (1985).
2039: \bibitem{MW}L. Mandel and E. Wolf, {\emph{Optical Coherence and Quantum Optics}},
2040: (Cambridge University Press, Cambridge, U.K., 1995). 
2041: \bibitem{Lou}W. Louisell, {\emph{Quantum Statistical Properties of Radiation}},
2042: (J. Wiley, New York, 1973).
2043: \bibitem{Buksp}E. Buks (private communication).
2044: \bibitem{Buks2}D. Sprinzak, E. Buks, M. Heiblum, and H. Shtrikman, 
2045:  {{Phys. Rev. Lett.}} {\bf{84}}, 5820 (2000). 
2046: \bibitem{AFM} R. M. D. Stevens, N. A. Frederick, B. L. Smith, D. E. Morse,
2047:  G. D. Stucky, and P. K. Hansma,  {{Nanotechnology}}
2048:   {\bf{11}}, 1  (2000).
2049: \bibitem{jack}J. D. Jackson, {{Am. J.  Phys.}}
2050: {\bf{68}}, 789 (2000). 
2051: \bibitem{nano} M.M.J. Treacy, T. W. Ebbesen, and J. M. Gibson,
2052:   {{Nature}} {\bf{381}}, 678 (1996).
2053: \bibitem{modes2}B. Reulet, A. Y. Kasumov, M. Kociak, R. Deblock,
2054: I. I. Khodos, Y. B. Gorbatov, V. T. Volkov, C. Journet, and H. Bouchiat,
2055:  {{Phys. Rev. Lett.}} {\bf{85}}, 2829 (2000). 
2056: \bibitem{modes}H.--J. Butt and M. Jaschke,  {{Nanotechnology}}
2057:   {\bf{6}}, 1  (1995).
2058: \bibitem{Mahan}G. Mahan, {\emph{Many--Particle Physics}}, 
2059: (2nd edition, Plenum Press, New York, 1990).
2060: \end{references}
2061: 
2062: 
2063: 
2064: \begin{figure}
2065: \caption{Magnitude of the transmission amplitude, $|\langle t_{\rm QD} \rangle|$
2066: at resonance, for a
2067:   cantilever with $\Gamma/\hbar\omega_0=10$ as a function of the dimensionless
2068:   coupling constant
2069: $ \kappa=\lambda/\hbar\omega_0$. The dashed curve is obtained without
2070: an average over incident electron energies, while the full curve includes
2071: the averaging. The amplitudes are normalized to one at $\kappa=0$.  }
2072: \label{fig:one}
2073: \end{figure}
2074: 
2075: 
2076: \begin{figure}
2077: \caption{Magnitude of the resonant transmission amplitude, $|\langle 
2078: t_{\mathrm QD}\rangle|$,
2079:  as a function of coupling constant $\kappa$, for  
2080: $\Gamma/\hbar\omega_0$=2 (dotted curve), 1 (dashed curve) and 0.5 (full curve).}
2081: \label{fig:two}
2082: \end{figure}
2083: 
2084: 
2085: \begin{figure}
2086: \caption{(a) Magnitude of the transmission amplitude, $|\langle 
2087: t_{\mathrm QD}\rangle |$, and
2088:   (b) the phase, for a
2089:   cantilever with $\Gamma/\hbar\omega_0=0.5$  as a function of coupling constant
2090: $ \kappa$ and the energy detuning 
2091: $\epsilon-\epsilon_{0}$. The average over the electron energy
2092:   distribution has not been performed.}
2093: \label{fig:three}
2094: \end{figure}
2095: 
2096: 
2097: 
2098: \begin{figure}
2099: \caption{(a) Magnitude of the transmission amplitude, $|\langle t_{\mathrm QD}\rangle|$, and
2100:   (b) the phase, for a
2101:   cantilever with $\Gamma/\hbar\omega_0=0.5$  as a function of coupling 
2102:   constant, including
2103:   the effect of the cantilever's environment with 
2104:   $Q=50$, as a function of $\kappa$ and  
2105:   $\epsilon-\epsilon_{0}$.}
2106: \label{fig:four}
2107: \end{figure}
2108: 
2109: \begin{figure}
2110: \caption{Magnitude of the resonant transmission amplitude, $|\langle 
2111: t_{\mathrm QD}\rangle|$, 
2112:  including the effect of the cantilever's environment as a function of 
2113:  $\kappa$, for  
2114: $\Gamma/\hbar\omega_0=0.5$ and $Q=50$ (dotted curve)
2115:  and  500
2116: (dashed curve). The latter curve has been shifted upwards by 0.5 and 
2117: the former shifted upwards by 1 for clarity.
2118: The case without environmental coupling where  
2119: $\gamma_{c}/\omega_0$=0 (full curve) is included for comparison.}
2120: \label{fig:five}
2121: \end{figure}
2122: 
2123: \vfill
2124: \eject
2125: \mbox{\epsfig{file=which-pathplots1.nb.ps, width=6in}}
2126: \mbox{\epsfig{file=which-pathplots2.nb.ps, width=6in}}
2127: \mbox{\epsfig{file=wp3a.ps, width=6in}}
2128: \mbox{\epsfig{file=wp3b.ps, width=6in}}
2129: \mbox{\epsfig{file=wp4a.ps, width=6in}}
2130: \mbox{\epsfig{file=wp4b.ps, width=6in}}
2131: \mbox{\epsfig{file=which-pathplots5.nb.ps, width=6in}}
2132: 
2133: 
2134: 
2135: 
2136: \end{document}
2137: