配色: 字号:
The quantum interference effect transistor
2023-03-21 | 阅:  转:  |  分享 
  
IOP PUBLISHING NANOTECHNOLOGY

Nanotechnology 18 (2007) 424014 (6pp) doi:10.1088/0957-4484/18/42/424014

The quantum interference effect transistor

Charles A Stafford

1

, David M Cardamone

2

and Sumit Mazumdar

1

1

Department of Physics, University of Arizona, 1118 E. 4th Street, Tucson, AZ 85721, USA

2

Department of Physics, Simon Fraser University, 8888 University Drive, Burnaby, BC,

V5A 1S6, Canada

E-mail: stafford@physics.arizona.edu

Received 3 May 2007, in final form 25 July 2007

Published 13 September 2007

Online at stacks.iop.org/Nano/18/424014

Abstract

We give a detailed discussion of the quantum interference effect transistor

(QuIET), a proposed device which exploits the interference between electron

paths through aromatic molecules to modulate the current flow. In the off

state, perfect destructive interference stemming from the molecular

symmetry blocks the current, while in the on state, the current is allowed to

flow by locally introducing either decoherence or elastic scattering. Details

of a model calculation demonstrating the efficacy of the QuIET are presented,

and various fabrication scenarios are proposed, including the possibility of

using conducting polymers to connect the QuIET with multiple leads.

1. Introduction

Despite their low cost and extreme versatility, modern

semiconductor transistors face fundamental obstacles to

continued miniaturization. First, top-down fabrication gives

them microscopic variability from device to device, which,

while acceptable at today’s length scales, renders them

unscalable in the nanometre regime. Second, these devices,

like all field-effect devices, function by raising and lowering

an energy barrier to charge transport of at least k

B

T ; each

device therefore dissipates energy of this magnitude into the

environment with every switching cycle

3

. At device densities

greater than the current state of the art, the cost and engineering

challenges associated with removing the resultant heat are

daunting [2]. While the first challenge can be met by utilizing

the bottom-up chemical fabrication of single-molecule devices

(e.g. [3]), this approach in itself does nothing to address the

need for a cooler switching mechanism.

An alternative paradigm to raising and lowering an energy

barrier is to exploit the wave nature of the electron to control

the current flow [4–9]. Traditionally, such interference-based

devices are modulated via the Aharanov–Bohm effect [10].

This, however, is incompatible with the small size of molecular

devices [4]: through a 1 nm

2

device, a magnetic field of over

600 T would be required to generate a phase shift of order

1 rad. Similarly, a device based on an electrostatic phase

shift [7] would require voltages incompatible with structural

stability. Previously [11], we have proposed a solution,

3

This is to be distinguished from the heating due to irreversible computation

of [1].

Figure 1. Artist’s conception of a quantum interference effect

transistor based on 1,3-benzenedithiol. The coloured spheres

represent individual carbon (green), hydrogen (purple), sulfur

(yellow), and gold (gold) atoms. In the ‘off’ state of the device,

destructive interference blocks the flow of current between the source

(bottom) and drain (right) electrodes. Decoherence introduced by the

scanning transmission microscope (STM) tip (upper left) suppresses

interference, allowing current flow. Image by Helen M Giesel.

called the quantum interference effect transistor (QuIET) (see

figure 1), which exploits a perfect destructive interference due

to molecular symmetry and controls quantum transport by

introducing decoherence or elastic scattering.

The purpose of this article is to communicate the details of

this proposal, including several potential chemical structures,

to facilitate fabrication and testing of this device. In section 2,

we describe the theoretical framework used to model the

device. Section 3 explains the QuIET’s operating mechanism.

0957-4484/07/424014+06$30.00 1 ? 2007 IOP Publishing Ltd Printed in the UK

Nanotechnology 18 (2007) 424014 C A Stafford et al

Section 4 discusses practical implementations of the device.

We conclude in section 5.

