1: \documentclass[aps,twocolumn,showpacs]{revtex4}% Physical Review B
2: %\documentstyle[prb,aps,twocolumn,eqsecnum,epsfig]{revtex}
3: %\documentstyle[prb,aps,eqsecnum,preprint,epsfig]{revtex}
4: \usepackage{graphicx}% Include figure files
5: \begin{document}
6: \title{Multi-particle effects in non-equilibrium electron
7: tunnelling and field emission}
8: \author{Andrei Komnik and Alexander O. Gogolin}
9: \affiliation{Department of Mathematics, Imperial College, 180 Queen's Gate,
10: London SW7 2BZ, United Kingdom}
11: \date{\today}
12: %\maketitle
13: \begin{abstract}
14: We investigate energy resolved electric current from various
15: correlated host materials under out-of-equilibrium conditions. We
16: find that, due to a combined effect of electron-electron
17: interactions, non-equilibrium and multi-particle tunnelling, the
18: energy resolved current is finite even above the Fermi edge of
19: the host material. In most cases, the current density possesses a
20: singularity at the Fermi level revealing novel manifestations of
21: correlation effects in electron tunnelling. By means of the
22: Keldysh non-equilibrium technique, the current density is
23: calculated for one-dimensional interacting electron systems and
24: for two-dimensional systems, both in the pure limit and in the
25: presence of disorder. We then specialise to the field emission
26: and provide a comprehensive theoretical study of this effect in
27: carbon nanotubes.
28: \end{abstract}
29: \pacs{03.65.X, 71.10.P, 73.63.F}
30:
31: \maketitle
32: \section{Introduction}
33: At the early stages of the development of the quantum theory it
34: became clear that electron tunnelling processes are of
35: fundamental importance for condensed matter physics \cite{mahan}.
36: Electron tunnelling became an essential concept in such fields as
37: semiconductor physics, particle transport in mesoscopic physics,
38: field emission, and many others. In a multitude of setups,
39: tunnelling processes have attracted attention over decades
40: and continue to do so.
41:
42: For the purposes of this work, we shall visualise tunnelling
43: events as taking place between a metallic host material and a
44: lead material (another metal, a semiconductor, or indeed vacuum
45: as in the field emission effect). The host and the lead are
46: separated from each other by a substantial potential barrier. The
47: lead is supposed to contain a detector measuring the current
48: density. The presence of a finite current immediately implies that
49: one is dealing with a non-equilibrium (steady-state) system.
50: Apart from the total current, the energy resolved current
51: $j(\omega)$, i.~e. the amount of current in the energy window
52: $\omega$ to $\omega+d\omega$, is an important characteristics of
53: the tunnelling process. The total current $J$ is then the
54: integral of $j(\omega)$ over all energies. In the field emission
55: (FE) setup, the energy resolved current have been measured for
56: different emitters over some 30 years (see \cite{plummer}, and
57: also references below). We shall therefore adopt the terminology
58: of the Gadzuk and Plummer's review article, \cite{plummer}, and
59: refer to $j(\omega)$ as the total energy distribution or TED. We
60: are not aware of any direct measurements of TEDs in tunnelling
61: contacts.
62:
63: From the physical point of view, the most interesting aspect of
64: the TED is the part of the spectrum above the Fermi edge,
65: $\omega>E_F$. At the leading order in the tunnelling amplitude,
66: $j(\omega)$ is simply proportional to tunnelling probability
67: times the electron energy distribution function in the host,
68: $n(\omega)$. The latter is identically zero (at zero temperature)
69: above the Fermi edge due to the Luttinger theorem
70: \cite{luttingertheorem}. The high-energy tail is due to a
71: combined effect of non-equilibrium, multi-particle tunnelling
72: processes, and mutual electron-electron interactions in the host
73: material (for non-equilibrium multi-particle effects {\it per se}
74: are not sufficient to smear the Fermi surface \cite{caroli}). So,
75: we shall also refer to this high-energy tail of the energy
76: resolved current distribution as `secondary' current, `primary'
77: current being the one below the Fermi edge (i.~e. the one which is
78: proportional to the tunnelling probability). In the grand
79: picture, the secondary current is akin to the Auger effect though
80: it is, of course, a more complex phenomenon taking place not in
81: an atom but in a fully interacting metallic host. Measurements of
82: the secondary current (the trivial thermal broadening of the
83: primary current being subtracted off) can thus provide a valuable
84: source of information about the electronic correlations in the
85: host \cite{plummer,my}. The effect was indeed first discovered
86: experimentally -- in FE measurements by Lea and Gomer
87: \cite{1stexp}. Theoretical analysis of this phenomenon was done
88: by Gadzuk and Plummer soon afterwards \cite{gadzuk}. Following
89: the pioneering paper by Fowler and Nordheim \cite{fowler} these
90: authors used the connection between the FE problem and the
91: tunnelling problem and studied Boltzmann-like equations for the
92: particle-hole balance in the low density approximation. To remove
93: such restrictions and to put the theory on the modern footing, so
94: that it becomes applicable to strongly-correlated emitters, we
95: embark in this paper on investigating the issue by employing
96: Keldysh diagram technique, appropriate for this non-equilibrium
97: situation \cite{keldysh,LLX}. This method allows us to
98: consistently write down series in the tunnelling amplitude for
99: all quantities of interest and for the TED in particular.
100:
101: Our motivation is in fact two-fold. On one hand, we think that
102: the secondary current phenomenon has not received sufficient
103: attention of theorists. The physics of the interplay between
104: non-equilibrium multi-particle tunnelling and electron
105: correlations is worth a deeper study. In particular, we
106: investigate in this paper what aspects of the electron
107: correlations are responsible for the current above the Fermi edge
108: and in what setups, other than the FE, such current can occur. On
109: the other hand, since the original work \cite{1stexp,gadzuk}
110: there have been considerable advances in the emitter technology.
111: Perhaps the most important recent development is the usage of
112: carbon nanotubes as field emitters. Carbon nanotubes display a
113: remarkable array of electronic and mechanical properties and are
114: of potential technological importance \cite{dekker}. Electron
115: transport in these systems has been thoroughly investigated.
116: While single-wall nanotubes (SWNTs) exhibit one-dimensional (1D)
117: Luttinger liquid (LL) type transport properties and are pretty
118: well understood, the theory \cite{sammlung1,sammlung2,sammlung3}
119: and experiment being in agreement \cite{tans1,bockrath1},
120: multi-wall nanotubes (MWNTs) are more complex systems subject to
121: intensive current debate. MWNTs are composed from many (at least
122: ten) concentric graphite shells. Current experiments on them are
123: consistent with two-dimensional (2D) diffusion with
124: characteristic weak localisation \cite{bachtold1} features and
125: zero-bias anomalies
126: \cite{schoenenberger,bachtold2,eg2001,mishchenko}. What concerns
127: us here is that, apart from other uses, carbon nanotubes are
128: expected to act as field emitters in high-resolution displays and
129: cathode tubes \cite{wong1,saito1}. While there have been several
130: experimental investigations of the FE from carbon nanotubes (see
131: \cite{french1} and the main text for more references), both from
132: SWNTs and MWNTs, the relevant theory has been lacking. So, the
133: second leg of our motivation is to discuss existing FE
134: experiments on carbon nanotubes and make further theoretical
135: predictions for these systems.
136:
137: Having in mind applications to carbon nanotubes and taking into
138: account the fact that FE usually occurs from a tip of the tube
139: (both for SWNTs and MWNTs) we narrow the following considerations
140: from a number of imaginable setups to an appropriate tip geometry.
141: Apart from this restriction, we intent to advance in this paper a
142: general discussion of non-equilibrium multi-particle tunnelling
143: from strongly-correlated 1D and 2D hosts paying special attention
144: to carbon nanotubes. Some of the results on 1D emitters
145: (applicable to SWNTs) have been announced in our recent letter
146: \cite{my}.
147:
148: The paper is organised as follows. In the next Section we present
149: some qualitative considerations concerning the physical nature of
150: the TED. Section \ref{general} contains general
151: (model-independent) results. We identify the relevant Keldysh
152: diagrams contributing to the TED above the Fermi edge and then
153: perform a spectral analysis of the involved correlation functions.
154: A simple application of the developed theory is contained in
155: Section \ref{seclocalinteraction}, where we treat an electron
156: system with local interactions confined to the vicinity of the
157: tunnelling point. In Section \ref{luttinger} we analyse the TED of
158: particles tunnelling from LLs. In subsequent Sections \ref{pure}
159: and \ref{disorder}, the same problem is studied for correlated 2D
160: electron systems, respectively in the pure limit and in the
161: presence of a disorder potential. In Section \ref{FE} we
162: specialise to the case of field emission from carbon nanotubes.
163: Finally, in Section \ref{LS} we discuss a more sophisticated
164: two-stage tunnelling via a localised state. Summary and
165: conclusions Section completes the paper.
166:
167:
168: \section{Physical picture} \label{secII}
169: Before proceeding with calculations, let us elaborate on
170: qualitative origins of the TED high-energy tails.
171: The simplest setting to start with involves two noninteracting electrodes
172: with a tunnelling contact between them, see Fig.~\ref{fig1}.
173: %The transmission coefficient does not depend on the energy of the particles.
174: Applying a finite voltage leads to a nonzero current through the
175: junction. Neglecting the charging effects and (for now) under the
176: assumption of constant electron densities of states and
177: transmission coefficient, the current is proportional to the
178: applied voltage $V$. On the right side of the contact, where the
179: chemical potential is supposed to be lower than on the left one,
180: the current is carried by the tunnelled electrons. On the left
181: side of the contact the current is carried by the holes moving in
182: the opposite direction away from the contact. As long as the
183: system is noninteracting, the TED of the electrons that tunnelled
184: out is uniform in the window between $E_F$ and $E_F-V$ and is
185: zero outside (this is evident but can be confirmed by simple
186: calculation in the spirit of \cite{caroli}, which we omit). From
187: now on we set $e=1$. $E_F$ is the Fermi level of the left
188: electrode.
189:
190: This picture changes drastically if interactions between the
191: electrons are switched on. We restrict our considerations to the
192: case when only the left contact -- the host -- is correlated (see
193: also discussion in the next Section). Then the difference between
194: the actual energy distribution function in the lead and the
195: noninteracting distribution function gives the TED of particles
196: tunnelling out of the host. The holes left behind by electrons
197: tunnelling out below the Fermi energy (`primary' electrons)
198: experience scattering from the electron sea, thereby creating
199: electron-hole pairs. Contrary to the problem of hot electron
200: relaxation (see e.~g.~\cite{meden}, and references therein), we
201: are dealing here with a \emph{flow} of hot holes in a steady
202: state. Also, we are interested in a different object: the energy
203: distribution functions rather than momentum distribution
204: functions (the latter do, of course, have a non-zero tail above
205: the Fermi momentum solely due to correlations, without a need to
206: include tunnelling \cite{agd}). The emerging `secondary'
207: electrons can also be carried over to the lead by a successive
208: tunnelling process, as shown in Fig.~\ref{fig1}. This phenomenon
209: can be regarded as condensed matter analogy of the Auger process
210: known from the atomic physics. Since the energy of the holes,
211: measured from the Fermi edge $E_F$, can not exceed $V$, the upper
212: limit for the energy the secondary electrons can acquire in the
213: electron-hole pair creation process is given by $V$ as well.
214: \begin{figure}
215: %\begin{center}
216: %\epsfxsize=0.6\columnwidth
217: %\hfill
218: %\epsffile{diag3.eps}
219: %\hfill
220: %\vspace*{0.3cm}
221: \includegraphics[scale=0.3]{setup1}
222: \caption[]{\label{fig1}
223: Tunnelling junction biased by finite voltage $V$. The left electrode is the
224: interacting one while the right one is uncorrelated. Tunnelling electrons with
225: energies higher than $E_F$ are due to hot hole (denoted by
226: $p_i$) scattering with creation of electron-hole pairs $e-p_c$ (see
227: inset).}
228: \end{figure}
229: At higher orders in the tunnelling probability this threshold is increased
230: so that the secondary electrons can, in fact, become more energetic than $V$.
231: In this case they ought to be successively scattered either from
232: another hot hole or from another hot electron.
233: This is, however, a process of higher order not only in the tunnelling but
234: also in the interaction.
235: Therefore secondary electrons with energies between $(n-1) V$ and $n V$
236: emerge in the processes of $2n$th order with respect to the tunnelling amplitude
237: as well as the interaction constant.
238: Obviously, in equilibrium $V=0$ all the high energy tails vanish, thereby
239: restoring the Fermi edge in accordance with the Luttinger theorem.
240: So, non-trivial interaction effects are encoded in the high-energy behaviour
241: of the non-equilibrium TED.
242:
243: Theoretically, the most interesting limiting case is the
244: behaviour of the TED just above the Fermi edge, when certain
245: universality can be expected. Out of the considerations of
246: Ref.~\cite{gadzuk} a divergent TED emerged, with (simplifying
247: matters) the singularity approximately of the form $j(\omega)\sim
248: 1/(\omega-E_F)$. This turned out to be roughly consistent with
249: measurements at the time \cite{1stexp}. As detailed in the
250: following Sections, the present study supports the view that the
251: limiting form of the TED strongly depends on the nature of the
252: host material and the geometry of the setup (within a given
253: material type though, some universality does set in, so for
254: point-contacted LLs we find a power-law behaviour, etc.).
255:
256: \section{Statement of the problem and general results} \label{general}
257: We now formalise the problem by writing down
258: the relevant tunnelling Hamiltonian:
259: \begin{equation} \label{ham0}
260: H = H[c] + H[\psi] + \gamma \left[ \psi^\dag(0) c(0) + c^\dag(0) \psi(0)
261: \right] \; .
262: \end{equation}
263: Here $\gamma$ is the tunnelling amplitude and $c$ and $\psi$ are the
264: annihilation operators for the electrons
265: in the lead and in the host, respectively.
266: The unperturbed part of the Hamiltonian $H_0=H[c] + H[\psi]$ describes
267: two decoupled electron systems at different chemical potentials
268: $\mu_\psi-\mu_c=V>0$.
269:
270: A clarification on the following points is in order.
271:
272: {\bf (i)} As we only want to consider a tip geometry, we have
273: explicitly assumed that the tunnelling occurs only locally at the
274: location of the tip: $x=0$. In reality, of course, there is a
275: small area over which the tunnelling takes place (strictly speaking,
276: a real-space integral is required in the tunnelling term in
277: Eq.(\ref{ham0})). As this has no qualitative influence, we shall
278: keep writing simple as long as we can. The locality assumption is
279: natural for carbon nanotubes because of the very shape of these
280: objects. It is however also justified for most bulk interfaces
281: and field emitters where, because of the roughness of the surface
282: and the pronounced exponential dependence of the penetration
283: coefficient on the distance between the electrodes, the main
284: contribution to the current comes from only few points between
285: the electrodes.
286:
287: {\bf (ii)} Related to the above is the question of the energy
288: (momentum) dependence of the tunnelling amplitude. In real
289: systems it is energy dependent. So, for the FE setup the relevant
290: energy scale is determined by a combination of the work function
291: and the applied field. It is important to keep in mind, however,
292: that as the tunnelling amplitude is a single electron property,
293: regarding its energy dependence there is no special significance
294: to the Fermi edge. Therefore, when addressing observables
295: determined by scattering processes taking place close to the
296: Fermi surface, it is quite safe to neglect the energy dependence
297: of the tunnelling amplitude and replace it by a constant,
298: $\gamma$ ($\gamma^2$ being proportional to the transmission
299: coefficient of the potential barrier at the Fermi energy: ${\cal
300: D}(E_F)$.) In the literature on the subject the tunnelling term is
301: often quoted in the momentum representation in the form
302: $\gamma_{pk}\psi^\dagger_p c_k$ plus conjugate with some
303: unspecified matrix elements $\gamma_{pk}$. We take the view that,
304: for calculating quantities related to the Fermi surface, this
305: would only complicate formulae without conceptual gain. There
306: will be, however, instances in the following when the energy
307: dependence of the tunnelling amplitude is important, like when
308: there is a localised state or when we discuss generalisations of
309: the Fowler--Nordheim relations for nanotubes. In these cases
310: (Sections \ref{FE} and \ref{LS}) we shall take the relevant
311: energy dependence fully into account.
312:
313: {\bf (iii)} Throughout this paper we assume that the lead is
314: non-interacting. Indeed one could not otherwise separate the
315: effect of correlations occurring in the host from those occurring
316: in the lead, which would seriously hamper meaningful
317: interpretation of measurements on such systems. This assumption is
318: justified for the FE setup, save for the Boersh effect, which is
319: not believed to be important for carbon nanotubes (see
320: \cite{fransen} and \cite{my}). For a general tunnelling junction
321: setup it would not be correct to {\it a priori} neglect
322: correlations in the lead. However, with care it is possible to
323: realise a reasonable setup involving, for instance, a nanotube in
324: contact with a low carrier density semiconductor or a quantum wire
325: opening up into a higher dimensional lead preferably screened by
326: a nearby gate (in fact most conductance measurements on quantum
327: wires are nowadays interpreted in terms of a junction with a
328: non-interacting lead \cite{tarucha,glazman}).
329:
330: In the bulk of this paper we shall be investigating the secondary
331: current $j(\omega>0)$ (we set $E_F=0$ from now on). It is a
332: plausible statement that the secondary current should be
333: proportional to the high-energy tail $n(\omega>0)$ of the
334: electron energy distribution function {\sl inside} the emitter tip
335: (calculated to all orders in the tunnelling amplitude). Indeed
336: the latter quantity corresponds to the amount of electrons
337: available for tunnelling at a given energy $\omega$, while the
338: proportionality factor contains (unessential) information about
339: how these electrons are then carried over to the detector. In the
340: first part of this Section we establish this statement.
341:
342: Since we are dealing with non-equilibrium phenomena we resort to the
343: Keldysh formalism. We denote by $G(x,x';t-t')$ the
344: generalised Green's function for electrons in the lead,
345: \begin{equation} \label{first}
346: G(x,x';t-t') = -i \langle T_C [c(x,t)
347: c^\dag(x',t')\,S_C] \rangle_0 \, ,
348: \end{equation}
349: the Green's function in the host, $g$, being defined by an analogous
350: formula.
351: Here $T_C$ stands for the ordering operation along the Keldysh
352: contour, shown in Fig.~\ref{contour}, and $x$ corresponds to a
353: set of coordinates specifying the electron states in the lead.
354: (It should be understood as a distance from the emitter tip plus
355: possibly a transverse channel index which we suppress as it plays
356: no important part in the following). The average in
357: Eq.~(\ref{first}) is taken over the ground-states of the
358: unperturbed Hamiltonian $H_0$ and the contour $S_C$-matrix is
359: responsible for tunnelling events:
360: \begin{eqnarray} \label{Smatrix} \nonumber
361: S_C = T_C \exp\left(-i \gamma \int\limits_C \, dt \, [ \psi^\dag(t) c(t) +
362: c^\dag(t) \psi(t) ]\right) \, .
363: \end{eqnarray}
364: \begin{figure}
365: \includegraphics[scale=0.3]{contour}
366: \caption[]{\label{contour} Schematic representation of the Keldysh contour.
367: The lower branch is time-ordered, while the upper
368: branch is anti-time-ordered.
369: }
370: \end{figure}
371: Note that because the system is not translationally invariant, all
372: Green's functions depend on both coordinates. However, all
373: functions still only depend on the time differences because the
374: system is in steady state. The time integration in
375: Eq.~(\ref{first}) is along the contour $C$. Keldysh
376: disentanglement of the time variables results in expressions with
377: integrations only along the real axis. Then, four different
378: Green's functions emerge in accordance with four possibilities to
379: arrange the times $t$ and $t'$ along the contour. This placement
380: of the time variables is reflected in additional superscripts of
381: the Green's functions. For instance, if the time $t$ lies on the
382: time-ordered part of the contour and $t'$ on the
383: anti-time-ordered one, see Fig.~\ref{contour}, evaluation of
384: Eq.~(\ref{first}) at the zeroth order in the tunnelling (indicated
385: by a subscript $0$) yields:
386: \begin{eqnarray}
387: G^{-+}_0(x,x';t-t') &=& -i \langle T_C [c(x,t)
388: c^\dag(x',t') ]\rangle_0 \nonumber \\ \nonumber &=& i \langle
389: c^\dag(x',t') c(x,t) \rangle_0 \, .
390: \end{eqnarray}
391: Its counterpart with an interchanged orientation
392: of time variables ($t$ on the $C_+$ and $t'$ on the $C_-$) is
393: \begin{eqnarray} \nonumber
394: G^{+-}_0(x,x';t-t') = -i \langle c(x,t) c^\dag(x',t') \rangle_0 \, .
395: \end{eqnarray}
396: Green's function in which both time variables lie on the same side
397: of the contour are the usual time-ordered and anti-time-ordered ones:
398: \begin{eqnarray}
399: G^{--}_0(x,x';t-t') &=& -i \langle T [c(x,t) c^\dag(x',t')
400: ]\rangle_0 \, , \nonumber \\ \nonumber G^{++}_0(x,x';t-t') &=& -i
401: \langle \widetilde{T}[ c(x,t) c^\dag(x',t') ]\rangle_0 \, ,
402: \end{eqnarray}
403: where $\widetilde{T}$ denotes the anti-time-ordering operation.
404:
405: The local electron energy distribution function $N(x,\omega)$ in
406: the lead, which we shall also call TED by abuse of terminology,
407: is given by the defining relation with one of the Keldysh Green's
408: functions (see e.~g.~\cite{LLX}):
409: \begin{equation} \label{defEDF}
410: N(x,\omega) = - i G^{-+}(x,x; \omega) \; .
411: \end{equation}
412: The lead being non-interacting, the TED of the tunnelling
413: particles is, up to a pre-factor, given by above
414: Green's function after subtracting off
415: the equilibrium distribution function.
416: Indeed, by examining the perturbative expansion in
417: the tunnelling amplitude $\gamma$ and keeping in mind
418: the fact that there is no correlations in the lead,
419: one can easily establish the following important identity,
420: \begin{eqnarray} \label{impform}
421: &~&~G(x,x'; t-t') = G_0(x,x';t-t') \\
422: \nonumber &~&~+\gamma^2 \int\limits_C\int\limits_C dt''dt'''
423: G_0(x,0;t-t'') g(0,0;t''-t''') \\
424: \nonumber &~&~\times G_0(0,x';t'''-t') \; .
425: \end{eqnarray}
426: The corresponding diagram is shown in Fig.~\ref{equation}.
427: The function $g(0,0;t-t')$ appearing in this relation is the
428: exact one, so the formula is valid for arbitrary
429: interactions in the host and to all orders in the tunnelling amplitude.
430: \begin{figure}
431: \includegraphics[scale=0.3]{equation}
432: \caption[]{\label{equation} Schematic representation of the equation
433: (\ref{impform}). Solid line denotes the electron's Green's function of the
434: lead and the dashed ones stand for particles in the host.
435: Thick lines they correspond to exact Green's functions to all
436: orders in tunnelling and interactions while thin lines represent Green's
437: functions with tunnelling neglected.
438: }
439: \end{figure}
440:
441: By disentangling the Keldysh indices and changing over to the
442: energy representation, we extract the Green's function of interest
443: from the general expression, (\ref{impform}),
444: \begin{eqnarray} \label{dis}
445: && G^{-+}(x,x'; \omega) = G^{-+}_0(x,x';\omega) \\
446: \nonumber &+& \gamma^2 \sum_{i,j=\pm} (ij)
447: \, G^{-i}_0(x,0;\omega) g^{ij}(0,0;\omega) G^{j+}_0(0,x';\omega)\, .
448: \end{eqnarray}
449: Plugging this expression into Eq.~(\ref{defEDF})
450: one obtains the complete TED at all energies.
451: Let us take a closer look at this relation.
452: All terms on the right-hand-side, apart of the one with
453: $(i,j)=(-,+)$, contain the unperturbed $G_0^{-+}$ function.
454: The latter is, in turn, proportional to the TED in equilibrium.
455: Therefore, no matter what the Green's functions of the interacting
456: fermions look like, all these contributions represent the TED below the
457: Fermi energy (where the first term is the dominant contribution).
458: The term in Eq.~(\ref{dis}) with $(i,j)=(-,+)$ is not constrained in such
459: way and can, in principle, contribute above the Fermi edge.
460: Therefore the high-energy part of the TED is given by
461: \begin{eqnarray} \label{EDFexp}
462: N_{>}(x;\omega) &=& -i\Theta(\omega) G^{-+}(x,x; \omega) \\ \nonumber &=&
463: - \gamma^2 G^{--}_0(x,0;\omega) g_>^{-+}(0,0;\omega) G^{++}_0(0,x;\omega)
464: \; ,
465: \end{eqnarray}
466: where $\Theta(\omega)$ is the Heaviside step function and
467: $g_>^{-+}(0,0;\omega)$ stands for the high energy part of the corresponding
468: Green's function.
469:
470: Similar properties can be established for the energy resolved
471: current $j(\omega)$, when we have
472: \begin{eqnarray} \nonumber
473: j(\omega) = \frac{1}{\pi} \int \, dk \, v_k \, N(k;\omega) \,
474: \end{eqnarray}
475: where $v_k$ is the velocity of the particle with wave number $k$.
476: Making use of Eqs.~(\ref{defEDF}) and (\ref{dis}) for the
477: high-energy part of the current one obtains,
478: \begin{eqnarray} \label{tokexp}
479: j_>(\omega) &=& -i \frac{\gamma^2}{\pi} \, g^{-+}_>(0,0;\omega) \\ \nonumber
480: &\times& \int \, dk \, k \int \, d(x-x') \, e^{i k (x-x')}\\
481: \nonumber &\times& \left[G^{--}_0(x,0;\omega)
482: G^{++}_0(0,x';\omega) \right] \, ,
483: \end{eqnarray}
484: where the product in brackets is actually translationally
485: invariant (only depends on $x-x'$), reflecting the fact that the
486: excess particles only travel in one direction, away from the
487: contact, in the (non-interacting) lead.
488:
489: We now observe that both Eq.~(\ref{EDFexp}) and Eq.~(\ref{tokexp})
490: are proportional to the high-energy part of the TED of particles
491: at the tip of the interacting host,
492: \begin{eqnarray} \label{tokexp1}
493: n_>(\omega) = -i g^{-+}_>(0,0;\omega) \, ,
494: \end{eqnarray}
495: thereby proving the statement put forward at the beginning of
496: this Section. We stress again that $n_>(\omega)$ is supposed to
497: be exact both with respect to the interaction and the tunnelling.
498: Indeed, as one can easily see, this quantity is zero for a
499: noninteracting system, even if tunnelling is taken into account
500: to all orders. It is still zero for interacting hosts if
501: tunnelling is neglected or if the system is in equilibrium
502: (Luttinger theorem). Situation changes dramatically in the case
503: of an interacting host, finite tunnelling and finite bias voltage.
504: In the rest of this Section we shall explore what statements can
505: be made about the TED without assuming a concrete model for the
506: host material (other than that there is a two-body interaction
507: present).
508:
509: As mentioned in the Introduction, there is a scattering process
510: that allows for creation of electrons above the Fermi energy. Let
511: us start with perturbative analysis. At the second order in the
512: interaction, the corresponding diagram for the Green's function
513: can be constructed by annihilating the particles in the same way
514: they were created. This essentially converts the scattering
515: amplitude for a given process into the corresponding probability.
516: The result is shown in Fig.~\ref{fig3}~a). Now we need to establish
517: the way the vertices can be decorated with the Keldysh
518: indices. Trivially the outmost points should have index $-$ on
519: the outward leg and $+$ on the inward leg in accordance with the
520: type of function ($-+$) we are calculating. Furthermore, there is
521: only one possibility to assign indices to interaction vertices.
522: In order to obtain a contribution above the Fermi energy the
523: inward Green's function should be of $++$ or anti-time ordered
524: type while the outward one has to carry $--$ indices, because all
525: other possibilities contain at least one $G^{-+}_0$ factor which
526: forces the diagram to vanish above $E_F$.
527: \begin{figure}
528: %\begin{center}
529: %\epsfxsize=0.6\columnwidth
530: %\hfill
531: %\epsffile{diag3.eps}
532: %\hfill
533: %\vspace*{0.3cm}
534: \includegraphics[scale=0.3]{diag3}
535: \caption[]{\label{fig3} a) The only second order diagram
536: contributing to the TED above the Fermi energy. Solid lines
537: correspond to host electrons and the dashed one to those in the
538: lead while crosses denote the tunnelling vertices. Wiggly lines
539: represent interactions. b) This diagram has to be inserted into
540: a) at points denoted by circles in order to obtain high-order
541: diagram for the TED. }
542: %\end{center}
543: \end{figure}
544: We are left with two cross vertices of the inserted Green's function.
545: Since we do not have \emph{a priori} knowledge about them,
546: we have checked all possibilities explicitly.
547: It turns out that the only possibility is to insert the function
548: $G^{+-}_0$ because in all other cases the diagram vanishes for
549: $\omega > 0$.
550: Hence the only second-order diagram contributing to the TED
551: above the Fermi energy is the one shown in Fig.~\ref{fig3}~a).
552:
553: Let us now consider all orders in the interaction but remain at
554: the second order in the tunnelling amplitude thereby also
555: formally justifying the above choice of the Keldysh indices.
556: The way to proceed is to analyse the relevant Green's functions
557: in the Lehmann-type spectral representation.
558: In the time domain the local (tip) Green's function of
559: interest $g^{-+}(t_1,t_2)$ is given by
560: \begin{eqnarray} \nonumber
561: g^{-+}(t_1,t_2) = i \langle \psi^\dag(t_2) \psi(t_1)
562: \rangle = -i \langle T_C [\psi(t_1^-)\psi^\dag(t_2^+)] \rangle \; ,
563: \end{eqnarray}
564: where the superscripts indicate that the time variables $t_1$ and $t_2$
565: lie on the time-ordered ($T$) and anti-time-ordered ($\widetilde{T}$) parts of
566: the Keldysh contour, respectively.
567: Performing the straightforward $S$-matrix expansion in powers of the tunnelling
568: amplitude $\gamma$ one obtains at the lowest non-vanishing order
569: (second order in $\gamma$):
570: \begin{eqnarray} \label{gsum}
571: \delta g^{-+}(t_2,t_1) &=& \gamma^2 \int \, dt_3 \, dt_4 \, \sum_{ij = \pm}
572: {\cal K}^{-+}_{ij}(t_2,t_1; t_3,t_4) \nonumber \\
573: &\times& G^{ij}(t_3,t_4) \, ,
574: \end{eqnarray}
575: where ${\cal K}^{-+}_{ij}(t_2,t_1; t_3,t_4)$ are the following four-point
576: correlation functions:
577: \begin{eqnarray} \label{gammadef}
578: {\cal K}^{-+}_{+-}(t_1,t_2; t_3,t_4) &=& \langle \widetilde{T}
579: [\psi^\dag(t_2)\psi^\dag(t_3)] T[\psi(t_1) \psi(t_4)] \rangle_0
580: \, , \nonumber \\
581: {\cal K}^{-+}_{-+}(t_1,t_2; t_3,t_4) &=& \langle \widetilde{T}
582: [ \psi(t_4) \psi^\dag(t_2)] T [ \psi(t_1) \psi^\dag(t_3)] \rangle_0
583: \, , \nonumber \\
584: {\cal K}^{-+}_{--}(t_1,t_2; t_3,t_4) &=& \langle \psi^\dag(t_2)
585: T [\psi(t_1) \psi^\dag(t_3) \psi(t_4)] \rangle_0 \, , \nonumber \\
586: {\cal K}^{-+}_{++}(t_1,t_2; t_3,t_4) &=& \langle \widetilde{T}
587: [ \psi^\dag(t_2) \psi^\dag(t_3) \psi^\dag(t_4)] \psi(t_1)
588: \rangle_0 \,
589: \end{eqnarray}
590:
591: We shall first find the spectral representation of the term
592: containing ${\cal K}^{-+}_{+-}$,
593: \begin{eqnarray} \label{EDFB}
594: \delta^{(1)} g^{-+}(t_1,t_2) = \gamma^2 \int \,
595: \frac{d\epsilon}{2 \pi} G^{+-}(\epsilon) \nonumber \\
596: \times \int \, dt_3 dt_4 \, e^{-i \epsilon (t_3-t_4)}
597: {\cal K}^{-+}_{+-}(t_1,t_2; t_3,t_4) \nonumber \\
598: = -i \gamma^2 \int_{-V}^{\infty} \, d\epsilon \,
599: \rho_c(\epsilon) P^{-+}_{+-}(t_1,t_2;\epsilon) \, ,
600: \end{eqnarray}
601: where we have used the
602: bare $-+$ Green's function in the lead,
603: \begin{eqnarray} \label{G0omega}
604: G^{+-}_0(\epsilon) = -i 2 \pi \Theta(\epsilon + V) \rho_c(\epsilon) \, ,
605: \end{eqnarray}
606: containing the local density of states $\rho_c(\epsilon)$ in the lead.
607: Partial Fourier transform of the correlation function appearing in
608: the above formula is defined by
609: \begin{eqnarray} \nonumber
610: P^{-+}_{+-}(t_1,t_2;\epsilon) = \int \, dt_3 \, dt_4 \,
611: e^{-i \epsilon (t_3-t_4)} {\cal K}^{-+}_{+-}(t_1,t_2; t_3,t_4) \, .
612: \end{eqnarray}
613: According to definition (\ref{gammadef}), this correlation function contains
614: two different time orderings. Writing them down explicitly and
615: inserting between every two $\psi$ operators a complete set of exact states,
616: it is possible to perform all time integrations.
617: This procedure is a straightforward generalisation of the standard one for
618: the equilibrium case \cite{agd}.
619: The result is
620: \begin{eqnarray} \label{pis}
621: P^{-+}_{+-}(t_1,t_2;\epsilon) = \sum_\mu e^{-i(E_\mu+\epsilon)(t_1-t_2)}
622: |{\cal B}^{(1)}_\mu (\epsilon)|^2 \, ,
623: \end{eqnarray}
624: with ${\cal B}^{(1)}_\mu (\epsilon)$ defined by
625: \begin{eqnarray} \nonumber
626: {\cal B}^{(1)}_\mu(\epsilon) = \sum_\nu a_{\mu \nu}a_{\nu
627: 0}\left[\frac{1}{E_\mu-E_\nu+\epsilon+i0} +\frac{1}{E_\nu+\epsilon-i0}\right]
628: \; , \nonumber
629: \end{eqnarray}
630: where Greek indices count all possible excited states of the
631: system with energies $E_{\nu, \lambda, \mu}$ and $a_{\mu \nu}$
632: stand for matrix elements of the operator $\psi$ , $a_{\mu
633: \nu}=\langle \mu| \psi | \nu \rangle$. In order to obtain the
634: actual correction to the TED we plug the two last equations back
635: into Eq.~(\ref{EDFB}) and compute the Fourier transform of the
636: latter with respect to the time difference $t_1-t_2$,
637: \begin{eqnarray} \label{specres1}
638: \delta n^{(1)}(\omega)= - i \int \, d(t_1-t_2) \, e^{i \omega (t_1-t_2)}
639: \, \delta^{(1)} g^{-+}(t_1,t_2) \nonumber \\
640: = - 2 \pi \gamma^2 \sum_\mu \Theta(V - E_\mu - \omega) |{\cal
641: B}^{(1)}_\mu(-V)|^2 \, .
642: \nonumber
643: \end{eqnarray}
644: Obviously, all $E_\mu$'s are larger than the ground state energy $E_0$
645: (which we have set to zero).
646: Therefore the upper boundary for $\omega$ is given by $V$.
647:
648: The remaining three terms in Eq.~(\ref{gsum}) can be treated in a
649: similar manner. Repeating steps leading to (\ref{pis}), we obtain
650: for the term containing the second four-point correlation
651: function in Eq.~(\ref{gammadef}):
652: \begin{eqnarray} \nonumber
653: P^{-+}_{-+}(t_1,t_2;\epsilon) =
654: \sum_\mu e^{i (E_\mu-\epsilon)(t_1-t_2)} |
655: {\cal B}^{(2)}_\mu(\epsilon) |^2 \, ,
656: \end{eqnarray}
657: with
658: \begin{eqnarray} \nonumber
659: {\cal B}^{(2)}_\mu (\epsilon) = \sum_\nu \left[ \frac{a_{0\mu}
660: a^*_{\mu \nu}}{E_\mu - \epsilon + i0} + \frac{a^*_{0 \mu}
661: a_{\mu \nu}}{E_\nu - E_\mu - \epsilon - i0} \right] \, ,
662: \end{eqnarray}
663: the corresponding contribution to the TED being
664: \begin{eqnarray} \label{specres2}
665: \delta^{(2)}n (\omega) = 2 \pi
666: \gamma^2 \sum_\mu \Theta(-V - E_\mu - \omega) |{\cal B}^{(2)}_\mu (V) |^2 \, .
667: \end{eqnarray}
668: Contrary to Eq.~(\ref{specres1}), this term is bounded by $-V$
669: from above and hence does not contribute to the high-energy part
670: of the TED. One can use an even simpler argument in order to
671: show that the last two four-point correlation functions in
672: (\ref{gammadef}) do not contribute above the Fermi edge either.
673: By inserting only one complete set of states between the
674: time-ordered operators (i.~e. at the break of the time-ordering)
675: one obtains (for simplicity we set $t_1=0$)
676: \begin{eqnarray} \nonumber
677: P^{-+}_{--}(0,t_2;\epsilon) &=& \int \int \, dt_3 \,
678: dt_4 \, e^{-i \epsilon (t_3-t_4)} \\ &\times& \sum_\nu
679: a^*_{0 \nu} e^{-i E_\nu t_2} \langle \nu | T [\psi(0)
680: \psi^\dag(t_3) \psi(t_4) ]| 0 \rangle \, \nonumber
681: \end{eqnarray}
682: for the third term.
683: The corresponding correction to the TED is then given by
684: \begin{eqnarray}
685: &~& \delta^{(3)} n(\omega) = i \gamma^2
686: \sum_\nu a^*_{0 \nu } \delta(-E_\nu - \omega) \int d\epsilon \,
687: G^{--}_0(\epsilon) \nonumber \\ \nonumber
688: &\times& \int dt_3 d t_4
689: \langle \nu | T [\psi(0) \psi^\dag(t_3) \psi(t_4)] | 0 \rangle \,
690: .
691: \end{eqnarray}
692: Since all $E_\nu$'s are always positive this expression is nonzero
693: only for negative energies $\omega<0$.
694: The same is true for the last four-point correlation function
695: ${\cal K}^{-+}_{++}$.
696:
697: The above approach is quite general (i.~e. valid for all kinds of
698: interacting host materials) but it is also inconclusive in the sense
699: that the actual energy dependence of the TED is determined by
700: spectral weights encoded in the structure of matrix elements
701: $a_{\mu \nu}$ that is different for different hosts. One positive
702: result of the spectral method, however, is that we have
703: identified the Keldysh four-point correlation function
704: responsible for the secondary current effect: ${\cal
705: K}^{-+}_{+-}$, see Fig.~\ref{fig4}. For some models we shall
706: calculate this correlation function exactly to all orders in the
707: interaction, for other models we'll resort to perturbative
708: expansions: note that diagram a) of Fig.~\ref{fig3} is a special
709: case (second-oder expansion) of the general Fig.~\ref{fig4} term.
710:
711: Let us conclude the present Section by making one more observation
712: of a general character.
713: Returning back to the real time representation of the lead Green's
714: function (\ref{G0omega})
715: and doing simple manipulations
716: with the time integrations, we arrive at the alternative representation
717: for the TED:
718: \begin{eqnarray} \label{com}
719: &~&~n_>(\omega) = - 2i \gamma^2 \int\limits_{-\infty}^\infty dt
720: \frac{e^{i (\omega-V)t}}{t+i \alpha} \int\limits_{-\infty}^\infty d\tau_1
721: \int\limits_{-\infty}^\infty d\tau_2e^{i\omega(\tau_1+\tau_2)} \nonumber \\
722: &~&~\times
723: \langle \widetilde{T}[\psi(\tau_1)\psi(0)]
724: T[\psi(t+\tau_1)\psi(t+\tau_1+\tau_2)]\rangle_0 \, ,
725: \end{eqnarray}
726: where $\alpha$ stands for the short-time (high-energy) cutoff
727: inversely proportional to the conductance band-width $D$. For the
728: sake of simplicity, we have assumed a constant density of states
729: in the lead but this assumption is not crucial.
730: \begin{figure}
731: \includegraphics[scale=0.4]{diag5}
732: \caption[]{\label{fig4}
733: Generic diagram giving contributing to the TED above the Fermi energy.
734: The four-point correlation function ${\cal K}^{-+}_{+-}(t_1,t_2;t_3,t_4)$ is
735: represented by the shaded square. }
736: %\end{center}
737: \end{figure}
738: We now observe that if the $T$-ordering operation in the above
739: formula were dropped, the expression would vanish. Indeed,
740: inserting some complete sets one finds that then the integrand in
741: the above, proportional to
742: \begin{equation} \nonumber
743: \sim\langle 0 |\psi(t+\tau_1)|\nu\rangle\langle \nu|
744: \psi(t+\tau_1+\tau_2)|0\rangle
745: \sim e^{iE_\nu\tau_2} \, ,
746: \end{equation}
747: becomes an analytic function of the time variable $\tau_2$ in the
748: upper half plane and hence has vanishing Fourier transform for
749: $\omega>0$. By re-arranging time integrations in a different way,
750: one can easily show that the same statement is true about the
751: $\widetilde{T}$-ordering operation. Subtracting off the unordered
752: correlation function we arrive at the following remarkable
753: representation for the TED:
754: \begin{eqnarray} \label{combis}
755: &~&~n_>(\omega) = - 2i \gamma^2 \int\limits_{-\infty}^\infty dt
756: \frac{e^{i (\omega-V)t}}{t+i \alpha} \int\limits_{0}^\infty d\tau_1
757: \int\limits_{0}^\infty d\tau_2e^{i\omega(\tau_1+\tau_2)} \nonumber \\
758: &~&~\times
759: {\cal R}(\tau_1,0;t+\tau_1,t+\tau_1+\tau_2) \, ,
760: \end{eqnarray}
761: with
762: \begin{eqnarray} \label{Rdef}
763: {\cal R} (t_1,t_2;t_3,t_4)= \langle \{
764: \psi^\dag(t_1),\psi^\dag(t_2) \} \{ \psi(t_3),\psi(t_4) \}
765: \rangle \, ,
766: \end{eqnarray}
767: \vspace*{0.0cm}
768:
769: \noindent where $\{.,.\}$ stands for the anti-commutator. The
770: advantage of this representation is that it is explicitly
771: vanishing in the non-interacting case as the field operators are
772: then anti-commuting at all times. The latter does not take place
773: in interacting systems. Generally, anti-commuting interacting
774: Fermi operators at different times would result in a complicated
775: object. However, there are systems for which the so-called
776: \emph{braiding relations} hold:
777: \begin{eqnarray} \label{braiding}
778: \psi(t_1) \psi(t_2) = e^{i \varphi(t_1-t_2)} \psi(t_2) \psi(t_1) \,.
779: \end{eqnarray}
780: The exact shape of the phase function $\varphi(t)$ depends on the
781: system in question but it is usually proportional to the
782: interaction constant. The braiding relation reflects the time
783: evolution of an anti-commutators in particular models and can be
784: used to simplify (\ref{Rdef}) to
785: \begin{widetext}
786: \begin{eqnarray} \label{braiding2}
787: {\cal R}(\tau_1,0 ; t+\tau_1,t+\tau_1+\tau_2)
788: = (1+e^{-i \varphi(\tau_1)})(1+e^{-i \varphi(-\tau_2)})
789: \langle \psi^\dag(\tau_1)\psi^\dag(0)
790: \psi(t+\tau_1)\psi(t+\tau_1+\tau_2) \rangle \,
791: \end{eqnarray}
792: \end{widetext}
793: from which formula it is clear that the expansion of $n_>(\omega)$
794: in the interaction constant starts (at least) at the second order.
795: In the next Section we present two models where such braiding
796: relations exist.
797:
798: In the following four Sections we apply the formalism developed
799: here to various physical systems.
800:
801: \section{Models with local interactions} \label{seclocalinteraction}
802: We start with the simplest (toy) model we could think of:
803: the interactions are only present at one point - the tip of the
804: emitter.
805: The host Hamiltonian entering Eq.~(\ref{ham0}) is then of the form
806: \begin{equation}
807: H[\psi]=H_0[\psi]+ H_i[\psi]
808: \end{equation}
809: with the Hubbard on-site repulsion for the interaction term
810: \begin{equation} \label{intdefinition}
811: H_i = U_0 \psi^\dag_\uparrow(0) \psi^\dag_\downarrow(0) \psi_\downarrow(0)
812: \psi_\uparrow(0) \, .
813: \end{equation}
814: We take a half-infinite electron system with linear
815: dispersion relation for the left electrode
816: \begin{eqnarray} \nonumber
817: H_0[\psi]= -i \sum_s \int\limits_0^\infty \, dx \,
818: \Big[ \psi_{sR}^\dag(x) \partial_x \psi_{sR}(x) -
819: \psi_{sL}^\dag(x) \partial_x \psi_{sL}(x)\Big] \; ,
820: \end{eqnarray}
821: and the same for the lead (but, of course, at chemical potential
822: of $-V$). Here $\psi_{sR(L)}$ stands for the annihilation operator
823: of the right(left)-moving electron species with spin $s$. To
824: simplify the formulae we have set $\hbar v_F=1$.
825:
826: At first sight such model seems unphysical.
827: However, it can be regarded as a special case of a
828: correlated quantum dot contacted by two
829: noninteracting electrodes, see e.~g. Ref.~[\onlinecite{wingreen}], where the
830: contact to one of them is nearly perfect while the contact to the
831: other one is weak.
832: One should then decorate the electron field operator with a tunnelling
833: channel (transverse quantisation) index.
834: Its presence does not affect qualitative results, so we
835: drop the channel index in this Section.
836:
837: For the half-infinite host, the bare Green's functions have
838: image structure of the form
839: \begin{equation} \label{goodformula}
840: g_{s0}(x,x';\omega) = g_0(x-x';\omega) - g_0(x+x';\omega) \; ,
841: \end{equation}
842: and are diagonal and $s\to -s$ symmetric in the spin space. The
843: Green's functions appearing on the right-hand-side in the above
844: formula are the translationally invariant (bulk) ones for the
845: corresponding spinless infinite system.
846: The simplest Green's functions are the retarded and advanced ones,
847: \begin{equation} \nonumber
848: g_0^{R}(x;\omega) = [g_0^A(x;\omega)]^* = -i e^{i (k_F + \omega) |x| } \,
849: \end{equation}
850: to which the diagonal Keldysh functions are related via
851: \begin{equation} \nonumber
852: g_0^{--}(x;\omega) = \Theta(\omega) g_0^{R}(x;\omega)+
853: \Theta(-\omega) g_0^{A}(x;\omega)
854: \end{equation}
855: and $g_0^{++}(x;\omega)=-[g_0^{--}(x;\omega)]^*$,
856: while the off-diagonal Keldysh functions are:
857: \begin{eqnarray} \nonumber
858: g_0^{+-(-+)}(x;\omega)=i 2 \Theta(\pm \omega) \cos \Big[(k_F +
859: \omega)x\Big] \; .
860: \end{eqnarray}
861:
862: Plugging these expressions into Eq.~(\ref{goodformula}) one
863: obtains all Keldysh components for the half-infinite system.
864: However, using them directly to evaluate $g^{-+}_{s>}(0,0;\omega)$
865: for electrons with spin orientation $s$ causes a technical problem
866: as at the origin all wave functions of the half-infinite system
867: vanish and so do the correlation functions. The reason for that
868: is the open boundary condition imposed on the wave functions.
869: Therefore we are in need of a regularisation. We now assume that
870: the underlying lattice model has a lattice constant $a_0$ so that
871: the tunnelling occurs between points with the spatial coordinate
872: $x<a_0$ in the left subsystem as well as in the right one.
873: Obviously, such regularisation does not influence the physics but
874: only creates some non-universal numerical factors in the TED. In
875: this regularisation all zeros in Eq.~(\ref{EDFexp}) are
876: understood to be substituted by $a_0$.
877: Suppressing this argument in the formulas, we obtain the following
878: expression for the second-order correction to the local (tip)
879: Green's function (diagram Fig.~\ref{fig3} with an appropriate
880: arrangement of the spin indices):
881: \begin{eqnarray} \label{bareint}
882: &\,& g^{-+}_{s>}(\omega) = -
883: \gamma^2 4 U_0^2 \Theta(\omega) g_{0}^{++}(\omega)
884: g_{0}^{--}(\omega) \nonumber \\ &\times&
885: %\sum_{pqr} (\delta_{sr} \delta_{qp}+\delta_{sq} \delta_{rp} )
886: \int d \epsilon_1 d \epsilon_2 \,
887: g_{0}^{-+}(\omega-\epsilon_1) \nonumber
888: \\ &\times& g_{0}^{-+}(\epsilon_1+\epsilon_2)
889: g_{0}^{--}(\epsilon_2) g_{0}^{++}(\epsilon_2)
890: G_{0}^{+-}(\epsilon_2) \;
891: \end{eqnarray}
892: where the local Green's functions in the lead ($G$'s) are understood
893: to be the same as in the host but with energy measured from $-V$.
894: To evaluate this diagram, it is useful to remember that
895: \begin{eqnarray} \nonumber
896: g_{0}^{++}(\omega) g_{0}^{--}(\omega) = - |g^{R}_{0}(\omega)|^2 =
897: - 4 \sin^2 \left[(k_F + \omega) a_0\right] \; .
898: \end{eqnarray}
899: As we restrict our considerations to energy scales much smaller
900: than the Fermi energy, we can regard these products essentially
901: constant. The same is justified for other Green's functions
902: except, of course, the step functions appearing in front of
903: the off-diagonal Keldysh functions. Then, calculating energy
904: integrals in Eq.~(\ref{bareint}), we obtain the following result
905: for the TED:
906: \begin{eqnarray} \nonumber
907: n_{s>}(\omega)=-ig^{-+}_{s>}(\omega)= C_0 \, \gamma^2 \, U_0^2 \,
908: \Theta(\omega)\Theta(V - \omega) (\omega-V)^2/2 \;
909: \end{eqnarray}
910: (with the non-universal constant $C_0 = 4^5 \sin^{10} (k_F a_0)$).
911:
912: Note that in accordance with the discussion in Section \ref{secII}
913: no particles can have energy exceeding $V$ at the second order in
914: the interaction constant. Such particles, however, appear if one
915: takes into account processes of higher order. The corresponding
916: diagrams can easily be constructed, see Fig.~\ref{fig3}. So, at
917: the fourth order in the Hubbard $U$ we found
918: \begin{eqnarray} \nonumber
919: g^{-+}_{s>}(\omega) = i \, C_0^\prime \, \gamma^4 \, U_0^4 \, \Theta(\omega)
920: \Theta(2V - \omega) (\omega-2V)^4/24 \;
921: \end{eqnarray}
922: ($C_0^\prime$ is again a non-universal numerical pre-factor).
923: It is not difficult to calculate the particle spectrum for arbitrary $\omega$.
924: In the energy window between $V(n-1)$ and $Vn$ ($n=1,2,...$) it is dominated
925: by the term of the order $[\gamma U_0]^{2n}$ and shows
926: $(nV - \omega)^{2n}$ decay.
927:
928: Another interesting local model is the local phonon model,
929: where the electron density operator is coupled to a local oscillator
930: at the tip $x=0$.
931: In the chiral formulation the Hamiltonian of the problem is
932: \begin{eqnarray} \label{phononH}
933: H &=& i \int \, dx \, \psi^\dag(x) \partial_x
934: \psi(x) + \Omega b^\dag b \nonumber \\
935: &+& \lambda \psi^\dag(0) \psi(0) (b^\dag + b) \,
936: \end{eqnarray}
937: where $b^\dag$ and $b$ are the creation and annihilation
938: operators of the local phonon. The chirality of fermions takes
939: care of the reflecting boundary condition at the origin (see
940: e.~g.~Ref.~\cite{fg}) where the tunnelling processes take place.
941: In the anti-adiabatic approximation (justified for high-frequency
942: phonons, $\Omega\gg\lambda/a_0$), this Hamiltonian can be solved
943: via the Fr\"ohlich transformation \cite{frolich}
944: \begin{eqnarray} \label{Qtrafo}
945: H' = e^{Q} H e^{-Q} \, ,
946: \end{eqnarray}
947: with
948: \begin{eqnarray} \label{Qdef}
949: Q = \frac{\lambda}{\Omega} \psi^\dag(0) \psi(0) (b^\dag - b) \, .
950: \end{eqnarray}
951:
952: Simple transformations of the field operators, which we omit here,
953: show that braiding relations are satisfied for this particular
954: model and the braiding phase $\varphi$ defined in
955: Eq.~(\ref{braiding}) is equal to:
956: \begin{eqnarray} \label{sinphase}
957: \varphi(t) = \pi + \frac{1}{2} \left(\frac{\lambda}{\Omega a_0}\right)^2
958: \sin [\Omega t] \, .
959: \end{eqnarray}
960: Using this result when evaluating Eqs.~(\ref{com}) and (\ref{Rdef})
961: we obtain for the TED:
962: \begin{widetext}
963: \begin{eqnarray}
964: n_>(\omega) &=& -2 i \gamma^2 \int\limits_{-\infty}^{\infty} \,
965: dt \, \frac{e^{i(\omega-V)t}}{(t+i \alpha)^2} \nonumber
966: \int\limits_0^\infty \, d\tau_1 \, \int\limits_0^\infty \,
967: d\tau_2 \, e^{i\omega(\tau_1+\tau_2)}(1-e^{i (\lambda/\Omega
968: a_0)^2 \sin[\Omega \tau_1]/2})
969: (1-e^{-i (\lambda/\Omega a_0)^2 \sin[\Omega \tau_2]/2}) \nonumber \\
970: &\times& \frac{\tau_1 \tau_2}{(\tau_1+\tau_2+t+i\alpha)(\tau_1+t+i
971: \alpha)(\tau_2+t+i \alpha)} \, . \nonumber
972: \end{eqnarray}
973: \end{widetext}
974: Expanding in powers of $\lambda/\Omega a_0$ the evaluation of
975: these integrals yields the main contribution to the TED
976: %(neglecting some logarithmic corrections)
977: of the form
978: \begin{eqnarray} \label{phononTED}
979: n_>(\omega) \approx \frac{\pi \gamma^2}{2} \left( \frac{\lambda}{\Omega a_0}
980: \right)^4 \Theta(V-\omega) \frac{\omega (V-\omega)}{\Omega^2-\omega^2} \, .
981: \end{eqnarray}
982: As expected, the emerging spectrum has a resonant character
983: having a sharp pole at the oscillator frequency $\Omega$. Under
984: appropriate conditions, the Fr\"ohlich transformation can also be
985: used for solving the bulk electron-phonon interaction where the
986: braiding relations still persist. We shall not pursue this issue
987: further in this paper.
988:
989: \section{Luttinger liquid model} \label{luttinger}
990: Staying with 1D hosts, we now move on to the next level of difficulty
991: and discuss the bulk interactions.
992: The simplest model here is the spinless Luttinger liquid model \cite{haldane}.
993: Subject to minor modifications (see below) the LL results will also be
994: applicable to quantum wires and SWNTs.
995: The interacting term in the host Hamiltonian then is
996: \begin{equation} \nonumber
997: H_i = \int \, dx \, dy \, \psi^\dag(x)
998: \psi^\dag(y) U(x-y) \psi(y) \psi(x) \, .
999: \end{equation}
1000: where $U(x)$ is the interaction potential.
1001:
1002: Although this model can be exactly solved,
1003: it is instructive to start with the perturbative expansion.
1004: Diagram a) in Fig.~\ref{fig3} still represents the only
1005: non-vanishing contribution to the TED above the Fermi edge.
1006: Compared to Eq.~(\ref{bareint}), the analytic expression for this diagram is
1007: slightly more complicated since all participating Green's
1008: functions acquire spatial dependence.
1009: Assuming local interaction, $U(x-y)\to U_0\delta(x-y)$, we write
1010: \begin{widetext}
1011: \begin{eqnarray} \label{bareint1}
1012: g^{-+}_>(\omega) &=& - \gamma^2 U_0^2 \Theta(\omega) \int_{a_0}^\infty \,
1013: dx_1 \, dx_2 \, g_0^{++}(x_2,a_0;\omega)
1014: g_0^{--}(a_0,x_1;\omega) \int d \epsilon_1 d
1015: \epsilon_2 \, g_0^{-+}(x_1,x_2;\omega-\epsilon_1) \nonumber \\
1016: &\times& g_0^{-+}(x_1,x_2;\epsilon_1+\epsilon_2)
1017: g_0^{--}(x_1,a_0;\epsilon_2)
1018: g_0^{++}(a_0,x_2;\epsilon_2) G_0^{+-}(a_0,a_0;\epsilon_2) \; .
1019: \end{eqnarray}
1020: %\end{widetext}
1021: We use the same regularisation as in Section \ref{seclocalinteraction}
1022: (that is why the lower boundary of the $x$-integration is the lattice
1023: constant $a_0$).
1024: Substituting the expressions for the corresponding bare Green's functions
1025: into this formula one obtains the following result:
1026: \begin{eqnarray} \label{LLfactorize}
1027: g^{-+}_>(\omega) &=& i \gamma^2 U_0^2 \, 8 C_0 \, \Theta(\omega)
1028: \Theta(V-\omega) \int_\omega^{V} \, d\epsilon_1
1029: \int_{-V}^{-\epsilon_1} \, d \epsilon_2 \, |
1030: H(\epsilon_1,\epsilon_2,\omega)|^2 \, ,
1031: \end{eqnarray}
1032: %\end{widetext}
1033: where $C_0$ is a non-universal numerical constant and the
1034: function $H$ is defined as
1035: \begin{eqnarray} \nonumber
1036: H(\epsilon_1,\epsilon_2,\omega) = \frac{1}{2} \int_0^\infty \, dx
1037: \, e^{i(\omega-\epsilon_2)x} \cos[(2 \epsilon_1 + \epsilon_2 -
1038: \omega)x] = -\frac{1}{4i}
1039: \frac{\epsilon_2-\omega}{\epsilon_1(\epsilon_1 + \epsilon_2 -
1040: \omega)} \, .
1041: \end{eqnarray}
1042: \end{widetext}
1043: We are ignoring terms containing rapidly oscillating integrands
1044: which are negligible on energy scales much smaller than $E_F$.
1045:
1046: Performing the last energy integrals we obtain
1047: for the TED above the Fermi edge
1048: \begin{eqnarray} \nonumber
1049: n_>(\omega) = \gamma^2 U_0^2 \, \Theta(\omega) \Theta(V - \omega) C_0 \Big\{
1050: \frac{V-\omega}{\omega} + F_0\Big(\frac{\omega}{V}\Big)
1051: \Big\} \, ,
1052: \end{eqnarray}
1053: where $F_0(x)$ is a function regular at $x=0$ and vanishing for
1054: $x>1$. While the TED vanishes smoothly in the vicinity of $\omega=V$,
1055: there is a sharp singularity towards the Fermi edge.
1056:
1057: Since the Luttinger liquid model is solvable for arbitrary
1058: interaction strength, \cite{haldane}, it is possible to go beyond
1059: the perturbative expansion. In particular, the four-point
1060: correlation function in Eq.~(\ref{com}) can be calculated exactly.
1061: We remind the reader that the system under consideration is a
1062: half-infinite spinless LL with an open boundary across which
1063: tunnelling processes occur. Using the standard bosonization
1064: scheme for open boundary LLs (technical details can be found in
1065: the literature, see \cite{fg,book}), we write
1066: \begin{equation} \label{bosrepr}
1067: \psi(x=0,t) = (2 \pi a_0)^{-1/2} \exp [i \phi(x=0,t)/\sqrt{g}] \,
1068: \end{equation}
1069: where $g$ is the LL parameter, defined by $g=(1+4 U_0/\pi)^{-1/2}$.
1070: The Gaussian chiral Bose field $\phi(x,t)$ has been rescaled
1071: and is governed by the LL Hamiltonian
1072: \begin{eqnarray}
1073: H_{LL}[\phi] = \frac{1}{4 \pi} \int_{-L}^{L} \, dx \, (\partial_x \phi)^2 \, .
1074: \end{eqnarray}
1075: The Bose field is periodic in $2L$, where $L$ is the system size.
1076: In order to obtain the correct analytic properties of correlation
1077: functions we first work with a system of a finite length $L$,
1078: which makes the energy quantisation equal to $\epsilon_0 =
1079: \pi/L$, and then send $L$ to infinity. Using representation
1080: (\ref{bosrepr}) we obtain the four-point correlation function in
1081: question:
1082: \begin{eqnarray} \label{GF}
1083: {\cal K}^{-+}_{+-}(t_1,t_2; t_3,t_4)= (2 \pi b)^{-2}
1084: \mbox{sgn}(t_2-t_3) \mbox{sgn}(t_1-t_4) \nonumber \\
1085: \times \Big[\frac{F(|t_2-t_3|)F(|t_1-t_4|)F^2(0)}{F(t_2-t_1)F(t_2-t_4)F(t_3-t_1)F(t_3-t_4)}
1086: \Big]^{1/g} \, ,
1087: \end{eqnarray}
1088: where the function $F(t)$ is defined by
1089: \begin{eqnarray} \nonumber
1090: F(t)=1-e^{i\epsilon_0 (t+i \delta)} \, .
1091: \end{eqnarray}
1092: In the thermodynamic limit $L\rightarrow \infty$, we expand Eq.~(\ref{GF})
1093: in powers of $\epsilon_0$.
1094: Then Eq.~(\ref{com}) can be brought into the following form
1095: \begin{widetext}
1096: \begin{eqnarray} \label{genLLres}
1097: n(\omega) &=& -2 i \gamma^2 e^{-i \pi/g} \cos^2(\pi/2g) \int dt
1098: \, \frac{e^{i(\omega-V)}}{(-t-i \alpha)^{1+1/g}} \int_0^\infty
1099: \int_0^\infty \, d\tau_1 d\tau_2 e^{i\omega(\tau_1+\tau_2)}
1100: \nonumber \\ &\times& \frac{(\tau_1
1101: \tau_2)^{1/g}}{[(-\tau_1-\tau_2-t-i\alpha) (-\tau_1 -t -i \alpha)
1102: (-\tau_2-t -i\alpha)]^{1/g}} \, , \nonumber
1103: \end{eqnarray}
1104: \end{widetext}
1105: where we used the fact that the LL field operators satisfy the
1106: braiding relations, (\ref{braiding}), with the braiding phase
1107: function
1108: \begin{eqnarray}
1109: \varphi(t) = \frac{\pi}{g}\, \mbox{sign} (t) \, .
1110: \end{eqnarray}
1111: Now we can perform the $\tau$-integrations
1112: using the integral representation
1113: \begin{eqnarray} \nonumber
1114: (\tau_1+\tau_2-t-i\alpha)^{-1/g} &=& \frac{e^{i \pi/2g}}{\Gamma(1/g)}
1115: \int_0^\infty dp \, p^{1/g-1} \\
1116: &\times& e^{-i(\tau_1+\tau_2-t)p} \, . \nonumber
1117: \end{eqnarray}
1118: Then
1119: \begin{eqnarray}
1120: &~& n_>(\omega) = 2 i \gamma^2 e^{-i \pi/g} \cos^2(\pi/2g) \\
1121: &\times& \int\limits_{-\infty}^\infty dt \,
1122: \frac{e^{i(\omega-V)}}{(-t-i \alpha)^{1+1/g}}
1123: \nonumber \\ &\times& \Gamma^2(1/g+1)
1124: \int_0^\infty \, dp \, p^{1/g-3} e^{i p t} \Psi^2(1/g,0; -i (p+\omega)t) \, , \nonumber
1125: \end{eqnarray}
1126: where $\Gamma$ and $\Psi$ are the gamma function and the Tricomi confluent
1127: hyper-geometric function, respectively \cite{bateman}.
1128: Deforming the $t$-integration contour from the real axis to
1129: the contour around the branch cut of the function $\Psi$ we arrive at a
1130: more convenient representation for the TED:
1131: \begin{eqnarray} \label{finalint}
1132: n_>(\omega) = A(g) \int_0^{V-\omega} dE \, \frac{E^{1/g-1}}{(E+\omega)^2}
1133: F_V(E+\omega) \, ,
1134: \end{eqnarray}
1135: where the spectral function is given by
1136: \begin{eqnarray} \label{psiintegral}
1137: F_V(p) = 2 \, \mbox{Im}\, e^{-i\pi/g} \int_\alpha^\infty d\xi \, \xi^{-1/g-1}
1138: e^{-(V-p)\xi} \nonumber \\
1139: \times \Psi^2(1/g,0,-p \xi+i\alpha) \, ,
1140: \end{eqnarray}
1141: and the pre-factor is
1142: \begin{eqnarray} \nonumber
1143: A(g) = \Gamma^2(1/g+1) \gamma^2 \alpha^{2/g} \frac{\cos^2[\pi/ 2 g]}{2 \pi
1144: a_0^2 \Gamma(1/g)} \; .
1145: \end{eqnarray}
1146: In the limit of small energies, $\omega/V \ll 1$, the function
1147: $\Psi$ is nearly constant and the limiting form
1148: of the spectral function can easily be established:
1149: \[
1150: F_V(p) \approx \frac{2 \pi}{\Gamma^3(1+1/g)} (V-p)^{1/g} \, .
1151: \]
1152: Therefore we obtain the following asymptotic form
1153: for the TED in vicinity of the Fermi edge
1154: \begin{eqnarray} \label{cresult}
1155: n(\omega) \approx C_2 (\lambda+1)^2 (\omega/V)^{\lambda} \, ,
1156: \end{eqnarray}
1157: where $\lambda = 1/g-2$ and $C_2$ is a
1158: numerical pre-factor regular at $\lambda=-1$.
1159:
1160: We investigated the behaviour of the TED near the upper threshold
1161: $\omega=V$ by numerical evaluation of Eq.~(\ref{finalint}). The
1162: result is given by another power law:
1163: \begin{equation} \label{exprelation}
1164: n(\omega) \sim (V-\omega)^\nu \,\, , \,\, \nu=\lambda+2 \, .
1165: \end{equation}
1166:
1167: In the limit of weak interactions, $g\rightarrow 1$, the exponent
1168: $\lambda$ approaches the perturbative result. Eq.~(\ref{cresult})
1169: can be regarded as consisting of two factors. The first factor is
1170: the universal $1/\omega$ divergence inherent to all interacting
1171: 1D systems. The second factor reflects power-law renormalisations
1172: occurring in the LLs and, in particular, contains the local
1173: density of states (LDOS) of the primary electron. The latter
1174: object is suppressed for repulsive interactions ($g<1$). So also
1175: the TED singularity is suppressed. At $g_c=1/2$ the LDOS
1176: suppression effectively wins over and the singularity disappears.
1177: It is noteworthy that in the case when the interaction constant is
1178: smaller than this critical value, the TED has a maximum between
1179: the upper and lower thresholds $\omega=0$ and $\omega=V$, as it
1180: is vanishing towards both limits.
1181:
1182: In real systems the tunnelling amplitude $\gamma$ is small. The
1183: secondary current is already calculated in the next-to-leading
1184: (compared to to the primary current) approximation in $\gamma$.
1185: Still, it is a valid question to ask what happens at higher
1186: orders in the tunneling amplitude. For the case when the TED is
1187: divergent at the Fermi level ($\lambda<0$), a small energy scale
1188: $\omega^*$ may emerge below which these higher order
1189: contributions become important. We do not currently have full
1190: answer to this question. Our preliminary calculations indicate,
1191: however, that the most divergent 4-th order diagram is the one
1192: for which Fig.~\ref{fig4} acts as self-energy. This would lead to
1193: the estimate $\omega^*\sim V\gamma^{-2/\lambda}$. This issue
1194: deserves further investigation.
1195:
1196: Thus far we have discussed the spinless LL. The above results can
1197: be straightforwardly generalised to spinful systems such as
1198: quantum wires. Due to the spin-charge separation the LDOS at the
1199: end of the wire is then given by the same expression as in the
1200: spinless case with $1/g$ substituted by $(1/g_c + 1/g_s)/2$,
1201: where $g_{c,s}$ are the interaction constants in the charge and
1202: spin sectors, respectively \cite{book}. Since the tunnelling
1203: processes conserve spin, the LDOS exponent is simply carried over
1204: to the TED, so that $\lambda$ in Eq.~(\ref{cresult}) should be
1205: changed to:
1206: \begin{equation} \label{lwires}
1207: \lambda\rightarrow\lambda=(1/g_c+1/g_s)/2-2\,.