2. Theoretical model

The QuIET consists of a central molecular element, two leads

chemically bonded to the molecule, and a third lead, which can

be coupled to the molecule either capacitively or via tunnelling.

The Hamiltonian of this system can be written as the sum of

three terms:

H = H

mol

+ H

leads

+ H

tun

. (1)

The first is the π-electron molecular Hamiltonian

H

mol

=

summationdisplay



ε

n

d

?



d



?

summationdisplay

nmσ

parenleftbig

t

nm

d

?



d



+ H.c.

parenrightbig

+

summationdisplay

nm

U

nm

2

Q

n

Q

m

, (2)

where d

?



creates an electron of spin σ =↑,↓ in the π-orbital

of the nth carbon atom, and ε

n

are the orbital energies. We use

a tight-binding model for the hopping matrix elements with

t

nm

= 2.2, 2.6 or 2.4 eV for orbitals connected by a single

bond, double bond, or within an aromatic ring, respectively,

and zero otherwise. The final term of equation (2) contains

intrasite and intersite Coulomb interactions, as well as the

electrostatic coupling to the leads. The interaction energies are

given by the Ohno parameterization [15, 16]:

U

nm

=

11.13 eV

radicalBig

1 + .6117

parenleftbig

R

nm

/

?

A

parenrightbig

2

, (3)

where R

nm

is the distance between orbitals n and m.

Q

n

=

summationdisplay

σ

d

?



d



?

summationdisplay

α

C



V

α

/e ? 1(4)

is an effective charge operator [17] for orbital n,wherethe

second term represents a polarization charge. Here C



is the

capacitance between orbital n and lead α, chosen consistent

with the interaction energies of equation (3) and the geometry

of the device, and V

α

is the voltage on lead α. e is the

magnitude of the electron charge.

Each metal lead α possesses a continuum of states, and

their total Hamiltonian is

H

leads

=

3

summationdisplay

α=1

summationdisplay

k∈α

σ

epsilon1

k

c

?



c



, (5)

where the epsilon1

k

are the energies of the single-particle levels in the

leads, and c

?



is an electron creation operator. Here leads 1

and 2 are the source and drain, respectively, and lead 3 is the

control, or gate electrode.

Tunnelling between molecule and leads is provided by the

final term of the Hamiltonian,

H

tun

=

summationdisplay

〈nα〉

summationdisplay

k∈α

σ

parenleftbig

V

nk

d

?



c



+ H.c.

parenrightbig

, (6)

where the V

nk

are the tunnelling matrix elements from a level

k within lead α to the nearby π-orbital n of the molecule.

Coupling of the leads to the molecule via molecular chains,

as may be desirable for fabrication purposes, can be included

in the effective V

nk

, as can the effect of substituents (e.g., thiol

groups) used to bond the leads to the molecule [18, 19].

We use the non-equilibrium Green function (NEGF)

approach [20, 21] to describe transport in this open quantum

system. The retarded Green function of the full system is

G(E) = [E ? H

mol

? Sigma1(E)]

?1

, (7)

where Sigma1 is an operator, known as the retarded self-energy,

describing the coupling of the molecule to the leads. The

QuIET is intended for use at room temperature, and it operates

in a voltage regime where there are no unpaired electrons in

the molecule. Thus lead–lead and lead–molecule correlations,

such as the Kondo effect, do not play an important role.

Electron–electron interactions may therefore be included via

the self-consistent Hartree–Fock method. H

mol

is replaced

by the corresponding mean-field Hartree–Fock Hamiltonian

H

HF

mol

, which is quadratic in electron creation and annihilation

operators, and contains long-range hopping. Within mean-field

theory, the self-energy is a diagonal matrix

Sigma1

nσ,mσ

prime (E) = δ

nm

δ

σσ

prime

summationdisplay

〈aα〉

δ

na

Sigma1

α

(E), (8)

with nonzero entries on the π-orbitals adjacent to each lead α:

Sigma1

α

(E) =

summationdisplay

k∈α

〈nα〉

|V

nk

|

2

E ? epsilon1

k

+ i0

+

. (9)