1208: \end{equation}
1209: The relation between the lower and upper threshold exponents,
1210: Eq.~(\ref{exprelation}), is still valid. When the spin SU(2)
1211: symmetry is preserved, it forces $g_s=1$. The residual
1212: spin-backscattering interaction renormalises to zero under
1213: renormalisation group transformations (see \cite{book} for a
1214: recent review on 1D physics and further references). However, the
1215: presence of an irrelevant operator should probably cause
1216: multiplicative logarithmic corrections to our power-law formula
1217: (\ref{cresult}). We have not attempted to calculate these
1218: correction though we expect that this can be done by standard
1219: methods \cite{affleck}.
1220:
1221: \section{Tunnelling from 2D systems: pure limit} \label{pure}
1222: In view of applications to MWNTs, we now consider tunnelling from
1223: a 2D electron system in tip geometry. Theoretically, this is a
1224: more complicated situation than in 1D. In some cases, relevant
1225: non-perturbative techniques exist, in other cases we shall present
1226: perturbative results. In this Section we discuss interacting
1227: 2D electron systems free from impurities (pure limit).
1228:
1229: From the mathematical point of view we still have to calculate the same
1230: diagram [(a) in Fig.~\ref{fig3}].
1231: To accomplish this task in 2D it is convenient to change over to
1232: the momentum representation.
1233: The host Hamiltonian is $H[\psi]=H_0[\psi]+H_i[\psi]$,
1234: where
1235: \begin{equation}\label{H02D}
1236: H_0[\psi]=\int\frac{d\vec{p}}{2\pi}\xi(\vec{p})\psi_{\vec{p}}^{\dag}
1237: \psi_{\vec{p}}
1238: \end{equation}
1239: describes free 2D electron gas with dispersion relation $\xi(\vec{p})$
1240: and $\psi_{\vec{p}}$ is the Fourier transform of the electron field operator.
1241: The interaction term has the standard form. In real space:
1242: \begin{equation}\label{Hi2D}
1243: H_i = \int \, d\vec{r} \, d\vec{r'} \, \psi^\dag(\vec{r})
1244: \psi^\dag(\vec{r'}) U_0(|\vec{r}-\vec{r'}|) \psi(\vec{r'}) \psi(\vec{r}) \, ,
1245: \end{equation}
1246: where $U_0(r)$ is the interaction potential to be specified later.
1247: The relevant diagram is given by the following expression:
1248: \begin{eqnarray} \label{wholecontrib}
1249: g^{-+}_>(\omega) &=& \int \, \frac{d^2 \vec{q_1}}{(2\pi)^2} \, \int
1250: \frac{d^2\vec{q_2}}{(2\pi)^2} \, \int \frac{d \Omega}{2 \pi}
1251: \int \frac{d^2\vec{p}}{(2\pi)^2} \, \nonumber
1252: \\ &\times& g^{--}_0(\omega,\vec{p}-\vec{q_1})
1253: g^{-+}_0(\omega-\Omega,\vec{p})
1254: \\ \nonumber &\times&
1255: U_0(q_1)\Pi(\Omega;\vec{q_1},\vec{q_2})U_0(q_2)
1256: g^{++}_0(\omega,\vec{p}-\vec{q_2})
1257: \end{eqnarray}
1258: Here $U_0(q)$ is the Fourier transform of the interaction potential,
1259: $\Pi(\Omega;\vec{q_1},\vec{q_2})$ stands for the inner polarisation
1260: bubble as in Fig.~\ref{fig3} (b) with vertex indices as in (a):
1261: \begin{eqnarray} \label{bubbledef}
1262: \Pi(\Omega;\vec{q_1},\vec{q_2}) &=& \int \frac{d \epsilon}{2 \pi}
1263: \int \frac{d^2
1264: \vec{p}}{(2\pi)^2} \, g^{--}_0(\epsilon,\vec{p}+\vec{q_1}) \nonumber
1265: \\
1266: &\times& G^{-+}_0(\epsilon) g^{++}_0(\epsilon,\vec{p}+\vec{q_2})
1267: g^{-+}_0(\epsilon+\Omega,\vec{p}) \, . \nonumber
1268: \end{eqnarray}
1269: Without the tunnelling insertions, this would be a Keldysh analogy
1270: to the standard 2D Lindhard functions \cite{benard}.
1271: We define $\vec{p} = P_\omega \vec{n}_\phi$ with a unit 2D vector
1272: $\vec{n}_\phi$ in direction parametrised by the angle $\phi$.
1273: The quantity $P_\omega$ gives the on-shell magnitude of the
1274: momentum $\vec{p}$ and, for a quadratic dispersion relation,
1275: satisfies the following equation,
1276: \begin{eqnarray} \label{Pdef}
1277: \frac{P_\omega^2}{2 m} = \frac{p_F^2}{2 m} + \omega \, ,
1278: \end{eqnarray}
1279: where $p_F$ is the Fermi momentum and $m$ is the effective mass.
1280:
1281: We change over from the momentum integration to the energy-angle
1282: integration in the standard manner:
1283: \begin{eqnarray} \nonumber
1284: \int \frac{d^2 \vec{p}}{(2 \pi)^2} \rightarrow \frac{1}{2 \pi} \int d \xi
1285: \rho(\xi) \int_0^{2 \pi} d \phi \, ,
1286: \end{eqnarray}
1287: where $\rho(\xi)$ is the density of states in the host.
1288: The polarisation bubble can then be written in the form
1289: \begin{widetext}
1290: \begin{eqnarray}
1291: \Pi(\Omega;\vec{q_1},\vec{q_2}) = \rho_c \Theta(\Omega-V)
1292: \int_{-V}^{-\Omega} \, d\epsilon \, \rho(\epsilon + \Omega)
1293: \nonumber \int_0^{2 \pi}
1294: d\phi g^{--}_0(\epsilon,P_{\epsilon+\Omega} \vec{n}_\phi + \vec{q_1})
1295: g^{++}_0(\epsilon,P_{\epsilon+\Omega} \vec{n}_\phi + \vec{q_2}) \, ,
1296: \nonumber
1297: \end{eqnarray}
1298: \end{widetext}
1299: where $\rho_c$ stands for the constant density of states in the lead.
1300: The diagonal bare Keldysh Green's functions are
1301: \begin{eqnarray} \nonumber
1302: & & g^{--(++)}_0(\epsilon,P_{\epsilon+\Omega}\, \vec{n}_\phi + \vec{q})
1303: \\ \nonumber
1304: &=& \mp
1305: \left[\Omega + \frac{1}{m}P_{\epsilon+\Omega}\,\vec{n}_\phi \vec{q} +
1306: \frac{1}{2 m} q^2 \pm i0 \right]^{-1} \; .
1307: \end{eqnarray}
1308: In this expression we have taken into account the fact that $\epsilon < 0$
1309: if we are only interested in $\omega>0$ part of $g^{-+}$.
1310: The off-diagonal Keldysh function is
1311: \begin{eqnarray} \nonumber
1312: g^{-+}_0(\omega-\Omega,\vec{p}) = i 2 \pi n_F(\vec{p})\, \delta(
1313: \omega-\Omega-\xi_{\vec{p}} ) \, ,
1314: \end{eqnarray}
1315: where $n_F(\vec{p})$ is the Fermi momentum distribution
1316: function. Using these expressions and relation (\ref{tokexp1})
1317: between the Green's function and the TED we obtain a formula
1318: similar to Eq.~(\ref{LLfactorize}) for the 1D case, namely
1319: \begin{eqnarray} \label{nbol}
1320: n_{>}(\omega) = \frac{\rho_c}{2 \pi} \int_\omega^V \, d \Omega \,
1321: \rho(\omega-\Omega) \int_{-V}^{-\Omega} \, d\epsilon \, \rho(\epsilon+\Omega)
1322: \\ \nonumber \times
1323: \int_0^{2 \pi} \, d\phi \, d\phi' |F_{\epsilon}(\Omega,\phi-\phi')|^2 \,
1324: ,
1325: \end{eqnarray}
1326: where the function $F$ is defined by
1327: \begin{eqnarray} \label{woher}
1328: & & F_{\epsilon}(\Omega,\phi) = \frac{1}{4 \pi^2} \int_0^\infty
1329: dq \, q \, U_0(q) \int_0^{2 \pi} d\theta
1330: \nonumber \\ \nonumber
1331: &\times& [\Omega + \frac{1}{m} P_{\epsilon+\Omega} q
1332: \cos(\theta-\phi/2) + \frac{1}{2m} q^2 + i0]^{-1}
1333: \\ \nonumber
1334: &\times&
1335: [\Omega + \frac{1}{m} P_{\epsilon+\Omega} q
1336: \cos(\theta+\phi/2) - \frac{1}{2m} q^2 + i0 ]^{-1} \, .
1337: \end{eqnarray}
1338:
1339: To proceed, we now need to specify the interaction potential. It
1340: is natural to start with a contact interaction term,
1341: $U(\vec{r}-\vec{r'}) = U_0\,\delta(\vec{r}-\vec{r'})$ or
1342: $U_0(q)=U_0$ in the momentum space. However, an unexpected
1343: technical difficulty arises. If we follow the standard practice
1344: and linearise the electron spectrum in the vicinity of the Fermi
1345: surface, then the momentum integration in Eq.~(\ref{woher}) can
1346: be easily done but the remaining angle integration diverges at
1347: $\varphi=\pm\pi$. This latter singularity comes from large
1348: momenta in the previous momentum integral and corresponds to the
1349: special case of hot hole back-scattering. To obtain a finite
1350: result for the TED one either needs to take into account the
1351: decay of $U_0(q)$ for large $q$ or the non-linearity of the
1352: electron spectrum. Accordingly, we present two different
1353: calculations in the rest of this Section.
1354:
1355: {\bf (i)} Contact potential: $U_0(q)=U_0$.
1356:
1357: First we scale the integration variable $q
1358: \rightarrow \Omega q$ and introduce dimensionless quantities
1359: $\gamma_{1,2}= P_{\epsilon
1360: \pm \Omega}/ p_F$ and $\kappa = \Omega/4 E_F$.
1361: Then,
1362: \begin{eqnarray} \label{transformed}
1363: & & F_{\epsilon}(\Omega,\phi) = \frac{U_0}{4 \pi^2 v_F^2}
1364: \int_0^\infty dq q \int_0^{2 \pi} d\theta
1365: \nonumber \\ \nonumber
1366: &\times& [1+ \gamma_1 q \cos(\theta - \phi/2) + \kappa q^2 + i0]^{-1}
1367: \\ \nonumber
1368: &\times& [1+ \gamma_2 q \cos(\theta + \phi/2) - \kappa q^2 + i0]^{-1} \, .
1369: \end{eqnarray}
1370: Notice that $\gamma$'s approach unity whereas $\kappa$ linearly
1371: tends to zero in the limit of small $\Omega$.
1372: Furthermore, if we set $\kappa=0$, the angle integration is still divergent.
1373: Therefore we set $\gamma_{1,2}=1$ and keep $\kappa$ finite.
1374: As a next step we expand the integrand in Eq.~(\ref{transformed})
1375: using the standard formula \cite{gelfand},
1376: \begin{eqnarray} \label{gshilov}
1377: \frac{1}{f(x)+i 0} = {\cal P}
1378: \frac{1}{f(x)} - i \pi \delta(f(x)) \, ,
1379: \end{eqnarray}
1380: where ${\cal P}$ denotes the principal value. This expansion
1381: produces three contributions: the first one contains the product
1382: of two principal parts and remains regular for $\Omega
1383: \rightarrow 0$, the second one contains products of one principal
1384: part and one delta function and is identically zero in the low
1385: energy regime, and the third contribution, $F^\delta_{\epsilon}$,
1386: containing a product of two delta functions is responsible for
1387: the divergency or the angle integral. Setting $v_F=1$ for the
1388: rest of this Section, we write
1389: \begin{widetext}
1390: \begin{eqnarray}
1391: F^\delta_{\epsilon}(\Omega,\phi) = \frac{U_0}{4 \pi^2}
1392: \int_0^\infty dq q \int_0^{2 \pi} d\theta
1393: \nonumber
1394: \delta[ 1+ q \cos(\theta - \phi/2) + \kappa q^2 ]
1395: \delta[ 1+ q \cos(\theta + \phi/2) - \kappa q^2 ] \, .
1396: \end{eqnarray}
1397: The evaluation of the angle integration leaves us with
1398: \begin{eqnarray}
1399: F^\delta_{\epsilon}(\Omega,\phi) = \frac{U_0}{4 \pi^2}
1400: \int_0^\infty dq \, q \,[q^2-(1+\kappa q^2)^2]^{-1/2}
1401: \nonumber \nonumber
1402: \delta[ 1-(1+\kappa q^2)\cos \phi + \sin \phi
1403: \sqrt{q^2-(1+\kappa q^2)^2}-\kappa q^2] \, .
1404: \end{eqnarray}
1405: \end{widetext}
1406: The argument of the delta function is zero for
1407: \begin{eqnarray} \nonumber
1408: q_0^2 = \frac{1}{2 \kappa^2} \sin^2(\phi/2)
1409: \Big( 1-\sqrt{1-16 \frac{\kappa^2}{\sin^2 \phi} } \Big) \, .
1410: \end{eqnarray}
1411: Therefore the remaining momentum integration yields
1412: \begin{eqnarray} \label{calculdF}
1413: F^\delta_{\epsilon}(\Omega,\phi) &=&
1414: \frac{U_0}{2 \pi^2} \Big[ \sin \phi + \cos \phi \Big( \sin \phi
1415: \nonumber \\ \nonumber &-& \sqrt{\sin^2 \phi - 16 \kappa^2} \Big)
1416: - 4 \kappa \sin \phi \Big]^{-1} \, .
1417: \end{eqnarray}
1418: Plugging this expression into Eq.~(\ref{nbol}) and taking into account
1419: that for small $\Omega$ (and hence small $\kappa$) the main contribution
1420: to the angle integration originates for $\phi$ close to $\pm\pi$
1421: one finds for the following limiting form for the TED:
1422: \begin{eqnarray} \nonumber
1423: n_>(\omega) \approx \frac{U_0^2}{8 \pi^4} \rho_c \nu^2 E_F V
1424: \ln \left(\frac{V}{\omega}\right) \, ,
1425: \end{eqnarray}
1426: where $\nu = \rho(0)$.
1427:
1428: {\bf (ii)} Yukawa potential: $U_0(q) = 4 \pi e^2/\sqrt{q^2+\lambda^2}$.
1429:
1430: Here $1/\lambda$ is the screening length. As $U_0(q)$ now tends
1431: to zero for large $q$ (which is, of course, always the case in
1432: real systems), we can linearise the dispersion relation. Then the
1433: function $F(\Omega,\phi)$ becomes independent of the energy
1434: variable $\epsilon$ and simplifies considerably:
1435: \begin{eqnarray} \label{newF}
1436: F(\Omega,\phi) = \frac{1}{4 \pi^2} \int_0^\infty \, dq \, q \,
1437: U_0(\Omega q) \, I(q,\phi) \, ,
1438: \end{eqnarray}
1439: where $q$ has again been scaled by $\Omega$ and the function $I(q,\phi)$
1440: is defined by
1441: \begin{eqnarray} \label{Idef}
1442: I(q,\phi) &=& \int_0^{2 \pi} \, d\theta \,\Big[(1+ q \cos(\theta-\phi/2)+i0)
1443: \nonumber \\ &\times& (1+ q \cos(\theta+\phi/2)+i0)\Big]^{-1} \, .
1444: \end{eqnarray}
1445: This integral can be calculated and is real
1446: \begin{eqnarray} \label{I1}
1447: I(q,\phi) = \frac{2 \pi}{[1-q^2 \cos^2(\phi/2)]\sqrt{1-q^2}} \, ,
1448: \end{eqnarray}
1449: for $q<1$ but contains an imaginary part
1450: \begin{eqnarray} \label{I2}
1451: I(q,\phi) &=& -i \frac{2 \pi}{[1-q^2 \cos^2(\phi/2)]\sqrt{q^2-1}}
1452: \\ \nonumber
1453: &-& \frac{\pi^2}{\sqrt{q^2-1} \cos(\phi/2)} \delta[q-1/\cos(\phi/2)] \, .
1454: \end{eqnarray}
1455: for $q>1$, as incoherent particle production takes place in the
1456: latter regime. Using Eqs.~(\ref{I1}) and (\ref{I2}) we compute the
1457: momentum integral in Eq.~(\ref{newF}) which yields
1458: \begin{eqnarray} \label{finalF}
1459: F(\Omega,\phi) &=& - \frac{2 e^2}{|\sin(\phi/2)|
1460: \sqrt{\Omega^2+\lambda^2 \cos^2(\phi/2)}} \nonumber \\
1461: &\times& \mbox{arccot} \Big[ \frac{\sqrt{\Omega^2 + \lambda^2
1462: \cos^2(\phi/2)}}{\lambda |\sin(\phi/2)|} \Big] \, .
1463: \end{eqnarray}
1464: Unfortunately we were not able to perform the last remaining angle
1465: integration in Eq.~(\ref{nbol}) in a closed form. However, to
1466: analyse the TED close to the Fermi surface $\omega\rightarrow 0$
1467: one merely requires the knowledge of $|F(\Omega,\phi)|^2$ for
1468: small $\Omega$'s. The latter can be easily read off
1469: Eq.~(\ref{finalF}) for $\Omega \ll \lambda$. There are two
1470: different regions in the parameter space $(\phi, \Omega)$: for
1471: most angles, $|\phi\pm \pi|> 2\Omega/\lambda$, we find
1472: \begin{eqnarray} \nonumber
1473: |F(\Omega,\phi)|^2 \approx \frac{4 e^4}{\lambda^2} \frac{\phi^2}{\sin^2 \phi}
1474: \end{eqnarray}
1475: whereas for large scattering angles $|\phi\pm \pi|< 2\Omega/\lambda$
1476: one obtains a singular in $\Omega$ behaviour,
1477: \begin{eqnarray} \nonumber
1478: |F(\Omega,\phi)|^2 \approx \frac{\pi^2 e^2}{\Omega^2} \, .