The imaginary parts of the self-energy matrix elements

determine the Fermi’s Golden Rule tunnelling widths

Gamma1

α

(E) ≡?2ImSigma1

α

(E) = 2π

summationdisplay

k∈α

|V

nk

|

2

δ (E ? epsilon1

k

) . (10)

As a consequence, the molecular density of states changes from

a discrete spectrum of delta functions to a continuous, width-

broadened distribution. We take the broad-band limit [20],

treating the Gamma1

α

as constants characterizing the coupling of the

leads to the molecule. Typical estimates [19] using the method

of [22] yield Gamma1

α

lessorsimilar 0.5 eV, but values as large as 1 eV have

been suggested [18].

The effective hopping and orbital energies in H

HF

mol

depend

on the equal-time correlation functions, which are found in the

NEGF approach to be

〈d

?



d



〉=

summationdisplay

〈aα〉

Gamma1

α



integraldisplay



?∞

dEG

nσ,aσ

(E)G

?

aσ,mσ

(E) f

α

(E),

(11)

where f

α

(E) ={1 + exp[(E ? μ

α

)/k

B

T ]}

?1

is the Fermi

function for lead α. Finally, the Green function is determined

by iterating the self-consistent loop, equations (7)–(11).

The current in lead α is given by the multi-terminal current

formula [23]

I

α

=

2e

h

3

summationdisplay

β=1

integraldisplay



?∞

dET

βα

(E)

bracketleftbig

f

β

(E) ? f

α

(E)

bracketrightbig

, (12)

where

T

βα

(E) = Gamma1

β

Gamma1

α

|G

ba

(E)|

2

(13)

is the transmission probability [21] from lead α to lead β,and

a (b) is the orbital coupled to lead α(β). Similar mean-field

NEGF calculations have been widely used to treat two-terminal

transport through single molecules [13].

2

Nanotechnology 18 (2007) 424014 C A Stafford et al

(a)

(b)

(c)

Figure 2. Effective transmission probability

?

T

12

of the device shown

in figure 1, at room temperature, with Gamma1

1

= 1.2eVandGamma1

2

= .48 eV.

Here ε

F

is the Fermi level of the molecule. (a) Sigma1

3

= 0; (b)

Sigma1

3

=?iGamma1

3

/2, where Gamma1

3

= 0 in the lowest curve, and increases by

.24 eV in each successive one; (c) Sigma1

3

is given by equation (15) with

a single resonance at ε

ν

= ε

F

+ 4eV.Heret

ν

= 0inthelowest

curve, and increases by 0.5 eV in each successive curve. Inset: full

vertical scale for t

ν

= 1 eV. Reprinted with permission from [11].

? 2006 American Chemical Society.

3. Switching mechanism

The QuIET exploits quantum interference stemming from the

symmetry of monocyclic aromatic annulenes such as benzene.

Quantum transport through single benzene molecules with two

metallic leads connected at para positions has been the subject

of extensive experimental and theoretical investigation [13];

however, a QuIET based on benzene requires the source (1)

and drain (2) to be connected at meta positions, as illustrated

in figure 1. The transmission probability T

12

of this device,

for Sigma1

3

= 0, is shown in figure 2. Due to the molecular

symmetry [8], there is a node in T

12

(E), located midway

between the highest occupied molecular orbital (HOMO) and

lowest unoccupied molecular orbital (LUMO) energy levels

(see figure 2(b), lowest curve). This mid-gap node, at the Fermi

level of the molecule, plays an essential role in the operation of

the QuIET.

The existence of a transmission node for the meta

connection can be understood in terms of the Feynman path

integral formulation of quantum mechanics [24], according

to which an electron moving from lead 1 to lead 2 takes all

possible paths within the molecule; observables relate only

to the complex sum over paths. In the absence of a third

lead (Sigma1

3

= 0), these paths all lie within the benzene ring.

An electron entering the molecule at the Fermi level has de

Broglie wavevector k

F

= π/2d,whered = 1.397

?

Ais

the intersite spacing of benzene (note that k

F

is a purely

geometrical quantity, which is unaltered by electron–electron

interactions [25]). The two most direct paths through the ring

have lengths 2d and 4d, with a phase difference k

F

2d = π,so

they interfere destructively. Similarly, all of the paths through

the ring cancel exactly in a pairwise fashion, leading to a node

in the transmission probability at E = ε

F

.

This transmission node can be lifted by introducing

decoherence or elastic scattering that breaks the molecular

symmetry. Figures 2(b) and (c) illustrate the effect of coupling

a third lead to the molecule, introducing a complex self-energy

Sigma1

3

(E) on the π-orbital adjacent to that connected to lead 1

or 2. An imaginary self-energy Sigma1

3

=?iGamma1

3

/2 corresponds to

coupling a third metallic lead directly to the benzene molecule,

as shown in figure 1. If the third lead functions as an

infinite-impedance voltage probe, the effective two-terminal

transmission is [12]

?

T

12

= T

12

+

T

13

T

32

T

13

+ T

32

. (14)

The third lead introduces decoherence [12] and additional

paths that are not cancelled, thus allowing current to flow, as

shown in figure 2(b). This quantum-mechanical effect of the

third lead is a fundamentally different switching mechanism

from other proposed molecular transistors, such as [3], which

rely on electrostatic gating to control the current. As a proof

of principle, a QuIET could be constructed using a scanning

tunnelling microscope (STM) tip as the third lead (cf figure 1),

with tunnelling coupling Gamma1

3

(x) to the appropriate π-orbital of

the benzene ring, the control variable x being the piezo-voltage

controlling the tip–molecule distance.

By contrast, a real self-energy Sigma1

3

introduces elastic

scattering, which can also break the molecular symmetry. This

can be achieved by attaching a second molecule to the benzene

ring, for example an alkene chain. The retarded self-energy

due to the presence of a second molecule is

Sigma1

3

(E) =

summationdisplay

ν

|t

ν

|

2

E ? ε

ν

+ i0

+

, (15)

where ε

ν

is the energy of the νth molecular orbital of the

second molecule, and t

ν

is the hopping integral coupling this

orbital with the neighbouring π-orbital of the benzene ring.

Figure 2(c) shows the transmission probability T

12

(E) in the

vicinity of the Fermi energy of the molecule, for the case of

a single side orbital at ε

ν

= ε

F

+ 4 eV. As the coupling

t

ν

is increased, the node in transmission at E = ε

F

is

lifted due to scattering from the side orbital. The side group

introduces Fano antiresonances [5, 26], which suppress current

through one arm of the annulene, thus lifting the destructive

interference. Put another way, the second molecule’s orbitals

hybridize with those of the annulene, and a state that connects

leads 1 and 2 is created in the gap (see figure 2(c) (inset)). In

practice, either t

ν

or ε

ν

might be varied to control the strength

of Fano scattering.

Tunable current suppression occurs over a broad energy

range, as shown in figure 2(b); the QuIET functions with any

metallic leads whose work function lies within the annulene

gap. Fortunately, this is the case for many bulk metals,

among them palladium, iridium, platinum, and gold [27].

Appropriately doped semiconductor electrodes [14] could also

be used.

We show in figure 3 the I–V characteristic of a QuIET

based on sulfonated vinylbenzene. The three metallic

3

Nanotechnology 18 (2007) 424014 C A Stafford et al

Figure 3. Room-temperature I –V characteristic of a QuIET based on

sulfonated vinylbenzene. The current in lead 1 is shown, where

V

αβ

= V

α

? V

β

. Here, Gamma1

1

= Gamma1

2

= 1eV.Gamma1

3

is taken as 0.0024 eV,

which allows a small current in the third lead, so that the device

amplifies current. A field-effect device with almost identical I –V can

be achieved by taking Gamma1

3

= 0. The curve for I

3

is for the case of

1.00 V bias voltage; I

3

for other biases look similar. Reprinted with

permission from [11]. ? 2006 American Chemical Society.

electrodes were taken as bulk gold, with Gamma1

1

= Gamma1

2

= 1eV,

while Gamma1

3

= 0.0024 eV, so that the coupling of the third

electrode to the alkene side group is primarily electrostatic.