1479: \end{eqnarray}
1480: The latter region, physically corresponding to hot hole
1481: back-scattering processes, is responsible for the leading
1482: contribution to the remaining integrals. With logarithmic
1483: accuracy we thus obtain the following threshold behaviour of the
1484: TED,
1485: \begin{eqnarray} \label{logresult}
1486: n_>(\omega) \approx \rho_c \nu^2 \frac{2 \pi e^4}{\lambda}
1487: \ln \left[ \frac{\mbox{max}(V,\lambda)}{\omega} \right] \, .
1488: \end{eqnarray}
1489: Clearly the same logarithmic singularity will persist for other
1490: functional forms of $U_0(q)$ as long as it decays for large $q$.
1491: (We note that in the purely Coulomb case, $\lambda=0$, an
1492: additional divergence at small $q$ is introduced. Calculation
1493: shows that all the divergences cancel exactly in this special
1494: case with the second order result for the TED being identically
1495: zero for all $\omega>0$. The physical meaning of this observation
1496: escapes us.)
1497:
1498: The net result is that we find a logarithmic singularity in the TED
1499: for 2D electron systems at the second order in the interaction.
1500: Such behaviour is in apparent contradiction with Ref.~\cite{gadzuk}
1501: where a stronger singularity was found.
1502: These differences may be due to the fact that
1503: Gadzuk and Plummer analysed a different set-up with a three-dimensional host
1504: and a 2D emitting surface (as opposed to our set-up with a 2D host and tip
1505: emission), or their stronger singularity may be an artifact of the
1506: low-density approximation.
1507: This issue as well as the role of higher-order scattering
1508: processes deserve
1509: more intensive study and will be discussed elsewhere.
1510: In view of applications to MWNTs, however, there are two
1511: more immediate questions: the role of the dynamical Coulomb screening
1512: and of the disorder potential.
1513: Both require a non-perturbative approach and will be discussed
1514: in the next Section.
1515:
1516: \section{Tunnelling from disordered 2D systems} \label{disorder}
1517: Contrary to the SWNTs, MWNTs can be described by
1518: the LL theory only in some special cases when the
1519: radii of outer shells are not too large \cite{eggeronly}.
1520: Otherwise their physics is best described in
1521: terms of a disordered 2D electron liquid with Coulomb interaction.
1522: Such systems have been extensively studied in the 80's
1523: using non-linear sigma model and renormalisation group
1524: \cite{finkel,castellani,belitz}.
1525: Recently they attracted more attention due to the discovery
1526: of a possible metal-insulator transition in MOSFETs
1527: \cite{kravchenko}, which stimulated further theoretical research.
1528: Kamenev and Andreev (KA) \cite{kamenev} have recently
1529: examined the Keldysh non-linear sigma model \cite{horbach}
1530: for the Coulomb case (long-range interaction).
1531: In this Section we apply their method to obtain the threshold
1532: behaviour of the TED for tunnelling from
1533: a 2D interacting disordered metal.
1534:
1535: The Hamiltonian of the previous Section, composed from (\ref{H02D})
1536: and (\ref{Hi2D}) with $U_0(q)=4\pi e^2/q$, should now be supplemented
1537: by the disorder term
1538: \begin{equation} \label{Hd2D}
1539: H_{d}[\psi]=\int d\vec{r} V(\vec{r})\psi^{\dag}(\vec{r})\psi(\vec{r})
1540: \end{equation}
1541: with a $\delta$-correlated Gaussian disorder potential
1542: $V(\vec{r})$.
1543:
1544: The formal procedure of deriving the Keldysh sigma model is
1545: described in the relevant literature (\cite{horbach,kamenev}, see
1546: also \cite{ludwig}) in detail. Therefore we only give a brief
1547: outline here. One starts with the Keldysh $S$-matrix. (This
1548: object is identically unity but it becomes a generating
1549: functional upon introducing auxiliary fields. If one wants to
1550: start with a diagonal bare Keldysh function, the contour is
1551: different from that in Fig.~\ref{contour}, see explanation in
1552: Refs. \cite{horbach,ludwig}). One can then integrate the disorder
1553: out obtaining a non-local in time four-fermion interaction term.
1554: The action is made quadratic in fermions by applying the
1555: Hubbard-Stratonovich transformation to the latter term as well as
1556: to the Coulomb interaction term. Hence there are two decoupling
1557: fields, $Q$ and $\Phi$ in the notation of Ref.~\cite{kamenev}. The
1558: fermions can now be formally integrated out resulting in the
1559: following expression for the generating functional \cite{kamenev}:
1560: \begin{eqnarray}\label{genfunc}
1561: \langle Z \rangle = \int {\cal D} \Phi e^{i \mbox{Tr}
1562: (\Phi^T U_0^{-1} \sigma_1 \Phi)} \int {\cal D} Q e^{i S[Q,\Phi]} \, .
1563: \end{eqnarray}
1564: with the sigma-model action (after Keldysh rotation)
1565: \begin{eqnarray}
1566: i S[Q,\Phi] &=& - \frac{\pi \nu}{4 \tau} \mbox{Tr} \, Q^2 \nonumber \\
1567: &+& \mbox{Tr} \, \ln [G_0^{-1} + \frac{i}{2 \tau_0} Q +
1568: \phi_\alpha \gamma^\alpha + \zeta ] \, , \nonumber
1569: \end{eqnarray}
1570: Here $\tau_0$ is the elastic mean free time and
1571: $U_0$ stands for the unscreened interaction potential.
1572: The decoupling field $Q$ depends on two time variables and
1573: is a 2$\times$2 matrix in the Keldysh space.
1574: The decoupling field $\Phi$ is a Keldysh doublet $(\phi_1, \phi_2)^T$.
1575: The vertex matrices are defined as follows:
1576: \begin{eqnarray} \nonumber
1577: \gamma^1 = \left( \begin{array}{rr}
1578: 1 & 0 \\
1579: 0 & 1
1580: \end{array}
1581: \right)
1582: \; , \;
1583: \gamma^2 = \sigma_1 = \left( \begin{array}{rr}
1584: 0 & 1 \\
1585: 1 & 0
1586: \end{array}
1587: \right) \, .
1588: \end{eqnarray}
1589: Doing the functional variation of the generating functional,
1590: (\ref{genfunc}), with respect to the auxiliary field $\zeta$
1591: and setting the latter to zero one obtains
1592: the complete set of single-particle Green's functions:
1593: \begin{eqnarray} \nonumber
1594: \left. \frac{\delta}{\delta \zeta}
1595: \langle Z \rangle \right|_{\zeta=0}
1596: = \left( \begin{array}{cc}
1597: G^R & G^K \\
1598: 0 & G^A
1599: \end{array} \right) = {\cal G} \, .
1600: \end{eqnarray}
1601: where the function $G^K = (G^R-G^A)(1-2 n(E))$ is
1602: related to the single-particle
1603: energy distribution function $n(E)$ and
1604: $G^{R(A)}$ are the retarded and advanced
1605: Green's functions, respectively.
1606:
1607: We need to calculate of the four-point correlation function
1608: ${\cal K}_{+-}^{-+}(t_2,t_3;t_1,t_4)$.
1609: (A calculation of a different four-point function for a
1610: non-equilibrium noise problem has recently been done within a similar
1611: framework in \cite{grabert}.)
1612: This can be achieved by double variation of the generating
1613: functional with respect to the auxiliary field $\zeta$:
1614: \begin{eqnarray} \label{doubleW}
1615: {\cal K}_{+-}^{-+}(t_2,t_3;t_1,t_4) = \int {\cal D}
1616: \Phi e^{i \mbox{Tr} (\Phi^T V_0^{-1} \sigma_1 \Phi)}
1617: \int {\cal D} Q e^{i S[Q,\Phi]} \nonumber \\
1618: \times \left[ W(t_1,t_4)\otimes W(t_3,t_2) + W(t_1,t_2)\otimes
1619: W(t_3,t_4) \right]_* \,
1620: \end{eqnarray}
1621: where
1622: \begin{eqnarray} \nonumber
1623: W(t_1,t_2) = \left[ G_0^{-1} + \frac{i}{2 \tau_0} Q +
1624: \phi_\alpha \gamma^\alpha \right]^{-1} \, .
1625: \end{eqnarray}
1626: A functional integral of this type with a single $W$ function
1627: yields the Green's function matrix ${\cal G}$ and was discussed
1628: in \cite{kamenev}. A product of two and more $W$ functions
1629: produces a complicated object, which is a tensor product
1630: containing all possible time orderings. The operation of
1631: extracting the particular component corresponding to ${\cal
1632: K}_{+-}^{-+}$ we denote by $[...]_*$. At this stage its exact
1633: definition is unimportant and we postpone it to the
1634: end of the Section.
1635:
1636: We now evaluate the functional integral with the product
1637: of two $W$-functions in the saddle point approximation
1638: following KA's approach \cite{kamenev}.
1639: In this case
1640: \begin{eqnarray} \label{Wsaddlpoint}
1641: W(t,t') = - i \pi \nu \, e^{i k_\alpha(t) \gamma^\alpha} \,
1642: \Lambda(t-t') \, e^{i k_\alpha(t') \gamma^\alpha} \, ,
1643: \end{eqnarray}
1644: where $\Lambda(t)$ denotes the mean-field non-interacting
1645: Green's function matrix in equilibrium (corresponding to
1646: non-crossing disorder diagrams \cite{agd})
1647: and $k_\alpha(t)$ is a linear functional of $\Phi$.
1648: The corresponding correlation matrix with respect
1649: to averaging over the $\phi$-fields is given by \cite{kamenev}:
1650: \begin{eqnarray} \label{corrmat}
1651: \langle k_\alpha(q,\omega)k_\beta(-q,-\omega) \rangle_\Phi =
1652: \frac{i}{2} {\cal V}_{\alpha \beta} (q,\omega) \, ,
1653: \end{eqnarray}
1654: \begin{eqnarray} \label{mainVs}
1655: {\cal V}_{\alpha \beta} (q,\omega) &=&
1656: \left( \begin{array}{cc}
1657: {\cal V}^R(q,\omega) & {\cal V}^R(q, \omega) \\
1658: {\cal V}^A(q,\omega) & 0
1659: \end{array} \right) \nonumber \\
1660: {\cal V}^{R(A)}(q,\omega) &=& -
1661: \frac{1}{(d q^2\mp i\omega)^2}\left[
1662: \frac{1}{V_0} + \frac{\nu d
1663: q^2}{d q^2 \mp i \omega}\right]^{-1} \\
1664: {\cal V}^K &=& n_B(\omega) ({\cal V}^R(q,\omega) -
1665: {\cal V}^A(q,\omega)) \, , \nonumber
1666: \end{eqnarray}
1667: where $n_B(\omega)$ is the Bose distribution function
1668: and $d$ is the electron diffusion constant.
1669: Using the saddle point representation in Eq.~(\ref{Wsaddlpoint})
1670: the product of two $W$-functions can be re-written as
1671: \begin{eqnarray} \label{4pis}
1672: &~& \langle \left[ W(t_1,t_4) \otimes \nonumber W(t_3,t_2)
1673: \right]_* \rangle_\Phi \\ \nonumber &=& \frac{1}{4} \sum_{\mu_i}
1674: \left[ (\gamma^{\mu_1} \Lambda(t_1,t_4) \gamma^{\mu_4}) \otimes
1675: (\gamma^{\mu_3}
1676: \Lambda(t_3,t_2) \gamma^{\mu_2}) \right]_* \nonumber \\
1677: &\times& \langle p_{\mu_1}(t_1) (p_{\mu_4}(t_4))^* p_{\mu_3}(t_3)
1678: (p_{\mu_4}(t_2))^* \rangle_\Phi \, .
1679: \end{eqnarray}
1680: The operation of taking the correct time ordering is now
1681: separated from of the averaging over the field $\Phi$, which
1682: only affects the objects
1683: \begin{eqnarray} \nonumber
1684: &~& p_\mu(t) = e^{i(k_1(t)+k_2(t))} + \mu
1685: e^{i(k_1(t)-k_2(t))} \\ \nonumber &=& s(t) + \mu \bar{s}(t) =
1686: e^{i \alpha (k_1(t)+ \beta k_2(t))} + \mu e^{i \alpha (k_1(t)-
1687: \beta k_2(t))} \, .
1688: \end{eqnarray}
1689: The product of four $p_\mu$-operators is a sum of
1690: 16 products of different $s(t_i)$-operators.
1691: The latter objects can be evaluated using
1692: the correlation matrix, Eq.~(\ref{corrmat}),
1693: \begin{eqnarray} \label{manyVs}
1694: && \langle e^{i \alpha_1 (k_1(t_1)+ \beta_1 k_2(t_1))}
1695: e^{-i \alpha_4 (k_1(t_4)+ \beta_4 k_2(t_4))} \nonumber \\
1696: &\times& e^{i \alpha_3 (k_1(t_3)+ \beta_3 k_2(t_3))}
1697: e^{-i \alpha_2 (k_1(t_2)+ \beta_2 k_2(t_2))} \rangle_\Phi \nonumber \\
1698: &=& \exp \Big\{ -\frac{i}{2} \Big[ \alpha_1 \alpha_3 {\cal V}_{13}
1699: (\beta_1,\beta_3)-\alpha_1 \alpha_4 {\cal V}_{14}
1700: (\beta_1,\beta_4) \nonumber \\
1701: &-& \alpha_4 \alpha_3 {\cal V}_{43}(\beta_4,\beta_3)
1702: + \alpha_4 \alpha_2 {\cal V}_{42}(\beta_4,\beta_2) \\
1703: &-& \alpha_3 \alpha_2 {\cal V}_{32}(\beta_3,\beta_2) -
1704: \alpha_1 \alpha_2 {\cal V}_{12}(\beta_1,\beta_2) +
1705: 2 {\cal V}_0(\beta_1,\beta_1) \Big] \Big\} \nonumber \, ,
1706: \end{eqnarray}
1707: where we use the following definition
1708: \begin{eqnarray} \nonumber
1709: {\cal V}_{ij}(\beta_i,\beta_j) = {\cal V}^K(t_i-t_j)
1710: + \beta_i {\cal V}^A(t_i-t_j) \nonumber \\
1711: + \beta_j {\cal V}^R(t_i-t_j) \, .
1712: \end{eqnarray}
1713: Taken separately these functions are divergent. Nevertheless
1714: Eq.~(\ref{manyVs}), being rewritten as a function of differences
1715: ${\cal V}_{ij}(\beta_i,\beta_j)-{\cal V}_{ii}(\beta_i,\beta_i)$,
1716: can be made convergent. In this case a special selection rule has
1717: to be fulfilled,
1718: \begin{eqnarray} \label{selrule}
1719: && \alpha_1 \alpha_3 {\cal V}_0(\beta_1,\beta_3) -
1720: \alpha_1 \alpha_4 {\cal V}_0(\beta_1,\beta_4) \nonumber \\
1721: &-& \alpha_4 \alpha_3 {\cal V}_0(\beta_4,\beta_3) +
1722: \alpha_4 \alpha_2 {\cal V}_0(\beta_4,\beta_2) \nonumber \\
1723: &-& \alpha_3 \alpha_2 {\cal V}_0(\beta_3,\beta_2) -
1724: \alpha_1 \alpha_2 {\cal V}_0(\beta_1,\beta_2) \nonumber \\
1725: &+& \frac{1}{2} \sum_{j=1,..,4} {\cal V}_0(\beta_j,\beta_j) = 0
1726: \, ,
1727: \end{eqnarray}
1728: otherwise (\ref{manyVs}) is divergent. Using Eqs.~(\ref{mainVs})
1729: one can calculate the asymptotic behaviour of ${\cal
1730: V}_{ij}(\beta_i,\beta_j)-{\cal V}_{ii}(\beta_i,\beta_i)$ for
1731: large time differences (to calculate the limiting Fermi edge
1732: asymptotics of the TED this knowledge is sufficient as we shall
1733: see shortly). Fortunately, the limiting form turns out to be
1734: independent of the arrangement of the $\beta$-indices, with the
1735: result
1736: \begin{eqnarray} \nonumber
1737: { \cal V}_{ij}(\beta_i,\beta_j)-{\cal V}_{ii}(\beta_i,\beta_i)
1738: \approx i \frac{e^2}{4 \pi f} \ln^2 \left| \frac{f^2 (t_i-t_j)}{d} \right| \, ,
1739: \end{eqnarray}
1740: where $f=2 \pi e^2 \nu d$.
1741: The last average in Eq.~(\ref{4pis}) is then given by
1742: \begin{eqnarray}
1743: \langle p_{\mu_1}(t_1) (p_{\mu_4}(t_4))^*
1744: p_{\mu_3}(t_3) (p_{\mu_4}(t_2))^* \rangle_\Phi \nonumber \\
1745: = \left(\prod_{j}(1+\mu_j)\right)\langle
1746: s(t_1)s^*(t_4)s(t_3)s^*(t_2)\rangle_\Phi \, . \nonumber
1747: \end{eqnarray}
1748: Obviously, the above is nonzero only if for all $i$ $\mu_i=+1$.
1749: That means that we can set all $\alpha_j$'s equal as well.
1750: Such choice automatically fulfils the selection rule, (\ref{selrule}).
1751: Therefore the full four-point correlation function is given by
1752: \begin{widetext}
1753: \begin{eqnarray} \label{fulldisK}
1754: {\cal K}_{+-}^{-+}(t_1,t_2; t_3, t_4) \approx Y(t_1,t_2; t_3, t_4)
1755: \exp \Big( - \frac{e^2}{8 \pi f} \Big\{ \ln^2 \left|
1756: \frac{f^4(t_1-t_4)(t_4-t_3)(t_3-t_2)(t_1-t_2)}{d^2
1757: (t_1-t_3)(t_4-t_2)} \right| \Big\} \Big)
1758: \end{eqnarray}
1759: \end{widetext}
1760: with
1761: \begin{eqnarray} \nonumber
1762: Y(t_1,t_2; t_3, t_4) \nonumber = \left[ (\gamma^{1} \Lambda(t_1,t_4)
1763: \gamma^{1})
1764: \otimes (\gamma^{1} \Lambda(t_3,t_2) \gamma^{1}) \right]_* \, .
1765: \end{eqnarray}
1766: The latter expression is basically the non-interacting four-point
1767: correlation function with special time ordering and is equal to
1768: $G^{+-}(t_1-t_4) G^{+-}(t_3-t_2)$. The second term in
1769: Eq.~(\ref{doubleW}) can be calculated in a similar manner
1770: resulting in the same, up to an interchange of time variables,
1771: exponential factor as in Eq.~(\ref{fulldisK}) and an appropriate
1772: pre-factor. The pre-factors can in principle be re-exponentiated
1773: but this would only results in logarithmic corrections which can
1774: be neglected in comparison to the main disorder part. In order to
1775: use formula (\ref{com}), we substitute $t_1=\tau_1$, $t_3=0$,
1776: $t_2=t+\tau_1$ and $t_4=t+\tau_1+\tau_2$. We are interested in
1777: the low-energy regime $\omega/V \ll 1$. Therefore the lower limit
1778: for the time integrations can be set to $1/D$ ($D$ being the
1779: conductance band-width) if it causes divergences and to zero
1780: otherwise. Thus we obtain
1781: \begin{eqnarray} \nonumber
1782: n_>(\omega) &\sim& \int_0^\infty \, d\tau_1 \, \int_0^\infty \,
1783: d\tau_2 \, e^{i \omega (\tau_1 + \tau_2)} \\ \nonumber
1784: &\times& \exp \left\{ - \frac{e^2}{8 \pi f}
1785: \ln^2 \left| \frac{f^4(\tau_1+\tau_2)}{d^2 D} \right| \right\} \, .
1786: \end{eqnarray}
1787: Estimating the $\tau$-integrals for small $\omega$ yields the
1788: final result for the TED in vicinity of the Fermi edge:
1789: \begin{eqnarray} \label{disorderresult}
1790: n_>(\omega) \sim \exp \left[ -\frac{e^2}{8 \pi f}
1791: \ln^2 \left( \frac{d^2 D}{f^4\omega}\right) \right] \, .
1792: \end{eqnarray}
1793: Thus we find that the combined effect of the disorder and the
1794: Coulomb interaction is to suppress the TED towards the Fermi
1795: surface. This result can be understood in terms of the well known
1796: Altshuler-Aronov-Lee anomaly \cite{aal} in the density of states
1797: of the primary electrons. Indeed, one of the important results of
1798: the KA's work \cite{kamenev} was to show that the negative
1799: logarithmic correction to the density of states can be
1800: exponentiated (and becomes a double-log due to the long-range
1801: Coulomb forces). Our formula (\ref{disorderresult}) achieves a
1802: similar exponentiation for the TED. As in \cite{kamenev}, this is
1803: only valid above certain energy scale $\epsilon^*$, due to the
1804: on-set of the fluctuation effects (corrections to the saddle-point
1805: approximation). In MWNTs the situation is further complicated by
1806: the dimensional cross-over effect; the corresponding energy scale
1807: $\epsilon^*\simeq d/L^2$ ($L$ being the tube's circumference) was
1808: estimated in \cite{eg2001}, see also \cite{kopietz2001} for
1809: further discussion.
1810:
1811: \section{Field emission from carbon nanotubes} \label{FE}
1812: Field emission setup is relatively simple. An electric field of
1813: strength $F$ is applied between a metallic tip of the emitter and
1814: a counter-electrode which are usually placed in a vacuum.
1815: Superposition of the confinement potential, the image potential
1816: and the electrostatic potential due to the external electric
1817: field leads to a formation of a barrier with a finite width
1818: allowing the electrons to tunnel out of the emitter even at zero
1819: temperatures, Fig.~\ref{FEsetup}.
1820: \begin{figure}
1821: \includegraphics[scale=0.25]{FEsetup}
1822: \caption[]{\label{FEsetup} Schematic representation of the field
1823: emission setup. Here $F$ is the applied electric field, $W$ is
1824: the work function and $D$ is the conduction band-width. The
1825: dashed line represents the confining potential without field and
1826: the dotted line stands for the image potential.}
1827: \end{figure}
1828: Therefore the FE process can be regarded as tunnelling into
1829: vacuum through a triangularly-shaped barrier whose upper part is
1830: rounded off by the image potential. However, at low temperatures
1831: the larger part of tunnelling electrons will have energies close
1832: to the Fermi energy, so the exact form of the barrier is
1833: unimportant.
1834:
1835: Quantities of interest are the energy-resolved current
1836: $j(\omega)$ and the total current $J$.
1837: If the tunnelling amplitude is small (which it is for
1838: all real systems), then $j(\omega)$ is proportional to the
1839: probability $n(\omega)$ for an electron to have energy $\omega$,
1840: to the energy dependent transmission coefficient $\cal D (\omega)$,
1841: and to a factor $\cal F$ responsible for the tip geometry:
1842: \begin{eqnarray} \label{fnappr}
1843: j(\omega) = {\cal F} {\cal D}(\omega) n(\omega) \, .
1844: \end{eqnarray}
1845: In the noninteracting case and neglecting higher-order tunnelling
1846: processes, $n(\omega)$ coincides with the Fermi distribution
1847: function multiplied by the LDOS. The (quasi-classical) transmission
1848: probability for a triangular barrier is given by \cite{landau}:
1849: \begin{eqnarray} \label{tansmprob}
1850: {\cal D}(\omega)\sim \exp( -4\sqrt{2m}(W-\omega)^{3/2}/3\hbar F ) \, ,
1851: \end{eqnarray}
1852: where $m$ is the electron mass and $W$ is the work function. As
1853: discussed e.~g. in Ref.~\cite{plummer}, the transmission
1854: coefficient is only slightly affected by the image potential and
1855: for the field strength $F \ll (W-E_F)^2/e^3$ the barrier can be
1856: regarded as strictly triangular. At zero temperatures, the
1857: emerging (primary) spectrum has a sharp threshold at the Fermi
1858: energy (no particles above $E_F$) and is essentially constant
1859: in its vicinity.
1860:
1861: The general picture does not change much even in the case of
1862: interacting electrons. For the LLs $n(\omega)$ is proportional to
1863: the relevant LDOS, which is known \cite{book,glazman} to be
1864: \begin{eqnarray}\label{LLraspr}
1865: n(\omega) = \Theta(-\omega) |\omega|^{1/g-1}/a_0 D^{1/g}
1866: \Gamma(1/g) \, .
1867: \end{eqnarray}
1868: Plugging this into Eq.~(\ref{fnappr}) one observes that
1869: at this lowest order in the tunnelling the TED above the Fermi energy
1870: is still zero, even for the strongly correlated LL.
1871: Below the Fermi energy the TED has a power-law singularity.
1872: Expanding the transmission coefficient in powers of $\omega/W$ (work function
1873: being the largest energy scale),
1874: and integrating over all energies we establish the
1875: Fowler-Nordheim (FN) formula for LLs:
1876: \begin{eqnarray}\label{FN}
1877: J = \frac{\cal F}{a_0 D^{1/g}}
1878: \left[ \frac{F^2}{4k_F W} \right]^{1/2g}
1879: \exp\left( - \frac{4 k_F^{1/2} }{3F}W^{3/2} \right) \, ,
1880: \end{eqnarray}
1881: relating the full current to the electric field's strength
1882: \cite{fowler,plummer}.
1883:
1884: The generalisation of these results for SWNTs is rather
1885: straightforward. Indeed, these systems are known to be described
1886: as four-channel LLs. Three channels $\phi_{c-}$ (charge-flavour),
1887: $\phi_{s+}$ (total spin), $\phi_{s-}$ (spin-flavour) are
1888: non-interacting. The fourth channel $\phi_{c+}$ (total charge, or
1889: the plasmon mode) possesses the LL parameter
1890: $K=(1+4U_0/\pi)^{-1/2}$, where $U_0$ is the zero Fourier
1891: component of the screened Coulomb potential. Note that though we
1892: now have four channels the field-operator actually factorises as
1893: \cite{sammlung1,sammlung2,sammlung3}
1894: \begin{equation} \label{factor}
1895: \psi\sim \exp\{i \phi_{c+}/(2\sqrt{K})+i(\phi_{c-}+
1896: \phi_{s+}+\phi_{s-})/2\} \, .
1897: \end{equation}
1898: Just as in the case of the spinless LL, the interaction constant $K$
1899: is included into the rescaled Bose fields and disappears from
1900: the Hamiltonian, which is given by
1901: \begin{eqnarray}
1902: H=\frac{1}{4\pi}
1903: \sum_{\delta=c,s j=\pm} \int \, dx \, (\partial_x \phi_{\delta j})^2 \, .
1904: \end{eqnarray}
1905: All the correlation functions also factorise. Hence the results
1906: for the LDOS \cite{book,glazman} as well as the results of Section
1907: \ref{luttinger}, Eqs.~(\ref{cresult}) and (\ref{exprelation}),
1908: are still valid for SWNTs given that the substitution
1909: \begin{equation}
1910: g^{-1}\rightarrow (K^{-1}+3)/4 \, .
1911: \label{subs}
1912: \end{equation}
1913: is made.
1914:
1915: The typical value for $K$ in SWNTs lies near $0.2$ ($0.15$ to
1916: $0.3$), which fixes the effective $g$ to $\approx 0.5$
1917: \cite{sammlung1,sammlung2,sammlung3,tans1,bockrath1}. As a
1918: result, Eq.~(\ref{FN}) is reasonably well approximated by the
1919: classical FN law. This fact offers one possible explanation as to
1920: why the experimental data of Refs.~\cite{french1,japan} can be
1921: relatively well fitted by the conventional FN curve. In addition,
1922: all the experiments currently available were made on SWNT films
1923: where the nanotubes build a network with essentially a random 2D
1924: geometry. This case, when extra complications are bound to arise,
1925: is beyond the scope of this paper. To straightforwardly reveal
1926: the correlation effects in the primary current, measurement on a
1927: {\it single} SWNT are more appropriate. To our knowledge, this
1928: has not been done yet.
1929:
1930: Let us now discuss the secondary current. The analysis of the
1931: secondary effects can also be made in terms of tunnelling into
1932: vacuum. Thereby we should bear in mind that the theory presented
1933: in Section \ref{general} assumes that the tunnelling amplitude is
1934: energy independent. Therefore, since the effective applied
1935: voltage is large in the FE setup (one has to send the chemical
1936: potential on the right electrode to minus infinity), some
1937: integrals inevitably diverge. To overcome this difficulty and to
1938: made the theory more quantitative, we now recall that there are
1939: two candidates that can be used as the effective voltage. The
1940: first one is the characteristic energy scale in the exponential
1941: transmission coefficient ${\cal A} = F/2(k_F W)^{1/2}$, see
1942: Eq.~(\ref{tansmprob}). The second one is the width of the emitter
1943: conductance band $D$. Therefore the high-energy cutoff
1944: playing the role of $V$ in the FE setup is either $D$ or ${\cal
1945: A}$, whichever is smaller. We shall use $D$ through the rest of
1946: the paper. With this modification all results of the previous
1947: Sections for the TED $n(\omega)$ are valid with the understanding
1948: that the tunnelling probability is proportional to the
1949: transmission coefficient of the barrier at the Fermi energy,
1950: $\gamma^2 \sim {\cal D}(E_F)$. According to Eqs.~(\ref{tokexp})
1951: and (\ref{tokexp1}) the secondary current is proportional to the
1952: TED. For SWNT the TED possesses all the singularities
1953: characteristic for LLs, which we have already discussed. One
1954: thing worth noting is that the critical value of the coupling now
1955: is $K_c=1/5$, which is actually within the experimental range.
1956: Therefore we do not make a specific prediction regarding the
1957: character of the singularity (divergent versus convergent TED at
1958: the Fermi edge). Instead we think that both the divergent and the
1959: vanishing TEDs can be observed depending on the experimental
1960: setup. Unfortunately, in most recent experiments, both
1961: on SWNTs and MWNTs, only the total current is measured. Where the
1962: energy-resolved measurements were actually made
1963: \cite{french1,french2,japan}, in all cases but one, the
1964: temperatures were too high for the secondary effects to be
1965: visible. To our knowledge, so far only the experiments of
1966: Ref.~\cite{fransen} contain high-energy tails which can be
1967: attributed to secondary emission. However, the quality of the
1968: presented data hinders us from attempting to actually fit the
1969: curves.
1970:
1971: \begin{figure}
1972: \includegraphics[scale=0.3]{scenario.eps}
1973: \caption[]{\label{scenario} A sketch of the energy resolved current in a field
1974: emission process from LLs. An additional Lorentz peak due to
1975: tunnelling through a localised level (if exists) is depicted in the inset.}
1976: \end{figure}
1977:
1978: We conclude this Section by summarising our scenario for the
1979: energy resolved current in the FE from LLs (and SWNTs). Well
1980: below the Fermi energy $j(\omega)$ is governed by the exponential
1981: growth of the transmission coefficient, $j(\omega) \sim
1982: \exp(\omega/{\cal A})$, see Fig.~\ref{scenario}, as is the case
1983: for any metallic emitter. Nearing the Fermi level, at energies
1984: given by $E_{0} \approx (1-1/g) {\cal A}$, the non-linear LL LDOS
1985: effects win over the exponential growth resulting in a maximum in
1986: the current profile $j(\omega)$. Further towards the Fermi energy
1987: it decreases according to the power-law: $\sim |\omega|^{1/g-1}$.
1988: The edge behaviour right above $E_F$ is given by $\sim
1989: \omega^{1/g-2}$. This changes shape depending on whether $g$ is
1990: smaller or larger than the critical coupling $g_c$. In the case
1991: of weak interactions, $1/2<g<1$, there is a singularity, while
1992: for strong correlations, $0<g<1/2$ this singular behaviour is
1993: suppressed and there is a power-law approach to zero. In the
1994: latter case $j(\omega)$ acquires an additional maximum at a
1995: cross-over energy $E^*$. For small applied fields, there is an
1996: upper threshold at energy $\mbox{min}(D,{\cal A})$ where the
1997: energy resolved current behaves as $\sim (\mbox{min}(D,{\cal
1998: A})-\omega)^{1/g}$.
1999:
2000: \section{Tunnelling via localised states} \label{LS}
2001: Recent luminescence spectra measurements in FE experiments on
2002: carbon nanotubes \cite{french1,fransen} suggest that, at least in some
2003: cases, the emission process cannot be understood simply
2004: in terms of tunnelling into vacuum.
2005: They rather seem to be compatible with a model where
2006: the emitted electron tunnels out of the system through one or more localised
2007: states at the tip of the emitter \cite{french1,fransen}.
2008: In this Section, we generalise our formalism in order
2009: to analyse such a multiple stage tunnelling.
2010:
2011: We start with the Hamiltonian of the system.
2012: Contrary to Eq.~(\ref{ham0}), the transfer of an electron between
2013: the host and the lead occurs in two stages.
2014: At the first stage, the electron populates a localised level with
2015: energy $-\Delta$, see Fig.~\ref{locur}.
2016: \begin{figure}
2017: \includegraphics[scale=0.25]{locur}
2018: \caption[]{\label{locur} A schematic representation of the
2019: electron tunnelling through a localised state $d$.}
2020: \end{figure}
2021: Let the tunnelling amplitude for this process be $\gamma_0$.
2022: At the second stage the electron tunnels from the
2023: localised state into the lead,
2024: \begin{eqnarray}
2025: H_2 &=& H[\psi] + H_0[c] + \gamma_0 \left[ \psi(0)^\dag d + d^\dag \psi(0)
2026: \right] \nonumber \\
2027: &+& \gamma \left[ d^\dag c(0) + c^\dag(0) d \right] - \Delta d^\dag d \; ,
2028: \end{eqnarray}
2029: where we retain the notation $\gamma$ for the tunnelling amplitude
2030: of the second process and $d^\dag,d$ are the creation and
2031: annihilation operators of the localised state, respectively. We
2032: assume that $0 < -\Delta < -V$. This is not restrictive but
2033: reasonable. Indeed, the first inequality reflects the fact that
2034: the electric field applied to the emitter causes an effective
2035: lowering of energy of the states close to the tip. The second
2036: inequality is a trivial one because if the energy of the state
2037: $d$ lied below both Fermi levels it would have always been
2038: populated. For simplicity, we
2039: neglect any electrostatic interaction between the localised state
2040: and the leads and the Hubbard term for $d$. Like in the case of
2041: the simple tunnelling, the lead is assumed to be uncorrelated.