The device characteristic resembles that of a macroscopic

transistor. As the voltage on lead 3 is increased, scattering from

the antibonding orbital of the alkene side group increases as it

approaches the Fermi energies of leads 1 and 2, leading to a

broad peak in the current. For Gamma1

1,2

greatermuch Gamma1

3

negationslash= 0, the device

amplifies the current in the third lead (dotted curve), emulating

a bipolar junction transistor. Alkene chains containing four

and six carbon atoms were also studied, yielding devices

with characteristics similar to that shown in figure 3, with

the maximum current I

1

shifting to smaller values of V

32

with increasing chain length. As evidence that the transistor

behaviour shown in figure 3 is due to the tunable interference

mechanism discussed above, we point out that if hopping

between the benzene ring and the alkene side group is set to

zero, so that the coupling of the side group to benzene is purely

electrostatic, almost no current flows between leads 1 and 2.

For Gamma1

3

= 0, I

3

= 0 and the QuIET behaves as a field-

effect transistor. The transconductance dI/dV

32

of such a

device is shown in figure 4. For comparison, we note that an

ideal single-electron transistor [28] with Gamma1

1

= Gamma1

2

= 1eVhas

peak transconductance (1/17)G

0

at bias voltage .25 V, and

(1/2)G

0

at bias 1 V, where G

0

= 2e

2

/h is the conductance

quantum. For low biases, the proposed QuIET thus has

a higher transconductance than the prototypical nanoscale

amplifier, while even for large biases its peak transconductance

is comparable. Likewise, the load resistances required for a

QuIET to have gain (load times transconductance) greater than

one while in its ‘on’ state are comparable to other nanoscale

devices, ~10/G

0

.

Operation of the QuIET does not depend sensitively on the

magnitude of the lead–molecule coupling

ˉ

Gamma1 = Gamma1

1

Gamma1

2

/(Gamma1

1

+

Gamma1

2

). The current through the device decreases with decreasing

ˉ

Gamma1, but aside from that, the device characteristic was found

to be qualitatively similar when

ˉ

Gamma1 was varied over one order

of magnitude. The QuIET is also insensitive to molecular

vibrations: only vibrational modes that simultaneously

alter the carbon–carbon bond lengths and break the six-

fold symmetry within the benzene component can cause

Figure 4. Transconductance dI/dV

32

of a QuIET based on

sulfonated vinylbenzene with Gamma1

3

= 0. The characteristic is similar to

that of a field-effect transistor, i.e. I

3

= 0 while I

1

=?I

2

= I .Asin

figure 3, Gamma1

1

= Gamma1

2

= 1 eV, and the calculation was done for room

temperature. Reprinted with permission from [11]. ? 2006

American Chemical Society.

decoherence in a benzene ‘interferometer’. Such modes are

only excited at temperatures greater than about 500 K.

The position of the third lead affects the degree to which

destructive interference is suppressed. For benzene, the most

effective location for the third lead is shown in figure 1.Itmay

also be placed at the site immediately between leads 1 and 2,

but the transistor effect is somewhat reduced, since coupling

to the charge carriers is less. The third, three-fold symmetric

configuration of leads completely decouples the third lead from

electrons travelling between the first two leads. For each

monocyclic aromatic annulene, one three-fold symmetric lead

configuration exists, yielding no transistor behaviour.

While H

mol

of equation (2) is well known to reproduce

the basic experimental features of conjugated molecules [29],

the QuIET’s characteristics, based on general principles

of quantum mechanics and symmetry, are qualitatively

independent of the particulars of the quantum chemical

method. Inclusion of σ electrons, for example, has little effect:

they form a separate system of localized bonds, and so cannot

contribute strongly to transport. We have verified via an all-

valence extended H¨uckel theory [30] that the QuIET’s tunable

coherent current suppression persists for such extensions of the

basis set.

The QuIET’s operating mechanism, tunable coherent

current suppression, occurs over a broad energy range within

the gap of each monocyclic aromatic annulene; it is thus a

very robust effect, insensitive to moderate fluctuations of the

electrical environment of the molecule. Although based on an

entirely different, quantum-mechanical, switching mechanism,

the QuIET nonetheless reproduces the functionality of

macroscopic transistors on the scale of a single molecule.

Furthermore, since the current flow is not blocked by an energy

barrier, which must be raised and lowered with each switching

cycle, heating of the environment is greatly reduced.

4. Implementations

As daunting as the fundamental problem of the switching

mechanism is the practical one of nanofabrication. The QuIET

requires a third lead coupled locally to the central molecule,

and, while there has recently been significant progress in

4

Nanotechnology 18 (2007) 424014 C A Stafford et al

(a) (b) (c)

(d)

Figure 5. Source–drain lead configurations possible in a QuIET based on [18]annulene. The bold lines represent the positioning of the two

leads. Each of the four arrangements has a different phase difference associated with it: (a) π;(b)3π;(c)5π;and(d)7π. The slight

deviations of [18]annulene’s hydrogen atoms from the molecular plane do not significantly affect the QuIET’s switching mechanism.

Reprinted with permission from [11]. ? 2006 American Chemical Society.

Figure 6. Schematic of various QuIETs based on a benzene ring. A, B, and C represent the various substituents which may be placed in series

between the ring and each lead. In particular, the conducting polymers like polyaniline and polythiophene may be useful in overcoming the

‘third lead’ problem.

V

a

V

b

R

Figure 7. Possible embodiment of a QuIET integrated with conventional circuitry on a chip. The source (1) and drain (2) electrodes are

connected via conducting polymers (in this case, polythiophene) to the central aromatic ring, while the gate electrode (3) is coupled

electrostatically to an alkene side group.

(This figure is in colour only in the electronic version)

that direction [14, 31, 32], to date, only two-lead single-

molecular devices, sometimes with global gating, have been

achieved [13]. With this in mind, we turn to potential practical

realizations of the device.

Using novel fabrication techniques, such as ultra-sharp

STM tips [31] or substrate pitting [32], it may soon be possible

to attach multiple leads to large molecules. Fortunately, the

QuIET mechanism applies not only to benzene, but to any

5

Nanotechnology 18 (2007) 424014 C A Stafford et al

monocyclic aromatic annulene with leads 1 and 2 positioned

so the two most direct paths have a phase difference of

π. Furthermore, larger molecules have other possible lead

configurations, based on phase differences of 3π,5π,etc;

as an example, figure 5 shows the lead configurations for a

QuIET based on [18]annulene. Other large ring-like molecules

that possess an aromatic conjugated electron system, such as

[14]annulene and the divalent metal-phthalocyanines, would

also serve well.

Another method of increasing the effective size of the

molecule is to introduce molecular wires linking the central

ring and leads (see figures 6 and 7). Conducting polymers,

such as polythiophene or polyaniline, are ideal for this task.

Such changes can be absorbed into the diagonal elements of

the self-energy Sigma1(E), and so only modify G(E) locally. As

such, while they can significantly modify the on-resonance

behaviour of a molecular device, the off-resonance function is

largely unaltered. In particular, the transmission node at the

centre of the gap is unaffected. An example of such a QuIET

integrated with conventional circuitry on a chip is shown in

figure 7.

5. Conclusions

The quantum interference effect transistor represents one way

to simultaneously overcome the problems of scalability and

power dissipation which face the next generation of transistors.

Because of the exact symmetry possible in molecular devices,

it possesses a perfect mid-gap transmission node, which serves

as the off state for the device. Tunably introduced decoherence

or elastic scattering can lift this quantum interference effect,

with the result of current modulation. Furthermore, a vast

variety of potential chemical structures possess the requisite

symmetry, easing fabrication difficulties. In particular,

molecular wires, such as conducting polymers, can be used to

extend the molecule to arbitrary size.

Acknowledgments

This work was supported in part by National Science

Foundation Grant Nos PHY0210750, DMR0312028, and

DMR0705163.

References

[1] Landauer R 1961 IBMJ.Res.Dev.5 183

[2] International Technology Roadmap for Semiconductors: 2006

Update http://public.itrs.net

[3] Di Ventra M, Pantelides S T and Lang N D 2000 Appl. Phys.

Lett. 76 3448

[4] Sautet P and Joachim C 1988 Chem. Phys. Lett. 153 511

[5] Sols F, Macucci M, Ravaioli U and Hess K 1989 Appl. Phys.

Lett. 54 350

[6] Joachim C and Gimzewski J K 1997 Chem. Phys. Lett. 265 353

Joachim C, Gimzewski J K and Tang H 1998 Phys. Rev. B

58 16407

[7] Baer R and Neuhauser D 2002 J. Am. Chem. Soc. 124 4200

[8] Yaliraki S N and Ratner M A 2002 Ann. New York Acad. Sci.

960 153

[9] Stadler R, Forshaw M and Joachim C 2003 Nanotechnology

14 138

Stadler R, Ami S, Forshaw M and Joachim C 2003

Nanotechnology 14 722

[10] Washburn S and Webb R A 1986 Adv. Phys. 35 375

[11] Cardamone D M, Stafford C A and Mazumdar S 2006 Nano

Lett. 6 2422

[12] B¨uttiker M 1988 IBMJ.Res.Dev.32 63

[13] Nitzan A and Ratner M A 2003 Science 300 1384 and

references therein

[14] Piva P G, DiLabio G A, Pitters J L, Zikovsky J, Rezeq M,

Dogel S, Hofer W A and Wolkow R A 2005 Nature 435 658

[15] Ohno K 1964 Theor. Chim. Acta 2 219

[16] Chandross M, Mazumdar S, Liess M, Lane P A, Vardeny Z V,

Hamaguchi M and Yoshino K 1997 Phys. Rev. B 55 1486

[17] Stafford C A, Kotlyar R and Das Sarma S 1998 Phys. Rev. B

58 7091

[18] Tian W, Datta S, Hong S, Reifenberger R, Henderson J I and

Kubiak C P 1998 J. Chem. Phys. 109 2874

[19] Nitzan A 2001 Annu. Rev. Phys. Chem. 52 681

[20] Jauho A-P, Wingreen N S and Meir Y 1994 Phys. Rev. B

50 5528

[21] Datta S 1995 Electronic Transport in Mesoscopic Systems

(Cambridge: Cambridge University Press) pp 293–342

[22] Mujica V, Kemp M and Ratner M A 1994 J. Chem. Phys.

101 6849

[23] B¨uttiker M 1986 Phys.Rev.Lett.57 1761

[24] Feynman R P and Hibbs A R 1965 Quantum Mechanics and

Path Integrals (New York: McGraw-Hill)

[25] Luttinger J M 1960 Phys. Rev. 119 1153

[26] Clerk A A, Waintal X and Brouwer P W 2001 Phys.Rev.Lett.

86 4636

[27] Marder M P 2000 Condensed Matter Physics (New York:

Wiley)

[28] Kastner M A 1992 Rev. Mod. Phys. 64 849

[29] Salem L 1966 The Molecular Orbital Theory of Conjugated

Systems (New York: Benjamin)

[30] Hoffman R 1963 J. Chem. Phys. 39 1397

[31] Rezeq M, Pitters J and Wolkow R 2006 J. Chem. Phys.

124 204716

[32] Mativetsky J M, Burke S A, Fostner S and Grutter P 2007 Small

3 818

6

献花(0)
+1
(本文系mc_eastian首藏)