2042:
2043: The problem of a localised state hybridised with a
2044: non-interacting continuum is, of course, exactly solvable
2045: \cite{mahan}. Therefore one way to proceed would be to eliminate
2046: the $\gamma$ hopping term from the Hamiltonian exactly and then
2047: apply perturbation theory in $\gamma_0$. However, in the FE setup
2048: $\Delta$ is expected to be much smaller than the conductance band
2049: width $D$. This implies that $\gamma_0 \gg \gamma$. Therefore we
2050: take an alternative route and first calculate the TED of the
2051: electrons on the level $d$ as a function of $g^{-+}(\omega)$ in
2052: the host ignoring tunnelling to the lead. (We did the
2053: alternative calculation as well and, at a given order of
2054: perturbation theory, obtained exactly the same results.) To
2055: proceed we define the Keldysh Green's function of the localised
2056: level:
2057: \begin{eqnarray} \label{Ddef}
2058: D(t) = -i \langle T_C [d(t) d^\dag(0) S_C ]\rangle \, ,
2059: \end{eqnarray}
2060: where the $S$-matrix includes only the tunnelling
2061: between the host and the localised level,
2062: \begin{eqnarray} \label{newSMat}
2063: S_C = T_C \exp\left( -i \gamma_0 \int_C \, dt \,
2064: \left[ \psi^\dag(0,t) d(t) + d^\dag(t) \psi(0,t)
2065: \right] \right) \; . \nonumber
2066: \end{eqnarray}
2067: Expanding the Green's function (\ref{Ddef}) and function $\langle
2068: T_C [d(t) \psi^\dag(0,t') S_C ]\rangle$ in powers of $\gamma_0$
2069: one obtains a set of coupled equations for them. Eliminating the
2070: latter object one then obtains the following Dyson-type equation
2071: for the Keldysh Green's function:
2072: \begin{eqnarray} \label{dysonequation}
2073: D(t,t') = D_0(t,t') + \int_C \, d t'' \, D(t,t'') \kappa(t'',t') \; ,
2074: \end{eqnarray}
2075: where the kernel is given by
2076: \begin{equation}
2077: \kappa(t'',t') = \gamma_0^2 \int_C \, d t \, g_0(t'',t) D_0(t,t') \, .
2078: \end{equation}
2079: Here $g_0(t,t')$ and $D_0(t,t')$ denote the Green's functions of
2080: the host and the localised level in the absence the hopping term.
2081: Disentangling the Keldysh indices and Fourier transforming we
2082: obtain the following set of equations:
2083: \begin{eqnarray}
2084: D^{-+}(\omega) = D^{-+}_0(\omega) + D^{--}(\omega) \kappa^{-+}(\omega) -
2085: D^{-+}(\omega) \kappa^{++}(\omega) \; , \nonumber \\
2086: D^{--}(\omega) = D^{--}_0(\omega) + D^{--}(\omega) \kappa^{--}(\omega) -
2087: D^{-+}(\omega) \kappa^{+-}(\omega) \; . \nonumber
2088: \end{eqnarray}
2089: The solution is given by
2090: \begin{widetext}
2091: \begin{equation} \nonumber
2092: D^{-+}(\omega) = \frac{D^{-+}_0 + \gamma_0^2 g^{-+}_0(\omega)
2093: |D^{--}_0(\omega)|^2}{1+2 \gamma_0^2 \mbox{Re} [ g^{--}_0(\omega)
2094: D^{--}_0(\omega)] + \gamma_0^4 |g^{--}_0(\omega) D^{--}_0(\omega)|^2} \, .
2095: \end{equation}
2096: \end{widetext}
2097: This equation can be further simplified if one takes into account
2098: that $\mbox{Re}[g^{--}_0(\omega)]/|D^{--}_0(\omega)|^2$ vanishes for
2099: our system. Then the TED on the localised level is proportional to
2100: \begin{equation} \label{correctD-+}
2101: D^{-+}(\omega) = \frac{\gamma_0^2 g^{-+}(\omega)}{(\omega+ \Delta)^2 +
2102: \gamma_0^4 |g^{--}(\omega)|^2} \, .
2103: \end{equation}
2104: In the last relation we omitted the subscript $0$ of $g(\omega)$
2105: functions. Strictly speaking this constitutes an approximation.
2106: The reason is that since the tunnelling onto the localised level
2107: affects the Green's function of the host they have to be
2108: calculated self-consistently. However, the renormalisations
2109: occurring close to the Fermi energy should not strongly affect
2110: the shape of the localised state and vice versa. That is unless
2111: the localised level becomes resonant ($\Delta\to 0$), which case
2112: we shall not consider in this paper.
2113: (We note though that the resonant level case can be approached
2114: via a mapping onto the Kondo problem \cite{resonantlevel}.)
2115: For a non-interacting host
2116: $|g^{--}(\omega)|^2$ is simply a constant equal to $1/4v_F^2$.
2117:
2118: As we know from Section \ref{general}, the TED in the
2119: non-interacting lead, is proportional to $D^{-+}(\omega)$ and the
2120: tunnelling probability $\gamma^2$:
2121: \begin{equation} \label{nf}
2122: n(\omega) = -i \gamma^2 D^{-+}(\omega) \, .
2123: \end{equation}
2124:
2125: We now briefly discuss the picture for a non-interacting emitter.
2126: According to Eq.~(\ref{correctD-+}) there is a Lorentzian peak at
2127: $\omega=-\Delta$ with height $4 v_F^2
2128: g^{-+}(-\Delta)/\gamma_0^2$ and width $\gamma_0^2/2 v_F$. At the
2129: second order in $\gamma^2$, there still is a sharp threshold at
2130: the Fermi energy. These results for non-interacting emitters are,
2131: of course, known (see e.~g.~\cite{plummer}). Here we have merely
2132: re-derived them using the Keldysh formalism.
2133:
2134: Let us now turn to the open question of how the presence of a
2135: localised state would modify the FE process from a LL. For the
2136: primary current we can still use formula (\ref{fnappr}), where we
2137: have to substitute the TED of the electrons in the host by that
2138: of the electrons on the localised state. The energy resolved
2139: current is then given by the following expression (we again set
2140: $v_F=1$),
2141: \begin{eqnarray} \label{LSFNbare}
2142: {\cal J}(\omega) &=& -i {\cal F} \, {\cal D}(\omega) \,
2143: D^{-+} (\omega) \nonumber \\
2144: &=& \frac{{\cal F} \, \gamma^2 \gamma_0^2 \, \Theta(-\omega)}{a_0
2145: \, D^{1/g}\, \Gamma[1/g]} \,
2146: \frac{|\omega|^{1/g-1}}{(\omega+\Delta)^2+\gamma_0^4/4} \nonumber \\
2147: &\times& \exp\left( - 4 k_F^{1/2} \frac{(W-\omega)^{3/2}}{3F} \right) \, .
2148: \end{eqnarray}
2149: In comparison to the situation without the localised level, there
2150: is an additional feature: the Lorentzian resonance
2151: at $\omega= - \Delta$. To compute the total emitted current we
2152: re-write the energy integral in terms of the dimensionless
2153: variable $\xi = \omega/{\cal A}$ and expand the exponent in powers
2154: of $\omega/W$ as the work function is large compared to other
2155: energy scales. The total current then is
2156: \begin{eqnarray} \label{LSFNintegral}
2157: {\cal J} &=& \frac{\gamma^2 \gamma_0^2 \, {\cal F}}{a_0 \, D^{1/g}
2158: \, \Gamma[1/g]} \, \exp \left( - 4 k_F^{1/2}\, \frac{W^{3/2}}{3F}
2159: \right) {\cal A}^{1/g-2} \nonumber \\&\times& \int_0^{\infty} \,
2160: d\xi \, \frac{\xi^{1/g-1}}{(\xi- \Delta/{\cal A})^2 +
2161: (\gamma_0^2/2 {\cal A})^2} \, e^{-\xi} \, .
2162: \end{eqnarray}
2163: This integral can be calculated exactly, resulting
2164: in the modified FN relation
2165: \begin{eqnarray} \nonumber
2166: {\cal J} = J \, \frac{2 \gamma_0^2}{\cal A} \, \mbox{Im} \,
2167: \left\{ w^{1/g-1} e^{w} \Gamma[1-1/g,w] \right\} \,
2168: \end{eqnarray}
2169: applicable for arbitrary parameters. Here $J$ is the total
2170: current in the absence of the localised state, see
2171: Eq.~(\ref{FN}), $\Gamma$ stands for the (incomplete) gamma
2172: function and we have introduced a dimensionless quantity $w$,
2173: defined by
2174: \begin{eqnarray} \nonumber
2175: w = \frac{1}{\cal A}\left( i \gamma_0^2/2 - \Delta \right) \, .
2176: \end{eqnarray}
2177:
2178: There two important limiting cases.
2179:
2180: {\bf (i)} The electric field is weak, $\Delta/{\cal A} \gg 1$, or
2181: $\Delta \gg F/2(k_F W)^{1/2}$. Then the resonance peak is far
2182: away from the Fermi edge in comparison with the characteristic
2183: decay scale of the transmission coefficient. The result of the
2184: integration in Eq.~(\ref{LSFNintegral}) is then essentially
2185: identical to the result for a system without a localised level
2186: Eq.~(\ref{FN}), up to numerical pre-factors. Therefore we would
2187: not expect that in this regime one can differentiate between the
2188: direct tunnelling and tunnelling via a localised state.
2189:
2190: {\bf (ii)} The electric field is sufficiently strong, so that
2191: $\Delta/{\cal A} \ll 1$. Now the distance from the resonance to
2192: the Fermi edge is much smaller than the characteristic energy
2193: scale ${\cal A}$. If also $\gamma^2\ll\Delta$ (which is to be
2194: expected) the total current is still given by Eq.~(\ref{FN}), apart from
2195: the overall pre-factor, but with a different exponent: $1/2g$
2196: should now be substituted by $1/2g-1$. This change can be
2197: important for the analysis of experimental data on the SWNTs. For
2198: SWNTs in this regime, we expect the modified FN plot to be linear in
2199: the coordinates $\ln [{\cal J}/F^{(1/K-5)/4}]$ and $1/F$. For $K
2200: \approx 0.2$ the total current depends exponentially on the field
2201: strength (not a stretched exponential).
2202:
2203: The calculation of the secondary current is more delicate.
2204: In the case of the localised state the
2205: insertion in the diagram for $g^{-+}(\omega)$,
2206: see Figs.~\ref{fig3} and \ref{fig4}, should be modified.
2207: It turns out that there is only one way to dress
2208: the Green's function $G^{+-}_0$ with
2209: the Green's functions of the localised level,
2210: $G^{+-}_0(\omega) \rightarrow D^{++}(\omega) G^{+-}_0(\omega) D^{--}(\omega)$,
2211: so that the diagram is non-zero above the Fermi edge.
2212: The time- and anti-time-ordered localised level Green's
2213: functions take the form
2214: \begin{eqnarray} \nonumber
2215: D^{--(++)}(\omega) = \mp \frac{1}{\omega+\Delta
2216: \pm i(\alpha_d+\gamma_0^2/2)} \, .
2217: \end{eqnarray}
2218: The imaginary part in the denominators reflects the fact that the
2219: natural width of the localised level $\alpha_d$ is now widened by
2220: the hybridisation with the continuum of states in the host. Using
2221: the Fourier transform of the new insertion, which must now be
2222: substituted instead of the Fourier transform of
2223: $G^{+-}_0(\omega)$ (given in Eq.~(\ref{com}) by
2224: $\exp{i(\omega-V)t}/(t+i \alpha)$), we construct the localised
2225: level counterpart of Eq.~(\ref{com}):
2226: \begin{widetext}
2227: \begin{eqnarray} \label{com2}
2228: n_>(\omega) &=& i \frac{\gamma^2}{2 \pi (\alpha_d + \gamma_0^2)}
2229: \int_{-\infty}^\infty dt e^{i (\omega+\Delta)t} (e^{(\alpha_d +
2230: \gamma_0^2)t} \Gamma[t(\alpha_d + \gamma_0^2 + i (\Delta+D))] -
2231: e^{-(\alpha_d + \gamma_0^2)t} \Gamma[t(-\alpha_d - \gamma_0^2 + i
2232: (\Delta+D))]) \nonumber \\ &\times& \int_0^\infty d\tau_1
2233: \int_0^\infty d\tau_2e^{i\omega(\tau_1+\tau_2)} {\cal
2234: K}(\tau_1,0;t+\tau_1,t+\tau_1+\tau_2) \, .
2235: \end{eqnarray}
2236: \end{widetext}
2237: The asymptotic behaviour of
2238: the $n_>(\omega)$ close the Fermi edge is not affected (apart from a
2239: numerical pre-factor) by the presence of the localised state.
2240: There is an important difference -- the upper threshold for the
2241: TED is now given by $\Delta$ instead of $D$. The reason is that in
2242: the presence of the localised state the upper energy limit of the
2243: hot hole is given by the energy of the $d$-level instead of the
2244: band-width $D$ (or the applied voltage $V$).
2245:
2246: Thus, in the case of tunnelling via a localised state the TED
2247: obtains an additional feature, namely a Lorentzian peak below the
2248: Fermi edge, see the inset in Fig.~\ref{scenario}. Recent
2249: luminescence experiments of Ref.~\cite{french1} suggest a similar
2250: structure -- two superimposed peaks in the spectrum of emitted
2251: light. The Lorentzian shape of the peaks was probably hidden by
2252: the intrinsic Gaussian broadening introduced by the apparatus (a
2253: Gaussian fit was adopted in Ref.~\cite{french1}). The calculation
2254: of the emerging luminescence spectra is beyond the scope of this
2255: paper and will be discussed elsewhere.
2256:
2257: \section{Summary and conclusions} \label{conclusions}
2258: In this paper, we studied the energy resolved current (or the
2259: TED) emitted from a tip of a correlated host material under
2260: out-of-equilibrium conditions. In the uncorrelated case the TED
2261: has a sharp threshold at the Fermi edge and there are no
2262: particles with energies above it. In the presence of interactions
2263: there is a finite high-energy tail due to the hot hole relaxation
2264: process. For most systems, the TED is singular right above the
2265: Fermi edge.
2266:
2267: In the repulsive Luttinger liquid model, where the LDOS is
2268: renormalised by interactions and vanishes towards the Fermi edge,
2269: the TED below the Fermi edge shows a power-law behaviour with the
2270: LDOS exponent. Above the Fermi energy, we also find a power-law
2271: but with a new exponent. There is a critical coupling and the TED
2272: is divergent for weak interactions but convergent for strong
2273: interactions. In the latter case, the TED has a maximum above the
2274: Fermi level. Although the high-energy tail still exists in
2275: systems with local interactions confined to the tip, the TED is
2276: then a regular function of the energy. Additional features in the
2277: TEDs arise in a situation when the tunnelling occurs in two
2278: stages via a localised state. In this case there is a Lorentzian
2279: peak centred at the energy of this localised level and of the
2280: width which depends on the hybridisation with the continuum
2281: states on both sides of the contact. All the singularities of the
2282: TED above the Fermi energy survive the introduction of the
2283: localised level. The singularities in high-energy tails are
2284: suppressed as soon as the disorder is introduced. We have
2285: calculated the TED of particles tunnelling from a correlated
2286: disordered 2D system. In this case the LDOS is exponentially
2287: suppressed and this suppression is carried over to the TED.
2288: Contrary, in the pure limit, we find a weak (logarithmic)
2289: singularity in the TED above the Fermi edge.
2290:
2291: We then specialised our results to the case of the FE from
2292: carbon nanotubes, both SWNTs and MWNTs.
2293: While they are in qualitative agreement with existing experiments,
2294: there is little data currently available to test our predictions
2295: for the secondary current in these systems.
2296: We hope that this work may encourage more measurements,
2297: especially at low temperatures ($\ll$ 300K).
2298: This could provide additional information about the nature of
2299: interactions in nanotubes.
2300:
2301: Throughout the paper we consistently worked at zero
2302: temperatures. All our results can be generalised to
2303: finite temperatures by standard methods (though
2304: this may not be of immediate interests until there
2305: is more experimental data).
2306: There are, of course, other open questions (like the role
2307: of higher orders in the tunnelling amplitude) and potentially
2308: interesting extensions of this work, which are detailed
2309: in the main text.
2310:
2311: Also interesting, though more difficult, could be experiments on
2312: tunnelling junctions if the TED in such setups were accessible
2313: e.~g. by means of a tunnel probe \cite{devoret}. In this case both
2314: threshold asymptotics could reveal essential details about the
2315: correlations. Even if direct measurements of the TED are not
2316: possible, there is another possibility to identify the secondary
2317: effects. Since the hot hole decay is a relaxation process, an
2318: electron-hole recombination can occur with an irradiation of a
2319: photon. Such a current-driven luminescence effect could be
2320: observable in quantum dot systems, where the corresponding
2321: experimental techniques became widespread in the recent years
2322: \cite{dekel}.
2323:
2324: \acknowledgments We have benefited from illuminating discussions
2325: with Nathan Andrei, David Edwards, Reinhold Egger, Fabian Essler,
2326: and Yang Chen. This work was supported by the EPSRC of the UK
2327: under grants GR/N19359 and GR/R70309. The authors also participate
2328: in the EC training network DIENOW.
2329:
2330:
2331: \bibliography{prb}
2332:
2333: \end{document}
2334: