Cuevas Book
Cuevas Book
Publishers’ page
i
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
Publishers’ page
ii
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
Publishers’ page
iii
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
Publishers’ page
iv
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
To our families
v
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
Preface
vii
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
1 See section 1.3 for a more detailed description of the structure and scope of the book.
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
Acknowledgments
It would not have been possible to write the book without the help of many
coworkers and colleagues. First of all, we want to thank Edith Goldberg
for encouraging one of us (JCC) to give a postgraduate course on molecular
electronics in the fall of 2008 in Santa Fe (Argentina). The excellent stu-
dents who attended that course demonstrated that, after a 50-hours course
and without any previous knowledge about this field, one can master the
basic concepts and techniques that now form the body of this monograph.
This fact provided the final boost that we needed to collect all our notes
and turn them into this book.
Similarly, for the experimental point of view of this book, the students
in the graduate course at Konstanz served as test candidates. Some of them
even got contaminated by this exciting field and went on asking questions
what finally resulted in contributions to this book. Very valuable input
came from my colleague Artur Erbe who was the real expert in molecular
electronics in our Department until he left to Dresden.
We also want to express our gratitude to Alvaro Martı́n Rodero, who
not only introduced one of us (JCC) to the exciting field of nanoelectronics,
but also contributed decisively to this manuscript with his personal notes,
which are the basis of several chapters of the theoretical background. The
same holds for Hilbert von Löhneysen and Cristián Urbina who sent the
other one of us (ES) to perform experiments with nanoelectronic circuits.
We would especially like to thank our coworkers Fabian Pauly, Janne K.
Viljas, Michael Häfner, Sören Wohlthat, Stefan Bilan, Linda A. Zotti, Cécile
Bacca, Stefan Bächle, Tobias Böhler, Uta Eberlein, Stefan Egle, Daniel
Guhr, Ning Kang, Thomas Kirchner, Christian Kreuter, Shou-Peng Liu,
Youngsang Kim, Hans-Fridtjof Pernau, Olivier Schecker, Christian Schirm,
Dima Sysoiev, Simon Verleger, and Reimar Waitz. They have contributed
ix
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
to this manuscript with many results, special figures and very important
suggestions and critical comments about the text.
Thanks go also sincerely to our colleagues who have read different parts
of the manuscript and have provided helpful comments: Douglas Natelson,
Abraham Nitzan, Wilson Ho, Latha Venkataraman, and Arunava Majum-
dar.
This monograph reflects our view of this field, which has emerged thanks
to the collaboration and exchange of ideas with many colleagues over the
years. So in this respect, we want to thank Alfredo Levy Yeyati, Gerd
Schön, Jan Heurich, Wolfgang Wenzel, Jan M. van Ruitenbeek, Nicolás
Agraı̈t, Gabino Rubio, Roel Smit, Oren Tal, Markus Dreher, Peter Nielaba,
Christoph Sürgers, Maya Lukas, Christoph Strunk, Sophie Géron, Richard
Berndt, Paul Leiderer, Wolfgang Belzig, Marcel Mayor, Thomas Huhn,
Andreas Marx, Ulrich Steiner, and Ulrich Groth.
We also want acknowledge the contribution of all the authors who have
kindly granted us the permission to reprint their work in this monograph.
Finally, I (JCC) want to thank my parents and brothers for being always
by my side. I also want to thank Ana for being so patient and share my
time with this book for too many nights and weekends. ES thanks her
family for continuous support and reminding me steadily of what is really
important in life.
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
Contents
Preface vii
Acknowledgments ix
xi
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
Theoretical background 75
Contents xiii
Contents xv
Contents xvii
Contents xix
Appendixes 621
Appendix A Second Quantization 623
A.1 Harmonic oscillator and phonons . . . . . . . . . . . . . . 624
A.1.1 Review of simple harmonic oscillator quantization 624
A.1.2 1D harmonic chain . . . . . . . . . . . . . . . . . 626
A.2 Second quantization for fermions . . . . . . . . . . . . . . 628
A.2.1 Many-body wave function in second quantization 628
A.2.2 Creation and annihilation operators . . . . . . . . 630
A.2.3 Operators in second quantization . . . . . . . . . 632
A.2.4 Some special Hamiltonians . . . . . . . . . . . . . 634
A.3 Second quantization for bosons . . . . . . . . . . . . . . . 637
A.4 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . 638
Bibliography 639
Index 697
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
PART 1
1
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
2
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
Chapter 1
How does the electrical current flow through a single molecule? Can a
molecule mimic the behavior of an ordinary microelectronics component or
maybe provide a new electronic functionality? How can a single molecule be
addressed and incorporated into an electrical circuit? How to interconnect
molecular devices and integrate them into complex architectures? These
questions and related ones are by no means new and, as we shall see later in
this chapter, they were already posed many decades ago. The difference is
that we are now in position to at least address them in the usual scientific
manner, i.e. by providing quantitative experimental and theoretical results.
The advances in the last two or three decades, both in nanofabrication
techniques and in the quantum theory of electronic transport, allow us now
to explore and to understand the basic properties of rudimentary electrical
circuits in which molecules are used as basic building blocks. It is worth
stressing right from the start that we do not yet have definitive answers for
the questions posed above. However, a tremendous progress has been made
in recent years and some concepts and techniques have already been firmly
established. In this sense, one of main goals of this book is to review such
progress, but more importantly, this monograph is intended to provide a
solid basis for the new generation of researchers that should take the field
of molecular electronics to the next level.
Molecular electronics, as used in this book, is defined as the field of
science that investigates the electronic and thermal transport properties of
circuits in which individual molecules (or an assembly of them) are used as
basic building blocks.1 Obviously, some of the feature dimensions of such
1 Molecular electronics, in the sense used here, should not be confused with organic
electronics, the field in which molecular materials are investigated as possible constituents
of a variety of macroscopic electronic devices.
3
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
MESOSCOPIC QUANTUM
PHYSICS CHEMISTRY
ELECTRICAL INORGANIC
ENGINEERING CHEMISTRY
MATERIAL ORGANIC
SCIENCE CHEMISTRY
BIOLOGY
molecular circuits are of the order of nanometers (or even less) and there-
fore, molecular electronics should be viewed as a subfield of nanoscience
or nanotechnology in which traditional disciplines like physics, chemistry,
material science, electrical engineering and biology play a fundamental role
(see Fig. 1.1). Molecular electronics, in the sense of a potential technology,
is based on the bottom-up approach where the idea is to assemble elemen-
tary pieces to form more complex structures, as opposed to the top-down
approach where the idea is to shrink macroscopic systems and components.
Molecular electronics has emerged from the constant quest for new tech-
nologies that could complement the silicon-based electronics, which in the
meantime it has become a true nanotechnology. It seems very unlikely
that molecular electronics will ever replace the silicon-based electronics,
but there are good reasons to believe that it can complement it by provid-
ing, for instance, novel functionalities out of the scope of traditional solid
state devices. More importantly, molecular electronics has become in recent
years a true field of science where many basic questions and quantum phe-
nomena are being investigated. In this sense, the importance of molecular
electronics is unquestionable and we are convinced that different traditional
disciplines will benefit from advances in this new field.
In the rest of this introductory chapter, we shall first try to answer the
questions of why it is worth pursuing molecular electronics research and
why it is interesting to work in a field like this. Then, in section 1.2 we
shall briefly review the complex history of this field to set the stage for
this book. Finally, in section 1.3 we shall clearly define the scope of this
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
This is probably the first time that the term molecular electronics was
used publicly, although it originally referred to a new strategy for the fab-
rication of electronic components, and it had yet little to do with the vision
of using individual molecules as electronically active elements. Fig. 1.2
summarizes the vision of colonel Lewis, where molecular electronics should
constitute be the next breakthrough in electronics, although it was not yet
clear what molecular electronics was supposed to mean.
The collaboration between Westinghouse and the US Air Force, which
started after the mentioned conference, lasted a few years and certain
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
Electronic
Equipment Conventioal Miniaturization Microminiaturization
Trend Component
Size Reduction (Transistor) (Integrated Circuit)
Size Breakthrough!!!
Weight
and
Space Asymptote for Present Technologies
Reduction
Molecular
Electronics
Asymptote for Molecular Technologies
Fig. 1.2 Graph presented by colonel Lewis of the US Air Force in the first conference
on molecular electronics held in November 1958. Here, one can see the trend in the
miniaturization of the electronic components during the 1940’s and 1950’s. According to
Lewis, molecular electronics should have constituted the next breakthrough in electronics
by the end of the 1950’s. Adapted from [3].
Fig. 1.3 Measurements of the low-bias tunneling conductivity (σt ) vs. the distance (d)
between the electrodes in Al/S(n)/Hg junctions. Here, S(n) stands for monolayers of
Cd salt of fatty acid CH3 (CH2 )n−2 COOH with different lengths (n ranges between 18
and 21). The solid line is a linear fit to the experiment data. The measurements were
performed at two different temperatures: 20 and -35 o C. Reprinted with permission from
[6]. Copyright 1971, American Institute of Physics.
Fig. 1.4 Principle of a local probe like the scanning tunneling microscope: The gentle
touch of a nanofinger. If the interaction between tip and sample decays sufficiently
rapidly on the atomic scale, only the two atoms that are closest to each other are able to
“feel” each other. Reprinted with permission from [12]. Copyright 1999 by the American
Physical Society.
indeed performed with the STM, but the experiment of Reed et al. is the first one realized
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
(a) (c)
Gold Gold
Gold wire electrode electrode
(b) (d)
Gold Gold
Gold wire
electrode electrode
SAM
Fig. 1.5 Schematics of the first transport measurements through single-molecule junc-
tions performed with the MCBJ technique [16]. (a) The gold wire of the break junc-
tion before breaking and tip formation. (b) After addition of benzene-1,4-dithiol, self-
assembled monolayers (SAMs) form on the gold wire surfaces. (c) Mechanical breakage
of the wire in solution produces two opposing gold contacts that are SAM-covered. (d)
After the solvent is evaporated, the gold contacts are slowly moved together until the
onset of conductance is achieved.
(a) (b)
(e)
(c) (d)
(f)
Fig. 1.6 Nanoscale molecular-switch crossbar circuits. (a) An optical microscope image
of an array of four test circuits, showing that each has 16 contact pads with micron-
scale connections leading to nanoscale circuits in the center. (b) An image taken with
a scanning electron microscope (SEM) showing two mutually perpendicular arrays of
nanowires connected to their micron-scale connections. (c) A SEM image showing that
the two sets of nanowires cross each other in the central area. (d) A 3D image of the
crossbar taken with an atomic force microscope. (e) Schematic representation of the
crossbar circuit structure in which monolayer of the [2]rotaxane is sandwiched between
an array of Pt/Ti nanowires on the bottom and an array of Pt/Ti nanowires on the top.
(f) Molecular structure of the bistable [2]rotaxane R. Reprinted with permission from
[26]. Copyright 2003 IOP Publishing Ltd.
By now molecular electronics is a very broad field with many different inter-
esting aspects and special topics. These topics can be divided in a natural
way into those related to the development and potential applications of
molecular devices and those concerning the novel physical phenomena that
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
body. All the topics are discussed in a didactic and self-contained manner
so that students without a previous knowledge on these topics should be
able, after reading this part, to follow the theory papers in this field. To
be precise, this part starts in Chapter 4 with an introduction to the scat-
tering (or Landauer) approach, which provides an appealing framework to
describe coherent transport in nanostructures. Then, we go on with several
chapters devoted to Green’s function techniques (Chapters 5-8), which pro-
vide powerful tools to compute equilibrium and nonequilibrium properties
of atomic-scale junctions beyond the capabilities of the scattering approach.
Finally, Chapters 9 and 10 deal with the two most widely used electronic
structure methods in molecular electronics, namely the tight-binding ap-
proach and density functional theory. These methods in combination with
the Green’s function techniques provide the starting point for the realistic
description of the transport properties of atomic and molecular junctions.
Let us emphasize that at the end of every chapter one can find several
exercises that have been chosen to illustrate the main concepts.
Part 3 presents a basic description of the physics of atomic-sized con-
tacts. Although this is not the main topic of the book, it is crucial to
have a basic knowledge about the transport properties of the metallic wires
that are then used as electrodes in molecular junctions. We have divided
this part into two chapters where we describe the physics of non-magnetic
atomic contacts (Chapter 11) and magnetic ones (Chapter 12).
Finally, Part 4 presents a detailed review on the transport through
molecular junctions. We have organized the material according to the phys-
ical mechanism which dominates the transport properties. Thus, we start
this part with two chapters devoted to the coherent transport in molecular
junctions (Chapters 13 and 14). Then, we discuss in Chapter 15 the physics
of the so-called molecular transistors, which are nothing but weakly coupled
molecular junctions where the transport is dominated by electronic corre-
lations that lead to phenomena like Coulomb blockade or the Kondo effect.
We then proceed to discuss in Chapters 16 and 17 the role of molecular
vibrations in the electrical current through molecular junctions. Chapter
19 is devoted to other transport properties beyond conductance and we
discuss there, in particular, shot noise and thermal transport in molecular
conductors. The optical properties of current-currying molecular junctions
are the subject of Chapter 20. Chapter 18 deals with the electronic trans-
port in long molecules where the hopping (or incoherent) transport regime
is realized. Finally, we conclude this part in Chapter 21 with a list of topics
that have not been addressed in this monograph and we indicate where to
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
Chapter 2
2.1 Introduction
In this chapter we shall present the most common methods which have
been developed during the last years for the fabrication of metallic atomic-
size contacts. Both the contacting methods and the physical properties
of atomic contacts found the basis for contacting single molecules. On
the other hand, these techniques have been further refined for contacting
molecules. These refinements are now also used for studying atomic con-
tacts. Therefore, the decision in which chapter one or the other method
is described is somewhat arbitrary. Manifold variations of the techniques
exist and are permanently improved further. The aim of this chapter is to
introduce into the most important principles and to compare the techniques
regarding their advantages and drawbacks.
As important as the sample preparation is the quality of the electronic
transport measurements. When dealing with tiny contacts, care has to be
taken to reduce the influence of the measurement onto the contact itself.
We will therefore end this chapter with a few brief remarks about the most
common measurement setups and possible artifacts.
One of the most versatile tools for the fabrication of atomic-size contacts
and atomic chains is the scanning tunneling microscope (STM) (for a re-
view, see Ref. [54]). It has been used for that purpose from the very be-
ginning of its invention [55]. While in the standard application of an STM
a fine metallic tip is held at distance from a counter electrode (in general
19
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
motion
Fig. 2.1 Working principle of the fabrication of atomic contacts with an scanning tun-
neling microscope (STM). The electron micrograph shows a STM tip. The width at half
length is in the order of 100 to 200 µm. The lower inset gives an artist’s view of the
atomic arrangement of an atomic contact. Courtesy of C. Bacca.
laser detector
(a) conductive (b)
cantilever
AFM cantilever
x-y motion
cantilever beam
metallic tip
metal sample
metal tip
conductive
sample
Fig. 2.2 Fabrication and characterization of atomic contacts with an atomic force mi-
croscope (AFM). (a) The conductive AFM uses a conductive cantilever and metallic
tip for recording the electrical signal. The deflection of the cantilever beam is detected
optically and used for recording the topographic information of sample. After Ref. [59].
(b) In the combined AFM-STM the sample is clamped to a cantilever. The metallic
contact is formed between the sample and the metal tip. The metal tip is part of an
STM and records the electrical signal. The deflection of the cantilever is recorded with a
separate AFM. This signal is used for measuring the force acting on the cantilever when
the atomic contact rearranges. After Ref. [58].
Transient atomic chains and contacts with lifetimes in the millisecond range
can also be fabricated in a table-top experiment first demonstrated by N.
Garcia and coworkers [60], which we call here “dangling-wire contacts”.
Two metal wires in loose contact to each other are excited to mechanical
vibrations, such that the contact opens and closes repeatedly. One end
of each wire is connected to the poles of a voltage source and the current
is recorded with a fast oscilloscope. This method is in principle particu-
larly versatile because it enables the formation of heterojunctions between
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
Fig. 2.3 Experimental setup used to visualize contacts between macroscopic metallic
electrodes inside a scanning electron microscope (SEM). Adapted with permission from
[61]. Copyright 1997 by the American Physical Society.
Another interesting method for preparing and imaging atomic contacts are
transient structures forming in a transmission electron microscope (TEM)
when irradiating thin metal films onto dewetting substrates [62, 63]. The
high energy impact caused by the intensive electron beam locally melts the
metal film causing the formation of constrictions which eventually shrink
down to the atomic size and finally pinch-off building a vacuum tunnel gap.
A typical system for these studies is Au on glassy carbon substrates. Several
variations of this principle have been developed that allows one to contact
both electrodes forming the contact, see Fig. 2.4. The high electron current
density necessary for imaging causes also high local temperatures resulting
in short lifetimes of these contacts. However, they offer the unique possi-
bility to simultaneously perform conductance measurements and imaging
with atomic precision. Similar results have been obtained with variations
of the STM inside a TEM [64]. This method enabled to directly prove
the existence of single-atom contacts, single-atom wide and several atom
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
Fig. 2.4 High resolution TEM images of short atomic wires fabricated with an STM
inside the vacuum chamber of the TEM. The arrows indicate the number of atomic rows.
In panel f the contact is broken and forms a tunnel contact. Reprinted by permission
from Macmillan Publishers Ltd: Nature [63], copyright 1998.
Already before the development of the first STM another technique en-
abling the fabrication of atomic-size contacts and tunable tunnel contacts
has been put forward. The first realizations include the needle-anvil or
wedge-wedge point contact technique pioneered by Yanson and co-workers
(for a review see [66]) and the squeezable tunnel junction method described
by Moreland and Hansma [67] and Moreland and Ekin [68] who used metal
electrodes on two separate substrates which are then carefully adjusted with
respect to each other. The needle-anvil technique was mainly used to form
contacts with diameters of typically several nanometers and thus having
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
L
counter supports
metal
wire
u
sacrificial layer
Fig. 2.5 Working principle of the MCBJ (not to scale) with the metal wire, the elastic
substrate, the insulating sacrificial layer, the pushing rod, the counter supports and the
dimensions used for calculating the reduction ratio (see text).
Fig. 2.6 The 100 nm wide gold wire is glued with epoxy resin (black) onto the substrate.
The electrical contact is made by thin copper wires glued with silver paint. The inset
shows a zoom into the notch region between the two black drops of epoxy resin. Reprinted
from [15]. Copyright 2003, with permission from Elsevier.
measurement circuit at both ends. The distance between the glue drops is
of the order of 50 µm to 200 µm.
Variations of this method have been put forward which enable contact-
ing of reactive or brittle materials out of which no wires can be formed [72].
For this purpose the sample preparation is performed in protective environ-
ment. A beam-shaped piece of the material is cut in a non-reactive liquid
such as dodecanol or other slowly evaporating alcohols, or glycerine. Four
holes are drilled into the metal and a wedge is cut in the middle between
the holes. An example is shown in Fig. 2.7. The beam is screwed with the
help of two electrically isolating bolts to the substrate, one on each side of
the wedge. The remaining two holes serve for screwing metallic wires to
the beam for the conductance measurements.
For a version which enables scanning the two electrodes with respect to
each other, at first two piezo tubes are glued to the substrate. The metal
wire is then glued on top of the piezos. After mechanically breaking the
wire, the piezos are polarized such that they are bent and the two parts
of the wire are sliding along each other [73]. This realization corresponds
to a high-stability STM, but with very restricted scan possibility. It is
therefore used only sparsely. Finally, simultaneous force and conductance
measurements are possible when adding a tuning fork like in AFMs. Details
of this very sophisticated method are given in Ref. [75].
Fig. 2.8 shows two examples of thin-film MCBJs, which were fabricated
using the usual techniques of nanofabrication, i.e. electron beam lithography
and metal deposition by evaporation. There are mainly two differences to
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
Fig. 2.7 Principle of the MCBJ technique adapted for reactive metals. Reprinted from
[15]. Copyright 2003, with permission from Elsevier.
(a) (b)
2 µm
100 nm
Fig. 2.8 Lithographic MCBJ. (a) Electron micrograph of a thin-film MCBJ made of
cobalt on polyimide taken under an inclination angle of 60o with respect to the normal.
The distance between the rectangular shaped electrodes is 2 µm, the thickness of the thin
film is 100 nm and the width of the constriction at its narrowest part is approximately
100 nm. (b) Electron micrograph of a thin-film MCBJ made of cobalt (medium grey) with
leads made of gold (light grey) taken under an inclination angle of 50o with respect to the
normal. The distance between the rectangular shaped electrodes is 2 µm, the thickness
of the Co film is 80 nm, of the Au film is 100 nm and the width of the constriction at its
narrowest part is approximately 100 nm. The sample has been fabricated using shadow
evaporation through a suspended mask such that two images of the mask exist. The Au
shadow of the bridge is broken off.
The metal should have a high elastic deformation limit. Typical metals
are bronze or spring steel. For particular purposes, in particular when
capacitive effects have to be minimized, the metal is replaced by a plastic
substrate. Both metal or plastic are thoroughly polished to reduce the
roughness to less than a micrometer. The remaining corrugations are then
filled with a thin layer of polyimide (thickness 1-2 µm), which is spin-coated
and hardbaked in vacuum. The polyimide also serves as electrical insulator
between the nanostructure and the substrate. Subsequently the electron
resist is spin-coated and thermally treated as required for electron beam
structuring. Fig. 2.9(c) shows an example in which a double-layer resist
is used. The double-layer is necessary for, e.g. evaporation of the metal
under arbitrary angle. The next step is electron-beam writing in a scanning
electron microscope equipped with a pattern generator or in a commercial
electron-beam writer. After development of the resist in a selective solvent
the resist mask remains on top of the polyimide layer. The mask itself may
be partially suspended when using a double-layer resist. Subsequently the
metal will be deposited either by evaporation, sputtering, chemical vapor
deposition or other means. Shadow evaporation, i.e. evaporation of several
materials under different angles can be used for forming contacts between
different metals or for supplying nanobridges of one metal with electrodes
made of another metal. The advantage of the shadow-evaporation technique
lies at first in its self-alignment property because the same mask is used
for all metal depositions. The second advantage is given by the fact that
all depositions can be made in a single vacuum step, which enables one to
fabricate clean interfaces between the metals. After the metal deposition
the mask is stripped in a more aggressive solvent. Finally the structure is
exposed to an isotropic oxygen plasma which attacks the polyimide layer.
Consequently its thickness is reduced and all narrow metal parts, like the
nanobridge become suspended like a bridge.
Both versions of the technique - the notched-wire MCBJs and the litho-
graphic MCBJs - share the idea of enhanced stability due to the formation
of the contact by breaking the very same piece of metal on a single sub-
strate and by transformation of the motion of the actuator into a much
reduced motion of the electrodes perpendicular to it. The small dimensions
of the freestanding bridge-arms give rise to high mechanical eigenfrequen-
cies, much higher than the ones of the setup. As a result the system is less
sensitive to mechanical perturbations by vibrations.
Assuming homogeneous beam-bending of the substrate we can calculate
the reduction ratio r between the length change of the bridge u and the
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
(a) (b)
(c) (d)
(e) (f)
(g) (h)
Fig. 2.9 Fabrication scheme of thin-film (lithographic) MCBJ. (a) The substrate (metal,
plastic) is polished mechanically. (b) the sacrificial wafer (polyimide) is spin-coated and
baked. (c) The resin (typically a bi-layer electron sensitive organic material) is spin-
coated and baked. (d) The resin is exposed in an electron beam writer or a scanning
electron microscope equipped with a pattern generator in the desired pattern. (e) The
chip is developed in a solvent which selectively removes the exposed parts of the resin.
The result is a mask, which resides on the sacrificial layer, in the shape of the exposed
pattern. (f) The metal is deposited by evaporation or sputtering. (g) The mask with the
metal on top of it is lifted-off in a more aggressive solvent which attacks the unexposed
parts of the resin. The result is a metal layer in the shape of exposed pattern. (h) Finally
the thickness of the sacrificial layer is reduced in an isotropic plasma. The narrow parts
of the metal pattern are suspended and form the bridge which will be broken in the
MCBJ mechanism.
6tu
r=, (2.1)
L2
where t is the thickness of the substrate, u the length of the free-standing
bridge arms and L the distance of the counter supports. This quantity
denotes the factor with which any motion of the pushing rod is reduced
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
to
to rotary axis
thermal bath
thread section
with pitch A
guiding wedge-ended
rod pushing rod
sample thread
t section
holder with pitch B
sample
Fig. 2.10 Sample holder with differential screw for thin-film MCBJ. A motor drives a
rotary axis which ends in a thread with two different pitches. Rotating the axis results
in varying distance between ground plate and sample holder. The sample resides on two
counter supports connected to the sample holder. It is bent by the pushing rod which
is attached to the ground plate. Three guiding rods (only one of which is shown) ensure
smooth and linear motion.
500 nm 500 nm
200 nm 200 nm
port changes from ohmic behavior, i.e. limited by scattering events of the
electrons to wave-like electronic transport, which can be described by the
Landauer picture (see Chapter 4).
ductor is locally reduced and its electrical resistance increases. The higher
resistance causes higher losses, enhanced dissipation, increasing tempera-
ture in the wire which further enhances the dissipation of ions. An impor-
tant role plays the temperature of the lattice because the diffusion and the
threshold current strongly depend on temperature. Electromigration has
become one of the most important origins of failures in integrated circuits,
due to the miniaturization of the metallic interconnects without reducing
the current by the same factor. Consequently, electromigration has widely
been studied in electrical engineering with the aim to achieve the highest
possible threshold current density for it to set in and the smallest diffusion
speed [85].
For the formation of atomic contacts a high threshold current is not
important but the possibility for controlling speed, shape and size of the
final structure. One of the most important preconditions is to define the
position at which the electromigration starts, and the contact forms. For
this purpose a short and thin metallic wire is fabricated by lithographic
methods as described in the previous section. Typical dimensions are a
length and width of 50 to 100 nm and a thickness of 10 to 20 nm. The
thin wire is connected to wider and thicker electrodes which consequently
have smaller resistivity. A convenient method to fabricate these structures
is shadow evaporation through a suspended mask as shown in Fig. 2.11.
First, thin layers of the metal (typical thickness 10 nm) are evaporated
under the angles Θ and −Θ. The angle is chosen such that both layers
slightly overlap underneath the suspended part of the mask. Afterwards a
thick layer of the electrode metal is deposited perpendicular to the substrate
plane. The ideal structure would consist of a single-crystalline wire in the
thin part of the wire, the boundaries of which are covered by the thick
electrodes in order to avoid electromigration of possible contaminants from
the grain boundaries. It is advantageous to work on a substrate with high
thermal conductivity in order to control the temperature.
The electromigration process itself is performed such that an electrical
current is continuously ramped up while the resistivity is monitored. As
soon as the resistance starts to increase a computer-controlled feedback
loop controls the current such that the rate of the resistance increase is
kept constant or slowed down. The resistance increase is partially due to
the temperature increase caused by the Joule heating of the driving current.
Although it has been shown that in the ohmic regime the current density is
the quantity which determines the diffusion of the ions, it is advantageous
to control the voltage in order to produce atomic size contacts. When the
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
resistance increases the current becomes smaller, which helps to limit the
migration speed. The low-resistive electrodes ensure that the voltage drops
locally making the driving force acting only locally as well. Consequently,
the dissipation and Joule heat generation are local as well. The procedure
should be stopped when the desired resistance is achieved. For the study
of atomic contacts the interesting regime is reached when the resistance
exceeds roughly one kiloohm. For usual metals this corresponds to contacts
with a narrowest cross section of roughly 10 atoms. An important finding
is that the behavior changes markedly when the size of the smallest cross
section corresponds to a few atoms. However, the exact position of the
position at which the wire finally breaks is difficult to predict. As will be
explained in Chapter 11 the electrical transport of contacts of this size is
determined by the wave properties of the electrons rather than by collisions
with defects. If this happens the resistance may start to decrease again
before the wire finally is burned through. This non-monotonous behavior
complicates the control scheme further. Several control schemes have been
put forward which are optimized for various sample geometries, metals and
working conditions such as vacuum or low temperature [21, 81–83, 86]. So
far only a few studies exist in which the electromigration process has been
imaged in detail, although these kind of studies are very insightful. One
example is shown in Fig. 2.11, where AFM images have been taken after
discrete electromigration steps. A particularly nice series of TEM images
showing that the most dramatic shape changes occur during the final phase
can be found in Ref. [82].
An important difference to STM techniques and MCBJs is the fact that
the wire forming the contact is in solid contact with a substrate. The ad-
vantages are at first ultimate stability which will become important when
studying atomic or molecular junctions as a function of external fields (see
Chapters 12 and 20). The second advantage lies in the fact that no par-
ticular requirements exist for the properties of the substrate, besides the
fact that it should be sufficiently insulating. Often silicon - the standard
substrate in microelectronics - is used. With suitable doping it can be used
as back-gate for inducing an electric potential and building a three-terminal
device. This technique is important for studying effects like Coulomb block-
ade, which will be explained in Chapters 11 and 15.
The main drawback of the electromigration technique is the fact that it
is a single-shot experiment: Once an atomic contact has been established
there is only limited possibility to fine tune its atomic configuration, in par-
ticular coming back to a larger contact is almost impossible. After burning
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
(a) (b)
Fig. 2.12 Electromigrated MCBJ with gate on silicon substrate. (a) Working principle
and (b) electron micrograph of an electromigrated MCBJ. The substrate is doped silicon
and can be used as back-gate. Reprinted with permission from [86]. Copyright 2005
American Chemical Society.
through the wire it cannot be closed again. As described before, the con-
trol of the final part of the electromigration process is tricky because the
character of the transport changes from ohmic to wave-like. A combina-
tion of electromigration with the lithographic MCBJ technique overcomes
this problem: a thin-film MCBJ is thinned-out by electromigration to a
narrow constriction with a cross section of less than 10 nm (see Fig. 2.12).
The substrate is then bent carefully for completely breaking the wire or
arranging single-atom contacts. This last step is reversible and repeatable
for studying small contacts [87] or trapped nanoobjects [86]. Because only
the very last part of the breaking requires mechanical deformation of the
substrate it is rather fast and enables the use of more brittle substrates
such as silicon.
Fig. 2.13 Setup for the electrochemical fabrication and control of atomic contacts. For
particular choices of the control potential the atomic contact can be switched between
defined conductance values and thus a ”switching current” is recorded. Reprinted with
permission from [89]. Copyright 2003 by the American Physical Society.
Fig. 2.14 MCBJ on silicon membranes. Top: Working principle of the membrane MCBJ
(not to scale). One or several lithographic MCBJs are defined on the front side of the
membrane. A glass or graphite tip is scanned along the rear side of the membrane with
the help of micromechanically controlled scan tables. The vertical motion of the tip
controls the deformation of the membrane. The close ups at the right side illustrate the
deformation of the membrane with a graphite tip, the rupture of the nanobridge, and
give an artist’s view of the atomic arrangement of a single-atom contact. The thickness
of the membrane is in the order of 300 nm, the lateral dimension of the membrane is
typically 1 mm × 1 mm. The length of the suspended bridge is smaller than the one for
lithographic MCBJs on massive substrates. The thickness of the sacrificial layer is in the
order of 100 nm only. When reducing the lateral size of the constriction first by electro-
migration, non-suspended metal bridges can be used. Bottom: optical micrograph of a
membrane carrying two MCBJs made of gold. The tip is positioned underneath the lower
bridge where the membrane is deformed. The size of the membrane is 0.6 mm × 0.6 mm.
the scope of this book is to serve as textbook for beginners in the field of
molecular electronics, we want to sensitize the reader to this issue. The par-
ticular facts which have to be taken into account in molecular conductance
measurements are the following:
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
a
G
S D
500 nm
b
gold
electrodes counter
polyimide supports
phosphor
bronze pushing rod
Isd
100 MΩ
Vg Vb
Fig. 2.15 MCBJ with gate electrode on bulk substrate. (a) Scanning electron micro-
graph of a lithographic MCBJ with gate electrode, (b) working principle of the MCBJ,
(c) and electronic circuit for the gated MCBJ. Reprinted with permission from [96].
Copyright 2009 American Chemical Society.
The typical signal sizes which have to be resolved are of the order of
a few nanovolts for the voltage and picoamperes for the current. For par-
11 Siemens is the inverse of 1 Ω = 1 Volt/Ampere and thus the unit of the conductance
in the international system of units (SI).
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
Fig. 2.16 Schematic experimental setup for measuring the voltage dependence of the
shot noise of an atomic contact. An atomic contact (double triangle symbol), of dynamic
resistance RD , is current biased through a resistance RB . The voltage V across the
contact is measured by two low noise preamplifiers through two nominally identical
lossy lines with total resistance RL in each line and the total capacitance C introduced
by the setup across the contact. The spectrum analyzer measures the cross-correlation
spectrum of the two voltage lines. The Si (i = B, Amp1 ,Amp2 ) are the known current
noise sources associated with the bias resistor and the two amplifiers. SI represents the
signal of interest, i.e. the shot noise associated with the current through the contact. Sv1
and Sv2 represent the voltage noise sources of each line (amplifier 1 connecting leads).
Reprinted with permission from [98]. Copyright 2001 by the American Physical Society.
2.11 Exercises
2.1 Vacuum: Estimate the number of gas atoms per area impinging on a surface
at normal pressure, in high vacuum (p = 10−6 mbar), and in ultra high vacuum
(p = 10−10 mbar) during one minute. Let us assume that all incoming gas atoms
stick to the surface. How thick is the gas layer after 10 minutes?
2.2 Nanowires and atomic contacts: Let us consider a cylindrical nanowire
made of Au. Au has a lattice constant of a = 0.41 nm.
(a) Estimate the number of atoms in the cross section for a wire with diameter
10 nm, 5 nm, and 1 nm.
(b) Estimate the number of surface atoms for these wires with a length of 5
nm.
(c) Calculate the ratio between surface atoms and bulk atoms in these
nanowires.
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
Fig. 2.17 Schematic description of the experimental setup for measuring thermoelec-
tric voltage based on an STM break junction. Individual molecules (symbolized by a
hexagon) are trapped between the Au STM tip kept at ambient temperature and a heated
Au substrate kept at temperature ∆T above the ambient. When the tip approaches the
substrate, a voltage bias is applied and the current is monitored to estimate the conduc-
tance. When the conductance reaches a threshold of 0.1 G0 , the voltage bias and the
current amplifier are disconnected. A voltage amplifier is then used to measure the in-
duced thermoelectric voltage, while the tip is gradually pulled away from the substrate.
Reprinted with permission from [102]. Copyright 2008 American Chemical Society.
Chapter 3
3.1 Introduction
In this chapter we shall present the most common methods for contacting
molecules. Although we are mainly interested in single molecule devices, we
shall also introduce the most basic methods which are in use for contacting
molecular ensembles, since many interesting effects in molecular electronics
have first been observed in devices containing these assemblies. Of course,
this list can never be complete because new methods and variations of
existing ones are constantly being developed. Let us remark that we shall
focus here on methods to contact molecules with metal electrodes. Devices
including at least one semiconductor electrode have also been realized and
examples will be briefly described in section 13.7. Finally, as in the previous
chapter, we shall compare the performance of the various techniques and
indicate their most common applications.
In the fabrication of molecular junctions not only the kind of the elec-
trodes used is crucial, but also the deposition method of the molecules.
Thus, any report about electric current through molecular junctions has
to address the “protocol”, i.e. the precise contacting scheme including the
way how, the moment when, and the conditions under which the molecules
are brought into electric contact with the electrodes. For this reason, we
shall introduce in this chapter the most common deposition methods, then
we shall turn to single-molecule contacting schemes and we shall end by
addressing the ensemble techniques.
Particularly interesting are techniques which enable the fabrication of
three-terminal devices. In these systems, two of the terminals serve to inject
the current and measure the voltage, while the third one acts as a gate that
controls the electrostatic potential in the molecule. The incorporation of
45
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
this third electrode is crucial for revealing the transport mechanism and it
allows us to tune the current through a molecular junction, very much like
in the transistors fabricated with the standard semiconductor technology.
For the sake of completeness, the first part of this chapter will be devoted
to introduce the standard molecules in use in molecular electronics as well
as to describe their basic properties.
Part of the fascination of molecular electronics lies in the fact that the
molecular toolbox is almost infinite, which makes us believe that it is pos-
sible to find an appropriate molecule for any imaginable application. So
far, however, only a few classes of molecules have been explored in molec-
ular electronics. In this section we shall introduce some of these molecules
and discuss their basic properties. But before doing that, it is convenient
to recall the most common functional elements in digital electronic circuits
that molecules are supposed to mimic. The main elements and their re-
quirements are the following:
When extending the scope to cover also logic circuits one additionally
has to consider:
Finally, since most of the existing devices containing molecules are com-
posite devices in which the molecules are connected to either metal or semi-
conductor electrodes yet another function has to be realized:
H H H H
H C C H C C H C C H
H H H H
Ethene Ethyne
Ethane
(Ethylene) (Acetylene)
Fig. 3.1 Examples of hydrocarbons. Left: Ethane with C-C single bond. Middle:
Ethene with one C-C double bond. Right: Ethyne with one C-C triple bond.
resistances have to be small. Since dissipation is already one of the most se-
vere problems in nowadays semiconductor devices, signal sizes, i.e. the level
of the current should be considerably smaller than in those devices. Since
our main interest lies in exploring the fundamental properties of molecular
electronic devices, we shall not pay attention to those requirements for the
rest of this book.
From the very beginning of molecular electronics, it has been become
clear that carbon-based molecules offer the required versatility to realize
most of these desired functionalities. Carbon is the basis of a great variety
of solid structures including graphite, diamond, graphene, and molecules
like the cage-shaped fullerenes and - last but not least - the quasi one-
dimensional nanotubes.
3.2.1 Hydrocarbons
Another very rich class of carbon-based molecules is the hydrocarbons with
the possibility to tune their degree of conjugation. The electronic richness
of both classes stems from the fact that the degree of hybridization of the
molecular orbitals depends on the conformation and the environment. The
carbon atom has four valence electrons which in the case of diamond are
sp3 hybridized corresponding to a tetrahedral arrangement of the bonds in
space. This conformation is realized in the saturated hydrocarbons with
the sum formula Cn H2n+2 which are called alkanes.1 Each carbon atom has
four direct neighbors, either C or H atoms and all bonds are σ-bonds, see
Fig. 3.1. Bigger alkanes with n ≥ 4 exist in several isomers, some of which
are ring-shaped (cycloalkanes). Since all electrons are used for forming
chemical bonds they are basically localized and the alkanes are insulating.
1 The transport through alkane-based molecular junctions will be discussed in section
14.1.2.
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
In graphite the valence electrons are sp2 hybridized in the graphite plane
with an angle of 120o between the bonds. The fourth electronic orbital has
p character with its lobes pointing perpendicular to the graphite plane. The
wave functions of neighboring carbon atoms overlap and form the electronic
π-system, which in case of graphite is responsible for the in plane and
the finite plane-to-plane conductance. The same situation takes place in
the alkene hydrocarbons containing one carbon-carbon double bond, see
Fig. 3.1. Interesting for molecular electronics are polyenes with the sum
formula Cn Hn+2 , which contain more than one double bond. When these
double bonds are alternating with single bonds, the wave function of π-
system is extended over the whole molecule. These molecules are called
conjugated or aromatic molecules. The criterion of aromaticity is 4n + 2
π-electrons.
The carbons in hydrocarbons may furthermore be triply bond in sp-
hybrids forming alkynes. When alternated with single-bonds these linear
bonds are very stable and give also rise to delocalized wave functions as in
the conjugated species with double bonds.
The delocalization of the wave function is broken when the double or
triple bonds do not alternate with single bonds. Furthermore, the con-
jugation can be tuned by introducing an angle between the planes of the
individual cyclic parts. The consequences of breaking the conjugation for
the conductance of a molecular junction will be discussed in section 13.5.
In a very common representation only the bonds are shown: single bonds
as single lines, double bonds as double lines, triple bonds as triple lines. The
carbon atoms themselves are not displayed. The positions of the carbon
atoms are at the kinks between these lines. Neither the hydrogen atoms
nor the bonds to them are drawn. The number and positions of them can
be deduced by fulfilling the valence four at each carbon. As an example we
show in Fig. 3.2(a) the polyene hexatriene (consisting of six carbons and
with three double bonds) in various representations.
As for the alkanes larger species of alkenes and alkynes arrive in several
isomers. When two doubly-bond carbon atoms are surrounded by different
groups one has to distinguish between the cis conformation, in which the
neighboring groups are on the same side of the double bond, and the trans
conformation with the neighbor groups being located on opposite sides of
the double bond. A cis-trans conformation change sets the basis for a class
of molecules with in-built switching functionality.2
2 The most popular species of molecular switches are those which can be addressed opti-
cally. Many realizations are based on two ground types of switching (cis/trans conforma-
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
(a) H H H H
(b) H H
H C H H C H
C C C C
C C C C C C
C C
, C C
H C H H C H
H H H H
H H
, :
H
H
H
H
H
H H
H
Hexatriene Benzene
Fig. 3.2 Various representations of the hexatriene and the benzene molecule. (a) The
polyene hexatriene is chosen as an example for a conjugated linear hydrocarbon molecule.
(b) The benzene molecule. Top and center panel: Because of the delocalization of the
π-electrons the positions of the double bonds are not defined. Therefore, they are often
symbolized by an inner ring.
N N S O
C60
Fig. 3.5 Line representation of the bonds of the fullerene molecule C60 .
facilitates the chemical adsorption on the surface. The most common com-
bination for molecular electronics devices is thiol-terminated molecules for
adsorption on gold surfaces. The molecules organize such that they form
ordered monolayers (see Fig. 3.6). This procedure sounds simple, but in
practice many parameters have to be well controlled for obtaining repro-
ducible SAM quality. A recent review of this technique is given in Ref. [105].
Another highly developed technique is monolayer formation via the
Langmuir-Blodgett (LB) technique [106, 107]. A LB film consists of one
or more monolayers of an organic material, deposited from the surface of
a liquid onto a metal surface by immersing the solid substrate into the liq-
uid. The molecules form a monolayer on the surface of the solution. The
monolayer is transferred to the substrate when dipping it into the solu-
tion. Upon repetition of the immersion a multilayer consisting of several
monolayers and, thus films with very accurate thickness can be formed (see
Fig. 3.6). The film formation relies on the fact that amphiphilic molecules
with a hydrophilic head and a hydrophobic tail are used. These molecules
assemble vertically onto the substrate. For other molecules a horizontal
adsorption may be favored, yielding low-density films. The density and or-
dering can be enhanced by concentrating the molecular layer on the surface
of the solution with a spatula before the substrate is dipped into it.
In particular, for the preparation of samples for low-temperature mea-
surements, remainders of the solvent may hamper the formation of clean
metal-molecule-metal junctions. Therefore, alternative “dry” deposition
methods have been developed. Gaseous molecules (like e.g. hydrogen, oxy-
gen, nitrogen, carbonmonoxide, methane) can be deposited directly from
the gas phase by condensation on the cold metal electrodes. Very stable
molecules, like the fullerenes or DNA bases may be evaporated thermally
from various sources including Knudsen cells or tungsten boats, which are
Joule heated by driving a current through them. More sensitive molecules
can be deposited using electrospray ionization (ESI). The method starts
with a solution in which the molecules to be ionized are dissolved. An
electrospray of this solution is created by a strong electric field, which orig-
inates from a voltage applied between the spray needle and the end of a
capillary. Due to the strong field at the tip apex, charged droplets are
created, which are directed towards the capillary, which forms the connec-
tion to a vacuum chamber where the already prepared metal electrodes are
located [108]. With this method well-controlled submonolayer molecular
films may be deposited onto substrates in ultra-high vacuum (UHV).
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
Au-covered substrate
Molecular solution
(thiol terminated)
Adsorption
Organization
Fig. 3.6 Top: Formation of a self-assembled monolayer (SAM) shown for two species of
alkanethiols on a gold-covered substrate. The substrate is immersed into the molecular
solution. The molecules adsorb assemble with the thiol-terminated end on the substrate.
After an incubation time a self-assembled monolayer is formed. Bottom: Fabrication of a
Langmuir-Blodgett (LB) film. The left panel shows a droplet of an amphiphilic molecule
dissolved in a volatile solvent. It is spread on the water-air interface of the trough. The
solvent evaporates and leaves a diluted and disordered monolayer behind which is then
compressed with the help of a moving barrier. The right panel shows how the monolayer
is transferred onto the substrate. Reprinted with permission from Ref. [106].
VSD
Au Au
Vgate
SiO2
Si
a b
c d
e f
Fig. 3.7 Three-terminal devices and possible artifacts in molecular contacts. Top panel:
Schematic diagram of electromigration gap and measurement configuration. Bottom
panel: Six models describing possible geometries formed within the electromigration gap
by molecule(s) and contaminant metal particles. (a) Single-molecule contact as desired.
The molecule is chemisorbed with both ends at the metal electrodes. (b) Single-molecule
in the vacuum gap between the electrodes. The molecule is not chemisorbed. (c) Metal-
nanoparticle bridging the gap between the electrodes. (d) Multi-molecule contact. (e)
The molecules are coupled indirectly via a metal nanoparticle to the electrodes. (f)
The molecules are not chemisorbed to the electrodes but to a metal nanoparticle. After
Ref. [109].
6 Thephysical results obtained with these devices are discussed in particular in Chapters
15 and 16.
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
(a) (c)
(b)
Fig. 3.9 (a) The structures of three molecules studied with the dumbbell tech-
nique: 1,4-benzenedimethanethiol (BDMT), 4,4′ -biphenyldithiol (BPD) and bis-(4-
mercaptophenyl)-ether (BPE). (b) The dimer contacting scheme. (c) TEM image of
a BDMT dimer made of 10-nm colloidal gold particles. The separation between the
two particles corresponds approximately to the BDMT length (0.9 nm). Adapted with
permission from MacMillan Publishers Ltd: Nature [114], copyright 2005.
C60 this can be performed via evaporation [115, 116]. The surface is then
scanned and a suitable molecule is selected. Depending on the physical
question to study, an isolated molecule or a member of a larger aggregate
can be chosen. As described in the previous chapter, for single-atom con-
tacts, the formation of a single-molecule chemical junction is signaled by a
sudden increase of the conductance. When this is achieved, the approach
can be stopped and spectroscopic investigations can be performed. From
the electronic point of view this contacting method usually results in asym-
metric contacts, meaning that the molecule is electronically better coupled
to the substrate than to the tip. This is important for the interpretation of
the transport properties, which will be discussed in Part 4. Often the cou-
pling to the substrate is in the “strong” regime while the electrons have to
tunnel from the molecule onto the tip and vice versa, i.e. it is in the “weak
coupling” regime. Therefore, this method is most suitable for molecules
which are only loosely bound to the substrate, e.g. by a single atom or a
few atoms, like for C60 , where the binding is given through one pentagon
or one hexagon of carbon atoms.
1 2 3 4
Fig. 3.10 Schematics of the contact formation process of a molecular junction with the
STM. Four stages of the contact formation during approach (1,2) and retraction (3,4) are
shown. At (3) the chemical bond between the contact atom and the substrate is broken
and the molecular wire is formed. Reprinted with permission from [119]. Copyright 2008
IOP Publishing Ltd.
able functionalization of the STM tip, e.g. with hydrogen molecules [122].
Recently, it has been demonstrated that molecular orbitals can be even
better resolved by atomic force microscopy when the tips are terminated
with carbon monoxide (CO) molecules [120, 121].
Finally, an elegant way to contact rod-like molecules is to embed the
molecules into a matrix of less conducting molecules, such that the long axis
of the molecules is almost perpendicular to the substrate, see Fig. 3.11(a).
With the techniques described in section 3.3 a self-assembled layer of weakly
conducting molecules is prepared. A standard combination would be alka-
nes with one thiol anchor group on a gold substrate. The thiol binds chem-
ically to the gold releasing the non-thiolated ends to the top of the SAM.
The properties of the SAM are chosen such that free places or defects ex-
ist at which the study molecule can be incorporated. When scanning the
sample with an STM tip the positions of the better conducting molecules
can be located and spectroscopic measurements can be performed [123].
In a variation of this technique the study molecules are equipped with
two highly reacting anchoring groups, e.g. thiols. One end attaches to
the gold surface, the other one pointing to the top of the SAM. These
thiols can the be used as binding places for gold nanoparticles (GNPs) , see
Fig. 3.11(b). Depending on the density of the study molecule and the size
of the GNPs, one or several molecules are contacted with the same GNP. In
this way a very stable molecular junction consisting of substrate, molecule
and GNP is fabricated. The prepared sample is then investigated with
an STM [124] or a conductive AFM [125]. The tip is either brought into
strong contact with the GNP, such that the tip-GNP contact has negligible
resistance. Or the transport properties due to the presence of the GNP have
to be incorporated in modeling the transport for deducing the properties of
the molecular junction. The obvious advantage of this latter method is the
high stability of the device. Both variations share the in-built possibility
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
(a) (b)
Tip Tip
metal
nanoparticle
Substrate Substrate
Fig. 3.11 (a) Scanning tunneling microscopy (STM) study of electron transport through
a target molecule inserted into an ordered array of reference molecules. (b) STM or
conducting atomic force microscopy (AFM) measurement of conductance of a molecule
with one end attached to a substrate and the other end bound to a metal nanoparticle.
After Ref. [40].
new position on the substrate can be chosen. The method is very suitable
for gathering statistical information about the preferred conductance values
of molecular junctions. However, the stability of the molecular junctions is
usually not sufficient for spectroscopic measurements [36, 126].9
electrical wiring
glass pipette contaiinng
molecular solution
spring-borne
contact
plug hosting pipette
and contacts
bolt
Fig. 3.12 A PDMS-sealed glass pipette, in which the molecular solution circulates, is
pressed onto the central part of MCBJ chip with the help of a plug screwed to the sample
holder. The electrical contacts are realized in this case via spring-borne contacts outside
the gasket.
One main problem in single-molecule studies lies in the fact that the
electronic transport depends crucially on the exact coupling between the
molecule and the metal electrodes, i.e. on the precise atomic arrangement
of the contacts.10 As a result pronounced sample-to-sample and junction-
to-junction variations are observed. Repeated measurements are needed to
deduce the typical behavior of a given metal-molecule system. The influence
of varying contact geometry averages out in devices containing ensembles
of molecules. Furthermore, these ensembles are contacted with rigid and
robust electrodes. These devices usually provide better mechanical stability
and longer life-times allowing long-time systematic measurements and the
variation of outer control parameters like temperature or magnetic field.
However, when interpreting data recorded on ensemble devices one has
to bear in mind possible interaction effects between the molecules them-
selves which might affect their electronic properties. Furthermore, also
without interaction effects it is not straightforward to infer the single-
molecule junction behavior from the ensemble because the number of
molecules which contribute to the transport may be smaller than the total
number of molecules in the ensemble, if not all are contacted equally. For
instance, some of the molecules forming the ensemble might be in strong
coupling to the electrodes while others are only weakly coupled. As a result
the transport characteristics may show superpositions of various transport
mechanisms. Furthermore, ensemble structures are necessarily larger in
space than single-molecule devices giving limits to their maximum integra-
tion density. From the point of view of fundamental research the most
promising strategy is to compare the results from single-molecule contact
schemes with ensemble measurements for revealing the robust properties
of the given molecule-metal system. We shall restrict ourselves to methods
suitable for small ensembles ranging from roughly a few hundred molecules
to several thousand molecules. Very efficient methods have been developed
for contacting large area molecular films, which are however, out of scope
of this monograph.
3.5.1 Nanopores
One technique which produces rather small ensembles of molecules uses
pores in thin freestanding membranes. The method has been used in
10 This issue will be addressed in Chapters 13 and 14.
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
the 1980’s and 1990’s for fabricating nanometer-sized metallic contacts for
point contact spectroscopy [129]. However, no single-atom contacts can be
achieved. A single crystalline silicon wafer is covered from both sides with
a thin layer of silicon nitride with a typical thickness of 50 nm to 100 nm.
The rear side of the wafer is patterned by optical lithography with squares
of typical lateral size of 100 µm. Using first plasma etching for attacking
through the nitride, then wet-etching in hydrofluoric acid the squares are
etched through the bulk of the silicon wafer. The wet etching process is
anisotropic. It attacks particular crystal orientations of the silicon much
faster than others. As a result inclined etch walls are formed thereby re-
ducing the size of the squares. The inclined walls become covered with a
native silicon oxide layer during the following process steps. Furthermore,
the acid attacks silicon much faster than silicon nitride. The process can
thus be stopped controllably when a suspended silicon nitride membrane is
obtained. Now the membranes are patterned from the front side via elec-
tron beam lithography with a small dot in each membrane. Using plasma
etching a small pore is drilled into the membrane with a typical diameter
of 10 to 50 nm.
The formation of molecular junctions requires three further steps [130].
First, a metal electrode - usually gold - is evaporated from the top side.
The device is then immersed into the molecular solution until a SAM has
formed. After a suitable reaction time which depends on the molecule-
metal combination the sample is rinsed and dried and the second metal
electrode is deposited by evaporation onto the rear side, see Fig. 3.13. Care
has to be taken that the SAM is not destroyed by thermal impact com-
ing from the metal atoms. With this technique thermally stable molecular
ensemble junctions are obtained which are particularly suitable for studies
of the temperature dependence of the transport properties. A difficulty of
the method lies in the fact that the quality of the first deposited electrode
cannot be characterized; it might be covered with water or other contami-
nants which could hamper the formation of a high-quality SAM. A similar
objection was made concerning the second molecule-metal interface: The
molecular layer is exposed to ambient conditions before the deposition of
the second electrode.11
11 We will discuss data recorded with this sample species in Chapter 13.
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
Au
Si3N4
Si
Si SiO2
Au
Au
Si3N4
Au
Alkanethiol
(a) Au (b)
SiO2
Fig. 3.14 Production of shadow mask on silicon substrate. (a) The shadow mask is
defined via electron-beam lithography in a Si3 N4 /SiO2 double layer using two dry etching
steps. (b) The bridge in the center of the structure is used to separate two metal
contacts, which are evaporated vertically onto the substrate. A SAM is deposited on
both electrodes. In a second step metal is evaporated under an angle that allows a small
overlap between this top electrode and one of the bottom electrodes. If this overlap is
small enough, transport through single or a few molecules can be possibly measured.
Reprinted with permission from [111]. Copyright 2005, American Institute of Physics.
shown that the smoothness of the first metal layer is mandatory for avoiding
shortcuts between both electrodes. A second problem of this method is the
risk of destroying the SAM by the heat impact during the evaporation of
the top electrode or of creating metal grains [109].
Fig. 3.15 Processing steps of a large-area molecular junction. (a) Gold electrodes are
vapor-deposited on a silicon wafer and a photoresist is spin-coated. (b) Holes are pho-
tolithographically defined in the photoresist. (c) An alkane dithiolSAM is sandwiched
between a gold bottom electrode and the highly conductive polymer PEDOT:PSS as
a top electrode. (d) The junction is completed by vapor-deposition of gold through a
shadow mask, which acts as a self-aligned etching mask during reactive ion etching of the
PEDOT:PSS. The dimensions for these large-area molecular diodes range from 10 to 100
mm in diameter. Reprinted with permission from MacMillan Publishers Ltd: Nature
[131], copyright 2006.
Si
(c) (d)
SAM
Substrate
Fig. 3.16 Production of nanoscale features by nano transfer printing (nTP). (a) The
features are defined by electron beam lithography in a polymethylmethacrylate (PMMA)
double layer on a silicon substrate. The elastomeric polydimethylsiloxane (PDMS) is cast
into the structures and cured at 60o C. Fluorination of the substrate before this step
ensures easy separation of PDMS and substrate after the curing. (b) Layers of 10-30 nm
metal gold are evaporated onto the PDMS stamp. (c) Alkanedithiols form a monolayer
on a GaAs substrate. The gold on the PDMS stamp binds to this monolayer and is
transferred to the substrate. (d) The patterned gold film that forms is transferred on
top of the GaAs substrate. Good binding to the monolayer is proved by the scotch
tape test. Reprinted with permission from [111]. Copyright 2005, American Institute of
Physics.
(a)
a l C8 OPE (b)
da
w
20mm 100nm
(c)
OPE
C8
C8 º S º S S OPE
Fig. 3.17 Contacting molecular networks with gold nanoparticles. (a) Electron mi-
croscopy image of a device: two square-shaped gold contacts were evaporated on top
of a nanoparticle array line of width w. (b) Electron micrographs of the array struc-
ture before and after OPE (oligo-phenylene-ethynylene) exchange. (c) Schematic of the
molecular-exchange process. Left: self-assembled alkanethiol-capped nanoparticles be-
fore exchange. Right: During the exchange process. The OPE molecules displace part of
the alkane chains and interlink neighboring nanoparticles to form a network of molecular
junctions. Adapted with permission from [133]. Copyright Wiley-VCH Verlag GmbH &
Co. KGaA.
are covered with a spherical ligand shell. The thickness of the ligand shell
corresponds to half the length of the molecules which shall be assembled
between the GNPs later. A dense-packed, well-ordered, two-dimensional
array with an approximate size of 10 µm × 20 µm of these dressed GNPs
is deposited onto a substrate which is subsequently patterned with metallic
electrodes for performing the contacts to the measurement circuit. The ar-
ray contains approximately a million nanoparticles. The molecules forming
the ligand shell can be replaced with an exchange reaction by the molecules
to be studied electrically. By using network analysis methods the typical
properties of an individual molecular junction can be at least partially de-
duced from the behavior of the network. Besides the particular stability
and in-built ensemble averaging, this method is suitable for the investiga-
tion of very small signals, such as electrical response to optical activation
of photochromic molecules [135], see section 20.7.
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
3.6 Exercises
PART 2
Theoretical background
75
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
76
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
Chapter 4
4.1 Introduction
77
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
diffusive ballistic
Fig. 4.1 Schematic illustration of a diffusive (left) and ballistic (right) conductor.
On the basis of Ohm’s law one would expect the conductance of a metallic
wire to scale as R2 , where R is its radius. Deviations from such a scaling law
were already discussed by Maxwell [138], who studied with classical argu-
ments the conductance of a diffusive constriction, where the contact radius
is large compared to the mean free path. He found that the conductance
scales linearly with the contact radius, i.e.
G = 2Rσ. (4.2)
where σ is the conductivity.
As we shrink a conductor to well below the mean free path, the con-
ductance departs from the value expected from the previous expression. In
1965 Sharvin [139] considered the propagation of electrical current through
a ballistic contact by approximating it with a classical problem of dilute
gas flow through an orifice. He reasoned that if the potential difference be-
tween the two half-spaces is eV , the conduction electrons passing through
the orifice should change their velocity by the amount ∆v = ±eV /pF , where
pF is the Fermi momentum.1 The net current will be I = ne∆vS, where
S = πR2 is the contact area and taking into account the Fermi-Dirac statis-
tics for electrons, n = 4πp3F /(3h3 ), one gets the conductance for a circular
ballistic point-contact
¶2 ¶2
2e2 πR 2e2 kF R
µ µ
G= = , (4.3)
h λF h 2
where e is the electron charge and h is the Planck’s constant. Notice that
for ballistic contacts the conductance is proportional to the contact area,
like in Ohm’s law, but the proportionality constant 2e2 /h has a quantum
nature. An important difference between the two lies in the fact that G is
1 This is just an approximation and the exact treatment includes an integration of
the projection of ∆v along the orifice axis over the solid angle of 2π. Anyway, the
phenomenological result is only a factor 8/3 different from the exact one [140].
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
where kF is the wave vector. This equation is valid for a contact in the
form of a wire. For an orifice the numerator of the last fraction should
be 1 instead of 2. Eq. (4.4), valid for contacts down to a few nanometers
in diameter [143], is often used to establish the relationship between the
conductance and the radius of a contact.
Due to limitations of the semiclassical approach, Eq. (4.4) does not
account for purely quantum effects which dominate when the size of the
contact becomes so small that the wave nature of an electron can no longer
be ignored. Rolf Landauer [137] showed, already back in the 1950’s, that
in the latter case “conductance is transmission”, i.e. in order to determine
the total conductance one has to solve the Schrödinger equation, find the
current-carrying eigenmodes, calculate their transmission values and sum
up their contributions. Mathematically, this is summarized by in the Lan-
dauer formula
N
2e2 X
G= Tn , (4.5)
h n=1
where the summation is performed over all available conduction modes and
Tn are their individual transmissions. If the transmission of a mode is per-
fect, it contributes exactly one quantum unit of conductance, G0 = 2e2 /h ∼
(12.9 kΩ)−1 . This formula shows that by changing the size of the contact,
one can change the number of modes contributing to the conductance and
thus the conductance itself in a step-like manner (see discussion below).
This is clearly at variance with the situations described above. The deriva-
tion of the Landauer formula and the discussion of its physical implications
is the subject of the rest of the next sections.
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
ikx −ikx
1111111111111111111111111111111111111111111
0000000000000000000000000000000000000000000
e + re
0000000000000000000000000000000000000000000
1111111111111111111111111111111111111111111
0000000000000000000000000000000000000000000
1111111111111111111111111111111111111111111
0000000000000000000000000000000000000000000
1111111111111111111111111111111111111111111
0000000000000000000000000000000000000000000
1111111111111111111111111111111111111111111
0000000000000000000000000000000000000000000
1111111111111111111111111111111111111111111
teikx
0000000000000000000000000000000000000000000
1111111111111111111111111111111111111111111
0000000000000000000000000000000000000000000
1111111111111111111111111111111111111111111
0000000000000000000000000000000000000000000
1111111111111111111111111111111111111111111
0000000000000000000000000000000000000000000
1111111111111111111111111111111111111111111
0000000000000000000000000000000000000000000
1111111111111111111111111111111111111111111
0000000000000000000000000000000000000000000
1111111111111111111111111111111111111111111
0000000000000000000000000000000000000000000
1111111111111111111111111111111111111111111
2 x
T = |t|
Fig. 4.2 Wave function (plane wave) impinging on a potential barrier. The wave is
partially reflected with a probability amplitude r and partially transmitted with a prob-
ability T = |t|2 .
Now, we can convert theRsum into an integral with the usual replacement:
P
(1/L) k g(k) → 1/(2π) g(k)dk. Thus,
e
Z
JL→R = dk v(k)T (k)fL (k)[1 − fR (k)]. (4.8)
2π
We now change from the variable k to energy, E, introducing the density
of states dk/dE = (dE/dk)−1 = m/(~2 k), since E = ~2 k 2 /(2m).2 Due to
the cancellation between the group velocity and the density of states, the
left-to-right current can be written as
e
Z
JL→R = dE T (E)fL (E)[1 − fR (E)]. (4.9)
h
Analogously, we can show that the current from right to left can be
written as
e
Z
JR→L = dE T (E)fR (E)[1 − fL (E)], (4.10)
h
2 Here, we are assuming that the conduction electrons can be described by a non-
interacting electron (or Fermi) gas.
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
where we have used the fact that the transmission probability is the same,
no matter in which direction the barrier is crossed.
Now, the total current3 I(V ) = JL→R − JR→L can be simply expressed
as
2e ∞
Z
I(V ) = dE T (E)[fL (E) − fR (E)]. (4.11)
h −∞
Here, we have introduced an extra factor 2 to account for the spin degener-
acy that usually exists in the systems that we shall analyze. This expression
is the simplest version of the so-called Landauer formula and it illustrates
the close relation between current and transmission. At zero temperature
fL (E) and fR (E) are step functions, equal to 1 below EF + eV /2 and
EF − eV /2, respectively, and 0 above this energy. If we moreover assume
low voltages (linear regime), this expression reduces to I = GV , where the
conductance is G = (2e2 /h)T , where the transmission is evaluated at the
Fermi energy.
This simple calculation demonstrates that a perfect single mode conduc-
tor between two electrodes has a finite resistance, given by the universal
quantity h/2e2 ≈ 12.9 kΩ. This is an important difference with respect to
macroscopic leads, where one expects to have zero resistance for the per-
fectly conducting case. The proper interpretation of this result was first
pointed out by Imry [144], who associated the finite resistance with the
resistance arising at the interfaces between the leads and the sample.
V(x)
V0
I II III
0 L x
where
√ p
2mE 2m(V0 − E)
k1 = k3 = and k2 = . (4.13)
~ ~
Note that we have assumed that the effective mass is the same everywhere
and we have discarded the incoming term (b3 e−ik3 x ) in ψIII because we are
considering here the problem of an a wave function impinging on the barrier
from the left.
Using now the continuity of the wave function and its first derivative at
x = 0 and x = L, we arrive at the following relationships
1 1
0.8
(a) (b)
0.01
T(E)
0.6 1
0.001 0.0001
0.4 1e-06 E = 1 eV
1e-06 E = 2 eV
0.2 1e-09 E = 3 eV
0 0.2 0.4 0.6 0.8 1
0 1e-08
0 1 2 3 4 0.2 0.4 0.6 0.8 1
E/V0 L (nm)
Fig. 4.4 (a) Transmission probability vs. energy for a symmetric potential barrier of
height V0 = 4 eV and width L = 1 nm. The inset shows a blow-up of the region E < V0 .
(b) Transmission as a function of the width of the potential barrier (V0 = 4 eV) for
different values of the energy. In both cases the mass is assumed to be the electron mass.
Solving these equations, we obtain the following expression for the energy
dependence of the transmission coefficient
¯ ¯2
¯ a3 ¯ 1 4E(V0 − E)
T = ¯¯ ¯¯ = ³ 2 2 ´2 = .
a1 k1 +k2
1 + 2k sinh2
(k L) 4E(V0 − E) + V02 sinh2 (k2 L)
1 k2
2
(4.15)
Proceeding in a similar way, one can compute the transmission for E >
V0 and the result is (see Exercise 4.2)
1 4E(E − V0 )
T = ´2 = . (4.16)
4E(E − V0 ) + V02 sin2 (k2 L)
³
k12 −k22
1+ 2k1 k2 sin2 (k2 L)
The energy and length dependence of the transmission of this potential
barrier are illustrated in Fig. 4.4. The most prominent feature is maybe
the exponential dependence of the transmission on the barrier width for
energies E < V0 , see Fig. 4.4(b). According
p to Eq. (4.15), this decay is
given by T ∝ exp(−2k2 L) = exp(−2L 2m(V0 − E)/~), i.e. the slopes
in Fig. 4.4(b) are mainly determined by the square root of the difference
between the electron energy and the barrier height. Since the transmission
determines the conductance, this model provides a natural explanation for
the exponential decay of the low-bias conductance as a function of the
distance between the electrodes in all kind of tunnel barriers. It also tells
us that such decay is simply governed by the work function of the metals
involved.
Landauer formula shows that the linear conductance at low tempera-
tures is determined by the transmission at the Fermi energy. However, the
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
(a) (b) eV
eV eV
Fig. 4.5 Rectangular potential barrier under the application of a voltage: (a) approxi-
mation and (b) actual potential profile.
Metal 1 Metal 2 eV sΒ
ϕ eV
Insulator 2
Fig. 4.6 Tunneling through a junction in which two metallic electrodes are separated
by a thin insulating film, which is modeled as a rectangular potential barrier. The three
panels show the three distinct voltage ranges discussed in the text.
with the electric field strength in the insulator F = V /s, where s is the
thickness of the insulating field.
In the case of vacuum tunneling (or tunneling through an insulator), we
should be aware of the fact that whilst the electron is in the tunnel gap,
it will induce image charges in the two electrodes. This serves to modify
the barrier potential. The net effect of this is to reduce the average barrier
height and hence increase the transmission probability. For an analysis of
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
V(x)
region 1 region 2
−L/2 0 +L/2 x
Fig. 4.7 The potential V (x) under consideration varies in an arbitrary way within the
interval −L/2 ≤ x ≤ +L/2 and goes to zero outside this interval.
these “image forces” for the case of the rectangular barrier discussed here,
see Ref. [147].
It is worth mentioning that the problem of the rectangular barrier under
an applied voltage, see Fig. 4.5(b), can be solved exactly using the full Airy
functions. This was done by Grundlach [148], who showed that the current
exhibits oscillations as a function of voltage that are superimposed in the
WKB result discussed above.
In the next section we shall present a more rigorous discussion of the scat-
tering formalism, where the concept of scattering matrix plays a central
role. The definition and properties of this matrix are described in many
quantum mechanics textbooks, but for the sake of completeness, we have
included here a brief discussion of this subject.
p
where k = 2mE/~2 , while in the region x > +L/2 (region 2) it has the
form
ψk (x) = a2 e−ikx + b2 eikx , (4.24)
Here, the different coefficients depend on k, as well as on the shape of the
potential under study. Notice that with our notation, the amplitudes ai
(i = 1, 2) correspond to the incoming waves impinging on the potential
region, whereas the amplitudes bi correspond to the outgoing waves.
The scattering matrix is defined as the 2 × 2 matrix that relates the
incoming and outgoing amplitudes as follows
µ ¶ µ ¶
b1 a1
= Ŝ , (4.25)
b2 a2
where Ŝ is usually written as
r t′
µ ¶
Ŝ = . (4.26)
t r′
Here, r and r′ are reflection amplitudes and t and t′ are the transmission
amplitudes associated to this potential.
Are all these four elements independent? What are the properties of the
scattering matrix? A first property of the S-matrix can be deduced from
the conservation of the current. Let us remind that in quantum mechanics,
the current associated with a wave function ψ(x) is given by
dψ ∗
· ¸
~ dψ
J(x) = ψ ∗ (x) − ψ(x) . (4.27)
2mi dx dx
Differentiating, we find
d2 ψ d2 ψ ∗
· ¸
d ~ ∗
J(x) = ψ (x) 2 − ψ(x) 2 . (4.28)
dx 2mi dx dx
Taking into account Eq. (4.22), we obtain
d
J(x) = 0. (4.29)
dx
Therefore, the current J(x) associated with a stationary state is the same
at all points of the x-axis. Note, moreover, that Eq. (4.29) is simply the
one-dimensional analog of the relation (continuity equation)
∇ · J(r) = 0, (4.30)
which is valid for any stationary state of a particle moving in three-
dimensional space. According to Eq. (4.29), the current J(x) has the same
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
V(x)
a1 b2 b3
b1 a2 a3
potential 1 potential 2
Fig. 4.8 Combination of two potential barriers of arbitrary shape. The coefficients ai
and bi represent the different incoming and outgoing amplitudes with respect to the
potential barrier i.
This result allows us to compute now very easily, for instance, the total
transmission through the combined structure. According to the previous
equations
T1 T2
T = |t|2 = √ , (4.44)
1 − 2 R1 R2 cos θ + R1 R2
where Ti = |t(i) |2 = |t′(i) |2 , Ri = |r(i) |2 = |r′(i) |2 and θ = phase(r′(1) ) +
phase(r(2) ) is the phase shift acquired in one round-trip between the scat-
terers.
This result can be used to study a very important phenomenon for us,
namely the resonant tunneling. In a double barrier system (or in a potential
well) one can have bound states in the region between the two scattering
centers. Then, the transmission probability in this system exhibits reso-
nances at energies close to the position of those bound states. The width
of the transmission peaks depends upon the transmissivity of the barriers,
while the distance between peaks is mainly determined by the distance be-
tween the barriers. These facts can be shown with the help of Eq. (4.44),
as it is illustrated in Exercise 4.8.
L ^
aL ^
aR
R
TL sample
TR
µL µR
^ ^
bL bR
Fig. 4.9 Two-terminal scattering problem for the case of one transverse channel.
of this formalism is included in Appendix A and it will be widely used in the following
chapters.
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
channel n in the left lead, which are incident upon the sample.5 In the
same way, the creation b̂†Ln (E) and annihilation b̂Ln (E) operators describe
electrons in the outgoing states. They obey anti-commutation relations
The creation operators ↠and b̂† obey a similar relation with the Hermitian
conjugated matrix Ŝ † .
The matrix Ŝ has dimensions (NL + NR ) × (NL + NR ). Its size, as well
as the matrix elements, depends on the total energy E. It has the block
structure
µ ′¶
r̂ t̂
Ŝ = . (4.48)
t̂ r̂′
The current operator in the left lead (far from the sample) is expressed
in a standard way,
· µ ¶ ¸
∂ ∂ †
Z
~e
IˆL (z, t) = dr⊥ Ψ̂†L (r, t) Ψ̂L (r, t) − Ψ̂L (r, t) Ψ̂L (r, t) ,
2im ∂z ∂z
(4.49)
where the field operators Ψ̂ and Ψ̂† are defined as
NL (E)
χLn (r⊥ )
Z X h i
−iEt/~ ikLn z −ikLn z
Ψ̂L (r, t) = dEe âLn e + b̂ Ln e
n=1
(2π~vLn (E))1/2
(4.50)
and
NL (E)
χ∗Ln (r⊥ )
Z h i
Ψ̂†L (r, t) = dEeiEt/~ â†Ln e−ikLn z + b̂†Ln eikLn z .
X
(2π~vLn (E)) 1/2
n=1
(4.51)
Here r⊥ is the transverse coordinate(s) and z is the coordinate along the
leads (measured from left to right), χL n are the transverse wave functions,
and we have introduced the wave vector, kLn = ~−1 [2m(E − ELn )]1/2
(the summation only includes channels with real kLn ), and the velocity of
carriers vn (E) = ~kLn /m in the n-th transverse channel.
After some algebra, the expression for the current can be cast into the
form6
eX
Z h i
IˆL (t) = dEdE ′ ei(E−E )t/~ â†Ln (E)âLn (E ′ ) − b̂†Ln (E)b̂Ln (E ′ ) .
′
h n
(4.52)
Using Eq. (4.47) we can now express the current in terms of the â and â†
operators alone,
e XX
Z
IˆL (t) =
′
dEdE ′ ei(E−E )t/~ â†αm (E)Amn ′
αβ (L; E, E )âβn (E ).
′
h mn
αβ
(4.53)
Here the indices α and β label the reservoirs and may assume values L or
R. The matrix A is defined as
X †
Amn ′
αβ (L; E, E ) = δmn δαL δβL − SLα;mk (E)SLβ;kn (E ′ ), (4.54)
k
and SLα;mk (E) is the element of the scattering matrix relating b̂Lm (E) to
âαk (E). Note that Eq. (4.53) is independent of the coordinate z along the
lead.
6 Here, we have used the fact that the velocities vn (E) vary with energy quite slowly,
typically on the scale of the Fermi energy, and neglected their energy dependence.
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
Let us now derive the average current from Eq. (4.53). For a system at
thermal equilibrium the quantum statistical average of the product of an
electron creation operator and annihilation operator of a Fermi gas is
†
âαm (E)âβn (E ′ ) = δαβ δmn δ(E − E ′ )fα (E).
®
(4.55)
Using Eq. (4.53) and Eq. (4.55) and taking into account the unitarity of
the scattering matrix Ŝ, we obtain
e ∞
Z
dE Tr t̂† (E)t̂(E) [fL (E) − fR (E)] .
£ ¤
I ≡ hIL i = (4.56)
h −∞
Here the matrix t is the off-diagonal block of the scattering matrix, tmn =
SRL;mn . In the zero-temperature limit and for a small applied voltage
Eq. (4.56) gives a conductance
e2 £ † ¤
G= Tr t̂ (EF )t̂(EF ) , (4.57)
h
where EF is the Fermi energy. Eq. (4.57) establishes the relation between
the scattering matrix evaluated at the Fermi energy and the conductance.
It is a basis invariant expression. The matrix t̂† t̂ can be diagonalized;
it has a real set of eigenvalues (transmission coefficients) Tn (E) (not to be
confused with temperature), each of them assumes a value between zero and
one. The corresponding eigenfunctions will be referred to as eigenchannels
or conduction channels. In this natural basis we have instead of Eq. (4.56)
eX ∞
Z
I= dE Tn (E) [fL (E) − fR (E)] . (4.58)
h n −∞
(a) (b)
(b) ky
y
Gate voltage
kF
x
π/W
2DEG W 2DEG kx
Fig. 4.11 Point contact conductance as a function of gate voltage at 0.6 K, demonstrat-
ing the conductance quantization in units of 2e2 /h. The constriction width increases
with increasing voltage on the gate (see inset). Reprinted with permission from [153].
Copyright 1988 by the American Physical Society.
The steps are near integer multiples of 2e2 /h, after correction for a gate-
voltage-independent series resistance from the wide 2DEG regions. This
phenomenon is referred to as conductance quantization.
An elementary explanation of this effect relies on two facts: (i) the
2DEGs are ballistic systems (at least along the constriction) and the only
scattering takes place against the potential walls defined by the split gates
and (ii) the momentum of the electron is quantized in the transverse direc-
tion giving rise to 1D subbands. Since every subband that contributes to
the transport (or conduction channel) has a perfect transparency and the
number of them is obviously an integer, it follows from the two-terminal
Landauer formula that the low temperature conductance G is quantized,
G = (2e2 /h)N, (4.60)
as observed experimentally. Here, N is the total number of open conduc-
tion channels and the prefactor 2 accounts for the spin degeneracy. This
number can be simply calculated assuming a square-well lateral confining
potential of width W . In the constriction, the electron momentum along
the transport direction (x-direction) can take any value, while the trans-
verse momentum ky is quantized and can only take the following values:
ky = ±nπ/W with n = 1, 2, ..., N , see Fig. 4.10(b). Since the current is only
carried by those electrons at the Fermi energy (or with momentum equal
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
n(E) n(E)
1 Tn 1
Tn
E E
0 0
EF EF+eV EF EF+eV
Fig. 4.12 In a quantum point contact with bias voltage, V , the transmission probability,
Tn , determines the distribution function, n(E), of a transmitted state as a function of its
energy, E. In the right reservoir, states with energy lower than the Fermi energy are all
occupied, while right-moving states with higher energy can only be coming from the left
reservoir, and therefore their average occupation is equal to the transmission probability,
Tn . This argument applies to every individual conduction channel.
∆n2 = n2 − n2 = Tn (1 − Tn ), (4.61)
where in the last step we used the fact that n2 = n, since for fermions n
is either zero or one. Hence, the fluctuations in the current are suppressed
both for Tn = 1 and for Tn = 0. According to Eq. (4.61) the fluctuations
will be maximal when the electrons have a probability of one half to be
transmitted. The shot noise is thus a non-linear function of the transmission
coefficients, as we anticipated above.
We shall now derive in a rigorous manner the main results concerning
shot noise in a two-terminal device within the scattering formalism. For
this purpose, we follow again Ref. [150]. Since are concerned with the
fluctuations of the current away from its average value, we then introduce
ˆ ≡ I(t)
the operators ∆I(t) ˆ − hIi, where Iˆ is the current operator evaluated
in a given reservoir, let us say, the left one. We define the correlation
function P (t − t′ ) of the current in a given contact as
1D ˆ ˆ ′ ) + ∆I(t
ˆ ′ )∆I(t)
ˆ
E
P (t − t′ ) ≡ ∆I(t)∆I(t . (4.62)
2
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
= δαδ δβγ δkn δml δ(E1 − E4 )δ(E2 − E3 )fα (E1 ) [1 − fβ (E2 )] . (4.64)
Here fα (E) is the corresponding Fermi distribution. Now, making use of
the current operator of Eq. (4.53) and of the expectation value of Eq. (4.64),
we arrive at the following expression for the noise power
e2 X X
Z
P (ω) = dEAmn nm
γδ (L; E, E + ~ω)Aδγ (L; E + ~ω, E)
h mnγδ
× {fγ (E) [1 − fδ (E + ~ω)] + [1 − fγ (E)] fδ (E + ~ω)} . (4.65)
Note that with respect to frequency, it has the symmetry properties P (ω) =
P (−ω). In the rest of this discussion, we shall only be interested in the
zero-frequency noise.7 For the noise power at ω = 0 we obtain
e2 X X
Z
P ≡ P (0) = dEAmn nm
γδ (L; E, E)Aδγ (L; E, E) (4.66)
h mnγδ
× {fγ (E) [1 − fδ (E)] + [1 − fγ (E)] fδ (E)} .
Eq. (4.66) can now be used to predict the low frequency noise properties of
arbitrary multi-channel phase-coherent conductors. But before presenting
the general result, let us first discuss two limiting cases of special interest:
Equilibrium noise. If the system is in thermal equilibrium at tem-
perature T , the distribution functions in both reservoirs coincide and are
equal to f (E). Using the property f (1 − f ) = −kB T ∂f /∂E and employing
the unitarity of the scattering matrix, one can arrive at the following result
P = 4kB T G, (4.67)
7 Zero-frequency noise actually means that the frequency is small in comparison with
the relevant frequency scales of the problem, but large enough to neglect the 1/f noise
that is present in almost any system.
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
The Fano factor assumes values between 0 (all channels are transparent)
and 1 (Poissonian noise). In particular, for one channel it becomes (1 − T ).
The general result for arbitrary temperature and voltage for the noise
power of the current fluctuations in a two-terminal conductor is
2e2 X ∞
Z
P = dE { Tn (E) [fL (1 − fL ) + fR (1 − fR )] +
h n −∞
o
2
Tn (E) [1 − Tn (E)] (fL − fR ) . (4.74)
Here the first two terms are the equilibrium noise contributions, and the
third term is the non-equilibrium or shot noise contribution to the power
spectrum. Note that this term is second order in the distribution function.
At high energies, in the range where the Fermi distribution function is well
approximated by a Maxwell-Boltzmann distribution, it is negligible com-
pared to the equilibrium noise described by the first two terms. According
to Eq. (4.74) the shot noise term enhances the noise power compared to
the equilibrium noise.
In the practically important case, when the scale of the energy depen-
dence of transmission coefficients Tn (E) is much larger than both the tem-
perature and applied voltage, these quantities in Eq. (4.74) may be replaced
by their values taken at the Fermi energy. We obtain then
" #
2e2
µ ¶X
X eV
P = 2kB T Tn2 + eV coth Tn (1 − Tn ) , (4.75)
h n
2kB T n
where V is again the voltage applied between the left and right reservoirs.
The full noise is a complicated function of temperature and applied voltage
rather than a simple superposition of equilibrium and shot noise. For low
voltages eV ≪ kB T one recovers the result of pure thermal noise, i.e. P =
4kB T G. Eq. (4.75) is the starting point for the analysis of experimental
results on noise in atomic and molecular junctions, see section 19.1.
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
The scattering formalism has been very successful explaining many basic
transport phenomena in a great variety of nanostructures. It has also been
extended to other situations of interest for the purpose of this book, such
as e.g. photon-assisted transport [165]. For space reasons we have to end
here our discussion of this formalism, and for more details we recommend
the the reviews of Refs. [150, 151, 166–168], the didactic book of S. Datta
[50] and the book on mesoscopic physics of Y. Imry [169].
In spite of its great success, the scattering approach is far from being a
complete theory of quantum transport. In this sense, it is important to be
aware of its limitations. Among them we want to emphasize two of special
interest for the scope of this book:
(i) The scattering approach gives no hints on how to compute the trans-
mission or, more generally, the scattering matrix. In particular, it does not
tell us how to determine the actual transmission of an atomic contact or
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
a molecular circuit. In this sense, one might think that this formalism
has merely replaced a problem by another. This would be, of course, un-
fair. The scattering approach can be combined with simple models, as we
showed in section 4.4, or with more sophisticated techniques like random
matrix theory [170] to predict the transport properties of a great variety
of systems such as diffusive wires, chaotic cavities, superconducting nanos-
tructures, resonant tunneling systems, tunnel junctions, etc.
(ii) The scattering picture is an one-electron theory which is valid only
as long as inelastic scattering processes can be neglected. In this formalism
one assumes that the electron propagation is a fully quantum coherent
process over the entire sample. According to normal Fermi-liquid theory,
such a description would be strictly valid at zero temperature and only
for electrons at the Fermi energy. At finite bias the coherent propagation
may be limited by inelastic scattering processes due to electron-phonon
and electron-electron collisions. The theoretical description of transport
in situations where inelastic interactions play an important role requires
more sophisticated methods like the Green’s function techniques that will
be described in the next chapters.
Let us mention that there is a phenomenological way of describing the
effect of inelastic or phase-breaking mechanisms within the scattering ap-
proach, which is due to Büttiker [171]. In this description the inelastic scat-
tering events are simulated by the addition of voltage probes distributed
over the sample. The chemical potential on these probes is fixed by im-
posing the condition of no net current flow through them. Thus, although
the presence of the probes does not change the total current through the
sample, they introduce a randomization of the phase which tends to destroy
phase coherence. The current in such a structure will contain a coherent
component, corresponding to those electrons which go directly from one
lead to the other, and an inelastic component, corresponding to those elec-
trons which enter into at least one of the voltage probes in their travel
between the leads.
4.10 Exercises
p p
where k1 = 2mE/~2 , k2 = 2m(E − V0 )/~2 and m is the electron mass.
V(x)
V0
0 x
V(x)
Va
I II
III
E
a b x
V0
4k1 k3 k22
T (E) = ,
k22 (k1 + k3 )2 cos (k2 L) + (k22
2 + k1 k3 )2 sin2 (k2 L)
V(x)
III II
V3 I
V1
V2
a b x
~2 ∂ 2
H=− + V0 δ(x),
2m ∂x2
where V0 is the strength of the delta potential that acts at x = 0.
(a) Demonstrate that the boundary conditions for the scattering states ψk (x),
k being the electron momentum, are: (i) continuity at x = 0 and (ii) ψk′ (x =
0+ ) − ψk′ (x = 0− ) = (2mV0 /~)ψ(x = 0), where the prime symbol indicates
derivative with respect to x.
(b) Use the previous result to show that the transmission probability through
this delta potential can be expressed as: T = 1/(1 + Z 2 ), where Z ≡ mV0 /(~2 k).
4.7 Scattering matrix:
(a) Show that in the presence of a magnetic field the scattering matrix fulfills
the property of Eq. (4.40).
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
where ǫ0 is the position of the resonance and ΓL,R are the scattering rates as-
sociated to the left and right potential barriers. Find an expression for these
rates in terms of the transmissions of the barriers. Hints: (i) The resonances are
well separated when the transmissions T1 and T2 are small (R1 , R2 ≈ 1). (ii)
The round-trip phase shift that appears in Eq. (4.44) is θ = 2kd, where k is the
electron momentum in the region between the two barriers.
4.9 Conductance quantization in a 2DEG: One of the most successful ap-
plications of the Landauer formula is the explanation of the conductance quan-
tization that takes place in split-gate constrictions (or quantum-point contacts)
in a two-dimensional electron gas (2DEG). A useful model to study the occur-
rence of conductance steps is the so-called saddle point model used by Büttiker in
Ref. [172]. In this model it is assumed that near the bottleneck of the constriction
the electrostatic potential can be expressed as
1 1
V (x, y) = V0 − mωx2 x2 + mωy2 y 2 . (4.91)
2 2
Here, V0 is the electrostatic potential at the saddle, ωx characterizes the curvature
of the potential barrier in the constriction and ωy the lateral confinement. Show
that for this potential the transmission probabilities are given by
1
Tn (E) = .
exp[π(E − V0 − (n + 1/2)ωx )/ωy ] + 1
Using this expression in combination with the Landauer formula, find the criteria
for the observation of well-defined conductance steps at low temperatures.
4.10 Shot noise and thermopower in a quantum point contact: Use the
saddle point model of the previous exercise to study the shot noise [155] and the
thermopower [161, 163] in a quantum point contact as a function of the Fermi
energy (or gate voltage).
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
Chapter 5
The discussion of the scattering formalism in the previous chapter has left
two basic questions open: (i) How to calculate the elastic transmission
of real systems such as atomic and molecular junctions? and (ii) how to
generalize Landauer formula to take into account correlation effects and
inelastic mechanisms? Indeed, both questions can be answered, at least
to a large extent, with the help of Green’s function techniques. For this
reason, we initiate here a series of three chapters devoted to this subject.
We are aware of the fact that at this point part of the readership will be
certainly tempted to jump to the next part of the book. The words Green’s
functions cause in many people an immediate rejection because they asso-
ciate them to some obscure theoretical techniques reserved to specialists.
We believe that this judgment is a bit unfair. The degree of difficulty of
the Green’s function techniques depends primarily on the type of prob-
lems addressed. Thus for instance, we shall show that what is required to
answer the first question posed above reduces to a standard problem of lin-
ear algebra that should be accessible to any student with a background in
quantum mechanics. The answer to the second question requires however
more elaborate methods, which will also be presented in this book. With
this distinction in mind, we shall guide you through the next three chapters
indicating the type of problems that we have in mind and we shall warn
you about the possible difficulties.
In our discussion on the Green’s function techniques we shall start in
this chapter by introducing the subject concentrating ourselves on the case
of electronic systems in equilibrium. This chapter is meant to give a first
insight into what Green’s functions in quantum mechanics are, what kind of
physical information they contain and how they can be calculated in some
simple situations. Having in mind the first question above, we shall focus
111
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
on the analysis of noninteracting systems. Then, the next chapter will deal
with the diagrammatic theory, which provides a systematic perturbative
approach to compute the Green’s functions of many-body systems where
correlations and inelastic mechanisms in general play a fundamental role.
Finally, since our final goal is the analysis of the transport properties of
atomic-scale junctions, we shall present in Chapter 7 the Keldysh formalism
that allows us to compute the Green’s functions of nonequilibrium systems.
Then, at the end of that chapter, we shall apply this formalism to the
calculation of the transmission in some illustrative examples.
This chapter is organized as follows. First, we shall remind the reader
of the basics of the Schrödinger and Heisenberg representations of quantum
mechanics. Then, we shall introduce the retarded and advanced Green’s
functions in energy space for a noninteracting electron system and show how
they can be computed in certain simple examples. We shall then define
the general (valid also for interacting systems) time-dependent retarded,
advanced and causal Green’s functions and analyze their main analytical
properties, their relation with the observables of interest and how they can
be computed, in principle, with the so-called equation-of-motion method.
One last comment before we get started. We shall constantly make
use of the second quantization formalism in our discussion of the Green’s
functions techniques. So, if you are not very familiar with this formalism,
we strongly recommend you to read Appendix A.
are defined as the inverse of differential operators. One can indeed pro-
ceed in a similar way with the Schrödinger equation, which is a second
order differential equation. As an illustration, let us consider the prob-
lem of an electron in an one-dimensional system, which is described by the
Schrödinger equation
H(x)Ψ(x) = EΨ(x). (5.7)
Now, we define the electron Green’s function (or propagator) as
[E − H(x)] G(x, x′ ) = δ(x − x′ ), (5.8)
where
1 ∂2
H(x) = − + V (x), (5.9)
2m ∂x2
V (x) being an external potential acting on the electron. Notice that the
Green’s function is a complex function that depends both on the spatial
coordinates and on the energy, E.
In the case of a free electron, V (x) = V0 = constant, the Green’s function
can be obtained exactly (see Exercise 5.2). Indeed, one can show that a
solution is given by
i ′
G(x − x′ , E) = − eik|x−x | , (5.10)
v
p
where k = 2m(E − V0 ), v = k/m and we have included the energy,
E, as an argument. As it will become clear later on, one can interpret the
Green’s function as the propagation amplitude of an electron. In this sense,
the previous expression corresponds to the propagation of a free electron
at energy E from the position x′ to the right (x − x′ > 0) or to the left
(x − x′ < 0).
It is important to notice that there is another solution that corresponds
to the time-reserved solution as compared with the previous one:
i −ik|x−x′ |
G(x − x′ , E) = e . (5.11)
v
This simply reflects the fact that the Green’s function is not completely de-
termined until we specify the boundary conditions for its differential equa-
tion.
Eq. (5.10) corresponds to the so-called retarded Green’s function, Gr ,
while Eq. (5.11) corresponds to the advanced Green’s function, Ga . Al-
though the time does not appear explicitly in these functions, we shall show
later that one can relate Gr [Eq. (5.10)] with the propagation of an electron
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
where from now on the limit limη→0 is implicitly assumed in all the ex-
pressions in which the parameter η appears. Are you able to show the
equivalence of Eqs. (5.13) and (5.14)? If not, see hints in Exercise 5.3.
Eq. (5.14) shows that the Green’s functions (for a noninteracting case)
have poles precisely at the eigenenergies, ǫn , of the system. This is the first
important piece of information contained in these functions.
From the previous equations, one can deduce a number of important
properties of the functions Gr,a . Let us discuss the most useful ones for our
purposes:
Property 1. The imaginary part of the Green’s functions is related to
the density of states of the system. To demonstrate this, let us remind that
the local density of states in a given position r can be written in terms of
the eigenstates of H as follows
X
ρ(r, E) = |hr|ψn i|2 δ(E − ǫn ). (5.15)
n
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
This is a consequence of the fact that the energy integral of ρi (E) is equal
to 1, i.e.
Z ∞
1 ∞
Z
dE ρi (E) = ∓ dE Im {Gr,a
ii (E)} = 1. (5.23)
−∞ π −∞
Property 5. As one can easily see from Eq. (5.13), the following simple
relation between Gr and Ga holds:
†
Gr (E) = [Ga (E)] . (5.24)
This means in practice that we only need to compute one of these two type
of functions.
Property 6. As a last issue, let us consider the case in which the
Hamiltonian H can be written as
H = H0 + V, (5.25)
In this section we shall apply what we have learned so far to the computa-
tion of the Green’s functions of several simple electronic systems described
in terms of tight-binding Hamiltonians.1 Such Hamiltonians, as we shall
see in the next chapters, play a fundamental role in the field of molecular
electronics. A generic tight-binding Hamiltonian adopts the following form
in the language of second quantization (see Appendix A)
X †
tij c†iσ cjσ .
X
H= ǫi ciσ ciσ + (5.29)
iσ i6=j;σ
Here, the indexes i and j run over of the sites (atoms) of the system and
σ represents the electron spin (σ =↑, ↓). The different operators have the
following meaning. For instance, c†iσ is the operator that creates an elec-
tron in the site i with spin σ, while ciσ annihilates such an electron. For
the sake of simplicity, we shall assume in this discussion that there is a
single relevant orbital per site. The parameters ǫi are the on-site energies,
while the hoppings tij describe the coupling between the different sites (see
Appendix A for a precise definition of all these parameters).
Our goal is the calculation of the different Green’s functions Gr,a
ij (E) in
this local basis representation. In principle, we have three methods at our
disposal: (i) the definition of Eq. (5.13), (ii) the spectral representation of
Eq. (5.14) and (iii) Dyson’s equation, see Eq. (5.28). We shall illustrate the
use of these different approaches with the analysis of three basic examples
that will be frequently used in subsequent chapters.
detail in Chapter 9. Here, we shall use the term tight-binding to refer to models or
Hamiltonians where the electronic structure is described in terms a local (atomic-like)
basis. We shall not discuss here how the matrix elements of such a Hamiltonian are
actually computed, and we shall just use them as parameters.
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
Gr,a
11 (E) (since the problem has spin degeneracy, we omit the spin indexes
in the Green’s functions). For symmetry reasons, this Green’s function is
equal to Gr,a
22 (E). In order to compute this function, we shall employ the
three methods mentioned above:
Method 1: Direct definition. According to the definition of Eq. (5.13),
the matrix Green’s function can be simply calculated by inverting the
Hamiltonian of Eq. (5.30). In the basis of the atomic states localized in
the hydrogen atoms, {|1i, |2i}, this Hamiltonian adopts the following ma-
trix form
µ ¶
ǫ0 t
H= , (5.31)
t ǫ0
and therefore the matrix Green’s function is given by
µ r,a ¶−1
E − ǫ0 −t
Gr,a (E) = , (5.32)
−t E r,a − ǫ0
where E r,a ≡ E ± iη, η being the infinitesimal imaginary part of the energy
appearing in the definition of Eq. (5.13). Thus, the element (1, 1) that we
are looking for reads
E r,a − ǫ0 1/2 1/2
Gr,a
11 (E) = = r,a + . (5.33)
(E r,a − ǫ0 )2 − t2 E − (ǫ0 + t) E r,a − (ǫ0 − t)
One can show that this expression fulfills the different properties of
a Green’s function discussed in the previous section. Thus for instance,
notice that Eq. (5.33) has precisely the form of the spectral representation
of Eq. (5.14). The poles in this case are nothing else but the energies ǫ± =
ǫ0 ± t of the bonding and antibonding orbitals of the hydrogen molecule,2
2 The hooping t is indeed a negative quantity and thus ǫ+ = ǫ0 + t corresponds to the
lowest energy level (bonding state).
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
see Fig. 5.1. Notice also that the sum of the weights (coefficients appearing
in the numerators) is equal to 1.
On the other hand, the density of states projected onto the site 1 is
given in this case by
1 1 1
ρ1 (E) = ∓ Im {Gr,a11 (E)} = δ(E − ǫ+ ) + δ(E − ǫ− ), (5.34)
π 2 2
i.e. it is a sum of delta functions evaluated at the molecular energies. This
is a consequence of the fact that we are dealing with a finite system. In a
similar way, one could demonstrate that the rest of the properties listed at
the end of the previous section are satisfied. In particular, properties 4 and
5 are rather obvious from Eq. (5.33).
Method 2: Spectral representation. Let us now use the spectral repre-
sentation of Eq. (5.14). To evaluate this expression we need both the eigen-
functions and the eigenvalues of the hydrogen molecule. For this purpose
we just need to diagonalize the Hamiltonian of Eq. (5.31). The eigenfunc-
tions are simply the bonding
√ (|ψ+ i) and antibonding (|ψ− i) states given
by: |ψ± i = (|1i ± |2i)/ 2 with the corresponding eigenvalues ǫ± . Thus,
the function Gr,a11 (E) is then given by
X h1|ψn ihψn |1i X |h1|ψn i|2
Gr,a
11 (E) = h1|G|1i = r,a
= . (5.35)
n=+,−
E − ǫn n=+,−
E r,a − ǫn
√
Using the fact that h1|ψ± i = 1/ 2, we arrive immediately at the expres-
sion of Eq. (5.33). Obviously, this method is not very practical in general
since it requires the knowledge of the eigenfunctions of the system, which
are typically unknown.
Method 3: Dyson’s equation. Now, our starting point is Eq. (5.28).
The first thing to do is to divide the Hamiltonian of Eq. (5.30) into the
unperturbed part H0 and the perturbation V. The natural choice is that
the perturbation be the coupling term between the two atoms (second term
in Eq. (5.30)). Thus, these two parts of the Hamiltonian adopt the following
matrix form
µ ¶ µ ¶
ǫ0 0 0t
H0 = ; V= . (5.36)
0 ǫ0 t0
To solve Dyson’s equation we also need the Green’s functions of the
unperturbed system, gr,a . These functions are simply given by
µ r,a ¶−1
r,a r,a −1 E − ǫ0 0 1
g = [E 1 − H0 ] = r,a = r,a 1. (5.37)
0 E − ǫ0 E − ǫ0
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
This equation can now be trivially inverted and using the explicit expression
of the unperturbed Green’s functions one arrives once more at the result of
Eq. (5.33).
We can use the discussion above to illustrate the concept of self-energy,
which was briefly mentioned at the end of the last section. In the previous
equation, we can identify the following energy-dependent function
This function describes how the properties of the atom 1 are modified via
the interaction with the second atom. This can be better seen by rewriting
Eq. (5.40) as
1
Gr,a
11 (E) = , (5.42)
E r,a − ǫ0 − Σr,a
11 (E)
(a) t t t t
........
1 2 3 4 5
1
(b)
0.5
a
t Re{G 11
(E)}
-0.5 a
t Im{G 11(E)}
-1
-3 -2 -1 0 1 2 3
(E - ε0)/t
Fig. 5.2 (a) Semi-infinite linear chain with a single orbital per site and only nearest-
neighbor couplings. (b) Real and imaginary parts of the advanced surface Green’s func-
tion, Ga
11 , of the semi-infinite chain as a function of the energy, see Eq. (5.46).
g11 (E) = 1/(E − ǫ0 ). On the other hand, the unperturbed function g22
is nothing else but the surface Green’s function of a semi-infinite chain,3
which is precisely what we are looking for, i.e. g22 = G11 . This allows us
to obtain the following closed equation for G11 (E)
(E − ǫ0 )G11 (E) = 1 + t2 G211 (E). (5.45)
This is a quadratic equation that possesses two possible solutions. In order
to choose the “physical” one, it is necessary to take into account the bound-
ary condition E → E r,a = E ± iη to distinguish between the retarded and
advanced solutions. As a practical advice, remember that the imaginary
part of these functions has a well-defined sign. The final solution adopts
the following expression
sµ
r,a r,a − ǫ
¶2
1 E − ǫ0 E 0
Gr,a
11 (E) =
− − 1 . (5.46)
t 2t 2t
The real and imaginary parts of the advanced function are depicted in
Fig. 5.2(b). Notice that the imaginary part, and therefore the density of
states, is only non-zero in the region |E − ǫ0 | < 2|t|, which defines the
3 The removal of an atom from the chain does not modify the fact that the remaining
chain is again a semi-infinite chain.
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
energy band of the linear chain. In this region, the Green’s function adopts
the following form
s
µ ¶2
1 E − ǫ0 E − ǫ0
Gr,a
11 (E) =
∓i 1− . (5.47)
t 2t 2t
0110 0110
(a) (b) 1010 1010
0000
1111 100 1010
1111
0000 0000
1111 ΓL 1 1010 ε0 1010 ΓR
0000
1111
0000
1111 0000
1111
0000
1111
t t 1010 1010
0000 L
1111 R0000
1111 1010 Γ
0000
1111
L
0000
1111 0000
1111
R
0000
1111
1010
0000 ε
1111 0000
1111 L 1010 1010 R
0000
1111
0000
1111 0 0000
1111
0000
1111 1010 1010
0000
1111
0000
1111 0000
1111
0000
1111 10 10
Fig. 5.3 (a) A single level of energy ǫ0 is coupled to two infinite electrodes via the hop-
pings tL and tR . (b) The corresponding energy scheme where one can see the continuum
of states in the electrodes filled up to the Fermi energy and the resonant level, which
has acquired a half width at half maximum equal to Γ = ΓL + ΓR due to the coupling
to the reservoirs.
uncoupled subsystems, i.e. the right hand side of the first line of Eq. (5.52).
Thus, the perturbation V is the term that describes the coupling between
the localized level and the electrodes (second line in Eq. (5.52)). Notice
that we are assuming that there is no direct coupling between the leads.
With this choice in mind, we take the element (0, 0) in Eq. (5.28) to
obtain
G00 (E) = g00 (E) + g00 (E)V0L GL0 (E) + g00 (E)V0R GR0 (E), (5.53)
where V0L = tL and V0R = tR and g00 (E) = 1/(E − ǫ0 ) is the unperturbed
Green’s function of the single-level system. As usual, to close this equation,
we have to determine the functions GL0 and GR0 . This can be done by
taking the corresponding elements in Dyson’s equation, i.e.
GL0 (E) = gLL (E)VL0 G00 (E)
GR0 (E) = gRR (E)VR0 G00 (E),
where VL/R0 = tL/R and gLL and gRR are the Green’s functions of the two
outermost sites of the left and right electrodes, respectively. Substituting
these expressions in Eq. (5.53), we obtain the following closed equation
G00 (E) = g00 (E) + g00 (E)V0L gLL (E)VL0 G00 (E) (5.54)
+ g00 (E)V0R gRR (E)VR0 G00 (E).
In this expression one can identify, as in the previous examples, the self-
energy Σ00 (E) = t2L gLL (E) + t2R gRR (E), which in this case is the sum of
two contributions associated to the two leads. In terms of the self-energy
we can express the function G00 (E) as
1
G00 (E) = , (5.55)
E − ǫ0 − Σ00 (E)
where we have used the expression of g00 (E). Here, we see once more that
the self-energy describes how the resonant level is modified by the inter-
action with the leads. In particular, its real part is responsible for the
renormalization of the level position, which becomes ǫ̃0 = ǫ0 + Re{Σ00 (E)},
while its imaginary part describes the finite energy “width” acquired by
the level via the interaction with the leads. This latter point becomes
more clear by using the following approximation. Let us assume that the
Green’s functions of the leads are imaginary for energies in the vicinity
of ǫ0 and that they do not depend significantly on energy in this region.4
r,a
Thus, we can approximate these functions by gLL,RR ≈ ∓i/WL,R , where
4 This approximation is usually known as wide-band approximation.
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
WL,R are energy scales related to the density of states of the leads at the
energy ǫ0 .5 For instance, if we modeled the electrodes by the semi-infinite
chains like in the previous example, WL,R would then be the bulk hop-
ping element of these chains. Within this approximation, the self-energy
becomes Σr,a 00 = ∓i (ΓL + ΓR ), where we have defined the scattering rates
ΓL,R ≡ t2L,R /WL,R . Obviously, with this approximation the level position
remains unchanged (see Exercise 5.9). Finally, the function G00 (E) adopts
in this case the form
1
Gr,a
00 (E) = , (5.56)
E r,a − ǫ0 ± i (ΓL + ΓR )
Thus, the local density of states that we wanted to calculate is given by
1 1 ΓL + ΓR
ρ0 (E) = ∓ Im {Gr,a 00 (E)} = , (5.57)
π π (E − ǫ0 )2 + (ΓL + ΓR )2
which is a Lorentzian function, where Γ = ΓL + ΓR is the half-width at
half-maximum (HWHM). This result shows clearly that the resonant level,
which originally had zero width (it was an eigenstate of the isolated central
system), acquires a finite width Γ via the coupling to the leads. This fact
is illustrated in Figs. 5.3(b). It is worth stressing that the width depends
both on the strength of the coupling to the electrodes (via t2L,R ) and on
the local electronic structure of the leads (via WL,R or, more generally, via
gLL,RR ). The time scale ~/Γ that can be interpreted as the finite lifetime of
the resonant level due to the interaction with the leads, or in other words,
as the time that an electron spends in the resonant level.
Thus, the take-home message of this example is that when an isolated
molecule (or an atom) is coupled to a continuum of states, its levels are,
in general, shifted and they acquire a width that depends on the strength of
the coupling and on the local electronic structure of the leads.
Let us finally say that we hope that the reader has realized that all the
calculations of this section involved simple algebraic manipulations. Indeed,
we shall show in the next chapters that, as long as we deal with systems
with only elastic interactions (described by mean-field Hamiltonians), the
evaluation of the Green’s functions, both in equilibrium and out of equilib-
rium, reduces to straightforward exercises of linear algebra. So maybe, this
Green’s function stuff is not so scary after all, don’t you think?
For more detailed discussion of Green’s functions in the framework of
tight-binding models, we recommend the book of Ref. [181], as well as the
exercises 5-9 at the end of this chapter.
5 This energy scales are simply given by WL,R = 1/[πρL,R (E = ǫ0 )], where ρL,R are
the local densities of states of the two outermost sites of the leads.
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
where |ΨH i = |ΨS (0)i is the wave function of the ground state of the system
(that can include interactions) and the operators are in Heisenberg picture.
We shall only include explicitly the spin index σ in Grij in those problems
where the spin symmetry is broken. In this definition, the step function, θ,
ensures that t > t′ and the symbol { , } stands for the anticommutator.
The Green’s functions are often defined using the basis {|ri} formed by
the eigenfunctions of the position operator. The corresponding creation and
annihilation operators in this representation are known as field operators
and they are denoted by Ψ†σ (r) and Ψσ (r), These operators are simply
related to c†iσ and ciσ by the basis transformation
φ∗i (r)c†iσ ,
X X
Ψσ (r) = φi (r)ciσ and Ψ†σ (r) = (5.59)
i i
where φi (r) are the basis wave functions of the discrete representation.
These field operators satisfy the standard type of anticommutation rela-
tions, i.e.
{Ψσ (r), Ψ†σ′ (r′ )} = δ(r − r′ )δσ,σ′ ; etc. (5.60)
In terms of these field operators, the retarded Green’s function is defined
as
Gr (rt, r′ t′ ) = −iθ(t − t′ )hΨH | Ψσ (r, t), Ψ†σ (r′ , t′ ) |ΨH i,
© ª
(5.61)
which is a complex function that depends on two spatial arguments and
two time arguments.
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
The advanced Green’s function has a similar definition, the only differ-
ence being that the propagation takes place backwards in time
n o
Gaij (t, t′ ) = iθ(t′ − t)hΨH | ciσ (t), c†jσ (t′ ) |ΨH i. (5.62)
Finally, it is convenient to define an additional Green’s function, namely
the one known as causal Green’s function, which is defined as follows
h i
Gcij (t, t′ ) = −ihΨH |T ciσ (t)c†jσ (t′ ) |ΨH i, (5.63)
where T is the time-ordering operator. It acts on a product of time-
dependent operators by ordering them chronologically from right to left.
Thus for instance, the previous function has the following explicit form
(
c ′ −ihΨH |ciσ (t)c†jσ (t′ )|ΨH i t > t′
Gij (t, t ) = (5.64)
ihΨH |c†jσ (t′ )ciσ (t)|ΨH i t′ > t.
Notice the sign change for t′ > t due to the anticommutation of fermion
operators.
So far, our discussion in this section has been a bit mathematical and
there are questions that arise naturally. The first one is: What is the
physical meaning of the Green’s functions? To answer this question no-
tice that these functions contain factors like hΨH |ciσ (t)c†jσ (t′ )|ΨH i. Here,
c†jσ (t′ )|ΨH i describes the creation (or injection) in the ground state of an
electron at time t′ in the state j. Then, the previous expectation value
yields the probability amplitude of finding such an electron at a later time
t in the state i. In other words, the Green’s functions simply describe the
probability amplitude of the occurrence of certain processes. The type of
processes described depends on the arguments of these functions. Thus for
instance, they can describe the propagation of electrons in time domain or
in energy space, propagation in real space, in momentum space or simply
in an atomic lattice.6
Another natural question is: What is the relation between this definition
of the Green’s functions and the one put forward in the previous section?
At a first glance, it seems that there is no relation at all. However, we
shall show below that if the system is noninteracting, the Fourier transform
with respect to the time arguments of these new Green’s functions fulfill
Eqs. (5.13) and (5.14), i.e. these two type of functions are equivalent.
Simple example: degenerate electron gas. To illustrate the previ-
ous definitions, we consider now the example of a free electron gas at zero
6 Inthis sense, it is not surprising that the elastic transmission of any real system can
be naturally expressed in terms of these functions.
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
Notice first that these functions depend on the difference of the time ar-
guments, which is a general property for equilibrium systems. Notice also
that they are diagonal in k-space. Having in mind the physical meaning of
the Green’s functions, it is easy to understand why they have such a simple
time dependence. Since we are injecting electrons in a state |kσi, which
is an eigenstate of the system, the probability of finding it at a later time
in such state must be equal to one. This is precisely what the previous
expressions illustrate.
It is instructive to make contact with the results of the previous section.
For this purpose we must now Fourier transform the previous functions with
respect to the time difference, i.e.
Z ∞
′
Gr,a,c (k, E) = dt Gr,a,c (k, t)e−iE(t−t ) . (5.68)
−∞
In the course of doing the Fourier transformations, one gets the impression
that the time integrals diverge. This can be cured by introducing a small
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
N +1
ihΨN +1
P
We now insert m |Ψm m | in the part for t > 0 and
N −1 N −1
< 0, where |ΨN +1
i and |ΨN −1
P
m |Ψ m ihΨm | in the part for t m m i are
the eigenfunctions of the system with one more and one less electrons, re-
spectively. The resulting expression reads
N +1
+1 † −E0N )t
X
Gcij (t) = −iθ(t) hΨN
0 |ciσ |Ψm
N +1
ihΨN
m |cjσ |ΨN
0 ie
−i(Em
m
X † N N −1
+iθ(−t) hΨN N −1
0 |cjσ |Ψm ihΨN
m
−1
|ciσ |ΨN
0 ie
−i(E0 −Em )t
.
m
part of electrons, we can add and subtract the energy of the ground state
with N + 1 electrons:
N +1
E − (Em − E0N ) = E − (E0N +1 − E0N ) − (Em
N +1
− E0N +1 ). (5.77)
The energy difference E0N +1− E0N in the limit N → ∞ is the chemical po-
tential µ of the system, N +1
while Em −E0N +1 is the energy of the excited state
of the system with N + 1 electrons. Repeating the same operations for the
hole part, one can finally write the Green’s functions in the thermodynamic
limit as (we only consider diagonal elements)
X |hΨN +1 |c† |ΨN i|2 X |hΨN −1 |ciσ |ΨN i|2
m 0 m 0
Gcii (E) = iσ
N +1
+ N −1
(5.78)
m
E − µ − ǫm + iη m
E − µ + ǫm − iη
X |hΨN +1 |c† |ΨN i|2 X |hΨN −1 |ciσ |ΨN i|2
Gr,a
ii (E) =
m iσ
N +1
0
+ m
N −1
0
, (5.79)
m
E − µ − ǫm ± iη m
E − µ + ǫm ± iη
where ǫNm
+1
= Em N +1
− E0N +1 and ǫN
m
−1
= EmN −1
− E0N −1 are the excitation
energies of the system with N + 1 and N − 1 electrons, respectively.
From the previous expressions one can show that the spectral represen-
tation reduces to Eq. (5.14) in the noninteracting case (this exercise is left
to the reader). This is one way to establish the connection between the
definitions introduced in this section and those of section 5.2.
From the general spectral representation, it is possible to derive the
following important properties of the exact Green’s functions of an arbitrary
electronic system, which are practically identical to those of section 5.2:
Property 1. It is possible to define a spectral density related to the
imaginary part of the Green’s functions as (we only write the diagonal
elements)
+1 †
X
ρi (E) = |hΨN
m |ciσ |ΨN 2
0 i| δ(E − µ − ǫm
N +1
)+ (5.80)
m
X
|hΨN
m
−1
|ciσ |ΨN 2 N −1
0 i| δ(E − µ + ǫm ).
m
In a case in which i stands for a site index in a tight-binding problem,
the previous expression represents the quasiparticle density of states of the
system projected onto that site. The relation of the previous function to the
imaginary part of the Green’s functions is obvious. Comparing Eq. (5.80)
with Eqs. (5.78) and (5.79), one obtains
1
ρi (E) = ∓ Im {Gr,aii (E)} (5.81)
π
1
ρi (E) = −sgn(E − µ) Im {Gcii (E)} . (5.82)
π
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
by means of
X
n(r) = −i Gcσ (rt, rt+ ), (5.90)
σ
where we have used the fact that the derivative of the step function is a
δ-function.
Now, in order to compute the time derivative of the annihilation oper-
ator appearing in the previous equation, we make use of the equation of
motion for operators in the Heisenberg picture, see Eq. (5.6). Thus,
∂ X
i ciσ = [ciσ , H] = i tik ckσ , (5.104)
∂t
k
where we have used Eq. (5.101) to obtain the last result. Substituting this
expression in Eq. (5.103), we arrive at
∂ X
i Grij,σ (t) = δ(t)δij + tik Grkj,σ (t). (5.105)
∂t
k
It is now convenient to Fourier transform to energy space to convert this
differential equation into an algebraic one. Thus, introducing
Z ∞ Z ∞
1 1
Grij,σ (t) = dE e−iEt Grij,σ (E) ; δ(t) = dE e−iEt (5.106)
2π −∞ 2π −∞
in Eq. (5.105), we obtain the following algebraic equation of the Green’s
function in energy space
X
EGrij,σ (E) = δij + tik Grkj,σ (E). (5.107)
k
This is nothing else but the element (i, j) of the matrix equation
−1
Gr (E) = [E1 − H] , (5.108)
which is precisely the expression that we used as a definition in section 5.2
[see Eq. (5.13)]. Thus, we have shown again the equivalence of the two type
of definitions for the case of noninteracting electron systems.
It is important to emphasize that the equation-of-motion method
illustrated above is by no means restricted to noninteracting system.
However, if the Hamiltonian contains two-electron terms (with four cre-
ation/annihilation operators), in general there is no straightforward way to
get a closed system of equations, as in the previous example. The problem
is that the equation of motion for the one-particle Green’s function couples
this function to higher-order ones containing an increasing number of oper-
ators and the resulting algebraic system has, strictly speaking, an infinite
dimension. In practice, one has to find an appropriate way of truncating
the system, which is not an easy task in general.
In order to illustrate what we meant in the previous paragraph, let us
consider the Anderson model that describes the interaction of a single level
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
where hn0σ i is the occupation of the level ǫ0 for spin σ, which in turn has
to be calculated with the full Green’s function of Eq. (5.112). Thus, in this
limit the Green’s functions exhibit poles at energies equal to ǫ0 and ǫ0 + U .
This tells us in particular that U is the energy that one has to supply to
accommodate a second electron in the level. The expression of Eq. (5.112)
can be used as an starting point to analyze the so-called Coulomb blockade
in quantum dots or molecular transistors (see Exercise 8.9).
Let us conclude this section by recommending Chapter 9 of Ref. [185]
for a more detailed discussion about the equation-of-motion method.
5.5 Exercises
Show that the time evolution of the operators c†kσ and ckσ in Heisenberg
picture is given by
Obtain the temporal evolution of the operators c1σ and c2σ in Heisenberg
picture.
5.2 Green’s function of a free electron in 1D: Let us consider the
Schrödinger equation of a free electron in a 1D potential
1 ∂2
» –
− + V 0 Ψ(x) = EΨ(x),
2m ∂x2
where V0 is a spatially constant potential. Show that the electron Green’s function
is given by the expressions detailed in section 5.2.
5.3 Equivalence of expressions (5.13) and (5.14): Show the equivalence
of Eq. (5.13) and Eq. (5.14). Hints: (i) Multiply bothPsides of Eq. (5.13) by
[(E ± iη) − H]. (ii) Introduce then the closure relation n |ψn ihψn | = 1, where
ψn are the eigenfunctions of H. (iii) Use H|ψn i = ǫn |ψn i and (iv) multiply by
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
the inverse of the operator on the left hand side of the Green’s function to obtain
Eq. (5.14).
5.4 Dyson’s equation: Starting from Eq. (5.26), show that the Green’s func-
tions fulfill the Dyson’s equation (5.28).
5.5 Semi-infinite tight-binding chain: Let us consider the Hamiltonian of
Eq. (5.43) for a semi-infinite chain. Calculate the off-diagonal retarded Green’s
functions Grn1 of the chain (where 1 is the first site and n an arbitrary one) and
demonstrate that it is given by the following expression for |E − ǫ0 | < 2|t|:
e−inφ
Grn1 (E) = where cos φ = (E − ǫ0 )/2t.
t
i −iφ|i−j|
Gaij (E) = e cos φ for |E − ǫ0 | < 2|t|.
|t|
(b) An infinite chain can be viewed as two coupled semi-infinite chains. In this
sense, consider the coupling between the semi-infinite chains as a perturbation
and use Dyson’s equation to obtain the diagonal advanced Green’s functions in a
site of the chain and demonstrate that it coincides with the result derived in (a)
for i = j.
5.7 Tight-binding chain with a defect: Let us consider an infinite chain as
in the previous problem in which a diagonal perturbation is introduced in one
of the sites, let us say in site i, such that its on-site energy becomes ǫ0 + ∆.
Calculate the local density of states in the site i and, in particular, investigate
the possibility of having a localized state outside the band. Study also the spatial
extension of such a state by calculating the occupation of this state in different
sites away from the one in which the defect is located.
5.8 Finite tight-binding chain: Let us consider a finite chain with N sites
and only nearest-neighbor interactions. Calculate the advanced Green’s function
Gan1 (E), where 1 refers to the atom in one of the extremes of the chain and n to
an arbitrary site. Demonstrate in particular that for |E − ǫ0 | < 2|t|
1 sin[(N − n + 1)φ]
Gan1 (E) = .
t sin[(N + 1)φ]
Chapter 6
In the previous chapter we have seen that the calculation of the zero-
temperature Green’s functions of a non-interacting system in equilibrium
reduces to solve an algebraic linear system, summarized in Dyson’s equa-
tion. This is practically all we need to tackle the problem of the determi-
nation of the elastic transmission of realistic systems. However, if we want
to go beyond and treat systems where the electron correlations or inelastic
interactions play a major role, we need many-body techniques. For this
reason, we present in this chapter a systematic perturbative approach for
the calculation of zero-temperature equilibrium Green’s functions.1 This
formalism is valid for any type of system and interaction and constitutes
the most general method for the computation of Green’s functions. More-
over, the nonequilibrium formalism introduced in the next chapter follows
closely the perturbative approach that we are about to describe.
The perturbative (or diagrammatic) approach is nicely explained in dif-
ferent many-body textbooks (see e.g. Refs. [173–175, 182–185]) and for this
reason, our description here will be rather brief.2 This approach is concep-
tually rather simple, but it contains several technical points that usually
make it rather obscure. In the spirit of this monograph, we shall avoid
very formal discussions and we shall provide instead simple plausibility ar-
guments or we shall simply refer the reader to the adequate literature.
Before the trees do not let us see the forest, let us give a brief overview
of what we are about to see. First, we shall learn how to write down a
perturbative series for the Green’s functions, i.e. how to express systemat-
ically the corrections to the Green’s function due to a perturbation such
1 In some sense, this approach is simply a generalization of the perturbation theory for
the wave functions that one studies in elementary courses of quantum mechanics.
2 This chapter is mainly based on Chapter 3 of Ref. [173]
143
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
Our goal now is the calculation of a generic causal Green’s function, which
in a discrete basis is given by
h i
hΨH |T ciσ (t)c†jσ (t′ ) |ΨH i
Gij (t, t′ ) = . (6.23)
hΨH |ΨH i
Here, the expectation value is evaluated in the ground state of the system
described by the Hamiltonian of Eq. (6.1) and the operators are written in
Heisenberg picture. Notice that we omit the superindex c to abbreviate the
notation and we introduce the denominator for normalization reasons that
will become clear later on.
As explained in the previous section, it is convenient to use the interac-
tion picture. We first transform the operators:
h i
(0) (0)†
hΨH |T S(0, t)ciσ (t)S(t, t′ )cjσ (t′ )S(t′ , 0) |ΨH i
Gij (t, t′ ) = . (6.24)
hΨH |ΨH i
Here, we have used the superindex (0) to emphasize that the operators in
the interaction picture correspond to Heisenberg operators of the unper-
turbed system. We now transform the wave function by using
|ΨH i = S(0, t)|ΨI (t)i, (6.25)
where t is an arbitrary time. Now, we want to relate the state |ΨI (t)i with
the unperturbed ground state (for V = 0), |φ0 i. This can be done using
the so-called adiabatic hypothesis. In this hypothesis, one assumes that
if the perturbation is switched on at an initial time, let us say t = −∞,
and grows slowly to its actual value at t = 0, the physics is not modified.
This adiabatic switch on is achieved by replacing the perturbation V by
Ve−ǫ|t| , where ǫ is an infinitesimally small positive parameter. In this
way, at t = ±∞ the perturbation vanishes and the system tends to the
unperturbed ground state
|ΨH i = S(0, −∞)|φ0 i. (6.26)
This procedure is not completely well-defined and one can show that
during the evolution of the ground state from t = −∞ to t = 0 with the
operator S, the wave function acquires a phase that diverges as ǫ tends to
zero. These phase factors are finally canceled by the terms in the denomi-
nator of the expectation value. The rigorous statement of this fact is known
as the Gell-Mann and Low theorem and for more information we refer the
reader to the book of Fetter and Walecka [173].
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
We now make use of Eq. (6.26) to write the causal Green’s function as
follows
h i
(0) (0)†
hφ0 |S(∞, 0)T S(0, t)ciσ (t)S(t, t′ )cjσ (t′ )S(t′ , 0) S(0, −∞)|φ0 i
Gij (t, t′ ) = .
hφ0 |S(∞, −∞)|φ0 i
(6.27)
Here, we have used the time symmetry of the problem that implies in par-
ticular that the ground state wave function is recovered at t = +∞ (apart
from a phase factor). On the other hand, it is obvious that in the previous
expression we can introduce the time-evolution operators appearing next to
the wave functions inside the time-ordered products. Thus, the expectation
value now reads
h i
(0) (0)†
hφ0 |T ciσ (t)cjσ (t′ )S(∞, −∞) |φ0 i
Gij (t, t′ ) = , (6.28)
hφ0 |S(∞, −∞)|φ0 i
where we have grouped all the pieces of the operator S since the operator
T ensures the proper ordering. Now, we use the expansion of Eq. (6.22) for
the operator S to write the expectation value as a perturbative expansion
"∞
1 X (−i)n Z ∞
′
Gij (t, t ) = dt1 ... dtn × (6.29)
hφ0 |S(∞, −∞)|φ0 i n=0 n! −∞
h i i
(0) (0)†
hφ0 |T ciσ (t)cjσ (t′ )V(0) (t1 ) · · · V(0) (tn ) |φ0 i ,
With the perturbative formalism that we have developed so far, the problem
of calculating a Green’s function or any expectation value of an operator
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
Notice that in the previous factorization one could have had additional
terms containing expectation values like for instance
h i h i
(0) (0) (0)† (0)†
hφ0 |T ciσ (t)ckσ (t1 ) |φ0 i, hφ0 |T cjσ (t′ )clσ (t2 ) |φ0 i
h i
(0) (0)†
or hφ0 |T ciσ (t)cj,σ̄ (t′ ) |φ0 i.
However, they usually vanish for different reasons. In the first two cases,
the combinations of operators do not conserve the number of electrons.
The third expectation value vanishes, unless the ground state is magnetic.
Thus, husually the onlyi terms that survive are those with the following form:
(0) (0)†
hφ0 |T ciσ (t)cjσ (t′ ) |φ0 i, i.e. those with a combination of a creation and
an annihilation operator. As a convention, we shall always place the cre-
ation operator on the right hand side in these factors.
To end this section, notice that the basic factor appearing in the de-
composition that results from Wick’s theorem is closely related to a single-
particle Green’s function of the unperturbed system
h i
(0) (0)† (0)
hφ0 |T ciσ (t)cjσ (t′ ) |φ0 i = iGijσ (t, t′ ). (6.36)
Thus for instance, the expectation value of the previous example can be
written as
h i
(0) (0)† (0) (0)†
hφ0 |T ciσ (t)cjσ (t′ )ckσ (t1 )clσ (t2 ) |φ0 i = (6.37)
(0) (0) (0) (0)
−Gijσ (t, t′ )Gklσ (t1 , t2 ) + Gilσ (t, t2 )Gkjσ (t1 , t′ ).
that appear in the perturbation theory. Thus for instance, the unper-
turbed causal Green’s functions, which appear in the perturbative expan-
sion through the application of Wick’s theorem, will be represented by a
solid line. This is shown in Fig. 6.1(a) for the function G(0) (rt, r′ t′ ) in real
space. For this case, the arrow points from the second set of arguments
(or event) to the first one (indicating the propagation of an electron from
r′ t′ to rt). If the problem depends explicitly on the spin, we would have to
label the different events with the corresponding spin. If we use a discrete
basis, the corresponding line will look like in Fig. 6.1(b).
Fig. 6.1 Basic elements of Feynman diagrams. (a) Propagator line between the events
r′ t′ to rt. (b) Propagator line between the states jσ ′ and iσ. (c) Full propagator line.
(d) Interaction line between the events r′ t′ to rt. (e) Interaction line for an external
potential.
The full (or dressed) Green’s function that corresponds to the total
amplitude for the electron propagation will be represented as a double
line, as shown in Fig. 6.1(c). On the other hand, the electron-electron
interaction between two events will be represented by a wavy line, as in
Fig. 6.1(d). Notice that, in general, the interaction is instantaneous and
therefore U (rt, r′ t′ ) ∝ δ(t − t′ ). In the case in which the perturbation is an
external potential, V (r), this will then be represented by a dashed line, see
Fig. 6.1(e).
The structure of perturbative series and the corresponding Feynman
diagrams depends on the type of perturbation under study. In what follows,
we shall illustrate the diagrammatic approach with the analysis of two
examples where the perturbation is (i) the electron-electron interaction and
(ii) an external static potential.
x x x
(1) (2) (3)
x’1 x’1
x1 x1
x1
x’1
x’ x’ x’
x (4) x (5) x (6)
x1
x’1
x1 x’1 x1 x’1
x’ x’ x’
causal Green’s function. This can be seen in Fig. 6.2, where we have num-
bered the terms from 1 to 6 following the order of Eq. (6.41).
Let us summarize some of the main features of these diagrams, which
are also found in higher-order contributions:
• The only thing that matters in the diagrams is their topology, i.e.
the way in which the different events are connected.
• The Green’s functions with equal time arguments are represented
by a closed loop and their value is equal to in(0) (r). If we used a
local representation {|ii}, then we would have
(0) (0)
Gii (t, t+ ) = ihniσ i. (6.44)
• Notice that all the intermediate events are linked by an interaction
line and they have an incoming and an outgoing propagator, which
correspond to the scattering process that the electron undergoes
due to the electron-electron interaction. These intermediate events
are known as vertexes (see Fig. 6.3).
• In Fig. 6.2 there are diagrams that have parts that are not con-
nected to the the rest of the diagram and, in particular, to the
initial and final events. Since there is an integration over the in-
termediate arguments appearing in these disconnected parts, they
4 We assume here that there is spin symmetry in the unperturbed problem. Thus, all
the Green’s functions are diagonal in spin space and we will not write explicitly their
spin index to abbreviate the notation.
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
Fig. 6.3 Vertex: point where two propagator lines and an interaction line meet.
Fig. 6.4 Some of the 10 second-order topologically distinct connected Feynman dia-
grams for the electron-electron interaction.
equal to
Z Z Z Z
2
−i ′
dx1 dx1 dx2 dx′2 G(0) (x, x1 )U (x1 , x′1 )G(0) (x′1 , x′2 )G(0) (x′2 , x′1 )
x
+ x + + .......
x
Fig. 6.5 Diagrammatic series for the propagator in the case of an external potential.
For the sake of simplicity, we have assumed that the potential does
not depend on the electron spin. In this case, the diagrammatic series
is very simple. Applying Wick’s theorem to Eq. (6.30), one obtains the
diagrammatic series shown in Fig. 6.5. This means that in the propagation
of the electron from the initial instance to the final one, one simply has a
series of sequential scattering events with the external potential. The rules
for computing the contribution to the nth-order correction of the causal
Green’s functions are very simple in this case:
(1) Draw the sequential diagrams like in Fig. 6.5 with n + 1 propagators
and n interaction lines.
(2) Associate the corresponding Green’s function to every propagator line.
(3) Assign the corresponding external potential to every interaction line.
(4) Integrate over the intermediate variables.
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
In spite of all the simplifications that we have introduced in the last section,
it is still very difficult to compute the different terms of the perturbative
series. This is due to the presence of the integrals over the intermediate
arguments. Thus for instance, a diagram of order 1 for the electron-electron
interaction contains up to six integrals.
The problem can be simplified by noticing first that in an equilibrium
situation the Green’s functions depend exclusively on the difference of the
time arguments. Thus, we can Fourier transform with respect to time and
work in the energy space. The introduction of the Fourier transformation
modifies the Feynman diagrams and we now study how this occurs in detail.
On the other hand, if the system is spatially homogeneous, the prob-
lem can be simplified even further since then the Green’s functions de-
pend only on the difference of the space coordinates. We shall first discuss
this case and later on, we shall generalize the results to an arbitrary non-
homogeneous system.
As we have just said, if the system is spatially homogeneous and in
equilibrium, the Green’s functions satisfy
transforms read
dp ipx dp ipx
Z Z
G(x) = e G(p); U (x) = e U (p), (6.51)
(2π)4 (2π)4
where dp ≡ d3 kdE is the volume element in (k, E)-space.
In order to illustrate how the diagrams are modified in energy space, we
choose a first-order diagram for the electron-electron interaction, namely
diagram 2 in Fig. 6.2. The contribution of this diagram, which we shall
denote as D(x − x′ ), is given by
Z Z
D(x − x′ ) = i dx1 dx′1 G(0) (x − x1 )U (x1 − x′1 ) (6.52)
dp ip(x−x′ )dq
Z Z
D(x − x′ ) = i e U (q)G(0) (p)G(0) (p − q)G(0) (p).
(2π)4 (2π)4
(6.56)
This implies that the Fourier transform of the diagram can be written as
dq
Z
D(p) = i U (q)G(0) (p)G(0) (p − q)G(0) (p). (6.57)
(2π)4
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
p−q
q
The previous derivation would be similar for any diagram. The key idea
is that the energy and the momentum are conserved in every vertex. Thus,
one can view the diagrams as flow diagrams in which the propagator lines
and the interaction lines carry momentum and energy. The momentum k
and the energy E carried by the initial propagator are also carried by the
final one, due to the conservation of momentum and energy in every vertex
of the diagram. This is illustrated in Fig. 6.7 with two first-order diagrams
and a second-order one. Notice that, since the interaction lines carry both
momentum and energy, one has to assign to them a direction, which is
indicated by an arrow in the diagram.
kE kE q E’’
kE
q E’
k−q k’+q
k’E’
E−E’’ E’+E’’
kE k−q,E−E’
kE q E’’
kE
(1) Draw all the topologically distinct connected diagrams with n interac-
tion lines and 2n + 1 propagator lines. These diagrams are the same as
in the ones in (r, t)-space.
(2) Assign the flow direction (arrows) of the momentum and energy to
every interaction and propagator line.
(3) The momentum and the energy must be conserved in every vertex.
(4) Every propagator with momentum k and energy E contributes with a
factor that is equal to the unperturbed causal Green’s function, which
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
0,σ
E E’’ _
0,σ
0,σ
E−E’’ E’ E’+E’’
0,σ _
E E’’ 0,σ
0,σ
Fig. 6.8 Second-order Feynman diagrams in energy space for the Anderson model.
(a)
X
= + X + + .......
X
(b)
= + + + .......
Fig. 6.9 Diagrammatic expansion for the propagator for (a) an external potential and
(b) the electron-electron interaction.
of Fig. 6.9 can be summarized in the following equation in real space (r-
representation)
Z Z
G(x, x′ ) = G(0) (x, x′ ) + dx1 dx2 G(0) (x, x1 )ΣI (x1 , x2 )G(0) (x2 , x′ ).
(6.60)
The equation in momentum-energy space (for a homogeneous case) reads
as follows
G(k, E) = G(0) (k, E) + G(0) (k, E)ΣI (k, E)G(0) (k, E). (6.61)
In the case of a localized basis (like in a tight-binding model), the previous
equation adopts the form:
(0)
X (0) (0)
Gij (E) = Gij (E) + Gik (E)ΣI,kl (E)Glj (E). (6.62)
kl
To avoid explicit reference to any particular representation or space, we
shall write the previous equation in matrix form:
G = G(0) + G(0) ΣI G(0) , (6.63)
111
000
000
111 000
111
000
111
000
111 000
111
000
111
000
111
000
111 000
111
000
111
000
111
000
111
000
111 111
000
000
111
000
111
000
111 000
111
ΣΙ
000
111
000
111 000
111
000
111
000
111
000
111 000
111
000
111
000
111
000
111
000
111
000
111
000
111
= 111
000
111
000
111
000
000
111
000
111
000
111
000
111 000
111
111
000
000
111
000
111
000
111 000
111
000
111
000
111
000
111 000
111
000
111
000
111
(a) X
111
000
000
111
000
111
000
111
000
111
000
111
000
111
000
111
X
000
111
000
111
000
111
000
111
000
111
000
111
000
111
000
111 = X + + X + .......
000
111
000
111
000
111
000
111
000
111
000
111
000
111
000
111
X
X
111
000
000
111
(b)
000
111
000
111
000
111
000
111
000
111
000
111
000
111
000
111
000
111
000
111
000
111
000
111
000
111
000
111
000
111
000
111
= + + + .......
000
111
000
111
000
111
000
111
000
111
000
111
000
111
000
111
Fig. 6.11 Diagrammatic expansion for the self-energy insertion. (a) External potential.
(b) Electron-electron interaction.
where the internal integrals and sums are implicitly assumed. It is possi-
ble to write this equation in a more convenient way by inspection of the
perturbative series of G or ΣI . Let us illustrate this fact first with the
example of an external potential. As we explained in previous sections, the
diagrammatic expansion has in this case the form of a geometrical series
where the diagram of order n is simply the repetition of n identical pieces.
If we define in this case the proper self-energy, Σ, as the part of the diagram
that includes only a single scattering process, which in this case is simply
the external potential, we have the following identity
X
000
111
000
111
111
000
000
111
000
111
000
111
000
111
000
111
X X
000
111
000
111
000
111
000
111
000
111
000
111
000
111
000
111
000
111
000
111
= X + + X + ....... =
000
111
000
111
000
111
000
111
000
111
000
111
X
X
Fig. 6.12 Relation between the self-energy insertion, ΣI and the proper self-energy, Σ.
The proper self-energy, or from now on just self-energy, does not con-
tain repetitions of the same process, but only one scattering event. Then,
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
(a)
(b)
Fig. 6.13 (a) Examples of irreducible self-energy diagrams for the electron-electron
interaction. (b) Reducible diagrams.
= + Σ
(a) Σ= X
(b) Σ= + + + .......
Fig. 6.15 Diagrammatic expansion for the proper self-energy. (a) External potential
and (b) electron-electron interaction.
• The Dyson’s equation relates directly the self-energy with the full
Green’s function. Therefore, the analytical properties of Σ(E) can
be derived from those of G(E).
• One can interpret Eq. (6.69) as a definition of Σ(E) in terms of G(E).
Thus, it is also possible to define a retarded and advanced self-energy.
• From Lehmann’s representation of the Green’s functions, one can de-
duce the following properties that we state here without any proof:
Im {Σrii (E)} ≤ 0 ; Im {Σaii (E)} ≥ 0 (6.71)
Im {Σcii (E)} ≥ 0 if E < µ ; Im {Σcii (E)} ≤ 0, if E > µ.
• ImΣii (E) and ReΣii (E) are related through a Hilbert transformation:
dE ′ Im {Σr,a
ii (E )}
′
Z
r,a
Re {Σii (E)} = ∓P (6.72)
π E − E′
dE ′ Im {Σcii (E ′ )} sgn(E ′ − µ)
Z
Re {Σcii (E)} = −P .
π E − E′
Apart from the Dyson’s equation, there exist other ways to include certain
diagrams in the expansion of the self-energy up to infinite order. By inspec-
tion of the set of diagrams that contribute to the self-energy, it is possible
to distinguish two types of diagrams. On the one hand, there are diagrams,
like the one shown in Fig. 6.16, in which in one of the propagators there
is a self-energy insertion. On the other hand, there exist diagrams that do
not contain insertions and they are called skeleton diagrams. An example
of a second-order skeleton diagram is shown in Fig. 6.15(b).
Analyzing the diagrammatic series of the self-energy, one realizes that if
we consider any skeleton diagram, there appear diagrams at higher orders
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
Fig. 6.16 Example of diagram with a self-energy insertion in one of the propagators.
with the same structure (or skeleton), but with all possible self-energy in-
sertions in their propagators. This fact makes possible to sum up to infinite
order all the diagrams that share the same skeleton, which leads to effective
diagrams like the one depicted in Fig. 6.17. Here, we have taken into ac-
count the fact that by adding all the diagrams with the same structure, the
propagator in the skeleton diagram can be replaced by the full (dressed)
propagator.
Σ= + + + ......
Σ HF = +
which is nothing else but the Hartree potential, where nσ (r) is the perturbed
electron density with spin σ that has to be determined self-consistently.
Analogously, the second diagram in Fig. 6.19 is given by (in the repre-
sentation |ri)
ΣX ′ ′ ′ +
σ (r, r ) = iU (r − r )Gσ (rt, r t ). (6.74)
One can show that this expression leads to the known nonlocal (Fock)
exchange potential. For this purpose, one just needs to expand the field
operators in the previous expression in terms of an arbitrary single-electron
basis and take into account that the ground state is noninteracting. This
leads to
X e2 φiσ (r′ )φiσ (r)
ΣX ′
σ (r, r ) = − . (6.75)
i
|r − r′ |
As an additional illustration of the Hartree-Fock approximation, we
discuss now the calculation of the energy bands in this approximation of a
homogeneous electron gas (see Exercise 6.5). In this case, it is not neces-
sary to do the self-consistency because it is automatically guaranteed due
homogeneity of the system with a constant density n = N/V . Instead of
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
using the expressions derived above, we compute now the self-energy in this
approximation in the (k, E)-space. Evaluating the Hartree-Fock diagrams
in this space, one arrives at
X Z dk′ Z dE ′ ′
ΣHσ = U (q = 0) 3
Gσ′ (k′ , E ′ )eiE η . (6.76)
(2π) 2π
′ σ
4πe2
ΣH
σ = n. (6.77)
µ2
Although this result diverges when µ → 0, it is exactly canceled in the jel-
lium model by the potential created by the uniform background of positive
charge. Thus, the only remaining contribution is the exchange one that can
be expressed as
dq dν dk′ 4πe2
Z Z Z
ΣXσ (k) = i 3
U (q)Gσ (k−q, E −ν) = − hnk′ σ i.
(2π) 2π (2π)3 |k − k′ |
(6.78)
Now using the Dyson’s equation in this representation, G(k, E) =
−1
[E − ǫk − Σ(k, E)] , we see that the energy bands in the Hartree-Fock
approximation are given by ǫk,HF = ǫk + ΣX (k). The explicit expression of
the dispersion relation is computed in Exercise 6.5.
The goal of this section is two-fold. On the one hand, we shall use the
Anderson model, already discussed in section 5.4.3 and Appendix A, to
illustrate the perturbative approach described in this chapter. On the other
hand, we shall use this model to get a flavor of the Kondo effect. This is
a many-body phenomenon which can appear in molecular junctions and it
will be described in much more detail in Chapter 15.
The Anderson model describes the interaction of a localized level with
electron-electron interaction with the continuum of states of a metallic sys-
tem. It was introduced by Anderson to describe a magnetic impurity in
a metal host, but it can also be used to describe a metal-molecule-metal
junction, which is the problem that we are interested in. In this model, the
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
function projected onto the level can written in terms of the self-energy as
1
Gr00,σ (E) = . (6.80)
E − ǫ0 + iΓ − Σr00,σ (E)
Taking now into account that the density of states in the level is given
by ρ0σ (E) = −(1/π)ImGr00,σ (E), the corresponding occupation can be ex-
pressed as
Z µ
1 µ 1
Z
hn0σ i = dE ρ0σ (E) = − dE . (6.81)
−∞ π −∞ E − ǫ0 + iΓ − Σr00,σ (E)
We can now use the relation
1 ∂
ln E − ǫ0 + iΓ − Σr00,σ (E) +
£ ¤
r =
E − ǫ0 + iΓ − Σ00,σ (E) ∂E
∂Σr00,σ (E)/∂E
(6.82)
E − ǫ0 + iΓ − Σr00,σ (E)
together with the Ward identity (see Exercise 6.6)
∂Σr00,σ (E)
Z µ
dE Gr00,σ (E) = 0, (6.83)
−∞ ∂E
to write the occupation as
Z µ
1 ∂
ln E − ǫ0 + iΓ − Σr00,σ (E) .
£ ¤
hn0σ i = − Im dE (6.84)
π −∞ ∂E
Integrating this expression we arrive at
ǫ0 − µ − ReΣr00,σ (µ)
· ¸
1 1
hn0σ i = − tan−1 . (6.85)
2 π Γ
Here, we have used the fact that in a Fermi liquid ImΣr00,σ (µ) = 0, which
physically means that the quasiparticles have an infinite lifetime at the
Fermi energy.
Thus, we can write the local density of states as
1 Γ + ImΣr00,σ (E)
ρ0σ (E) = £ ¤ . (6.86)
π E − ǫ0 − ReΣr (µ) 2 + Γ + ImΣr (E) 2
¤ £
00,σ 00,σ
Using Eq. (6.85), we can relate the exact density of states at the Fermi
energy with the occupation of the level as follows
1
ρ0σ (µ) = sin2 [πhn0σ i] , (6.87)
πΓ
which is known as Friedel sum rule. In a case with electron-hole symmetry
and hn0σ i = 1/2, the previous expression reduces to
1
ρ0σ (µ) = . (6.88)
πΓ
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
0,σ E’’ −
0,σ
(a) (b)
E’ E’ E’+E’’
E−E’’
0,σ −
0,σ
0,σ −
E’’ 0,σ
Fig. 6.20 First (a) and second (b) order self-energy diagrams in the Anderson model.
Notice that this equation implies that in the symmetric case, the density
of states at the Fermi energy coincides with the corresponding one in the
(0)
unperturbed problem, i.e. ρ0σ (µ) = ρ0σ (µ).
Friedel sum rule implies the appearance of a narrow peak in the density
of states in the limit U/Γ → 0. Let us discuss how this comes about. In
section 5.4.3 we saw that the level Green’s function in the limit U/Γ → 0
(atomic limit) is given by Eq. (5.112). This equation suggests that when
U ≫ Γ, the density of states consists mainly of two subbands (of width
∼ Γ) around ǫ0 and ǫ0 + U , which have most of the total spectral weight.
However, Eq. (6.88) tells us that there is a finite density at the Fermi
energy. Therefore, the exact density of states must exhibit a narrow peak
at the Fermi energy, known as Kondo peak or Kondo resonance, the width
of which tends to zero in the limit U/Γ → 0. Indeed, it can be shown that
this weight decays exponentially in this limit.
tion as
1
G00,σ (E) = , (6.90)
E − ǫ0 + iΓsgn(E) − U hn0σ̄ i
where we have set µ = 0. Notice that the role of the interaction is to
shift the position of the resonant level, which moves to ǫ0 + U hn0σ̄ i. In
the special case in which ǫ0 = −U/2, known as the symmetric case, the
self-consistent solution, assuming that there is no magnetic solution, is
hn0σ i = hn0σ̄ i = 1/2. The problem exhibits in this case electron-hole
symmetry around µ = 0 and the density of states is still described by a
Lorentzian of width Γ.
Let us now analyze the contribution of the second-order diagram, see
Fig. 6.20(b). Such contribution is given by
Z ∞
dE ′′ ∞ dE ′ (0)
Z
(2) 2 (0) (0)
Σ00,σ (E) = U G00σ (E − E ′′ )G00σ̄ (E ′ )G00σ̄ (E ′ + E ′′ ).
−∞ 2π −∞ 2π
(6.91)
This expression is not easy to evaluate, but the main features of this self-
energy can be reproduced in a simple analytical calculation in which one
assumes a constant density of states for the unperturbed problem (see Ex-
ercise 6.7).
If in the diagram of Fig. 6.20(b) the Green’s function line is dressed with
the Hartree diagram and one considers the symmetric case (ǫ0 = −U/2), the
second-order approximation preserves the electron-hole symmetry around
(0)
µ = 0 and one has hn0σ i = hn0σ i. Moreover, in this case one can show that
(2) (2) (0)
ReΣ00,σ (µ) = ImΣ00,σ (µ) = 0. This implies that ρ0σ = ρ0σ and therefore
the Friedel sum rule is satisfied. This is one of the reasons why this second-
order approximation gives an excellent description in the symmetric case,
even if U is not too small in comparison with Γ.
In order to illustrate the effect of the electron-electron interaction in the
density of states, we have computed it numerically in the symmetric case
using the second-order self-energy of Eq. (6.91). The results for different
values of the ratio U/Γ are shown in Fig. 6.21.8 As one can see, as the U/Γ
increases, the density of states exhibits two subbands around ǫ0 and ǫ0 + U
and a narrow peak at the Fermi energy (the Kondo peak). Notice that the
height of this peak remains constant and it is equal to 1/(πΓ), as in the
case without electron-electron interaction. The appearance of this peak at
8 In this figure we explore cases in which U is considerably larger than Γ, which in
principle should be out of the scope of this second-order approximation. However, as
stated above, this approximation works nicely in the symmetric case and it reproduces
the main features of the exact solution [651].
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
1
1
U/Γ = 0.0
U/Γ = 5.0 0.8
0.8
U/Γ = 10.0 0.6
U/Γ = 15.0
DOS (1/πΓ)
0.4
0.6 0.2
0
-1 -0.5 0 0.5 1
0.4
0.2
0
-10 -8 -6 -4 -2 0 2 4 6 8 10
(E-µ)/Γ
Fig. 6.21 Density of states projected onto the localized level as a function of the energy
in the Anderson model for ǫ0 = −U/2 and different values of the ratio U/Γ. The
calculation has been done including the self-energy diagrams up to second order. The
inset shows a blow-up of the energy region close to the Fermi energy.
the Fermi energy has very important consequences for the low-temperature
transport properties of molecular junctions. This will be discussed in detail
in section 15.6.2.
address in this monograph and for those readers interested in this topic we
recommend Refs. [173, 174, 182, 185].
Finally, we would like to emphasize that at this stage the reader is ready
to study many important topics in solid state physics which are out of the
scope of this book. For instance, the formalism detailed in this chapter
is the starting point to understand the Fermi liquid theory, which is very
important to get a deeper insight into the physics of metals. The reader is
now also prepared to study the physics of the homogeneous electron gas,
which is a model system where one can learn many important lessons related
to the relevance of electronic correlations. Again, Refs. [173, 174, 182, 185]
are very recommendable for studying these topics.
6.11 Exercises
hφ0 |n1↑ n1↓ |φ0 i = hφ0 |n1↑ |φ0 ihφ0 |n1↓ |φ0 i
hφ0 |n2↑ n2↓ |φ0 i = hφ0 |n2↑ |φ0 ihφ0 |n2↓ |φ0 i.
6.2 Wick’s theorem II: Starting from the results of Exercise 5.1(b) about the
time evolution of the creation and annihilation operators of the two-sites system,
show without applying Wick’s theorem that
h i
(0) (0)
hφ0 |T c1σ (t)c†2σ̄ (t)c†2σ (t′ )c1σ̄ (t′ ) |φ0 i = −G12σ (t − t′ )G12σ̄ (t′ − t),
1 + +
Fig. 6.22 Diagrammatic expansion of the denominator of the Green’s function up to
first order in the electron-electron interaction.
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
~k2 2e2 kF 1
» ˛ ˛–
1 − k0 ˛˛ 1 + k0 ˛˛
ǫk,HF = − − ln ˛ ,
2m π 2 4k0 1 − k0 ˛
where k0 ≡ k/kF . Show also that the derivative of the dispersion relation exhibits
a logarithmic divergence at k = kF .
6.6 Ward identity: Demonstrate the Ward identity of Eq. (6.83).
6.7 Density of states and Kondo resonance in the Anderson model:
Compute the second-order contribution to the retarded self-energy in the Ander-
son model, see Eq. (6.91), in the symmetric ǫ0 = −U/2 by assuming that the
unperturbed density of states adopts the form
(0) 1/W, −W/2 < E < W/2
ρ0σ (E) =
0, |E| > W/2
where W is a constant. Use this result to plot the density of states in the level
as a function of energy for different values of the ratio U/Γ. Hint: Use first the
spectral representation to write the unperturbed Green’s function appearing in
(0)
Eq. (6.91) in terms of the density of states ρ0σ .
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
Chapter 7
So far we have shown how the Green’s function techniques can help us to
understand the physics of systems in equilibrium. Since our goal is the
analysis of the transport properties of different nanocontacts, we have to
generalize those techniques to deal with situations in which the systems
are driven out of equilibrium. This is precisely the goal of this chapter
in which we shall discuss the so-called nonequilibrium Green’s function for-
malism (NEGF). This formalism was developed independently by Kadanoff
and Baym [186] and Keldysh [187] in the early 1960’s. Here we shall follow
Keldysh formulation of this approach and we shall refer to it as the Keldysh
formalism. This formalism is a natural extension of the diagrammatic the-
ory that we have presented in the previous chapter. The importance of
the Keldysh formalism lies in the fact that it allows us to go beyond the
usual linear response in a systematic manner. Since its appearance, it has
been used in a great variety of topics (see Refs. [188, 189] and references
therein). In particular, it has been applied to the study of electronic trans-
port in many types of nanoscale devices and it constitutes a basic tool that
will be used throughout the rest of the book.
Apart from the original paper [187], there exist a number of excellent
reviews devoted to the Keldysh formalism in the literature [188–191]. We
try to explain it here in a didactic manner, concentrating ourselves on its
application to the problems of molecular electronics that we have in mind,
rather than entering into very technical discussions about its foundation.
Bearing this in mind, we have organized this chapter as follows. We first
present the general ideas of the Keldysh formalism. Then, we shall briefly
discuss how to perform the diagrammatic expansion within this formalism.
We shall finish the formal discussion by reviewing both the main properties
of the functions appearing in this nonequilibrium formalism and the main
179
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
practical equations. Finally, the last part of this chapter is devoted to the
application of the Keldysh formalism to some simple transport problems.
8
−
8
still order the time arguments along a modified time contour. This contour
is referred to as the Keldysh contour and it is depicted in Fig. 7.1.
On this contour, the time runs from −∞ to +∞ in the upper branch,
whereas it does it backwards in the lower one, i.e. from +∞ to −∞. In
order to indicate in which branch the time arguments lie, we introduce
a subindex that will be equal to + for the upper branch and − for the
lower one. With this notation, we can write now the expectation value of
Eq. (7.4) as
hφ0 |S− (−∞, ∞)S+ (∞, t)AI (t)S+ (t, −∞)|φ0 i
hAi = , (7.5)
hφ0 |S− (−∞, ∞)S+ (∞, t)S+ (t, −∞)|φ0 i
if t lies in the upper branch or
hφ0 |S− (−∞, t)AI (t)S− (t, ∞)S+ (∞, −∞)|φ0 i
hAi = , (7.6)
hφ0 |S− (−∞, t)S− (t, ∞)S+ (∞, −∞)|φ0 i
if t lies in the lower one. Defining the operator Tc that orders the time
arguments along the Keldysh contour, we can rewrite the expectation value
as
hφ0 |Tc [AI (t)S− (−∞, ∞)S+ (∞, −∞)] |φ0 i
hAi = . (7.7)
hφ0 |S− (−∞, ∞)S+ (∞, −∞)|φ0 i
This expression can be in turn rewritten in a more familiar way by defining
the operator that describes the time-evolution along the Keldysh contour
Sc (∞, −∞) ≡ S− (−∞, ∞)S+ (∞, −∞). (7.8)
With this definition we can finally write the expectation value hAi as
follows
hφ0 |Tc [AI (t)Sc (∞, −∞)] |φ0 i
hAi = . (7.9)
hφ0 |Sc (∞, −∞)|φ0 i
Analogously, one can express the expectation value of any operator product.
The expectation value of Eq. (7.9) has formally the same structure as
in an equilibrium situation. The main difference is the fact that one has
to keep track of the branch in which the time arguments lie (t+ and t− ).
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
This implies that when defining the propagators in this formalism, there
are four different possibilities depending on the two time arguments. These
definitions are analogous to those of the causal function in the equilibrium
formalism
h i
hΨH |Tc ciσ (tα )c†jσ (t′β ) |ΨH i
Gij (tα , t′β ) = −i (7.10)
hΨH |ΨH i
h i
hΨH |Tc Ψσ (rtα )Ψ†σ (r′ t′β ) |ΨH i
G(rtα , r′ t′β ) = −i , (7.11)
hΨH |ΨH i
depending on whether we use the representation |ii or |ri. The subindexes
α and β take the values + and − and indicate in which branch the time
arguments lie. Let us now discuss in detail the expression for the four
possible functions:
where the operator T̄ orders the time arguments in the opposite way as
compared with the usual time-ordering operator T, i.e. in a antichrono-
logical order.
where the check symbol (ˇ) indicates that we are dealing with a 2 × 2
matrix in Keldysh space. The perturbative expansion couples the different
components of this matrix, which effectively leads to an enlargement of the
propagator space in a factor of 2. This enlargement is indeed quite natural
since in an out-of-equilibrium situation we have to determine not only the
states, the information of which is contained in the causal function, but also
the distribution function that describes how such states are occupied. This
latter information is provided by the off-diagonal functions in Eq. (7.16).
Formally speaking, the perturbative expansion is very similar to the
equilibrium one, and one has only to keep track of the matrix structure. A
additional complication is that in time-dependent problems, the products
are replaced by convolutions over intermediate arguments, which makes the
calculations considerably more complicated. Fortunately, transport prob-
lems often admit a stationary solution and then, the application of the
nonequilibrium formalism is not more complicated than the equilibrium
one.
As stated above, apart from the matrix structure introduced by the
Keldysh formalism, the rest of the perturbative approach is very similar to
the equilibrium one. To derive the perturbative expansion of the matrix
propagator of Eq. (7.16), one can use the expression of Eq. (7.9) and expand
the operator Sc . Let us recall that Sc (∞, −∞) ≡ S− (−∞, ∞)S+ (∞, −∞)
and the perturbative expansions of both time-evolution operators are given
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
by
∞
(−i)n ∞
X Z Z ∞
S+ (∞, −∞) = dt1 · · · dtn T [VI (t1 ) · · · VI (tn )] (7.17)
n=0
n! −∞ −∞
∞
(−i)n −∞
X Z Z −∞
S− (−∞, ∞) = dt1 · · · dtn T̄ [VI (t1 ) · · · VI (tn )] .
n=0
n! ∞ ∞
After expanding the operators S+ and S− , one applies the Wick’s theo-
rem in the standard way. Therefore, the resulting diagrammatic structure
is analogous to the one in equilibrium, the main difference being the en-
largement of the space that is encoded in the indexes α and β. We shall
discuss the peculiarities of the nonequilibrium diagrammatic expansion in
the next section.
Finally, since the structure of the diagrammatic expansion is identical to
the equilibrium one, such an expansion can be also summarized in a Dyson’s
equation, which in the nonequilibrium case has the following matrix form
Z Z
Ǧ(t, t′ ) = ǧ(t, t′ ) + dt1 dt2 ǧ(t, t1 )Σ̌(t1 , t2 )Ǧ(t2 , t′ ). (7.18)
The diagrams in this case are trivial because, as in the case of a static
potential, they consist of the repetition of identical scattering events.
The matrix self-energy is therfore given by (see Exercise 7.2)
µ ¶
V (r, t) 0
Σ̌(r, t) = . (7.22)
0 −V (r, t)
It is interesting to note that for this single-electron perturbation the
components Σ+− and Σ−+ vanish. The existence of off-diagonals com-
ponents of the self-energies in the Keldysh space is only possible in
the case of inelastic mechanisms such as electron-electron interaction
or electron-phonon interaction (see next case).
• Case 2: Electron-electron interaction.
Let us consider an electronic system where the electron-electron in-
teraction is assumed to be the perturbation. The system might be
out of equilibrium due to, for instance, the presence of a current. For
the sake of concreteness, let us assume that the unperturbed system
can be described by a tight-binding Hamiltonian and the interaction is
Hubbard-like (see Appendix A)
X
H = H0 + U ni↑ ni↓ . (7.23)
i
The diagrams are topologically identical to the equilibrium ones and the
only difference is the fact that one has to indicate where the time argu-
ments reside on the Keldysh contour. In this respect, every equilibrium
diagram gives rise to several diagrams for the different components of
the self-energy in Keldysh space. We illustrate this fact in Fig. 7.2,
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
+ + − +
iσ i−σ iσ i−σ
iσ i−σ iσ
+ + + + i−σ
Fig. 7.2 Examples of second-order self-energy diagrams in the Keldysh space for the
electron-electron interaction. The indexes + and − indicate in which branch the time
arguments lie.
In the previous section we have seen that the Dyson’s equation has acquired
an additional 2 × 2 matrix structure, which gives the impression that one
has to solve four times more equations than in the equilibrium case. Indeed,
one can show that the different functions in the 2 × 2 matrix of Eq. (7.16)
are not independent and the number of equations that one has to solve in
practice can be reduced to only two. In this sense, the goal of this section
is to derive those equations and to discuss the general properties of the
Keldysh-Green’s functions.
Analogously,
G−− ′ ′ +− ′ ′ −+ ′
ij (t, t ) = θ(t − t )Gij (t, t ) + θ(t − t)Gij (t, t ). (7.27)
Adding these two equations, we obtain the relation stated above.
On the other hand, from this relation and using the Dyson’s equation
in Keldysh space, see Eq. (7.18), one can show the following relation be-
tween the different elements of the self-energy matrix in Keldysh space (see
Exercise 7.3)
Σ++ + Σ−− = − Σ+− + Σ−+ .
¡ ¢
(7.28)
Other important relations are those between the Keldysh-Green’s func-
tions and the advanced and retarded functions Ga and Gr . Such relations
can be found as follows. Using the expression of Eq. (7.26), one obtains
G++ +−
′ ′ ′
£ +− ′ −+ ′
¤
ij (t, t ) − Gij (t, t ) = −θ(t − t ) Gij (t, t ) − Gij (t, t ) , (7.29)
and using the definitions of G+− and G+− , we arrive at
† †
G++ ′ +− ′ ′ ′ ′
ij (t, t ) − Gij (t, t ) = −iθ(t − t )hciσ (t)cjσ (t ) + cjσ (t )ciσ (t)i
= Grij (t, t′ ) (7.30)
Proceeding in an analogous way, one can show the following relations
Gr = G++ − G+− = G−+ − G−− (7.31)
a ++ −+ +− −−
G =G −G =G −G . (7.32)
These relations are crucial for the discussion of next section.
be easily deduced from those of g++ (E). Thus, we concentrate now on the
analysis of the functions g+− (E) and g−+ (E). From its definition in the
time domain (and in a discrete basis)
†
G+−
ij (t) = ihcjσ (0)ciσ (t)i, (7.49)
This implies that G+− ii (E) = 2πiρi (E)f (E), where f (E) is the Fermi func-
tion and ρi (E) is the local density of states in the site i. In the same way,
one can show that G−+ ii (E) = −2πiρi (E)[1 − f (E)]. Taking into account
this result, it is clear that G+− ∝ f (E) and G−+ ∝ 1 − f (E). This fact
together with the general relation
where we have assumed, without loss of generality, that the hopping ele-
ments tij are real. Our first task is to derive an expression for the electrical
current operator in this local basis. For this purpose, we first consider the
simple case of a tight-binding chain with only nearest-neighbor hoppings,
denoted by t. Such a chain is schematically represented in Fig. 7.3. Let us
compute now the current between the sites k and k + 1. Without doing
any calculation, one can guess that the operator must adopt somehow the
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
following form2
Xh i
I∝t c†kσ (t)ck+1σ (t) − c†k+1σ (t)ckσ (t) , (7.57)
σ
where the first term in the sum represents the current flowing in one direc-
tion and second one corresponds to the current flowing in the opposite one.
Let us see if a rigorous calculation confirms our intuition.
B A
t
........ ........
k−1 k k+1
Fig. 7.3 Schematic representation of a linear chain with only nearest-neighbor hoppings.
Notice that this has exactly the intuitive form that we had anticipated
above.
This expression can be easily generalized to any 3D system described by
a tight-binding Hamiltonian as in Eq. (7.56). The electrical current through
an arbitrary surface that separates two regions A and B is given by
ie X X n † o
I(t) = tij ciσ (t)cjσ (t) − c†jσ (t)ciσ (t) . (7.62)
~ σ
i∈A;j∈B
Let us now compute the expectation value of the current operator, for
instance, for the case of the chain. According to Eq. (7.61), one can write
(dropping the subindex A)
iet X n † o
hI(t)i = hckσ (t)ck+1σ (t)i − hc†k+1σ (t)ckσ (t)i . (7.63)
~ σ
The expectation values appearing in the previous equation can be expressed
in terms of the Keldysh functions G+− as follows
e X n +− o
hI(t)i = t Gk+1,k (t, t) − G+−
k,k+1 (t, t) , (7.64)
~ σ
and there is a similar expression for the most general case of Eq. (7.62).
In many situations, for instance when there is a constant voltage applied
in a junction, the problem admits a stationary solution and the Green’s
functions depend exclusively on the difference of the time arguments. In
those cases, Eq. (7.64) can be written in terms of the Green’s functions in
energy space as
e X ∞ dE n +−
Z o
hIi = t Gk+1,k (E) − G+−
k,k+1 (E) . (7.65)
~ σ −∞ 2π
We are now in position to discuss the electronic transport in some simple
examples of special interest.
where HL and HR are the Hamiltonians describing the left and right elec-
trodes, respectively. We assume that there is a bias voltage V applied across
the contact and that the potential drops abruptly in the interface region.
The task in this example is to compute the current-voltage characteris-
tics. According to Eqs. (7.63-7.65), the current evaluated at the interface
between the electrodes is given by3
2et ∞
Z
dE G+− +−
£ ¤
I = hIi = RL (E) − GLR (E) , (7.67)
h −∞
3 We assume that the voltage is time-independent and therefore the problem admits a
stationary solution. This allows us to write the current in terms of the Fourier transform
of the Green’s functions with respect to the difference of the time arguments.
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
G+− +− a a r r +−
LR = gLL ΣLR GRR + gLL ΣLR GRR , (7.71)
G+−
RL = G+− a a
RR ΣRL gLL + +−
GrRR ΣrRL gLL . (7.72)
4 Thisdoes not mean that the unperturbed system is out of equilibrium since in the
absence of coupling, there is no current and the electron distributions in both leads are
the equilibrium one.
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
where f (E) is the Fermi function and ρL/R is the local density of states of
the leads projected onto the sites L and R. Notice that we have already
taken into account the relative shift of the chemical potentials due to the
bias voltage V .
Using Eqs. (7.79) and (7.80), we can finally write the current as follows5
2e ∞ 4πt2 ρL (E − eV )ρR (E)
Z
I= dE [f (E − eV ) − f (E)] .
h −∞ |1 − t2 gLL (E − eV )gRR (E)|2
(7.81)
Notice that Eq. (7.81) has exactly the form of the Landauer formula,
i.e.
2e ∞
Z
I= dE T (E, V ) [f (E − eV ) − f (E)] , (7.82)
h −∞
where we can identify T (E, V ) as an energy and voltage-dependent trans-
mission probability given by
4πt2 ρL (E − eV )ρR (E)
T (E, V ) = . (7.83)
|1 − t2 gLL (E − eV )gRR (E)|2
As it can be seen, the transmission depends primarily on the coupling
element t and the local electronic structure of the leads.
For sufficiently low voltages, there is a linear regime where the current is
proportional to the voltage. In this limit, the conductance is given by G =
(2e2 /h)T (EF , V = 0), where T (EF , V = 0) is the zero-bias transmission at
the Fermi energy given by
4πt2 ρL (EF )ρR (EF )
T (EF , V = 0) = . (7.84)
|1 − t2 gLL (EF )gRR (EF )|2
One can often consider that the Green’s functions are constant around the
Fermi energy and one can also neglect their real part (this is the wide-band
approximation introduced in Chapter 5). This means that the lead Green’s
functions can be approximated by
i
gLL ≈
, (7.85)
W
where W = 1/πρL/R (EF ) (we are assuming a symmetric contact (gLL =
gRR ) for simplicity). Within this approximation, one obtains the following
expression for the transmission
4t2 /W 2
T = . (7.86)
(1 + t2 /W 2 )2
5 This expression for the current was first derived in Ref. [194] for a more realistic model.
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
This expression illustrates the transition from the tunnel regime, when
the electrodes are separated by a large distance, to the contact regime at
small distances. In the former limit, the transmission given in Eq. (7.86)
can be approximated by 4t2 /W 2 . This means that the dependence of the
transmission on the distance between the electrodes, and therefore that of
the linear conductance, is determined by t2 . At large distances, a hopping
element is roughly proportional to the overlap of the atomic orbitals and
decays exponentially with the distance between the corresponding atoms.
This is how the exponential length dependence, which we already discussed
in section 4.4, comes about from an atomistic point of view. From the scat-
tering approach, see section 4.4, we concluded that the length dependence
of a metallic tunnel junction is determined by the metal work function.
However, with this simple model, we get the impression that such a de-
pendence is governed by a local property, namely the coupling between the
outermost orbitals of the electrodes. These two pictures, which at first
glance look contradictory, can indeed be reconciled. This is, however, a
subtle issue that is out of the scope of this book and we refer the reader to
Ref. [195] for a discussion of this question.
When the electrodes approach each other the hopping t becomes of the
same order as the energy scale W and the transmission can reach unity
and in turn the conductance approaches the quantum of conductance G0 =
2e2 /h. The transition from tunnel to contact was first discussed within
this type of atomistic models in Ref. [194] in connection with the first
experiment that explored such a transition [55]. For an overview on recent
experiments exploring the tunnel-to-contact transition both in single atoms
and molecules, see Refs. [196, 197].
Let us now study in more detail the tunnel limit (t → 0). In this case,
the non-linear current of Eq. (7.81) can be approximated by
8πe 2 ∞
Z
I= t dE ρL (E − eV )ρR (E) [f (E − eV ) − f (E)] , (7.87)
h −∞
which tell us that the current in this limit is determined by the convolution
of the local density of states of both electrodes. This well-known expression
is a fundamental result for the theory of STM and provides a simple inter-
pretation of the STM images. Assuming that the left electrode represents
a STM tip with a constant density of states around the Fermi energy, the
differential conductance at low temperatures is simply given by
dI 2e2
G(V ) = = 4πt2 ρL (EF )ρR (EF + eV ), (7.88)
dV h
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
i.e. the conductance is a measure of the local density of states of the sample
(or right electrode in our case).
If we now substitute the expressions for I(t) and hI(t)i for an atomic
contact and we write the result in terms of the Green’s functions, we obtain
2e2 ∞
Z
dE G+− −+ +− −+
£
P (0) = LR (E)GRL (E) + GRL (E)GLR (E)−
h −∞
G+− −+ +− −+
¤
LL (E)GRR (E) − GRR (E)GLL (E) . (7.91)
Here, in order to obtain this expression, we have made use of Wick’s theo-
rem to decouple the averages of four operators (let us remind that this is
valid since our electron system is noninteracting).
At this stage the calculation of the shot noise has been reduced to
the computation of the different Keldysh-Green’s functions that appear in
Eq. (7.91). These functions can be calculated following exactly the same
procedure detailed in the previous subsection. If we now assume zero tem-
perature and use the wide-band approximation of Eq. (7.85) for the un-
perturbed Green’s functions, we can obtain the following expression (see
Exercise 7.5)
4e2
P (0) = T (1 − T )V, (7.92)
h
which is the result derived in section 4.7 using the scattering approach.
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
where ǫ0 is the position of the resonant level, which in principle can also
depend on the bias voltage, and tL,R are the matrix elements describing
the coupling to the reservoirs. Here, L and R denote the outermost sites
of the left and right electrodes, respectively. On the other hand, we now
assume that there is a constant bias voltage across the system and our task
is to compute the current-voltage characteristics.
We start by evaluating the current at the interface between the left
electrode and the level, which in terms of the Green’s functions G+− can
be written as follows
2etL ∞
Z
dE G+− +−
£ ¤
I= L0 (E) − G0L (E) . (7.94)
h −∞
In order to determine the Green’s functions in the previous expression,
we use again the Keldysh formalism and we treat the coupling terms be-
tween the level and the electrodes, i.e. the second line in Eq. (7.93), as
a perturbation. With this choice the only non-vanishing elements of the
self-energy are: Σr,a r,a r,a r,a
L0 = Σ0L = tL and ΣR0 = Σ0R = tR .
Following now the same steps as in section 7.4.1, we can write the current
in terms of diagonal elements of the Green’s functions as
2etL ∞
Z
£ +−
(E)G−+ −+ +−
¤
I= dE gLL 00 (E) − gLL (E)G00 (E) . (7.95)
h −∞
Now, to determine the full Green’s functions, we use the Dyson’s equa-
tion, Eq. (7.47), to write
+−/−+ +−/−+
G00 = (1 + Gr Σr )00 g00 (1 + Σa Ga )00 + (7.96)
+−/−+ a +−/−+ a
Gr00 Σr0L gLL ΣL0 Ga00 + Gr00 Σr0R gRR ΣR0 Ga00 .
If we now substitute this expression into the current formula, the term
+−/−+ +−/−+
containing gLL is canceled. Moreover, the term proportional to g00
+−/−+
does not contribute either. The reason is that g00 (E) ∝ δ(E − ǫ0 ) and
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
the prefactor of this term vanishes at E = ǫ0 .6 Thus, the current can now
be expressed as Z
∞
2e
I = 4π 2 t2L t2R dE ρL (E)ρR (E)|Gr00 (E)|2 [fL (E) − fR (E)] , (7.97)
h −∞
where it is implicitly assumed that the density of states (and distribution
function) of the left electrode is shifted by eV . Notice that we have already
used the expression of the lead Green’s functions in terms of the local
density of states and Fermi functions.
At this point, the only remaining task is the calculation of Gr00 (E), but
this is something that we have already done in section 5.3.3 and we just
recall here the result
1
Gr00 (E) = r (E) − t2 g r (E) . (7.98)
E − ǫ0 − t2L gL R R
Therefore, the currentZ adopts again the form of the Landauer formula
2e ∞
I= dE T (E, V ) [f (E − eV ) − f (E)] , (7.99)
h −∞
where this time the transmission T (E, V ) is given by
4π 2 t2L t2R ρL (E − eV )ρR (E)
T (E, V ) = r (E − eV ) − t2 g r (E)|2 . (7.100)
|E − ǫ0 − t2L gL R R
To simplify this expression, we use now as in section 5.3.3 the wide-
band approximation and neglect the energy dependence introduced by the
r
leads. This way, gL/R ≈ −iπρL/R (EF ) and we define the scattering rates
2
ΓL/R = πtL/R ρL/R (EF ). In this approximation the transmission can be
written as
4ΓL ΓR
T (E, V ) = . (7.101)
(E − ǫ0 )2 + (ΓL + ΓR )2
In this case, the voltage dependence of the transmission may only stem from
the eventual voltage dependence of the level position. This expression is
the well-known Breit-Wigner formula that was derived in Chapter 4 within
the scattering approach (see Exercises 4.5 and 4.8) and it will be used
extensively in later chapters.
Again, in the linear regime the low-temperature conductance is simply
given by G = (2e2 /h)T (EF , 0). This expression shows that the maximum
conductance is reached when EF = ǫ0 , which is the resonant condition.
In the symmetric case (ΓL = ΓR ), this maximum is equal to G0 = 2e2 /h,
irrespectively of the value of the scattering rates. These facts are illustrated
in Fig. 7.5. The non-linear current-voltage characteristics of this model will
be discussed in detail in Chapter 15.
6 Physically speaking, it is quite reasonable that this term does not contribute to the
current. It makes no sense that the current depends on the occupation of the level before
being coupled to the electrodes.
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
1 1
0.8 (a) 0.8
(b)
G/G0
0.6 0.6
0.4 0.4
0.2 0.2
0 0
-10 -5 0 5 10 0 0.2 0.4 0.6 0.8 1
(EF - ε0)/Γ ΓL/ΓR
Fig. 7.5 Zero-temperature linear conductance in the resonant tunneling model. (a)
Linear conductance (normalized by G0 = 2e2 /h) as a function of the level position, ǫ0
for a symmetric contact ΓL = ΓR = Γ. (b) Linear conductance at resonance (ǫ0 = EF )
as a function of the ratio between the scattering rates.
7.5 Exercises
N
X
V(t) = V (ri , t). (7.102)
i=1
Apply Wick’s theorem to demonstrate that the self-energy is given by Eq. (7.22).
7.3 Properties of the Keldysh-Green’s functions:
(a) Demonstrate the property of Eq. (7.28). Hint: Use the property of
Eq. (7.25) and the Dyson’s equation in Keldysh space.
(b) Demonstrate Eq. (7.46).
7.4 Shot noise in a single-channel point contact:
Derive the expression of the zero-frequency shot noise of a single-channel point
contact following the discussion of the example of section 7.4.2 and demonstrate
that it is given by
4e2
P (0) = T (1 − T )eV,
h
where T is the energy-independent transmission coefficient of the contact given
by Eq. (7.86).
7.5 Electrical current through a linear chain: Consider the electronic trans-
port in a finite one-dimensional system formed by a tight-binding chain with N
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
sites such that the site 1 is connected to the left electrode through a hopping tL
and the site N is connected to the right electrode with a hopping tR . Show that
the current formula in this case is given by
2e 2 2 2 ∞
Z
I= 4π tL tR dE ρL (E − eV )ρR (E)|Gr1N (E)|2 [f (E − eV ) − f (E)] .
h −∞
For the sake of simplicity, consider that in the chain there are only hoppings
between nearest-neighbor atoms, t, and that the on-site energy is given by ǫ0 .
Study the linear conductance of this system as a function of the number of sites
N in the chain and show that it may exhibit parity oscillations, depending on
whether N is even or odd.
7.6 Thermopower of a single-channel point contact: Using the model
of section 7.4.1, derive the expression for the thermopower for a single-channel
contact and show that it coincides with the result obtained with the scattering
approach in section 4.8.
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
Chapter 8
In the previous chapter we showed how the Keldysh formalism can be com-
bined with simple Hamiltonians to compute the current in model systems.
In this chapter we shall exploit this technique and derive some general
expressions for the electrical current that can be combined with realistic
methods for the determination of the electronic structure. To be precise,
we shall address three basic issues:
This chapter is rather technical and it can be skipped by those who are
not so interested in the algebra behind the current formulas. Anyway, we
recommend to read the next section about the derivation of the Landauer
formula, since the expression obtained there for the elastic transmission will
be frequently used in subsequent chapters.
205
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
the only difference being the expression for the transmission coefficient. In
this section we shall demonstrate that this was not a coincidence and we
shall derive a general expression for the elastic current valid for any type
of atomic and molecular junction.
Let us consider a contact with arbitrary geometry like the one depicted
in Fig. 8.1. Such a contact can be either an atomic contact or a molecular
junction. Since we shall ignore inelastic interaction in this discussion, one
can describe the system in terms of the following generic tight-binding
Hamiltonian
where i, j run over the atomic sites and α, β denote the different atomic
orbitals. The number of orbitals in each site can be arbitrary. For the sake
of simplicity, we assume that the local basis is orthogonal. Later in this
section, we shall generalize the results to the case of nonorthogonal basis
sets. Notice also that we are assuming that matrix elements are independent
of the spin, i.e. for the moment we do not consider magnetic situations.
We now distinguish three different parts in this contact: the reservoirs
L and R, and a central region that can have arbitrary size and shape.
In principle, the reservoirs L and R could also have an arbitrary shape
and we assume that an electron in these subsystems has a well-defined
temperature and chemical potential. In other words, these regions play the
role of electron reservoirs, in the spirit of the scattering approach of Chapter
4. The separation of the contact in these three subsystems is somewhat
arbitrary, especially in the linear response regime, and one can play with
that, as we shall discuss below. We also assume that there is no direct
coupling between the reservoirs. With this assumption the Hamiltonian
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
where i runs over the atoms of the left electrode which are connected with
the atoms in the central region C, and j runs over the atoms of the central
region coupled to the left electrode (in principle, all of them). The indexes
α and β indicate the different atomic orbitals in every site.
Following the ideas of the last section of the previous chapter, we make
use of nonequilibrium Green’s function techniques to calculate the current.
First of all, we express the expectation values appearing in the current
expression in terms of the Keldysh-Green’s function G+− . This function
gives information about the distribution function of the system and in a
local basis it adopts the following form
′
+−,σσ
Giα,jβ (t, t′ ) = ihc†jβ,σ′ (t′ )ciα,σ (t)i. (8.4)
Using this expression one can write the current as
e X h
+−,σσ +−,σσ
i
I= tiα,jβ Gjβ,iα (t, t) − tjβ,iα Giα,jβ (t, t) . (8.5)
~
i∈L;j∈C;α,β,σ
where Tr denotes the trace over atoms and orbitals in the central region C.
The prefactor 2 comes from the sum over spins, since for the moment we
do not consider any magnetic situation. For the same reason, we drop the
superindex σ in the Green’s functions.
This transport problem admits a stationary solution and therefore, the
different Green’s functions only depend on the difference of time arguments.
Thus, we can Fourier transform with respect to the difference of the time
arguments and write the current as
2e ∞
Z
dE Tr G+− +−
£ ¤
I= CL (E)tLC − tCL GLC (E) . (8.8)
h −∞
Notice that the current is expressed in terms of the trace of a matrix whose
dimension is the number of orbitals in the central region, which we denote as
NC . At this stage, the problem has been reduced to the determination of the
functions G+− in terms of matrix elements of the Hamiltonian of Eq. (8.1).
We shall calculate these functions considering the coupling terms between
the electrodes and the central region as a perturbation. Then, starting from
the Green’s functions for the three isolated systems, we shall determine the
corresponding functions for the whole system. With this choice, the self-
energies of the problem are the hopping matrices defined in Eq. (8.6) and
the equivalent ones for the interface between the central region and the
right electrode R.
We now follow the ideas of section 7.4.3 and make use of Dyson’s equa-
tion in Keldysh space, see Eq. (7.46), to write the functions G+− as follows1
G+− +− a r +−
LC = gLL tLC GCC + gLL tLC GCC (8.9)
G+−
CL = G+− a
CC tCL gLL + +−
GrCC tCL gLL ,
r,a
where gXXare the (retarded, advanced) Green’s functions of the uncoupled
reservoirs (X = L, R). Introducing this equation in the current expression
and making use of the relation G+− − G−+ = Ga − Gr , we obtain
2e ∞
Z
dE Tr G−+ +− +− −+
£ ¤
I= CC tCL gLL tLC − GCC tCL gLL tLC . (8.10)
h −∞
Then, we determine G+−/−+ by means of the relation
G+−/−+ = (1 + Gr t) g+−/−+ (1 + tGa ) . (8.11)
Taking the element (C,C) in this expression we obtain
+−/−+ +−/−+ +−/−+
GCC = GrCC t̂CL gLL tLC GaCC +GrCR tCR gRR tRC GaCC . (8.12)
1 In order to abbreviate the notation, we do not write the energy argument E explicitly.
Moreover, since there are no inelastic processes involved in this model, the self-energies
Σ+− associated with them vanish.
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
+−/−+
Notice that there is an additional contribution containing gCC that was
left out in the previous expression. The reason for this is that, in analogy
with our discussion of the resonant tunneling model in section 7.4.3, one
can show that such a term does not contribute to the final expression of
the current.
Substitution of the previous equation in the expression of the current
yields
2e ∞
Z
−+ +−
dE Tr GrCC tCR gRR tRC GaCC tCL gLL
£
I= tLC −
h −∞
+− −+
GrCC tCR gRR tRC GaCC tCL gLL
¤
tLC . (8.13)
Let us recall that the unperturbed functions g+− and g−+ satisfy the
following relations2
g+− = (ga − gr ) f = 2i Im (ga ) f
−+ (8.14)
g = (ga − gr ) (f − 1) = 2i Im (ga ) (f − 1),
where f is the Fermi function. Thus, the current can be expressed as
8e ∞
Z
I= dE Tr [GrCC tCR Im {gRR
a
} tRC GaCC tCL Im {gLL
a
} tLC ]
h −∞
× (fL − fR ) . (8.15)
Here, fL/R is the Fermi function of the corresponding electrode, which takes
into account the shift of the chemical potential induced by the voltage.
One can further simplify the expression of the current by defining
Σr,a r,a r,a r,a
L = tCL gLL tLC and ΣR = tCR gRR tRC , (8.16)
These matrices are nothing else but the self-energies of this problem for the
subspace of the central region. These self-energies describe the influence of
the reservoir in the central region and they depend both on the coupling
between the reservoirs and the central region and on the local electronic
structure of the leads. Notice that these matrices have a dimension equal
to the number of orbitals in the central region. Using these definitions, the
current can now be rewritten in the following familiar form
2e ∞
Z
I= dE T (E, V ) (fL − fR ) , (8.17)
h −∞
where T (E, V ) is the energy- and voltage-dependent total transmission
probability of the contact given by
T (E, V ) ≡ 4Tr [ΓL GrCC ΓR GaCC ] . (8.18)
2 Noticethat in Eq. (8.14) we have assumed that that Hamiltonian is real, i.e. there is
time reversal symmetry. One can easily show that this implies that gr (E) = [ga (E)]∗ .
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
where we have defined the scattering rate matrices as ΓL,R ≡ Im{ΣaL,R }.3
The voltage dependence of the transmission comes through the scattering
rates (i.e. via the leads), but also through the possible voltage dependence
of the Hamiltonian matrix elements of the central region.
We can further symmetrize this expression by £ using † the cyclic
¤
property
£ † of the trace
¤ and write T (E, V ) = Tr t(E, V )t (E, V ) =
Tr t (E, V )t(E, V ) , where
1/2 1/2
t(E, V ) = 2ΓL GrCC ΓR (8.19)
1/2
is the transmission matrix of the system. The existence of Γ as a real
matrix is warranted by Γ being positive definite (see Exercise 8.1).
Finally, the current adopts the form
2e ∞
Z
dE Tr t† (E, V )t(E, V ) (fL − fR ) ,
£ ¤
I= (8.20)
h −∞
valid for arbitrary bias voltage. In the linear regime this expression reduces
to the standard Landauer formula for the zero-temperature conductance
N
2e2 £ † ¤ 2e2 X
G= Tr t (EF , 0)t(EF , 0) = Ti , (8.21)
h h i=1
where Ti are the eigenvalues of t̂† t (or tt̂† ) at the Fermi level. As one can
see, in principle the number of channel would be NC , which is the dimension
of the matrix t† t. However, as we stated at the beginning of this section, the
separation in three subsystems in somewhat arbitrary and one can evaluate
the current at any point. Thus, it is evident that the actual number of
channels is controlled by the narrowest part of the junction. This fact will
be very important in our discussion of the conduction channels in metallic
single-atom contacts, see section 11.5. Notice also that in this formulation,
the conduction channels , defined as the eigenfunctions of t† t̂, are linear
combinations of the atomic orbitals in the central system.
As a result of the discussion above, we have not only re-derived the
Landauer formula, but more importantly, we have also obtained an explicit
formula for the transmission as a function of the microscopic parameters of
the system. As one can see in Eq. (8.18) or in Eq. (8.19), the determination
of the transmission requires the calculations of both the retarded/advanced
Green’s functions of the central system and the scattering rate matrices.
These functions can be determined from their Dyson’s equation
¤−1
GaCC = (GrCC )† = (E − i0+ )1 − HCC − ΣaL − ΣaR
£
, (8.22)
3 We have assumed without loss of generality that the hopping matrix elements are real.
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
where HCC is the Hamiltonian of the central region and the self-energies
ΣX (X = L, R) are given by Eq. (8.16).
On the other hand, the calculation of the scattering rate matrices,
which are the imaginary part of the self-energies of Eq. (8.16), requires
the knowledge of the Green’s functions of the uncoupled reservoirs, gXX
(with X = L, R). The leads are semi-infinite systems and thus they cannot
possess in practice a very complicated geometry. A typical option is to
describe these leads as ideal surfaces of the corresponding material and the
unperturbed Green’s functions are then computed using special recursive
techniques like the so-called decimation [198].
Let us end this section with a brief technical discussion. The quantity
t(E, V ) appearing in Eq. (8.19) has been called transmission matrix without
a real justification. We should demonstrate that this matrix fulfills the
properties of a transmission matrix. In particular, we should at least prove
that the eigenvalues of tt† are bounded between 0 and 1. Indeed, this
property can be shown using a few algebraic manipulations (see Exercise
8.2).
Another way of showing that t(E, V ) in Eq. (8.19) is indeed the trans-
mission matrix of the contact is via the so-called Fisher-Lee relation [199],
which expresses the elements of the scattering matrix in terms of Green’s
functions. For the readers interested in this route, we recommend the orig-
inal work of Ref. [199] and the discussion on this matter in Chapter 3 of
Ref. [50].
where tσ is the transmission matrix of the spin sector σ and Tnσ are the
corresponding transmission coefficients. The transmission tσ is given by
Eq. (8.19), where all the quantity are referred to the spin band σ.
The previous current formula describes any (elastic) situation where
there is no mixing of the two spin bands. This is what occurs in most of
the atomic-scale junctions that we have in mind, where the system size is
clearly smaller than the spin-diffusion length. However, this is no longer
true if, for instance, there is a small domain wall of atomic size in the
junction or a strong spin-orbit interaction is present. Let us show how the
formula for the elastic current is modified in those situations.
A system in which the majority and minority spin bands are mixed can
be generically described by the following tight-binding Hamiltonian
†
X ′
H= hσσ
iα,jβ ciασ cjβσ ′ , (8.30)
ijαβσσ ′
where i, j run over the atomic sites, α, β denote the different atomic orbitals,
and σ = ↑, ↓ the spin. Within this model, the current can be computed fol-
lowing the same steps as in the case with spin symmetry and we only sketch
here the main idea and the final result. Briefly, the atomic-scale contact is
divided into three parts, a central region C containing the constriction and
the left/right (L/R) leads. The retarded Green’s functions of the central
part read4
−1
GrCC = [ESCC − HCC − ΣrL − ΣrR ] , (8.31)
where ΣrX = (tCX − ESCX )gXX r
(tCX − ESCX )† are the lead self-energies
(X = L, R). Here, tCX and SCX are the hoppings and overlaps between the
r
C region and the lead X, and gXX is a lead Green’s function. Notice that
the dimension of all the matrices in the previous equation is equal to the
total number of orbitals in the central region multiplied by two. This factor
4 Notice that we take into account the possibility of using non-orthogonal basis sets.
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
two comes from the structure in spin space. As before, the transmission
1/2 1/2
matrix is given by t = 2ΓL GrCC ΓR , but this time the scattering rate
matrices are given by where ΓX = i[ΣrX − (ΣrX )† ]/2. The reason for this is
r a
that, in general, the Hamiltonian is not real and gXX = (gXX )† . Finally,
the current then adopts the standard Landauer form of Eq. (8.28), but now
the trace includes not only a sum over the orbitals in the central part, but
also over spins. Finally, the low-temperature linear conductance can be
written as G = (e2 /h) n Tn , where Tn are the transmission coefficients,
P
obtain
+−/−+ +−/−+ +−/−+
GCC = gCC + GrCC t̂CL gLL tLC GaCC + (8.33)
+−/−+ +−/−+ a
GrCR tCR gRR tRC GaCC − GrCC ΣCC GCC .
Here we have used the fact that the interactions are restricted to the central
region, which in practice means that the inelastic self-energies Σ+−/−+
have only a (CC) component. Introducing now these Green’s functions in
Eq. (8.10), one can readily show that the current can be written as the sum
of two contributions: I = Iel + Iinel , where
8e ∞
Z
Iel = dE Tr [GrCC ΓR GaCC ΓL ] (fL − fR ) (8.34)
h −∞
Z ∞
4ie
dE Tr GaCC ΓL GrCC (fL − 1)Σ+− −+
© £ ¤ª
Iinel = CC − fL ΣCC (8.35).
h −∞
Again, the trace in these expressions has to be understood as a sum over all
the orbitals in the central region. The first term, Iel , represents the elastic
current and it has the same form as the Landauer formula derived in the
previous section. The second term, Iinel , which we call inelastic current,
is the new contribution due to the inelastic interactions. Notice that this
term has a rather asymmetric form, which is a consequence of our choice of
computing the current in the left interface. If wanted, one can symmetrize
this expression by combining it with the inelastic current evaluated in right
interface5 and using current conservation to define the inelastic current as
L R
Iinel = (Iinel + Iinel )/2.
From Eq. (8.35) it is not obvious that the inelastic current vanishes at
zero bias. However, this can be shown by using the general relations for a
system in equilibrium
Σ+− (E) = (Σr − Σa ) f (E); Σ−+ (E) = (Σr − Σa ) (f (E) − 1), (8.36)
GaCC = (GrCC )† = [(E − i0+ )1 − HCC − ΣaL − ΣaR − ΣaCC ]−1 , (8.37)
5 Such expression reads
4ie ∞
Z n h io
R +− −+
Iinel =− dE Tr Ga r
CC ΓR GCC (fR − 1)ΣCC − fR ΣCC .
h −∞
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
eV
λ hω
L
ε0 R
lator of energy ~ω by
µ ¶
1
Hvib = ~ω b† b + , (8.39)
2
where the creation and annihilation operators b† and b satisfy the bosonic
commutation relations, e.g. [b, b† ] = 1. Finally, the interaction between the
vibration mode and the conduction electrons is described by the following
Hamiltonian [174]
He−vib = λc†0 c0 (b† + b), (8.40)
where λ is the electron-vibration coupling constant and c†0 and c0 are the
fermionic operators related to the electronic level.6
In this simple model, the central region consists of a single site and
therefore the Green’s functions, scattering rates and self-energies appearing
in the current formulas of Eqs. (8.34) and (8.35) are just scalars. Such
formulas reduce to the following expressions
Z ∞
8e
Iel = ΓL ΓR dE |Gr |2 (fL − fR ), (8.41)
h −∞
Z ∞
4ie
dE |Gr |2 (fL − 1)Σ+− −+
£ ¤
Iinel = ΓL e−vib − fL Σe−vib . (8.42)
h −∞
Here, the Green’s function Gr (E) refers to the central site or resonant
level. Moreover, as usual, we have assumed that the scattering rates, ΓL,R ,
that describe the strength of the coupling between the resonant level and
the leads are independent of the energy. Now, we have to determine the
6 The spin does not play any role in this problem and we have dropped it in the previous
expression.
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
This gives a constant contribution that simply renormalizes the position of the resonant
level and we ignore it in what follows.
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
eV hω eV
L L
ε0 R ε0 R
hω eV hω eV
L L
ε0 R ε0 R
Fig. 8.3 Schematic representation of the elastic (a) and inelastic (b-d) tunneling pro-
cesses that can occur in the model in which an electronic level is coupled to a single
vibration mode. Here, we have assumed that the electron-phonon interaction is weak
and the processes (b-d) are responsible for the inelastic correction to the elastic current
up to order λ2 .
electronic Green’s functions can be neglected, i.e. G̃r (E) = G̃r (EF ). This
means in practice that we assume that both the local density of states
and the transmission are energy-independent. This is a good approxima-
tion in two cases: (i) when the coupling to the leads is so strong that
ΓL + ΓR >> ~ω, eV, |EF − ǫ0 | and (ii) when the resonant level is far away
from the Fermi energy, i.e. |EF − ǫ0 | ≫ ΓL,R , eV, ~ω. With this approxi-
mation the different terms can be computed analytically. At temperatures
well below the vibrational energy, the correction to the elastic current is a
competition between the emission term in Iinel and the elastic correction
δIel . Assuming a symmetric junction, ΓL = ΓR = Γ, the three contri-
butions to the zero-temperature differential conductance are given by (see
Exercise 8.6)
G0el
= T,
G0
λ2
½ 2
δGel (V ) T (1 − T )/2; |eV | ≤ ~ω
= 2
G0 Γ T 2 (1 − 2T )/2; |eV | > ~ω
λ2
½
Ginel (V ) 0; |eV | ≤ ~ω
= 2 2 , (8.49)
G0 Γ T /4; |eV | > ~ω
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
_ _ _ _
− hω + hω eV − hω + hω eV
where T = 4Γ2 |G̃r (EF )|2 is the elastic transmission in the absence of
electron-vibration interaction. Notice that δGel has a discontinuity (step
down) at eV = ±~ω proportional to −T 3 /2, while the emission term con-
tributes to this jump as ∼ +T 2 /4. This means that the sign of the con-
ductance jump depends on the junction transmission and it is given by:
(λ2 /Γ2 )T 2 (1 − 2T )/4. Notice that the magnitude is determined by the ra-
tio of the two relevant coupling constants, λ and Γ, which has been assumed
to be small. On the other hand, the conclusion of this analysis is that the
electron-vibration interaction in this simple model is reflected in the appear-
ance of a jump in the low-temperature conductance at eV = ±~ω. This
jump is seen as a step up in conductance for T < 1/2 and as step down
for T > 1/2. This conclusion is summarized schematically in Fig. 8.4. The
signature of the vibration modes can be seen more clearly in the second
derivative of the current, d2 I/dV 2 , where it appears as a peak or as a dip
depending on the junction transmission.8 The results of this model will be
discussed in much more detail in section 17.1.1.
Using the general relations G+− + G−+ = G++ + G−− and G+− − G−+ =
Ga − Gr , it is straightforward to show that the current evaluated at the
left interface, IL , is given by
4ie ∞
Z
dE Tr ΓL G+− r a
© £ ¤ª
IL = CC + (GCC − GCC )fL , (8.51)
h −∞
where the scattering rate ΓL is defined in the usual way.
Analogously, one can obtain the expression of the current, evaluated
this time at the right interface, IR . Writing then the current in a more
symmetric manner as I = (IL + IR )/2, one arrives at
2ie ∞
Z
dE Tr (ΓL − ΓR )G+− r a
© ª
I= CC + (fL ΓL − ΓR fR )(GCC − GCC ) .
h −∞
(8.52)
This is the Meir-Wingreen formula in its most general form. It is completely
equivalent to the expression derived above and in the non-interacting case
it reduces to the Landauer formula (see Exercise 8.7). The “popularity”
of this formula is due to the fact that it takes an appealing form in the
case in which the couplings to the leads differ only by a constant factor,
ΓL (E) = λΓR (E). In this case, the current reads
8e ∞
Z
I= dE Tr {ΓA} (fL − fR ), (8.53)
h −∞
where Γ ≡ ΓL ΓR /(ΓL + ΓR ) and A ≡ i(GrCC − GaCC )/2 is the spectral
function of the central region. The division in the expression of Γ has to
be understood as multiplication by the inverse of the matrix appearing in
the denominator. The nice thing about this formula is that the current is
expressed in terms of the spectral function, A. Unfortunately, the condition
of proportionality of the scattering rates is quite restrictive and most cases
it is not really fulfilled. For applications of this latter formula, see Exercises
8.9 and 8.10.
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
We shall calculate the current for an arbitrary potential profile in the central
region (encoded in the functions Ui (t)), the actual shape of which should
in principle be obtained self-consistently [165].
In order to derive the current formula in this situation, we shall follow
the same steps taken in section 8.1 and we shall emphasize here only the
main differences with respect to that calculation. Our starting point is the
expression of the time-dependent current evaluated at the left interface,
which can be written in terms of the Green’s functions as follows
2e £
I(t) = Tr G+− +−
¤
CL (t, t)tLC − tCL GLC (t, t) . (8.55)
~
To determine the Green’s functions we follow the same perturbative ap-
proach as in section 8.1. The essential difference now is that the Green’s
functions depend explicitly on two time arguments (rather than on their dif-
ference), which introduces an extra complication, as we are about to show.
Using the Dyson’s equation [see Eq. (7.46)] we can express the functions
appearing in the current as10
G+− ′
© +− a r +−
ª ′
LC (t, t ) = gLL ◦ tLC ◦ GCC + gLL ◦ tLC ◦ GCC (t, t ) (8.56)
G+− ′ +− a r +− ′
© ª
CL (t, t ) = GCC ◦ tCL ◦ gLL + GCC ◦ tCL ◦ gLL (t, t ),
product, the derivation still follows the same steps as in section 8.1. Thus,
we can easily arrive at the following expression for the current
2e £ r
I(t) = Tr GCC ◦ Σ−+ a +−
R ◦ GCC ◦ ΣL −
~
GrCC ◦ Σ+− a −+
¤
R ◦ GCC ◦ ΣL (t, t), (8.57)
which is the analog of Eq. (8.13). Here, we have define the “lead self-
energies”
h i
+−/−+ +−/−+
ΣX (t, t′ ) = tCX ◦ gXX ◦ tXC (t, t′ ), (8.58)
where X = L, R.
The lead Green’s functions have now a more complicated time depen-
dence. Due to the ac voltage they oscillate on time as follows11
′
c
gX (t, t′ ) = e−iφX (t) gX
c
(t − t′ )eiφX (t ) , (8.59)
where c = r, a, +−, −+. Here, ∂φX (t)/∂t = µX (t)/~, where µX (t) is the
chemical potential of the corresponding electrode. Therefore, φX (t) =
dc ac
(UX /~)t + αX cos(ωt), with αX = UX /(~ω).
As usual, it is more convenient to work in energy space and for this
reason we now Fourier transform with respect to the two time arguments
1
Z Z
′ ′
c ′
gX (t, t ) = dE dE ′ e−iEt/~ eiE t /~ gX
c
(E, E ′ ). (8.60)
2π
From Eq. (8.59) it is easy to show that the lead Green’s functions admit a
Fourier expansion of the form
dE −iE(t−t′ )/~ c
X Z
′
c
gX (t, t′ ) = eimωt e gX (E, E + m~ω). (8.61)
m
2π
c
In other words, the functions gX (E, E ′ ) satisfy the following relation
X
c
gX (E, E ′ ) = c
[ĝX ]0,n (E)δ(E − E ′ + n~ω), (8.62)
n
c c
where [ĝX ]0,n (E) ≡ gX (E, E +n~ω). Other Fourier components are related
c c
by [gX ]n,m (E) = [gX ]0,m−n (E + n~ω). These Fourier components can be
seen as the matrix elements of the Green’s functions in energy space. We
11 This time dependence can be shown by solving the Dyson’s equation for the lead
denote the matrices in this space with a “hat” symbol. The previous rela-
tion is the mathematical expression of the fact that all physical quantities
in this problem oscillate in time with the driving frequency and all its har-
monics.
With the help of the relation
X
eiα cos(ωt) = im Jm (α)eimωt , (8.63)
m
where Jm is the Bessel function of first kind of order m, one can show that
the Fourier components of the lead Green’s functions are given by
c,eq
X
c
[ĝX ]n,m (E) = im−n Jn−l (αX )Jm−l (αX )gX dc
(E − UX + l~ω), (8.64)
l
c,eq
where gX are the equilibrium Green’s functions of the lead X, i.e. the
usual lead Green’s function for us. With these expressions, it is straight-
forward to show that the self-energies, like the ones in Eq. (8.58), and the
corresponding scattering rates are related to the corresponding equilibrium
quantities as follows
X c(l) X (l)
[Σ̂cX ]m,n = [Σ̂X ]m,n , [Γ̂X ]m,n = [Γ̂X ]m,n , (8.65)
l l
current in Eq. (8.57), in which the two time arguments are equal, has the
following time dependence
X
I(t) = Im eimωt , (8.68)
m
i.e. as anticipated, it oscillates with the external frequency and all its har-
monics. We are only interested here in the dc component, I0 , which from
now on we will simply denote as I.
Using the generic Fourier expansion of Eq. (8.61) for all the quantities
appearing in the current expression, see Eq. (8.57), it is easy to show that
the dc current can be written in terms of the different Fourier components
in energy space defined above as
8e ∞ X
Z n o
(n) (n′ ) (n′ ) (n)
I= Tr [Ĝr ]0,k [Γ̂R ]k,l [Ĝa ]l,m [Γ̂L ]m,0 (fL − fR ),
h −∞
k,l,m,n,n
′
(8.69)
(n) dc
where fX (E) = f (E − UX+ n~ω). At this stage it is already obvious
that in the absence of an ac field, this formula reduces to the Landauer
formula derived in section 8.1. We can write the current in numerous
ways by changing summation indices and the integration variable. Thus
for instance, it is not difficult to show that the dc current can be expressed
as follows12
∞ Z ∞
2e X (k)
I(V ; α, ω) = dE [TRL (E, V ; α, ω)fL (E) − (8.70)
h −∞
k=−∞
(k)
TLR (E, V ; α, ω)fR (E)],
dc
where fX (E) = f (E − UX ), the parameter α = αL − αR = eVac /~ω is
the strength of the ac drive and the coefficients appearing inside the energy
integral are given by
(k) (k) (0)
TRL (E) = 4Trω [Ĝr (E)Γ̂R (E)Ĝa (E)Γ̂L (E)], (8.71)
(k) a (k) (0)
TLR (E) = 4Trω [Ĝ (E)Γ̂L (E)Ĝr (E)Γ̂R (E)], (8.72)
where trace Trω includes a summation over the “harmonic” indexes, i.e.
over the Fourier components in energy space, and over the usual site and
(k)
orbital indexes of the central region. Here TRL (E) can be interpreted as
a transmission coefficient that describes processes taking an electron from
left (L) to right (R), under the absorption of a total of k energy quanta
12 For the sake of clarity, we make explicit the dependence of the current on the dc
voltage, V , the frequency, ω, and the strength of the ac drive, α = αL − αR = eVac /~ω.
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
(k)
~ω. The coefficient TLR (E) has a similar interpretation. By the way,
these interpretations are the reason why one usually talks about photo-
assisted processes in this problem, although there is indeed no quantized
electromagnetic field interacting with the conduction electrons in our model.
Let us summarize the discussion above. The current of Eq. (8.70) de-
scribes the dc current in the presence of an oscillating potential and it
adopts a form similar to the standard Landauer formula. The main differ-
ence is that all the quantities have now a matrix structure in an extended
Hilbert space, which includes both the orbital and the energy space. The
appearance of off-diagonal elements in energy space is a natural conse-
quence of the occurrence of the inelastic processes that take place in this
problem. In those inelastic tunneling processes, a certain number of en-
ergy quanta (multiples of ~ω) can be either absorbed or emitted. The
retarded/advanced Green’s functions appearing in the current formula are
determined by solving the matrix equation (8.67), while the scattering rates
are given by Eq. (8.66). All these matrices have, in principle, an infinite
dimension in energy space, but they can be truncated in practice and their
actual dimension is governed by the amplitude of the ac drive, α.
The formalism above has been recently used to discuss both the photon-
assisted transport in atomic [221] and molecular wires [222]. This formalism
is a bit cumbersome and numerically demanding due to the large size of
the matrices involved. However, the current formula above can be greatly
simplified in the case in which we can ignore the energy dependence in the
leads, which is frequently a very good approximation. In this situation the
self-energies Σ̂X become diagonal (see Exercise 8.8)
[Σ̂X ]n,m (E) = ΣX (E)δn,m . (8.73)
If in addition we assume that the ac potential profile is such that it is
constant in the central region (i.e. the drops occur at the interfaces), the
current formula reduces to [165, 211, 214]
∞
2e X h ³ α ´i2
Z
I(V ; α, ω) = Jl dE T (E + l~ω)[fL (E) − fR (E)], (8.74)
h 2
l=−∞
0.6 α = 0.0
I(V,α,ω)/(eω/π)
G(V,α,ω)/G0
0.4 α = 2.0
0.4 α = 4.0
α = 0.0 0.2
0.2 α = 1.0
α = 2.0
α = 4.0
0 0
6 8 10_ 12 14 6 8 10_ 12 14
eV/hω eV/hω
G(V=0,α,ω)/G(ω=0)
1
200 α = 1.0 α = 0.0
G(V=0,α,ω)/G0
(c) α = 2.0 0.8 (d) α = 1.0
150 α = 4.0 α = 2.0
0.6 α = 4.0
100
0.4
50 0.2
0 0
0 0.5 _ 1 1.5 -3 -2 -1 0 _1 2 3
hω/ε0 (ε0-EF)/hω
Fig. 8.5 Photon-assisted transport in the resonant tunneling model. In this example we
consider a symmetric junction with ΓL = ΓR = 0.1~ω and all the results are obtained at
zero temperature. (a) Current as a function of the dc bias voltage V for ǫ0 − EF = 5~ω
and different values of α = eVac /~ω. (b) The differential conductance corresponding to
the I-V curves of panel (a). (c) Photoconductance normalized by the conductance in the
absence of radiation as a function of the radiation frequency for different values of α.
(d) Photoconductance versus level position measured with respect to the Fermi energy.
8.4 Exercises
8.1 Scattering rate matrices: Show that the scattering rate matrices defined
in section 8.1 as ΓX = Im{ΣX } (X = L, R), where ΣX are the self-energies of
Eq. (8.16), are positive definite and therefore their square roots are well-defined.
8.2 Transmission matrix: The goal of this exercise is to show that the matrix
defined in Eq. (8.19) has indeed the basic properties of a transmission matrix. For
this purpose, it must be shown that the eigenvalues of tt† are bounded between
0 and 1. Demonstrate this property following the next steps:
(i) Using the result of the previous exercise, show that tt† is positive definite
and therefore all its eigenvalues are real and positive.
(ii) Use the definition of the scattering rate matrices and Dyson’s equation
for the retarded and advanced Green’s functions to prove the following relation
i
GrCC [ΓL + ΓR ] GaCC = [GrCC − GaCC ] .
2
1 = rr† + tt† ,
1/2 1/2
r = 1 − 2iΓL GrCC ΓL .
(iv) Using this last relation, show that the all eigenvalues of tt† are less than
(or equal to) one.
8.3 Formula for the current through an atomic chain: Consider the model
for an atomic chain described in Exercise 7.5. Use the general expression of
Eq. (8.20) to re-derive the formula for the electrical current obtained in that
exercise.
8.4 Phonon Green’s functions: The phonon Green’s functions are defined in
analogy with the electronic ones as
h i h i
Dr (t, t′ ) = −iθ(t − t′ )h A(t), A† (t′ ) i, Da (t, t′ ) = −iθ(t′ − t)h A(t), A† (t′ ) i,
D+− (t, t′ ) = ihA† (t′ )A(t)i, D−+ (t, t′ ) = −ihA(t)A† (t′ )i,
where A = b + b† and the creation and annihilation operators b† and b satisfy
the bosonic commutation relations (see Appendix A). Show that for the case of
a free phonon (or vibration) mode, described by the Hamiltonian of Eq. (8.39),
these functions are by given Eq. (8.45). Hint: Compute first the time evolution
of the bosonic operators by solving the equation of motion of an operator in the
Heisenberg picture.
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
Here, the subindex c indicates that the Green’s functions can be any of the four
components in Keldysh space depending upon where the time arguments, tα and
tβ (α, β = +, −), lie on the Keldysh contour. The integrations above have to be
understood as follows
Z Z ∞ Z ∞
dti = dti,+ − dti,− .
c −∞ −∞
(ii) Apply Wick’s theorem to the previous expression and keep only the con-
tributions of topologically distinct connected diagrams. Show that the only two
relevant self-energy diagrams are the ones shown in Fig. 8.6.
(iii) Evaluate the contribution of the diagrams of Fig. 8.6 to the different com-
ponents of the self-energy in energy space. Show that the contributions coming
from the diagram on the left hand side lead to the results of Eq. (8.43). Discuss
also the relevance of the contributions coming from the other diagram.
8.6 Signature of a vibrational mode in the differential conductance:
Consider the model used in section 8.2.1 to understand the signature of a vibration
mode in the current through a single resonant level. Assume that the density of
states in that level and the corresponding transmission are energy-independent
and show that the zero-temperature differential conductance is given by Eq. (8.49)
in the case of a symmetric junction.
Hint: The only complicated term in the expression for the current is the elastic
correction, which contains the self-energy Σre−vib . Separating the contributions
of the real and imaginary part of the retarded phonon Green’s function Dr , this
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
The first term, which has to be understood as principle value, does not contribute
to the conductance in the case of a symmetric junction, while the others are
responsible for the contribution of the elastic correction to Eq. (8.49).
8.7 The Meir-Wingreen formula:
(i) Follow the steps indicated in section 8.2.2 to show that the current through
an interacting region is given by Eq. (8.52).
(ii) Show that the current given by Eq. (8.52) vanishes in equilibrium.
(iii) Demonstrate that in the noninteracting case the Meir-Wingreen formula
of Eq. (8.52) reduces to the Landauer formula derived in section 8.1.
(iv) Assume that the scattering rates fulfill ΓL (E) = λΓR (E) and prove that
the Meir-Wingreen formula adopts the form given in Eq. (8.53).
8.8 Photo-current formula in the wide-band approximation:
(i) Show that the general formula of Eq. (8.70) for the current in a nanocontact
under an ac field reduces to Eq. (8.74) when (i) the energy dependence of the
density of states in the leads can be neglected (wide-band approximation) and
(ii) the ac potential is assumed to be flat in the central region of the system.
(ii) Starting from Eq. (8.75), show that in the limit of α ≪ 1 and small
frequencies, the conductance correction induced by the ac drive is a measure of the
second derivative of the transmission around the Fermi energy, i.e. demonstrate
Eq. (8.76).
Hint: Use the following properties of Bessel’s functions
∞
X ∞
X
Jn+l (x)Jm+l (x) = δnm , [Jl (x)]2 = 1,
l=−∞ l=−∞
(±x/2)l (±x/2)l+2
Jl (x ≪ 1, l > 0) ≈ − .
l! (l + 1)!
tures by combining the Meir-Wingreen formula of Eq. (8.53) and the single-level
Anderson model of Eq. (5.109). For this purpose, carry out the following tasks:
(i) Adapt the Meir-Wingreen formula to the case of a single-level Anderson
model and derive an expression for the linear conductance in terms of the spectral
function in the resonant level.
(ii) Use the approximation of Eq. (5.112) to compute the spectral function in
the weak coupling limit.
(iii) Combine the results of (i) and (ii) to obtain the gate voltage and temper-
ature dependence of the linear conductance and show that this model reproduces
the two signatures described above.
Hint: This problem was addressed by Meir et al. in Ref. [632].
8.10 Kondo effect in molecular transistors: Unitary limit. The Kondo
effect in molecular junctions is manifested in the appearance of a pronounced
resonance in the density of states at the Fermi energy. This many-body effect is
usually described with the help of the Anderson model (see section 6.9). Apply
the Meir-Wingreen formula to this model and show that in the Kondo regime the
low-temperature linear conductance in a symmetric junction (ΓL = ΓR ) is equal
to the conductance quantum (G0 ). This is referred to as the unitary limit. Hint:
Use the Friedel sum rule discussed in section 6.9.1.
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
Chapter 9
The main idea of the tight-binding approach was already introduced in Ap-
pendix A and indeed it has been extensively used in the previous chapters
devoted to the Green’s function techniques. Anyway, let us now define
237
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
where |iαi denotes the state that corresponds to the localized orbital α
that is centered around Ri , i.e. hr|iαi = φiα (r) = φα (r − Ri ). This generic
form for the Hamiltonian implies that either the many-body interactions
such as the electron-electron interaction are neglected or they are taken
into account in a mean field manner by an appropriate choice of the matrix
elements. In the former case, the matrix elements are rigorously defined as
~2 2
Z · ¸
∗
Hiα,jβ = dr φα (r − Ri ) − ∇ + V (r) φβ (r − Rj ), (9.2)
2m
where V (r) is the potential that describes the Coulomb interaction between
the electrons and ions. Finally, in the tight-binding approach, as it is used
in this book, the matrix elements are not determined from first principles,
i.e. from a direct evaluation of the integral in Eq. (9.2), but they are used
merely as parameters that may be derived approximately or may be fitted
to experiment or to other theories. Thus, by tight-binding model we mean
here a model in which the system is described in terms of a single-particle
Hamiltonian written in a local basis, the elements of which are determined
in a empirical or semi-empirical way. The different tight-binding models
differ in the way in which these parameters are obtained.
There are two situations where the wave function associated to a tight-
binding model can be determined in a straightforward manner. The first
one corresponds to the case of a small finite system such as a molecule and
1 This Hamiltonian in our usual second quantization language reads
Hiα,βj c†iα cjβ .
X
H=
ij,αβ
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
This leads immediately to the following set of equations for the coefficients
(see Exercise 9.1)
X
[Hiα,jβ − ESiα,jβ ] ciα,jβ = 0, (9.4)
jβ
is the overlap between the states |iαi and |jβi. Here, we have taken into
account the possibility that the localized orbitals centered in different atoms
can be non-orthogonal. These equations have non-trivial solutions if
det (H − ES) = 0, (9.6)
where the symbol “det” denotes the determinant of the matrix appearing
inside the brackets. The roots of this secular equation yield the eigenen-
ergies or energy levels of the finite problem and the eigenfunctions are the
corresponding waves functions (or molecular orbitals) of this system. The
dimension of the matrices in Eq. (9.6) is simply the total number of local-
ized orbitals in the problem and therefore, the solution of the generalized
eigenvalue problem of Eq. (9.6) requires typically to resort to numerics.
In the case of an infinite periodic system, typical of solid state physics,
one can diagonalize the Hamiltonian making use of Bloch’s theorem (see
for instance Ref. [223]). The idea goes as follows. Consider a periodically
replicated unit cell, where the lattice vectors are denoted as Rm , with a set
of atoms i located at positions bi in each unit cell. Associated with each
atom is a set of atomic-like orbitals φiα , where α denotes both the orbital
and angular quantum number of the atomic state. The Hamiltonian can
be easily diagonalized in reciprocal space as follows. We first construct the
following wavefunctions (Bloch sums)
1 X
Φkiα (r) = √ exp(ik · Rn )φiα (r − Rn − bi ), (9.7)
N n
where k is the Bloch wave vector, which is restricted to the Brillouin zone,
and N is the number of unit cells in the sum. The solution to Schrödinger
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
equation for wave vector k then requires the diagonalization of the Hamil-
tonian matrix using the basis functions of Eq. (9.7). Since the Hamilto-
nian has the periodicity of the lattice, this basis will block-diagonalize the
Hamiltonian, with each block having a single value of k. Within one of
these blocks, the matrix elements can be written in the form
X Z
Hiα,jβ (k) = exp(ik · Rn ) φ∗iα (r − Rn − bi )Hφjβ (r − bj )d3 r, (9.8)
n
where we have used the translation symmetry of the lattice to remove one of
the sums over the lattice vector R (see Exercise 9.2). In the same way, one
can also define the overlap matrix in reciprocal space where the different
elements adopt the form
X Z
Siα,jβ (k) = exp(ik · Rn ) φ∗iα (r − Rn − bi )φjβ (r − bj )d3 r. (9.9)
n
The corresponding secular equation reads this time
det (H(k) − ES(k)) = 0. (9.10)
The solution of this generalized eigenvalue problem yields the different en-
ergy bands, ǫµ (k) of the solid and the corresponding eigenvectors Qµ (k).
Notice that the number of bands, i.e. the number of solutions of Eq. (9.10),
equals the number of atoms in the unit cell times the number of orbitals per
atom. Thus, in some simple cases the solution can be found analytically
and, in general, this problem can be easily solved numerically.
An important quantity for many purposes is the density of states (DOS)
per unit energy E. The local DOS projected onto a given atom, orbital and
spin (summarized by the index ν) is defined in terms of the energy bands
ǫµ (k) as follows
1 X
ρν (E) = |Qν,µ (k)|2 δ(ǫµ (k) − E) (9.11)
Nk
k,µ
Ωcell X
Z
= dk |Qν,µ (k)|2 δ(ǫµ (k) − E),
(2π)d µ BZ
where BZ denotes the Brillouin zone, Ωcell is the volume of the unit cell
and d is the dimensionality of the system.
In the case of infinite non-periodic systems, like the atomic-scale junc-
tions that we are interested in, the determination of the wavefunction is
literally impossible. However, the use of the Green’s function techniques
described in Chapter 5 allows to extract most of the relevant information
about the electronic structure from a tight-binding Hamiltonian.
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
where r = |r|, r̂ = r/r and n indicates different functions with the same
angular momentum. We shall work frequently with real basis functions √ that
+ ∗
can be defined using the√ real angular functions S lm = (Ylm + Y lm )/ 2 and
− ∗
Slm = (Ylm − Ylm )/(i 2). The examples of real s (l = 0), p (l = 1) and
d (l = 2) orbitals are given in Fig. 9.1. The analytical expressions of the
angular dependence of these real orbitals can be found in many textbooks,
see e.g. Chapter 1 of Ref. [224] or Chapter 3 of Ref. [227].
The key problem in a tight-binding model is the determination of the
matrix elements (or integrals) that appear both in Eq. (9.8) and Eq. (9.9).
Those matrix elements can be divided into one-, two-, and three-center
terms. The simplest is the overlap matrix in Eq. (9.9), which involves only
one center if the two orbitals are on the same site and two centers otherwise.
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
m=0 m =+
−1
m=0 m =+
−1 m =+
−2
Fig. 9.1 Boundary surfaces for real s-, p-, and d-orbitals. The index m indicates the
quantum number corresponding to the z-component of the orbital angular momentum.
~2 2 X
H=− ∇ + Vk (r − Rn − bk ), (9.14)
2m
nk
where the first term is the usual kinetic energy and the second is the po-
tential decomposed into a sum of spherical terms centered on each site k in
the unit cell. The kinetic part of the Hamiltonian matrix element always
involves one or two centers. However, the potential terms may depend upon
the positions of other atoms; they can be divided into the following.
• One-center, where both orbitals and the potential are centered on the
same site. These terms have the same symmetry as an atom in free
space.
• Two-center, where the orbitals are centered on different sites and the
potential is on one of the two. These terms have the same symmetry
as other two-center terms.
• Three-center, where the orbitals and the potential are all centered on
different sites. These terms can also be classified into various symme-
tries based upon the fact that three sites define a triangle.
• A special class of two-center terms with both orbitals on the same site
and the potential centered on a different site. These terms add to the
one-center terms above, but depend upon the crystal symmetry.
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
ppσ
ppπ
pdσ
pdπ
ddπ ddδ
ddσ
Fig. 9.2 The 10 irreducible SK-parameters for the s, p and d orbitals, which are classi-
fied by the angular momentum about the axis with the notation σ (m = 0), π (m = 1)
and δ (m = 2). The orbitals shown are the real combinations of the angular momen-
tum eigenstates. Positive and negative lobes are denoted by solid and dashed lines,
respectively.
z
y
spx sp σ
x
R R
ppσ
px pz R
R ppπ
In Fig. 9.2 we show the orbitals for the non-zero σ, π, and δ matrix
elements for s, p, and d orbitals. The orbitals shown are actually the real
±
basis functions Slm defined as combinations of the ±m angular momentum
eigenstates. These are oriented along the axes defined by the line between
the neighbors and two perpendicular axes. All states except the s state
have positive and negative lobes. Note that states with odd l are odd under
inversion. Their sign must be fixed by convention (typically one chooses
the positive lobe along the positive axis). The direction of the displacement
vector is defined to lie between the site denoted by the first index and that
denoted by the second index. For example, in Fig. 9.2, the Kspσ matrix
element in the top center has the negative lobe of the p function oriented
toward the s function. Interchange of the indices leads to Kpsσ = −Kspσ
′
and, more generally, to Kll′ m = (−1)l+l Kl′ lm .
An actual set of basis functions is constructed with the quantization
axis fixed in space, so that the functions must be transformed to utilize the
standard irreducible form of the matrix elements. Examples of two-center
matrix elements of s and pi = px , py , pz orbitals for atoms separated by
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
Table 9.1 Table of two-center matrix elements for either the overlap or the Hamil-
tonian, with real orbitals s and px , py , pz . The vector R between sites, as shown
in Fig. 9.3, is defined to have direction components R̂ ≡ x, y, z. The matrix el-
ements are then expressed in terms of these coordinates and the four irreducible
matrix elements: Kssσ , Kspσ , Kppσ and Kppπ . Other matrix elements can be
found by permuting elements.
Element Expression
Ks,s Kssσ
Ks,px xKspσ
Kpx ,px x2 Kppσ + (1 − x2 )Kppπ
Kpx ,py xy(Kppσ − Kppπ )
Kpx ,pz xz(Kppσ − Kppπ )
the displacement vector R are shown in Fig. 9.3. Each of the orbitals on
the left-hand side can be expressed as a linear combination of orbitals that
have the standard form oriented along the rotated axes, as shown on the
right. An s orbital is invariant and a p orbital is transformed to a linear
combination of p orbitals. The only non-zero matrix elements are the σ
and π matrix elements, as shown. The top row of the figure illustrates the
transformation of the px orbital needed to write the matrix element Ks,px
in terms of Kspσ and the bottom row illustrates the relation of Kpx ,pz to
Kppσ and Kppπ . Specific relations for all s and p matrix elements are given
in Table 9.1. Expressions for d orbitals are given in Refs. [234, 224, 225].
Now we are in position to describe the Slater and Koster approach [234].
These authors proposed that the Hamiltonian matrix elements can be ap-
proximated with the two-center form and fitted to theoretical calculations
(or empirical data) as a simplified way of describing and extending cal-
culations of electronic bands. Within this approach, all matrix elements
have the same symmetry as for two atoms in free space (see Fig 9.3 and
Table 9.1). This is a great simplification that leads to an extremely useful
approach to understanding electrons in materials.
Slater and Koster gave extensive tables for matrix elements, including
the s and p matrix elements given in Table 9.1. In addition, they presented
expressions for the d states and analytical formulas for bands in several
crystal structures. Examples of the latter are presented in the next sec-
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
tion to illustrate the useful information that can be derived. However, the
primary use of the SK approach in electronic structure has become the de-
scription of complicated systems, including the bands, total energies, and
forces for relaxation of structures and molecular dynamics. These applica-
tions have very different requirements that often lead to different choices of
SK parameters.
For the bands, the parameters are usually designed to fit selected eigen-
values for a particular crystal structure and lattice constant. For example,
the extensive tables derived by Papaconstantopoulos [235] are very useful
for interpolation of results of more expensive methods. It has been pointed
out by Stiles [236] that for a fixed ionic configuration, effects of multi-center
integrals can be included in two-center terms that can be generated by an
automatic procedure. This makes it possible to describe any band struc-
ture accurately with a sufficient number of matrix elements in SK form.
However, the two-center matrix elements are not transferable to different
structures.
On the other hand, any calculation of total energies, forces, etc., requires
that the parameters be known as a function of the position of the atoms.
Thus, the choices are usually compromises that attempt to fit a large range
of data. Such models are fit to structural data and, in general, are only
qualitatively correct for the bands. Since the total energy depends only
upon the occupied states, the conduction bands may be poorly described
in these models. Of particular note, Harrison [224, 225] has introduced a
table that provides parameters for any element or compound. The forms
are chosen for simplicity, generality, and ability to describe many properties
in a way that it is instructive and useful. The basis is assumed to be
orthonormal, i.e. Smm′ = δmm′ . The diagonal Hamiltonian matrix elements
are given in a table for each atom. Any Hamiltonian matrix element for
orbitals on neighboring atoms separated by a distance R is given by a factor
′
times 1/R2 for s and p orbitals and 1/Rl+l for l > l′ .
Many other SK parameterizations have been proposed, each tailored
to particular elements and compounds. Some additional examples can be
found in Chapter 14 of Ref. [226].
Let us illustrate the tight-binding approach with the analysis of some simple
situations.
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
4 ε0+ 2t
(a) (b) (6)
−t −t
5
3
(4)
ε0+ t ε0+ t (5)
−t −t Energy
(2)
ε0− t ε0− t (3)
2 6
−t −t ε0− 2t
(1)
1
Fig. 9.4 (a) Schematic representation of the Hückel model for the benzene molecule,
as described in the text. (b) Energy level diagram of benzene as obtained from this
approximation. The levels are labeled from 1 to 6 following Eq. (9.16). We also show
charge-density plots of the molecular orbitals obtained from a density-functional-theory
calculation to show that indeed the Hückel approximation reproduces the character of
the orbitals, see Eq. (9.17). The two colors indicate different signs of the wavefunctions.
The ground state is obtained by doubly occupying the three lowest energy levels.
0.3 0.2
1 (a) 1D (b) 2D (c) 3D
DOS (1/t)
0.2
0.1
0.5
0.1
0 0 0
-2 -1 0 1 2 -4 -2 0 2 4 -6 -4 -2 0 2 4 6
E/t E/t E/t
Fig. 9.5 Density of states per spin (DOS) vs. energy for an s-band in (a) a one-
dimensional line, (b) a two-dimensional square, and (c) a three dimensional simple cubic
lattice with nearest neighbor interactions t.
tice with spacing a the general expressions (9.8) and (9.10) reduce to (see
Exercise 9.4)
The bands for the square lattice in the xy-plane are given by this expression,
omitting the kz term; for a line (or chain) in the x-direction, only the kx
term applies. From this expression one can easily deduce several interesting
consequences. First, the bands are symmetric about ǫ(k) = 0 in the sense
that every state at +ǫ has a corresponding state at −ǫ. This can be seen in
Fig. 9.5, where we show the density of states (DOS) for one, two and three
dimensions. The shapes can be found analytically in this case (see Exercise
9.6). Notice that the bandwidth is determined by the hopping element t.
In the case of a metal, this parameter has a value of around 1 eV.
In the square lattice, the energy ǫ(k) = 0 at a face zone k = (π/a, 0).
This is a saddle point since the slope vanishes and the bands curve upward
and downward in different directions. This leads to a density of states with
a logarithmic divergence at ǫ = 0. Furthermore, for a half-filled band (one
electron per cell), the Fermi surface is at ǫ(k) = 0. This leads to the result
that the Fermi surface is a square rotated by π/4 with half the volume of
the Brillouin zone, and the density of states diverges at ǫ = EF as shown
in Fig. 9.5(b). If there are second-neighbor interactions, the symmetry of
the bands in ±ǫ is broken and the Fermi surface is no longer square.
Let us assume now that the states are no longer orthogonal, but the
overlap between nearest neighbors is equal to s. Then the solution for the
bands, Eq. (9.18), is generalized to (Exercise 9.7)
H(k) 2t [cos(kx a) + cos(ky a) + cos(kz a)]
ǫ(k) = = . (9.19)
S(k) 1 + 2s [cos(kx a) + cos(ky a) + cos(kz a)]
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
and a is the lattice constant. This is readily solved to yield the bands
h √ i1/2
ǫ(k) = ±t 1 + 4 cos( 3ky a/2) cos(kx a/2) + 4 cos2 (kx a/2) . (9.22)
The most remarkable feature of the graphene bands is that they touch at
the corners of the hexagonal Brillouin zone, e.g. the points denoted K± =
(kx = ±4π/3a, ky = 0). Note also that the bands are symmetric in ±ǫ.
Since there is one π electron per atom, the band is half-filled and the bands
touch with finite slope at the Fermi energy, i.e. a Fermi surface consisting of
points. Indeed, one can show (see Exercise 9.8) that the dispersion relation
around the points K± (Dirac points) is √ linear, i.e. ǫ(q) = ~v|q|, where
q = k − K± with a velocity given by v = ( 3a/2~)t. This linear dispersion
relation resembles that of Dirac’s massless fermions and it is the origin of
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
0.3
a1
DOS
0.2
a2
0.1
0
-3 -2 -1 0 1 2 3
E/t
Fig. 9.6 (a) Honeycomb lattice for a graphene sheet: the lattice is triangular and there
√
are two atoms per unit cell. Two primitive vectors are ~a1 = a(1, 0) and ~a2 = a/2(1, − 3),
where a is the lattice constant. (b) Local density of states (per spin) projected onto an
atom of the unit cell as a function of the energy normalized by the hopping parameter t.
Notice that the DOS vanishes at E = 0 and there are van Hove singularities at E = ±t.
more practical information about this parameterization, visit the web page: http://cst-
www.nrl.navy.mil/bind/.
4 The only exception was the method put forward by Harrison that was briefly mentioned
in section 9.4.
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
Note that V0 depends upon the structure of the crystal, as well as the
original method for determining the energy zero. Notice also that the ǫ′ (k)
are in some sense “universal”. That is, if any two band structure methods
are sufficiently well converged, they will give the same total energy, and the
eigenvalues derived from the two methods will differ by only a constant.
Then the definition of V0 for each method will be such that the shifted
eigenvalues ǫ′ (k) are identical.
In the NRL method, the authors construct a first-principles5 database of
eigenvalues ǫ(k) and total energies E for several crystal structures at several
volumes. Then, they find V0 for each system, and shift the eigenvalues.
Next, they attempt to find a set of parameters which will generate non-
orthogonal, two-center SK Hamiltonians which will reproduce the energies
and eigenvalues in the database.
Let us now describe how the TB parameters for elemental systems are
constructed. One assumes that the on-site terms are diagonal and sensitive
to the environment. For single-element systems one assigns atom i in the
crystal an embedded-atom-like “density”
X
ρi = exp(−λ2 Rij )F (Rij ), (9.28)
j
where the sum is over all the atoms j within a range Rc of atom i; λ is the
first fitting parameter, squared to ensure that the contributions are greater
from the nearest neighbors; and F (R) is a cut-off function,
F (R) = θ(Rc − R)/ {1 + exp [(R − Rc )/l + 5]} , (9.29)
where θ(z) is the step function. Typically one takes Rc between 10.5 and
16.5 Bohr and l between 0.25 and 0.5 Bohr.
One then defines the angular-momentum-dependent on-site terms by
2/3 4/3
hil = al + bl ρi + cl ρi + dl ρ2i , (9.30)
where l = s, p, or d. These (a, b, c, d)l form the next 12 fitting parameters.
In the spirit of the two-center approximation, one assumes that the
hopping integrals depend only upon the angular momentum of the orbitals
and the distance between the atoms. As we showed in section 9.3.1, all
the two-center (spd) hopping integral can then be constructed from ten
independent parameters, the SK parameters, Hll′ m , where
(ll′ m) = ssσ, spσ, ppσ, ppπ, sdσ, pdσ, pdπ, ddσ, ddπ, and ddδ. (9.31)
5 Thefirst-principle methods used by the authors are typically the augmented plane
wave method (APW) or the linearized augmented plane wave method (LAPW).
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
where R is the separation between these atoms and F (R) is the cut-off
function defined above. The parameters (ell′ m , fll′ m , gll′ m , hll′ m ) constitute
the next 40 fitting parameters.
Since this is a non-orthogonal calculation, one must also define a set of
SK overlap functions. These represent the overlap between two orbitals sep-
arated by a distance R. They have the same angular momentum behavior
as the hopping parameters:
The parameters (pll′ m , qll′ m , rll′ m , sll′ m ) make up the final 40 fitting pa-
rameters for a monoatomic system, giving in total 93 fitting parameters
that are chosen to reproduce the contents of the first-principles database,
as noted above.6
So in summary, this parameterization uses an analytical set of two-
center integrals, nonorthogonal parameters and on-site parameters that de-
pend on the local environment. The method reproduces not only the band
structure, but also the total energy of the system. It has been demonstrated
that this method reproduces very well structural energy differences, elas-
tic constants, phonon frequencies, vacancy formation energies, and surface
energies for both transition metal and noble metals.
As an application of this parameterization, we have computed the bulk
density of states of six different metals that play an important role in molec-
ular electronics.7 The results can be seen in Fig. 9.7. Notice that in the
cases of Ag and Au (noble metals), the Fermi energy lies in the region
where the DOS is dominated by the s band. In the case of Al and Pb,
the s and p bands dominate the DOS around the Fermi energy. The main
difference between these two metals is that Pb has 4 valence electrons and
therefore, the Fermi energy lies well inside the p band. Finally, Nb and Pt
are examples of transition metals, where the d band dominates the DOS at
the Fermi energy and for this reason, the d orbitals play a fundamental in
the transport properties of these metals.
6 These parameters for many different elementary solids can be found in the following
web page: http://cst-www.nrl.navy.mil/bind/.
7 In particular, we shall analyze in Chapter 11 the conductance of single-atom contacts
3 3
Ag Au
DOS (1/eV)
5s 6s
2 5p 2 6p
4d 5d
1 1
0 0
-10 -5 0 5 10 15 -10 -5 0 5 10 15
Al 3s Pb 6s
DOS (1/eV)
0.4 0.8
3p 6p
3d
0.2 0.4
0 0
-10 -5 0 5 10 15 -10 -5 0 5 10 15
3 3
Nb 5s Pt 6s
DOS (1/eV)
2 5p 2 6p
4d 5d
1 1
0 0
-10 -5 0 5 10 15 -10 -5 0 5 10 15
E-EF (eV) E-EF (eV)
Fig. 9.7 Bulk DOS as a function of energy for Ag, Au, Al, Pb, Nb, and Pt computed
using the NRL-tight-binding parameterization. The DOS is projected onto the s, p, and
d orbitals that give rise to the bands around the Fermi energy (EF ).
In this final section we shall explain how the tight-binding approach is used
in practice to describe the transport properties of atomic-scale junctions
and we shall also review very briefly the impact of this approach in the
field of molecular electronics to date.
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
and the lack of periodicity complicates the calculation of their Green’s func-
tions. There are different solutions for this problem. For instance, one can
describe the electrodes with simple structures like Bethe lattices [246]. A
more satisfactory solution to avoid artificial interface resistances is to de-
scribe the leads as ideal surfaces and compute the Green’s functions with
recursive methods like the decimation technique of Ref. [198].8
Step 4: Computation of the current. The final step is the calculation
of the current from the knowledge of the Green’s functions, which is done
using Eqs. (8.18-8.20).
In general, the “recipe” described above has to be carried out numeri-
cally, but the computer codes can be developed by a single person in a few
weeks. Moreover, the tight-binding approach is extremely efficient, com-
putationally speaking, and the calculations of the transport properties of
realistic systems can be done in standard PC’s. Of course, the level of accu-
racy of these calculations depends on the quality tight-binding parameters,
which in turn depends on the system under study.
Also in the middle of the 1990’s, Datta and coworkers employed the tight-
binding approach to describe the current-voltage characteristics of different
organic molecules and to establish a detailed comparison with the experi-
ments [260, 261].
In the context of metallic atomic-sized contacts,10 the tight-binding for-
malism was first used, in combination with molecular dynamics simula-
tions, to elucidate the origin of the conductance jumps observed during
the formation of these nanowires [262]. Tight-binding models were then
used to establish the relation between the conduction channels of single-
atom contacts and the detailed chemistry of the metal atoms [263, 264].
The tight-binding approach has also been extended to calculate changes
to interatomic forces under electrical current flow in atomic-scale conduc-
tors [265] and this formalism has been used to model electromigration and
current-induced fracture of atomic wires [266, 267].
9.8 Exercises
9.1 Secular equation for a finite system: Use the Schrödinger equation to
show that the coefficients of the expansion of Eq. (9.3) satisfy the set of equations
given by Eq. (9.6).
9.2 Bloch’s theorem: Using the translational invariance in a Bravais lattice,
show that matrix elements of the Hamiltonian with basis functions Φkiα and
Φk′ jβ are non-zero only for k = k′ , and derive the expression of Eq. (9.8).
9.3 Energy spectrum of benzene: Solve analytically the secular equation
(9.15) and show that within the Hückel approximation the energy levels and the
corresponding molecular orbitals are given by Eq. (9.16) and Eq. (9.17), respec-
tively.
9.4 Molecular orbital structure of butadiene: The 1,3-butadiene molecule
(C4 H6 ) shown in Fig. 9.8 is a simple conjugated diene, i.e. it is a hydrocarbon
which contains two double bonds. Determine its energy levels and molecular
orbitals using the Hückel approximation.
9.5 Energy bands of s-bands in line, square and cubic Bravais lattices:
Show that for an s-band in a line, square lattice, and simple cubic lattice with
only nearest neighbor Hamiltonian matrix elements, the energy bands are given
by Eq. (9.18).
9.6 Density of states of s-bands in line, square and cubic Bravais lat-
tices: Reproduce the results of Fig. 9.5 for the density of states of s-band in a line,
square lattice, and simple cubic lattice with only nearest neighbor Hamiltonian
matrix elements.
9.7 Energy bands of s-bands in line, square and cubic Bravais lattices
in a non-orthogonal model: Show that the expression for bands with non-
orthogonal basis orbitals, Eq. (9.19), is correct. Why are the bands in this case
no longer symmetric about ǫ = 0?
9.8 Electronic structure of graphene: Consider the model for graphene de-
tailed in section 9.5.3. Carry out the following tasks: (i) Determine the Brillouin
zone of the honeycomb lattice, (ii) show that the energy bands are given by
Eq. (9.22), (iii) demonstrate that the dispersion relation around the Dirac’s point
is linear, and (iv) compute the local density of states and show that it is given
by the result of Fig. 9.6(b).
9.9 The NRL tight-binding method: An interesting project for graduate
students and advanced undergraduate students is to write a computer code (in
whatever language) to calculate the energy bands and bulk density of states [see
Fig. 9.7] of elementary solids within the NRL tight-binding method described in
section 9.6.
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
Chapter 10
263
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
and Kohn’s Nobel lecture [270]. Among the DFT reviews, we recommend
the one of Ref. [271]. Finally, let us say that this chapter is based on the
monographs of Refs. [272, 226] and specially on that of Ref. [273]. Those
readers familiar with DFT who only want to know how it is applied in
molecular electronic are advised to jump directly to the last section of this
chapter.
In order to pave the way for the understanding of the basic formulation of
density function theory, we start reminding some basic issues in quantum
mechanics.1
charges in units of the electron charge, ~ is the unit of action, the energy is measured in
Hartrees (27.211 eV) and the length unit is the Bohr (0.52910 Å).
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
The Schrödinger equation [Eq. (10.1)] can be simplified using the Born-
Oppenheimer approximation. Due to their masses the nuclei move much
slower than the electrons, which implies that we can consider the electrons
as moving in the field of fixed nuclei, i.e. the nuclear kinetic energy is
zero and their potential energy is merely a constant. Thus, the electronic
Hamiltonian reduces to
N N M N N
1 X 2 X X ZA X X 1
Helec = − ∇i − + = T + VN e + Vee . (10.3)
2 i=1 i=1
riA i=1 j>i rij
A=1
In principle, our main problem now is to solve Eq. (10.4), which is simply
impossible to accomplish in general.4
For a system of N electrons and given nuclear potential Vext , the vari-
ational principle defines a procedure to determine the ground-state wave
function Ψ0 , the ground-state energy E0 [N, Vext ], and other properties of
interest. In other words, the ground state energy is a functional of the
number of electrons N and the nuclear potential Vext
E0 = E[N, Vext ]. (10.8)
where
· ¸
1
Z
Hi ≡ ψi∗ (~x) − ∇2 + Vext (~x) ψi (~x) d~x (10.12)
2
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
defines the contribution due to the kinetic energy and the electron-nucleus
attraction and
1 ∗
Z Z
Jij = ψi (~x1 )ψi∗ (~x1 ) ψ (~x2 )ψj (~x2 )d~x1 d~x2 , (10.13)
r12 j
1
Z Z
Kij = ψi∗ (~x1 )ψj (~x1 ) ψi (~x2 )ψj∗ (~x2 )d~x1 d~x2 . (10.14)
r12
The integrals are all real, and Jij ≥ Kij ≥ 0. The Jij are called Coulomb
integrals, the Kij are called exchange integrals. We have the property Jii =
Kii .
The variational freedom in the expression of the energy [Eq. (10.11)] is in
the choice of the orbitals. TheR minimization of the energy functional with
the normalization conditions ψi∗ (~x)ψj (~x)d~x = δij leads to the Hartree-
Fock differential equations (see Exercise 10.2)
f ψi = ǫi ψi , i = 1, 2, ..., N. (10.15)
These N equations have the appearance of eigenvalue equations, where
ǫi are the eigenvalues of the operator f . The Fock operator f is an effective
one-electron operator defined as
M
1 X ZA
f = − ∇2i − + VHF (i). (10.16)
2 riA
A
The first two terms are the kinetic energy and the potential energy due
to the electron-nucleus attraction. VHF (i) is the Hartree-Fock potential,
the average repulsive potential experienced by the i-th electron due to the
remaining N -1 electrons, and it is given by
N
X
VHF (~x1 ) = (Jj (~x1 ) − Kj (~x1 )) , (10.17)
j
1
Z
Jj (~x1 ) = |ψj (~x2 )|2 d~x2 . (10.18)
r12
The Coulomb operator J represents the potential that an electron at
position ~x1 experiences due to the average charge distribution of another
electron in spin orbital ψj .
The second term in Eq. (10.17) is the exchange contribution to the HF
potential. It has no classical analog and it describes the modification of the
energy that can be ascribed to the effects of spin correlation. It is defined
through its effect when operating on a spin orbital
1
Z
Kj (~x1 ) ψi (~x1 ) = ψj∗ (~x2 ) ψi (~x2 ) d~x2 ψj (~x1 ). (10.19)
r12
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
In this section we shall introduce the electron density, which is the funda-
mental quantity in DFT, and we shall briefly review some early attempts
to develop a density functional theory.
The electron density is defined as the integral over the spin coordinates
of all electrons and over all but one of the spatial variables (~x ≡ ~r, s)
Z Z
ρ(~r) = N ... |Ψ(~x1 , ~x2 , ..., ~xN )|2 ds1 d~x2 ...d~xN . (10.20)
The electron density ρ(~r) determines the probability of finding any of the N
electrons within volume element d~r. Clearly, ρ(~r) is a non-negative function
of only the three spatial variables which vanishes at infinity and integrates
to the total number of electrons, i.e.
Z
ρ(~r → ∞) = 0; ρ(~r)d~r = N. (10.21)
This functional was combined with the classical expression for the electron-
nuclei potential and the electron-electron potential to write down the fol-
lowing functional for the energy of an atom
3
Z
2 2/3
ET F [ρ(~r)] = (3π ) ρ5/3 (~r)d~r
10
ρ(~r) 1 ρ(~r1 )ρ(~r2 )
Z Z Z
−Z d~r + d~r1 d~r2 . (10.23)
r 2 r12
Notice that the energy is given completely in terms of the electron density.
In order to determine the correct density to be included in Eq. (10.23),
they employed a variational principle. They assumed that the ground state
of the system is connected
R to the ρ(~r) for which the energy is minimized
under the constraint of ρ(~r)d~r = N . The obvious question at this stage is:
does this variational principle make sense? The Hohenberg-Kohn theorems
discussed in the next section will prove that this approach can be rigorously
justified.
R
EN e [ρ] = ρ(~r)VN e (~r)d~r, from those which are universal, FHK [ρ]. This
functional FHK [ρ] is the holy grail of density functional theory. If it was
known, we would be able to solve the Schrödinger equation exactly and for
any system. This functional contains the functional for the kinetic energy
T [ρ] and that for the electron-electron interaction, Eee [ρ]. The explicit
form of both these functionals is unknown. However, from the latter we
can extract at least the classical part J[ρ],
1 ρ(~r1 )ρ(~r2 )
Z Z
Eee [ρ] = d~r1 d~r2 + Encl = J[ρ] + Encl [ρ]. (10.26)
2 r12
Encl is the non-classical contribution to the electron-electron interaction:
self-interaction correction, exchange and Coulomb correlation. The explicit
form of the functionals T [ρ] and Encl [ρ] is the major challenge of DFT.
Let us now address the following question: how can we be sure that a
certain density is the ground-state density that we are looking for? The
second Hohenberg-Kohn theorem answers this question. This theorem
states that FHK [ρ], the functional that delivers the ground state energy of
the system, delivers the lowest energy if and only if the input density is the
true ground state density. This is nothing but the variational principle
E0 ≤ E[ρ̃] = T [ρ̃] + EN e [ρ̃] + Eee [ρ̃]. (10.27)
In other words, this means that for any trial density ρ̃(~
R r), which satisfies
the necessary boundary conditions such as ρ̃(~r) ≥ 0, ρ̃(~r)d~r = N , and
which is associated with some external potential Ṽext , the energy obtained
from the functional of Eq. (10.24) represents an upper bound to the true
ground state energy E0 . E0 results if and only if the exact ground state
density is inserted in Eq. (10.27).
Let us summarize what we have learned so far and some basic conse-
quences of the previous theorems:
• The explicit form of the functional FHK [ρ] is unknown and this remains
as the major challenge of DFT.
We have seen that the ground state energy of a system can be written as
µ Z ¶
E0 = min FHK [ρ] + ρ(~r)VN e d~r , (10.29)
ρ→N
where the universal functional FHK [ρ] contains the contributions of the ki-
netic energy, the classical Coulomb interaction and the non-classical portion
FHK [ρ] = T [ρ] + J[ρ] + Encl [ρ]. (10.30)
Of these, only J[ρ] is known. The main problem is to find the expressions
for T [ρ] and Encl [ρ]. The Thomas-Fermi model of section 10.2 provides an
example of density functional theory. However, its performance is really
bad due to the poor approximation of the kinetic energy. To solve this
problem Kohn and Sham proposed in 1965 [269] the following approach.
They suggested to calculate the exact kinetic energy of a non-interacting
reference system with the same density as the real, interacting one
N N X
1X X
TS = − hψi |∇2 |ψi i, ρS (~r) = |ψi (~r, s)|2 = ρ(~r), (10.31)
2 i i s
problem, we write down the expression for the energy of the interacting
system in terms of the separation described in Eq. (10.32)
E[ρ] = TS [ρ] + J[ρ] + EXC [ρ] + EN e [ρ], (10.34)
where
1 ρ(~r1 )ρ(~r2 )
Z Z Z
E[ρ] = TS [ρ] + d~r1 d~r2 + EXC [ρ] + VN e ρ(~r)d~r
2 r12
N N N Z Z
1X 1 XX 1
=− hψi |∇2 |ψi i + |ψi (~r1 )|2 |ψj (~r2 )|2 d~r1 d~r2
2 i 2 i j r12
N Z X
M
X ZA
+EXC [ρ] − |ψi (~r1 )|2 d~r1 . (10.35)
i
r1A
A
The only term for which no explicit form can be given is EXC . We now
apply the variational principle and ask: What condition must the orbitals
{ψi } fulfill in order to minimize this energy expression under the usual
constraint hψi |ψj i = δij ? The resulting equations are the Kohn-Sham
equations:
µ ¶
1
− ∇2 + Vef f (~r1 ) ψi = ǫi ψi , (10.36)
2
where the effective potential Vef f (~r1 ) is given by
M
ρ(~r2 ) ZA
Z X
Vef f (~r1 ) = d~r2 + VXC (~r1 ) − . (10.37)
r12 r1A
A
state with the highest eigenvalue; since the density is assumed to be exact, so must the
eigenvalue be exact. No other eigenvalue is guaranteed to be correct.
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
• The accuracy of the LDA for the exchange energy is typically within
10%, while the normally much smaller correlation energy is generally
overestimated by up to a factor 2. The two errors typically cancel
partially.
• Experience has shown that the LDA gives ionization energies of atoms,
dissociation energies of molecules and cohesive energies with a fair ac-
curacy of typically 10-20%. However, the LDA gives bond lengths of
molecules and solids typically with an astonishing accuracy of ∼ 2%.
• The moderate accuracy that LDA delivers is insufficient for most appli-
cations in chemistry. For this reason, for many years, where LDA was
the only approximation for the exchange-correlation functional, DFT
was mostly used by solid-state physicists and it hardly had any impact
in quantum chemistry.
GGA
In practice, EXC is usually split into its exchange and correlation con-
GGA GGA GGA
tributions, EXC = EX + EC , and approximations for the two terms
GGA
are sought separately. With respect to the exchange part EX , it can be
written as
X Z
GGA LDA
EX = EX − F (sσ )ρ4/3
σ (~r) d~r. (10.43)
σ=↑,↓
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
The argument of the function F is the reduced density gradient for spin σ
|∇ρσ (~r)|
sσ = 4/3
. (10.44)
ρσ
Numerous forms for the function F above have been given. We just
mention here three of the most widely used ones that were proposed by
Becke in 1986 (B86) [277], Perdew also in 1986 (P) [278], and Perdew, Burke
and Ernzerhof in 1996 (PBE) [279]. In all these cases, F is a complicated
rational function of the reduced density gradient that we shall not write
here explicitly.
The corresponding gradient-corrected correlation functionals have even
more complicated analytical forms and cannot be understood by simple
physically motivated reasoning. Among the most widely used choices is
the correlation counterpart of the 1986 Perdew exchange functional [278],
usually referred to as P or P86. This functional employs an empirical
parameter, which was fitted to the correlation energy of the neon atom. A
few years later Perdew and Wang [280] refined their correlation functional,
leading to the parameter free PW91. Another, nowadays even more popular
correlation functional is due to Lee, Yang, and Parr (LYP) [281]. This
functional was derived from an expression for the correlation energy of
the helium atom. The LYP functional contains one empirical parameter
and it differs from other GGA functionals in that it contains some local
components.
In principle, each exchange functional could be combined with any of the
correlation functionals, but only a few combinations are currently in use.
The exchange part is usually chosen to be Becke’s functional which is either
combined with Perdew’s 1986 correlation functional or the LYP one. These
combinations are termed BP86 and BLYP, respectively. Sometimes also
the PW91 correlation functional is employed, corresponding to BPW91. It
is worth stressing that these combinations lead to results that are of very
similar quality.
As a general statement about the performance of GGA-based function-
als, let us say that they have reduced the LDA errors of, in particular,
atomization energies of standard set of small molecules by a factor 3-5.
This improved accuracy has made DFT one of the most widely used tools
in quantum chemistry.
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
If we now multiply this equation from the left with an arbitrary basis
function ηµ and integrate over space we get L equations6
XL Z L
X Z
cνi ηµ (~r1 )f KS (~r1 )ην (~r1 )d~r1 = ǫi cνi ηµ (~r1 )ην (~r1 )d~r1 , (10.51)
ν=1 ν=1
where i runs from 0 to L.
The integrals on the left hand side of this equation define the Kohn-
Sham matrix, FKS , with the corresponding elements defined as
Z
KS
Fµν = ηµ (~r1 )f KS (~r1 )ην (~r1 )d~r1 , (10.52)
and on the right hand side we can identify the overlap matrix, S, the ele-
ments of which are given by
Z
Sµν = ηµ (~r1 )ην (~r1 )d~r1 . (10.53)
Up to this point, exactly the same formulas also apply in the Hartree-
Fock case. The difference is only in the exchange-correlation part. In the
Kohn-Sham scheme this is represented by the integral
Z
XC
Vµν = ηµ (~r1 )VXC (~r1 )ην (~r1 )d~r1 , (10.60)
The original motivation for contracting was that the contraction coefficients
daτ can be chosen in a way that the CGF resembles as much as possible a
single STO function. In density functional theory, CGF basis sets enjoy a
strong popularity.
A fourth type of basis functions are the numerical basis functions.
In this case, the orbitals are represented numerically on atomic centered
grids. These functions can be generated, for instance, by numerically solv-
ing the atomic KS equations with a given approximation for the exchange-
correlation functional. Obviously, in this approach the different integrals
are computed numerically.
Irrespective of the type of functions used, the basis sets can be classified
in the following simple way that already gives a hint about their quality.
The simplest (and smallest) basis functions are those that use a single basis
function for each atomic orbital up to and including the valence orbitals.
These basis sets are called, for obvious reasons, minimal basis sets. A
typical representative is the STO-3G basis set, in which three primitive
GTO functions are combined into one CGF. For carbon, this basis set
consists of five functions, one describing the 1s atomic orbital, another one
for the 2s orbital and three more for the 2p shell. One should expect no
more than only qualitative results from minimal sets and nowadays they
are hardly used anymore.
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
In the next level of sophistication are the double-zeta basis sets. Here,
the set of functions is doubled, i.e. there are two functions for each orbital.
If only the valence orbitals are doubled, and each core atomic orbitals is
still described by a single function, the resulting basis set is called split-
valence basis set. Typical examples are the 3-21G or 6-31G Gaussian basis
sets. In most applications, such basis sets are augmented by polarization
functions, i.e. functions of higher angular momentum than those occupied in
the atom. Polarized double-zeta or split valence basis sets are the mainstay
of routine quantum chemical applications since usually they offer a balance
compromise between accuracy and efficiency. Finally, it is obvious how
these schemes can be extended by increasing the number of functions in
the various categories. This results in triple- or quadruple-zeta basis sets
which are augmented by several sets of polarization functions.
If the molecules or solids of interest contain elements heavier than, say
krypton, one usually employs effective core potentials, also called pseudopo-
tentials, to model the core electrons. For a detailed discussion of the theory
of pseudopotentials see Ref. [226].
Our discussion about density functional theory would not be complete with-
out answering, at least partially, the most obvious question at this stage,
namely: how much should one trust DFT? In other words, what is the
accuracy of DFT at present, i.e. with the existent approximations for the
exchange-correlation functional? A detailed answer to this question is out
of the scope of this book and we just pretend to give here a flavor about
DFT’s performance in the case of the systems of interest in molecular elec-
tronics.
Let us remind again that the standard DFT, as presented here, gives
only results for the ground state energy and density of a system and related
properties. In the context of chemistry (or molecular physics), this means in
practice that one can expect from DFT information about the structure of
molecules, vibrational frequencies, atomization energies, dipole moments,
reactions paths and other similar properties. In what follows, we shall illus-
trate DFT’s performance with a very brief discussion of its predictions for
some basic molecular properties. We follow here Ref. [273], where an excel-
lent discussion of “goodness” of DFT in the context of quantum chemistry
can be found.
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
Table 10.1 Calculated and experimental bond lengths for different bonding
situations [Å]. The LDA calculations were done with the 6-31G(d) basis set
and the GGA ones with the 6-311++G(d,p) basis set.
BLYP 0.9945 45 10
BP86 0.9914 41 6
B3LYP 0.9614 34 6
B3P86 0.9558 38 4
B3PW91 0.9573 34 4
In this final section we shall discuss how DFT is used in practice in the field
of molecular electronics. For this purpose, we shall first discuss how DFT
can be combined with the nonequilibrium Green’s function (NEGF) tech-
niques presented in previous chapters to describe the electronic transport
in atomic-scale junctions. Then, we shall end this section with some com-
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
fronted with the following two questions: (i) how to compute the charge
density? and (ii) how to make finite the dimension of the problem? Both
questions can be answered with the help of Green’s function methods as
follows. First, we divide the junction into three parts: the left (L) and right
(R) electrodes and a central part or “extended molecule” that contains the
narrowest part of the junction (the molecule in Fig. 10.1) and part of the
electrodes.8 Second, within the LCAO approach, the charge density is com-
puted in terms of the density matrix, see Eqs. (10.57) and (10.58), which
in turn can be computed in terms of Green’s functions in the following
way. Let us assume that the system is in equilibrium. The retarded and
advanced Green’s functions Gr,a µν referred to the local basis functions µ and
ν can be written via their spectral representation [see Eq. (5.14)] as follows
X cµi c∗iν
Gr,a
µν (E) = . (10.65)
i
E ± iη − Ei
Here, the c’s are the coefficients of the expansion of the system eigenfunc-
tions (or molecular orbitals) in terms of local orbitals and Ei are the cor-
responding eigenenergies. Notice that i cµi c∗iν is nothing but the element
P
Pµν of the density matrix, see Eq. (10.58).9 Therefore, the density matrix
in the central part of the junction can be calculated from the retarded or
advanced Green’s function matrix of this part of the system as10
1 ∞
Z
P=∓ dE Im {Gr,a (E)} f (E), (10.66)
π −∞
where f (E) is the Fermi function that ensures that only the occupied states
contribute to the electron density. Now, these Green’s functions can be
computed via their Dyson’s equation [see Eq. (8.26)]
−1
Gr,a = [(E ± iη)S − H − Σr,a r,a
L − ΣR ] , (10.67)
where S is the overlap matrix, H is the one-electron Kohn-Sham Hamilto-
nian of the central part and Σr,aL/R are the left and right self-energies (see
section 8.1). The calculation of these self-energies requires the computation
of the Hamiltonian and Green’s functions of the electrodes. This issue will
be discussed in detail below.
This discussion shows that DFT can be applied to describe nanoscale
junctions by using Eq. (10.66) for determining the density matrix, rather
8 The reason for dividing the system in this way will become clear below.
9 In Eq. (10.58), we assumed that the c’s were real, but in principle, these coefficients
can be complex numbers and then, i cµi c∗iν is the most general definition of Pµν .
P
10 In what follows, we shall not write explicitly the subindexes CC to refer to the central
Im(z)
C
R
Re(z)
Fig. 10.2 The integral of a retarded Green’s function Gr (z), considered as a function
of a complex variable z, is the same along the contour C and along the real axis R.
However, Gr (z) is much smoother away from the real axis and for this reason, it is
advantageous to integrate the Green’s functions in Eq. (10.66) along a contour like C.
The lower limit of this contour has to be below the lowest lying states of the system,
while the upper limit should be the chemical potential of the system.
than solving the Kohn-Sham equations. The evaluation of the density ma-
trix requires the calculation of the Green’s functions via Eq. (10.67) from
the knowledge of the Kohn-Sham Hamiltonian of the central part. Since this
Hamiltonian depends on the charge density (or density matrix), Eqs. (10.66)
and (10.67) are coupled and they have to be solved in a self-consistent man-
ner. Finally, when these equations are solved, one can compute the different
equilibrium properties of a junction such as charge density, total energy, lo-
cal density of states, etc.
A technical comment is pertinent at this point. Usually Green’s func-
tions vary rapidly as a function of energy, which complicates the integration
appearing in Eq. (10.66). One can get around this problem by making use
of the fact that the Green’s functions are analytical functions and they
can be extended into the complex plane. This means in practice that the
integral in Eq. (10.66) can be done by integrating along a contour in the
complex plane, see Fig. 10.2, where these functions are very smooth. Thus,
one needs a much smaller number of points to carry out the numerical
integration.
The previous discussion also suggests a straightforward way of general-
izing this approach to nonequilibrium situations. In this case, the density
matrix can be expressed in terms of the Keldysh-Green’s functions as
Z ∞
1
P= dE G+− (E), (10.68)
2πi −∞
where G+− can be computed in terms of the retarded and advanced func-
tions as [see Eq. (8.12)]
G+− (E) = 2iGr [ΓL fL + ΓR fR ] Ga . (10.69)
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
Here, the scattering rates ΓL/R are the imaginary part of the self-energies
ΣaL/R (see section 8.1) and fL/R are the Fermi functions of the left and
right electrodes that include the energy shift caused by the applied bias
voltage. In this latter equation, the retarded and advanced functions can
be calculated from a Dyson’s equation like Eq. (10.67) taking into account
the presence of the bias voltage. Again, Eqs. (10.69) and (10.69) are coupled
and they have to be solved self-consistently. Once this is done, the different
transport properties can be computed as described in Chapter 8. It is
worth mentioning that again the integration in Eq. (10.68) can be done
more efficiently in the complex plane, although this time the integration
close to the Fermi energy requires to modify the contour shown in Fig. 10.2
(see Ref. [287] for details).
The key step to make our generic problem finite was the division of
the system into three parts, see Fig. 10.1. In this division one assumes
that the electrodes are not perturbed by the central part and therefore,
their Hamiltonians and charge densities can be obtained from a separate
(bulk-like) calculation, which only needs to be done once. This assumption
is based on the idea that deep inside a solid the Kohn-Sham potential
approaches the bulk potential. This approximation is often referred to as
the screening approximation and it provides natural boundary conditions
for the potential of the open system. In any calculation, it should be checked
that the potential of the central part actually matches that of the bulk
calculation. Such a check defines in practice the size of the central part and
this size depends on the nature of the electrodes.
The practical implementations of the DFT-NEGF combination differ
mainly in the way in which the electrodes Green’s functions are determined
and how the potential and Hamiltonian of the central part are forced to
match the corresponding ones in the leads. Roughly speaking, one can
grouped all the existent approaches into the following two families:
1.- Methods based on quantum chemistry software. In order to take ad-
vantage of the powerful and well-tested existent quantum chemistry codes,
several groups have implemented the DFT-NEGF approach as follows. The
diagonalization in these codes of the Kohn-Sham Hamiltonian of the finite
central system is replaced by Eqs. (10.68) and (10.69), which are solved
self-consistently. The self-energies required to compute the retarded and
advanced Green’s functions appearing in Eq (10.69) are obtained from a
separate calculation, from which one extracts the bulk Hamiltonian as well
as the coupling matrix elements between the central part and the leads.
This separate calculation can be done at different levels of sophistication.
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
DFT DFT
ρ, P H ρ, P H
H
ρ, P H ρ, P ΣL , ΣR
Fig. 10.3 Schematic description of the self-consistent loop in DFT for the determina-
tion of the electronic structure of a finite system (left panel) and an infinite non-periodic
system (right panel) [293]. For an isolated system, the density matrix is constructed by
occupying the states of the Kohn-Sham Hamiltonian H with N electrons. For an infi-
nite system, the density matrix is computed from the nonequilibrium Green’s functions,
see Eq. (10.68), which requires the determination of the self-energies from a separate
calculation.
Thus for instance, some authors describe the leads in terms of simple pa-
rameterized tight-binding methods [288–292] and others extract the bulk
Hamiltonian from DFT calculations of finite clusters [293, 294, 247]. The
bulk parameters are then used to construct surface Green’s functions using
recursive methods like those described in Refs. [198, 295, 296]. Following
Damle et al. [293], we summarize in Fig. 10.3 this approach and emphasize
the main differences with the standard method used for finite equilibrium
systems.
In this approach it is implicitly assumed that the central Hamiltonian
is a functional of the charge density only in the central system. This is
indeed the case for the contributions coming from the kinetic energy and the
electron-nuclei interaction, see Eq. (10.56). It is also true for the exchange-
correlation potential, see Eq. (10.59), but it is not really the case for the
classical Coulomb contribution, see Eq. (10.60). In this latter term, there
are non-local contributions coming from the leads, which are not easy to
describe correctly in the approach discussed in the previous paragraph. The
lack of these non-local contributions causes sometimes severe problems in
the convergence procedure in this method. However, it seems that, as shown
in Ref. [247], if the central system is sufficiently large, those additional
contributions do not play a major role in the physical quantities of interest,
and the self-consistent loop is not crucial in equilibrium systems.
2.- Methods based on solid state software. In the implementations of
the DFT-NEGF method based on the computer codes specially designed
for the description of (infinite) solid states systems, the Hartree potential
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
Calculation of the
bulk Hamiltonians of
the electrodes
Fig. 10.4 Flowchart of the self-consistent loop for the solution of the nonequilibrium
transport problem based on the solution of the Poisson equation [286].
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
Table 10.3 List of implementations of the combination of DFT and Green’s func-
tion techniques for the description of equilibrium and nonequilibrium properties of
nanoscale junctions. We provide the name of the code, if any, some characteristics,
the reference equilibrium DFT code in which it is based on, and a reference where
details about it can be found. Methods 1 and 2 refer to the methods described in
the text, BC means boundary conditions and TB corresponds to tight-binding.
10.9 Exercises
PART 3
293
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
294
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
Chapter 11
295
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
of Ref. [15].
8
7
Conductance (2e /h) 6
Gold
2
5
4
3
2
1
0
0 50 100 150 200 250 300
Piezo-voltage (V)
Fig. 11.1 Three typical recordings of the conductance G measured in atomic-size con-
tacts for gold at helium temperatures, using the MCBJ technique. The electrodes are
pulled apart by increasing the piezo-voltage. The corresponding displacement is about
0.1 nm per 25 V. After each recording the electrodes are pushed firmly together, and
each trace has new structure. Reprinted with permission from Ref. [74].
As we shall see later in this chapter, the actual number channels that give
a significant contribution to the conductance depends on the geometry of
the narrowest part of the contacts and on the number of valence orbitals of
the atoms of the corresponding metal.
The first question that we want to address is: what is the conductance
of a metallic atomic contact? As we discussed in Chapter 2, a metallic
contact of atomic size can be fabricated with various techniques, but the
most widely used ones are the scanning tunneling microscope (STM) and
the mechanically controllable break junction (MCBJ). In Fig. 11.1 one can
see some typical examples of the conductance measured during breaking of
a gold contact at low temperatures, using a MCBJ device.4 Notice that the
conductance decreases by sudden jumps, separated by “plateaus”, which
have a negative slope, the higher conductance the steeper. Some of the
plateaus are remarkably close to multiples of the conductance quantum,
G0 ; in particular the last plateau before loosing contact is nearly flat and
very close to 1 G0 .5 This behavior resembles the conductance quantization
4 In these atomic contacts the current-voltage characteristics are typically linear at low
voltages (below, let us say, 100 mV) and for this reason we shall mainly talk about the
linear conductance as the central transport property.
5 As it will become clear later in this chapter, the last conductance plateaus most likely
correspond to contacts with one atom in cross section and, in particular, long plateaus,
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
3 4
conductance (2e /h)
(a) Al (b) Pb
2
3
2
2
1
1
0 0
0 2 4 6 8 10 12 14 16 0 2 4 6 8 10 12 14 16
displacement (Å) displacement (Å)
Fig. 11.2 Evolution of conductance vs tip-sample relative displacement for several rep-
resentative nanocontacts of Al and Pb in STM experiments at low temperatures (4.2 K
for Al and 1.5 K for Pb). The black and grey curves correspond to elongation (open-
ing of the contact) and contraction (closing of the contact), respectively. Adapted with
permission from [264]. Copyright 1998 by the American Physical Society.
that occurs in point contacts defined in 2D electron gases (2DEG), see sec-
tion 4.6.1 and in particular Fig. 4.11. Indeed, different authors interpreted
the step-like evolution of the conductance as an evidence of conductance
quantization in atomic contacts. However, closer inspection of Fig. 11.1
shows that many plateaus cannot be identified with integer multiples of
the quantum unit, and the structure of the steps is different for each new
recording. Also, the height of the steps is of the order of the quantum unit,
but they can vary by more than a factor of 2, where both smaller and larger
steps are found.
The conductance traces not only change from realization to realiza-
tion, but they are also clearly distinct for different metals. In Fig. 11.2
we show several examples of conductance curves for aluminum and lead
wires obtained in the last stages of the breaking of contacts formed with a
STM at low temperatures. In the case of aluminum, one finds that many
plateaus have an anomalous slope: the conductance increases when pulling
the contact, in contrast to the results for gold. For aluminum, the last
plateau before breaking is still close to the quantum conductance, but one
frequently observes the conductance diving below this value, and then re-
covering to nearly 1 G0 , before contact is lost. Lead, on the other hand,
has a last conductance value, which is clearly above 1 G0 and the slope is
positive, i.e. the conductance is reduced upon stretching.
It is worth mentioning that, as one can see in the examples of Fig. 11.2,
like the one in the left curve in Fig. 11.1, are a signature of the formation of a monoatomic
chain.
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
20
conductance (2e²/h) 15
10
5
(a)
0
2
(b)
0
∆F
force (nN )
-2
-4
-6
-8
-1 0
0 .0 0 .5 1 .0 1 .5
tip displacem ent (nm )
Fig. 11.3 Simultaneous measurement of force and conductance on atom scale point
contacts for Au. The sample is mounted on a cantilever beam and the force between tip
and sample is measured by the deflection of the beam using an AFM. The measurements
are done in air at room temperature. Reprinted with permission from [58]. Copyright
1996 by the American Physical Society.
the conductance traces recorded when opening the contacts differ from
those recorded during the closing of the contacts. The reason for this lies
in the different atomic arrangements which can be achieved when stretching
as opposed to the ones when pushing the electrodes together. Furthermore,
as we shall explain later, the shape of the conductance traces also depends
on the technique used for fabricating the nanowires.
The previous examples raise several basic questions. The main one is
related to the origin of the conductance steps. In the case of point contacts
defined in 2DEGs, these steps are due to continuous change in the number
of conduction channels as the width varies. In that case the abruptness of
the jumps depends in particular on the shape of the confinement potential.
However, in the case of atomic contacts the cross section cannot be changed
continuously. Early molecular dynamic simulations [303, 304, 262] already
suggested that these jumps could be due to sudden atomic rearrangements.
The idea goes as follows. Upon stretching of the contact, the stress accumu-
lates elastic energy in the atomic bonds over the length of a plateau. This
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
(a) (b)
Fig. 11.4 (a) Conductance histograms of gold at 4.2, 77 and 295 K using the notched-
wire MCBJ technique. The inset shows the first peaks on the expanded scale. (b)
Conductance histograms of gold built from 2000 traces recorded at 1.25 V bias and
12 K (gray). The low temperature histogram (4.2 K) from the left panel is shown for
comparison (black). Note that the vertical axis is in logarithmic scale. Reprinted with
permission from [315].
(a) (b)
For alkali metals (Na, K, etc.) one finds histograms at low temperatures
with peaks near 1, 3, 5 and 6 times G0 [306]. An example for sodium is
shown in the upper right panel of Fig. 11.6. The fact that peaks near 2
and 4 G0 are absent points at an interpretation in terms of a smooth, near-
perfect cylindrical symmetry of the sodium contacts. The alkali metals can
be described to a good approximation as free electron systems. Within this
framework, it can be shown that in smooth cylindrical contacts with con-
tinuously adjustable contact diameter [143, 317], the conductance increases
from zero to 1 G0 as soon as the diameter is large enough, so that the first
conductance mode is occupied. When increasing the diameter further, the
conductance increases by two units because the second and third modes are
degenerate. In a similar way, one can explain the absence of a peak at 4 G0
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
Ag Na
Al Pb
Nb Pt
0 1 2 3 4 5 6 7 0 1 2 3 4 5 6 7
2 2
Conductance (2e /h) Conductance (2e /h)
Fig. 11.6 Conductance histograms of several metals obtained using the MCBJ tech-
nique. All the histograms were recorded at 4.2 K, except for Nb, which was obtained at
16 K. The conductance was measured at 20 mV for Ag and Nb, 10 mV for Na and Al
and 100 mV for Pb and Pt. Adapted with permission from [315].
• With the exception of alkali metals, the highest peak is always lying at
the lowest conductance value.
• The position of this peak for all the elements falls in the range between
0.7 and 2.3 G0 . There is no structure related to metallic conductance
in the histograms below the position of the first peak.
• For free electron-like alkali metals the first peak is extremely sharp and
is located almost exactly at 1 G0 . This statement also extends to the
almost free electron-like noble metals.
• For divalent metals (zinc, magnesium) and trivalent ones (aluminum)
the first peak is rather sharp and located slightly below 1 G0 . Other
multivalent metals, and in particular transition metals, exhibit a broad
first peak located well above 1 G0 and in some case like niobium it lies
even above 2 G0 .
As the Landauer formula indicates [see Eqs. (11.1) and (11.2)], the con-
ductance measurements gives us only access to the total transmission
P
T = n Tn at the Fermi energy. Obviously, the experimental determina-
tion of the individual transmission coefficients, Tn , could provide a valuable
insight into the origin of the differences between atomic contacts of different
metals. From a mathematical point of view, it is clear that the extraction
of the set {Tn } requires the analysis of transport properties that depend on
the transmission coefficients on a non-linear manner. As we saw in section
4.7, the shot noise is an example of such a quantity. Indeed, the experimen-
tal study of shot noise has provided very important information about the
conduction channels of both atomic contacts and single-molecule junctions.
This is discussed in detail in Chapter 19.
In this section we shall focus our attention on the first method that was
used to extract the individual transmission coefficients of an atomic con-
tact and which continues to be the most precise one. This method was put
forward by Scheer et al. [77] and it is based on the analysis of the subgap
structure in superconducting contacts. Let us explain this idea in certain
detail. Many simple metals, like Al, Pb, Nb, etc., are superconducting be-
low a critical temperature of the order of a few K. In the superconducting
state these metals exhibit a gap in density of states, ∆, which is typically
between 0.1 and 1 meV. This gap strongly influences the transport prop-
erties of superconducting contacts (including atomic junctions) leading to
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
1e 2e 3e
(a) (b) (c)
e h e
e
h
2∆ e 2∆ 2∆ h
e e
1
0 1
0 1
0 1
0 11
00 1
0
1
0 1
0 1
0 1
0 11
00 1
0
e e e
> 2∆/1
eV − − 2∆/2
eV > − 2∆/3
eV >
Fig. 11.7 Schematic representation of the multiple Andreev reflection (MAR) that take
place in a contact between two superconductors with gap ∆. We have sketched the
density of states of both electrodes, which exhibits a singularity at the gap edges. In order
to simplify these graphical representations, we have not shifted the DOS of the leads with
bias voltage, but equivalently we have taken into account the fact the quasiparticles gain
an energy eV every time they cross the junction. (a) This panel describes the process in
which a single electron tunnels through the system overcoming the gap due to a voltage
eV ≥ 2∆. (b) Andreev reflection process in which an electron is reflected as a hole
transferring a Cooper pair to the other electrode. This process has a threshold voltage
equal to ∆/e and its probability is proportional to T 2 . (c) MAR of order 3 in which a
quasiparticle is reflected twice before it finds an available state in the right electrodes.
In this process three electron charges are transferred across the junction, the threshold
voltage is 2∆/3e and its probability is proportional to T 3 . Higher-order processes with
contributions proportional to T n can also occur when the bias voltage is larger than
2∆/ne (with n integer).
5 7
(a) 6 (b) T = 0.2
4 0.95 T = 0.4
1.0 5 T = 0.8
eI/GN∆
G/GN
3 4
0.8
2 3
0.6
0.4 2
1 0.2 1
0.01
0 0
0 0.5 1 1.5 2 2.5 3 0 0.5 1 1.5 2 2.5 3
eV/∆ eV/∆
Fig. 11.8 (a) Zero-temperature I-V characteristics of a single channel superconducting
quantum point contact for different values of the normal transmission coefficient (indi-
cated in the graph). Notice that the current has been normalized with the normal state
conductance GN = G0 T to see all the curves in the same scale. (b) The corresponding
differential conductance G = dI/dV for three different values of the transmission. As
a guide for the eyes, the vertical dotted lines indicate the position eV = 2∆/n with
n = 1, . . . , 6. From Ref. [326].
and its onset causes a step in the current at V = ∆/e. The height of
the current step is smaller than the step at 2∆/e by a factor T , since the
probability for two particles to tunnel is T 2 . Depending on the junction
transparency, similar processes of order n involving the transfer of n parti-
cles can occur. These processes give rise to current onsets at eV = 2∆/n
with a step height proportional to T n . An example for n = 3 is illustrated in
Fig. 11.7(c). These processes are referred to as multiple Andreev reflections
(MARs) [323].10 The microscopic theory of MARs for a single-channel point
contact was developed in the late 1980’s and in the 1990’s by several groups
independently [324–328]. In Fig. 11.8 we show the zero-temperature I-V
curves and the corresponding differential conductance for different values of
the normal transmission coefficient. Notice the appearance of a pronounced
structure in the I-Vs close to voltages V = 2∆/ne (with n integer) as a re-
sult of the onset of the different MAR processes. This structure, which is
known as subharmonic gap structure, is more clearly seen in the differential
conductance a series of maxima, see Fig. 11.8(b).
The subharmonic gap structure had been measured in the context of
atomic contacts by several authors [69, 329, 330], but Scheer et al. [77]
10 These multiple processes were first described by Schrieffer and Wilkins [322] in the
limit of low transparent junctions. These authors coined the name multiple particle
tunneling (MPT) for these tunnel events. It is now understood that the concepts of
MAR and MPT are indeed equivalent.
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
Fig. 11.9 Current–voltage characteristics for four atom-sized contacts of aluminum us-
ing a lithographically fabricated MCBJs at 30 mK (symbols). The right inset shows
the typical variation of the conductance, or total transmission T = G/G0 , as a func-
tion of the displacement of the electrodes, while pulling. The data in the main panel
have been recorded by stopping the elongation at the last stages of the contact (a–c) or
just after the jump to the tunneling regime (d) and then measuring the current while
slowly sweeping the bias voltage. The current and voltage are plotted in reduced units,
eI/G∆ and eV /∆, where G is the normal state conductance for each contact and ∆ is
the measured superconducting gap, ∆/e = (182.5 ± 2.0)µV. The solid lines have been
obtained by adding several theoretical curves for a single channel contact and optimiz-
ing the set of transmission values. The curves are obtained with: (a) three channels,
P
T1 =0.997, T2 =0.46, T3 =0.29 with a total transmission Tn =1.747, (b) two channels,
P
T1 =0.74, T2 =0.11, with a total transmission Tn =0.85, (c) three channels, T1 =0.46,
P
T2 =0.35, T3 =0.07 with a total transmission Tn =0.88. (d) In the tunneling range a
P
single channel is sufficient, here Tn = T1 =0.025. Reprinted with permission from [77].
Copyright 1997 by the American Physical Society.
were the first to realize that the highly non-linear dependence of the su-
perconducting I-Vs on the transmission coefficient offers the possibility to
extract the transmission coefficients of a few-atom thick contacts. The
principle is illustrated in Fig. 11.9. Using lithographic MCBJs, Al atomic
contacts were formed at very low temperatures (30 mK). During the break-
ing of the Al wires, I-V at low bias (. 1 mV) were recorded along the
conductance plateaus. Examples of these I-Vs can be seen in the main
panel of Fig. 11.9 for different realizations of the contacts. Notice in par-
ticular that curves (b) and (c) correspond to similar values of the normal
state conductance (i.e. for voltages much larger than the Al gap). This
indicates that while these two junctions are almost indistinguishable in
the normal state, they exhibit clearly distinct superconducting I-Vs, which
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
means that their set of transmission coefficients are very different. The
I-V curves were fitted very accurately with the single-channel I-V curves of
Fig. 11.8(a) using as adjustable parameters both the number of conduction
channels and the transmission coefficients. The authors of Ref. [77] showed
that for the smallest contacts, the set of transmission probabilities can be
unambiguously determined.
The most important finding in these experiments was that in the last
“plateau” in the conductance, just before the breaking of the contact, typ-
ically three channels with different T ’s are required for a good description.
This is surprising since the conductance for such contacts is typically below
1 G0 (see Fig. 11.2(a) and the Al histogram in Fig. 11.6), and it would in
principle require only a single conductance channel. Contacts at the verge
of breaking are expected to consist of a single atom, and this atom would
then admit three conductance channels, but each of the three would only be
partially open, adding up to a conductance close to 1 G0 . This very much
contradicts a simple picture of quantized conductance in atomic-size con-
tacts, and poses the question as to what determines the number of channels
through a single atom.
In order to answer the question posed at the end of the previous section,
Cuevas, Levy Yeyati and Martin-Rodero [263] put forward a minimal model
to compute the conductance of atomic contacts within the framework of
Landauer approach. This model is based on a combination of a simple tight-
binding (TB) model and nonequilibrium Green’s functions techniques, in
the spirit of what we have discussed in Chapters 7-9, and it contains the fol-
lowing three basic ingredients. First, a proper description of the electronic
structure of atomic contacts, and in turn of their transport properties, re-
quires the inclusion in the TB model of at least the atomic orbitals that
give the major contribution to the bulk density of states at the Fermi en-
ergy. As one can see in Fig. 9.7, this means in practice to include the s
orbitals for alkali and noble metals, the s and p orbitals for metals like Al
and Pb and the s and d orbitals in the case of transition metals. Second,
since often we do not have direct information about the geometry of the
atomic contacts, it is important to study the influence of the precise atomic
arrangements. Finally, metals often exhibit local charge neutrality due to
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
11 Strictly speaking, this is only true if the hopping elements in the TB Hamiltonian are
approach of Ref. [263] as a linear combination of the atomic orbitals of the central atom.
13 In simple terms, this antisymmetric combination in the central atom is almost orthog-
onal to the incoming states from the leads which results in a very weak effective coupling
and the corresponding negligible contribution to the total conductance.
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
Transmission
2.5 T 2 = T3
2 T 4 = T5
T6
1.5
1
0.5
0
-5 -4 -3 -2 -1 0 1 2 3 4 5
(b) (d) E-EF (eV)
3 1 s
Bulk DOS (1/eV)
px = py
LDOS (1/eV)
pz
0.8 d3z2-r2
2 dxy = dx2-y2
4d 0.6
5s dyz = dzx
5p 0.4
1
0.2
0 0
-5 0 5 10 15 -5 -4 -3 -2 -1 0 1 2 3 4 5
E-EF (eV) E-EF (eV)
Fig. 11.10 (a) Ideal geometry of a single-atom Ag contact grown along the [111] di-
rection (taken as the z-axis). The distances are set to bulk distances and the last two
layers on both sides correspond to those atoms in the infinite surfaces used to model
the leads that are coupled to the atoms in the constriction. (b) Bulk density of states
(DOS) projected onto the s, p and d orbitals as a function of energy (measured with
respect to the Fermi energy EF ). (c) Total transmission and transmission coefficients
of the contact of panel (a) as a function of energy. (d) Local density of states (LDOS)
at the central atom projected onto the different atomic orbitals as a function of energy.
Courtesy of Michael Häfner [331].
aluminum atom. Moreover, the results for the number of channels were
shown to be robust against changes in the atomic configuration, whereas
the total conductance was found to vary depending on the exact atomic
geometry. Finally, this analysis was extended to the case of transition met-
als (in particular Nb) showing that for these metals up to 5 channels can
be expected for a single-atom contact. Again, the sixth channel that could
potentially contribute in a transition metal is actually closed for symmetry
reasons.
Before turning to the analysis of the experiments that confirmed these
ideas, we now want to illustrate them in more detail. In what follows, we
shall make use of the NRL tight-binding method of section 9.6 and the
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
formulas derived in section 8.1. The NRL method provides a very accurate
TB parameterization of the bulk properties of elementary solids that is also
well suited for low-dimensional structures (see discussion below). More-
over, this parameterization takes into account long range hopping matrix
elements and it includes up 9 orbitals in the basis set (the s, p and d closest
to the Fermi energy). Thus, this parameterization is more accurate than
that used in Ref. [263] and it serves us to test the conclusions drawn above.
Let us start by analyzing the conductance of an ideal single-atom contact
of Ag. The geometry of this ideal contact is shown in Fig. 11.10(a). It
is constructed by starting from a central atom and including the nearest
neighbors in the successive layers of the fcc lattice along the [111] direction.
The leads are modeled as two infinite surfaces grown along the same direc-
tion. The bulk density of states (DOS) of this metal computed from this
TB parameterization is shown in Fig. 11.10(b). Notice that the d bands
are filled, while the p bands have little weight at the Fermi energy. There-
fore, one expects the s band to dominate the transport properties of this
monovalent metal. Moreover, from the arguments above, one also expects
to have a single conduction channel in the case of one-atom contacts. This
is indeed confirmed by the calculations, as one can see in Fig. 11.10(c).
This figure shows both the total transmission and individual transmission
coefficients as a function of energy for this geometry. Notice in particular
that the transmission at the Fermi energy, which determines the conduc-
tance, is largely dominated by a single channel. One can get insight into
the nature of the conductance channels of single-atom contacts by analyz-
ing the corresponding local density of states (LDOS) at the central atom.
This LDOS projected onto the different atomic orbitals for the geometry
of panel (a) can be seen in Fig. 11.10(d). The first thing to notice is the
presence of true energy bands that, although are narrower than those of
the bulk solid, have widths of several electronvolts. This illustrates the
fact that the central atom is strongly coupled to the electrodes and there
is huge hybridization between its orbitals and those of the leads. Notice
also that there is a clear correlation between the energy dependence of the
transmission and that of the LDOS. In particular, one can see that the
transmission at the Fermi energy arises from a resonance of the s band, as
expected from the arguments above.
On the other hand, we can use this example to anticipate the results for
single-atom contacts of other metals in the periodic table. The idea goes
as follows. Most metals have similar energy bands and the main difference
is the position of the Fermi level, which is determined by the number of
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
bcc [001]
fcc [111] T1
T2 = T3
T4
T5 = T6
T7
1 1
Transmission
0.8 Ag 0.8 Au
0.6 0.6
0.4 0.4
0.2 0.2
0 0
-5 -4 -3 -2 -1 0 1 2 3 4 5 -5 -4 -3 -2 -1 0 1 2 3 4 5
1 1
Transmission
0.8 0.8 Pb
0.6 Al 0.6
0.4 0.4
0.2 0.2
0 0
-5 -4 -3 -2 -1 0 1 2 3 4 5 -5 -4 -3 -2 -1 0 1 2 3 4 5
1 1
Transmission
0.8 Nb 0.8 Pt
0.6 0.6
0.4 0.4
0.2 0.2
0 0
-5 -4 -3 -2 -1 0 1 2 3 4 5 -5 -4 -3 -2 -1 0 1 2 3 4 5
E-EF (eV) E-EF (eV)
Fig. 11.11 Individual transmission coefficients as a function of energy for dimer contacts
of Ag, Au, Al, Pb, Nb, and Pt computed with the NRL TB parameterization. The
structures of the contacts are shown in the upper part of the graph. For all the cases the
geometries are grown along the [111] direction of a fcc lattice (see left structure), except
for Nb, which is grown along the [001] direction of a bcc lattice. Courtesy of Michael
Häfner [331].
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
4
Pb
3
6 3 3 1
2
1
0
-1.0 -0.8 -0.6 -0.4 -0.2 0.0
6
Al
conductance (2e /h)
4
5
2
2 ≥8 6 3 21
0
-1.2 -1.0 -0.8 -0.6 -0.4 -0.2 0.0
6
Nb
4
≥7 5
2 3 1
0
-0.2 -0.1 0.0 0.1 0.2
4
5 4 Au
3
8 7 6 3 2 1 1
2
1
0
-2.0 -1.5 -1.0 -0.5 0.0
∆x (nm)
Fig. 11.12 Conductance curves measured as a function of contact elongation for Al, Pb,
Nb and Au. The number of channels contributing to the conductance was determined at
each point in the curves by recording the current-voltage relation and fitting the curves
with the theory for superconducting subgap structure. The numbers along the curves in
the figure indicate the number of channels obtained in this way. The number is constant
over a plateau, and usually jumps to a smaller value at the steps in the conductance.
Reprinted by permission from Macmillan Publishers Ltd: Nature [335], copyright 1998.
smaller than usually found, see Figs. 11.1 and 11.6. This was tentatively attributed to
the strong scattering in the nanofabricated device.
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
there were no means to obtain direct information about the contact ge-
ometries. Later, it became possible to directly image atomic-size contacts
by means of high resolution transmission electron microscopy (HR-TEM),
see e.g. Refs. [63, 336, 62, 65]. This technique has allowed to confirm the
existence of single-atom contacts and the formation of monoatomic chains
(see below).
Ttotal Ttotal
(a) TB calculation (b) DFT calculation
T1 T1
3.5
T 2 = T3 T 2 = T3
3
Transmission
T4 T4
2.5 T5 T 5 = T6
2 T6 T7
1.5 T 8 = T9
1
0.5
0
-8 -6 -4 -2 0 2 4 -8 -6 -4 -2 0 2 4
E-EF (eV) E-EF (eV)
A B C
5 5 4
5 4 4
1 1 2
1 3 2 3
2
3
3.5
Total conductance
A st
1 channel
3.0 nd
2 channel
rd
B 3 channel
th
2.5 4 channel
D E F th
5 channel
2.0 D th
G/G0
6 channel
5 5 1.5 C
5 4 4 4
1 1
1 2 2 2 1.0 F
3 E
3 3 0.5
0
0.4 0.8 1.2 1.6 2.0 2.4 2.8 3.2
stretching displacement (Å)
Fig. 11.14 Left panel: Snapshots of the structure of an Al nanowire during a stretching
process. The atoms involved in the important bonding rearrangements related to dis-
continuous changes in total energy, force, and conductance are labeled 1-5. Right panel:
Total conductance in units of G0 and the channel contribution along the stretching path.
Adapted with permission from [332]. Copyright 2003 by the American Physical Society.
tions due to both electronic and atomic shell effects [72, 345, 346].15 In the
case of gold, it has become clear that the pronounced peak close to 1 G0 is
related to the formation of monoatomic chains which sustain a single almost
fully open channel (see section 11.8). Several suggestions for the origin of
the peaks in the low-temperature histograms of some multivalent metals
have been made [347].16 An interesting idea was put forward by Hasmy et
al. [351] who performed molecular dynamics simulations to study the his-
tograms of the minimum cross section for Al contacts. At low temperatures
they obtained peaks at multiples of the cross section of a single atom, which
led them to an interpretation of the conductance histogram peaks based on
preferential geometrical arrangements of nanocontact necks. Dreher et al.
[342, 343] have corroborated the existence of well-defined peaks in the min-
imum cross section histograms. However, they have shown that those peaks
15 This is a very interesting topic that will not be further discussed here because our
interest is focused on the smallest contacts. For a detailed discussion of the shell effects
we recommend Refs. [15, 315].
16 Let us mention that more recently room-temperature conductance histograms of Al
and noble metals have been interpreted as an evidence of electronic and atomic shell
effect in these metals [348–350].
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
2.0
a
1.5
σGV (Go/V)
1.0
0.5
0.0
b
# points (x 10 )
20 53000
3
10
0
0 1 2 3 4
2
G (2e /h)
Fig. 11.15 (a) Standard deviation of the voltage dependence of the conductance versus
conductance for 3500 curves for gold measured with the notched-wire MCBJ technique
at 4.2 K. All data points in the set were sorted as a function of the conductance after
which the rms value of dG/dV was calculated from a fixed number of successive points.
The circles are the averages for 300 points, and the squares for 2500 points. The solid
and dashed curves depict the calculated behavior for a single partially-open channel and
a random distribution over two channels respectively. The vertical grey lines are the
corrected integer conductance values (see text). (b) Conductance histogram obtained
from the same data set. The peak in the conductance histogram at G0 extends to
53000 on the vertical scale. The insets shows a schematic diagram of the configuration
used in the analysis. The dark lines with arrows show the paths, which contribute to
the conductance fluctuations in lowest order. Reprinted with permission from [312].
Copyright 1999 by the American Physical Society.
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
depends on the phase difference between the two waves, and this phase dif-
ference depends on the phase accumulated by the wave during the passage
through the diffusive medium. The probability amplitude an is a sum over
all possible trajectories, and the phase for such a trajectory of total length
L is simply kL, k being the wave vector of the electron. The wave vector
can be influenced by increasing the voltage over the contact, thus launching
the electrons into the other electrode with a higher speed. The interference
of the waves changes as we change the bias voltage, and therefore the total
transmission probability, or the conductance, changes as a function of V .
This describes the dominant contributions to the conductance fluctuations,
and from this description it is clear that the fluctuations are expected to
vanish either when tn = 0, or when rn = 0.
Based on this model, Ludoph et al. [312] obtained the following analyt-
ical expression for σGV ,
³ ~/τ ´3/4 s
2.71 e G0 e
X
σGV = √ Tn2 (1 − Tn ) , (11.3)
~kF vF 1 − cos γ eVm n
where kF and vF are the Fermi wave vector and Fermi velocity, respectively,
τe = le /vF is the scattering time. The shape of the contact is taken into ac-
count in the form of the opening angle γ (see the inset in Fig. 11.15), and Vm
is the applied voltage modulation amplitude. The solid lines in Fig. 11.15(a)
are obtained from Eq. (11.3), assuming a single partially-open channel at
any point, i.e. assuming that channels open one-by-one as the conductance
increases. In agreement with the results discussed in previous sections, the
conductance for the smallest gold contacts is very well described by this
simple approximation. The amplitude of the curves is adjusted to fit the
data, from which a value for the mean free path is obtained, le = 5 ± 1 nm.
Similar experiments [312, 313] for copper and silver and for sodium also
show the quantum suppression of conductance fluctuations observed here
for gold. However, this suppression is not observed in the cases of alu-
minum or niobium [313], which clearly indicates that the transport in these
multivalent metals is governed by partially open channels. Thus, the con-
ductance fluctuation measurements confirm the overall picture described in
previous sections in which the transport in these nanowires is determined
by the valence orbitals of the corresponding material.
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
As we have seen in previous sections, all evidence shows that for an one-
atom thick contact of monovalent metals the current is carried by a single
mode, with a transmission probability close to one. Guided by this knowl-
edge in experiments on gold Yanson et al. [356] discovered that during the
contact breaking process the atoms in the contact form stable chains of
single atoms, up to 7 atoms long. Independently, Ohnishi et al. [63] discov-
ered the formation of chains of gold atoms at room temperature using an
instrument that combines a STM with a transmission electron microscope,
where an atomic strand could be directly seen in the images. Similar results
were also obtained in Refs. [336, 357].
Some understanding of the underlying mechanism can be obtained from
molecular dynamics simulations. Already before the experimental observa-
tions, several groups had observed the spontaneous formation of chains of
atoms in computer simulations of contact breaking [358, 359]. The au-
thors argue that the interatomic potentials used in the simulation may not
be reliable for this unusual configuration. However, the stability of these
atomic wires has now been confirmed by various more advanced calculations
[360–364].
Only three metals are known to form purely metallic atomic chains,
namely Au, Pt, and Ir [365]. They are neighbors in the sixth period of the
periodic table of the elements and they share another property: they make
similar reconstructions of the surface atoms on clean [100], [110], and [111]
surfaces. A common origin for these two properties has been suggested in
terms of a relativistic contribution to the linear bond strength [365].
There are many interesting aspects of the physics of metallic atomic
chains that could be discussed in detail such as the formation mechanism,
their stability or the fundamental limits for their length. However, we
shall focus our attention here on the analysis of their transport properties
and, in particular, of the so-called parity oscillation of the conductance
because it nicely illustrates how the electrical conduction takes place in
these remarkable 1D systems.
Let us start our discussion by briefly describing the original observa-
tions of the parity oscillations reported by Smit et al. [366]. These authors
investigated the changes of conductance in the process of pulling atomic
chains of Au, Pt and Ir using a STM and MCBJs. In Fig. 11.16 we show a
typical conductance trace obtained during the breaking of an Au contact.
As we have discussed in previous sections, the last conductance plateau
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
15
C o n d u c ta n c e [2 e /h ]
1 .1
2
1 .0
10
0 .9
0 .8
0. 0 0 .5 1 .0 1 .5
0
-1.5 -1.0 -0.5 0. 0 0 .5 1 .0 1 .5 2 .0
Fig. 11.16 Evolution of the conductance while pulling a contact between two gold elec-
trodes (measured with the notched-wire MCBJ technique at 4.2 K). In the inset, an
enlargement of the plateau of conductance at ∼ 1 G0 is shown. Variations to lower
conductance and back up by about 10–15% can be noticed when the atomic chain is
stretched. Reprinted with permission from [366]. Copyright 2003 by the American
Physical Society.
1.1
1.05 Au
1
0.95
conductance (2e /h)
2
2
Pt
1.5
2.2
Ir
2
1.8
1.6
0 0.2 0.4 0.6 0.8 1 1.2 1.4
length (nm)
Fig. 11.17 Averaged conductance traces for chains of atoms of Au, Pt, and Ir (measured
with the notched-wire MCBJ at 4.2 K). Each of the curves are made by the average of
individual traces of conductance while pulling atomic contacts or chains. Histograms of
the plateau lengths for the three metals obtained from the same set of data are shown by
the filled curves. Reprinted with permission from [366]. Copyright 2003 by the American
Physical Society.
to the peak spacing in the length histogram in Fig. 11.17. The latter is
obtained by taking as a starting point of the chain a conductance drop-
ping below 2.4 G0 . Ir shows a similar behavior although somewhat less
pronounced and it is more difficult to obtain good length histograms.
The simplicity of the atomic chains had stimulated numerical simula-
tions of their transport properties well before their experimental observa-
tion [369]. In particular, various groups [369–374] had found oscillations
in the conductance as a function of the number of atoms for calculations
of sodium atomic chains, where this metal was selected because it has the
simplest electronic structure. Sim et al. [370], using first-principles calcu-
lations and exploiting the Friedel sum rule, found that the conductance for
an odd number of atoms is equal to G0 , independent of the geometry of
the metallic banks, as long as they are symmetric for the left and right
connections. On the other hand, the conductance is generally smaller than
G0 and sensitive to the lead structure for an even number of atoms. The
odd-even behavior follows from a charge neutrality condition imposed for
monovalent-atom wires. These predictions agree nicely with the results
found for the Au chains.
As explained by the authors of Ref. [366], the odd-even behavior is
essentially an interference effect and it can be easily understood in the
frame of a simplified one-dimensional free-electron model, see Exercise 11.1
and Refs. [366, 375]. Instead, we shall provide here an argument from our
usual “atomistic” point of view. We shall analyze the parity effect in gold
chain with the help of the simple model described in Exercise 7.5, which
is represented schematically in Fig. 11.18. In this model we describe the
gold chain with a tight-binding Hamiltonian with a single orbital per atom
and with hopping elements, t, only between nearest neighbors. We assume
that the on-site energy, ǫ0 , is the same for all the atoms in the chain and
we set it to zero. We describe the leads by two identical semi-infinite linear
chains with, for simplicity, the same parameters as in the finite chain (bulk
hopping t and on-site energy ǫ0 ). Finally, the coupling between the chain
and the leads is described by the hoppings tL and tR that can be different
from the intra-chain hopping t.
The calculation of the transmission in this model, and therefore of the
zero-bias conductance, is a simple exercise that we proceed to sketch.18
From the general formula of Eq. 8.18, it is easy to show that the zero-bias
18 Itis not necessary to follow this calculation to understand the main conclusions that
will be drawn from this toy model.
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
1111
0000
0000
1111 1111
0000
0000
1111
0000
1111
tL t t t t tR1111
0000
0000
1111
0000
1111 0000
1111
1111
0000
L
0000
1111 R
0000
1111
0000
1111
0000 1
1111 ε0 0000
1111
0000
1111
0000
1111 N 1111
0000
0000
1111
0000
1111 0000
1111
0000
1111
Fig. 11.18 Schematic representation of the simple tight-binding model used to analyze
the parity effect in gold chains. In this model the chain has N atoms with a single
orbital per site and with an on-site energy ǫ0 = 0. There is only coupling between
nearest neighbors inside the chain, t. The coupling to the leads is given by the matrix
elements tL,R . The leads are modeled in practice by two identical semi-infinite chains
with the same parameters as the finite central chain.
1 1
Transmission
0.8 0.8
0.6 tL = tR = t
0.6
0.4 tL = tR = 0.9t 0.4
tL = tR = 0.8t
0.2 tL = tR = 0.7t 0.2
0 0
-2 -1 0 1 2 -2 -1 0 1 2
E/t E/t
Fig. 11.19 Transmission as a function of energy (normalized by the intra-chain hopping
t) for two chains with 4 (a) and 5 (b) atoms. The different curves correspond to different
values of the interface hopping tL,R , as indicated in the legend. The vertical dotted lines
indicate the position of the Fermi energy, which is zero in this case.
(and in any chain with an even number of atoms), those resonances produce
a minimum of the transmission at the Fermi energy, whereas they lead to a
maximum for the chains with an odd number of atoms. This result explains
qualitatively the parity effect discussed above for a monovalent metal.
The presence of transmission maxima at the Fermi energy for odd num-
ber of atoms in the chain and the minima for the chains with even number
of atoms can be understood as follows (see Exercise 11.1). The maxima of
the transmission appear at the position of the levels of the decoupled chain.
In the case of odd N there is always a level in the chain spectrum exactly
at the Fermi energy (E = 0) for symmetry reasons, which together with
the charge neutrality leads to a maximum of the linear conductance. On
the contrary, when N is even there is no chain level at the Fermi energy
and therefore these chains exhibit a lower conductance. To conclude this
discussion, we show in Fig. 11.20 the transmission at the Fermi energy as
a function of the number of chain atoms. As one can see, the amplitude of
the even-odd oscillations depends on the quality of the interfaces.19
The simple explanation presented above can account qualitatively for
the experimental behavior in the case of Au, characterized by a full 5d
band and a nearly half-filled 6s band. However, for the case of Pt and Ir, in
which the contribution of 5d orbitals to the conductance is important, there
19 The conductance does not decay with length in this case because the Fermi energy
lies inside the “band” formed by the states of the finite chain. For energies outside this
energy window, the conductance decays exponentially with length. This is what happens
in the case of molecular junctions (see discussion in section 13.4).
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
1
Transmission (EF)
0.8
0.6
tL = tR = t
0.4
tL = tR = 0.9t
tL = tR = 0.8t
0.2 tL = tR = 0.7t
0
2 3 4 5 6 7 8 9 10
Number of chain atoms
Fig. 11.20 Transmission at the Fermi energy as a function of the number of atoms in
a linear chain. Notice the even-odd effect. The different curves correspond to different
values of the interface hopping tL,R , as indicated in the legend.
is no reason why this simple picture should still hold. The experiments of
Smit et al. [366] triggered off new the theoretical analyses of the conduc-
tance of these monoatomic wires, see for instance Refs. [342, 343, 376–381].
In particular, de la Vega et al. [376] presented an appealing comparative
study that we now proceed to describe. These authors studied the con-
ductance of ideal chain geometries of Au, Pt and Ir, in which the atomic
chain is connected to bulk electrodes represented by two semi-infinite fcc
perfect crystals along the (111) direction. Using the Green’s function tech-
niques detailed in Chapters 7 and 8 and a parameterized self-consistent
tight-binding model, they obtained the evolution of the conductance with
the number of atoms in the chain depicted in Fig. 11.21(a). Notice that
this evolution is rather sensitive to the elongation, especially in the case
of Pt and Ir (for Au the conductance exhibits small amplitude even-odd
oscillations, which remain practically unaffected upon stretching).
The main features and the differences between Au, Pt and Ir are more
clearly understood by analyzing the local density of states and the energy
dependence of the transmission, shown in Fig. 11.21(b) for a N = 5 chain of
these metals at an intermediate elongation. The Au chains are characterized
by a single conduction channel around the Fermi energy with predominant
s character. The transmission of this channel lies close to one and exhibits
small oscillations as a function of energy resembling the behavior of the
single band TB model discussed above.
In the case of Pt the contribution from the almost filled 5d bands be-
comes important for the electronic properties at the Fermi energy. There
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
(a) a=2.70 Å
(b)
1.02 0.8 3
a=3.0 Å
Au
1.00 0.6 Au Au
2
0.98 0.4
1
0.96 0.2
0.94 0.0 0
Total transmission
2.5 0.8 3
Pt
2.0 0.6 Pt Pt
G(EF)/G0
DOS
2
1.5 0.4
a=2.70 Å
1
1.0 a=2.80 Å 0.2
a=2.90 Å
0.5 0.0 0
Ir a=2.50 Å 3
a=2.75 Å 0.6
4 Ir Ir
a=2.85 Å
0.4 2
3
0.2 1
2
0.0 0
1 2 3 4 5 6 7 8 0 5 10 0 5 10
N Energy (eV) Energy (eV)
Fig. 11.21 (a) Evolution of the conductance with N for different values of the inter-
atomic distance a. (b) Local density of states (LDOS) at the central atom and total
transmission for Au, Pt and Ir chains N = 5 at an intermediate elongation. The LDOS
is decomposed in s (full line), d (dotted line) and p (dashed line) orbitals with the same
normalization in the three cases. Reprinted with permission from [376]. Copyright 2004
by the American Physical Society.
(a) (b) 1
0.8
4 0.6
a = 2.33 Å m=0 m = 0 up
2 Pt 0.4
E-EF (eV)
1 m = -1,1
T(EF)
0
0.8
-2 m = -2,2
0.6
-4 0.4
m = -1,1
-6 m=0 0.6
-8 0.4 m = 0 down
0 0.2 0.4 0.6 0.8 1
0.2
ka/π 0
0 5 10 15 20 25 30
N
Fig. 11.22 (a) Band structure of the infinite Pt chain. The bands are classified by the
quantum number m corresponding to the projection of the angular momentum on the
chain axis. The arrows indicate the crossing of the Fermi level for the m = 0 and the
m = ±1 bands. (b) Channel decomposition for Pt chains as a function of N . The legends
indicate the symmetry of the corresponding bands in the infinite chain. Courtesy of A.
Levy Yeyati.
channels of the chains is realized when analyzing the evolution of the con-
ductance and its channel decomposition for even longer chains than in
Fig. 11.21 (N > 8). This is illustrated in Fig. 11.22(b). As it can be
observed, the decrease of the total conductance of Pt for N < 7 − 8 cor-
responds actually to a long period oscillation in the transmission of the
two nearly degenerate channels associated with the m = 1± bands. This
period can be related to the small Fermi wave vector of these almost filled
d bands, as indicated by the arrows in Fig. 11.22(a). In addition, the up-
per m = 0 band crossing the Fermi level is close to half-filling giving rise
to the even-odd oscillatory behavior observed in the transmission of the
channel with predominant s character. The lower m = 0 band tends to be
completely filled and the corresponding channel is nearly closed for short
chains. However, one can appreciate a very long period oscillation in its
transmission, rising up to ∼ 0.5 G0 , for N ∼ 13-14.
The general rule that emerges from the above analysis is that the
transmission corresponding to each conduction channel oscillates as ∼
cos2 (kF,i N a), where kF,i is the Fermi wave vector of the associated band in
the infinite chain. In the case of Pt the total conductance for short chains
(N < 7-8) exhibits an overall decrease with superimposed even-odd oscil-
lations in qualitative agreement with the experimental results. For even
longer chains (not yet attainable in experiments) these calculations pre-
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
11.10 Exercises
V2
k1 k2 < k1 k3 = k1
V1 V3
x=0 x=L x
Fig. 11.23 One-dimensional model for the potential landscape describing an atomic
chain of length L.
(i) Compute the transmission through this potential barrier for energies higher
than the step height as a function of k1 and k2 . Hint: the solution is given in
Eq. (4.16).
(ii) Show that the transmission exhibits oscillations as a function of the chain
length where the maxima are given by Tmax = 1 and minima by Tmin = 4γ 2 /(1 +
γ 2 )2 , where γ = k2 /k1 .
(iii) To determine the value of k2 relevant for the transport, one might be
tempted to fix it to kF of an infinite chain. Show that assuming that there is
an electron per atom this Fermi wave vector is given by kF = π/(2a), where
a = L/N is the interatomic distance, N the number of atoms in the chain and
L its length. Show also that with this choice for k2 this 1D model predicts that
the conductance maxima should appear for chains with even number of atoms,
contrary to the model explained in section 11.8.
(iv) The problem found in (iii) can be solved by computing k2 in the following
more appropriate manner. Since the chain is finite, there is a limited set of
possible values for k2 , namely k2i = (π/a)i/N with i = 1, ..., N (can you explain
why?). Then, imposing charge neutrality one obtains k2 = (π/2a)N/(N + 1) for
the Fermi wave vector. Use this value for k2 in the expression of the transmission
to show that now the model reproduces the correct phase of the conductance
oscillations.
11.2 Even-odd effect in gold atomic chains:
Let us consider the chain model discussed in section 11.8 to explain the even-
odd effect in the conductance of gold atomic wires.
(i) Reproduce the results of Figs. 11.19 and 11.20.
(ii) Diagonalize the Hamiltonian of the uncoupled finite chain for different
number atoms, N , to obtain its energy spectrum. Show that for N odd there is
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
Here, ΓL,R are the scattering rates at the Fermi energy given by ΓL,R = tL,R ,
where we have assumed that the semi-infinite chains describing the leads have
the same hopping as the finite central chain.
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
Chapter 12
Spin-dependent transport in
ferromagnetic atomic contacts
The use of the spin degree of freedom of the electron in conventional charge-
based electronic devices has lead to the discovery of many fundamental
effects and, in some cases, to new technological applications [387, 388].
The emblematic physical effects in this new field, already known as spin-
tronics, like the giant magnetoresistance (GMR), tunneling magnetoresis-
tance (TMR) or anisotropic magnetoresistance (AMR) stem from the spin-
sensitivity of the scattering mechanisms that dominate the transport prop-
erties in electronic devices made of magnetic materials.1 In recent years,
a great effort has been devoted to understand how these fundamental ef-
fects are modified when the dimensions of a magnetic device are reduced of
the way down to the atomic scale. Contrary to the case of non-magnetic
atomic contacts, the physics of their ferromagnetic counterparts is not so
well established and there are still basic open problems. The goal of this
chapter is to provide a brief introduction to the transport properties of
ferromagnetic atomic-size contacts and to draw the attention to problems
that could be soon analyzed in the context of molecular junctions.
There are many different topics in this field that one could address. In
order to illustrate the interesting physics of ferromagnetic atomic contacts,
we have chosen to discuss three issues that are attracting a lot of atten-
tion. The first one concerns the conductance of these atomic contacts in
the absence of domain or external magnetic fields and, in particular, the
possibility of observing conductance quantization. The second problem is
related to the magnetoresistance of these atomic-scale conductors, which
has been shown to be enormous in comparison with the one found in larger
devices made of the same materials. Finally, we shall address the issue of
1 Fora basic explanation of all these magnetoresistive effects, see Ref. [388] or chapter
15 in Ref. [389].
335
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
3000
Fe
2000
1000
0
0 1 2 3 4 5
3000
Co
2000
Counts
1000
0
0 1 2 3 4 5
4000
3000 Ni
2000
1000
0
0 1 2 3 4 5
2
Conductance (2e /h)
Fig. 12.1 Conductance histograms for Fe, Co and Ni atomic contacts obtained with the
notched-wire MCBJ technique without magnetic field (thin curve) and when a magnetic
field of 5 T parallel to the current direction was applied (thick curve). The conductance
was measured using a dc bias voltage of 20 mV and a temperature of 4.2 K. Reprinted
with permission from [403]. Copyright 2004 by the American Physical Society.
ferromagnetic contacts. In this work the calculations are based on the com-
bination of the NRL tight-binding method of section 9.6 (see also Ref. [428])
and nonequilibrium Green’s function techniques, which was already used
in the last chapter.3 Let us stress that in this discussion we shall neglect
both the spin-orbit interaction and we shall assume that there are no do-
main walls present in the contacts.4 With these assumptions, the transport
properties of ferromagnetic contacts can be described in terms of two inde-
pendent contributions coming from both spin bands. In particular, in the
framework of Landauer’s approach the linear conductance at low tempera-
ture can be expressed as follows
e2 X
G= Tσ (EF ), (12.1)
h σ
where Tσ (E) is the total transmission for spin σ =↑, ↓ at energy E and EF
is the Fermi energy. We also define the spin-resolved conductances Gσ =
(e2 /h)Tσ (EF ), such that G = G↑ + G↓ . The transmissions are obtained as
follows
X
Tσ (E) = Tr[t†σ (E)tσ (E)] = Tn,σ (E), (12.2)
n
where tσ (E) is the transmission matrix and Tn,σ (E) are the individual
transmission eigenvalues for each spin σ (see section 8.1.3 for more details
on the calculation of these transmission matrices).
An important quantity in our discussion will be the spin polarization P
of the current, which we define as
G↑ − G↓
P = × 100%. (12.3)
G↑ + G↓
Here, we shall assume that spin up denotes the majority spins, while spin
down corresponds to the minority ones.
In order to understand the results described below, it is instructive to
first discuss the bulk density of states (DOS). The spin- and orbital-resolved
bulk DOS of these materials around EF , as calculated from the NRL tight-
binding method, is shown in Fig. 12.2. The common feature for the three
ferromagnets is that the Fermi energy for the minority spins lies inside the
d bands. This fact immediately suggests that the d orbitals may play an
important role in the transport. For the majority spins the Fermi energy lies
3 The technical details of the calculation of the current in ferromagnetic contacts have
3d
1.5 1.5 1.5 4p
4s
1 1 1
0 0 0
1.5 1.5 1.5
1 1 1
0 0 0
-6 -4 -2 0 2 -6 -4 -2 0 2 -6 -4 -2 0 2
E (eV) E (eV) E (eV)
Fig. 12.2 Bulk density of states (DOS) of Fe, Co, and Ni, resolved with respect to the
individual contributions of 3d, 4s, and 4p orbitals, as indicated in the legend. The upper
panels show the DOS for the majority spins (spin up) and the lower ones the DOS for
minority spins (spin down). The vertical dotted lines indicate the Fermi energy, set to
zero. Reprinted with permission from [427]. Copyright 2008 by the American Physical
Society.
close to the edge of the d band. The main difference between the materials
is that for Fe there is still an important contribution of the d orbitals, while
for Ni the Fermi level is in a region where the s and p bands become more
important. The calculated values of the magnetic moment per atom (in
units of the Bohr magneton) of 2.15 for Fe, 1.3 for Co, and 0.45 for Ni are
reasonable agreement with the literature values [429].
Let us turn now to the analysis of the conductance of Fe, Co and Ni
contacts. We consider ideal one-atom thick contact geometries with a cen-
tral dimer as shown in the upper part of Fig. 12.3. In this figure we also
present the total transmission for majority spins and minority spins as a
function of energy as well as the individual transmission coefficients for
those geometries. As one can see in Fig. 12.3(a), for the case of Fe one
finds 3 channels for the majority spins, yielding G↑ = 1.24e2 /h, while for
the minority spins 3 channels contribute to G↓ = 0.70e2 /h. The total con-
ductance is 0.94 G0 and the polarization P = +28%. For the Co contact,
see Fig. 12.3(b), one finds G↑ = 0.90e2 /h and G↓ = 2.23e2 /h, summing up
to a total conductance of 1.6 G0 . The transmission is formed by 3 channels
for the majority spins (with one clearly dominant) and 6 channels for mi-
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
Fig. 12.3 Transmission for the three single-atom contacts of Fe, Co, and Ni containing a
dimer in the central part of the contact. The geometries are shown in the upper graphs.
The distances are set to bulk distances and the atoms of the last two layers correspond
to the atoms of the leads (semi-infinite surfaces) that are coupled to the central atoms in
the model. We present the total transmission (black solid line) for both majority spins
and minority spins as well as the transmission of individual conduction channels that give
the most important contribution at Fermi energy, which is indicated by a vertical dotted
line. The channels corresponding to τ1 , τ2 , and τ3 are two-fold degenerate. Reprinted
with permission from [427]. Copyright 2008 by the American Physical Society.
Fe to Ni, the Fermi energy lies more and more outside of the d band for
the majority spins, which implies that the number of channels is reduced
for this spin species. In particular, for Ni a single majority spin channel
dominates and in some sense, this material behaves as a monovalent and
a transition metal combined in parallel. Finally, the conductance values
for single-atom contacts lie typically above 1 G0 , in agreement with the
experimental results of Fig. 12.1.
The analysis of Häfner et al. [427] also put forward two additional im-
portant conclusions. First, as a consequence of the contribution of the d
bands, the value of the conductance and the current polarization are very
sensitive to the contact geometry and to disorder. Second, in the tunneling
regime one can have a much higher current polarization reaching in some
cases values close to 100%.
These ideas and conclusions can be further illustrated with the results
of Ref. [343] where conductance calculations were combined with classical
molecular dynamics simulations to determine the contact geometries. In
Fig. 12.4 we show the formation of a single-atom Ni contact containing a
dimer in its central part just before rupture. Moreover, this figure shows the
corresponding conductance and channel transmissions for both spin compo-
nents, the strain force necessary to break the contact, the spin polarization
of the current and the contact geometries. As one can see in this figure, in
the last stages of the stretching the conductance is dominated by a single
channel for the majority spins, while for the minority spin band there are
still up to 4 open channels. In particular, in the very final stages (regions
of 3 or 1 open channels for G↑ ) the spin-up conductance lies below 1.2e2 /h,
while for spin down it is close to 2e2 /h, adding up to a conductance of
around 1.2-1.6 G0 .
It is worth discussing the behavior of the spin polarization of the current,
P . Notice that at the beginning of this contact evolution it takes a value
around −40%, which is indeed close to the spin polarization of the bulk DOS
at the Fermi energy (−40.5% in these model calculations). However, as the
contact evolves, P fluctuates and even increases to positive values, which
cannot be simply explained in terms of the bulk DOS. Notice also that
P reaches the value of +80% in the tunneling regime, when the contact
is broken. Such a huge value in this regime is due to the fact that the
couplings between the d orbitals of the two Ni tips decrease much faster
with distance than the corresponding s orbitals. As a result there is a great
reduction of the spin-down conductance and in turn in a large positive value
of P .
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
8
force (nN)
6
4
2
0
0.2 0.4 0.6 0.8
2
G↑ (5) (4)
MCS
1
0
80
spin polarization (%)
9 60
rad. of the min. cross-section (Å)
40
8 20
0
7
conductance (e /h)
-20
6 -40
2
-60
5 (18) (17) -80
-100
0 0.2 0.4 0.6 0.8
4 elongation (nm) (7)
(6)
3 (5)
(4)
2 G↓
MCS (9) (8)
1
0
0 0.2 0.4 0.6 0.8
elongation (nm)
As a final comment on these results, let us to point out that the contri-
bution of the minority spin component to the conductance is more sensitive
to changes in the contact geometry, as one can see in Fig. 12.4. Again, this
is a consequence of the fact that the minority spin contribution is domi-
nated by the bands arising from the d orbitals, which are anisotropic and
therefore more sensitive to disorder than the s states responsible for the
conductance of the majority spins.
R(P )]/R(AP ) × 100%, which in the usual situations has an upper bound of 100%.
7 The anisotropic MR will be the subject of the next section.
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
Fig. 12.5 (A) Scanning electron micrograph of a device where gold electrodes are used
to contact two permalloy thin-film magnets (inset) on top of an oxidized aluminum
gate. (B) Micromagnetic modeling showing antiparallel magnetic alignment across the
tunneling gap in an applied magnetic field of H= 66 mT. Reprinted with permission
from [438]. Copyright 2006 American Chemical Society.
2
40
(a) (b)
∆R/RP (%)
1.5 30
20
1
10
0.5 0
-10
100 200 300 10
3
10
4
Resistance (Ω) Resistance (Ω)
Fig. 12.6 (a) Magnetoresistance as a function of resistance in the range less than 400
Ω (device I). (b) Magnetoresistance as a function of resistance in the range 60 Ω - 15 kΩ
(device II). Adapted with permission from [438]. Copyright 2006 American Chemical
Society.
advantage of this device geometry is that the magnets are attached rigidly
to a non-magnetic substrate with no suspended parts, so that the influence
of magnetostriction and magnetostatic forces on the contact are expected to
be negligible. Moreover, these experiments were conducted at low temper-
atures (4.2 K) to have the required thermal stability. Although it cannot
be taken for granted that electromigration of alloys would maintain the
stoichiometry down to the atomic scale, the magnetic properties seem not
to have changed during the final phase of the electromigration process.
Let us now summarize the main findings of this work. When the resis-
tance of a device is low (< 400 Ω), it increases smoothly as electromigration
proceeds. The cross-section of the constriction varies from 100 × 30 nm2
(60 Ω) to approximately 1 nm2 (400 Ω). In this regime small (< 3%) pos-
itive MR was found which increases as the constriction is narrowed, see
Fig. 12.6(a), as expected from the semiclassical theory of Levy and Zhang
[432]. In this theory, the resistance of the domain wall scales inversely with
its width and the MR ranges typically from 0.7% to 3% for bulk ferromag-
nets.
In the resistance range from 400 Ω to 25 kΩ, corresponding to a crossover
between transport through just a few atoms and tunneling, the value of
MR exhibits pronounced dependence on the resistance of the device, see
Fig. 12.6(b). The MR has a minimum for resistances above 1 kΩ, and
typically changes sign here to give negative values. As the resistance is
increased further into the kΩ range, the MR increases gradually to positive
values of 10-20%. The observed MR values in the point contact regime are
smaller than expected from scaling results of the semiclassical theory [432],
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
which is not surprising since the current is transmitted through just a few
channels.
Finally, in the tunnel regime, when the resistance of a device becomes
greater than tens of kΩ, MR values in the range from -10% to a maximum
of 85% where observed. These large fluctuations clearly indicate that the
MR is sensitive to the details of the atomic structure near the tunnel gap.
The tunneling current is flowing through just a few atoms on each of the
electrodes, and the electronic structure at these atoms does not necessarily
reflect the same degree of spin polarization as in the bulk of the ferromagnet.
Experiments like the ones just described raise several basic questions,
most of them related to the role of a domain wall scattering in the mag-
netoresistance of atomic-size contacts and whether or not it can be re-
sponsible for the huge MR values reported in some experiments. These
questions have been addressed theoretically by numerous authors, see e.g.
Refs. [439–441, 409, 410, 413, 412, 414, 415]. In order to elucidate these
issues a theory should incorporate three basic ingredients: (i) a proper de-
scription of the electronic structure of ferromagnetic atomic contacts, (ii)
an adequate description of the domain wall or magnetization profiles that
can appear in atomic-scale junctions and (iii) an analysis of realistic atomic
geometries. One of the few works that meets these requirements is that of
Jacob et al. [415] in which the authors studied the magnetoresistance of Ni
atomic contacts using ab initio transport calculations. In Fig. 12.7 we repro-
duce results from this work for the transmission as a function of energy for a
single-atom Ni contact for both P and AP orientations. These results were
obtained using local spin density approximation (LSDA). In the AP case
the self-consistent magnetization reverses abruptly between the tips atoms,
i.e. this calculation confirms the possibility of having atomically-abrupt do-
main walls. However, despite this fact, the MR acquires a moderate value
of 23%, which suggests that the domain wall scattering does not account
for the large MR in Ni single-atom contacts.
The quantitative result above was found to be very sensitive to the func-
tional used in the DFT calculations, but in no case very large MR values
were found. According to the authors, the reason for the moderate MR val-
ues is two-fold. First, in the AP configuration the resistance is never too low
because of the robust contribution of the s orbitals, which is of the order of
G0 for a single-atom contact. Second, in the P configuration the resistance
never reaches the minimum value of the ballistic case because, as we saw in
the previous section, the transport in these ferromagnetic contacts is not
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
Fig. 12.7 Conductance per spin channel in the P configuration for the model nanocon-
tact shown in the inset calculated with the local spin density approximation. (b) Same
as in (a), but for the AP configuration. Reprinted with permission from [415]. Copyright
2005 by the American Physical Society.
really ballistic8 and the d bands contribute with partially open channels.
Another interesting finding of this work is the fact that the MR in atomic
contacts can become negative, as in the experiments described above. This
shows once more that the usual classical or semiclassical arguments do not
apply to the transport in ferromagnetic atomic-size contacts.
(a)
(b) (c)
Fig. 12.8 (a) Zero-bias differential resistance vs angle of applied magnetic field at dif-
ferent field magnitudes at 4.2 K, illustrating bulk AMR for a permalloy constriction size
of 30×100 nm2 and resistance R0 = 70 Ω. The inset shows a scanning electron micro-
graph of a typical device. (b) Evolution of AMR as the device resistance R0 is increased
from 56 to 1129 Ω. (c) AMR for a device with R0 = 6 kΩ exhibiting 15% AMR, and a
R0 = 4 MΩ tunneling device, exhibiting 25% AMR. All measurements were made at a
field magnitude of 800 mT at 4.2 K. Inset in panel (b): AMR magnitude as a function
of R0 for 12 devices studied into the tunneling regime. Adapted with permission from
[446]. Copyright 2006 by the American Physical Society.
infinite chain in the absence of spin-orbit interaction [panel (a)] and in the
presence of spin-orbit interaction for magnetizations both parallel to the
chain axis [panel (b)] and perpendicular to it [panel (c)]. The key idea is
that by rotating the magnetization one can change the number of bands
crossing the Fermi energy, EF . Since in a ballistic conductor the number of
bands at EF is equal to the number of conduction channels (all of them with
perfect transparency), this change is reflected in an abrupt change of the
corresponding linear conductance. In the particular example of Fig. 12.10,
the conductance would change from 6 e2 /h to 7 e2 /h and back upon rotation
of the magnetization.
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
The results of Fig. 12.9 definitively resemble the expected BAMR be-
havior described above, although in general the conductance jumps do not
occur between quantized values, as can easily be seen in the representative
traces However, as we have seen in previous sections, realistic ferromagnetic
contacts made of transition metals are not ballistic and thus, the interpre-
tation of the conductance jumps in terms of BAMR is at least question-
able. Indeed, Shi and Ralph [451] have suggested that these jumps might
originate from two-level fluctuations due to changes in atomic configura-
tions [452].
From the above discussion, one can see that at present there is still a
controversy about AMR in atomic contacts, concerning the origin of the
enhanced amplitude, the anomalous angular dependence, the occurrence
of conductance jumps and the voltage dependence. Different theoretical
groups have tried recently to shed new light on this problem. Thus for
instance, it has been proposed that the presence of resonant states local-
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
ized in the electrodes near the junction break could give rise to a strong
dependence of the conductance on the magnetization direction [453, 454].
This is an appealing explanation, but as we have seen in section 12.1, the
transmission of ferromagnetic contacts is usually very smooth around the
Fermi energy on the scale of a few meV. On the other hand, Autes et al.
[455] have proposed an alternative explanation of the conductance jumps
in terms of the existence of giant orbital moments in the contacts.
More recently, Häfner et al. [456] have put forward a simple explanation
for the anomalous AMR in terms of the reduced symmetry of the atomic
contacts as compared with bulk samples. The idea goes as follows. The
AMR stems from the scattering between the s and d energy bands induced
by the spin-orbit interaction and therefore, the AMR signal may reflect
the symmetry of the lattice. In bulk samples the final signal is a result of
the average over many impurities, but in the extreme case of an atomic-
scale contact, such signal strongly depends on the local geometry. This is
particularly clear in the case of a single-atom contact where all the current
must flow through a single bond. Thus, it is not so strange to observe,
depending on the contact realization, a large amplitude or an anomalous
angular dependence as compared with bulk samples. This idea is illustrated
in Fig. 12.11, where we reproduce the results of Ref. [456]. Here, one can
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
Fig. 12.11 Contact evolution of a Ni junction grown in the fcc [001] direction as ob-
tained from classical molecular dynamics simulations. (a) Spin-projected, G↑,↓ , and
total conductance in the absence of spin-orbit interaction and total conductance aver-
aged over θ, φ in the presence of spin-orbit interaction. Vertical lines correspond to the
contact geometries in (b). Inset: relative AMR amplitude ∆G/hGiθ = (Gmax,θ (φ) −
Gmin,θ (φ))/hG(θ, φ)iθ vs. inverse averaged conductance. (c) Conductance vs. θ for the
geometries in (b) and with φ in steps of π/6. (d) Same as (c) for the thick contact with
324 atoms shown in this panel. Reprinted with permission from [456]. Copyright 2009
by the American Physical Society.
see the evolution of the conductance and AMR signal10 as a function of the
polar angle θ and azimuthal angle φ during the formation of a Ni atomic
contact. This formation was simulated by means of molecular dynamics
(see Ref. [456] for technical details).
In Fig. 12.11(d) one can see that in the limit of thick contacts, these
model calculations recover the bulk behavior with an AMR amplitude of
0.45%. However, in the case of small contacts, one can observe clear devi-
ations from the cos2 θ-behavior and an enhancement of the amplitude [see
Fig. 12.11(a-c)]. Notice in particular that the signal in this case also de-
pends strongly on the azimuthal angle φ, contrary to the bulk case. Finally,
the statistical analysis of the data of these simulations reveals strong fluc-
tuations in the AMR signal and an increase to 2% on average in the last
steps before breaking, see inset of Fig 12.11(a).
On the other hand, in the analysis of these realistic geometries, Häfner
et al. [456] did not find signs of BAMR or the presence of pronounced
resonances in the local density of states of the electrodes. These authors
argued finally that the voltage dependence observed in the experiments
10 In this case the AMR signal is defined in terms of the conductance, see caption of
Fig. 12.11.
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
tute the starting point for the investigation the spin-dependent transport
through single-molecule junctions. As we shall see in the next part of book,
experiments of that kind have been already reported. Some of them are
exploring the spin injection in molecules with the use of ferromagnetic elec-
trodes, while others investigate how the molecular magnetism is reflected
in the transport properties of molecular junctions with non-magnetic elec-
trodes.
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
PART 4
355
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
356
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
Chapter 13
1 The basic properties of the main molecules explored so far in molecular electronics, as
well as their possible functionalities, are described in section 3.2.
357
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
2 Let us remind the reader at this stage that empirical methods like extended Hückel and
its descendents have played a fundamental role in quantum chemistry, but one cannot
expect these methods to give quantitative answers to the key questions in molecular
electronics, such as the position of the molecular levels, hybridization with the extended
states of the metallic leads, metal-molecule charge transfer, etc.
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
Table 13.1 Possible conduction mechanisms. Here, J is the current density, V is the bias
voltage, ϕB is the barrier height, d is the barrier length and T the temperature.
Direct “ √ ”
tunneling J ∼ V exp − 2d
~
2mϕB none J ∼V
Fowler-Nordheim „ √ 3/2
«
4d 2mϕB
tunneling J ∼ V 2 exp − 3q~V
none ln( VJ2 ) ∼ 1
V
Thermionic „ √ « “ ”
ϕ −q qV /4πǫd J 1
emission J ∼ T 2 exp − B k T ln T2
∼ T
ln(J) ∼ V 1/2
B
Hopping “ ” “ ”
conduction J ∼ V exp − kϕBT ln J
V
∼ 1
T
J ∼V
B
have in common that the I-V’s are rather insensitive to temperature and
they only differ in the voltage dependence.
The third mechanism, thermionic emission, is a process that takes place
when the electrons are excited over a potential barrier, as opposed to tun-
neling through it. This clearly has a very strong temperature dependence,
and will become significant when the potential barrier is relatively small.
Notice that, strictly speaking, thermionic emission is also a coherent mech-
anism since the electrons proceed elastically through the barrier without
losing their phase memory.
Hopping conduction is a mechanism in which electrons are localized at
certain points within the molecule, and can hop between those points. This
will also be a thermally activated process.5 This mechanism dominates the
transport properties of long molecules, except in some remarkable cases
such as carbon nanotubes.
Based on whether thermal activation is involved, the conduction mecha-
nisms fall into two distinct categories: (i) thermionic or hopping conduction,
which has temperature-dependent I-V characteristics, and (ii) direct tunnel-
ing or Fowler-Nordheimer tunneling, which does not exhibit temperature-
dependent I-V curves. According to this slightly oversimplified discussion,
one can conclude that if the I-V curves are temperature independent, the
dominant conduction mechanism is (coherent) tunneling. Moreover, the
transport regime can be discriminated by the analysis of the shape of the
I-V characteristics. It is important to recall that most experimental tech-
niques, especially those designed to work with single molecules, are not
suitable for temperature-dependent measurements. Thus, it may not be
easy to carry out the test proposed above to elucidate the transport mech-
anism.
The working principle stated in the previous paragraph has been used
in many different investigations to establish the conduction mechanism.
In Fig. 13.1 one can see an example taken from Ref. [130]. In this case,
the authors studied the transport through thiolated alkanes of different
length using the nanopore technique (see section 3.5.1). In this experiment
the transport through a self-assembled monolayer (SAM) was investigated.
Although our main interest is on single-molecule junctions, this experiment
is specially illustrative and it will be used several times in this chapter.
As one can see in Fig. 13.1(b,c), the current is rather insensitive to the
temperature and thus it was concluded that the conduction mechanism
(a) (b)
(c)
Fig. 13.1 (a) Schematics of a nanometer-scale device used in the experiments [130].
The structure of octanethiol is shown as an example. (b) Temperature-dependent I-V
characteristics of dodecanethiol (C12). I-V data at temperature from 300 to 80 K with
20 K steps are plotted on a logarithmic scale. (c) Arrhenius plot generated from the
I-V data in panel (b), at voltages from 0.1 to 1.0 V with 0.1 V steps. Reprinted with
permission from [130]. Copyright 2003 by the American Physical Society.
In spite of the quality of the fit shown in Fig. 13.2, there are a few
things that are not very satisfactory. First, an attempt to fit the results
with a rectangular barrier fails to describe the high-bias regime, see dashed
line in Fig. 13.2. This conclusion has been drawn in several analyses of
the transport through alkanethiol [458, 459]. This is the reason why α
was used above as an adjustable parameter, although its physical meaning
is not really clear. Second, the value obtained for the barrier height is
certainly small as compared with the expectations. This height, ϕB , is
in principle the distance between the Fermi energy of the electrodes and
the nearest molecular energy level in the molecule. For the combination of
Au contacts and alkanes, this distance is expected to be between 4 and 5
eV [299]. A possible way out for these problems has been pointed out by
Akkerman et al. [460]. These authors have shown that the description of
the transport through SAMs of alkenedithiols can be improved by including
the effect of image charges in the Simmons model (see brief discussion of
the role of image charges in section 4.4). They were able to describe the
transport in their experiments up to 1 V by using a single effective mass
and a barrier height. The barrier heights found were in the order of 4-5 eV
and, irrespective of the length of the molecules, an effective mass of 0.28 m
was determined in agreement with theoretical predictions [299].
Simmons model has been used in many other examples in molecular
electronics to interpret the observed I-V characteristics. For instance, an-
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
other beautiful example can be seen in Ref. [461], where the authors used
this model to explain the measured I-V curves in metal-molecule-metal
junctions formed from π-conjugated thiols, which were consistent with a
change in transport mechanism from direct tunneling to field emission.
Tunneling models, like Simmons one, borrowed from the field of metal-
lic tunnel junctions and semiconductor devices will continue to play an
important role in molecular electronics. However, their use is at least ques-
tionable. For instance, one may argue that one should use at least a double-
barrier model to describe a metal-molecule-metal junction since we have two
interfaces. Of course, such models are available, as we showed in Chapter
4. However, one could still argue that the bound states of a simple double-
barrier structure do not necessarily resemble those of a molecule. One could
go on trying to refine even further such barrier models, but it seems more
natural to use models that already incorporate the molecular features right
from the start. This is precisely the strategy that we are going to follow in
the rest of this chapter, where we shall introduce simple molecular-based
models to describe the transport in molecular junctions. In particular, we
shall start in the next section by studying the main conclusions that can
be drawn from the simple resonant tunneling model.
(a) (b)
11
00
00
11 11
00
00
11 1
0
0
1 1
0
0
1
00
11
00
11 00
11
00
11 0
1
0
1 0
1
0
1
00
11 00
11 0
1 0
1
00
11
00
11 00
11
00
11 ΓL 0
1
0
1 ε0 0
1
0 ΓR
1
00
11
00
11 00
11
00
11 0
1
0
1 0
1
0
1
00
11
00
11 00
11
00
11 0
1
0
1 0
1
0
1
00
11 00
11 0
1 0
1
00
11
00
11 00
11
00
11 0
1
0
1 0
1
0
1
00
11
00
11 00
11
00
11 L 0
1
0
1 0
1
0
1
L 00
11 00
11 R 0
1 0 R
1
00
11
00
11 00
11
00
11 0
1
0
1 0
1
0
1
00
11 00
11 0
1 0
1
00
11 00
11 0
1 0
1
Fig. 13.3 (a) Level scheme of a molecular junction. The molecule has a series of sharp
resonances corresponding to the different molecular orbitals, whereas the metal possess
a continuum of states that is filled up to the Fermi energy of the metal. (b) The same
as in panel (a) for a situation where the transport is dominated by a single level, ǫ0 .
7.4.1, following the spirit of the Landauer approach, the I-V characteristics
in this model can be computed from the following expression
2e ∞
Z
I(V ) = dE T (E, V ) [f (E − eV /2) − f (E + eV /2)] , (13.1)
h −∞
where the factor 2 is due to the spin symmetry of the problem, f (E) is the
Fermi function and T (E, V ) is the energy- and voltage-dependent trans-
mission coefficient given by the Breit-Wigner formula
4ΓL ΓR
T (E, V ) = . (13.2)
[E − ǫ0 (V )]2 + [ΓL + ΓR ]2
Here, the scattering rates are assumed to be energy- and voltage-
independent. This assumption can be easily relaxed, but it is usually a
good approximation for noble metals like gold with a rather flat density of
states around the Fermi energy. Notice also that we assume that the voltage
is applied symmetrically between the left and right electrode. Obviously,
this is irrelevant and the current only depends on the different of the chem-
ical potentials. The previous simple expressions will be our starting point
to discuss a few basic issues in the next subsections.
0.5
40 (a) (b)
0.4
20
I (µA)
G/G0
0.3
0
ΓL = ΓR = 0.1eV 0.2
-20 ΓL = ΓR = 0.05eV
ΓL = ΓR = 0.02eV 0.1
-40
0
-4 -2 0 2 4 -4 -2 0 2 4
V (V) V (V)
Fig. 13.4 (a) Current vs. bias voltage in the resonant tunneling model for a level position
ǫ0 = 1 eV (measured with respect to the Fermi energy of the electrodes) and at room
temperature (kB T = 0.025 eV). The different curves correspond to different values of the
scattering rates that are assumed to be equal for both interfaces. (b) The corresponding
differential conductance G = dI/dV normalized by G0 = 2e2 /h.
has a characteristic shape where one can distinguish three different regions.
We focus on the positive bias part. The first region is at low bias, when the
voltage is much smaller than |ǫ0 |, see Fig. 13.5(a). In this case the current is
quite low, specially if Γ is rather small. The second region is defined by the
resonant condition: eV /2 = ǫ0 (V ), i.e. eV = 2ǫ0 , see Fig. 13.5(b), where
the level is aligned with the chemical potential of one of the electrodes.
Here, when the voltage approaches this condition, the current is greatly
enhanced. Finally, when the voltage is larger than 2|ǫ0 | + Γ, the current
saturates to the value given by Isat obtained above, see Fig. 13.5(c).
As one can see in Fig. 13.4(b), the corresponding differential con-
ductance, G = dI/dV , exhibits two peaks at the resonant conditions
Fig. 13.5 Voltage dependence of the level alignment in the resonant tunneling model
for symmetric coupling. (a) Zero bias region, (b) resonant situation where the level is
aligned with the chemical potential of one of the electrodes and (c) large bias region
where the current saturates. The level has a finite broadening given by Γ = ΓL + ΓR .
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
(a) (b)
Fig. 13.6 (a) I-V curves of octanedithiol-based junctions measured with the appara-
tus shown in the inset. Here octanedithiol molecules are placed inside a monolayer of
octyl chains on top a gold surface. The molecules are contacted with a gold nanopar-
ticle, which in turn is contacted by the gold tip of a conducting AFM. From [125].
Reprinted with permission from AAAS. (b) I-V curves (solid lines) measured in molec-
ular junctions formed with the break junction technique at room temperature where a
trans-platinum(II) complex is contacted with gold electrodes (see inset). The curves
were fitted with a model for a rectangular barrier of height 2.5 eV (circles). Reproduced
with permission from [462]. Copyright Wiley-VCH Verlag GmbH & Co. KGaA.
0.15
6
4 (a) (b)
2 0.1
I (µA)
G/G0
0
-2 ΓL = ΓR = 0.1eV
ΓL = ΓR = 0.05eV
0.05
-4 ΓL = ΓR = 0.02eV
-6
0
-1.5 -1 -0.5 0 0.5 1 1.5 -1.5 -1 -0.5 0 0.5 1 1.5
V (V) V (V)
Fig. 13.7 The same as in Fig. 13.4 for low bias (|eV | < ǫ0 (V )).
Such tunneling curves can also be described with the resonant tunneling
model. If in Fig. 13.4 we focus on the low bias regime, i.e. before the
resonant condition is reached, one obtains the current and conductance
shown in Fig. 13.7. The similarity with the experimental curves is rather
obvious and by adjusting the parameters ǫ0 and Γ, one can in principle
fit those I-V curves. Anyway, it is important to emphasize that what it
is usually called a tunnel-like curve is nothing but a cubic function of the
form: I(V ) = AV + BV 3 , where A and B are constants. Almost any
tunneling model that produces symmetric I-V curves gives rise to such a
voltage-dependence at low bias and therefore it is suitable for fitting the
I-V characteristics in this regime.8 For this reason, if the I-V curves have
no much structure, one must be careful in interpreting the fits and one
should make sure that the values of the parameters obtained from the fits
are sensible.
1 2
(a) 1.5 (b)
0.8 kBT = 2 meV 1 I (nA)
kBT = 10 meV 0.5
kBT = 25 meV
I (µA)
0.6 0
0 0.25 0.5 0.75 1
V (V)
0.4 30
25 (c)
20
0.2 15 G (µS)
10
5
0 0
1.5 2 2.5 1.8 1.9 2 2.1 2.2
V (V) V (V)
Fig. 13.8 (a) Current-voltage characteristics computed with the resonant tunneling
model for different temperatures. The parameter values are: ΓL = ΓR = 2 meV and
ǫ0 = 1 eV. (b) Blow-up of the low bias region. Notice that the current is independent of
the temperature. (c) The corresponding differential conductance vs. voltage.
Eq. (13.1), tell us that the temperature dependence of the current or of the
conductance is determined by the energy dependence of the transmission
coefficient, which is usually not very pronounced. Thus, the temperature
dependence in the coherent regime, if any, is typically a power law, which
is clearly at variance with, for instance, the exponential behavior in the
incoherent hopping regime that takes place in very long molecules. In the
particular case of the resonant tunneling model, it is easy to see that if the
transmission is fairly energy-independent in the energy window controlled
by the voltage, then the current is insensitive to the temperature. This is
precisely what occurs at low bias when the level lies well above (or below)
the equilibrium Fermi energy of the system. Therefore, we can conclude
that the current (and also the conductance) is temperature independent in
an off-resonant situation.
The situation changes when the transport takes place on resonance. In
this case, if the temperature is comparable or larger than Γ, the current
depends on temperature. This is illustrated in Fig. 13.8 where we show
the I-V curves and the corresponding differential conductance for temper-
atures larger than the width of the resonance. As one can see, the current
and conductance depend on temperature for voltages around the resonant
condition, while at low bias they are insensitive to its value, see Fig. 13.8(b).
To be more precise, let us now study the temperature dependence of
the conductance in the linear regime. From Eqs. (13.1) and (13.2), one can
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
10
7.5 ΓL = ΓR
ΓL = 0.5ΓR
5 ΓL = 0.2ΓR
2.5 ΓL = 0.1ΓR
I (µA)
0
2
-2.5 1
-5 0
ΓL = 2 meV; ΓR = 20 meV
-1 ΓL = 20 meV; ΓR = 2 meV
-7.5
-2
-2 0 2
-10
-4 -2 0 2 4
V (V)
Fig. 13.9 Current-voltage characteristics in the resonant tunneling model for an asym-
metric situation for ǫ0 = 1 eV, ΓR = 20 meV and at room temperature (kB T = 25
meV). The different curves correspond to different values of the left scattering rate. The
inset shows very asymmetric situations where the scattering rates have bee interchanged.
Notice the that the I-V curves exhibit a clear rectification behavior.
coupled to the left and right electrodes. If the scattering rates ΓL and
ΓR are different, it is reasonable to assume that the voltage drops at the
interfaces accordingly to the ratio of the scattering rates. This can be simply
modeled by assuming that the voltage dependence of the level position is of
the form: ǫ0 (V ) = ǫ0 + (eV /2)(ΓL − ΓR )/Γ. This expression simply reflects
the fact that if one of the rates is much greater than the other, the level
follows the shift of the chemical potential of the electrode that is better
coupled.
With this simple model, we can now compute the I-V curves and an
example is shown in Fig. 13.9. Here, the different curves correspond to
different values of the ratio ΓL /ΓR . As we can see, when this ratio clearly
differs from one, the I-V curves become very asymmetric and the desired
rectification behavior becomes apparent. Notice that the polarity of the
curves can be controlled by exchanging the values of the scattering rates in
an asymmetric situation, as it is shown in the inset of Fig. 13.9.
It is easy to understand the shape of the I-V curves in Fig. 13.9. For
instance, if we focus on the situation where ΓL ≪ ΓR , the level is shifted
with the bias as ǫ0 (V ) = ǫ0 − eV /2, i.e. it follows the chemical potential
of the right electrode. Then, the resonant condition is reached for positive
voltages when the Fermi energy of the left electrode is aligned with the
level, i.e. when eV /2 = ǫ0 − eV /2, which implies ǫ0 = eV . For negative
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
voltages, since the level follows the right electrode, the resonant condition
is never reached and then the current for this polarity is much lower than
for positive voltages. These arguments explain the curve in Fig. 13.9 for
ΓL = 0.1ΓR . Using similar arguments, one can easily explain the other
curves in this figure.
It is worth pointing out that the asymmetry in the coupling can be due
to extrinsic factors, like a different coupling between left and right due to
an asymmetric configuration of the molecular junction, or it can be due to
something intrinsic, like the geometry of the molecule under investigation.
Thus for instance, an asymmetric molecule has molecular orbitals with an
asymmetric charge distribution. This induces a different coupling with the
electrodes, which can lead in turn to an asymmetric voltage profile. In both
cases, the final result is the observation of asymmetric I-V curves. For an
illustrative experimental example, we refer to the reader to Ref. [469].
Fig. 13.10 I-V curves of tolane-based molecules junctions measured with the micro-
fabricated MCBJ technique at room temperature and under liquid environment [472].
The molecules investigated are shown in the upper part: 4,4′ -bisthiotolane (BTT), 4,4′ -
bisnitrotolane (BNT) and 4,4′ -biscyanotolane (BCT). The black lines in the different
panels correspond to the experimental results, while the lighter lines are the fits to the
resonant tunneling model. The parameters used in the fits of these symmetric curves
(Γ = ΓL = ΓR ) are: Γ = 42 meV and ǫ0 = 404 meV for BTT, Γ = 93 meV and ǫ0 = 271
meV for BNT and Γ = 1.8 meV and ǫ0 = 558 meV for BCT. Courtesy of Artur Erbe.
anchoring group used to bind the molecules. Let us also stress that Zotti et
al. showed by means of ab initio DFT-based calculations that the use of the
resonant tunneling model was justified. To be precise, they showed that the
transport in these molecules is indeed dominated by a single molecular or-
bital that gives rise to a Breit-Wigner resonance close to the Fermi energy.
In particular, the transport was found to be dominated by the HOMO in
the case of the thiolated molecule, while the LUMO was found to be re-
sponsible for the conduction in the other two cases with nitro and cyano
(or nitril) groups. The implications of this work for the role of anchoring
groups in the transport through molecular junctions will be discussed in
section 14.2.
In the previous section we have assumed that the coherent transport was
completely dominated by a single molecular level. Of course, this is not
always the case. For instance, the Fermi level may lie more or less in the
middle of the HOMO-LUMO gap and then both molecular orbitals would
contribute to the transport. In other situations, we can have other levels
very close to the HOMO or to the LUMO contributing significantly to the
transport. For these reasons, we want to refine the resonant tunneling
model to include a second level. Our goal is to learn how the conductance
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
0110 11
00
00
11
10 00
11
1010 00
11
00
11
1010 00
11
00
11
10 t tH t11
00
1010 00
11
00
11
1010 ε0 ε0 00
11
00
11
10 00
11
1010 00
11
00
11
L 1010 00
11
00
11
R
10 00
11
1010 00
11
00
11
10 00
11
Fig. 13.11 Schematic representation of a two-level model where two sites with on-site
energy ǫ0 are coupled via a hopping tH . Each site is coupled to its closest electrode by
a hopping t (the same for both leads).
depends on the distance between the two levels and on the strength of the
coupling to the electrodes.
The model that we are about to describe is inspired by an important
example in molecular electronics, namely the transport through a hydrogen
molecule [569]. As we shall see in section 14.1.3, Smit and coworkers [127]
investigated the transport through hydrogen molecules with Pt contacts us-
ing the break junction technique. These authors concluded that a hydrogen
molecule can form a stable bridge between Pt electrodes and that such a
bridge has typically a conductance very close to the conductance quantum
G0 = 2e2 /h. Obviously, in this situation only two molecular levels can
participate in the transport, namely the bonding and antibonding state of
the hydrogen molecule.
With the hydrogen molecule in mind, we now proceed to analyze the
transmission in the model represented schematically in Fig. 13.11. In this
model we consider that the molecule is formed by two atoms with a single
relevant orbital per site. The on-site energy is denoted by ǫ0 and it is
assumed to be the same in both sites. The two sites are connected by a
hopping tH , while the symmetric coupling to the electrodes is described by
the hopping t. Notice that, for simplicity, we assume that the electrodes are
only coupled to its closest atom. The hopping tH is related to the splitting
between the bonding (ǫ+ ) and the antibonding state (ǫ− ) of the molecule,
namely ǫ± = ǫ0 ± tH . Thus, the HOMO-LUMO gap is simply 2tH in this
case. Obviously, within this model the conductance is made up of a single
channel because there is only one distinct path to cross the molecule.
The calculation of the zero-bias transmission is a simple exercise for
those who have followed the theoretical background (see Exercise 13.1).
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
1 Γ = 1.0tH
(a) (b) Γ = 0.6tH
0.8 1.5
DOS(E)
Γ = 0.4tH
T(E)
0.6 Γ = 0.2tH
1 Γ = 0.05tH
0.4
0.2 0.5
0 0
-2 -1 0 1 2 -2 -1 0 1 2
(E - ε0)/|tH| (E - ε0)/|tH|
Fig. 13.12 (a) Transmission as a function of the energy for the two-level model. The
different curves correspond to different values of the scattering rate Γ. (b) Total density
of states (DOS) projected onto the molecule, i.e. the sum of the local DOS in both sites
vs. energy.
one even in the energy region between the two molecular states. The first
limit describes the typical situation in many organic molecules in which
the Fermi energy lies somewhere in the HOMO-LUMO gap and the broad-
ening of the levels (0.01-0.5 eV) is clearly smaller than the gap (3-8 eV).
This is the reason why most organic molecules, even those with delocalized
orbitals, are poorly conductive. The opposite limit describes the situation
that occurs in strongly coupled systems such as the hydrogen molecule [127]
and other short organic molecules coupled to transition metals [473, 474],
where the linear conductance can be as high as 1 G0 . In these cases the
strong hybridization between the molecules and the electrodes (made of
Pt) provides a broadening to the molecular levels of several electronvolts,
which is in some cases comparable to the gap of the molecules or it simply
facilitates the resonant condition for the relevant orbital for transport (see
sections 14.1.3 and 14.1.4). Thus, almost irrespective of the exact position
of the Fermi energy, the transmission reaches a value close to unity. This
is, in simple terms, the explanation for the high conductance observed in
those examples.
Another simple two-level model is that in which the transmission is
assumed to be the sum of two independent Lorentzian functions. We shall
make use such a model in section 19.3 in our discussion of the thermopower
of molecular junctions.
One of the most studied issues in molecular electronics is the length de-
pendence of the conductance of molecular junctions. Typically the experi-
ments are restricted to low bias, but there are also studies of the influence
of a finite bias on this length dependence. Series of molecules like alkanes,
oligophenylenes, oligothiophenes, etc., have been extensively studied with
different techniques (see Ref. [41] for exhaustive list of references). The
most common finding is that conductance decays exponentially with the
length of the molecule, L, as
G(L) = Ae−βL , (13.8)
where the attenuation factor β depends on the particular type of molecule,
the presence of side groups, eventually on the bias voltage and not so much
on the anchoring group. Here, A is just a prefactor that determines the
order of magnitude of the conductance. Typical values of β range from 0.2-
0.4 Å−1 for conjugated molecules to 0.8-1.2 Å−1 for aromatic compounds.
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
0110 0110
1010 1010
1010 1010
1010 t 2,3 t N−1,N 1010 t
1010 t L t1,2 1010 R
1010 ε2 εΝ−1 εΝ
1010
1010 ε1 1010
1010 1010
L 1010 1010 R
1010 1010
1010 1010
Fig. 13.14 Schematic representation of the bridge model to explain the exponential
length dependence of the conductance. For further explanations, see text.
following the expectation of Simmons model and as it can be seen, the fit
is satisfactory. On the other hand, in order to compare with other results
reported in the literature, the authors also performed a fit to Eq. (13.8).
They obtained a β value from 0.83 to 0.72 Å−1 in the bias range from 0.1
to 1.0 V, which is comparable to results reported previously with other
techniques [458, 475, 476].
From an atomistic point of view, the exponential length dependence
of the conductance can be understood using a simple tight-binding model,
often used in the field of electron transfer [38]. Let us briefly explain the
main idea. The model is schematically represented in Fig. 13.14. In this
model a molecular bridge formed by N sites (or segments) with on-site
energies ǫi (only one orbital per side) is coupled to two metallic leads via the
hoppings tL,R . In the bridge we only consider nearest-neighbor hoppings
denoted by ti,i+1 . Notice that this model is simply the inhomogeneous
version of the model that we have used to explain the even-odd effect in
gold atomic chains in section 11.8.
Let us briefly remind how the transmission through the molecular bridge
can be calculated.11 Using the result of the Exercise 7.5 or the general
formulas derived in section 8.1, the zero-bias transmission coefficient can
be written as
where the ΓL,R are the scattering rates determining the strength of the cou-
pling to the metallic electrodes. Usually they do not have a very significant
energy dependence and we assume here that they are constant. Moreover,
Ga1N is the (advanced) Green’s function connecting the first and last site in
11 Those readers not familiar with the Green’s function techniques described in the second
part of the book can skip this discussion and go directly to Eq. (13.14).
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
the molecular bridge. In this sense, |Ga1N (E)|2 can be seen as the proba-
bility for an electron to propagate along the molecular wire. This function
can be calculated by taking the element (1, N ) of the following matrix (see
section 11.8)
−1
Ga (E) = [E a 1 − Hbridge − ΣaL − ΣaR ] , (13.12)
where E a = E − i0+ and Hbridge is the Hamiltonian of the molecular
bridge. Here, the only non-vanishing elements of the matrix self-energies
a
are (ΣaL )11 = t2L gL and (ΣaR )N N = t2R gR
a a
, where gL,R are the lead Green’s
functions (their exact expressions are irrelevant for our present discussion).
The scattering rates are giving by ΓL,R = t2L,R Im{gL,R a
}.
Rather than inverting exactly the previous N × N matrix, we compute
the first non-vanishing contribution to Ga1N . Obviously, this lowest-order
contribution corresponds to the sequential tunneling along the bridge with-
out any reflection. This is a good approximation to the exact expression in
the weak coupling regime, where max{ti,i+1 } ≪ min{|E − ǫi |}. Mathemat-
ically, this contribution can be written as
N −1
1 Y ti,i+1
Ga1N (E) ≈ . (13.13)
E a − ǫN i=1 E a − ǫi
For the sake of simplicity, we now assume that all bridge segments are
identical, i.e. ti,i+1 = t and ǫi = ǫ. Substituting the previous result into the
expression of the transmission, one obtains for the homogeneous bridge
¯ ¯2N
4ΓL ΓR ¯¯ t ¯¯
T (E) ≈ . (13.14)
|t|2 ¯ E − ǫ ¯
This result implies a simple form for the attenuation parameter of Eq. (13.8)
¯ ¯
2 ¯¯ E − ǫ ¯¯
β(E) = ln ¯ , (13.15)
a t ¯
where a measures the segment size, so that the bridge length is N a. Notice
that β is independent of the coupling to the leads and it is just determined
by intrinsic properties of the molecular bridge. The exponential dependence
on the bridge length is a manifestation of the tunneling character of this
process. Again, remember that the relevant energy for the linear conduc-
tance is the Fermi energy, EF . For typical values, e.g. |(EF − ǫ)/t| = 10
and a = 5 Å, Eq. (13.15) yields β = 0.92 Å−1 .
So in short, the general conclusion of our discussion is that the expo-
nential length dependence of the conductance is a signature of coherent
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
Fig. 13.15 (a) Structures of a subset of the biphenyl series studied, shown in order of
increasing twist angle or decreasing conjugation. (b) Conductance histograms of the
different molecules obtained with an STM at a bias voltage of 25 mV. (c) Position of
the peaks for all the molecules studied plotted against cos2 θ, where θ is the calculated
twist angle for each molecule, see Ref. [477] for more details. Reprinted by permission
from Macmillan Publishers Ltd: Nature [477], copyright 2006.
As we have discussed in sections 13.2 and 13.3, in most cases the coherent
transport through molecular junctions is determined by Breit-Wigner reso-
nances that originate from the different molecular orbitals. However, these
are not the only transmission line shapes that can be expected in molec-
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
0110 11
00
00
11 1
1010 00
11
(a) 00
11 (b)
1010 00
11
00
11
1010 ε 00
11
00
11
0.1
1010 ΓL ΓR11
00
t 00
11
T(E)
1010 00
11
00
11 0.01
1010 ε0 00
11
00
11
t = 0.0 eV
1010 00
11 t = 0.1 eV
00
11 t = 0.3 eV
L 1010 00
11 R 0.001
00
11 t = 0.6 eV
1010 00
11
00
11
1010 00
11
00
11 0.0001
-1 0 1 2 3
E (eV)
Fig. 13.16 (a) Schematic representation of a simple that illustrates the physics of Fano
resonances. Here, the resonant tunneling model of section 13.2 is modified by introducing
an additional level (ǫ) that is coupled to the resonant level, but not to the leads. (b)
Zero-bias transmission as a function of the energy for the model of panel (a) for ǫ0 = 2.0
eV, ǫ = 0.0 eV, ΓL = ΓR = 0.1 eV and different values of the coupling t.
ular junctions. In the last years, different authors have discussed the role
of quantum interference [259, 479–482] and, in particular, Fano resonances
[483–487] in the transport through molecular contacts. It has been shown
that these phenomena can give rise of transmission line shapes that dif-
fer significantly from the standard Breit-Wigner resonances of section 13.2.
As an example, in this section we shall briefly discuss the physics of Fano
resonances in molecular wires.
In 1961 U. Fano showed that in the context of the excitation spectra of
atoms and molecules, the interference of a discrete autoionized state with a
continuum gives rise to characteristically asymmetric peaks [488], which are
nowadays referred to as Fano peaks or resonances. The appearance of this
type of resonances in transport experiments have been discussed in several
contexts in mesoscopic physics ranging from one-dimensional waveguides to
Kondo impurities. In the context of molecular junctions, a Fano resonance
can appear in the transmission, for instance, due to the interplay between
extended molecular orbitals and states that are localized in a side group of
the molecule which is decoupled from the electrodes [484].
Following Ref. [484], we shall use the toy model schematically repre-
sented in Fig. 13.16(a) to explain the origin of Fano resonances. This model
is based on the resonant tunneling model and the ingredient is the presence
of an additional site (or energy level) that represents a side group that is
not directly connected to the electrodes. The coupling to the resonant level
is given by the hopping t and the level position of this “side group” is de-
noted by ǫ. The calculation of the zero-bias transmission in this model is a
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
simple exercise (see Exercise 13.5) and the final result reads
4ΓL ΓR
T (E) = 2 , (13.17)
[E − ǫ0 − t2 /(E − ǫ)] + Γ2
where Γ = ΓL + ΓR . This equation reduces to the Breit-Wigner formula
of Eq. (13.2) when the coupling element t vanishes. The main new feature
in this model is the appearance of an antiresonance at E = ǫ where the
transmission vanishes. This feature stems from a destructive quantum in-
terference between the direct path crossing the resonant level and a path in
which the electron “visits” the side group. Apart from this antiresonance,
the transmission exhibits two maxima at E = ǫ± , where ǫ± are given by
1n p o
ǫ± = (ǫ + ǫ0 ) ± (ǫ − ǫ0 )2 + 4t2 . (13.18)
2
In the limit t ≪ |ǫ − ǫ0 |, i.e. when the “side group” is weakly coupled to
the central backbone, the transmission exhibits a Breit-Wigner resonance
of width Γ in the vicinity of E = ǫ0 . Moreover, a Fano peak occurs near
the antiresonance (E = ǫ) separated from it by a distance of approximately
t2 /|ǫ−ǫ0 |. Thus, in this limit the hybridization with the weakly coupled side
group leads to the appearance of a peculiar asymmetric structure formed
by a peak followed by an antiresonance, which is the main fingerprint of
this phenomenon. Examples of those asymmetric line shapes are shown in
Fig. 4.73(b) in the limit of weak coupling (small t). Notice in particular
the dramatic change in the transmission that can go from 1 all the way
down to zero by changing slightly the energy. Obviously, in order to have
an impact in the transport properties, the Fano resonances needs to be
located close to the Fermi energy. If this is the case, they can give rise
to a pronounced structure in the I-V curves [480] or they can significantly
modify the thermoelectric properties of a molecular junction [487].
It is worth mentioning that an experimental situation that closely mim-
ics our simple model was reported in Ref. [489]. In this work, an artificial
quantum structure consisting of a single CO molecule adsorbed on a Au
chain was assembled by manipulating single Au atoms on NiAl(110) at 12
K with a STM. It was shown that the CO disrupts the delocalization of
electron density waves in the chain, as it suppresses the coupling between
neighboring chain atoms. In a subsequent paper, Calzoni et al. [490] showed
theoretically that the electronic properties of this system can be tuned by
the selective adsorption of small molecules. In particular, they showed that
a single CO group induces a quantum interference pattern that modulates
the electronic wave functions and modifies the coherent transport properties
of the system.
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
00
11 01
00
11 10 15
(a) 00
11
00
11
00
11
1010
10 (b) 1
11
00 10
10
00
11 10 0
00
11
00
11 1010 5
00
11 ∆ 1010 Gap
-1
00
11
I (µA)
00
11 10 -2 -1 0 1 2 3
00
11 ΓR 1010
0
00
11 ΓL
00
11
00
11 ε0 1010 -5
ΓL = ΓR = 10 meV
L 00
11
00
11 1010 R ΓL = ΓR = 20 meV 2∆/e
00
11 1010 ΓL = ΓR = 30 meV
00
11
-10
00
11 1010
00
11 -15
-4 -2 0 2 4
V (V)
and coworkers [512]. These authors showed that indeed one can have NDR
in these systems by means of coherent resonant tunneling. The presence
of a semiconductor band-edge leads to NDR when the molecular levels are
driven by the external potential into the semiconducting band-gap. We
now proceed to illustrate this mechanism with a simple model.
Let us consider once more the resonant tunneling model of section 13.2.
In order to describe a metal-molecule-semiconductor junction, we now as-
sume that there is a gap in, let us say, the right electrode, see Fig. 13.17(a).
The size of this gap is denoted by ∆. In a heterojunction like this one, it is
important to describe correctly the band-bending in the semiconductor and
the overall level alignment. We shall ignore these important details, in order
to emphasize the basic conceptual issues. We assume that the equilibrium
band-alignment is as shown in Fig. 13.17(a). Here, the Fermi energy lies
near the semiconductor valence band-edge, i.e. we assume that the semicon-
ductor is heavily p-type doped. The presence of a gap in the right electrode
strongly modifies the scattering rate, which in particular vanishes inside the
gap. We model this situation by a rate, ΓR , that is constant outside the
gap region and equal to zero at energies E ∈ [EF − eV /2, EF − eV /2 + ∆].
Here, we have already taken into account the shift of the chemical potential
of the right electrode induced by the bias voltage. The energy dependence
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
The goal of this chapter has been to describe and illustrate some basic
concepts related to the coherent transport through molecular junctions. It
is often believed in the context of molecular electronics that the theory is
unable to reproduce the experimental observations. We hope to have shown
that this judgment is unfair. We have been able to explain qualitatively a
variety of effects by simply using toy models and handwaving arguments.
A different story is our quantitative understanding that, as we shall see in
the next chapter, is not yet that satisfactory.
We also want to stress that there are other basic issues related to the
coherent transport that we have not covered in this chapter. Probably the
most important one is the issue of the electrostatic potential profile. We
have learned in this chapter that the position of the energy levels plays
a crucial role determining the current through a molecular junction. At
finite bias the energy levels are shifted with the voltage in a way that de-
pends on the exact electrostatic profile across the junction. Therefore, the
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
13.9 Exercises
13.1 Resonant tunneling model: Let us consider the resonant tunneling model
of section 13.2 for a symmetric situation (ΓL = ΓR = Γ/2). Calculate the current
at zero temperature up to third order in the bias voltage, i.e. determine the
relation I(V ) = AV + BV 3 at low bias and express the constants A and B as a
function of the two parameters of the model, namely Γ and ǫ0 .
13.2 Two-level model: Let us consider the two-level model of section 13.3.
(a) Use the general expressions derived in section 8.1, see Eq. (8.18) or (8.19),
to show that the zero-bias transmission is given by Eq. (13.7). Discuss also under
which conditions one recovers the expression of the transmission of the (single-
level) resonant tunneling model.
(b) Compute the I-V characteristics within the two-level model and discuss
the results. Hint: Assume that there is no voltage drop inside the molecular
bridge and that the scattering rates are independent of the energy.
13.3 The cos2 θ-law: Show that a matrix element (or hopping) between two
π-orbitals is proportional to cos θ, where θ is the angle formed by the axes of the
two orbitals. Hint: See discussion about two-center matrix elements in section
9.3.1.
13.4 Length and energy dependence of the transmission in molecu-
lar wires: In a series of papers Mujica and coworkers studied the conduction
through molecular wires using an effective tight-binding Hamiltonian (equivalent
to a Hückel model) [255–259]. They obtained the following interesting results for
the linear conductance of a molecular junction:
(1) The conductance achieves large (but bounded) values in the vicinity of any
of the wire energy eigenvalues.
(2) The conductance oscillates as a function of both injection energy and of wire
length when the electron is injected within the wire’s energy band.
(3) The conductance decreases exponentially with length when the electron is
injected outside the band of the wire.
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
Use the model for a molecular bridge discussed in section 13.4, see Fig. 13.14,
to demonstrate the previous conclusions. For this purpose, solve the model ex-
actly inverting Eq. (13.12), rather than using perturbation theory as we did in
our discussion in section 13.4. On the other hand, model the leads as semi-infinite
linear chains and use Eq. (5.46) for the Green’s functions of the outermost atoms
of the chains that are coupled to the molecular bridge.
13.5 Fano resonances: Show that the transmission in the model introduced
in section 13.6 is given by Eq. (13.17). Hints: (i) Use the general expression of
Eq. (8.18) to show that the zero-bias transmission can be written as T (E) =
4ΓL ΓR |Ga00 (E)|2 , where Ga00 (E) is the advanced Green’s function in the resonant
level. (ii) Show that Ga00 (E) = 1/{E − ǫ0 − t2 /(E − ǫ) − iΓ}. This result leads
directly to Eq. (13.17).
Finally, investigate the impact of Fano resonances on the current through a
molecular junction by computing the I-V curves within this model. For simplicity,
assume that the system is symmetrically coupled to the leads and that the voltage
drops occur at the metal-molecule interfaces.
13.6 NDR in metal-molecule-semiconductor junctions: Using the model
of section 13.7 show that one can encounter NDR at negative voltages using
a heavily n-type doped semiconductor. For this purpose, (i) assume that in
equilibrium the Fermi level lies near the edge of the conduction band of the
semiconducting lead and (ii) assume that in equilibrium the resonant level lies
above the semiconductor gap.
13.7 Transmission of a benzene junction: The goal of this exercise is to com-
pute the transmission as a function of energy for a metal-benzene-metal junction.
Use for this purpose the Hückel approximation for the benzene molecule described
in section 9.5.1 with ǫ0 = 0 for the on-site energy of the π-orbital in each carbon
atom and t = −2.5 eV for the hopping between neighboring atoms. Assume that
the benzene molecule is coupled to the leads through a single carbon atom in each
side [e.g. atoms 1 and 4 in Fig. 9.4(a)] and describe the strength of the coupling
with a scalar and energy-independent scattering rate Γ (the same for both inter-
faces). Calculate the zero-bias transmission as a function of energy within this
model for different values of Γ. Determine also the linear conductance assuming
that the Fermi energy is EF = 0 (i.e. it lies in the middle of the HOMO-LUMO
gap of the benzene molecule) and estimate the value of Γ necessary to reach a
conductance larger than 0.1 G0 .
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
Chapter 14
In the previous chapter we have learned that the coherent transport through
molecular junctions is determined by the strength of the metal-molecule
coupling as well as by the intrinsic properties of the molecules, including
their length, conformation, the HOMO-LUMO gap and the alignment of
this gap to the metal Fermi level. Moreover, we have shown that in many
cases the experimental observations can be explained by means of very
simple qualitative arguments. In this chapter we shall go on discussing the
coherent transport in single-molecule junctions, but from a more quanti-
tative point of view. Our goal is twofold. On the one hand, we want to
calibrate our present level of understanding and for this purpose, we shall
compare different experimental and theoretical results for various test sys-
tems. On the other hand, we shall illustrate some of the basic concepts
discussed in the previous chapter in more quantitative terms.
Bearing these goals in mind, we shall discuss in the next section the
results obtained so far for some test-bed molecules of special interest in
molecular electronics. Then, we shall review recent advances in the un-
derstanding of the role of the metal-molecule interface and the efforts to
chemically tune the conductance with the use of side-groups. Moreover,
we shall briefly describe a set of controlled experiments performed with
the STM, in which the junctions are fully characterized providing thus im-
portant test systems. We shall finish this chapter with a summary of the
main conclusions and some comments about the future challenges and open
problems.
Before getting started, let us say that the current status of the under-
standing of the electronic transport through molecular junctions has been
reviewed several times in this decade. In particular, we recommend the
following articles by Lindsay and collaborators [524–529].
391
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
1 The temperature dependence of the I-V curves in Refs. [534, 535] cannot be easily
explained within a coherent transport picture, unless such changes are related to the
thermal stability of the contacts.
2 These DFT results were obtained with the code TURBOMOLE v5.7 [575] using a split
(a)
(a) (b)
−1.42 eV LUMO
1.6 (b)
(c) WFs (0.28)
DZP (0.24)
SZP (0.21)
Transmission
1.2
0.8
−4.95 eV
0.4
HOMO
0
-6 -5 -4 -3 -2 -1 0 1 2
ε−εF (eV)
Fig. 14.2 (a) Frontier orbitals of a benzenedithiol (BDT) molecule as obtained from a
DFT calculation (see footnote 2). (b) Supercell used to model the central region of a
Au(111)-BDT-Au(111) junction with S at the fcc hollow site. (c) The calculated trans-
mission functions with two different methods and different basis sets. The transmission
at the Fermi level is indicated in the parentheses following the legends. Reprinted with
permission from [545]. Copyright 2008, American Institute of Physics.
shown in Fig. 14.2(c). In this case the leads are ideal Au(111) surfaces and
the S atoms were place at the minimum energy positions in the fcc hollow
sites. The linear conductance obtained in this case is ∼ 0.28 G0 in line with
the naive expectation and clearly higher that in the experiments.
we restrict ourselves to the linear alkanes. A brief discussion of the properties of these
molecules can be found in section 3.2.
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
Table 14.1 Values of the three main peaks (low, medium and high)
of the conductance histograms of Au-alkanedithiol-Au junctions. The
integer N indicates the number of C atoms in the molecule.∗
N =5 2.45×10−5 8.26×10−4 -
N =6 3.16×10−5 2.58×10−4 1.22×10−3
N =8 1.14×10−5 5.68×10−5 2.71×10−4
N =9 6.06 ×10−6 2.58×10−5 1.27×10−4
N = 10 2.84 ×10−6 5.81×10−6 2.17×10−5
the three peaks. The first one was found at Gl = 1.2 × 10−5 G0 ,5 which was
attributed to the conductance of a single-molecule junction. The other two
peaks appear at Gm = 4.5×10−5 G0 , and Gh = 2.3×10−4 G0 , see Fig. 14.3.
They found that the Gm has the strongest statistical weight, whereas Gh
is only observed in a non-contact mode, in which the electrodes do not get
into contact before each new molecular junction formation. They proposed
that these two values reflect the formation of several molecular junctions
in parallel between the electrodes.
Then, what is the linear conductance of Au-alkanedithiol-Au? In Table
14.1 we have reproduced the experimental results of Wandlowski’s group
obtained with STM break junctions for the three main peaks in the conduc-
tance histograms of alkanedithiol molecules of different length [555]. These
values show an exponential decay of the linear conductance as a function of
the number of C atoms (or length) for both the medium and the high peaks
with exponents of 0.94 and 0.96 per carbon atom (βN ), respectively. How-
ever, the low peak does not exhibit such an exponential decay, see Ref. [555]
for further details.
Let us discuss now how the transport takes place through alkane
molecules. As we explained in the previous chapter, the analysis of the I-
V characteristics in experiments involving alkane SAMs have shown clearly
that the transport mechanism is coherent tunneling [41, 130]. This has also
been confirmed in single-molecule experiments [557]. This is indeed what
is naively expected from the electronic structure of these carbon chains. In
Fig. 14.4 we have summarized some of the main features of such electronic
structure, as obtained from DFT-based calculations (see footnote 2). As
one can see, these molecules exhibit a very large HOMO-LUMO gap of
5 This peak was followed by several ones at multiples of Gl .
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
2
C8
0
LUMO
+1.09 eV
LUMO 9.2
Gap (eV)
Energy (eV)
-2
8.8
8
−7.28 eV 5 10 15
-6 N
HOMO
HOMO
-8
2 4 6 8 10 12 14 16
N
Fig. 14.4 Electronic structure of alkane molecules as computed from DFT (see text).
(a) Frontier orbitals (HOMO and LUMO) of octane (C8). (b) HOMO and LUMO levels
for alkanes of different length (N is the number of carbon atoms). The dashed line
indicates the approximate position of the Fermi energy of gold. The inset shows the
HOMO-LUMO gas vs. N.
more than 8 eV. The HOMO lies around 2-3 eV below the Fermi energy (or
negative work function) of gold.6 Thus, it is reasonable to assume that the
transport in Au-alkanedithiols-Au junctions takes place through the tails of
the HOMO of these molecules. This simple picture is basically confirmed
by the existent DFT-based calculations of the linear conductance of these
junctions [562, 563, 555]. However, there are still significant discrepancies
between the different theoretical studies, as we now proceed to explain.
The DFT-based study of Ref. [555] indicates that the conductance of
these junctions strongly depends on the binding geometry. These authors
proposed values of 0.83 and 0.88 for the attenuation factor per C atom
(βN ) for the medium and high conductance peaks, respectively, which is in
fair agreement with the experimental results reported in that work. They
also indicated that these exponents are sensitive to the functional used in
the DFT calculations and differences up to 20% between functional can be
expected. On the other hand, the estimates based on a complex band struc-
ture analysis, performed by Tomfohr and Sankey [299] and by Picaud et al.
[564], suggested βN ≈ 1.0 and 0.9, respectively. However, their estimates
for the tunneling barrier (distance between the HOMO of the molecule and
the gold Fermi energy) of 3.5-5.0 eV exceed the values of Ref. [555] by
a factor of 2. Another study by Müller [563] reported a comprehensive
6 The inclusion of thiol groups at the end of the carbon chains introduces states close
to the gold Fermi energy. These states are mainly localized in the sulfur atoms and
therefore, they are not expected to play a role in the conduction, at least for long
molecules. The situation may be different in the case of short alkanes.
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
Bonds
2
15.6 Ang
Angles
60
-60
1
Transmission
-2
10
-4
10
-6
10 0 5 10 15 20
11.0 13.0 15.0 17.0 19.0 Distance [Ang]
Fig. 14.5 (left) Snapshots of the formation of an octanedithiol molecular junction sim-
ulated using DFT-based molecular dynamics. As the junction is being stretched, the
molecule migrates into the junction and pulls out a short gold chain before finally break-
ing. (right) Calculated electron transmission probability as a function of stretching
distance. The number of Au-S bonds (defined by rAu−S < 3.3 Å) and dihedral angles
(0o ∼ straight molecule; 60o ∼ gauche defect) for the S-C8 -S chain are also shown.
Reprinted with permission from [566]. Copyright 2009 American Chemical Society.
along the last “plateau” is very sensitive to the contact geometry, and one
can observe upon stretching large variations in the conductance of an or-
der of magnitude when gauche defects are present. We show an example
of these simulations in Fig. 14.5 for octanedithiol (C8), where one can see
both the evolution of the contact geometry upon stretching and the cor-
responding transmission probability. From these simulations, the authors
constructed rudimentary conductance histograms from which they deduced
a value of βN = 1.19 and values of the conductance peaks of 2.2 ×10−3 ,
3.1 ×10−4 , and 2.0 ×10−5 in units of G0 for C6, C8, and C10, respectively.
Notice that these values tend to overestimate the experimental results of
Table 14.1 (see Ref. [566] for further details).
Fig. 14.6 Conductance curves and histograms for clean Pt, and Pt in a H2 atmosphere.
The inset shows a conductance curve for clean Pt (black) at 4.2 K recorded with a bias
voltage of 10 mV, before admitting H2 gas into the system. About 10000 similar curves
are used to build the conductance histogram shown in the main panel (black). After
introducing hydrogen gas the conductance curves change qualitatively as illustrated by
the grey curve in the inset, recorded at 100 mV. This is most clearly brought out by
the conductance histogram (grey; recorded with 140 mV bias). Reprinted by permission
from Macmillan Publishers Ltd: Nature [127], copyright 2002.
The physics behind the signatures of vibration modes and the impor-
tance of transport properties like the shot noise will be discussed in detail
in subsequent chapters, where we shall come back to this example.
All the observations detailed above offer a very stringent test to the the-
ory, which should explain consistently all the experimental results. Before
reviewing the existent work in the literature, it is interesting to discuss the
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
1.4 8 k-points
1.2 Γ-point
Transmission
1
0.8
Fermi level
0.6
0.4
0.2
0
-8 -6 -4 -2 0 2 4
Energy (eV)
Fig. 14.7 Calculated transmission for the molecular hydrogen contact shown in the in-
set. For comparison both the k-point sampled transmission and the Γ-point transmission
are shown. The wide plateau with T ≈ 1 extending across the Fermi level indicates a
single, robust conductance channel with nearly perfect transparency. Reprinted with
permission from [571]. Copyright 2005 by the American Physical Society.
Fig. 14.8 Conductance histograms for a Pt junction (black), and for Pt after introducing
benzene (filled) measured with the MCBJ technique. Each conductance histogram is
constructed from more than 3000 conductance traces recorded with a bias of 0.1 V during
repeated breaking of the contact. Reprinted with permission from [473]. Copyright 2008
by the American Physical Society.
o Ttot o
4.9 A 6.5 A
2 T1 1.2
T2 1
Transmission
1.5
T3 0.8
1 0.6
0.4
0.5
0.2
0 0
LDOS (1/eV)
HOMO1
0.4 HOMO2 0.4
LUMO1
LUMO2
0.2 0.2
0 0
-10 -8 -6 -4 -2 0 -10 -8 -6 -4 -2 0
E (eV) E (eV)
Fig. 14.9 Transmission and density of states (DOS) as a function of energy in a Pt-
benzene-Pt junction as calculated in Ref. [473]. The left panels show the results for
a geometry where the outermost Pt atoms were separated a distance of 4.9 Å, while
the right ones show the corresponding results for a separation of 6.5 Å. The contact
geometries are shown on top of these panels. The transmission plots show both the total
transmission and its decomposition into individual transmission coefficients, Ti . The
local DOS has been projected onto the four benzene frontier orbitals, which are shown
in the upper part of the figure. The vertical dashed lines indicate the position of the
Fermi energy (-5.4 eV). Courtesy of Sören Wohlthat.
to two channels and the frontier orbitals of the benzene acquired a large
broadening due to the strong interaction with the metallic leads. When
the elongation of the contact proceeds, the reduction of the metal-molecule
coupling becomes apparent in the transmission curve with the appearance
of a pseudo-gap around the Fermi energy (see right panels in Fig. 14.9).
In this case, the transport is dominated by a single conduction channel.
The reason for this is not really obvious from the information of the local
DOS. Then, what determines the number of channels in this case? As a
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
decouple the molecule from the leads, like in the case of the molecular transistors (see
next chapter).
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
Fig. 14.10 (a) Sample conductance traces measured with STM Au break junc-
tion without molecules and with 1,4-benzenediamine, 1,4-benzenedithiol, and 1,4-
benzenediisonitrile shown on a semilog plot. (b) Conductance histograms constructed
from over 3000 traces measured in the presence of the three molecules shown on a log-log
plot. The control histogram of Au without molecules is also shown. Inset: same data on
a linear plot showing a Gaussian fit to the peak (black curve). Adapted with permission
from [126]. Copyright 2006 American Chemical Society.
0
10 (a) (b)
120
100
-1
10
/h)
counts (ms/trace)
2
ctance (2e
10
-2 80
BDC60
60
10
-3 BDA
40
condu
-4
10
20 BDT
10
-5
0 clean solv ent
0 5 10 15 20 25 30 35 10
-5
10
-4
10
-3
10
-2
10
-1
10
0
ctance (2e
2
time (s) condu /h)
(a) (b) 1
0.8
G(θ)/G(θ=0)
0.6
S
0.4 N
N Exp.
0.2
0
0 0.2 0.4 0.6 0.8 1
2
cos θ
Fig. 14.12 Theoretical results on the conductance of biphenyl derivatives. (a) Molecules
studied capped with NH2 . The dark vertex in the backbone of molecule 3 corresponds
to N and the side groups of molecules 6 and 7 (other than H) correspond to F and
Cl atoms, respectively. (b) Zero-bias conductance in a hollow configuration, for sulfur
contact (circles), nitrogen contact (squares) and values from Ref. [477] (triangles). All
cases have been normalized to the θ = 0 value. Adapted with permission from [589].
Copyright 2008 IOP Publishing Ltd.
to flow through the system [478]. In this sense, Pauly et al. [546] have
shown theoretically that the conductance of oligophenylenes of different
length remains finite when the molecules are modified with methyl side-
groups, although these substituents induce a rotation of the neighboring
phenyl rings of about 90o . The typical reduction of the conductance, in
comparison with the conjugated molecules, is about two order of magni-
tude. Recently, Lörtscher et al. [590] have shown experimentally that such
non-conjugated molecules are still conductive.
The role of the conjugation in the conduction through molecular systems
can also be illustrated without resorting to side-groups. Thus for instance,
the comparison of the conductance through alkanes to that through proto-
typical molecular wires with extended π-electron states, like oligophenyle-
neethynylene (OPE) or oligophenylenevinylene (OPV), shows substantially
higher conductance through the conjugated molecules and a rational de-
pendence on the HOMO-LUMO gap [475, 591–593].
As mentioned above, the second main effect of side-groups is to shift
the frontier orbitals of a molecule. In this sense, side-groups can be used
to improve the usually bad alignment between the molecular levels and
the Fermi energy of the metallic electrodes. In other words, and using
terminology of semiconductor physics, one can use side-groups to “dope”
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
G (nS)
0.8
−6.62 −0.59 3
2
0.7
4
0.6 1
3 −5.23 +0.09
0.5
0.4 2
4 −5.97 +0.12
-7 -6.5 -6 -5.5 -5
HOMO (eV)
in-situ about the level alignment with the electrode Fermi energy, along
the lines of Ref. [461]. This is of course extremely challenging in the case
of single-molecule junctions, although not impossible as we shall see in
Chapter 20.
Another example of this doping effect was presented by Venkataraman
et al. [558]. In this case, the authors studied the single molecule con-
ductances of a series of very short conjugated molecules (substituted 1,4-
diaminobenzenes) using an STM-based break junction technique. They
found that electron donating substituents resulted in higher molecular con-
ductances, and there was an approximate correlation between the conduc-
tance and the Hammett σp parameter,9 consistent with hole transport (i.e.
transport dominated by the HOMO of the molecules). Another interest-
ing example related to the influence of side-groups has been reported by
Baheti et al. [102]. In this work the thermopower of molecular junctions
based on several 1,4-benzenedithiol (BDT) derivatives was investigated.
The BDT molecule was modified by the addition of electron-withdrawing
or -donating groups such as fluorine, chlorine, and methyl on the benzene
ring. It was found that the substituents on BDT generated small and pre-
dictable changes in conductance depending on their character. Moreover,
the authors showed that by replacing the thiol end groups by cyanide end
groups the transport changes radically and it turns out to be dominated by
the LUMO of the molecule. These results will be discussed in more detail
in Chapter 19 in the context of thermoelectricity in molecular junctions.
One of the major problems in most of the experiments that we have dis-
cussed so far is the fact that it is not easy to prove that one is dealing with
a single molecule. In principle, the STM constitutes an ideal tool to resolve
this issue.10 The STM can be utilized to perform controlled transport
experiments through individual molecules that have been deposited onto
metal surfaces by bringing the metallic tip into contact to the molecule. The
obvious advantage of the STM is that the structure under investigation–a
molecule along with its substrate–can be imaged with submolecular preci-
sion prior to and after taking conductance data. In this way, parameters
9 Roughly speaking, the Hammett parameter (or constant) describes the change in re-
action rates upon introduction of substituents. For a precise definition, see Ref. [595].
10 The STM as a tool to fabricate molecular junctions has extensively described in section
3.4.4.
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
Fig. 14.15 Schematic illustration of the I(s) STM method of forming molecular wires.
(A) A low coverage of the studied molecule is formed on the Au(111) surface, and the
set-point current is increased. (B) Attachment of the molecule at one end to the Au STM
tip is achieved, and then (C-D) the tip is retracted from the surface while recording the
current. The graph shows the conductance decay with distance for a clean Au substrate
(lower curve) and for a molecule on Au (upper curve). In the latter curve the different
stages of the contact formation are indicated (B, C and D). Notice the presence of a
plateau before rupture. Courtesy of Edmund Leary.
and subsequently break. It has been shown that both the I(s) and the I(t)
method result in the same single-molecule conductance for alkanedithiols
[550]. A nice application of the I(t) method can be found in Ref. [599],
where the authors showed, in combination with ab initio transport calcu-
lations, that the tilt-angle dependence of the electrical conductance is a
sensitive spectroscopic probe, providing information about the position of
the Fermi energy.
The use of methods in the spirit of the I(s) and the I(t) ones are opening
new ways of looking at molecular conductance. Thus for instance, Temirov
et al. [119] have reported beautiful results on a complex system, PTCDA
(4,9,10-perylenetetracarboxylic-dianhydrid), on a Ag(111) surface. They
demonstrated that one can controllably contact the molecule to the STM
tip at one of the four oxygen corner groups and peel the molecule gradually
from the surface. The conductance clearly varies in the process of peeling,
but when pulled to an upright position the conductance is approximately
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
Although in this chapter we have only talked about some concrete aspects
of single-molecule conduction, we can already draw a few general conclu-
sions and point out some of the main challenges for the near future. It
is clear that in the last years a significant progress has been made in the
experimental approaches to study single-molecule junctions as well as in
the qualitative understanding of their transport properties. From the ex-
perimental side, the introduction of statistical methods to determine the
conductance has partially eliminated the discrepancies between different
experimental results which appear when only individual traces are com-
pared. The use of new techniques to measure other transport properties
such as shot noise or thermopower provides very valuable additional infor-
mation that is not contained in the standard conductance measurements
(see Chapter 19). The use of low temperatures and the improvement in
the stability of the devices allow now making use of the inelastic tunneling
spectroscopy (see Chapter 16), which gives an essential information about
the presence of the molecules and the geometry of the junctions.
From the theory side, the development of ab initio methods makes now
possible to study both the mechanical and the electrical properties in a
much more reliable way. In particular, DFT-based calculations provide, for
instance, a detailed information about the possible structure of the con-
tacts, the relevant vibration modes and the conductance of the junctions.
These theoretical methods are now able to describe the general experimen-
tal trends, see e.g. Ref. [579], although they still fail in general to describe
quantitatively the transport results.
So in short, there are good reasons to be optimistic about the develop-
ment of this field. However, it has to be acknowledged that there are still
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
Chapter 15
15.1 Introduction
1 In
the resonant tunneling model, the injection gap is simply the energy ǫ0 of the level,
measured with respect to the Fermi energy.
423
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
time, which covers the different situations.2 Thus, if ~/τ is larger than the
energy scales associated with inelastic interactions like electron-electron,
U , or electron-vibration, λ, then the transport is mainly coherent. At the
contrary, if a molecule is weakly coupled to the electrodes (Γ < max{U, λ})
and the system is brought close to resonance (∆E ≈ 0), the transport will
very likely be dominated by the Coulomb interaction in the molecule or by
the excitation of internal degrees of freedom like vibrational modes.
There are by now many examples in nanophysics in which the elec-
tronic transport through a small object which is weakly coupled to metal-
lic electrodes has been explored. Let us mention, for instance, the cases
of semiconductor quantum dots, carbon nanotubes or metallic nanopar-
ticles. In all these systems, the transport in the weak-coupling regime is
governed by single-electron tunneling processes that lead to phenomena like
the Coulomb blockade effect. Moreover, if the coupling is not so weak, other
interesting many-body phenomena, like the Kondo effect, can show up at
show temperatures. We shall see in this chapter that these phenomena also
appear in single-molecule junctions. These effects have been understood
in great detail in different devices with the help of a gate electrode. This
third terminal is only capacitively coupled to the small object and it allows
to tune its energy level spectrum and to explore different charge (or redox)
states. The gate electrode allows us in turn to control the current that flows
through the system with an external field, very much like in the case of field-
effect transistors in microelectronics. Due to this analogy and also to the
fact the transport is usually dominated by single-electron processes, these
weakly coupled systems are known as single-electron transistors (SETs). In
the last decade it has become possible to incorporate a gate electrode into
single-molecule devices. We shall refer to these three-terminal molecular
devices as single-molecule transistors (SMTs).
The goal of this chapter is to discuss the electronic transport through
SMTs with special emphasis in the role of the Coulomb interaction in the
molecules. The role of the vibrational modes in these systems will be dis-
cussed in the next chapter. With this idea in mind, we shall first review
briefly the general conditions necessary to observe charging effects and we
shall recall the basic signatures of these effects in the transport characteris-
tics. Then, in section 15.3 we shall recall the main experimental techniques
that have been used so far to fabricate SMTs. The experimental results in
SMTs are often analyzed in the light of the “orthodox” theory of Coulomb
2 See Ref. [38] and references therein for a detailed discussion about the tunneling traver-
sal time.
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
Source Drain
Gate
V VG
Fig. 15.1 Schematic representation of a generic three-terminal device. The sphere rep-
resents the dot (or island), which is weakly coupled to the source and drain electrodes
by tunnel junctions. Finally, a third electrode (the gate) is capacitively coupled to the
island.
where m is the electron mass and N the number of electrons. The charac-
teristic energy scale is thus ~2 π 2 /(mL2 ). For a 1D box, the level spacing
grows for increasing N , in 2D it is constant, while in 3D it decreases as N
increases. The level spacing of a 100 nm 2D dot is ∼ 0.03 meV, which is
large enough to be observable at dilution refrigerator temperatures of ∼ 100
mK. Thus, dots made in semiconductor heterostructures are true artificial
atoms, with both observable quantized charge states and quantized energy
levels. Using 3D metals to form a dot, one needs to make nanoparticles as
small as ∼ 5 nm in order to observe atom-like properties. In the case of
molecular junctions, the spacing ∆E, which is basically the HOMO-LUMO
gap, is typically of the order of several electronvolts. Therefore, level quan-
tization should be easily observable in SMTs even at room temperature.
Now that we have identified the relevant scales for the occurrence of
charging effects, let us now see how they are revealed in the transport
characteristics. The tunneling of a single charge changes the electrostatic
energy of the island by a discrete value, a voltage VG applied to the gate
(with capacitance CG ) can change the island’s electrostatic energy in a con-
tinuous manner. In terms of charge, tunneling changes the island’s charge
by an integer while the gate voltage induces an effective continuous charge
q = CG VG that represents, in some sense, the charge that the dot would
like to have. This charge is continuous even on the scale of the elementary
charge e. If one sweeps VG , the build up of the induced charge will be com-
pensated in periodic intervals by tunneling of discrete charges onto the dot.
This competition between continuously induced charge and discrete com-
pensation leads to the so-called Coulomb oscillations in a measurement
of the current (or conductance) as a function of gate voltage at a fixed
source-drain voltage.
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
(a) (b)
(c) (d)
Fig. 15.2 Coulomb blockade in a single-wall carbon nanotube. (a) AFM image of a
carbon nanotube on top of a Si/SiO2 substrate with two 15-nm-thick Pt electrodes, and
a corresponding circuit diagram. The total length of the tube is 3 µm, with a section
of 140 nm between the contacts to which a bias (source-drain) voltage is applied. A
gate voltage Vgate applied to the third electrode in the upper-left corner of the image is
used to vary the electrostatic potential of the tube. (b) Current versus gate voltage at
Vbias = 30 µV. Two traces are shown that were performed under the same conditions.
(c) Current-voltage curves of the tube at a gate voltage of 88.2 mV (trace A), 104.1
mV (trace B) and 120.0 mV (trace C). (d) Conductance G = I/Vbias versus ∆Vgate
at low bias voltage Vbias = 10 µV and different temperatures. Solids lines are fits of
G ∝ cosh−2 (e∆Vgate /α2kB T ), corresponding to the model of a single molecular level
that is weakly coupled to two electrodes. The factor α is the gate coupling parameter (see
text) and for this peak equals 16. Reprinted by permission from Macmillan Publishers
Ltd: Nature [613], copyright 1997.
Most of the experiments described in the previous two chapters in our dis-
cussion of the coherent transport have been performed with two-terminal
devices fabricated with the break-junction technique and the STM. These
techniques have several advantages, but it is however very difficult to in-
corporate a gate electrode in their set-ups. This is an important drawback
since, as discussed in the previous section, a gate electrode allows us to
extract much more information about the junctions. Thus for instance, the
gate makes possible to study the conduction through molecules in differ-
ent transport regimes by bringing the energy levels into and out of reso-
nance with the Fermi energy. This way, one can also probe excited states
and different charge states can be accessed. Excited states can either be
vibrational [22, 617, 678], electronic [618], or related to spin transitions
[619, 620]. These excitations serve as a fingerprint of the molecule under
study.
An important parameter in three-terminal devices is the gate coupling
parameter, α. This parameter quantifies the shift of the orbital levels that
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
a b Gold deposition
angle
Mask
Molecule
Gold
SOURCE DRAIN
Al2O3 electrodes
Al Gate
Al2O3
Al Gate
c d
Molecule Gold
Countersupport
electrodes Gold
particle Molecule
SiO2
SOURCE DRAIN
Doped Si Pushing rod
SiO2
Doped Si Gate
Fig. 15.3 Schematic diagrams of different three-terminal device techniques. (a) Elec-
tromigrated thin metal wire on top of a Al/Al2 O3 gate electrode. (b) Angle evaporation
technique to fabricate planar electrodes with nanometer separation on top of a Al/Al2 O3
gate electrode. (c) Gated mechanical break junction. (d) The dimer contacting scheme
(see text). Reprinted with permission from [606]. Copyright 2008 IOP Publishing Ltd.
fined on top of a gate electrode [see Fig. 15.3(d)]. According to the authors,
this dimer-based contacting scheme provides several advantages such as the
ability to fabricate single-molecule devices with high certainty in which the
contacts to the molecule are well defined. The gold particles in this set-up,
however, efficiently screen the gate potential. Moreover, at low tempera-
tures spectroscopic features of the gold particles were sometimes observed
to be superimposed on the characteristics of the molecule conduction.
the name quantum dot to refer generically to a small island weakly coupled to the source
and drain.
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
11
00
00
11 11
00
00
11
L 00
11
00
11 00
11
00
11 R
00
11
00
11 00
11
00
11
00
11 00
11
EF 00
11
00
11 00
11
00
11
00
11
00
11 Ep 00
11
00
11
00
11 00
11
00
11
00
11 00
11
00
11
00
11 00
11
00
11
00
11 00
11
00
11
EF
00
11
00
11 00
11
00
11
eV 00
11 00
11
00
11 00
11
ηeV−eVG
Fig. 15.4 Schematic drawing of the energy level diagram and electrostatic potential
profile of a generic quantum dot. The dot possesses a single-particle spectrum with
discrete levels, Ep . The Fermi levels in the left and right reservoirs are indicated. We
measure the levels Ep with respect to EF , which from now on we set to zero. The single
particle spectrum may be shifted by the external potential. Here, η is the portion of the
bias voltage that drops at the right interface and VG corresponds to the gate voltage.
This expression for the total energy summarizes the constant interaction
model, in which the capacitance C does not vary with N .
It is important to clarify the meaning of the external potential in
Eq. (15.5). This can be done with the help of the equivalent circuit shown
in Fig. 15.5. Elementary electrostatics gives the following relation between
the different potentials and the charge Q on the island:
CVdot − CS VS − CD VD − CG VG = Q, (15.7)
where C = CS + CD + CG . Comparing this expression with Eq. (15.4) we
arrive at the following result for the external potential
Vext = (CS VS + CD VD + CG VG )/C. (15.8)
Thus, we see that the potential on the dot depends on the induced potential
Vext of the source, drain and gate. Notice that the change in the external
potential due to a change in the gate voltage carries a factor α = CG /C,
which is the gate coupling parameter that was mentioned in the previous
section. On the other hand, assuming that the drain is grounded as in
Fig. 15.5, the factor η introduced in Fig. 15.4 can now be simply expressed
as the capacitance ratio η = CS /C.
A key quantity determining whether the current can flow through the
dot is its chemical potential, which is the minimum energy required to add
an extra electron to the dot. From Eq. (15.6) it is easy to see that this
chemical potential is given by
1 e2
µ ¶
µdot (N ) = Edot (N ) − Edot (N − 1) = N − − eVext + EN . (15.9)
2 C
Before discussing the main predictions of this theory, it is important to
be aware of the conditions for which the constant interaction model gives
CS CD
Source Vdot Drain
CG
Gate
VS VG
Fig. 15.5 Schematic representation of the capacitance model of a quantum dot. The dot
is connected to source and drain electrodes with tunnel junctions and the gate electrode
shifts the electrostatic potential of the dot. Here, we assume that drain electrode is
grounded (VD = 0).
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
a reliable description of the device. This is first of all weak coupling to the
leads. A second condition is that the size of the device should be sufficiently
large to make a description with single values for the capacitances possible.
Finally, the single-particle spectrum Ep should not vary with the charge
N residing on the dot. The constant interaction model works well for
weakly coupled quantum dots for which it is very often used. However,
the previous conditions are not fulfilled in general in molecular devices.
Neither the charging energy nor the energy level spectrum are expected
to be independent of the number of electrons in the molecule, specially
for small ones. Thus, the constant interaction model should be used with
caution in this case.
e2 e2
∆µ(N ) = µdot (N + 1) − µdot (N ) = + EN +1 − EN = + ∆E, (15.11)
C C
where ∆E = EN +1 − EN is the level spacing mentioned in section 15.2.
In the absence of charging effects, the addition energy, ∆µ(N ), is deter-
mined by the irregular spacing ∆E of the single-electron levels in the quan-
tum dot. The charging energy e2 /C regulates the spacing, once e2 /C & ∆E.
If there is spin degeneracy of the levels, it is lifted by the charging energy.
In the limit (e2 /C)/∆E → 0, Eq. (15.11) is the usual condition for reso-
nant tunneling. In the limit (e2 /C)/∆E → ∞, Eq. (15.11) describes the
periodicity of the classical Coulomb-blockade oscillations in metallic islands
where the level spacing is negligible [609]. In molecular physics the related
energies are defined as A = Edot (N ) − Edot (N + 1) for the electron affinity
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
Fig. 15.6 Potential landscape through a quantum dot. The states in the contacts are
filled up to the electrochemical potentials µL and µR , which are related by the external
voltage VSD = (µL − µR )/e. The discrete single-particle states in the dot are filled
with N electrons up to µdot (N ). The addition of one electron to the dot raises µdot (N )
(i.e. the highest solid curve) to µdot (N + 1) (i.e. the lowest dashed curve). In (a) this
addition is blocked at low temperatures. In (b) and (c) the addition is allowed since here
µdot (N + 1) is aligned with the reservoir potentials by means of the gate voltage. (b)
and (c) show two parts of the sequential tunneling process at the same gate voltage. (b)
shows the situation with N and (c) with N + 1 electrons on the dot.
(a)
current
)
−V
C
VG
Eadd /e
β(
∆E
VSD
0
α Eadd/e
γ ( VC
N N+1
−VG
)
(b)
VG
Fig. 15.7 (a) Generic two-dimensional plot of the current as a function of bias and
gate voltage (stability diagram) for a quantum dot in the Coulomb blockade regime. For
small bias current only flows in the three degeneracy points indicated with circles. Upper
shadow region: positive currents. Lower shadowed region: negative currents. White:
blockade, no current. The dotted lines indicate the presence of excitations (see text).
(b) Measured stability diagram of a metallic single-walled carbon nanotube showing the
expected fourfold shell filling. Blockade regime is white. Reprinted with permission from
[626]. Copyright 2005 by the American Physical Society.
see dotted lines in the upper panel of Fig. 15.7. At such a line, a new
(electronically or vibrationally) excited state enters the bias window, cre-
ating an additional transport channel. The result is a step-wise increase of
the current and a corresponding peak in the differential conductance. The
energy of an excitation, ∆E in Fig. 15.7(a), can be determined by reading
off the bias voltage of the intersection point between the excitation line and
the Coulomb diamond edge through the same argument that we used for
finding addition energies. The excitations correspond to the charge state of
the Coulomb diamond they end up in [see Fig. 15.7(a)]. The width of the
lines in the dI/dVSD plot (or, equivalently, the voltage range over which the
step-wise increase in the current occurs) is determined by the larger one
of the energies kB T and Γ, which in the Coulomb blockade regime must
be the first one. In practice, this means that sharp lines and thus accurate
information on spectroscopic features are obtained at low temperatures and
for weak coupling to the leads.
There are other important issues like the role of the asymmetry in the
coupling that can be discussed at a qualitative level. For more details,
recommend the review of Ref. [605].
of occupation numbers, one for each energy level. Notice that the restric-
tion kB T, ∆E ≫ Γ results in the conductance being much smaller than
e2 /h. We also assume conservation of energy in the tunneliing processes,
thus neglecting contributions of higher order in Γ from tunneling via virtual
intermediate states in the quantum dot. We finally assume that inelastic
scattering takes place exclusively in the reservoirs — not in the quantum
dot. The effect of inelastic scattering in the quantum dot is considered in
Ref. [624] (see also Exercise 15.4).
Energy conservation upon tunneling from an initial state p in the quan-
tum dot (containing N electrons) to a final state in the left reservoir at
energy Epf,l (in excess of the local electrostatic potential energy), requires
that7
Epf,l (N ) = Ep + U (N ) − U (N − 1) − (1 − η)eV. (15.15)
Here η is the fraction of the applied voltage V which drops over the right
barrier,8 see Fig. 15.4. The energy conservation condition for tunneling
from an initial state Epi,l in the left reservoir to a final state p in the quantum
dot is
Epi,l (N ) = Ep + U (N + 1) − U (N ) − (1 − η)eV, (15.16)
where N is the number of electrons in the dot before the tunneling event.
Similarly, for tunneling between the quantum dot and the right reservoir
one has the conditions
Epf,r (N ) = Ep + U (N ) − U (N − 1) + ηeV, (15.17)
Epi,r (N ) = Ep + U (N + 1) − U (N ) + ηeV, (15.18)
where Epi,r
and Epf,r
are the energies of the initial and final states in the
right reservoir.
The stationary current through the left barrier equals that through the
right barrier, and is given by
∞
e X X (p)
ΓL P ({ni }) δnp ,0 f (Epi,l (N )) − δnp ,1 f¯(Epf,l (N )) . (15.19)
£ ¤
I=
~ p=1
{ni }
where we have used the shorthand notation f¯(E) ≡ 1 − f (E). The second
summation is over all realizations of occupation numbers {n1 , n2 , . . .} ≡
7 Let us remind that the energies E are measured with respect to the Fermi energy of
p
the leads.
8 Notice that this definition differs from the one of Ref. [624]. This change has been
introduced to preserve the convention our convention about the current direction and
the sign of the bias voltage.
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
{ni } of the energy levels in the quantum dot, each with stationary proba-
bility P ({ni }). here, the numbers ni can take only the values 0 and 1. In
equilibrium, this probability distribution is the Gibbs distribution in the
grand canonical ensemble " ̰ !#
1 1 X
Peq ({ni }) = exp − Ei ni + U (N ) − N EF , (15.20)
Z kB T i=1
P
where N ≡ i ni , and " Z is the
̰ partition function given!#by
X 1 X
Z= exp − Ei ni + U (N ) − N EF . (15.21)
kB T i=1
{ni }
The non-equilibrium probability distribution P is a stationary solution
of the kinetic equation
∂
~ P ({ni }) = 0
∂t h i
(p) (p)
X
=− P ({ni })δnp ,0 ΓL f (Epi,l (N )) + ΓR f (Epi,r (N ))
p
h i
(p) (p)
P ({ni })δnp ,1 ΓL f¯(Epf,l (N )) + ΓR f¯(Epf,r (N ))
X
−
p
X
+ P (n1 , . . . np−1 , 1, np+1 , . . .)δnp ,0
p
h i
(p) (p)
× ΓL f¯(Epf,l (N + 1)) + ΓR f¯(Epf,r (N + 1))
X
+ P (n1 , . . . np−1 , 0, np+1 , . . .)δnp ,1
p
h i
(p) (p)
× ΓL f (Epi,l (N − 1)) + ΓR f (Epi,r (N − 1)) , (15.22)
The kinetic equation, Eq. (15.22), for the stationary distribution function is
equivalent to the set of detailed balance equations (one for each p = 1, 2, . . .)
(p) (p)
P (n1 , . . . np−1 , 1, np+1 , . . .)[ΓL f¯(Epf,l (Ñ + 1)) + ΓR f¯(Epf,r (Ñ + 1))]
(p) (p)
= P (n1 , . . . np−1 , 0, np+1 , . . .)[ΓL f (Epi,l (Ñ )) + ΓR f (Epi,r (Ñ ))], (15.23)
P
with the notation Ñ ≡ i6=p ni . A similar set of equations formed the basis
for the work of Averin, Korotkov, and Likharev on the Coulomb staircase
in the non-linear I-V characteristics of a quantum dot [629–631].
Eq. (15.22), together withX the normalization condition
P ({ni }) = 1, (15.24)
{ni }
form a set of linear algebraic equations that can be easily solved numerically.
For those readers not familiar with rate or master equations, we recommend
Exercise 15.1, in which a single-level dot is considered.
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
natural to recover this result, since in the Coulomb blockade theory detailed
in this section we have only considered elastic tunneling processes. Indeed,
it is sometimes difficult to distinguish experimentally between Coulomb
blockade effect and coherent transport through a weakly coupled system.
From the theory side, there is an obvious difference. While in the coherent
case the electrons tunnel through the single-particle levels of the dot, in
the Coulomb blockade regime the resonances are also determined by the
charging energy.
15
(b)
e/C + ∆E/e
dI/dV (µS)
10
0 1 2
(a) 5
E2
∆E
L E1 R 0
0 50 100 150 200 250 300
αVG (mV)
0.3
0.2 (c)
0.1
I (µA)
0
αVG = 0 mV
-0.1 αVG = 50 mV
-0.2 αVG = 100 mV
-0.3
-400 -200 0 200 400 600
VSD (mV)
Fig. 15.8 (a) Two-level model to illustrate the transport characteristics of a quantum
dot in the Coulomb blockade regime within the constant interaction model. The single-
particle energies are E1 = 50 meV and E2 = 80 meV (measured with respect to the
equilibrium chemical potential of the leads). The excitation energy is thus ∆E = E2 −
E1 = 30 meV. The charging energy is chosen to be e2 /C = 100 meV and all tunneling
rates are assumed to be equal to Γ = 1 meV. The temperature is kB T = 2.5 meV (i.e.
T ≈ 30 K) and η = 0.6. (b) Differential conductance vs. the gate voltage (including the
gate coupling constant α) corresponding to the model of panel (a). The source-drain (or
bias) voltage is 20 µV (linear regime). The numbers 0, 1 and 2 indicate the number of
electrons in the different regions separated by the Coulomb peaks. (c) Corresponding
non-linear current-voltage characteristics for several gate voltages.
series of steps, which correspond to the opening of new channels when the
reservoir chemical potentials cross the different resonances in the dot. Thus
for instance, at VG = 0 and positive bias voltage, the current is blocked
until µdot (N = 1) equals the chemical potential of the left reservoir, i.e.
µdot (N = 1) = e2 /2C + E1 + ηeVSD = eVSD . This occurs at VSD = 250
mV. Then, the next step corresponds to the crossing of the excited state
without changing the net charge in the dot. This requires an additional bias
voltage equal to ∆E/e(1 − η) = 75 mV, which explains the appearance of
a step at VSD = 325 mV. Following this line of reasoning, one can explain
the position of all the steps in the I-V curves.
Two additional features in Fig. 15.8(c) are worth mentioning. First,
notice that the gap in the low-bias voltage region can be completely closed
by increasing the gate voltage, in accordance with the Coulomb oscillations.
Second, notice that the I-V curves are not symmetric with respect to the
inversion of the bias voltage. This is simply due to the fact that we have
chosen an asymmetric electrostatic profile with η 6= 0.5.
(iii) Stability diagrams: As we discussed in section 15.4.3, the different
energy scales and capacitances of the problem can be extracted from the
so-called stability diagrams, where either the current or the differential
conductance are plotted as a function of both the gate voltage and the
bias voltage. In Fig. 15.9 we show the stability diagrams for our two-level
example, which nicely illustrate the main conclusions of our qualitative
discussion in section 15.4.3. In particular, notice that the addition energy
can be extracted from the height of the middle diamond or from its width
(distance between two consecutive degeneracy points). On the other hand,
the energy of the excitation, ∆E = 30 meV, can be read off from the
bias voltage of the intersection point between the excitation line and the
Coulomb diamond edge. The excitation line is particularly visible in the
diagram of the differential conductance. Finally, notice that the diamonds
are “inclined” due to the asymmetric potential profile (η = 0.6).
There are two assumptions of the constant interaction model that are not
met in general in SMTs. Neither the charging energy nor the level spectrum
are expected to be independent of the number of electrons in a molecule. In
this sense, the orthodox theory of Coulomb blockade may not be adequate
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
Fig. 15.9 Stability diagrams corresponding to the example of Fig. 15.8. (a) Current
vs. gate voltage and source-drain (or bias) voltage. (b) Differential conductance vs. gate
voltage and source-drain voltage.
for SMTs. The natural question is now how to generalize the standard
theory to deal with molecular devices. It is worth mentioning that this
question has also emerged in the other contexts like few-electron quantum
dots [628] and ultrasmall metallic grains [633].
Part of the answer to this question is rather simple, at least conceptually
speaking. Any theory of Coulomb blockade in SMTs should include an
appropriate description of the molecular many-body spectrum as a function
of the number of electrons in the molecule. In other words, we need as a
starting point the ground state and excited states of the molecules not
only for the neutral species, but also for the stable cations and anions.
In principle, this requires the use of ab-initio (post Hartree-Fock) quantum
chemistry methods like, for instance, the configuration interaction approach
[176]. In practice, both model Hamiltonians and approximate methods like
density functional theory10 have been used for this purpose. Once the
many-body spectrum for different number of electrons is known, one needs
to solve a master equation to determine the occupation of the different
states and finally their contribution to the current. By now many authors
10 Density functional theory is not designed to give the level spectrum of a system and,
although it may give reasonable results for neutral molecules, it has severe problems with
charged species (anions and cations) [273].
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
11 This will be done in section 17.3.1 in the context of the study of the influence of the
where the matrix W̃ ˆ is identical to Ŵ but with one (arbitrarily chosen) row
s0 replaced with (Γ, · · · , Γ) and ~v is a vector, vs = Γδs,s0 . The transition
rates Wss′ , with s 6= s′ (in the absence of bosonic coupling) are the sum
r
P
Wss′ = r Wss ′ of the Golden rule rates for the tunneling of an electron
to/from electrode r = L, R:
X fr (Es − Es′ )¯ P tr hs|c† |s′ i¯2 Ns′ < Ns
( ¯ ¯
r i i iσ
Wss′ = 2πρe 2 , (15.31)
f¯r (Es′ − Es )¯ i tri hs|ciσ |s′ i¯ Ns′ > Ns
¯ P ¯
σ
20 0.3
(a) U/e 0.2 (b)
dI/dV (µS)
15
0.1
I (µA)
10 0 1 2 0
αVG = 0 mV
-0.1 αVG = 25 mV
5
-0.2 αVG = 50 mV
0 -0.3
0 50 100 150 200 -400 -200 0 200 400
αVG (mV) V (mV)
Fig. 15.10 (a) Differential conductance vs. the gate voltage (including the gate coupling
constant α) for a molecular transistor described with the Anderson model. Here, ΓL =
ΓR = 1 meV, ǫ0 = 50 meV and U = 100 meV. The temperature is kB T = 2.5 meV (i.e.
T ≈ 30 K) and the bias voltage is 20 µV (linear regime). The numbers 0, 1 and 2 indicate
the number of electrons in the different regions separated by the Coulomb peaks. (b)
Corresponding non-linear current-voltage characteristics for several gate voltages. The
voltage is assumed to drop symmetrically at the interfaces.
to the spin degeneracy, the I-V curves only exhibit two plateaus. Notice
that the two steps (for a given voltage polarity) are separated by a distance
U/(eη).
In this example the transport characteristics are very similar to those
obtained with the constant interaction model (orthodox theory). The differ-
ences become more pronounced when there are more charge states involved
or additional quantum numbers play a fundamental role.
So far in this chapter, we have focused on the limit where the tunnel cou-
pling, Γ, is much smaller than any other energy scale in the problem. If
this coupling strength is increased, higher-order tunneling processes begin
to give a significant contribution to the transport properties [649]. In the
opposite limit (strong coupling regime) where Γ ≫ e2 /C, ∆E, kB T , the
electronic states in the molecule and electrodes are strongly hybridized. In
that case, as we have discussed in previous chapters, the elastic coherent
tunneling dominates transport and signatures of the Coulomb blockade are
washed out by quantum fluctuations of the molecular charge. Between the
weak coupling and strong coupling regime one can identify a third regime
which we shall refer to as the intermediate coupling regime. In this regime
it is still possible to observe Coulomb diamonds, but higher-order processes
lead to a non-negligible current inside the blockade regions. In this section
we shall discuss three different types of higher-order tunneling processes:
elastic and inelastic cotunneling and spin-flip cotunneling. This latter pro-
cess is behind the appearance of the Kondo effect.
Fig. 15.11 Elastic cotunneling process. The N th electron on the dot jumps to the drain
(virtual state) to be immediately replaced (final state) by an electron from the source
(black arrow sequence). A similar process involves the unoccupied state (light arrow
sequence). In both examples, an electron is effectively transported from source to drain.
olate energy conservation. The final state then has the same energy as the
initial one, but one electron has been transported through the molecule.
This elastic cotunneling process is analogous to the superexchange mecha-
nism in chemical electron transfer theory [582]. It occurs at arbitrarily low
bias as the energy of the tunneling electron and the molecule are unchanged
and leads to a nonzero background conductance in the blockade regions.
A cotunneling event that leaves the molecule in an excited state is called
inelastic. An example of such a process is depicted in Fig. 15.12. As one can
see, the onset of the cotunneling event occurs at eVSD = ∆E, the condition
dictated by the energy conservation principle. In transport measurements,
inelastic cotunneling appears in the stability diagram inside the Coulomb
diamonds as two symmetric lines running parallel to the gate axis as repre-
sented by grey lines in Fig. 15.14(a). Their energy, ∆E, is the distance of
the excitation to the zero-bias axis as illustrated in Fig. 15.14(a). Further-
more, the inelastic cotunneling line is expected to intersect, at the diamond
boundary, the corresponding excitation line inside the single-electron tun-
neling region [650].
It is worth stressing that that higher-order coherent processes appear as
sharp spectroscopic features as the conductance of the i-th order process is
proportional to Γi , while for first-order incoherent single-electron tunneling,
the current is proportional to Γ.
If the electron spin is taken into account, one can encounter another
elastic cotunneling process connected to the Kondo effect. This will be
analyzed in detail in the next subsection.
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
Fig. 15.12 Inelastic cotunneling process. For eVSD ≥ ∆E, the N th electron on the dot
may jump from the ground state to the drain (virtual state) to be immediately replaced
by an electron from the source (final state), which enters the excited state.
energy
10 ε0 10 10 0
1 0
1 0
1
10 10 10 0
1 0
1 0
1
10 10 10 0
1 0
1 0
1
1010 1010 1010 0
1 0
1 0
1
L 1010 µdot (N) 1010 R L 1010 µdot (N) 1
0
0
1 R L 0
1
0
1 µdot
(N)
0
1
0
1 R Γ
1
0 0
1 0
1
1010 1010 1010 0
1 0
1 0
1
1010 1010 1010 0
1
0
1 0
1
0
1 0
1
0
1
0
1 0
1 0
1
Fig. 15.13 (a) Spin-flip cotunneling. A spin-up electron jumps out of the dot (virtual
state) to be immediately replaced by a spin-down electron (final state). (b) Kondo
resonance in the density of states that appears as a consequence of the spin-flip tunneling
processes.
for the new electron to have spin opposite to the first to be very high.
This spin exchange qualitatively changes the energy spectrum of the
system. When many such processes are taken together, one finds that a
new state, known as Kondo resonance, is generated exactly at the Fermi
level, see Fig. 15.13(b).12 In this situation, the localized spin is completely
screened and the many-boy ground state turns out to be a singlet state
(S = 0). It is important to note that the Kondo state is always “on reso-
nance” since it is fixed to the Fermi energy. Even though the system may
start with an energy, ǫ0 , which is very far away from the Fermi energy, the
Kondo effect alters the energy of the system so that it is always on reso-
nance. For this reason, these many-body correlations can lead to a great
enhancement of the conductance. The only requirement for this effect to
occur is that the system is cooled to sufficiently low temperatures below
the Kondo temperature TK (see next paragraph).
The width of the Kondo resonance is proportional to the characteristic
energy scale for Kondo physics, the so-called Kondo temperature TK . For
ǫ0 ≫ Γ, TK is given by [668]
√ · ¸
ΓU πǫ0 (ǫ0 + U )
kB TK = exp . (15.35)
2 ΓU
Here, Γ is the coupling strength and U can be seen as the charging en-
ergy, e2 /C. Typical Kondo temperatures are TK ∼ 1 K for semiconductor
quantum dots [655], TK ∼ 1 − 10 K for carbon nanotubes [660, 661] and
TK ∼ 20 − 50 K for molecular devices [23, 24, 662, 664]. This increase of
TK with decreasing dot size can be understood from the prefactor, which
12 Insection 6.9 we presented a discussion of the origin and description of the Kondo
resonance in the framework of the Anderson model.
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
Fig. 15.14 Schematic representation of the main characteristics of the Kondo effect in
electron transport through a molecular quantum dot. (a) In the stability diagram, the
Kondo effect results in a zero-bias resonance (white line) for an odd number of electrons
in the dot. Inelastic cotunneling excitations appear as lines running parallel to the gate
axis at finite bias. (b) For T ≪ TK , the full width at half-maximum (FWHM) of the
Kondo resonance is ∼ kB TK . (c) Temperature dependence of the Kondo-peak height in
the middle of the Coulomb diamond. Reprinted with permission from [606]. Copyright
2008 IOP Publishing Ltd.
contains the charging energy (U = e2 /C). Notice that the Kondo temper-
ature depends on the position of the level and therefore it can be tuned in
three-terminal devices by means of the gate voltage.
The theoretical description of the Kondo effect is very challenging. The
reason is that below TK , high order spin-flip processes contribute signifi-
cantly to both the electronic structure and the transport properties. This
implies that one needs to employ non-perturbative methods to describe
properly this phenomenon. Different many-body methods have been used
to account for the Kondo correlations in quantum dots and related struc-
tures. The description of such techniques is out of the scope of this book
and in the rest of this subsection we shall concentrate ourselves on the dis-
cussion of its main transport characteristics and refer the interested reader
to Refs. [651, 652, 669, 670, 665, 666] for more details about the theory.
The Kondo effect is manifested in the stability diagrams as a zero-
bias resonance in the differential conductance, dI/dVSD , versus VSD , in-
side the Coulomb diamond connecting both degeneracy points as shown
in Fig. 15.14(a). For an even number of electrons with all spins paired,
S = 0 and there is no Kondo resonance. This even-odd asymmetry is very
helpful in assigning the parity of the charge state which can then add extra
information to the understanding of the spectroscopic features observed in
the stability plots. In the low temperature limit (T ≪ TK ), the full width
at half-maximum (FWHM) of the Kondo resonance, as observed in a plot
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
tances were measured in the 10-100 nS range and different kinds of nonlinear
I-V curves with steplike features were reproducibly obtained. An example
of the results of these measurements can be seen in Fig. 15.15. Notice
that the I-V curves resemble very much the Coulomb staircase observed in
quantum dots. Indeed, the authors were able to fit the experimental results
within the framework of the ortodox Coulomb blockade theory described
in section 15.4, taking into account the discrete nature of the electronic
spectrum of the molecule. Let us mention that charging energies of the
order of 0.2 eV were used in the fits (see Ref. [676] for more details).
As we have discussed in previous sections, an unambiguous confirmation
that characteristics like the ones shown in Fig. 15.15 are a consequence of
the occurrence of charging effects requires the implementation of a gate elec-
trode, which is very challenging in the case of MCBJs. To our knowledge,
the first three-terminal single-molecule experiment was reported by Park et
al. [22] in 2000. These authors prepared single-C60 junctions by depositing
a dilute toluene solution of C60 onto a pair of connected gold electrodes
fabricated using e-beam lithography. A gap of 1 nm between these elec-
trodes was then created by electromigration [21]. The entire structure was
defined on a SiO2 insulating layer on top of a degenerately doped silicon
wafer which served as a gate electrode that modulates the electrostatic po-
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
Fig. 15.16 I-V curves obtained at T = 1.5 K from a single-C60 transistor fabricated with
the electromigration technique. The curves corresponds to five different gate voltages.
The inset shows a schematic diagram of an idealized single-C60 transistor. Reprinted by
permission from Macmillan Publishers Ltd: Nature [22], copyright 2000.
Fig. 15.17 Different conductance plots as a function of the bias voltage V and the gate
voltage Vg obtained from four different devices. The dark triangular regions correspond
to the conductance gap, and the bright lines represent peaks in the differential conduc-
tance. The arrows mark the point where the conductance lines intercept the conductance
gap. Reprinted by permission from Macmillan Publishers Ltd: Nature [22], copyright
2000.
molecule in this geometry can exceed 150 meV. This value is much larger
than in semiconductor quantum dots.
Notice that in the stability diagrams of Fig. 15.17 there are running
lines that intersect the main diamonds or conductance gap regions. As
we explained in previous sections, this indicates the presence of internal
excitations of the C60 molecules. The energies of these excitations (of a few
meV) are too small to correspond to electronic excitations. Moreover, some
of these lines are observed for both charge states and multiple excitations
with the same spacing are observed (see Fig. 15.16). These observations
suggest that these lines may correspond to the excitation of vibration modes
of the C60 molecules. The lowest-energy mode is known to have around
33 meV and this could explain some of the lines seen in the experiment.
However, internal vibrational modes cannot account for the observed 5-
meV features in Fig. 15.17. The authors of Ref. [22] suggested that this
line could correspond to the excitation of the center-of-mass oscillation of
C60 within the confinement potential that binds it to the gold surface.14
14 The signatures of the excitation of vibration modes in the transport characteristics
will be discussed in detail in the next two chapters.
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
(a) (b)
V (mV)
V (mV)
Vg (V) Vg (V)
nomena that were well-known in other nanodevices, but they have also
made possible to access new transport regimes and to discover novel physi-
cal phenomena. In the context of Kondo physics, we would like to mention
the work of Pasupathy et al. [663] where the Kondo effect in the presence
of ferromagnetism has been reported for the first time. In this work the
authors measured the transport through single-C60 transistors with ferro-
magnetic nickel electrodes. They showed that Kondo correlations persisted
despite the presence of ferromagnetism, but the Kondo peak in the differ-
ential conductance was split by an amount that decreased (even to zero)
as the spin polarizations in the two electrodes were turned from parallel to
antiparallel alignment. Although, the reported splitting was too large to
be explained by a local magnetic field, the voltage, temperature, and mag-
netic field dependence of the signal agreed with predictions for an exchange
splitting of the Kondo resonance [680, 681].
SMTs have also allowed to study the interplay between Kondo physics
and the electron-vibration interaction. The signatures of vibrational modes
have been shown to persist in the Kondo regime [617, 664, 682] and we
shall discuss this issue in certain detail in the next chapter. It is also worth
mentioning that although most of the experiments on the Kondo effect in
molecular junctions have been performed with the electromigration tech-
nique, the Kondo physics has also been studied with breakjunctions by
Ralph’s group [682]. Although in this case a gate electrode was not opera-
tive, the authors could tune the Kondo resonance in a single-C60 junction
by adjusting the metal-molecule distance that is a capability of the break-
junction technique that is lacking in electromigration-based experiments.
Changing the metal-molecule coupling the authors were able to tune the
Kondo temperature and showed that the temperature dependence of the lin-
ear conductance agreed with the scaling function expected for the S = 1/2
Kondo problem [682].
SMTs have also been used to explore other basic aspects of the Kondo
physics. Thus for instance, Roch et al. [684] have recently reported the
observation of a quantum phase transition between a singlet and a triplet
spin state at zero magnetic field in a single-C60 transistor. The analysis
of the transport through three-terminal molecular devices has also allowed
to study the fundamental scaling laws that govern the non-equilibrium the
standard S = 1/2 Kondo effect [683].
Another aspect to which SMTs have contributed enormously is the un-
derstanding of the role of vibrational modes in the transport through single
molecules. The signatures of the excitation of vibronic degrees of freedom
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
Fig. 15.21 (a) Side view of a Mn12 molecule with tailor made ligands containing acetyl-
protected thiol end groups (R=C6 H4 ). The diameter of the molecule is about 3 nm.
(b) Schematic drawing of the Mn12 molecule (circle) trapped between electrodes. A
gate can be used to change the electrostatic potential on the molecule enabling energy
spectroscopy. (c) Scanning electron microscopy image of the electrodes. The gap is not
resolvable. Scale-bar corresponds to 200 nm. Reprinted with permission from [619].
Copyright 2006 by the American Physical Society.
are specially visible in the transport characteristics in the limit of weak cou-
pling between the molecule and the metallic electrodes [22, 685]. Moreover,
in this regime the electron-vibration interaction can lead to a great variety
of novel physical phenomena. This subject will be discussed in detail in the
next chapter.
We now turn to a class of experiments where the transport through
single-molecule magnets (SMMs) has been investigated (see Ref. [686] for a
progress article on this subject). This type of molecules exhibits magnetic
hysteresis due to their large spin and high anisotropy barrier, which ham-
pers magnetization reversal [687, 688]. The first transport experiment on
a SMM was performed by the group of van der Zant [619]. These authors
studied the prototypical SMM, Mn12 acetate, which has a total spin S = 10
and an anisotropy barrier of about 6 meV. The molecules that were inves-
tigated were [Mn12 O12 (O2 C-R-SAc)16 (H2 O)4 ] (Mn12 from now on), where
R={C6 H4 , C15 H30 }, see Fig. 15.21(a). These molecules were designed with
thiol groups in the outer ligand shell to ensure a strong affinity for gold
surfaces. On the other hand, the ligands are believed to serve as tunnel
barriers, so that the molecules are only weakly coupled electronically to the
gold and their magnetic properties are preserved. The molecules were in-
corporated in a SMT geometry with gold electrodes using electromigration,
see Fig. 15.21(b).
In Fig. 15.22(a) we reproduce some of the results of this work for the
differential conductance as a function of gate (Vg ) and bias voltage (Vb )
for one of the devices (T = 3 K, R=C6 H4 ). The lines separating the con-
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
Fig. 15.22 (a) Differential conductance (gray-scale) as a function of gate voltage (Vg )
and bias voltage (Vb ) (T = 3 K, R=C6 H4 ). A region of complete current suppression
(left degeneracy point, arrow) and low-energy excitations with negative differential con-
ductance (right degeneracy point) are observed. The dashed line near the left degeneracy
point indicates the suppressed diamond edge. (Gray-scale from -0.8 nS [black] to 1.4 nS
[white]). (b) I − Vb at the gate voltage indicated in (a) with a line. NDC is clearly
visible as a decrease in |I| upon increasing |Vb |. Upon applying a magnetic field, current
is increased for negative bias. Reprinted with permission from [619]. Copyright 2006 by
the American Physical Society.
15.8 Exercises
15.1 Rate equations in a single-level model: For those who are not familiar
with rate (or master) equations, it is convenient to start by analyzing the following
situation. Let us consider a quantum dot with a single (non-degenerate) level of
energy E1 , which is measured with respect to the equilibrium Fermi energy of
the leads, which we set to zero. (The energy E1 can depend on the gate voltage,
the exact electrostatic profile and the charging energy). This dot has only two
possible configurations with n1 = 0 (empty dot) and n1 = 1 (one electron in level
E1 ). We shall denote the corresponding probabilities as P0 and P1 , respectively.
As usual, we denote the left and right tunneling rates (in units of energy) as ΓL
and ΓR , respectively, and we assume them to be energy-independent.
(a) Write down the kinetic equation for the probability distribution and show
that in the stationary case the probabilities Pi are given by
ΓL f¯L + ΓR f¯R ΓL fL + ΓR fR
P0 = , P1 = .
ΓL + Γ R ΓL + Γ R
Here, fL,R = f (E1 ∓ eV /2), where V is the bias voltage and f (E) is the Fermi
function.
(b) Use the previous solution to show that the current through the dot can
be written as
e ΓL ΓR
I= [fL − fR ] .
~ ΓL + Γ R
Notice that this expression coincides with the expression for the current obtained
in the single resonant tunneling model in the limit of weak coupling.
(c) Using the previous expression, show that the linear conductance is given
by Eq. (15.26).
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
„ «
eV
P ({ni }) ≡ Peq ({ni }) 1 + Ψ({ni }) .
kB T
Linearize the detailed balance equation (15.23) and solve it to show that Ψ can
be written as
∞
!
(i)
X ΓR
Ψ({ni }) = constant + ni (i) (i)
−η ,
i=1 ΓL + Γ R
where the constant first term takes care of the normalization of P to first order
in V and it does not need to be determined explicitly. Hint: Use the following
relations:
where the prime symbol in the Fermi function stands for derivative with respect
to its argument.
(b) Linearize the formula for the current in Eq. (15.19) and use the expression
for Ψ({ni }) to obtain Eq. (15.26).
15.3 Coulomb oscillations, Coulomb staircase and stability diagrams:
The goal of this exercise is to compute transport characteristics in the Coulomb
blockade regime within the two-level model discussed in section 15.4.4.2.
(a) As a first step, compute the occupation probabilities of the four possible
configurations of the dot. For this purpose, show that the stationary kinetic
equation, Eq. (15.22), together with the normalization condition of Eq. (15.24)
can be written in the following matrix form: Ŵ p ~ = ~v . Here, p~ is the column
vector containing the probabilities of the four configurations of the dot, i.e. p ~T =
(P1 , P2 , P3 , P4 ), where 1 ≡ (0, 0), 2 ≡ (1, 0), 3 ≡ (0, 1) and 4 ≡ (1, 1). The vector
~v is simply given by ~v T = (1, 0, 0, 0) and the different elements of the matrix Ŵ
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
(p)
Here, the ΓL,R (p = 1, 2) are the tunneling rates, while the expressions for the
energies appearing in the arguments of the Fermi functions can be found in section
15.4.4.
(b) Solve numerically the 4 × 4 system Ŵ p ~ = ~v and use the expression of
Eq. (15.19) to reproduce the results of Figs. 15.8 and 15.9.
(c) In a molecular transistor the level splitting ∆E may be larger than the
charging energy, e2 /C. Study how the stability diagrams in this case differ from
those shown in Figs. 15.9. Choose for instance e2 /C = 30 meV and ∆ = 100
meV, while keeping the other parameters equal to those in the example of section
15.4.4.2.
(d) An important experimental issue is that for a particular charge state
lines are often only visible on one side of the Coulomb diamond. This is due to
an asymmetry in the coupling. Illustrate this fact with the example of section
15.4.4.2 by choosing very different tunneling rates for the left and right barriers.
15.4 Effects of inelastic scattering in the Coulomb blockade regime: In
the Coulomb blockade theory described in section 15.4.4 inelastic scattering was
assumed to take place exclusively in the reservoirs. One of the effects of inelastic
scattering in the dot is the thermalization of the electrons inside the dot. In the
limiting case of full thermalization, the probability distribution function P ({ni })
is given by the equilibrium expression of Eq. (15.20). Use this expression in the
example of section 15.4.4.2 (and of the previous exercise) to study the effect of
inelastic scattering in the different transport characteristics (Coulomb oscillations,
Coulomb staircase and stability diagrams).
15.5 Coulomb blockade theory for single-molecule transistors: The goal
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
r
W21 = Γr fr (E2 − E1 )
r
= −Γr f¯r (E2 − E1 ) + fr (E4 − E2 )
ˆ ˜
W22
r
W24 = Γr f¯r (E4 − E2 )
r
W31 = Γr fr (E3 − E1 )
W33 = −Γr f¯r (E3 − E1 ) + fr (E4 − E3 )
r ˆ ˜
r
W34 = Γr f¯r (E4 − E3 )
r r
W42 = Γr fr (E4 − E2 ) = W43
W44 = −Γr f¯r (E4 − E2 ) + f¯r (E4 − E3 )
r ˆ ˜
r r r r
W14 = 0 = W23 = W32 = W41 ,
Chapter 16
Vibrationally-induced inelastic
current I: Experiment
16.1 Introduction
1 The term “phonon” in this chapter is used for vibrational modes associated with any
nuclear motion.
473
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
should have a clear idea about: (i) what are the basic experimental signa-
tures of vibrational modes in the current through single-molecule junctions,
(ii) what are the physical mechanisms giving rise to those signatures and
(iii) what are the main open problems related to this subject.
In this first chapter we shall describe some of the main experiments that
have illustrated the role of vibrations in the electrical conduction through
molecular junctions. We have grouped these experiments in three differ-
ent categories. First of all, we shall discuss situations where the electron-
phonon interaction is weak, in the sense explained above, and the electron
tunneling is off-resonant. The analysis of the vibronic signatures in this
regime is known as inelastic electron tunneling spectroscopy (IETS) for the
historical reasons that will be explained in section 16.2. Then, we shall fo-
cus our attention in section 16.3 on the case of highly conductive junctions,
where the electron-phonon interaction is also weak. In this regime, and
again for historical reasons, the study of the vibrational modes is known as
point-contact spectroscopy (PCS). Section 16.4 is devoted to a discussion
of the relation between IETS and PCS. In section 16.5 we shall discuss the
third group of experiments that correspond to the regime sometimes known
as resonant inelastic electron tunneling. This regime corresponds to a situ-
ation where the transport is resonant and the electron-phonon interaction
can be very strong. This regime is realized, in particular, in the molecular
transistors described in the previous chapter. The discussion below will
end with a brief summary of the main vibrational signatures that can be
observed in the different transport regimes. If you are an impatient reader
(as we are), please feel free to jump directly to section 16.6 and then come
back to this point.
The recent progress in the understanding of vibrational effects in molec-
ular transport junctions has been thoroughly described by Galperin, Ratner
and Nitzan in the review of Ref. [695], which contains close to 500 references
related to the main subject of these two chapters. For those who prefer a
quick overview, we recommend them the shorter review of Ref. [696] of the
same authors.
Fig. 16.1 Recorded traces of d2 I/dV 2 versus voltage for three Al-Al oxide-Pb junctions
taken at 4.2 K. The zero of the vertical scale is shifted for each curve, and all three are
normalized to the same arbitrary units. The largest peaks represent increases of 1% in
the conductance. Also indicated are intervals associated with the energy of IR-active
molecular vibrational modes. Curve A is obtained from a “clean” junction. Curves B and
C are obtained from junctions exposed to propionic acid [CH3 (CH2 )COOH] and acetic
acid (CH3 COOH), respectively. The peaks positions are independent of the polarity.
Reprinted with permission from [697]. Copyright 1966 by the American Physical Society.
EF
−hω hω eV
hω eV
(c) 2 2
d I/dV
−hω
hω eV
Vacuum/
Tip Molecule Metal
Fig. 16.2 (a) Schematic representation of the inelastic tunneling above the threshold
for a vibrational excitation. An electron can tunnel losing part of its energy which is
employed to excite a vibration mode of energy ~ω. This process is only possible when
eV ≥ ~ω. (b) The opening of the inelastic channel gives rise to an increase in the
conductance at eV = ±~ω. (c) The onset of the inelastic process is seen in the second
derivative of the current, d2 I/dV 2 , as a peak (dip) for positive (negative) bias.
(2) Spectral linewidth: The full width at half maximum (FWHM) of the
d2 I/dV 2 vibrational peak is given by W = [(1.7Vm )2 + (5.4kB T /e)2 +
WI2 ]1/2 , where Vm is the modulation voltage in the lock-in technique,
kB is the Boltzmann constant, T is the temperature, and WI is the
intrinsic width (due to the finite phonon lifetime) [705, 706].
(3) Selection rules: Although there are no selection rules in IETS as there
are in infrared (IR) and Raman spectroscopy, certain selection prefer-
ences have been established. According to the IETS theory [707, 708],
molecular vibrations with net dipole moments perpendicular to the in-
terface of the tunneling junction have larger peak intensities than vibra-
tions with net dipole moments parallel to the interface (for dipoles close
to the electrodes). For a more complete description of the propensity
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
Fig. 16.3 STM inelastic tunneling spectra of acetylene molecules. The plot shows back-
ground difference d2 I/dV 2 spectra for C2 H2 (1) and C2 D2 (2), taken with the same STM
tip. Notice the presence of peaks at 358 mV and 266 mV, respectively. The difference
spectrum (1 − 2) yields a more complete background subtraction. From [709]. Reprinted
with permission from AAAS.
Soon after the invention of the STM, it was clear that this tool could
serve to extend IETS all the way down to single molecules. However, this
turned out to be very challenging since it requires the use of low tempera-
tures (∼ 4 K) and very high mechanical stability. The breakthrough came
from Ho’s group that reported in 1998 the first study of the vibrational
spectra for a single molecule adsorbed on a solid surface [709]. To be pre-
cise, these authors measured the inelastic electron tunneling spectra for an
isolated acetylene (C2 H2 ) molecule adsorbed on the copper (100) surface
using a STM under UHV conditions at a temperature of 8 K. They observed
an increase in the tunneling conductance at 358 mV, which was attributed
to the excitation of the C-H stretch mode. The increase in conductance is
typically rather small (around 3-6% in these experiments depending on the
tip) and for this reason the features related to the vibrational modes are
better seen in the second derivative of the current, d2 I/dV 2 , where they
appear as peaks (for positive bias), very much like in the IETS in planar
tunnel junctions. We show an example of the original data in Fig 16.3.
To confirm the interpretation of the origin of the peak in d2 I/dV 2 , the
authors used isotopic substitution, i.e. they replaced the hydrogen atoms
by deuterium ones in the molecules. In the case of the deuterated acetylene
(C2 D2 ), they showed that the peak in d2 I/dV 2 is shifted to 266 mV, which
corresponds to the expected change in energy of the C-H stretch mode.
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
Indeed, these values are in close agreement with those obtained by electron
energy loss spectroscopy (EELS). This experiment inspired an enormous
amount of work in which the chemical sensitivity of the single-molecule
IETS has been exploited. This has led in the last years to a better under-
standing and control of surface chemistry at the atomic level. The activities
of the first years on STM-IETS have been reviewed by Ho in Ref. [710].
The first STM-IETS experiments raised several fundamental questions
related, for instance, to the selection (or propensity) rules that apply in
this case. With respect to the tunneling process that gives rise to the peaks
seen in the spectra, it was believed that there is no fundamental difference
with respect to the traditional IETS in oxide tunnel junctions. In other
words, the process responsible for the vibrational signatures was believed
to be the phonon emission process described in Fig. 16.2. However, it is
worth stressing that the electron-phonon interaction in these systems does
not always lead to an increase of the conductance at the phonon energies.
Thus for instance, Ho and coworkers have reported in Ref. [711] STM-
IETS studies that revealed two vibrational modes showing a decrease in the
conductance at 682.0 and 638.3 mV for single oxygen molecules chemisorbed
on the fourfold hollow sites of an Ag(110) surface at 13 K. These results can
be seen in Fig. 16.4, where one can observe the presence of two well-defined
dips at positive bias. It is worth remarking that in this case the change in
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
Fig. 16.5 IETS spectrum of a C11 junction formed with gold cross wires. The dashed
line is a simple polynomial background and is presented as a guide to the eye. Mode
assignments are from comparison to previous experimental results. Reprinted with per-
mission from [712]. Copyright 2004 American Chemical Society.
in molecular transport junctions, but they were the first ones that explored the regime
discussed in this section, where the transport through the junctions takes place in a
non-resonant manner and the current probes the vibrational modes of the ground state
of the molecule.
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
Fig. 16.6 IET spectrum of a C8 dithiol SAM measured with the nanopore technique.
The spectrum was obtained from lock-in second-harmonic measurements with an ac
modulation of 8.7 mV (rms value) at a frequency of 503 Hz (T = 4.2 K). Reprinted with
permission from [713]. Copyright 2004 American Chemical Society.
Fig. 16.7 Single-molecule IETS measurements using STM break junctions. (a) Semilog
conductance histogram with a peak at G0 , and an additional peak at 6 × 10−3 G0 ,
which is attributed to the conductance of propanedithiol. (b) A conductance curve with
steps for a single molecule measurement. The four symbols represent four stretching
distances where the bias was swept and the I-V and first derivative were recorded. (c) The
corresponding four first derivative curves (offset for clarity), and (d) the corresponding
IET spectra obtained numerically. The curves are antisymmetric, and certain features
are very reproducible along the conductance plateau. Reprinted with permission from
[722]. Copyright 2008 American Chemical Society.
that the observed spectra were indeed valid IETS data by examining the
peak width as a function of temperature. This important test is usually
very difficult to carry out with other techniques.
IETS has become quite popular in the field of molecular electronics
over the last years and it has distinguished itself as a unique spectroscopic
probe of molecular junctions. From comparison between experiments and
computations, IETS can be useful for characterizing numerous aspects of
molecular junctions such as the confirmation of the presence of the molecule,
information on the nature of the interfaces, the orientation of the molecule
and even electronic pathways can be identified. For further experimental
examples of the use of the of IETS in the regime described in this section
see Refs. [574, 715–721].
The experiments that we have just described correspond to situations
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
d2 V/dI 2 (VA )
−2
Au
voltage (mV)
Fig. 16.8 An example of an electron-phonon spectrum measured for a gold point contact
by taking the second derivative of the voltage with respect to the current. The long-
dashed curve represents the phonon density of states obtained from inelastic neutron
scattering. Reprinted with permission from [728]. Copyright 1980 IOP Publishing Ltd.
bringing a needle of a metal gently into contact with a metal surface. With
this technique stable contacts are typically formed having resistances in
the range from ∼ 0.1 to ∼ 10 Ω, which corresponds to contact diameters
between d ≃ 10 and 100 nm. The elastic and inelastic mean free path
can be much longer than this length d, when working with clean metals
at low temperatures, and the ballistic nature of the transport in such con-
tacts has been demonstrated in many experiments. The main application of
the technique has been to study the electron-phonon interaction in metals.
Here, one makes use of the fact that the (small but finite) probability for
back-scattering through the contact is enhanced as soon as the electrons ac-
quire sufficient energy from the electric potential difference over the contact
that they are able to excite the main phonon modes of the material. The
differential resistance, dV /dI, of the contact is seen to increase at the char-
acteristic phonon energies of the material. Notice that this is at variance
with the typical signature in IETS. A spectrum of the energy-dependent
electron-phonon scattering can be directly obtained by measuring the sec-
ond derivative of the voltage with respect to the current, d2 V /dI 2 , as a
function of the applied bias voltage. An example is given in Fig. 16.8.
Peaks in the spectra are typically observed between 10 and 30 mV, and
are generally in excellent agreement with spectral information about the
phonons of the corresponding metal obtained from other experiments (e.g.
neutron scattering), and with calculated spectra.
Traditionally, electron-phonon spectroscopy in large metallic contacts
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
µ2
µ1
µ1
kF
µ2 kF
Fig. 16.9 Electron distribution function in the vicinity of the orifice. Here, kF is the
equilibrium Fermi wave vector, µ1 and µ2 are the chemical potentials for each side,
which, far from the orifice and in the presence of an applied potential V , are equal to
EF − eV /2 and EF + eV /2, respectively.
1
conductance (2e /h)
2
dG / dV (arb. units)
0 .9 9 T
L
0 .9 8 L
T
-4 0 -2 0 0 20 40
voltage (m V)
0 .9 7
-0.4 -0 .2 0 0 .2 0.4
voltage (V)
Fig. 16.10 Differential conductance as a function of the applied bias voltage for a one-
atom Au contact at 4.2 K. The contact was tuned to have a conductance very close
to 1 G0 , which suppresses the amplitude of the conductance fluctuations. This allows
the observation of a phonon signal, which is seen as a maximum at zero bias. Inset: By
taking the derivative of the conductance the transverse (T) and longitudinal (L) acoustic
branches can be recognized symmetrically positioned around zero. Note the expanded
scale of the voltage axis in the inset. Reprinted with permission from [731]. Copyright
2000 by the American Physical Society.
phonon interaction, and η is a function of the scattering angle that takes the
geometry into account, such that only backscattering through the contact is
effective, η(θ) = (1 − θ/ tan θ)/2. From this expression, and by considering
Fig. 16.9, one can see that the contribution of scattering events far away
from the contacts is suppressed by the effect of the geometric angle at
which the contact is seen from that point. The probability for an electron
to return to the contact decreases as (a/d)2 , with a the contact radius
and d the distance from the contact. This implies that the spectrum is
dominantly sensitive to scattering events within a volume of radius a around
the contact, thus the effective volume for inelastic scattering in the case of
a clean opening (the contact) between two electrodes is proportional to
a3 . Clearly, this effective volume must depend on the geometry of the
contact. For a long cylindrical constriction, the electrons scattered within
the constriction will have larger return probability, the effective volume, in
this case, increases linearly with the length [724].
The point-contact spectroscopy has been extended in recent years to
atomic-sized contacts. As the contact becomes smaller, the signal comes
from scattering on just a few atoms surrounding the contact. The spectrum
no longer measures the bulk phonons, but rather local vibrational modes of
the contact atoms. In attempting to measure the phonon signal for small
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
1.01
2
a 1.00 d
G (2e /h)
short wire long wire 0.99
2
conductance (2e /h)
L
2
0.98
M
0.97
1
3
long wire (L)
rupture rupture
S
2
0
0 5 0 5 10 15 20 25
displacement (Å) 1
G0 dG/dV (V )
-1
1.01
0.92
0.91
b 1.00 c 0
G0 dG/dV (V ) G (2e /h)
0.99
2
0.90
-1
0.98 -1
0.89 short wire (S) short wire (M)
0.97
-1
0.8 0.8
0.4 0.4 -2
0.0 0.0
-0.4 -0.4 -3
-1
-0.8 -0.8
-30 -20 -10 0 10 20 30 -30 -20 -10 0 10 20 30 -30 -20 -10 0 10 20 30
bias voltage (mV) bias voltage (mV) bias voltage (mV)
Fig. 16.11 Point contact spectroscopy of gold atomic chain. (a) Short and long atomic
wire, ∼ 0.4 and ∼ 2.2 nm, respectively, as given by the length of the conductance plateau.
Panels (b–d) show the differential conductance and its derivative at points S, M , and
L, respectively. The various curves in (b–d) were acquired at intervals of 0.03, 0.03 and
0.05 nm, respectively. Note that the vertical scales for the last thee panels are chosen
to be identical, which brings out the relative strength of the electron-phonon interaction
for the longer chains. The wire in (d) has a length of about 7 atoms. Reprinted with
permission from [732]. Copyright 2002 by the American Physical Society.
contact sizes one encounters the problem that the phonon signal intensity
decreases, according to Eq. (16.2), while the amplitude of the conductance
fluctuations5 remains roughly constant, or slightly increases. The result
is that the phonon signal is sometimes hidden in the conductance fluctua-
tions for the smallest contacts. A solution to this problem is obtained for
the special and interesting case of a contact made up of a single channel
with nearly perfect transmission probability, where these fluctuations are
suppressed. Under these conditions the features due to phonon scattering
become clearly visible. This is illustrated in Fig. 16.10 where we show an
example of the point contact spectrum of a gold one-atom contact [731].
Surprisingly, one observes a spectrum (see inset of the figure) that still
closely resembles the bulk phonon spectrum, although the relative intensi-
ties of the features in the spectrum are different.
5 These fluctuations were described in detail in section refsec-cond-fluct.
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
The point contact spectroscopy was pushed to study the phonon modes
in Au atomic chains by Agraı̈t and coworkers [732, 733]. As we described
in section 11.8, atomic chains of certain metals can be formed with the
STM and break junction techniques. These chains constitute in some sense
the simplest molecules that one can think of. Thus, the PCS of gold chains
of Refs. [732, 733] is of special interest for us. In these experiments, the
differential conductance was measured using a lock-in detection with a 1
mV modulation voltage, from which dG/dV was calculated numerically.
The energy resolution was limited by the temperature of 4.2 K to 2 meV.
The results for the differential conductance and its derivative for a long
atomic chain (∼ 7 atoms) are shown in Fig. 16.11(d). Notice that at ±15
mV bias the conductance exhibits a rather sharp drop by about 1%. In the
second derivative d2 I/dV 2 this produces a pronounced single peak, point-
symmetric about zero bias. The chains of Au atoms have the fortuitous
property of having a single nearly perfectly transmitted conductance mode,
which suppresses conductance fluctuations that would otherwise mask the
phonon signal. Some asymmetry that can still be seen in the conductance
curves is attributed to the residual elastic scattering and interference con-
tributions.
The fact that only one conductance drop is clearly seen was interpreted
by the authors as follows. By energy and momentum conservation the
signal can only arise from electrons that are back-scattered, changing their
momentum by 2kF . With ~ω2kF the energy for the corresponding phonon,
the derivative of the conductance is expected to show a single peak at
eV = ±~ω2kF . The transverse phonon mode cannot be excited in this one-
dimensional configuration and only the longitudinal mode is visible. We
shall see in the next chapter that this argument is, strictly speaking, only
valid for infinite chains, while it is approximate for the finite chains realized
in the experiments.
Another interesting feature of the point-contact spectra of gold atomic
chains is that the position of the peak in dG/dV shifts as a function of the
strain in the wire. As one can see in Fig. 16.11(d), the frequency of the
mode associated to the peak decreases as a function of the tension because of
the decreasing bond strength between the atoms. However, the amplitude
(peak height) increases, until an atomic rearrangement takes place, signaled
by a small jump in the conductance (not shown here). At such points the
amplitude and energy of the peak in dG/dV jump back to smaller and
larger values, respectively. This is consistent with the phonon behavior
of Au atomic chains found in ab initio calculations [360]. The growing
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
50
D
40 2
30
dI/dV (2e /h)
20
0.98
2
10
dI/dV (2e /h)
0.96
2
0.95
60
0.94
50 HD
0.96
40
0.93
(arb.unit)
Counts
0.1
30
0. 0
20
d I/dV (arb.unit)
0.2 -0.1
2
10
d I/dV
-0.2
-80 -60 -40 -20 0 20 40 60 80
2
Energy (meV)
0. 0 30 H
2
20
2
-0.2 10
0 20 40 60 80 100
Energy (meV) Energy (meV)
Fig. 16.12 Left panel: Differential conductance curve for D2 contacted by Pt leads.
The dI/dV curve (top) was recorded over 1 min, using a standard lock-in technique with
a voltage bias modulation of 1 meV at a frequency of 700 Hz. The lower curve shows the
numerically obtained derivative. The spectrum for H2 in the inset shows two phonon
energies, at 48 and 62 meV. Right panel: Distribution of vibrational energies observed
for H2 , HD, and D2 between Pt electrodes, with a bin size of 2 meV. The peaks in
the distribution for H2 are marked by arrows and their widths byp error margins. These
p
positions and widths were scaled by the expected isotope shifts, 2/3 for HD and 1/2
for D2 , from which the arrows and margins in the upper two panels have been obtained.
Reprinted with permission from [567]. Copyright 2005 by the American Physical Society.
the heavier isotopes D2 and HD. The positions of the peaks in the spec-
tra of d2 I/dV 2 vary within some range between measurements on different
junctions, which can be attributed to variations in the atomic geometry of
the leads to which the molecules bind. Fig. 16.12 (right panel) shows his-
tograms for the vibrational modes observed in a large number of spectra for
each of the three isotopes. Two pronounced peaks are observed in each of
the distributions, that scale approximately as the square root of the mass of
the molecules, as expected. The two modes can often be observed together,
as in the inset of the left panel of Fig. 16.12. For D2 an additional mode
appears near 90 meV. This mode cannot easily be observed for the other
two isotopes, since the lighter HD and H2 mass shifts the mode above 100
meV where the junctions become very unstable. For a given junction with
spectra as in Fig. 16.12 (left panel), it is often possible to stretch the contact
and follow the evolution of the vibrational modes. The frequencies for the
two lower modes were seen to increase with stretching, while the high mode
for D2 is seen to shift downwards. This unambiguously identifies the lower
two modes as transverse modes and the higher one as a longitudinal mode
for the molecule. This interpretation agrees well with DFT calculations for
a configuration of a Pt-H-H-Pt bridge in between Pt pyramidally shaped
leads [567, 571]. The fact that the vibrational modes observed for HD that
are intermediate between those for H2 and D2 confirms that the junction
is formed by a molecule, not an atom.
A drop in the conductance as a fingerprint of the presence of a molecule
in highly conductive junctions has also been reported, for instance, in
Ref. [385]. In this work, PCS was used to identify the presence of oxygen
intercalated in Au atomic chains. More recently, similar vibration-induced
steps down in the conductance have been also observed in various small
molecules directly bonded to Pt electrodes [474].
fact that the tunneling is typically off-resonant, and therefore the injection
energy ∆E is rather large. In the PCS case, however, this occurs because
the molecule is strongly coupled to the leads and thus Γ is very large. In
view of this similarity, one may wonder whether there is any fundamen-
tal difference between IETS and PCS in molecular junctions. As we shall
discuss in the next chapter, recent theoretical work has shown that IETS
and PCS are indeed two sides of the same coin and they can be described
in a unified manner. In other words, these two techniques are based on
the same underlying physics and they simply refer to two different limiting
cases depending on the junction transparency.
In recent years, different experiments on highly conductive single-
molecule junctions, but with conductances not to close to 1 G0 , have clearly
suggested the idea that there is a smooth crossover between IETS and PCS.
Thus for instance, experiments on Pt-H2 junctions [734], Ag atomic wires
decorated with oxygen [605] and Pt-benzene junctions [473] with conduc-
tances between 0.1 and 0.4 G0 have shown that the signature of vibrational
modes is a step up in the conductance at the vibrational energies, i.e. exactly
like in the standard IETS case. The experiment that has finally clarified
this issue was reported recently by Tal et al. [735] and we now proceed to
describe it in certain detail.
In Ref. [735] the authors presented PCS and shot noise measure-
ments across a single-molecule junction formed by Pt electrodes and H2 O
molecules. The Pt/H2 O molecular junctions were formed using a MCBJ
setup at about 5 K. The formation of a clean Pt contact was verified by
conductance histograms, which exhibited a single peak around 1.4 G0 , pro-
viding so a fingerprint of a clean Pt contact [127]. Water molecules were
then introduced to the junction through a heated capillary, while the Pt
junction was broken and formed repeatedly. Following the introduction of
water, the typical Pt peak in the conductance histogram was suppressed
and contributions from a wide conductance range were detected with mi-
nor peaks around 0.2, 0.6, and 1.0 G0 . The continuum in the conductance
counts implies a variety of stable junction configurations that the authors
exploited for spectroscopy measurements on junctions with different con-
ductance.
In Fig. 16.13 we reproduce the results for the differential conductance as
a function of the voltage across the Pt/H2 O junction at two different linear
conductance values: 1.02 ± 0.01 G0 (a) and 0.23 ± 0.01 G0 (b). Junctions
with different zero-bias conductance were formed by altering the distance
between the Pt contacts or by re-adjusting a new contact. The steps in
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
(a) (b)
1. 02 0.28 Step up
dI/dV [G0]
dI/dV [G0]
0.26
1.00
0.24
0.98
Step down
0.22
-80 -60 -40 -2 0 0 20 40 60 80 -80 -60 -40 -20 0 20 40 60 80
Bias Voltage [mV] Bias Voltage [mV]
Fig. 16.13 Differential conductance (dI/dV ) as a function of the bias voltage for two
different Pt-H2 O-Pt junctions with linear conductance of 1.02 ± 0.01 G0 (a) and 0.23
± 0.01 G0 (b). Reprinted with permission from [735]. Copyright 2008 by the American
Physical Society.
above, see Fig. 16.2. The difference is now that this process is much more
probable when the energy of the electron surpasses the energy of the elec-
tronic level by an amount that is equal to ~ω (corrected by a factor that
depends on how the voltage drops across the junction). The reason is that
an electron with that energy may lose an energy equal to ~ω (by emitting
a vibrational mode) and then it crosses the molecule exactly at resonance
with the molecular level. The enhanced probability of this inelastic process
gives rise to a peak in the differential conductance at a bias voltage of the
order of ~ω/e away from the Coulomb blockade peak. From this argument,
it is also easy to understand that in order to observe a pronounced current
step, the width of the electronic resonance, Γ, must be smaller than ~ω. In
any case, the bias voltage must be larger than ~ω for this inelastic process
to take place.
As it was also explained in section 15.7, the signature of a vibrational
mode can be also seen in the stability diagrams, see Fig. 15.17. In these
plots, the peaks in the conductance, which correspond to the step-like fea-
tures in Fig. 15.16, show up as lines. In particular, the vibration mode
with energy 5 meV appears there as running lines that intersect the main
diamonds or conductance gap regions. The energy of this excitation is too
small to correspond to an electronic excitation. Moreover, some of these
lines are observed for both charge states, which would be very unlikely for
an electronic excitation. Even more convincing is the fact that multiple
excitations with the same spacing are observed, see Fig. 15.17(d). This
corresponds most likely to the excitation of several vibrational quanta of
the same mode, i.e. multi-phonon processes. Let us also say to conclude
this discussion that signatures of intrinsic vibrational modes of the C60
molecules were also observed in the stability diagram of some devices, see
in particular Fig. 3 in Ref. [22].
The experiment just described was followed by other experiments with
weakly coupled molecules where signatures of the vibrational modes in
the transport characteristics were also observed. For instance, Zhitenev et
al. [738] reported transport measurements through a small self-assembled
monolayer of thiolated organic molecules in which the conductance exhib-
ited a series of equally spaced peaks, the position of which could be con-
trolled by a gate voltage. These peaks were attributed to the lowest molec-
ular vibrations of the molecules. The most surprising thing in this exper-
iment was the observation of a large number of conductance peaks with
slowly decreasing amplitudes. This would mean that phonon processes of
very high-order were taking place in these junctions. On the other hand,
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
Fig. 16.14 Stability diagrams (dI/dV vs. V and Vg ) for four C140 SMTs fabricated
with the electromigration technique. White arrows indicate excited levels at 11 and
22 meV. dI/dV is represented by a color scale from black (zero) to white (maximum),
with maximum values 200 nS (device I), 600 nS (II), 15 nS (III), and 100 nS (IV).
Measurements were done at 1.5 K for I-III and 100 mK for IV. Reprinted with permission
from [685]. Copyright 2005 American Chemical Society.
Fig. 16.15 Stability diagrams of a three-terminal junction with OPV-5, measured at 1.6
K. Plotted in (a) is dI/dV as measured with a lock-in technique (modulation amplitude
0.4 mV) and in (b) the numerically calculated second derivative, which serves to highlight
the fine structure of the excitations. The current levels are the same near both degeneracy
points, which is a strong indication that they belong to the same molecule. Three
different charge states are probed. The N + 1 state is not indicated; for low bias voltages
it starts at gate voltages larger than 2.2 V. The data yield an addition energy of 210
meV and a gate coupling of 0.05. Reproduced with permission from [678]. Copyright
Wiley-VCH Verlag GmbH & Co. KGaA.
Kondo resonance and vibrational sidebands [23, 662]. The first work in
which this coexistence was studied in detail was reported by Natelson’s
group [617]. Using electromigration-based SMT junctions they analyzed
the transport through a molecule comprising a single Co ion coordinated
by conjugated ligands. In many devices they observed the Kondo effect
and a Kondo temperature of ∼ 40 K was deduced from the temperature
dependence of the zero-bias conductance. Moreover, in some cases, the
conductance in the classically blockaded region and/or outside the Kondo
resonance was large enough to allow clean measurements of ∂ 2 I/∂VSD 2
. In
Fig. 16.16 we reproduce results from Ref. [617] where maps of this quan-
tity are shown as a function of VSD (source-drain voltage) and VG (gate
voltage) in two different devices at 5 K. The left panel shows mainly a
diamond-like region corresponding to a charge state exhibiting standard
Coulomb blockade, while the right one focuses on the next diamond where
the Kondo resonance is visible at zero bias. Two prominent features within
the blockaded (Kondo) regime are indicated with black arrows. Features in
∂ 2 I/∂VSD
2
of opposite sign are symmetrically located around zero source-
drain bias, consistent with inelastic tunneling expectations.
The ∂ 2 I/∂VSD2
features in the blockaded region occur at essentially con-
stant values of VSD until VG is varied such that the feature approaches the
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
hω
− hω hω eV − hω hω eV eV
2 2
2
d I/dV
2 d I/dV dI/dV hω
− hω hω
hω eV − hω eV eV
Fig. 16.17 Summary of the main vibrational signatures in the transport characteristics
of molecular junctions in various transport regimes.
Let us now briefly summarize the main vibrational signatures that we have
shown to appear in the different transport regimes (see Fig. 16.17):
The reader should bear in mind that this summary is slightly oversim-
plified and more complex signatures are also possible. Thus for instance,
Thijssen et al. [743] have reported the observation of anomalous spikes in
the differential conductance of a variety of junctions, which were attributed
to vibrationally induced two-level systems. On the other hand, vibronic
effects can also be responsible for other strong non-linearities in the I-V
characteristics (see section 8 of Ref. [695] for a detailed discussion of this
issue).
conduction channels. However, one can have situations in which several channels combine
to give a conductance close to G0 . In this case, the signature of the vibrational modes
can be a step down in the conductance, depending on the precise value of transmission
coefficients.
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
Chapter 17
Vibrationally-induced inelastic
current II: Theory
In this section we shall address the limit in which the traversal time is much
smaller than the time needed for an electron to feel the molecular vibrations.
In this case, the usual approach is to treat the electron-electron interaction
at a mean field level and to make a perturbative expansion in the electron-
phonon interaction. Our discussion of this regime will be divided into two
501
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
subsections. In the first one, we shall discuss in detail the results obtained
from the resonant tunneling model including in addition the coupling to a
single phonon mode. This model will help us to understand the origin of the
different vibrational signatures in this regime. Then, the next subsection
will be devoted to a description of the ab initio methods that have been
developed so far to elucidate the propensity rules in this regime and to
establish a quantitative comparison with experimental results.
H = He + ~ω b† b + 1/2 + λd† d b† + b .
¡ ¢ ¡ ¢
(17.1)
hω eV hω eV
L L
ε0 R ε0 R
kBT = 0.0
_
1 kBT = 0.01hω
G/G0 and Gel/G0
_
kBT = 0.025hω
1 _
0.995 kBT = 0.05hω
_
G/G0
kBT = 0.2hω
Gel
0.99 G 0.99
(a) 0.985
(c)
0.98
0 0.5 1_ 1.5 2 0.5 0.75 1_ 1.25 1.5
eV/hω eV/hω
δGel/G0 and δGinel/G0
0
2
(1/eG0) d I/dV
0.01
δGinel -0.05
0
2
-0.1
-0.01
δGel
-0.02 (b) -0.15 (d)
0 0.5 1_ 1.5 2 0.5 0.75 1_ 1.25 1.5
eV/hω eV/hω
Fig. 17.2 Results of the single-phonon (or Holstein) model for a highly transmissive
contact: ǫ0 − EF = 0, λ = 2~ω and ΓL = ΓR = 10~ω. (a) Zero-temperature total
conductance (G = dI/dV ) and elastic conductance (Gel = dIel 0 /dV ) as a function of the
bias voltage. (b) Elastic (δGel = dδIel /dV ) and inelastic (δGinel = dIinel /dV ) conduc-
tance corrections vs. voltage for the parameters of (a). (c) Temperature dependence of
the total conductance. (d) The corresponding d2 I/dV 2 vs. voltage for the temperatures
considered in (c).
exhibits an abrupt step down (of about 1%) at eV = ~ω. This result re-
produces the typical signature observed in the gold atomic chains or in the
Pt-H2 junctions discussed in section 16.3. As one can see in Fig. 17.2(b),
the step down in the conductance is due to the dominant negative con-
tribution coming from the elastic correction (δGel = dδIel /dV ). In other
words, the elastic correction gives rise in this limit to a finite backscatter-
ing that reduces the conductance of the junction [745]. After all, this is
natural because the (elastic) transmission is already close to one and thus,
an incoming electron can only be backscattered.
In Fig. 17.2(c) and (d) we show for this high transmission case the tem-
perature dependence of the differential conductance and the corresponding
d2 I/dV 2 , respectively. First, notice that the signature of the inelastic cur-
rent in d2 I/dV 2 is a dip and second, notice also that for temperatures of
the order of 0.2~ω/kB the signature is no longer visible.
We now consider a low-transmissive situation by simply shifting the level
away from the Fermi energy (ǫ0 − EF = 80~ω), but keeping the values of
the scattering rates of the previous example unchanged. Thus, the (elastic)
zero-bias conductance is equal to 0.059 G0 . In Fig. 17.3(a) we show the
contributions δGel and δGinel versus the voltage, as well as the sum of the
two (δG). In this case, we have assumed that the temperature is kB T =
0.05~ω. Notice that there several basic differences with respect to the
previous example. First, the change in the conductance at eV = ~ω is
dominated this time by the phonon emission process giving rise to a step
up. Second, the contribution δGel is positive for every voltage, but it
decreases slightly at the phonon energy. Third, the emission term has no
abrupt onset because of the finite temperature.
As one can see in Fig. 17.3(b), the signature of the vibrational mode is
barely visible in the differential conductance and one has to resort to its
derivative to see it clearly, see Fig. 17.3(c). Of course, the order of mag-
nitude of the inelastic current depends primarily on the electron-phonon
coupling constant, λ, which we have chosen small in comparison with the
scattering rates to ensure the validity of the perturbative approach. On the
other hand, notice that d2 I/dV 2 exhibits a linear background, typically
seen in the experiments, which is due to the contribution of the elastic
current, which contains a tiny cubic term (∝ V 3 ).
The model also describes the crossover between the two situations just
described, as we illustrate in Fig. 17.4. In this example, we have kept
constant the values of the scattering ΓL = ΓR = 10~ω (symmetric junction)
and changed the level position. As one can see in this figure, the vibrational
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
0.062
0.0014
0
0 0.5 1_ 1.5 2 0 0.5 1_ 1.5 2
eV/hω eV/hω
Fig. 17.3 Results of the single-phonon model for a low transmissive contact: ǫ0 −
EF = 80~ω, λ = 2~ω, ΓL = ΓR = 10~ω and kB T = 0.05~ω. (a) Elastic (δGel ),
inelastic (δGinel ) and total (δG = δGel +δGinel ) conductance corrections versus voltage.
(b) Corresponding total conductance and elastic conductance versus voltage. (c) The
corresponding d2 I/dV 2 .
0
(a) ε0 = 0.0 0.03 _
-0.02 (b) ε0 = 10hω
0.02
Gel(V=0) = G0 Gel(V=0) = 0.8G0
-0.04 0.01
0 0.5 1_ 1.5 2 0 0.5 1_ 1.5 2
eV/hω eV/hω
2
(1/eG0) d I/dV
0.028 _
(c) ε0 = 20hω
2
0.027
0.026
Gel(V=0) = 0.5G0
0.025
0 0.5 1_ 1.5 2
eV/hω
_
(d) ε0 = 40hω
_
(e) ε0 = 80hω
0.009
0.0015
0.0085 Gel(V=0) = 0.2G0
0.0014 Gel(V=0) = 0.059G0
0.008
0 0.5 1_ 1.5 2 0 0.5 1_ 1.5 2
eV/hω eV/hω
Fig. 17.4 Results of the single-phonon model for the crossover between PCS and IETS:
λ = 2~ω, ΓL = ΓR = 10~ω and kB T = 0.05~ω. Second derivative of the current as a
function of the voltage for different values of the level position (measured with respect
to EF ) as indicated in the different panels. We also indicate in the panels the value of
the zero-bias elastic conductance.
τ = 1/2. Moreover, one can show that while the phonon emission term
gives a contribution equal to +(λ2 /Γ2 )τ 2 /4 to the conductance jump, the
elastic correction gives a negative contribution equal to −(λ2 /Γ2 )τ 3 /2. No-
tice that this result suggests that for very low transparencies, and if one
is only interested in the signature at the phonon energy, the contribution
of the elastic correction can be ignored, which is usually done in the IETS
context.
With respect to the temperature dependence of the phonon signature,
Paulsson et al. [736] have shown that for a symmetric contact, and ignoring
the energy dependence of the elastic transmission, the full width at half
maximum (FWHM) of the peak in d2 I/dV 2 is approximately 5.4kB T , i.e.
like in the standard IETS case [705].
Let us address now the typical situation realized in the STM contacts,
where there is a large asymmetry in the couplings between the molecule
and the surface and the molecule and the STM tip. In Fig. 17.5 we show
IET spectra for a junction in which ΓL = 10~ω and ΓR = 0.01ΓL . In
this figure the level position has been varied from a resonant case in panel
(a) to an off-resonant situation in panels (c) and (d). As one can see in
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
0
(a) 0.0022 (b)
2
-0.002
(1/eG0) d I/dV
ε0 = 0.0 ε0 = 10hω
_
-0.004
0.002
2
-0.006
(c) (d)
0.0008 0.00013
_
_
ε0 = 20hω 0.00012 ε0 = 40hω
0.0007
0.00011
0 0.5 1_ 1.5 2 0 0.5 1_ 1.5 2
eV/hω eV/hω
Fig. 17.5 Results of the single-phonon model for a very asymmetric contact simulating
the situation typically realized in STM experiments: λ = 2~ω, ΓL = 10~ω, ΓR = 0.01ΓL
and kB T = 0.05~ω. Second derivative of the current as a function of the voltage for
different values of the level position (measured wit respect to EF ) as indicated in the
different panels.
Fig. 17.6 Phase diagram for the single-phonon level model discussed in this section
(inset) illustrating the sign of the conductance change at the onset of phonon emission.
At a given asymmetry factor α the elastic transmission τ has an upper bound τmax (solid
line), and the inelastic conductance change undergoes a sign change at τcrossover =
τmax /2 (dashed line). Reprinted with permission from [751]. Copyright 2008 by the
American Physical Society.
we will follow here Paulsson et al. [736] and describe this effect at a phe-
nomenological level. The simplest way to include non-equilibrium heating
is to write down a rate equation for the phonon occupation, n, including
an external damping rate γd of the phonons [736]
P
ṅ = + γd [nB (~ω) − n] , (17.3)
~ω
where P is the power dissipated into the phonon mode and nB is the
Bose function. The external damping can be due to either the interac-
tion with the phonons of the electrodes or the electron-phonon interaction
in the molecule. From this equation, the steady state occupation n is easily
found. To complete the calculation we need now an expression for both the
power and the current in terms of the nonequilibrium phonon occupation.
Assuming that the transmission is energy-independent and considering a
symmetric junction (ΓL = ΓR = Γ), Paulsson et al. [736] showed that these
quantities can be expressed within the LOE approximation as follows
γeh π~
P LOE = γeh ~ω [nB (~ω) − n] + P, (17.4)
4 ~ω
2e2 1 − 2τ π~ Sym
I LOE = τ V + eγeh I , (17.5)
h 4 e~ω
where γeh = (ω/π) λ2 τ 2 /Γ2 is the electron-hole damping rate.4 Here, P and
I Sym are universal functions of the voltage, phonon frequency, temperature
and phonon occupation given by
h ³ ´ i ³ ´ ³ ´
eV ~ω eV
~ω cosh kB T − 1 coth 2kB T ~ω − eV sinh kB T
P= ³ ´ ³ ´ (17.6)
π~ ~ω
cosh kB T − cosh kB T eV
à !
Sym 2e ~ω − eV ~ω + eV
I = 2eV n + ~ω−eV − ~ω+eV . (17.7)
h e kB T − 1 e kB T − 1
Eq. (17.5) reproduces the zero-temperature transmission dependence
of the conductance jump discussed above. However, there is a small dis-
crepancy between these two results, namely the zero-temperature inelastic
conductance in Eq. (17.5) vanishes for eV < ~ω, while this is not the case
in the results presented above. The origin of this little difference is unclear
to us.
Eqs. (17.3)-(17.5) were used by Paulsson et al. in Ref. [736] to fit the
experimental results of Pt-H2 junctions [127, 567]. As one can see in
4 There is difference of a factor 4 in the expression of γeh with respect to Ref. [736]
because of the different definition of the scattering rates.
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
60 40 20 0 20 40 60
1 b) Au
dI/dV (G0)
Exp.
γd = 0
γd = 3 γeh
0.98
γd = 10 γeh
-20 -10 0 10 20
Bias (mV)
Fig. 17.7 Single level model [Eqs. (17.4) and (17.5)] fitted to the experimentally mea-
sured conductance through a deuterium molecule [567]. The parameters used for the
fit are ~ω = 50 meV, τ = 0.9825, γeh = 1.1 × 1012 s−1 , and T = 17 K. (b) A simple
model (see Ref. [736] for details) fitted to the measured conductance through an atomic
gold wire (experimental data from Ref. [732]). The fit yields the following parameters:
~ω = 13.8 meV, T = 10 K, γeh = 12 × 1010 s−1 , and γd = 3γeh . Reprinted with
permission from [736]. Copyright 2005 by the American Physical Society.
d†i λα
XX
†
He−vib = ij dj (bα + bα ). (17.8)
ij α
Here ωα are the vibrational frequencies, Hij = hi|H|ji are the matrix ele-
ments of the single-particle electronic Hamiltonian H in the atomic-orbital
5 This list is by no means complete, but it should be easy to trace back from it the
ship. The reader not interested in this theoretical discussion can jump directly to the
description of the propensity rules in section 17.1.2.4.
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
kµ
where Mij = hi|∇kµ H|Q=0
~ |ji. From Eq. (17.10) one can see that the
calculation of the coupling constants requires to compute derivatives of the
Hamiltonian matrix elements with respect to the atomic position. Indeed,
since the employed basis sets are usually nonorthogonal, things are slightly
more complicated and the coupling constants are often calculated using the
ideas of Head-Gordon and Tully [780], see e.g. Ref. [775].
∞
4ie
Z
dE Tr Ga ΓL Gr (fL − 1)Σ+− −+
© £ ¤ª
Iinel = e−vib − fL Σe−vib ,
h −∞
where sharp resonances are present, but it turns out to be quite reasonable
in many situations of interest. With this assumption, the retarded and
advanced Green’s functions, as well as the scattering rates, can be evaluated
at the Fermi energy and some of the integrals can be done analytically,
which simplifies enormously the calculations. The detailed formulas for the
current within this approximation can be found in Refs. [736, 202].
Another important issue is the expression of the phonon occupation
that enters in the current formula via the electron-vibration self-energies.
The simplest approximation, which is fully consistent with the LOE, is to
assume that the phonons are in thermal equilibrium at the bath tempera-
ture. Heating effects, due to the nonequilibrium established at finite bias,
can be described in a various ways. For instance, as we explained for the
single-phonon model, the authors of Refs. [763, 736, 775] determine the
phonon occupation in a self-consistent manner by imposing that the power
transferred by electrons from the leads into to the device is balanced by the
power transferred from the device electrons to the phonons. Another phe-
nomenological way of introducing the nonequilibrium effects is discussed in
Ref. [777].
them to compute the line shapes of the IET spectra, it provides a conve-
nient way to determine the intensities of the peaks in off-resonant situations.
More importantly, this appealing formulation has allowed the authors to
get a deeper insight into the propensity rules. With this approach, Troisi
and Ratner have again emphasized the importance of modes with large
component in the tunneling direction. Thus for instance, they have shown
that for a linear chain with one orbital per atom, only totally symmetric
modes contribute to IETS signal. For molecules with side chains any nor-
mal mode dominated by side chain motion will contribute only weakly to
IETS. The authors have also employed group theory to identify the main
normal modes for planar conjugated molecules with C2h symmetry.
Gagliardi et al. [776] have also presented a detailed study of the propen-
sity rules in the case of low-transmissive junctions, extending the work of
Troisi and Ratner. The approach of this work is based on the idea that
both the elastic and inelastic current can be expressed as the sum of a
small number of essentially noninteracting paths or conduction channels
through the device.
More recently, Paulsson et al. [751] have reported a method to determine
the propensity rules in junctions with arbitrary transparency (within the
weak electron-phonon coupling regime). Similar to Ref. [751], the key idea
in this work is to analyze the inelastic transport in terms of just a few
selected electronic scattering states, namely those belonging to the most
transmitting channels at the Fermi energy. These scattering states typically
have the largest amplitude inside the junction and thus account for the
majority of the electron-phonon scattering.
Fig. 17.8 IET spectra of an anthracene thiol junction. The upper curve corresponds to
the experimental spectra and the lower curve to the computed one. Labels refer to the
normal modes of the molecule computed in the absence of metal and numbered from the
lowest energy vibration. Reprinted with permission from [718].
Fig. 17.9 Calculated (black lines) and experimental (blue lines) IETS. (a) OPE molecule
with Au(111) leads, (b) Au chain connected to Au(100) leads, (c) O2 molecule on
Ag(110), and (d) CO molecule on Cu(111). In case (c) the Fermi energy has been
shifted manually to match the experiment (dashed red line). The experimental data
originates from Refs. [712, 732, 711, 781]. For the STM configurations (c) and (d), the
calculated IETS is compared with a rescaled d2 I/dV 2 . Reprinted with permission from
[751]. Copyright 2008 by the American Physical Society.
regime. This agreement of the calculated and measured IET spectra makes
this spectroscopy, in combination with theory, a very useful diagnostic tool.
-7
( 10 )
-8
( 10 )
5
dI/d (A/V)
2
dI/d (A/V)
1
3
Fig. 17.10 Differential conductance versus source-drain voltage calculated with the
EOM method applied to the single-phonon model of Eq. (17.1). The parameters have
the following values: ΓL = ΓR = 0.02 eV, T = 10 K, ǫ0 = 2 eV, ~ω = 0.2 eV, and
λ = 0.01 eV. The solid line corresponds to the self-consistent result and the dashed line
to the zero-order result (see Ref. [789] for details). The inset shows a blow-up of the
phonon absorption peak that appears on the left of the main resonance. Reprinted with
permission from [789]. Copyright 2006 by the American Physical Society.
hω eV ε0 hω
eV
L L
ε0
R R
appearance of satellite peaks on the right hand side of the elastic peak.
Notice that these peaks are separated by a distance ∼ 2~ω/e. The factor 2
is again due to the choice of the voltage profile.
As we explained earlier, the fact that vibrations are in this case mani-
fested as peaks in the conductance, rather than peaks in d2 I/dV 2 , is simply
due to the fact that we are now dealing with a resonant situation. As shown
schematically in Fig. 17.11(a), the probability of the phonon emission tun-
neling process is greatly enhanced when the energy of an incoming electron
is such that by emitting a phonon it loses exactly the energy necessary
to cross the molecular level on resonance. This implies the appearance of
a peak in the differential conductance when the bias exceeds the voltage
necessary to see the resonant level in a quantity equal to ~ω/e times a cor-
rection factor that accounts for the shift of the level due to the voltage (this
factor equals 2 in Fig. 17.10). This argument applies for a single-phonon
process. If the electron-phonon coupling is large enough, emission of sev-
eral vibrational quanta becomes possible and it results in the appearance of
additional peaks in the conductance separated in this example by a voltage
equal to ∼ 2~ω/e, see Fig. 17.10. It is important to emphasize that in order
to resolve such satellite vibronic peaks, both the electronic coupling (Γ) and
the thermal energy (kB T ) must be smaller than ~ω, as in the example of
Fig. 17.10. Of course, at very low temperatures the voltage must be larger
than ~ω/e for the emission process to happen at all.
Another remarkable feature in Fig. 17.10 is the appearance of an ad-
ditional peak at Φ ∼ 3.25 V (see inset). As we show schematically in
Fig. 17.11(b), the resonant absorption of phonon could lead in this case
to the appearance of a peak on the left side of the elastic one. However,
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
already described at the end of the 1980’s [794, 795]. In the context of
molecular transistors, Boese and Schoeller [796] were the first to analyze
vibronic effects in the Coulomb blockade regime. Motivated by the ex-
periments on C60 SMTs by Park et al. [22], these authors generalized the
many-body master equations described in section 15.5.1 to include vibronic
effects. Since then, many authors have used rate equations to study differ-
ent aspects of this problem, see e.g. Refs. [747, 797–807]. In what follows,
we shall first describe how this transport problem is formulated in terms of
rate equations and then, we shall briefly discuss some of the main physical
effects that have been predicted to occur in this regime.
The different formulations of rate (or master) equations differ only in
minor details and we have chosen to follow Ref. [747]. The starting point
is the single-phonon model (Holstein model) that we have extensively dis-
cussed in previous sections, but now with the inclusion of the electron-
electron interaction in the molecule. In this model, often referred to as
Anderson-Holstein model, the transport through the molecule is assumed
to be dominated by a single level of degeneracy dg with energy ε in the pres-
ence of one vibrational mode with frequency ω0 . This system is described
by the Hamiltonian H = Hmol + Hleads + Ht , where9
U
Hmol = εnd + nd (nd − 1) + λ~ω0 (b† + b)nd + ~ω0 (b† b + 1/2),
2
X X
Hleads = ǫp c†apσ capσ ,
a=L,R p,σ
X X¡
ta c†apσ diσ + h.c. .
¢
Ht = (17.13)
a=L,R; i=1,dg p, σ
Here, Hmol describes the molecular degrees of freedom, Hleads the leads and
Ht the tunneling between the leads and the molecule. The Coulomb block-
ade is taken into account via the charging energy U . We focus on the regime
of strong Coulomb blockade, U → ∞, appropriate when eV, kB T ≪ U . The
operator diσ (d†iσ ) annihilates (creates) an electron with spin projection σ
on degenerate level i of the molecule and nd = i=1,dg ; σ d†iσ diσ denotes
P
where the transformed phonon operator b̃ = b − λ i,σ d†iσ diσ , so that the
P
of H′mol with occupancy n − 1). This rate is equal to the transition rate
involving hopping of an electron from the lead a to the dot by changing the
phonon occupancy from q (measured relative to the ground state of H′mol
with occupancy n − 1) to q ′ (measured relative to the ground state of H′mol
with occupancy n). More explicitly
2
Γaq′ ,q = Γa |hq ′ |X|qi| . (17.18)
The matrix elements hq ′ |X|qi are known as the Franck-Condon matrix ele-
ments because they also govern the transitions between different vibrational
states in molecular physics. They can be computed by standard methods
[174] and their absolute value |hq|X|q ′ i|2 ≡ Xqq2
′ , which are symmetric un-
′
der interchange of q and q , are given by (see Exercise 17.1)
¯X (−λ2 )k (q!q ′ !)1/2 λ|q−q′ | e−λ2 /2 ¯2
¯ q ¯
2
Xq<q′ = ¯ ¯ . (17.19)
¯ ¯
¯ (k)!(q − k)!(k + |q ′ − q|)! ¯
k=0
0.1
0.05
Equilibrated Phonons
I Unequilibrated phonons
0
−0.05
−0.1
−5 −4 −3 −2 −1 0 1 2 3 4 5
0.03
0.02
Equilibrated Phonons
0.01 Unequilibrated Phonons
I 0
−0.01
−0.02
−0.03
−5 −4 −3 −2 −1 0 1 2 3 4 5
Vsd
Fig. 17.12 Current (I) vs source-drain voltage Vsd for coupling constant λ = 1.0, ~ω0 =
1 and kB T = 0.05. Upper panel is for Vg = 0.0, while lower panel is for Vg = Vsd /2, µR =
0. I is in units of ekB T /~. Reprinted with permission from [747]. Copyright 2004 by
the American Physical Society.
This equation also gives a simple rule of thumb to estimate how many
phonon sidebands are expected for a given coupling constant λ. In particu-
lar, multiple steps arise only if λ is of the order of 1 or larger. Let us mention
that Sapmaz et al. [793] reported I-V characteristics of suspended single-
wall carbon nanotube quantum dots exhibiting a series of steps equally
spaced in voltage. These features were attributed to the excitation of the
stretching mode of the nanotubes. By comparing the I-V curves with the
model above for equilibrated phonons, a reasonable agreement was found
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
Fig. 17.13 (a) I-V characteristics for intermediate (λ = 1) and strong (λ = 4) electron-
phonon coupling for ε′ = 0 and kB T = 0.05~ω0 for equilibrated and unequilibrated
phonons. The strong electron-phonon coupling leads to a significant current suppres-
sion at low bias voltages. This Franck-Condon blockade arises from the behavior of the
Franck-Condon rates for phonon transitions q1 → q2 plotted in (b). The rates Γq1 q2 ,
shown for λ = 1 (left) and λ = 4 (right), are given in units of the ordinary electronic
coupling Γ (here a symmetric junction is considered). For strong electron-phonon cou-
pling, transitions between low-lying phonon states are exponentially suppressed. The
corresponding current suppression cannot be lifted by a gate voltage, which may serve
as a fingerprint of FC blockade. This is depicted in the plot of dI/dV in the V –Vg plane
for unequilibrated phonons with λ = 4 (c). The case of intermediate coupling with λ = 1
(d) is shown for comparison. Reprinted with permission from [801]. Copyright 2005 by
the American Physical Society.
1 0.04
G/ (ΓR / ΓL ) 2 e 2 / h
eV / ω 0
0 0.035
−1 0.03 c
a b c
−2
b
0.025 a
0.8 1 1.2 −2 −1 0 1 2
(Vg Cg /e) / (1−λ2ω 0 / Ec ) eV /ω0
Fig. 17.14 Left panel: ∂ 2 I/∂V 2 vs. bias and gate voltage, for λ2 = 3, N (0)|tL |2 =
0.1~ω0 , U = 16~ω0 , and kB T = 0.01~ω0 . The junction is considered to be very asym-
metric (ΓL ≫ ΓR ). Black/white indicates large negative/positive values. Right panel:
Conductance vs. bias voltage for three values of Vg corresponding to the vertical black
lines (a,b,c) in the upper panel. The lower curve (a) corresponds to the symmetric point
N = 1 − λ2 ~ω0 /EC . Reprinted with permission from [818]. Copyright 1999 by the
American Physical Society.
17.5 Exercises
17.2 Phonon sidebands in the Coulomb blockade regime: Solve the rate
equations both for equilibrated and unequilibrated phonons to reproduce the
results of Fig. 17.12. Using the parameters of the upper panel of Fig. 17.12,
compute the corresponding stability diagram.
17.3 Franck-Condon blockade: Use the rate-equation formalism described
in section 17.3.1 to study the Franck-Condon blockade in the regime of strong
electron-phonon coupling. In particular, solve the master equations for both
for equilibrated and unequilibrated phonons to compute the different transport
characteristics and reproduce the results of Fig. 17.13.
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
Chapter 18
537
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
1 2 N−1 N
∆E
L R
Fig. 18.1 Schematic representation of the model discussed in text to describe the inco-
herent tunneling through a molecular junction. Here, N sites with the same energy are
connected via nearest-neighbor transfer rates kj,j±1 . The continua on the left and right
correspond to the metallic states in the electrodes and ∆E is the activation energy.
The question that we want to address in this section is: How can we identify
the occurrence of the hopping regime in an experiment? As we saw in
Chapter 13, the coherent transport in off-resonant situations is manifested
in the linear conductance as an exponential dependence on the length of
the molecule and as an independence on the temperature. The hopping
regime is however characterized by the following two main signatures:
• The conductance decays linearly with the length of the molecular wire.
• The conductance depends exponentially on the temperature as
exp(−∆E/kB T ), where ∆E is an activation energy that depends on
the system under study.
1 We assume that only nearest-neighbor sites are directly connected, i.e. the only non-
zero rates are kj,j±1 .
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
kB T /(e2 kL,R ), and the N − 1 connections between the bridge sites, with a resistance
kB T /(e2 k) each.
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
(a) (b)
Fig. 18.2 (a) A set of I-V curves measured at different temperatures in Au-molecule-Au
junctions fabricated with the electromigration technique. The molecule is shown in the
upper inset, where a junction is schematically represented. (b) Arrhenius plots of Ln
current (amperes) versus inverse T (K−1 ) at different bias voltages showing a transition
in conductance from T -independent tunneling behavior at low T to a thermally activated
process at high T . The bias increment between curves is 0.1 V, and the bias of the lowest
curve is 0.1 V. The transition temperatures between coherent and incoherent behavior
are marked by the intersection between lines; see, for example, the arrow for 0.3 V.
Reprinted with permission from [841]. Copyright 2004 American Chemical Society.
this molecule. This suggests that the rate-limiting process in the hopping
mechanism is not thermal population of electrons/holes from the electrodes
into the first hopping site, but rather an intramolecular hopping process
(along the molecule).
Although the evidence presented in Ref. [841] is rather convincing, one
cannot completely exclude an interpretation of the data of Fig. 18.2 in terms
of coherent tunneling. As we explained in section 13.2, coherent tunneling
can also lead to a pronounced temperature dependence of the I-V charac-
teristics (see discussion below). This has been illustrated by Poot et al.
[471], who reported data similar to those of Fig. 18.2(b) in a three-terminal
device fabricated with the electromigration technique. In particular, these
authors investigated the gate and temperature dependence of the current
in molecular junctions containing sulfur end-functionalized tercyclohexyli-
denes. In Fig. 18.3 we reproduce some of the results of Ref. [471] in which
one can see the current as a function of temperature for four different bias
voltages at two gate voltages on a semilog scale. Notice that at low bias
the curves of Fig. 18.3 show thermally activated transport at high tem-
perature and temperature-independent transport at low temperature, i.e.,
very much like in Fig. 18.2(b). The crossover temperature is about 150
K in Fig. 18.3(a) and it decreases slightly as the bias is increased. The
slope of the exponential increase above this crossover temperature yields
and activation energy of 120 meV at low bias and this value decreases with
increasing bias.
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
The authors of Ref. [471] use the simple resonant tunneling model de-
scribed in detail in section 13.2 to analyze their data. Let us recall that in
this model the temperature dependence comes from the Fermi distribution
function in the leads and that the current becomes temperature-dependent
when kB T is not too small in comparison with the injection energy (or level
position measured with respect to the Fermi energy), which is the case in
this experiment and in the previous one described above. As one can see
in Fig. 18.3, the authors were able to fit the experimental data with this
model using as adjustable parameters the level position, ǫ0 , and the scatter-
ing rates ΓL and ΓR . It is important to remark that the two gate voltages
in Fig. 18.3 were far away from the degeneracy points, i.e. the transport is
not completely at resonance. In the fits, the values found for ǫ0 were very
similar to the activation energy mentioned above of 120 meV, which shows
the consistency of the fits. The total broadening of the level, Γ = ΓL + ΓR ,
was found to be in the range from 0.1 to 5 meV and it increased with
increasing bias voltage.5 In addition, the ratio Γ/ǫ0 was found to range
between 10−3 and 10−2 .6
The previous discussion shows that an unambiguous identification of
the hopping regime requires additional information beyond the tempera-
ture dependence of the current. As we explained in the previous section,
another key signature of the hopping regime is the linear decay of the con-
ductance/current with the length of the molecule. To our knowledge, this
signature, which is well-known in the context of electron transfer (see e.g.
Ref. [842]), has not yet been reported in single-molecule junctions. How-
ever, Choi et al. [843] have reported recently the transition from coherent
to hopping regime as a function of the molecular length in junctions based
on monolayers of conjugated oligophenyleneimine (OPI) molecules ranging
in length from 1.5 to 7.3 nm. The OPI wires were grown on a gold substrate
and contacted by a metal-coated AFM as a second electrode. In Fig. 18.4(a)
we reproduce the results of this experiment concerning the resistance (R)
versus molecular length (L) for a series of OPI molecules with different
numbers of phenyl units (n). As one can see, there is a clear transition of
the length dependence near 4 nm (OPI 5). In short wires, the linear fit in
Fig. 18.4(a) indicates that the data are well described with the standard
formula of coherent non-resonant tunneling: R = R0 exp(βL). The β value
5 Here, Γ corresponds to the full width of the resonance at half maximum, while in
section 13.2 it represents the half width at half maximum.
6 The temperature dependence of the current within the resonant tunneling model is
Fig. 18.4 Measurements of molecular wire resistance with a conducting probe AFM. A
gold coated tip was brought into contact with an OPI monolayer on a gold substrate.
(a) Semilog plot of R versus L for the gold/wire/gold junctions. Each data point is the
average differential resistance obtained from 10 I-V traces in the range -0.3 to +0.3 V.
Straight lines are linear fits to the data according to R = R0 exp(βL). The inset shows
a linear plot of R versus L, demonstrating linear scaling of resistance with length for the
long OPI wires. (b) Arrhenius plot for OPI 4, OPI 6, and OPI 10. Each data point is
the average differential resistance obtained at six different locations on samples in the
range -0.2 to +0.2 V. Straight lines are linear fits to the data. From [843]. Reprinted
with permission from AAAS.
was found to be 0.3 Å−1 , which is within the range of β values of typical
conjugated molecules.
For long OPI wires, there is a much flatter resistance versus molecular
length relation (β ∼ 0.09 Å−1 ). The extremely small β suggests that the
principal transport mechanism is hopping. As one can see in the inset of
Fig. 18.4(a), a plot of R versus L for long wires is linear, which is consis-
tent with hopping. The change in transport mechanism was also verified
by the temperature dependence. Fig. 18.4(b) shows that the resistance
for OPI 4 is independent of temperature from 246 to 333 K, as expected
for non-resonant coherent tunneling. However, both OPI 6 and OPI 10
display the strongly thermally activated transport that is characteristic of
hopping. The activation energies determined from the slopes of the data
are identical at 0.28 eV for both OPI 6 and OPI 10. Concerning the ques-
tions on the nature of the hopping sites and the origin of this activation
energy, the authors suggested that three-repeat conjugated subunits are
the charge-hopping sites in the long wires and that the hopping activation
energy corresponds to the barrier for rotation of the aromatic rings, which
transiently couples the conjugated subunits.
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
2 nm
major groove minor groove
Fig. 18.5 Double helical structure of DNA in B-conformation. Taken from Wikipedia
Commons.
the B-conformation the long axis of neighboring base pairs are twisted with
respect to each other by an average angle of 36o such that 10 base pairs
make a full turn. This conformation is stabilized by water molecules and
other counter ions. In dry conditions (less than 5 H2 O molecules per base
pair) the stable conformation is the so-called A-conformation that differs
from the B-conformation by an inclination of the base pairs with respect
to the molecule’s axis and a somewhat weaker twist [850].
With respect to the conduction mechanism in DNA, it is generally ac-
cepted that the electron transfer in DNA takes place via the overlap between
the π-orbitals of neighboring base pairs. This is similar to what happens
in certain stacked aromatic crystals, like the Beechgard salts, which are in-
deed metallic. This electron transfer mechanism in DNA suggests that the
base sequence can be very important since the π-system of the individual
bases may be different. Moreover, the conformation is important because
it determines the overlap between the base pairs. Finally, in order to have
a measurable electrical current in a DNA junction, it is crucial to make
sure that the π-system hybridizes strongly with the metallic states of the
electrodes.
Now, we turn to the discussion of the transport experiments in DNA-
based junctions. Let us start by summarizing the main findings. Most of
the transport measurements on single DNA molecules reported so far can be
divided into three classes. First, there are experiments showing that DNA
is an insulator for lengths larger than 40 nm at room temperature, with
essentially no discernible conductance up to 10 V. This suggests that the
electronic states of DNA are completely localized [851, 852]. Second, some
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
(a) (d)
(e)
(b)
(f)
(c)
Fig. 18.6 The left panels show conductance histograms of three DNA duplexes mea-
sured with the STM break-junction technique: (a) 5′ -CGCGCGCG-3′ -thiol linker, (b)
5′ -CGCGATCGCG-3′ -thiol linker, and (c) 5′ -CGCGAATTCGCG-3′ -thiol linker. (d)
Schematic illustration of a single DNA conductance measurement. (e) Natural loga-
rithm of GCGC(AT)m GCGC conductance vs. length (total number of base pairs). The
solid line is a linear fit that reflects the exponential dependence of the conductance on
length. (f) Conductance of (GC)n vs. 1/length (in total base pairs). Reprinted with
permission from [862]. Copyright 2004 American Chemical Society.
et al. [842], where the charge transfer rate in DNA molecules was measured.
Some of these results are shown in Fig. 18.7. As compared to the experi-
ment just described, Giese et al. found a weaker length dependence of the
transfer rates when several AT base pairs were inserted between CG base
pairs. These experiments showed the existence of two different processes for
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
Fig. 18.7 Sequence dependence of the charge transfer in DNA. Plot of log(PGGG /PG )
(PGGG /PG is proportional to the charge transfer rates) against the number n of the AT
base pairs. Each experiment was performed three times, and their relative errors are
within ±10 − 20%. The steep line corresponds to the coherent superexchange charge
transfer (tunneling). The flat line is drawn in order to make clear the weak distance de-
pendence. The arrows in the depicted DNA strands indicate the superexchange charge
transfer between the guanine radical cation G+ 22 and the GGG sequence for short dis-
tances (n = 2), or the hopping mechanism for long distances (n = 5), where - in addition
adenines act as charge carriers. For clarity, only the double strands with n = 2 and n = 5
are shown. The nucleotides in grey indicate all charge carriers. Reprinted by permission
from MacMillan Publishers Ltd: Nature [842], copyright 2001.
to gold electrodes at both ends of the molecules [855, 862]. Also, rel-
atively high currents have been measured through dsDNA molecules
covalently attached to single-walled carbon nanotubes (SWNT) [875]
and through duplex DNA coupled to SWNT electrodes via amide link-
ages [876].
• The sequence: The exact base sequence determines finally the level of
the current that can flow through DNA molecules [842, 862, 864]. In
particular, a sequence rich in CG pairs is expected to exhibit a higher
current.
18.4 Exercises
Chapter 19
In the previous chapters we have addressed the main transport regimes that
are realized in molecular junctions. In our discussion so far, we have fo-
cused our attention on the analysis of the electrical conductance. However,
there are many other transport properties that provide valuable informa-
tion, which often is not contained in the conductance. A paradigmatic ex-
ample is the current fluctuations or noise. Its investigation has contributed
decisively to our understanding of the transport mechanisms in a great va-
riety of mesoscopic and nanoscale devices [150]. On the other hand, the
charge transport is not the only important aspect in the context of conduc-
tion in molecular junctions. Thermal transport is also a key issue in the field
of molecular electronics from a fundamental as well as a from a practical
viewpoint. Molecular-scale contacts provide a new territory to study heat
conduction in regimes never explored before and, issues like heating will
have to be faced and understood, if molecular electronics wants to become
a viable technology. Obviously, the study of thermoelectric phenomena
in molecular junctions, resulting from the interplay between electrical and
thermal transport, can also give a new insight into the physics of these
nanocircuits.
For these reasons, we shall put aside the electrical conductance for a
while, and in this chapter we shall concentrate on the discussion of other
transport properties. To be precise, in section 19.1 we shall discuss the
basic physics of noise in molecular junctions and describe the first noise
experiments in this field. Then, we shall turn our attention to thermal
transport and in section 19.2 we shall present a detailed discussion of heat-
ing and heat conduction in molecular wires. Finally, section 19.3 is devoted
to the analysis of the thermopower, which is becoming a vital source of
novel information on molecular transport junctions.
553
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
Since the shot noise depends on the sum over the second power of the
transmission coefficients, this quantity is independent of the conductance,
P
G = G0 n τn , and the simultaneous measurement of these two quantities
1 Throughout this chapter we shall denote the transmission as τ in order to avoid con-
fusions with the temperature.
2 Here, we have taken into account the spin degeneracy that will be assumed in our
1.0
Transmission
0.8
0.3 0.6
0.4
x
Excess noise (2eI)
0.2
0.0
0 1 2 3 4
2
Conductance (2e/h)
0.2
0.1
20%
10%
x=0% 5%
0.0
1 2 3 4
2
Conductance (2e /h)
Fig. 19.1 Noise measurements on Au atomic contacts using the MCBJ technique. The
symbols correspond to the measured excess noise values for 27 contacts at 4.2 K with
a bias current of 0.9 µA. The different lines show the calculations using Eq. (19.5) in
the case of one single partially transmitted channel (full curve) and for various amounts
of contributions of other modes according to the model described in the inset (dashed
curves). In the limit of zero conductance, these curves all converge to full shot noise,
i.e. 2.9 10−25 A2 /Hz. Inset: transmission of modes in the case of x=10% contribution
from neighboring modes. Reprinted with permission from [885]. Copyright 1999 by the
American Physical Society.
Shot noise in atomic-scale contacts was first measured by van den Brom
and J.M. van Ruitenbeek using the MCBJ technique [885]. The measure-
ments were conducted at low temperatures to reduce the thermal noise.
However, in these experiments the noise level of the pre-amplifiers in gen-
eral exceeds the shot noise to be measured. Using two sets of pre-amplifiers
in parallel and measuring the cross-correlation, this undesired noise is re-
duced. By subtracting the zero-bias thermal noise from the current-biased
noise measurements, the pre-amplifier noise, present in both, is further
eliminated. For currents up to 1 µA the shot noise level was found to
have the expected linear dependence on current. For further details on the
measurement technique, we refer to [885].
In Fig. 19.1 we show the results of Ref. [885] for the noise of gold atomic
contacts as a function of the conductance of the junctions. The measured
shot noise is given relative to the classical shot noise value 2eI. All data are
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
2.5
0.006 (a) (b)
2.0
0.005
〈S I〉 (pA /Hz)
0.004 1.5
T (K)
2
*
0.003
1.0
0.002
0.5
0.001
0.000 0.0
-2 -1 0 1 2 -3 -2 -1 0 1 2 3
eV/2k B (K) eV/2k B (K)
strongly suppressed compared to the full shot noise value, with minima close
to 1 and 2 times the conductance quantum. For contacts with conductance
below 1 G0 the data are consistent with a single conduction channel having
a transmission probability τ = G/G0 , as expected for this monovalent
metal. For larger contacts there is a tendency for the channels to open
one-by-one, but admixture of additional channels grows rapidly. There
is a very strong suppression, down to F = 0.02, for G = 1 G0 , which
unambiguously shows that the current is carried dominantly by a single
channel. It needs to be stressed that this holds for gold contacts. There is a
fundamental distinction between this monovalent metal and the multivalent
metal aluminum, which shows no systematic suppression of the shot noise
at multiples of the conductance quantum, and the Fano factors lie between
about 0.3 and 0.6 for G close to G0 [886].
Shot noise measurements by Cron et al. [98] have provided a very strin-
gent experimental test of the multichannel character of the electrical con-
duction in Al atomic contacts. In these experiments the set of transmis-
sions τn were first determined independently by the technique of fitting the
subgap structure in the superconducting state, discussed in section 11.4.
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
Fig. 19.3 Shot noise measurements in Pt-D2 break junctions. The left panel shows the
point-contact spectroscopy (PCS) signal for this junction, with a clear vibrational mode
at 76 meV. The right panel shows the excess noise (the white noise in the current above
the thermal noise) as a function of the current. The noise is strongly suppressed below
the full Schottky noise for a tunnel junction. After each measurement of noise at a given
current, the PCS was measured again to verify that the contact had not changed. The
total conductance for this junction is G = 1.021 G0 , and the shot noise can be fitted with
two channels, τ1 = 1.000, and τ2 = 0.021, giving a Fano factor of F = 0.020. Adapted
with permission from [568]. Copyright 2006 American Chemical Society.
More recently, shot noise measurements have also been used to charac-
terize Pt-H2 O-Pt junctions [735] and Pt-benzene-Pt junctions [473]. In the
former case, the noise results indicated that for conductance below 1 G0
there are typically two conduction channels, although one clearly domi-
nates the transport. These results were very important to understand the
crossover between PCS and IETS and to test the so-called 1/2-rule (see
section 16.4). In the case of Pt-benzene junctions (see section 14.1.4),
the analysis of the shot noise results showed that for conductances around
1 G0 (and also well below) several channels contribute significantly to the
transport, while when the conductance is reduced to 0.2 G0 , the number of
channels is eventually reduced to one. As opposed to Pt-H2 O-Pt junctions,
in this case there is no dominant transmission channel when more than a
single channel exists. It was shown theoretically in same work [473] that
the number of channels is roughly determined by the number or carbon
atoms directly coupled to the Pt electrodes.
So far the shot noise measurements in molecular junctions have been
used to extract the channel transmissions in highly conductive junctions,
where the transport is supposed to be coherent. Notice that this applica-
tion is restricted to junctions with a high conductance, let us say above
0.1 G0 . Below that, the quadratic term in the transmission coefficients is
negligible and the shot noise becomes proportional to the conductance (i.e.
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
linear in the transmission coefficients). Anyway, the shot noise can provide
very important information also in other transport regimes. For instance,
shot noise measurements in the Coulomb blockade regime [887, 888] or in
the Kondo regime [889] have been only reported very recently in the con-
texts of carbon nanotubes and semiconductor quantum dots. In this sense,
weakly coupled molecular junctions (molecular transistors) can be an ideal
playground to further explore the noise in these transport regimes.
On the other hand, the vibrational effects discussed in previous chap-
ters can be further investigated with the help of the noise. For instance,
it has been predicted that the Franck-Condon blockade (see section 17.3.1)
is characterized by remarkably large Fano factors (102 -103 for realistic pa-
rameters), which arise due to avalanche-like transport of electrons [801].
The vibrationally-induced inelastic effects on noise properties of molec-
ular junctions in different transport regimes have been studied using NEGF
techniques by Zhu and Balatsky [784] and by Galperin and coworkers [890].3
Very recently, several theoretical groups have discussed the noise induced
by vibrations in the limit of weak electron-phonon coupling [891–893]. One
of the central issues of these papers was the discussion of the sign of the
inelastic noise as a function of the transmission, which is related to our dis-
cussion of the sign of the inelastic conductance in this regime (see section
16.4). The predictions of these papers could be in principle tested in the
type of experiments discussed above.
claimed that it misses vertex corrections even at the lowest order in the electron-phonon
interaction.
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
have dimensions that are much smaller than the inelastic mean free path
for phonons, even at room temperature, and thus they offer the possibility
to study phonon transport (and their contribution to thermal conduction)
in a very special regime.
Heat generation and heat conduction are intimately connected. Indeed,
heat conduction is an essential ingredient in the balance of the processes
that determines local heat generation. For this reason, we have chosen to
organize this section in the following way. After some general comments
about the problem of describing heating and heat conduction in molec-
ular junctions, we shall briefly review the work done so far on thermal
conductance. Then, we shall discuss the issue of heat generation in molec-
ular junctions and, in particular, we shall describe the main experiments
reported to date on this topic.
The subject of thermal properties of molecular junctions is presently
dominated by the theory. The experiments are rather scarce due to the
difficulties of measuring thermal transport at the nanoscale (for a review
on this subject see Ref. [894]). Although it is a very interesting subject,
we shall not discuss here in depth the theoretical techniques to describe
heat transport in molecular junctions, and we shall merely point out the
main ideas and challenges. For a more detailed discussion on the theory,
see section 9 of Ref. [695].
-11
x 10
20
15
K (W/K)
10
0
5 10 15 20
N
Fig. 19.4 Theoretical results for the heat transport coefficient (heat flux per unit T
difference between hot and cold bath) displayed as a function of alkane bridge length,
for a particular model of molecule-heat bath coupling at 50 K (full line), 300 K (dotted
line) and 1000 K (dashed line). Reprinted with permission from [903]. Copyright 2003,
American Institute of Physics.
edge of the heat burst traveled ballistically along the chains at a velocity
of 1 kilometer per second. The molecular conductance per chain was 50
pW/K.
The thermal conductance of alkane-based junctions was indeed ad-
dressed theoretically by Segal et al. [903] a few years before the realization
of the experiments mentioned above. These authors computed the phonon
contribution to the heat current, which should be the dominant one in these
low transmissive junctions. To be precise, they computed the heat flux for
a harmonic molecule characterized by a set of normal modes and coupled
through its end atoms to harmonic heat reservoirs. They have also per-
formed classical mechanics simulations in order to assess the role played by
anharmonicity. The general conclusions of this work are: (i) At room tem-
perature and below, molecular anharmonicity is not an important factor
in the heat transport properties of alkanes of length up to several tens of
carbon atoms. (ii) At room temperature, the efficiency of heat transport by
alkane chains decreases with chain size above 3-4 carbons, then saturates
and becomes length independent for moderate sizes of up to a few tens
of carbon atoms (this prediction agrees with the observations of Ref. [923]
mentioned above). (iii) At low temperature, the heat transport efficiency
increases with chain length. This is a quantum effect: at low temperatures
only low frequency modes can be populated and contribute to phonon trans-
port, however such modes are not supported by short molecules and become
available only in longer ones. In Fig. 19.4 we reproduce the results for the
thermal conductance of Segal et al. [903] that illustrate these conclusions.
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
60
50 C8
40 C10
T (K)
30 C12
20
10 C18
0
0 100 200 300 400 500
Bias (mV)
Fig. 19.5 Estimated junction temperature as a function of bias in alkane-dithiol junc-
tions of various chain lengths. Reprinted with permission from [766]. Copyright 2005
American Chemical Society.
(a) (b)
Temperature (K)
Temperature (K)
Bias voltage (V) Bias voltage (V)
Fig. 19.6 Measurements using surface enhanced Raman spectroscopy of the mode-
specific effective temperature [Teff (ν)] as a function of bias for two representative Ag-
SAMBPDT-Ag junctions. (a) Plot of Teff (ν) as a function of bias voltage for a mode
with 1,585 cm−1 (triangles) and 1,083 cm−1 (squares) modes (532 nm laser). (b) Plot of
Teff (ν) as a function of bias voltage for a mode with 1,585 cm−1 (triangles), 1,280 cm−1
(circles) and 1,083 cm−1 (squares) modes (671 nm laser). Reprinted by permission from
Macmillan Publishers Ltd: Nature Nanotechnology [938], copyright 2008.
ence.
6 Strictly speaking, this expression is only approximate. The numerator should be the
difference in electrochemical potential divided by −e, not the electric potential, see
Eq. (4.79). However, the chemical potential is often relatively constant as a function of
temperature, so using electric potential alone is in these cases a very good approximation.
7 Any thermal gradient gives rise to the transport of heat by the phonons, while an
electric current, though carried by the electrons, cannot fail to transfer some of its
momentum to the lattice vibration, and drag them along with it.
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
Fig. 19.7 Left panel: Schematic diagram of the modified MCBJ configuration, used
for the simultaneous measurement of conductance and thermopower of metallic atomic
contacts. Right panel: Density plot of thermopower of gold atomic contacts against con-
ductance constructed from 220 breaking curves for for two samples. Black represents no
data points and white more than 100. Reprinted with permission from [942]. Copyright
1999 by the American Physical Society.
HOMO LUMO
1 60
(a) 40 (b)
S (µV/K)
0.1 20
τ(E)
0
0.01 -20
-40
0.001 -60
-8 -7 -6 -5 -4 -3 -2 -8 -7 -6 -5 -4 -3 -2
EF (eV) EF (eV)
Fig. 19.8 (a) Transmission as a function of the Fermi energy computed from Eq. (19.12).
The values of the parameter are: ǫ1 = −7 eV, ǫ2 = −3 eV and ΓL = ΓR = 100 meV
(solid line) and ΓL = ΓR = 30 meV (dashed line). (b) Corresponding thermopower as a
function of the Fermi energy calculated from Eqs. (19.12) and (19.11).
on the position of the Fermi level, which we leave as a free parameter. Us-
ing Eq. (19.11) one can compute the thermopower and the result is shown
in Fig. 19.8(b). Notice that depending on the position of the Fermi energy
with respect to molecular levels, the thermopower can be either positive or
negative. If EF is closer to the HOMO, the sign is positive and one talks
about hole-dominated transport. If at the contrary, the LUMO is closer
to EF , then the thermopower is negative and one has electron-dominated
transport. Notice also that the thermopower is in the range of µV /K (or
larger), like in the case of atomic contacts, and therefore it should be mea-
surable. Finally, when EF is not too close to one of the frontier orbitals,
the thermopower is very similar for the two cases shown in Fig. 19.8(b),
although the broadenings differ by a factor of three. Indeed, if the Fermi
energy is located between the HOMO and LUMO and far away from them,
it is easy to show from Eqs. (19.12) and (19.11) that, to first order, the
thermopower is independent of the metal-molecule coupling [944].
The first experiment measuring the thermopower in single-molecule
junctions was reported by Reddy et al. [101]. These authors used STM
break junctions to trap molecules between two gold electrodes with a tem-
perature difference across them. In this way they were able to measure
the thermopower (or Seebeck coefficient) of 1,4-benzenedithiol (BDT), 4,4′ -
dibenzenedithiol (DBDT), and 4,4′′ -tribenzenedithiol (TBDT) in contact
with gold at room temperature and found the values +8.7 ± 2.1 µV/K,
+12.9 ± 2.2 µV/K, and +14.2 ± 3.2 µV/K, respectively. As explained
above, the positive sign indicates p-type (hole) conduction in these hetero-
junctions, i.e. the transport is dominated by the HOMO of the molecules.
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
It was also observed that S grows roughly linearly with the number N of
the phenyl rings in the molecule. These results are illustrated in Fig. 19.9.
This pioneering experiment motivated new theoretical work on this sub-
ject. Thus for instance, Pauly et al. [546] presented an ab initio (DFT-
based) study of the thermopower in metal-molecule-metal junctions made
up of dithiolated oligophenylenes contacted to gold electrodes. It was found
that, in agreement with the experiment, the transport is dominated by
the HOMO of these molecules. Moreover, it was shown that while the
conductance decays exponentially with increasing molecular length, the
thermopower increases linearly as in the experiments of Ref. [101]. This
is illustrated in Fig. 19.10, where the conductance and thermopower for
oligophenylenes with up to 4 phenyl rings are shown in panel (c) and (d),
respectively. Notice that the transmission functions for these molecules, see
Fig. 19.10(a), resemble those obtained with the simple model of Eq. (19.12).
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
R1 R1
EF R2 R2
-2
10
-10 EF
-3
10 (a) (b)
-4 -20
10 -7 -6 -5 -4 -3 -2 -1 -6 -5 -4 -3 -2
E (eV) E (eV)
0
10 40
-1
(c) R, exp (d)
10 30 R
S (µV/K)
G / G0
-2
10 20
-3
10 10
-4
10 0 0
1 2 3 4 0 1 2 3 4
N N
Fig. 19.10 Ab initio calculations for the conductance and thermopower of dithiolated
oligophenylenes contacted to gold electrodes. N is the number of phenyl rings in the
molecules. (a,b) Transmission function and the negative of its logarithmic derivative.
(c,d) The corresponding conductance (G) and thermopower (S). The experimental data
in (d) are form Ref. [101]. The straight lines are the best fits to the numerical results.
Adapted with permission from [546]. Copyright 2008 by the American Physical Society.
As one can see in Fig. 19.10(b), there is a qualitative agreement with the ex-
perimental results which, taking into account the usual theory-experiment
disagreement for the conductance, is certainly encouraging.
Pauly et al. also explained in simple terms the origin of the linear
increase of the thermopower with the length of the molecules (see also
Ref. [945]). As mentioned above, the transport in oligophenylenes proceeds
through the tail of the HOMO and the off-resonant tunneling is reflected
in the typical exponential decay of the linear conductance: G/G0 ∼ e−βN ,
where N is the number of phenyl rings. This off-resonant transport is
the origin of the linear increase in the thermopower. The idea goes as
follows. Assuming that the transmission around E = EF is of the form
τ (E) = α(E)e−β(E)N , then Eq. (19.11) yields S = SC + βS N , where
π 2 kB
2
T π 2 kB
2
T ′
SC = − [ln α(EF )]′ and βS = β (EF ). (19.13)
3e 3e
It is important to notice that, while SC depends on the prefactor α(E), βS
does not. Since α(E) contains the most significant uncertainties related to
the contact geometries, one expects βS to be described at a higher level
of confidence than SC . This linear dependence of the thermopower on
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
Fig. 19.11 Measurements of the Seebeck coefficient vs. molecular length for N -unit
phenylenedithiols (N = 1, 2, 3), phenylenediamines (N = 1, 2, 3), and alkanedithiols
(N = 2, 3, 4, 5, 6, 8). Fit lines to the data indicate that thermopower increases with
length at a similar rate (βS ) for phenylenediamines and phenylenedithiols but decreases
with length for alkanedithiols. Reprinted with permission from [947]. Copyright 2009
American Chemical Society.
the Fermi energy. As shown in section 13.4, one can use simple bridge
models to obtain an expression of the energy dependence of the β-factor.
Assuming off-resonant tunneling, we obtained in section 13.4 the expres-
sion of Eq. (13.15) for β. Such a simple model predicts that a positive
thermopower is accompanied by an increase with length. This explains the
trend for the phenylenes, but not for the alkanes.8 The authors of Ref. [947]
suggested an explanation of this peculiar behavior in terms of gold-sulfur
metal induced gap states residing between the HOMO and the LUMO. As
we mentioned in section 14.1.2, several theoretical groups have concluded
that the transport in alkanes can be influenced by states that originate from
the hybridization of gold and sulfur orbitals with localized orbitals of the
alkane chain [948–950]. Indeed, evidence of the existence of these states has
been reported in STM experiments [951]. These hybrid states are localized
at the interfaces and therefore, they are expected to have a major impact
on the conductance for short molecules, while for long ones the transport is
expected to be dominated by the HOMO of the alkane chains. The different
length dependence of the metal induced gap states and the HOMO of the
chains could be the origin of the decreasing thermopower [947].
As we already discussed in section 13.6, the transmission function of
a molecular contact can exhibit lineshapes that completely differ from the
double-peak structure shown in Fig. 19.8(a). One way to increase the ther-
mopower is by “engineering” a much more pronounced energy dependence
of the transmission function. As shown recently by Finch et al. [487], some
molecules can exhibit sharp Fano resonances very close to the Fermi energy
that in turn can lead to a huge thermopower in molecular junctions.
In the discussion so far we have focused on the thermopower in the co-
herent transport regime. However, this transport property can also provide
very valuable information in many other transport regimes. For instance,
Koch et al. [800] have shown theoretically that the thermopower of weakly
coupled molecular junctions can give access to the electronic and vibrational
excitation spectrum of the molecule even in a linear-response measurement.
To summarize, we have shown in this section that thermopower mea-
surements in molecular junctions provide very important information not
contained in the conductance. In this sense, we believe that measurements
of this thermoelectric property will play a crucial role in the immediate
future of molecular electronics.
8 The simple bridge model of section 13.4 suggests that a decreasing S with length can
only be obtained when the transport is dominated by the LUMO and therefore, the
thermopower is negative.
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
Chapter 20
579
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
With those questions in mind, we have organized the rest of this chapter
as follows. First, we shall describe recent experiments in which molecu-
lar junctions have been characterized using surface-enhanced Raman spec-
troscopy. Then, we shall discuss the physical mechanisms that are expected
to play a major role in irradiated atomic and molecular junctions. In par-
ticular, we shall pay special attention to the so-called photon-assisted tun-
neling or current rectification.1 Section 20.4 presents a description of some
recent experimental results on the electronic transport through irradiated
atomic and molecular contacts. In section 20.5 we shall briefly discuss the
phenomenon of current amplification and other novel transport phenom-
ena that have been predicted to appear in ac driven molecular junctions.
Section 20.6 is devoted to the analysis of fluorescence of current carrying
junctions. Finally, in section 20.7 we shall review some of the experiments
in which the optical properties of certain molecules have been exploited to
design primitive molecular optoelectronic devices.
With respect to the first question posed above, it is obvious that combined
optical and transport experiments on molecular transport junctions could
reveal a wealth of additional information beyond that available from purely
electronic measurements. It is, however, very challenging to use conven-
tional optical spectroscopies to obtain local information about molecular
junctions. First of all, it is not easy to inject light into slits of molecu-
lar size between two metal leads and second, the molecular emission may
be strongly damped because of the proximity to a metal surface. Fortu-
nately, recent work has shown that surface-enhanced Raman spectroscopy
(SERS) can offer a way out of these problems [953–956]. The idea is based
on the fact that metallic nanostructures, similar to those used to form the
electrodes of molecular junctions, can act as effective plasmonic antennas,
leading to a dramatic enhancement of the electric field locally at the junc-
tion region (see e.g. Ref. [957]). This enhanced field can then be used to
perform Raman spectroscopy of objects placed in these nanogaps (for a
review on SERS, see e.g. Ref. [958]). This idea has been explored recently
in the context of molecular electronics, in particular, by Natelson’s group
1 Thesetwo terms are sometimes believed to refer to two different physical mechanisms.
However, we shall show in section 20.3 that they are indeed identical.
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
Fig. 20.1 The upper panel shows the Raman spectrum (1 s integration) for a single
FOPE molecule in a gold junction formed by electromigration. The lower panel shows
the correlated measurement of the conductance of this junction. The Raman mode ob-
served between 1950 and 2122 cm−1 is believed to be for the same 2122 cm−1 mode
associated with the C≡C stretch of the FOPE molecule. The large spectral shifts ob-
served for this mode are attributed to interactions between the molecule and its nanogap
environment. Clear correlations between the Raman structure and conductance can be
seen. In particular in region B and for part of region E the Raman spectrum is observed
to disappear while the conductance drops to zero. Reprinted with permission from [607].
Copyright 2008 IOP Publishing Ltd.
Raman and IET spectra, which, unfortunately, was not yet possible in the
experiments just described. As we explained in the previous chapter, IETS
requires cryogenic temperatures, at which it is not easy to operate a Raman
microscope. However, in the experiment of Ioffe et al. [938], which was dis-
cussed in section 19.2.3, a comparison between vibrational modes revealed
by Raman scattering and IETS could be established, although these two
type of measurements were not performed at the same time. Let us re-
mind that these authors reported SERS measurements of junctions based
on biphenyldithiol SAMs3 . In this experiment the Raman spectra were
acquired at room temperature, while the IET spectra were obtained in
transport measurements at 4 K. Interestingly, all Raman active vibrational
modes were revealed in the IETS measurements, in spite of the fact that the
selection/propensity rules are different in these two types of spectroscopies.
Let us also recall that the main goal of this work was to measure the voltage
dependence of the effective temperature of these current-carrying junctions.
This was achieved by measuring both the Stokes and anti-Stokes compo-
nents of the Raman scattering. Then, the effective temperature Tef f (ν) for
each mode was calculated at each bias using the following expression4
IAS (νL + νν )4 σAS A2AS
= exp (−hνν /kB Tef f (ν)) , (20.1)
IS (νL − νν )4 σS A2S
where IAS(S) is the intensity of the anti-Stokes (Stokes) Raman mode,
νL(ν) is the frequency of the laser (Raman mode), σAS(S) is the anti-Stokes
(Stokes) scattering cross-section of the adsorbed molecules and AAS(S) is
the average local field enhancement at the molecules at the anti-Stokes
(Stokes) frequency. Strictly speaking, this expression is only valid in ther-
modynamical equilibrium and one may wonder whether this relation still
holds at a finite bias voltage. For a discussion of this issue we refer the
reader to Refs. [959, 960], where a detailed theoretical study of Raman
scattering in current-carrying molecular junctions is presented.
can play a role in this problem, which is by no means a trivial task. The
theoretical and experimental work reported so far on this subject suggests
that the main “suspects” are the following:
This list is not complete, and other effects can also play an important
role. In particular, some readers might miss surface plasmons in this list.
In this respect, we would like to say that for optical frequencies there is no
doubt that plasmons play a key role. Surface plasmons are responsible for
the local field distribution at the junction and, in particular, for its enhance-
ment with respect to the incident field. In this sense, we can consider that
plasmons determine the effective amplitude (and frequency dependence) of
the ac potential induced in the junction, but the transport mechanism is
still PAT or current rectification. In other words, we rather prefer to say
that plasmons play an important role in the PAT mechanism than to say
they constitute a different mechanism. After all, an ac field would also
appear in the absence of plasmons and their role is only to modify the field
distribution.
As we have already mentioned above, the importance of the different
mechanisms depends primarily on the radiation frequency. For instance, in
the microwave range PAT (or current rectification) largely dominates the
transport. This has been firmly established in a great variety of mesoscopic
structures [214] and more recently in atomic and molecular junctions (see
discussion below). In the optical range, however, the other three mecha-
nisms can also be very important.
In the this section we shall present a description of the PAT theory for
several reasons. First, this mechanism is likely to operate in almost any
situation since when a junction is illuminated most of the radiation indeed
impinges on the electrodes. Second, it is believed to be the dominant one
at low frequencies and finally, recent experiments in atomic and molecular
contacts seem to suggest that this mechanism is the dominant one even at
optical frequencies. In what follows, we shall first present the basic theory
of PAT and then, we shall discuss the basic predictions of this theory for
atomic and molecular junctions.
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
so far. The total time-dependent voltage will always be denoted as V (t), i.e. including
explicitly the time argument.
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
..
.
J2 (α) exp(+2i ω t)
J−2(α) exp(−2i ω t)
..
.
Fig. 20.2 Virtual energy levels generated according to Eq. (20.3) by adiabatic modula-
tion of the energy Ei (t) for each quasiparticle state within the ungrounded electrode of
a junction in the presence of an ac field.
Gordon formula of Eq. (20.4) under the following assumptions: (i) The
energy dependence of the lead density of states is negligible and (ii) the
ac potential does not vary spatially along the central part of the junction
(i.e. across the molecule in the case of molecular junctions). While the first
assumption is often justified, the second one may seem rather restrictive.
However, Viljas et al. [221, 222] have shown that if the amplitude of the
ac voltage is not too large, the precise shape of the profile does not play a
crucial role.
In the derivation of Eq. (20.4) it was assumed that the ac voltage is
applied to one of the electrodes, while the other is grounded. If we consider
that the ac voltage is applied symmetrically, i.e. it drops equally at both
interfaces, the current can be then written as
X∞ h ³ α ´i2
I(V ; α, ω) = Jn I0 (V + 2n~ω/e). (20.5)
n=−∞
2
Here, assuming that the transport in the absence of ac drive is elastic, the
current I0 is given by the standard Landauer formula. At low temperatures
and in the linear response regime (vanishing dc bias), the conductance takes
the particularly simple form
X∞ h ³ α ´i2
G(V = 0; α, ω) = G0 Jn τ (EF + n~ω), (20.6)
n=−∞
2
d2 I0
µ ¶
1 2
∆Idc (V ; α, ω) ≈ Vac . (20.8)
4 dV 2
This expression reproduces the classical result for rectification [212] and
illustrates the connection between PAT and current rectification. This is a
very important relation since it provides a direct way to test whether the
dominant transport mechanism is indeed PAT/current rectification. Such
test requires to measure independently the induced dc current and the
second derivative of the current with respect to the bias in the absence
of radiation. Moreover, according to Eq. (20.8), the ratio of these two
quantities gives the amplitude of the ac bias, which is typically unknown.
This amplitude gives information about the field enhancement locally in
the junction region.
Notice that Eq. (20.8) suggests that if the I-V characteristics in the
absence of radiation exhibit an asymmetry at vanishingly bias voltage due
to material and/or geometrical asymmetries (i.e. if d2 I0 /dV 2 6= 0 at V = 0),
a radiation-induced current can flow in the system even in the absence
of any dc bias voltage. This phenomenon of rectification at zero dc bias
voltage was predicted by Cutler et al. in 1987 [964] and it was first reported
by Walther’s group in 1991 in laser-driven STM experiments on graphite
surfaces [965] (for a detailed discussion of this phenomenon, see the review
of Ref. [215]). The current generated by the ac field in the absence of dc
bias is often referred to as photocurrent.9
The classical expression of Eq. (20.8) is probably valid in a wide range
of molecular junctions for microwave frequencies, while in the optical range
significant deviations from this expression are likely to appear and it has
to be replaced by its quantum version of Eq. (20.7) [see Exercise 20.2(ii)].
On the other hand, it is interesting to derive similar expressions for the
linear conductance. Defining the induced linear conductance correction as
∆Gdc (α, ω) ≡ G(V = 0; α, ω)−G(ω = 0), where G(ω = 0) = G = G0 τ (EF ),
9 Laterin this chapter we shall discuss the so-called ratchet effect in molecular junctions,
which is just another name for rectification at zero bias.
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
3 3
(a) Au contact τtotal (c) Pt contact τtotal
2.5 τ1 2.5 τ1
Transmission
Transmission
τ2 τ2
2 2
τ3 τ3
1.5 τ4 1.5 τ4
1 1
0.5 0.5
0 0
-5 -4 -3 -2 -1 0 1 2 3 4 5 -5 -4 -3 -2 -1 0 1 2 3 4 5
E-EF (eV) E-EF (eV)
λ (nm) 800 400 λ (nm) 800 400
Gdc(ω) / G0
Fig. 20.3 Theoretical results for the photoconductance of Au and Pt atomic contacts.
(a) Equilibrium transmission τtotal and its decomposition into conduction channels
τ1,2,3,4 for an Au dimer contact. (b) Zero-temperature photoconductance for several
values of α as a function of frequency ω computed using Eq. (20.6). In (b) the wave-
lengths λ with a tick spacing of 400 nm are shown. The range of visible light is indicated
by vertical dotted lines. (c-d) The same as in panels (a-b) but for a Pt contact. Adapted
with permission from [221]. Copyright 2007 by the American Physical Society.
tact10 of Au and Pt. Here, we just reproduce the results obtained with the
simple approximation of Eq. (20.6), which was usually found to reproduce
qualitatively the more rigorous results obtained with the NEGF formalism
[221]. In the case of Au, as can seen in Fig. 20.3(a), the conductance for
ω = 0 is equal to 1 G0 with a single open channel arising from the contri-
bution of the 6s orbitals. Moreover, notice that the transmission around
EF is very flat. Due to this flatness, for frequencies up to ~ω ≈ 1.5 eV
(λ ≈ 827 nm) the effect of radiation is practically negligible. In the red
part of the visible range (~ω . 2 eV) ∆Gdc > 0 and it can reach up to 20%
depending on the value of α.11 This increase in the conductance is due to
the contribution of the 5d bands located 2 eV below EF , where the number
of open transmission channels is higher than at EF .
10 The exact geometry of this dimer contact can be seen in Fig. 1 of Ref. [221]. This
type of geometry is typically responsible for the last conductance plateau in the breaking
process of an atomic contact.
11 It is important to remark that for the case of Au it was found that the results were
quite sensitive to the exact profile of the ac voltage. For the other metals the profile did
not play a major role.
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
Fig 20.3(c-d) show the corresponding results for Pt. In the absence
of radiation the conductance is close to 2.1 G0 due to the contributions
of mainly three conduction channels, which originate from the 6s and 5d
orbitals. In this case, and in general in the contact regime for Pt, the effect
of the radiation is always a significant reduction in conductance. This
is understandable, since EF lies at the edge of the d band, and photon
absorption leads to an energy region where less open transmission channels
are available and τtotal is smaller.
To conclude, the key message is that the photoconductance simply re-
flects the energy dependence of the transmission of the contacts. The sign
of the induced correction can be both positive (like for Au) and negative
(like for Pt) depending on the material. With respect to the order of mag-
nitude of the correction, it can reach up to 50%-100% in some special cases
depending on the geometry, frequency and power of the radiation, but it is
usually below those values.
1.5 1
(a) R1
(a) EF 0.1
Gdc / G0
1 G2
T(E)
0.01
D = 0.97 nm
0.5 0.001 (f) G1
R2
0 0.0001
-8 -6 -4 -2 1 2 3 4
E (eV) n
1
(b) α = 1.8
Gdc(ω) / G0
D = 1.40 nm (c)
R3 0.1
0.01 α = 0.2 R1 R2
D = 1.83 nm
R4 1
Gdc(ω) / G0
0.01
D = 2.27 nm R3 R4
(b) 0.001
V ac
0 0.5_ 1 1.5 2 0 0.5_ 1 1.5 2
h ω (eV) h ω (eV)
ω
Z 2π/ω h √ i
I(VB , VJ ) = I VB + 2VJ cos(ωt) dt. (20.10)
2π 0
Here, the function I(V ) in the integrand has the same form as the static
I-V characteristics measured in the absence of microwave, but with a time
dependent argument. An example of the fits is shown in Fig. 20.5(b) for
VIN = 5 mV. Notice the high accuracy of the fit, which provides a strong
support for the interpretation of the results in terms of rectification. More-
over, from the fits the value of VJ could be extracted. A plot of VJ versus
VIN is shown in Fig. 20.5(c), yielding a slope 45.5 from the best linear fit.
This slope gives a direct information about the field enhancement at the
contact.
On the other hand, it was also shown that the induced dc current as
a function of voltage for a single Mn atom follows closely the d2 I/dV 2
spectrum in the absence of microwaves (see Fig. 2 of Ref. [971])12 . This
can be understood from Eq. (20.8), which tells us that the correction to
the dc current is proportional to the d2 I/dV 2 spectrum without radiation.
Notice that such relation can also be derived from Eq. (20.10) in the limit
of small VJ by expanding the integrand up to second order in VJ . The close
relation between the induced dc current and the d2 I/dV 2 spectra was also
found in the case of transport through individual MnCO molecules, which
constitutes a convincing proof of the fact that the rectification mechanism
dominates the transport in irradiated atomic-scale junctions at microwave
frequencies.
12 The main difference between this experiment and previously reported ones was the use
of low temperatures (∼ 18 K) that made possible to measure directly d2 I/dV 2 spectra.
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
13 Itwas found that continuous irradiation of the devices with λ = 488 nm for several
seconds with a power of a few mW results in irreversible conductance changes.
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
16
150
14
P = 1.9 mW
laser
= 515 nm
ctance
12 sample condu
LIS
10 100
G/G [%]
G0]
8 150
G [
i
i
6
6 100 50
4 4
50
2 2
0 0
0 0
880 900 920 940 960 980 1000
Time [s]
Fig. 20.6 Conductance and light-induced relative conductance change ∆G/Gi versus
time when opening the break junction continuously. Inset: close-up of the few-atom
region with Gi < 9 G0 . Reprinted with permission from [973]. Copyright 2007 by the
American Physical Society.
in which the gold wires are fully anchored onto Si/SiO2 substrates [975]. As
a result the atomic junctions are mechanically highly stable even at room
temperature, and under irradiation their heat dissipation characteristics
are far more efficient than those of suspended MCBJs, resulting in only
residual heating. Thanks to this method the authors were able to carry out
a detailed study of the influence of laser light in the conductance of gold
atomic contacts with conductances equal to 1 G0 .
In this work, the junctions were irradiated with three different lasers
with wavelengths of 532 nm (2.33 eV), 658 nm (1.88 eV), 781 nm (1.58 eV).
The maximum used power of the lasers was ∼ 20 mW, all measurements
were performed under ambient conditions at room temperature and the
junctions were placed with their long axis parallel to the laser polarization.
Fig. 20.7 shows representative results of the conductance as a function
of laser intensity for two different contacts and two different wavelengths.
Notice that in the absence of light the conductance (measured at 30 mV)
is ∼ 1 G0 . In all cases the conductance is enhanced by laser irradiation and
the relative changes, which increase with decreasing wavelength, are below
10%. In order to establish a comparison with the results of PAT theory,
Eq. (20.6) was used with the transmission curve of Fig. 20.3(a). The results
for different values of α = eVac /~ω are shown in Fig. 20.3(b). Using α
as an adjustable parameter the authors were able to fit the experimental
results with a reasonable accuracy, see solid lines in Fig. 20.7. Notice in
particular the nonlinear behavior, which is a remnant of the Bessel functions
of Eq. (20.6).
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
where ∆ is the hopping matrix elements that describes the coupling of each
orbital to its nearest neighbors and ǫn stands for the one-site energies. The
time-dependent part in the second term of this Hamiltonian describes the
coupling to an oscillating dipole field that causes time-dependent level shifts
xn a(t), where xn = (N + 1 − 2n)/2 denotes the scaled position of site n.
The energy a(t), which is periodic in time, is determined by the electrical
field strength multiplied by the electron charge and the distance between
two neighboring sites.
At a first glance, one might have the impression that models based on
the Hamiltonian of Eq. (20.12) describe very different physics from the
Tien-Gordon PAT theory detailed above. However, one can show that
if the spatial dependence of the field-matter interaction in Eq. (20.12) is
neglected, i.e. if the driving shifts all the wire levels simultaneously, it is
possible to map the driving field by a gauge transformation to oscillating
chemical potentials. In other words, models based on Hamiltonians like
the one in Eq. (20.12) reproduce the simple Tien-Gordon-like results in the
limiting case of spatially homogeneous field-matter interaction.
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
1
0
0
1 hω 0110
0
1 10
0
1
0
1 1010
0
1 ∆ 1010 ΓR
ΓL 1010 ∆
∆ ∆ 10
0
1
0
1 ε2 ε3 ε5
1010
0
1
0
1 ε1 ε4 1010
0
1
0
1 1010
L 101010 1010
1010
R
0
1
0
1
0
1 −2 −1 0 1 2
1010
0
1 x 10
Fig. 20.8 Schematic representation of the level structure of a molecular bridge with
N = 5 sites coupled to metallic electrodes.
sion of ac forces without any net bias into directed motion. In the context
of molecular junctions the question is whether it is possible to induce a
dc current with an ac field in the absence of a dc bias voltage. As we
commented in section 20.3.1, in the context of STM it is well-known that
this is indeed possible if there are left-right asymmetries in the junctions
[964, 965, 215, 216]. There, the ratchet effect is referred to as rectification
at zero dc bias. In the context of molecular junctions, this issue has been
extensively studied by Lehmann et al. [982–984] and the main results have
been reviewed in Ref. [211]. These authors have shown that it possible
to generate a dc current with a pure ac driving (i.e. a photocurrent) by
introducing certain asymmetries in the problem.15 For instance, using a
conductor with an asymmetric level structure, one can generate a dc cur-
rent even with a purely harmonic dipole driving. Another possibility is to
use a driving field in which several frequencies are mixed. In this case, a
dc current is generated even in spatially symmetric molecules bridges.
Related to the ratchet effect, Galperin and Nitzan [977] have predicted
that light-induced current in unbiased junctions (i.e. photocurrents) can
flow when the bridging molecule is characterized by a strong charge-transfer
transition. Such a current reaches its maximum when the light frequency
matches the internal transition frequencies of the molecule. Using realistic
estimates of molecule-lead coupling and molecule-radiation field interaction,
these authors showed that such an effect should be observable.
Coherent destruction of tunneling.– As we saw in our discussion of PAT
in section 20.3, the current as a function of the amplitude of the ac driv-
ing is modulated according to the behavior of the Bessel functions, see
Eqs. (20.4) and (20.5). If the parameter α is such that J0 vanishes, there
is a pronounced reduction of the current [see Exercise 20.2(iv)]. This phe-
nomenon appears in many different ac driven systems and it is known as
coherent destruction of tunneling [985]. For a detailed discussion of this
phenomenon in the context of molecular wires, see section 7 in Ref. [211]
and references therein.
Role of electron excitation in the leads.– As we discussed in section 20.2,
apart from modulating the electronic levels in the leads, the electromagnetic
field can produce hot electrons in the leads by direct photon absorption.
These electronic excitations can in turn contribute to the transport. Sim-
ple estimates of the contribution of these inelastic processes to the total
current have been put forward long ago in the context of the STM [961].
15 Tobe precise, the generation of a photocurrent requires the breaking of the so-called
generalized parity [211].
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
More recently, Galperin and coworkers have studied in detail the influ-
ence of electronic excitations in the leads on the current through molecular
transport junctions [962]. These authors have concluded that in certain
situations such excitations can give a significant contribution to the cur-
rent and moreover, this contribution can be distinguished from the direct
current because it scales differently with the distance between the molecule
and the leads.
Let us conclude this section by saying that the different physical effects
discussed in the previous paragraphs can also have a big impact in other
transport properties like shot noise. The theoretical activities along these
lines have been reviewed by Kohler et al. in Ref. [211].
In this section we shall face the question number 5 of our list in the in-
troduction of this chapter. When a sufficiently high bias voltage is applied
to a molecular junction, two molecular orbitals can be partially populated
and then optical transitions between them become, in principle, possible
with the subsequent light emission (or fluorescence). Is the light emission
from a single molecule measurable? If so, what can we learn about the
junctions from this local optical spectroscopy? The goal of this section is
to briefly describe the recent experimental and theoretical efforts devoted
to answer these and other basic questions related to the current-induced
light emission from single molecules in transport junctions.
The fact that electron tunneling can lead to emission of light was first
discovered by Lambe and McCarthy in 1976 in the context of metal-oxide-
metal tunnel junctions [986]. In the context of atomic-scale junctions, light
emission has been frequently observed in STM experiments. Thus for in-
stance, it has been reported in clean metal [987, 988] and semiconductor
surfaces [989], as well as for atomic and molecular adsorbates on metal sub-
strates [990–993]. However, often the reported photon emission spectra do
not show identifiable molecule-related features [992]. On a metal surface,
the electronic levels of a molecule are considerably broadened whereas light
emission is strongly quenched, making it difficult to detect and identify any
molecule-specific emission.
In recent years, it has been demonstrated by means of STM experi-
ments that electric-current flow through a molecule may indeed cause the
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
EF B
hω
A
eV
hω
EF
Fig. 20.9 Diagram showing the two major processes contributing to STM-excited light
emission from a molecule adsorbed on a surface. In process A, the inelastic electron
tunneling channel, an electron tunnels from the Fermi level of the STM tip into an
unoccupied molecular orbital with simultaneous excitation of a plasmon. In process B,
the fluorescence channel, an electron tunnels into the higher unoccupied orbital of the
molecule. The charged molecule typically relaxes to a lower vibrational level of the same
electronic level, with subsequent radiative (excitation of a plasmon) transition to the
lower electronic level. The final step involves tunneling of this extra electron into the
substrate.
750 1
1500
3
500
1000 NiAl
1
2 2
500 3 250
Photon counts
NiAl 4
4
oxide oxide
0 0
C D
Counts
Time has come to address the last question posed in the introduction.
One of the dreams in molecular electronics is to use the amazing optical
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
(a) (b)
Fig. 20.12 Left panel: Schematic cross section of the device layout of a large-area
molecular junction in which the diarylethene is sandwiched between Au and poly(3,4-
ethylenedioxythiophene): poly(4-styrenesulphonic acid) (PEDOT:PSS)/Au. Using UV
(312 nm) illumination the open, nonconjugated isomer can be converted to the closed,
conjugated isomer. Visible irradiation of 532nm reverses the photoisomerization process.
Right panel: Current density (J) versus voltage (V ) of the closed (circles) and open
(squares) isomers as self-assembled in the molecular junctions, and J-V characteristics
of the junctions with the open isomer self-assembled and subsequently photoisomerized
to the closed isomer with UV irradiation (triangles). Averaged data (at least 35 de-
vices) from devices with diameters of 10-100 mm. Error bars by standard deviation.
Reproduced with permission from [1010]. Copyright Wiley-VCH Verlag GmbH & Co.
KGaA.
clusively in the open or closed states. The conductance through the closed,
more-conducting state of the switch is shown to be 16 times higher at 0.75
V bias. On the other hand, devices with the open isomer were illuminated
for 15 min with 312 nm UV irradiation to convert the molecular switches in
the devices to the closed isomer. The J-V characteristics of the converted
open-state isomer after UV irradiation show an increase of the conductance
through the monolayer, as expected from the devices with closed isomers
present, see Fig. 20.12 (right panel). Following UV irradiation and consecu-
tive measurements, these devices were illuminated with 532 nm irradiation
(visible light) to achieve ring opening of the switches in the SAM. The
observed J-V characteristics show a significant decrease (by a factor 3)
of the conductance upon visible light irradiation, but the conductance of
the devices with the open isomer is not fully recovered. The origin of this
behavior was not fully understood (see Ref. [1010] for more details).
Let us mention that light-controlled conductance switching of molec-
ular devices based on photochromic diarylethene molecules has also been
demonstrated by van der Molen et al. [135]. In this case, the devices con-
sisted of ordered, two-dimensional lattices of gold nanoparticles, in which
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
states in a molecular junction. So far, there have been very few experi-
ments where molecular signatures have been unambiguously identified in
the fluorescence spectra and more experiments are needed. For the theory
the challenge is now to provide quantitative predictions that can be directly
compared with the experiments.
The last topic that we have discussed in this chapter is very important
from the technological point of view. Molecular electronics is often seen as a
field that aims at reproducing the standard microelectronic components and
devices, but at a smaller scale. However, the future of molecular electronics
depends crucially on our ability to provide devices with new functionalities
out of the scope of more traditional technologies. The optical properties
of many molecules may offer a down-to-earth possibility for the future. In
principle, they can be used in transport junctions to control the current at
will and many researchers believe by now that molecular optoelectronics
will soon grow as a field of its own. However, so far it has been difficult
to take advantage of those optical properties and the only successful imple-
mentations have made use of photochromic molecular switches. At present,
it is not clear whether the difficulties encountered so far are just of techni-
cal nature or there are true fundamental limitations. In any case, the next
years will be certainly exciting for the scientists working on this subject.
20.9 Exercises
2
X 4ΓL ΓR
τ (E) = ,
i=1
(E − ǫi )2 + (ΓL + ΓR )2
18 This model was extensively used in section 19.3 to describe the thermopower in molec-
ular junctions.
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
were ǫ1 and ǫ2 is the energy of HOMO and LUMO, respectively, and ΓL and
ΓR the broadenings by contacts L and R. For the sake of concreteness, we
shall assume throughout this exercise that ǫ1 = −7.0 eV, ǫ2 = −3.0 eV and
ΓL = ΓR = 30 meV.
(i) Use the second equality in Eq. (20.9), i.e. the classical rectification formula,
to analyze the low-frequency radiation-induced correction to the conductance
(∆Gdc ) as a function of the Fermi level position (EF ). In particular, show that
∆Gdc is negative only when EF is very close to one of the frontier orbitals.
(ii) Let us now assume that the Fermi level lies in the middle of the HOMO-
LUMO gap, i.e. EF = −5.0 eV and consider the limit of small ac amplitudes
(α ≪ 1). Compute ∆Gdc as a function of photon energy and show that for
~ω & 0.5 eV, the classical rectification formula fails to describe the correction to
the conductance given by the first equality in Eq. (20.9).
(iii) Now assume that EF = −6.0 eV and use Eq. (20.6) to compute ∆Gdc
as a function of the photon energy in the interval ~ω ∈ [0 eV, 4 eV] for different
values of α. For α ≪ 1 you will find the appearance of two peaks at 1 and 3 eV.
What is the origin of these peaks? Finally, compare the results obtained with the
exact formula of Eq. (20.6) and with the approximation of Eq. (20.9) to establish
in which range of α this approximation is valid.
(iv) One of the key signatures of PAT is the appearance of additional steps
in the I-V characteristics. Assume that EF = −6.0 eV and ~ω = 0.5 eV and
use Eq. (20.5) or Eq. (8.74) to compute the I-V curves and the corresponding
differential conductance for α = 0, 1, 2, 4. Discuss the origin of current steps (or
the corresponding peaks in the differential conductance) induced by the radia-
tion. Finally, analyze the phenomenon of coherent destruction of tunneling by
computing the I-V curves and differential conductance for values of α for which
J0 in Eq. (20.5) vanish, i.e. α/2 = 2.405, 5.520, . . . .
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
Chapter 21
anisms by which the electron is transferred through the bridge/molecular wire are essen-
tially the same.
617
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
2 It is not obvious what is the role of inter-molecular interactions in these systems and in
some cases, the transport characteristics may differ significantly from the corresponding
ones of a single-molecule device.
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
PART 5
Appendixes
621
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
622
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
Appendix A
Second Quantization
623
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
It is easy to show from their definitions that these operators satisfy the
following commutation relations
[a, a† ] = 1 ; [a, a] = 0 ; [a† , a† ] = 0. (A.7)
With these relations, the Hamiltonian adopts the following form
· ¸
~ω £ † †
¤ † 1
H= aa + a a = ~ω a a + . (A.8)
2 2
The three commutators above plus this Hamiltonian completely specify
the harmonic oscillator problem in terms of operators. With these four
relationships, one can show that the eigenvalue spectrum is indeed that of
Eq. (A.4). The eigenstates are
(a† )n
|ni = √ |0i, (A.9)
n!
where |0i is the ground state which obeys
a|0i = 0 (A.10)
and where the n! is for normalization. Operating on this state by a creation
operator gives
(a† )n+1
a† |ni = √ |0i = (n + 1)1/2 |n + 1i (A.11)
n!
the state with the next highest integer. In the same way, one can show that
a|ni = (n)1/2 |n − 1i, (A.12)
which shows that the annihilation operator a lowers the quantum number.
Then operating by the sequence
a† a|ni = a† (n)1/2 |n − 1i = n|ni (A.13)
gives an eigenvalue n, which verifies the eigenvalue relation A.4. Further-
more, using the original definition of Eq. (A.5) permits us to express x and
p in terms of these operators as
µ ¶1/2
~
x= (a + a† ) (A.14)
2mω
µ ¶1/2
m~ω
p=i (a† − a). (A.15)
2
The description of the harmonic oscillator in terms of operators is equiv-
alent to the conventional method of using wave functions ψn (ξ) of position.
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
The systems that we will be dealing with are composed of many identi-
cal particles such as electrons or phonons. In quantum mechanics those
identical particles are indistinguishable. Thus for instance, no electron can
be distinguished from another electron, except by saying where it is, what
quantum state it is in, etc. Internal quantum-mechanical consistency re-
quires that when we write down a many-identical-particle state, we make
that state noncommittal as to which particle is in which single-particle
state. For example, we say that we have electron 1 and electron 2, and we
put them in states a and b respectively, but exchange symmetry requires
(since electrons are fermions) that a satisfactory wave-function has the form
Φ(r1 , r2 ) = A [φa (r1 )φb (r2 ) − φa (r2 )φb (r1 )] , (A.30)
i.e. the wave function is antisymmetric with respect to the exchange of the
two electrons: Φ(r1 , r2 ) = −Φ(r2 , r1 ). This is a consequence of Pauli’s
exclusion principle that states that there cannot be two fermions in the
same quantum sate.
If we have N particles, the wave functions must be either symmetric or
antisymmetric under exchange depending on the nature of the particles:
ΦB (r1 , ..., ri , ..., rj , ..., rN ) = ΦB (r1 , ..., rj , ..., ri , ..., rN ) (Bosons)
ΦF (r1 , ..., ri , ..., rj , ..., rN ) = −ΦF (r1 , ..., rj , ..., ri , ..., rN ) (Fermions).
In particular, in the fermionic case the antisymmetry of the wave func-
tion can be ensured by using Slater determinants. But, can we satisfy the
antisymmetry principle without using Slater determinants? Second quan-
tization is a formalism in which the antisymmetry property of the wave
function has been transferred onto the algebraic properties of certain oper-
ators. Second quantization introduces no new physics. It is just another,
although very elegant, way of treating many-electron systems, which shifts
the emphasis away from N -electron wave functions to the one- and two
electron matrix elements of the different operators. This has been illus-
trated already in the case of bosons with the analysis of the phonons in a
1D chain in the previous section, and we shall now concentrate on the case
of fermions.
That is, the Φk1 ,...,kN (r1 , ..., rN ) for the non-interacting system are the basis
states used to describe the interacting system.
Now these are rather clumsy expressions to carry around, so it would
be desirable to have a more compact way of writing them. This may be
achieved by noting that since all particles are indistinguishable, the essential
information in Eq. A.31 is just how many particles there are in each single-
particle state. Therefore, we could equally well specify the state of the
non-interacting system by writing Φ as
Φk1 ,...,kN (r1 , ..., rN ) = Φnp1 ,np2 ,...,npi ,... (r1 , ..., rN ). (A.33)
For short, we shall represent this as
Φnp1 ,np2 ,...,npi ,... (r1 , ..., rN ) ≡ |np1 , np2 , ..., npi , ...i (A.34)
meaning: np1 particles in state φp1 , np2 in φp2 , etc., where nk = 0 or 1
by the Pauli principle. This is called “occupation number notation”. For
brevity, from now on we shall drop the p’s and just use the numerical
subscripts. Then
Φ = |n1 , n2 , ..., ni , ...i. (A.35)
It is important to remember that the |n1 , n2 , ..., ni , ...i are orthonormal
because the Φk1 ,...,kN are, and we may write this in the following way
hn′1 , n′2 , ..., n′i , ...|n1 , n2 , ..., ni , ...i = δn′1 ,n1 δn′2 ,n2 ...δn′i ,ni ... (A.36)
Up to this point we have been dealing with systems containing a fixed
number of particles. Now we take an important step, and, even though the
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
0 Φ0 |000...i
1 Φ1 , Φ2 , Φ3 , ... |100...i, |010...i, |001...i, ...
2 Φ12 , Φ13 , Φ23 , ... |1100...i, |1010...i, |0110...i, ...
.. .. ..
. . .
N Φk1 ,k2 ,...,kN |n1 , n2 , ..., ni , ...i
.. .. ..
. . . (A.37)
where
That is, we get a factor of (−1) for each particle (i.e., each occupied state)
standing to the left of the state i in the wave function. For example
ci |..., 0i , ...i = 0 , c†i |..., 1i , ...i = 0
c3 |11111000...i = +|11011000...i
c†4 |1110100...i = −|11111000...i
c†2 c3 c†1 c2 c†3 c1 |1100...i = −|1100...i. (A.42)
One of the nice properties of the c†i operators is that by applying them
repeatedly to the true vacuum state (state with no particles in it), it is
possible to generate all other states, thus:
|n1 , n2 , ...i = (c†1 )n1 (c†2 )n2 ...|0000...i. (A.43)
For example
|011000...i = c†2 c†3 |0000...i. (A.44)
Another important property of the c†i , ci operators is that they are
hermitian adjoint of each other, i.e. c†i = (ci )† . The demonstration is left
to the reader. This property shows that c†i , ci are non-hermitian and are
therefore not observables. It is, however, easy to construct a hermitian
operator from c†i and ci as follows. The combination
|n1 , ..., ni , ...i are just simultaneous eigenfunctions of the number operators
n̂1 , ..., n̂i , ....
The c†i , ci operators obey the following important fermion commutation
rules:
{cl , c†k } = cl c†k + c†k cl = δlk ; {cl , ck } = {c†l , c†k } = 0. (A.47)
These can be easily proved from the definitions of Eqs. (A.39) and (A.40).
Thus for instance, the second relation can be shown as follows:
cl ck |n1 , ..., nl , ..., nk , ...i = (−1)Σk nk cl |n1 , ..., nl , ..., nk − 1, ...i (A.48)
Σk +Σl
= (−1) nk nl |n1 , ..., nl − 1, ..., nk − 1, ...i
Σl
ck cl |n1 , ..., nl , ..., nk , ...i = (−1) nl ck |n1 , ..., nl − 1, ..., nk , ...i
= (−1)(−1)Σk +Σl nk nl |n1 , ..., nl − 1, ..., nk − 1, ...i,
where the extra (−1) on line four comes from the fact that there is one less
particle to the left of state k. Adding the two equations yields the second
rule in Eq. (A.47). The other rules may be established in a similar fashion.
The importance of the above set of anti-commutation relations lies in
the fact that all the antisymmetry properties are built into them. Therefore,
by using them in the right places, we do not have to worry either about the
symmetry of the wave functions themselves, or even about the awkward
(−1)Σ factors.
where h(i) is any operator involving only the ith electron. These operators
represent dynamic variables that depend only on the position or momentum
of the electron in question, independent of the position or momentum of
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
other electrons. Examples are operators for the kinetic energy, attraction
of an electron to a nucleus, dipole moment, and most of the other operators
that one encounters. The second type of operator is a sum of two-electron
operators
N
1X
O2 = V (i, j), (A.50)
2
i6=j
where the sums run over the set {ψi }. Here, the different matrix elements
are defined as follows
Z
hij ≡ dr1 ψi∗ (r1 )h(r1 )ψj (r1 ) (A.54)
Z Z
Vijkl ≡ dr1 dr2 ψi∗ (r1 )ψj∗ (r2 )V (r1 , r2 )ψk (r1 )ψl (r2 ). (A.55)
Let us now demonstrate that we can also recover this result using the second
quantization expression for the operator O1 . In this case,
hij hΨ|c†i cj |Ψi.
X
hΨ|O1 |Ψi = (A.57)
ij
Since both cj and c†i are trying to destroy an electron (cj to the right and
c†i to the left), the indices i and j must belong to the set {a, b, ...} and thus
X
hΨ|O1 |Ψi = hab hΨ|c†a cb |Ψi. (A.58)
ab
Using
c†a cb = δab − cb c†a (A.59)
to move c†a to the right, we have
hΨ|c†a cb |Ψi = δab hΨ|Ψi − hΨ|cb c†a |Ψi. (A.60)
The second term on the right is zero since c†a
is trying to create an electron
in ψa , which is already occupied in |Ψi. Since hΨ|Ψi = 1, we finally have
X X
hΨ|O1 |Ψi = hab δab = haa . (A.61)
ab a
with the electrons interacting with a potential U (r), such as the lattice
potential in a solid, and with each other through particle-particle interac-
tions V (ri − rj ), typically the Coulomb interaction. As we have learned
above, this Hamiltonian can be written in terms of the fermionic creation
and annihilation operators as
1X
hij c†i cj + Vijkl c†i c†j cl ck ,
X
H= (A.63)
ij
2
ijkl
where
· 2 2 ¸
~ ∇
Z
hij = dr ψi∗ (r) − + U (r) ψj (r)
2m
Z Z
Vijkl = dr1 dr2 ψi∗ (r1 )ψj∗ (r2 )V (r1 − r2 )ψk (r1 )ψl (r2 ).
The basic premise is to get rid of the atoms and to replace them with
a uniform positive background charge of density n0 . The homogeneous
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
electron gas is also called jellium model. One can think of taking the
positive charge of the ions and spreading it uniformly about the unit cell
of the crystal. Of course, the homogeneous electron gas has no crystal
structure. To preserve charge neutrality, the average particle density of the
electron gas must also be n0 . This model, although a bit academic, has
played a key role to understand basic issues about the Coulomb interaction
of a many-particle system. A detailed discussion of this model can be found
in Ref. [174].
The plane-wave model is often a poor approximation of electron be-
havior in solids where the electrons are localized on atomic sites and only
occasionally hop to neighboring sites. This behavior is described by the
tight-binding model, where the basis is formed by localized atomic-like or-
bitals.3 One simple form of this model is bilinear in the operators:
tij c†iσ cjσ .
X
H= (A.67)
ijσ
where
e2
Z Z
Vijkl = dr1 dr2 φ∗ (r1 − Ri )ψ ∗ (r2 − Rj ) ψ(r1 − Rk )ψ(r2 − Rl ).
|r1 − r2 |
The four orbitals could be centered on four different sites. These are called
four-center integrals. They are usually small and often neglected in many-
body calculations.
3 Tight-binding models and their used in molecular electronics are the subject of Chapter
9.
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
The Hubbard model [177] retains only the Coulomb integral which is the
largest, namely that in which all four orbitals φ(r) are centered on the same
site. This term describes the interaction between two electrons which are
on the same atom. Since two electrons cannot be in the same state, the two
on the same atom must be in different atomic states. In the simplest model,
which considers only a single orbital state on each atom, the two electrons
must have different spin configurations. One has spin up, while the other
has spin down. The Hubbard model considers the following Hamiltonian
e2
Z Z
U = Viiii = dr1 dr2 |φ∗ (r1 )|2 |φ∗ (r1 )|2 . (A.71)
|r1 − r2 |
The second quantization formalism for bosons was already outlined when
we discussed the physics of phonons in section A.1.2. Anyway, for the sake
of completeness, we summarize here the main results of this formalism for
the case of bosons:
(1) The many-body wave functions for a bosonic system has to be sym-
metric with respect to the particle exchange. In this sense, the Slater
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
(4) The one- and two-body operators are expressed in terms of the creation
and annihilation operators in the same way as in the fermion case.
A.4 Exercises
Bibliography
639
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
Bibliography 641
Bibliography 643
[69] C.J. Muller, J.M. van Ruitenbeek, L.J. de Jongh, Experimental observation
of the transition from weak link to tunnel junction, Physica C 191, 485 (1992).
[70] J.M. Krans, C.J. Muller, I.K. Yanson, Th.C.M. Govaert, R. Hesper, J.M.
van Ruitenbeek, One-atom point contacts, Phys. Rev. B 48, 14721 (1993).
[71] J.M. van Ruitenbeek, A. Alvarez, I. Piñeyro, C. Grahmann, P. Joyez, M.H.
Devoret, D. Esteve, C. Urbina, Adjustable nanofabricated atomic size con-
tacts, Rev. Sci. Instrum. 67, 108 (1996).
[72] A.I. Yanson, I.K. Yanson, J.M. van Ruitenbeek, Observation of shell struc-
ture in sodium nanowires, Nature 400, 144 (1999).
[73] N. van der Post, Superconductivity and magnetism in nano-scale junctions,
PhD thesis, Leiden, The Netherlands (1997).
[74] J.M. Krans, Size effects in atomic-scale point contacts, PhD thesis, Leiden,
The Netherlands (1996).
[75] A.M.C. Valkering, A.I. Mares, C. Untiedt, K.B. Gavan, T.H. Oosterkamp,
J.M. van Ruitenbeek, A force sensor for atomic point contacts, Rev. Sci.
Instrum. 76, 103903 (2005).
[76] S.A.G. Vrouwe, E. van der Giessen, S.J. van der Molen, D. Dulic, M.L.
Trouwborst, B.J. van Wees, Mechanics of lithographically defined break junc-
tions, Phys. Rev. B 71, 35313 (2005).
[77] E. Scheer, P. Joyez, D. Esteve, C. Urbina, M.H. Devoret, Conduction channel
transmissions of atomic-size aluminum contacts, Phys. Rev. Lett. 78, 3535
(1997).
[78] L. Grüter, M.T. González, R. Huber, M. Calame, C. Schönenberger, Elec-
trical conductance of atomic contacts in liquid environments, Small 1, 1067
(2005).
[79] H.S.J. van der Zant, Y.-V, Kervennic, M. Poot, K. O’Neill, Z. de Groot,
J.M. Thijssen, H.B. Heersche, N. Stuhr-Hansen, T. Bjørnholm, D. Vanmaekel-
bergh, C.A. van Walree, L.W. Jenneskens, Molecular three-terminal devices:
fabrication and experiment, Faraday Discuss. 131, 347 (2006).
[80] Z.M. Wu, M. Steinacher, R. Huber, M. Calame, S.J. van der Molen, C.
Schönenberger, Feedback controlled electromigration in four-terminal nano-
junctions, Appl. Phys. Lett. 91, 053118 (2007).
[81] M.L. Trouwborst, S.J. van der Molen, B.J. van Wees, The role of Joule
heating in the formation of nanogaps by electromigration, J. Appl. Phys. 99,
114316 (2006).
[82] H.B. Heersche, G. Lientschnig, K. O’Neill, H.S.J. van der Zant, H.W. Zand-
bergen, In situ imaging of electromigration-induced nanogap formation by
transmission electron microscopy, Appl. Phys. Lett. 91, 072107 (2007).
[83] R. Hoffmann, D. Weissenberger, J. Hawecker, D. Stöffler, Conductance of
gold nanojunctions thinned by electromigration, Appl. Phys. Lett. 93, 043118
(2008).
[84] A. Lodder, Calculations of the screening of the charge of a proton migrating
in a metal, Phys. Rev. B 74, 045111 (2006).
[85] K.-N. Tu, J.W. Mayer, L.C. Feldman, E lectronic Thin Film Science for
Electrical Engineers and Materials Scientists, (McMillan, New York, USA,
1992).
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
[86] A.R. Champagne, A.N. Pasupathy, D.C. Ralph, Mechanically adjustable and
electrically gated single-molecule transistors, Nano Lett. 5, 305 (2005).
[87] C. Schirm, H.-F. Pernau, E. Scheer, Switchable wiring for high-resolution
electronic measurements at very low temperatures, Rev. Sci. Instrum. 80,
024704 (2009).
[88] J. Heinze, Cyclic voltammetry-electrochemical spectroscopy, Angew. Chem.
Int. Ed. 23, 831 (1984).
[89] F.-Q. Xie, L. Nittler, Ch. Obermair, Th. Schimmel, Gate-controlled atomic
quantum switch, Phys. Rev. Lett. 93, 128303 (2004).
[90] K. Terabe, T. Hasegawa, T. Nakayama, M. Aono, Quantized-conductance
atomic switch, Nature 433, 47 (2005).
[91] A.F. Morpurgo, C.M. Marcus, D.B. Robinson, Controlled fabrication of
metallic electrodes with atomic separation, Appl. Phys. Lett. 74, 2084 (1999).
[92] M. Reyes-Calvo, A.I. Mares, V. Climent, J.M. van Ruitenbeek, C. Un-
tiedt, Formation of atomic-sized contacts controlled by electrochemical meth-
ods, phys. stat. sol. (a) 204, 1677 (2007).
[93] C.Z. Li and N.J. Tao, Quantum transport in metallic nanowires fabricated
by electrochemical deposition/dissolution, Appl. Phys. Lett. 72, 23 (1998).
[94] R. Waitz, O. Schecker, E. Scheer, Nanofabricated adjustable multicontact
devices on membranes, Rev. Sci. Instrum. 79, 093901 (2008).
[95] R. Waitz, R. Hoffmann, E. Scheer, unpublished.
[96] C.A. Martin, R.H.M. Smit, H.S.J. van der Zant, J.M. van Ruitenbeek, A
nanoelectromechanical single-atom switch, Nano Lett. 9, 2940 (2009).
[97] A. Kumar, L. Saminadayar, D.C. Glattli, Y. Jin, B. Etienne, Experimental
test of the quantum shot noise reduction theory, Phys. Rev. Lett. 76, 2778
(1996).
[98] R. Cron, M.F. Goffman, D. Esteve, C. Urbina, Multiple-charge-quanta shot
noise in superconducting atomic contacts, Phys. Rev. Lett. 86, 4104 (2001).
[99] R. Aguado and L.P. Kouwenhoven, Double quantum dots as detectors of
high-frequency quantum noise in mesoscopic conductors, Phys. Rev. Lett. 84,
1986 (2000).
[100] R. Deblock, E. Onac, L. Gurevich, L.P. Kouwenhoven, Detection of quan-
tum noise from an electrically driven two-level system, Science 301, 203
(2003).
[101] P. Reddy, S.-Y. Jang, R. Segalman, A. Majumdar, Thermoelectricity in
molecular junctions, Science 315, 1568 (2007).
[102] K. Baheti, J.A. Malen P. Doak, P. Reddy, S.-Y. Jang, T.D. Tilley, A. Ma-
jumdar, R. A. Segalman, Probing the chemistry of molecular heterojunctions
using thermoelectricity, Nano Lett. 8, 715 (2008).
[103] P.J.F. Harris, Carbon Nanotube Science - Synthesis, Properties and Appli-
cations, (Cambridge University Press, Cambridge, UK, 2009).
[104] D. Krüger, H. Fuchs, R. Rousseau, D. Marx, M. Parinello, Pulling
monatomic gold wires with single molecules: An ab initio simulation, Phys.
Rev. Lett. 89, 186402 (2002).
[105] J.Ch. Love, L.A. Estroff, J.K. Kriebel, R.G. Nuzzo, G.M. Whitesides, Self-
assembled monolayers of thiolates on metals as a form of nanotechnology.
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
Bibliography 645
Bibliography 647
[158] J.C. Cuevas, A. Martin-Rodero, A. Levy Yeyati, Shot noise and coherent
multiple charge transfer in superconducting quantum point contacts, Phys.
Rev. Lett. 82, 4086 (1999).
[159] Y. Naveh and D.V. Averin, Nonequilibrium current noise in mesoscopic dis-
ordered superconductor-normal-metal-superconductor junctions, Phys. Rev.
Lett. 82, 4090 (1999).
[160] U. Sivan and Y. Imry, Multichannel landauer formula for thermoelectric
transport with application to thermopower near the mobility edge, Phys. Rev.
B 33, 551 (1986).
[161] P. Streda, Quantised thermopower of a channel in the ballistic regime, J.
Phys.: Condens. Matter 1, 1025 (1989).
[162] P. N. Butcher, Thermal and electrical transport formalism for electronic mi-
crostructures with many terminals, J. Phys.: Condens. Matt. 2, 4869 (1990).
[163] H. van Houten, L.W. Molenkamp, C.W.J. Beenakker, C.T. Foxon, Thermo-
electric properties of quantum point contacts, Semicond. Sci. Technol. 7, B215
(1992).
[164] L.D. Landau and E.M. Lifschitz, Statistical Physics, (Pergamon Press, New
York, USA, 1959).
[165] M. H. Pedersen and M. Büttiker, Scattering theory of photon-assisted elec-
tron transport, Phys. Rev. B 58, 12993 (1998).
[166] M. Büttiker and T. Christen, Basic Elements of Electrical Conduction, in B.
Kramer (Ed.), Quantum Transport in Semiconductor Submicron Structures,
NATO ASI Series E, Vol. 326, p. 63, (Kluwer Academic Publishers, Dordrecht,
NL, 1996).
[167] M. Büttiker and T. Christen, Admittance and nonlinear transport in quan-
tum wires, point contacts, and resonant tunneling barriers, in L. L. Sohn,
L. P. Kouwenhoven, G. Schön (Eds.), Mesoscopic Electron Transport, NATO
ASI Series E, Vol. 345. p. 259., (Kluwer Academic Publishers, Dordrecht, NL,
1997).
[168] Y. Imry and R. Landauer, Conductance viewed as transmission, Rev. Mod.
Phys. 71, S306 (1999).
[169] Y. Imry, Introduction to Mesoscopic Physics, (Oxford University Press,
Oxford, UK, 1997).
[170] C.W.J. Beenakker, Random-matrix theory of quantum transport, Rev. Mod.
Phys. 69, 731 (1997).
[171] M. Büttiker, Role of quantum coherence in series resistors, Phys. Rev. B
33, 3020 (1986).
[172] M. Büttiker, Quantized transmission of a saddle-point constriction, Phys.
Rev. B 41, 7906 (1990).
[173] A.L. Fetter and J.D. Walecka, Quantum Theory of Many-Particle Systems,
(McGraw-Hill, New York, USA, 1971).
[174] G.D. Mahan, Many-Particle Physics, (Plenum Press, New York, USA,
1990).
[175] R.D. Mattuck, A Guide to Feynman Diagrams in the Many-Body Problem,
(Dover Publications, New York, USA, 1992).
[176] A. Szabo and N. Ostlund, Modern Quantum Chemistry: Introduction to
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
Bibliography 649
Bibliography 651
[216] A. Levy Yeyati and F. Flores, Photocurrent effects in the scanning tunneling
microscope, Phys. Rev. B 44, 9020 (1991).
[217] S. Datta and M.P. Anantram, Steady-state transport in mesoscopic systems
illuminated by alternating fields, Phys. Rev. B 45, 13761 (1992).
[218] A.-P. Jauho, N.S. Wingreen, Y. Meir, Time-dependent transport in inter-
acting and noninteracting resonant-tunneling systems, Phys. Rev. B 50, 5528
(1994).
[219] T. Brandes, Truncation method for Green’s functions in time-dependent
fields, Phys. Rev. B 56, 1213 (1997).
[220] B. Wang, J. Wang, H. Guo, Current partition: A nonequilibrium Green’s
function approach, Phys. Rev. Lett. 82, 398 (1999).
[221] J.K. Viljas and J.C. Cuevas, Role of electronic structure in photoassisted
transport through atomic-sized contacts, Phys. Rev. B 75, 075406 (2007).
[222] J.K. Viljas, F. Pauly and J.C. Cuevas, Modeling elastic and photoassisted
transport in organic molecular wires: Length-dependence and current-voltage
characteristics, Phys. Rev. B 77, 155119 (2008).
[223] N.W. Ashcroft and N.D. Mermin, Solid State Physics, (Saunders, USA,
1976).
[224] W.A. Harrison, Electronic Structure and the Properties of Solids, (Dover
Publication, New York, USA, 1989).
[225] W.A. Harrison, Elementary Electronic Structure, (World Scientific Publish-
ers, Singapore, 1999).
[226] R.M. Martin, Electronic Structure: Basic Theory and Practical Methods,
(Cambridge University Press, Cambridge, UK, 2004).
[227] P.W. Atkins and R.S. Friedman, Molecular Quantum Mechanics (3rd edi-
tion), (Oxford University Press, Oxford, UK, 1997).
[228] R. Hoffmann. An Extended Hückel Theory. I. Hydrocarbons, J. Chem. Phys.
39, 1397 (1963).
[229] E. Hückel, Quantentheoretische Beiträge zum Benzolproblem. I Die Elek-
tronenkonfiguration des Benzols und verwandter Verbindungen, Zeitschrift für
Physik 70, 204 (1931).
[230] E. Hückel, Quantentheoretische Beiträge zum Benzolproblem. II Quanten-
theorie der induzierten Polaritäten, Zeitschrift für Physik 72, 310 (1931).
[231] E. Hückel, Quantentheoretische Beiträge zum Benzolproblem. III Quan-
tentheoretische Beiträge zum Problem der aromatischen und ungesättigten
Verbindungen., Zeitschrift für Physik 76, 628 (1932).
[232] C.A. Coulson, B. O’Leary, R.B. Mallion, Hückel Theory for Organic
Chemists, Academic Press, San Diego, USA, 1978).
[233] M. Wolfsberg and L.J. Helmholz, The spectra and electronic structure of the
tetrahedral ions MnO− −− −
4 , CrO4 , and ClO4 , J. Chem. Phys. 20, 837 (1952).
[234] J.C. Slater and G.F. Koster. Simplified LCAO method for the periodic po-
tential problem, Phys. Rev. 94, 1498 (1954).
[235] D.A. Papaconstantopoulos, Handbook of Electronic Structure of Elemental
Solids, (Plenum Press, New York, USA, 1986).
[236] M.D. Stiles, Generalized Slater-Koster method for fitting band structures,
Phys. Rev. B 55, 4168 (1997).
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
[237] P.R. Wallace, The band theory of graphite, Phys. Rev. 71, 622 (1947).
[238] K.S. Novoselov, A.K. Geim, S.V. Morozov, D. Jiang, Y. Zhang, S.V.
Dubonos, I.V. Grigorieva, A.A. Firsov, Electric field effect in atomically thin
carbon films, Science 306, 666 (2004).
[239] A.K. Geim and K.S. Novoselov, The rise of graphene, Nature Mat. 6, 183
(2007).
[240] A.K. Geim and A.H. MacDonald, Graphene: Exploring carbon flatland,
Physics Today 60, 35 (2007).
[241] A.K. Geim and P. Kim, Carbon Wonderland, Scientific American 290, 90
(2008).
[242] R. E. Cohen, M. J. Mehl, D. A. Papaconstantopoulos, Tight-binding total-
energy method for transition and noble metals, Phys. Rev. B 50, 15694 (1994).
[243] M.J. Mehl and D.A. Papaconstantopoulos, Applications of a new tight-
binding total energy method for transition and noble metals: Elastic constants,
vacancies, and surfaces of monatomic metals, Phys. Rev. B 54, 4519 (1996).
[244] M. J. Mehl and D. A. Papaconstantopoulos, Tight-binding parametrization
of first-principles results, in Topics in Computational Materials Science, p.
169., C. Y. Fong (Ed.) (World Scientific Publishers, Singapore, 1998).
[245] D. A. Papaconstantopoulos and M. J. Mehl, The Slater-Koster tight-binding
method: a computationally efficient and accurate approach, J. Phys.: Con-
dens. Matter 15, R513 (2003).
[246] P.L. Pernas, A. Martin-Rodero, F. Flores, Electrochemical-potential varia-
tions across a constriction, Phys. Rev. B 41, 8553 (1990).
[247] F. Pauly, J.K. Viljas, U. Huniar, M. Häfner, S. Wohlthat, M. Bürkle,
J.C. Cuevas, G. Schön, Cluster-based density-functional approach to quantum
transport through molecular and atomic contacts, New J. Phys. 10, 125019
(2008).
[248] T.N. Todorov, Tight-binding simulation of current-carrying nanostructures,
J. Phys.: Condens. Matter 14, 3049 (2002).
[249] A. Martin-Rodero, F. Flores, N.H. March, Tight-binding theory of tunneling
current with chemisorbed species, Phys. Rev. B 38, 10047 (1988).
[250] P. Sautet and C. Joachim, Calculation of benzene on rhodium STM images,
Chem. Phys. Lett. 185, 23 (1991).
[251] N. Mingo, L. Jurczyszyn, F.J. Garcia-Vidal, R. Saiz-Pardo, P.L. de Andres,
F. Flores, S.Y. Wu, W. More, Theory of the scanning tunneling microscope:
Xe on Ni and Al, Phys. Rev. B 54, 2225 (1996).
[252] J. Cerdá, M.A. Van Hove, P. Sautet, M. Salmeron, Efficient method for the
simulation of STM images. I. Generalized Green-function formalism, Phys.
Rev. B 56, 15885 (1997).
[253] P. Sautet and C. Joachim, Electronic interference effects induced by a ben-
zene ring embedded in a polyacetylene chain, Chem. Phys. Lett. 153, 511
(1988).
[254] P. Sautet and C. Joachim, Electronic transmission coefficient for the single-
impurity problem in the scattering matrix approach, Phys. Rev. B 38, 12238
(1988).
[255] V. Mujica, M. Kemp, M.A. Ratner, Electron conduction in molecular wires.
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
Bibliography 653
2745 (1984).
[276] D.M. Ceperly and B.J. Alder, Ground state of the electron gas by a stochas-
tic method, Phys. Rev. Lett. 45, 566 (1980).
[277] A.D. Becke, Density functional exchange-energy approximation with correct
asymptotic bahavior, Phys. Rev. A 38, 3098 (1986).
[278] J.P. Perdew, Density-functional approximation for the correlation energy
of the inhomogeneous electron gas, Phy. Rev. B 33, 8822 (1986).
[279] J.P. Perdew, K. Burke, and M. Ernzerhof, Generalized gradient approxi-
mation made simple, Phys. Rev. Lett. 77, 3865 (1996); Erratum: Phys. Rev.
Lett. 78, 1396 (1997).
[280] J.P. Perdew and Y. Wang, Accurate and simple analytic representation
ofthe electron-gas correlation energy, Phy. Rev. B 45, 13244 (1992).
[281] C. Lee, W. Yang, R.G. Parr, Development of the Colle-Salvetti correlation-
energy formula into a functional of the electron density, Phys. Rev. B 37, 785
(1988).
[282] A.D. Becke, A new mixing of Hartree-Fock and local density-functional the-
ories, J. Chem. Phys. 98, 1372 (1993).
[283] A.D. Becke, Density functional thermochemistry III. The ole of exact ex-
change, J. Chem. Phys. 98, 5648 (1993).
[284] P.J. Stephens, J.F. Devlin, C.F. Chabalowski, M.J. Frisch, Ab initio cal-
culations of vibrational absorption and circular dichroism spectra using SCF,
MP2, and density functional theory force fields, J. Phys. Chem. 98, 11623
(1994).
[285] A.P. Scott and L. Radom, Harmonic vibrational frequencies: An evaluation
of Hartree-Fock, Moller-Plesset, quadratic configuration interaction, density
functional theory, and semiempirical scale factors, J. Phys. Chem. 100, 16502
(1996).
[286] A. Pecchia and A. Di Carlo, Atomistic theory of transport in organic and
inorganic nanostructures, Rep. Prog. Phys. 67, 1497 (2004).
[287] M. Brandbyge, J.-L. Mozos, P. Ordejón, J. Taylor, K. Stokbro, Density-
functional method for nonequilibrium electron transport, Phys. Rev. B 65,
165401 (2002).
[288] S.N. Yaliraki, A.E. Roitberg, C. Gonzalez, V. Mujica, M.A. Ratner, The
injecting energy at molecule/metal interfaces: Implications for conductance of
molecular junctions from an ab initio molecular description, J. Chem. Phys.
111, 6997 (1999).
[289] Y. Xue, S. Datta, and M.A. Ratner, Charge transfer and ”band lineup”
in molecular electronic devices: A chemical and numerical interpretation, J.
Chem. Phys. 115, 4292 (2001).
[290] J.J. Palacios, A.J. Pérez-Jiménez, E. Louis, E. SanFabián, J.A. Vergés,
First-principles approach to electrical transport in atomic-scale nanostruc-
tures, Phys. Rev. B 66, 035322 (2002).
[291] J. Heurich, J.C. Cuevas, W. Wenzel, G. Schön, Electrical transport through
single-molecule junctions: from molecular orbitals to conduction channels,
Phys. Rev. Lett. 88, 256803 (2002).
[292] Y. Xue and M.A. Ratner, Microscopic study of electrical transport through
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
Bibliography 655
individual molecules with metallic contacts. I. Band lineup, voltage drop, and
high-field transport, Phys. Rev. B 68, 115406 (2003).
[293] P.S. Damle, A.W. Ghosh, S. Datta, Unified description of molecular con-
duction: From molecules to metallic wires, Phys. Rev. B 64, 201403 (2001).
[294] P.S. Damle, A.W. Ghosh, S. Datta, First-principles analysis of molecular
conduction using quantum chemistry software, Chem. Phys. 281, 171 (2002).
[295] M.P. López Sancho, J.M. López Sancho, J. Rubio, Quick iterative scheme
for the calculation of transfer matrices: application to Mo (100), J. Phys. F:
Met. Phys. 14, 1205 (1984).
[296] M.P. López Sancho, J.M. López Sancho, J. Rubio, Highly convergent
schemes for the calculation of bulk and surface Green functions, J. Phys.
F: Met. Phys. 15, 851 (1985).
[297] J. Taylor, H. Guo, J. Wang, Ab initio modeling of quantum transport prop-
erties of molecular electronic devices, Phys. Rev. B 63, 245407 (2001).
[298] A. Di Carlo, M. Gheorghe, P. Lugli, M. Sternberg, G. Seifert, T. Frauen-
heim, Theoretical tools for transport in molecular nanostructures, Physica B
314, 86 (2002).
[299] J.K. Tomfohr and O.F. Sankey, Complex band structure, decay lengths, and
Fermi level alignment in simple molecular electronic systems, Phys. Rev. B
65, 245105 (2002).
[300] K.S. Thygesen and K.W. Jacobsen, Molecular transport calculations with
Wannier functions, Chem. Phys. 319, 111 (2005).
[301] A.R. Rocha, V.M. Garcı́a-Suárez, S. Bailey, C. Lambert, J. Ferrer, S. San-
vito, Spin and molecular electronics in atomically generated orbital landscapes,
Phys. Rev. B 73, 085414 (2006).
[302] J.M. van Ruitenbeek, Conductance quantization in metallic point contacts,
in K.H. Meiwes-Broer (Ed.), Metal Clusters on Surfaces: Structure, Quantum
Properties, Physical Chemistry, p. 175 (Springer-Verlag, Heidelberg, Berlin,
D, 2000).
Available in cond-mat/9910394.
[303] A.P. Sutton and J.B. Pethica, Inelastic flow processes in nanometre volumes
of solids, J. Phys.: Condens. Matter 2, 5317 (1990).
[304] U. Landman, W.D. Luedtke, N.A. Burnham, R.J. Colton, Atomistic mech-
anisms and dynamics of adhesion, nanoindentation and fracture, Science 248,
454 (1990).
[305] L. Olesen, E. Lægsgaard, I. Stensgaard, F. Besenbacher, J. Schiøtz, P.
Stoltze, K.W. Jacobsen, J.K. Nørskov, Reply on comment on quantized con-
ductance in an atom-sized point contact, Phys. Rev. Lett. 74, 2147 (1995).
[306] J.M. Krans, J.M. van Ruitenbeek, V.V. Fisun, I.K. Yanson, L.J. Jongh,
The signature of conductance quantisation in metallic point contacts, Nature
375, 767 (1995).
[307] N. Agraı̈t, J.G. Rodrigo S. Vieira, Conductance steps and quantization in
atomic-size contacts, Phys. Rev. B 47, 12345 (1993).
[308] J.I. Pascual, J. Méndez, J. Gómez-Herrero, A.M. Baró, N. Garcı́a, V.
Thien Binh, Quantum contact in gold nanostructures by scanning tunneling
microscopy, Phys. Rev. Lett. 71, 1852 (1993).
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
Bibliography 657
Bibliography 659
Bibliography 661
[410] A. Smogunov, A. Dal Corso, E. Tosatti, Complex band structure with ul-
trasoft pseudopotentials: fcc Ni and Ni nanowire, Surf. Sci. 532, 549 (2003).
[411] A. Delin and E. Tosatti, Magnetic phenomena in 5d transition metal
nanowires, Phys. Rev. B 68, 144434 (2003).
[412] J. Velev and W.H. Butler, Domain-wall resistance in metal nanocontacts,
Phys. Rev. B 69, 094425 (2004).
[413] A.R. Rocha and S. Sanvito, Asymmetric I-V characteristics and magne-
toresistance in magnetic point contacts, Phys. Rev. B 70, 094406 (2004).
[414] A. Bagrets, N. Papanikolaou, I. Mertig, Magnetoresistance of atomic-sized
contacts: An ab initio study, Phys. Rev. B 70, 064410 (2004).
[415] D. Jacob, J. Fernández-Rossier, J.J. Palacios, Magnetic and orbital blocking
in Ni nanocontacts, Phys. Rev. B 71, 220403 (2005).
[416] M. Wierzbowska, A. Delin, E. Tosatti, Effect of electron correlations in Pd,
Ni, Co monowires, Phys. Rev. B 72, 035439 (2005).
[417] H. Dalgleish and G. Kirczenow, Theoretical study of spin-dependent electron
transport in atomic Fe nanocontacts, Phys. Rev. B 72, 155429 (2005).
[418] P.A. Khomyakov, G. Brocks, V. Karpan, M. Zwierzycki, P.J. Kelly, Con-
ductance calculations for quantum wires and interfaces: Mode matching and
Green’s functions, Phys. Rev. B 72, 35450 (2005).
[419] J. Fernández-Rossier, D. Jacob, C. Untiedt, J.J. Palacios, Magnetic and
orbital blocking in Ni nanocontacts, Rev. B 72, 224418 (2005).
[420] G. Autès, C. Barreteau, D. Spanjaard, M.C. Desjonquères, Magnetism of
iron: from the bulk to the monatomic wire, J. Phys. Condens. Matter 18,
6785 (2006).
[421] A. Smogunov, A. Dal Corso, E. Tosatti, Ballistic conductance and mag-
netism in short tip suspended Ni nanowires, Phys. Rev. B 73, 75418 (2006).
[422] D. Jacob and J.J. Palacios, Orbital eigenchannel analysis for ab initio quan-
tum transport calculations, Phys. Rev. B 73, 075429 (2006).
[423] K. Xia, M. Zwierzycki, M. Talanana, P.J. Kelly, G.E.W. Bauer, First-
principles scattering matrices for spin transport, Rev. B 73, 64420 (2006).
[424] A.R. Rocha, T. Archer, S. Sanvito, Search for magnetoresistance in excess
of 1000% in Ni point contacts: Density functional calculations, Phys. Rev. B
76, 054435 (2007).
[425] J.C. Tung and G.Y. Guo, Systematic ab initio study of the magnetic and
electronic properties of all 3d transition metal linear and zigzag nanowires,
Phys. Rev. B 76, 094413 (2007).
[426] A. Bagrets, N. Papanikolaou, I. Mertig, Conduction eigenchannels of
atomic-sized contacts: Ab initio KKR Green’s function formalism, Phys. Rev.
B 75, 235448 (2007).
[427] M. Häfner, J.K. Viljas, D. Frustaglia, F. Pauly, M. Dreher, P. Nielaba,
J.C. Cuevas, Theoretical study of the conductance of ferromagnetic atomic-
sized contacts, Phys. Rev. B 77, 104409 (2008).
[428] M.J. Mehl, D.A. Papaconstantopoulos, I.I. Mazin, N.C. Bacalis, W.E. Pick-
ett, Applications of the NRL tight-binding method to magnetic systems, Vol.
90, p. 6680, (American Institute of Physics, USA, 2001).
[429] D.R. Lide, CRC Handbook of Chemistry and Physics, 79th Edition, (CRC
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
Bibliography 663
Bibliography 665
Bibliography 667
Bibliography 669
Bibliography 671
molecules with different contact geometries on gold, Phys. Stat. Sol. (b) 243,
2021 (2006).
[563] K. H. Müller, Effect of the atomic configuration of gold electrodes on the
electrical conduction of alkanedithiol molecules, Phys. Rev. B 73, 045403
(2006).
[564] F. Picaud, A. Smogunov, A. Dal Corso, E.J. Tosatti, Complex band struc-
tures and decay length in polyethylene chains, J. Phys.: Condens. Matter 15,
3731 (2003).
[565] G. Fagas, P. Delaney, J.C. Greer, Independent particle descriptions of tun-
neling using the many-body quantum transport approach, Phys. Rev. B 73,
241314 (2006).
[566] M. Paulsson, C. Krag, T. Frederiksen, M. Brandbyge, Conductance of alka-
nedithiol single-molecule junctions: A molecular dynamics study, Nano Lett.
9, 117 (2009).
[567] D. Djukic, K.S. Thygesen, C. Untiedt, R.H.M. Smit, K.W. Jacobsen, J.M.
van Ruitenbeek, Stretching dependence of the vibration modes of a single-
molecule Pt-H2-Pt bridge, Phys. Rev. B 71, 161402 (2005).
[568] D. Djukic and J.M. Ruitenbeek, Shot noise measurement on a single
molecule, Nano Lett. 6, 789 (2006).
[569] J.C. Cuevas, J. Heurich, F. Pauly, W. Wenzel, G. Schön, Theoretical de-
scription of the electrical conduction in atomic and molecular junctions, Nan-
otechnology 14, R29 (2003).
[570] Y. Garcı́a, J.J. Palacios, E. SanFabián, J.A. Vergés, A.J. Pérez-Jiménez, E.
Louis, Electronic transport and vibrational modes in a small molecular bridge:
H2 in Pt nanocontacts, Phys. Rev. B 69, 041402 (2004).
[571] K.S. Thygesen, K.W. Jacobsen, Conduction mechanism in a molecular hy-
drogen contact, Phys. Rev. Lett. 94, 036807 (2005).
[572] V.M. Garcı́a-Suárez, A.R. Rocha, S.W. Bailey, C. Lambert, S. Sanvito, J.
Ferrer, Single-channel conductance of H2 molecules attached to platinum or
palladium electrodes, Phys. Rev. B 72, 045437 (2005).
[573] S. Csonka, A. Halbritter, G. Mihály, O.I. Shklyarevskii, S. Speller, H. van
Kempen, Conductance of Pd-H nanojunctions, Phys. Rev. Lett. 93, 016802
(2004).
[574] D.P. Long, J.L. Lazorcik, B.A. Mamtooth, M.H. Moore, M.A. Ratner, A.
Troisi, Y. Yao, J.W. Ciszek, J.M. Tour, R. Shashidhar, Effects of hydration
on molecular junction transport, Nature Mat. 5, 901 (2006).
[575] R. Ahlrichs, M. Bär, M. Häser, H. Horn, C. Kölmel, Electronic structure cal-
culations on workstation computers: The program system turbomole, Chem.
Phys. Lett. 162, 165 (1989).
[576] F. Pauly, J.K. Viljas, U. Huniar, M. Häfner, S. Wohlthat, M. Burkle, J.C.
Cuevas, Gerd Schön, Cluster-based density-functional approach to quantum
transport through molecular and atomic contacts, New J. Phys. 10, 125019
(2008).
[577] A. Ulman, Formation and structure of self-assembled monolayers, Chem.
Rev. 96, 1533 (1996).
[578] S.Y. Quek, L. Venkataraman, H.J. Choi, S.G. Louie, M.S. Hybertsen, J.B.
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
Bibliography 673
ASI Series E, Vol. 345, p. 105, (Kluwer Academic Publishers, Dordrecht, NL,
1997).
[609] D.V. Averin and K.K. Likharev, in B.L. Al’tshuler, P.A. Lee, R.A. Webb
(Eds.), Mesoscopic Phenomena in Solids, (Elsevier, Amsterdam, NL, 1991).
[610] H. Grabert and M.H. Devoret (Eds.), Single Charge Tunneling: Coulomb
Blockade Phenomena in Nanostructures, NATO ASI Series B, Vol. 294,
(Plenum Press, New York, USA, 1992).
[611] J.P. Bird, Electron Transport in Quantum Dots, (Kluwer Academic Pub-
lishers, Dordrecht, NL, 2003).
[612] D.C. Ralph, C.T. Black, J.M. Hergenrother, J.G. Lu, M. Tinkham, in L.
L. Sohn, L. P. Kouwenhoven, G. Schön (Eds.), Mesoscopic Electron Trans-
port, NATO-ASI Series E, Vol. 345, p. 453, (Kluwer Academic Publishers,
Dordrecht, NL, 1997).
[613] S.J. Tans, M.H. Devoret, H. Dai, A. Thess, R.E. Smalley, L.J. Geerligs,
C. Dekker, Individual single-wall carbon nanotubes as quantum wires, Nature
386, 474 (1997).
[614] P.L. McEuen, M. Fuhrer, H. Park, Single-walled carbon nanotube electron-
ics, IEEE Trans. on Nanotechnol. 1, 78 (2002).
[615] M.T. Björk, B.J. Ohlsson, T. Sass, A.I. Persson, C. Thelander, M.H.
Magnusson, K. Deppert, L. R. Wallenberg, L. Samuelson, One-dimensional
steeplechase for electrons realized, Nano Lett. 2, 87 (2002).
[616] M.T. Björk, C. Thelander, A.E. Hansen, L.E. Jensen, M.W. Larsson, L.R.
Wallenberg, L. Samuelson, Few-electron quantum dots in nanowires, Nano
Lett. 4, 1621 (2004).
[617] L.H. Yu, Z.K. Keane, J.W. Ciszek, L. Cheng, M.P. Stewart, J.M. Tour,
D. Natelson, Inelastic electron tunneling via molecular vibrations in single-
molecule transistors, Phys. Rev. Lett. 93, 266802 (2004).
[618] E.A. Osorio, K. O’Neill, M. Wegewijs, N. Stuhr-Hansen, J. Paaske, T.
Bjørnholm, H.S.J. van der Zant, Electronic excitations of a single molecule
contacted in a three-terminal configuration, Nano Lett. 7, 3336 (2007).
[619] H.B. Heersche, Z. de Groot, J.A. Folk, H.S.J. van der Zant, Electron trans-
port through single Mn12 molecular magnets, Phys. Rev. Lett. 96, 206801
(2006).
[620] M.H. Jo, J.E. Grose, K. Baheti, M.M. Deshmukh, J.J. Sokol, E.M. Rum-
berger, D.N. Hendrickson, J.R. Long, H. Park and D.C. Ralph, Signatures of
molecular magnetism in single-molecule transport spectroscopy, Nano Lett. 6,
2014 (2006).
[621] T. Taychatanapat, K.I. Bolotin, F. Kuemmeth, D.C. Ralph, Imaging elec-
tromigration during the formation of break-junctions, Nano Lett. 7, 652
(2007).
[622] S. Kubatkin, A. Danilov, M. Hjort, J. Cornil, J.-L. Bredas, N. Stuhr-
Hansen, P. Hedegard, T. Bjørnholm, Single-electron transistor of a single
organic molecule with access to several redox states, Nature 425, 698 (2003).
[623] S. Kubatkin, A. Danilov, M. Hjort, J. Cornil, J.-L. Bredas, N. Stuhr-
Hansen, P. Hedegard, T. Bjørnholm, Single electron transistor with a single
conjugated molecule, Curr. Appl. Phys. 4, 554 (2004).
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
Bibliography 675
[642] J. Lehmann and D. Loss, Sequential tunneling through molecular spin rings,
Phys. Rev. Lett. 98, 117203 (2007).
[643] R. Stadler, V. Geskin, J. Cornil, Towards a theoretical description of molec-
ular junctions in the Coulomb blockade regime based on density functional
theory, Phys. Rev. B 78, 113402 (2008).
[644] R. Stadler, V. Geskin, J. Cornil, Screening effects in a density func-
tional theory based description of molecular junctions in the Coulomb blockade
regime, Phys. Rev. B 79, 113408 (2009).
[645] L. Michalak, C.M. Canali, M.R. Pederson, V.G. Benza, M. Paulsson, The-
ory of tunneling spectroscopy in a Mn12 single-electron transistor by DFT
methods, arXiv:0812.1058.
[646] K. Kaasbjerg and K. Flensberg, Strong polarization-induced reduction of
addition energies in single-molecule nanojunctions, Nano Lett. 8, 3809 (2008).
[647] J. König, H. Schoeller, G. Schön, Zero-bias anomalies and boson-assisted
tunneling through quantum dots, Phys. Rev. Lett. 76, 1715 (1996).
[648] J. König, J. Schmid, H. Schoeller, G. Schön, Resonant tunneling through
ultrasmall quantum dots: Zero-bias anomalies, magnetic-field dependence,
boson-assisted transport, Phys. Rev. B 54, 16820 (1996).
[649] D.V. Averin and Y.V. Nazarov, Macroscopic quantum tunneling of charge
and co-tunneling in Single Charge Tunneling: Coulomb Blockade Phenomena
in Nanostructures, H. Grabert and M.H. Devoret (Eds.), NATO-ASI Series
B, Vol. 294, p. 217 (Plenum Press, New York, USA, 1992).
[650] S. de Franceschi, S. Sasaki, J.M. Elzerman, W.G. van der Wiel, S. Tarucha,
L.P. Kouwenhoven, Electron cotunneling in a semiconductor quantum dot,
Phys. Rev. Lett. 86, 878 (2001).
[651] A. C. Hewson, The Kondo Problem to Heavy Fermions, (Cambridge Uni-
versity Press, Cambridge, UK, 1993).
[652] J. Kondo, Resistance minimum in dilute magnetic alloys, Prog. Theor.
Phys. 32, 37 (1964).
[653] L. Kouwenhoven and L. Glazman, Revival of the Kondo effect, Physics
World 14, 33 (2001).
[654] D. Goldhaber-Gordon, H. Shtrikman, D. Mahalu, D. Abusch-Magder, U.
Meirav, M.A. Kastner, Kondo effect in a single-electron transistor, Nature
391, 156 (1998).
[655] D. Goldhaber-Gordon, J. Göres, M.A. Kastner, H. Shtrikman, D. Mahalu,
U. Meirav, From the Kondo regime to the mixed-valence regime in a single-
electron transistor, Phys. Rev. Lett. 81, 5225 (1998).
[656] S.M. Cronenwett, T.H. Oosterkamp, L.P. Kouwenhoven, A tunable Kondo
effect in quantum dots, Science 281, 540 (1998).
[657] J. Li, W.-D Schneider, R. Berndt, B. Delley, Kondo scattering observed at
a single magnetic impurity, Phys. Rev. Lett. 80, 2893 (1998).
[658] V. Madhavan, W. Chen, T. Jamneala, M.F. Crommie, N.S. Wingreen,
Tunneling into a single magnetic atom: Spectroscopic evidence of the Kondo
resonance, Science 280, 567 (1998).
[659] H.C. Manoharan, C.P. Lutz and D.M. Eigler, Quantum mirages: The co-
herent projection of electronic structure, Nature 403, 512 (2000).
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
Bibliography 677
[660] J. Nygard, D.H. Cobden and P.E. Lindelof, Kondo physics in carbon nan-
otubes, Nature 408, 342 (2000).
[661] M.R. Buitelaar, A. Bachtold, T. Nussbaumer, M. Iqbal and C.
Schönenberger, Multiwall carbon nanotubes as quantum dots, Phys. Rev. Lett.
88, 156801 (2002).
[662] L.H. Yu and D. Natelson, The Kondo effect in C60 single-molecule transis-
tors, Nano Lett. 4, 79 (2004).
[663] A.N. Pasupathy, R.C. Bialczak, J. Martinek, J.E. Grose, L.A.K. Donev,
P.L. McEuen, D.C. Ralph, The Kondo effect in the presence of ferromag-
netism, Science 306, 86 (2004).
[664] L.H. Yu, Z.K. Keane, J.W. Ciszek, L. Cheng, J.M. Tour, T. Baruah, M.R.
Pederson, D. Natelson, Kondo resonances and anomalous gate dependence in
the electrical conductivity of single-molecule transistors, Phys. Rev. Lett. 95,
256803 (2005).
[665] M. Pustilnik, L.I. Glazman, D.H. Cobden, L.P. Kouwenhoven, Magnetic
field-induced Kondo effects in Coulomb blockade systems, Lect. Notes Phys.
579, 3 (2001); Preprint cond-mat/0010336.
[666] M. Pustilnik and L. Glazman, Kondo effect in quantum dots, J. Phys.:
Condens. Matter 16, R513 (2004).
[667] M. Grobis, I.G. Rau, R.M. Potok, D. Goldhaber-Gordon, Kondo effect in
mesoscopic quantum dots in H. Kronmuller (Ed.), Handbook of Magnetism
and Advanced Magnetic Materials, Vol. 5, (Wiley, New York, USA, 2007).
[668] F.D.M. Haldane, Scaling theory of the asymmetric Anderson model, Phys.
Rev. Lett. 40, 416 (1978).
[669] L.I. Glazman and M.E. Raikh, Resonant Kondo transparency of a barrier
with quasilocal impurity states, JETP Lett. 47, 452 (1988).
[670] T.A. Costi, A.C. Hewson and V. Zlatic, Transport coefficients of the An-
derson model via the numerical renormalization group, J. Phys.: Condens.
Matter 6, 2519 (1994).
[671] W.G. van der Wiel, S. De Franceschi, T. Fujisawa, J.M. Elzerman, S.
Tarucha, L.P. Kouwenhoven, The Kondo effect in the unitary limit, Science
289, 2105 (2000).
[672] P. Jarillo-Herrero, J. Kong, H.S.J. van der Zant, C. Dekker, L.P. Kouwen-
hoven, S. De Franceschi, Orbital Kondo effect in carbon nanotubes, Nature
434, 484 (2005).
[673] S. Sasaki, S. De Franceschi, J.M. Elzerman, W.G. van der Wiel, M. Eto,
S. Tarucha and L.P. Kouwenhoven, Kondo effect in an integer-spin quantum
dot, Nature 405, 764 (2000).
[674] M. Pustilnik, Y. Avishai and K. Kikoin Quantum dots with even number of
electrons: Kondo effect in a finite magnetic field, Phys. Rev. Lett. 84, 1756
(2000).
[675] M. Eto and Y.V. Nazarov, Enhancement of Kondo effect in quantum dots
with an even number of electrons, Phys. Rev. Lett. 85, 1306 (2000).
[676] C. Kergueris, J.-P. Bourgoin, S. Palacin, D. Esteve, C. Urbina, M. Ma-
goga, C. Joachim, Electron transport through a metal-molecule-metal junction,
Phys. Rev. B 59, 12505 (1999).
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
Bibliography 679
(2004).
[714] L.H. Yu, C.D. Zangmeister, J.G. Kushmerick, Origin of discrepancies in
inelastic electron tunneling spectra of molecular junctions, Phys. Rev. Lett.
98, 206803 (2007).
[715] L.T. Cai, M.A. Cabassi, H. Yoon, O.M. Cabarcos, C.L. McGuiness, A.K.
Flatt, D.L. Allara, J.M. Tour, A.S. Mayer, Reversible bistable switching in
nanoscale thiol-substituted oligoaniline molecular junctions, Nano Lett. 5,
2365 (2005).
[716] L.H. Yu, C.D. Zangmeister, J.G. Kushmerick, Structural contributions to
charge transport across Ni-octanedithiol multilayer junctions, Nano Lett. 6,
2515 (2006).
[717] W. Wang and C.A. Richter, Spin-polarized inelastic electron tunneling
spectroscopy of a molecular magnetic tunnel junction, Appl. Phys. Lett. 89,
153105 (2006).
[718] A. Troisi, J.M. Beebe, L.B. Picraux, R.D. van Zee, D.R. Stewart, M.A.
Ratner, J.G. Kushmerick, Tracing electronic pathways in molecules by us-
ing inelastic tunneling spectroscopy, Proc. Natl. Acad. Sci. USA 104, 14255
(2007).
[719] J. M. Beebe, H.J. Moore, T.R. Lee, J.G. Kushmerick, Vibronic coupling
in semifluorinated alkanethiol junctions: Implications for selection rules in
inelastic electron tunneling spectroscopy, Nano Lett. 7, 1364 (2007).
[720] D.P. Long and A. Troisi, Inelastic electron tunneling spectroscopy of alkane
monolayers with dissimilar attachment chemistry to gold, J. Am. Chem. Soc.
129, 15303 (2007).
[721] W. Wang, A. Scott, N. Gergel-Hackett, C.A. Hacker, D.B. Janes, C.A.
Richter, Probing molecules in integrated silicon-molecule-metal junctions by
inelastic tunneling spectroscopy, Nano Lett. 8, 478 (2008).
[722] J. Hihath, C.R. Arroyo, G. Rubio-Bollinger, N.J. Tao, N. Agraı̈t, Study of
electron-phonon interactions in a single molecule covalently connected to two
electrodes, Nano Lett. 8, 1673 (2008).
[723] H. Song, Y. Kim, J. Ku, Y.H. Jang, H. Jeong, T. Lee, Vibrational spectra
of metal-molecule-metal junctions in electromigrated nanogap electrodes by
inelastic electron tunneling, Appl. Phys. Lett. 94, 103110 (2009).
[724] I.K. Yanson, O.I. Shklyarevskii, Point-contact spectroscopy of metallic al-
loys and compounds, Sov. J. Low Temp. Phys. 12, 509 (1986).
[725] A.M. Duif, A.G.M. Jansen and P. Wyder, Point-contact spectroscopy, J.
Phys.: Condens. Matter 1, 3157 (1989).
[726] Yu. G. Naidyuk, I.K. Yanson, Point-Contact Spectroscopy, Springer Se-
ries in Solid-State Sciences, Vol. 145 (Springer-Verlag, Heidleberg, Berlin, D,
2004).
[727] I.K. Yanson, Nonlinear effects in the electric conductivity of point junc-
tions and electron-phonon interaction in metals, Zh. Eksp. Teor. Fiz. 66,
1035 (1974) [Sov. Phys. JETP 39, 506 (1974)].
[728] A.G.M. Jansen, A.P. van Gelder, P. Wyder, Point contact spectroscopy in
metals, J. Phys. C 13, 6073 (1980).
[729] A.V. Khotkevich, I.K. Yanson, Atlas of Point Contact Spectra of Electron-
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
Bibliography 681
Bibliography 683
Bibliography 685
Bibliography 687
(2001).
[858] A.Y. Kasumov, D.V. Klinov, P.-E. Roche, S. Guéron, H. Bouchiat, Thick-
ness and low-temperature conductivity of DNA molecules, Appl. Phys. Lett.
84, 1007 (2004).
[859] T. Heim, D. Deresmes, D. Vuillaume, Conductivity of DNA probed by
conducting-atomic force microscopy: Effects of contact electrode, DNA struc-
ture, surface interactions, J. Appl. Phys. 96, 2927 (2004).
[860] H.X. He, X.L. Li, N.J. Tao, Discrete conductance switching in conducting
polymer wires, Phys. Rev. B 68, 45302 (2003).
[861] J.F. Leger, G. Romano, A. Sarkar, J. Robert, L. Bourdieu, D. Chatenay,
J.F. Marko, Structural transitions of a twisted and stretched DNA molecule,
Phys. Rev. Lett. 83, 1066 (1999).
[862] B. Xu, P. Zhang, X. Li, N. Tao, Direct conductance measurement of single
DNA molecules in aqueous solution, Nano Lett. 4, 1105 (2004).
[863] J. Jortner, M. Bixon, T. Langenbacher, M.E. Michel-Beyerle, Charge trans-
fer and transport in DNA, Proc. Nat. Acad. Sci. USA 95, 12759 (1998).
[864] R. Roche, Sequence dependent DNA-mediated conduction, Phys. Rev. Lett.
91, 108101 (2003).
[865] D.J. Patel, S. Bouaziz, A. Kettani, Y. Wong, Structures of guanine-rich and
cytosine-rich quadruplexes formed in vitro by telomeric, centromeric, triplet
repeat disease DNA sequences, in S. Neidle (Ed.), Oxford Handbook of Nucleic
Acid Structure, p. 389, (Oxford University Press, Oxford, UK, 1999).
[866] H. Cohen, T. Sapir, N. Borovok, T. Molotsky, R. di Felice, A. Kotlyar, D.
Porath, Polarizability of G4-DNA observed by electrostatic force microscopy
measurements, Nano Lett. 7, 981 (2007).
[867] E. Shapir, H. Cohen, A. Calzolari, C. Cavazzoni, D.A. Ryndyk, G. Cu-
niberti, A. Kotlyar, R. Di Felica, D. Porath, Electronic structure of single
DNA molecules resolved by transverse scanning tunnelling spectroscopy, Na-
ture Mat. 7, 68 (2008).
[868] H. Tanaka and T. Kawai, Partial sequencing of a single DNA molecule with
a scanning tunnelling microscope, Nature Nanotech. 4, 518 (2009).
[869] S.S. Mallajosyula, J.C. Lin, D.L. Cox, S.K. Pati, R.R.P. Singh, Sequence
dependent electron transport in wet DNA: Ab initio and molecular dynamics,
Phys. Rev. Lett. 101, 176805 (2008).
[870] B. Song, M. Elstner, G. Cuniberti, Anomalous conductance response of
DNA wires under stretching, Nano Lett. 8, 3217 (2008).
[871] M.L. Bennink, O.D. Schärer, R. Kanaar, K. Sakata-Sogawa, J.M. Schins,
J.S. Kanger, B.G. de Grooth, J. Greve, Single-molecule manipulation of
double-stranded DNA using optical tweezers: Interaction studies of DNA with
RecA and YOYO-1, Cytometry 36, 200 (1999).
[872] P. Cluzel, A. Lebrun, C. Heller, R. Lavery, J.-L. Viovy, D. Chatenay, F.
Caron, DNA: An extensible molecule, Science 271, 792 (1996).
[873] S.B. Smith, Y. Cui, C. Bustamante, Overstretching B-DNA: The elastic
response of individual double-stranded and single-stranded DNA molecules,
Science 271, 795 (1996).
[874] P. Maragakis, R. L. Barnett, E. Kaxiras, M. Elstner, T. Frauenheim, Elec-
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
Bibliography 689
Bibliography 691
[914] B. Li, L. Wang, G. Casati, Thermal diode: rectification of heat flux, Phys.
Rev. Lett. 93, 184301 (2004).
[915] D. Segal and A. Nitzan, Spin-boson thermal rectifier, Phys. Rev. Lett. 94,
034301 (2005).
[916] D. Segal and A. Nitzan, Heat rectification in molecular junctions, J. Chem.
Phys. 122, 194704 (2005).
[917] B. Li, J. Lan, L. Wang, Interface thermal resistance between dissimilar
anharmonic lattices, Phys. Rev. Lett. 95, 104302 (2005).
[918] K. Saito, Asymmetric heat flow in mesoscopic magnetic system, J. Phys.
Soc. Japan 75, 034603 (2006).
[919] D. Cahill, K. Goodson, A. Majumdar, Thermometry and thermal transport
in micro/nanoscale solid-state devices and structures, J. Heat Transfer 124,
223 (2002).
[920] L. Shi and A. Majumdar, Thermal transport mechanisms at nanoscale point
contacts, J. Heat Transfer 124, 329 (2002).
[921] C. Yu, L. Shi, Z. Yao, D. Li, A. Majumdar, Thermal conductance and
thermopower of an individual single-wall carbon nanotube, Nano Lett. 5, 1842
(2005).
[922] N. Mingo and D.A. Broido, Carbon nanotube ballistic thermal conductance
and its limits, Phys. Rev. Lett. 95, 096105 (2005).
[923] R.Y. Wang, R.A. Segalman, A. Majumdar, Room temperature thermal con-
ductance of alkanedithiol self-assembled monolayers, Appl. Phys. Lett. 89,
173113 (2006).
[924] Z. Wang, J.A. Carter, A. Lagutchev, Y.K. Koh, N.-H. Seong, D.G. Cahill,
D.D. Dlott, Ultrafast flash thermal conductance of molecular chains, Science
317, 787 (2007).
[925] H.E. van den Brom, A.I. Yanson, J.M. van Ruitenbeek, Characterization
of individual conductance steps in metallic quantum point contacts, Physica
B 252, 69 (1998).
[926] M. Tsutsui, S. Kurokawa, A. Sakai, Bias-induced local heating in atom-sized
metal contacts at 77 K, Appl. Phys. Lett. 90, 133121 (2007).
[927] R.H.M. Smit, C. Untiedt, J.M. van Ruitenbeek, The high-bias stability of
monatomic chains, Nanotechnology 15, S472 (2004).
[928] Z. Huang, B.Q. Xu, Y.C. Chen, M. Di Ventra, N.J. Tao, Measurement of
current-induced local heating in a single molecule junction, Nano Lett. 6, 1240
(2006).
[929] Z. Huang, F. Chen, R. D’Agosta, P.A. Bennett, M. Di Ventra, N.J. Tao, Lo-
cal ionic and electron heating in single-molecule junctions, Nature Nanotech.
2, 698 (2007).
[930] T.N. Todorov, Local heating in ballistic atomic-scale contracts, Phil. Mag.
B 77, 965 (1998).
[931] M.J. Montgomery, T.N. Todorov, A.P. Sutton, Power dissipation in
nanoscale conductors, J. Phys.: Condens. Matter 14, 5377 (2002).
[932] Y.C. Chen, M. Zwolak, M. Di Ventra, Local heating in nanoscale conduc-
tors, Nano Lett. 3, 1691 (2003).
[933] R. D’Agosta, N. Sai, M. Di Ventra, Local electron heating in nanoscale
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
Bibliography 693
Hou, Q.S. Zhu, What can a scanning tunneling microscope image do for the
insulating alkanethiol molecules on Au(111) substrates?, J. Chem. Phys. 117,
851 (2002).
[952] M. Galperin and A. Nitzan, Optical properties of current carrying molecular
wires, J. Chem. Phys. 124, 234709 (2006).
[953] A.M. Nowak and R.L. McCreery, In situ Raman spectroscopy of bias-
induced structural changes in nitroazobenzene molecular electronic junctions,
J. Am. Chem. Soc. 126, 16621 (2004).
[954] J. Tian, B. Liu, X.L. Li, Z.L. Yang, B. Ren, S.T. Wu, N.J. Tao and Z.Q.
Tian, Study of molecular junctions with combined surface-enhanced Raman
and mechanically controllable break junction method, J. Am. Chem. Soc. 128,
14748 (2006).
[955] D.R. Ward, N.K. Grady, C.S. Levin, N.J. Halas, Y. Wu, P. Nordlander, D.
Natelson, Electromigrated nanoscale gaps for surface-enhanced Raman spec-
troscopy, Nano Lett. 7, 1396 (2007).
[956] D.R. Ward, N.J. Halas, J.W. Ciszek, J.M. Tour, Y. Wu, P. Nordlander, D.
Natelson, Simultaneous measurements of electronic conduction and Raman
response in molecular junctions, Nano Lett. 8, 919 (2008).
[957] P. Mühlschlegel, H.-J. Eisler, O.J.F. Martin, B. Hecht, D.W. Pohl, Reso-
nant optical antennas, Science 308, 1607 (2005).
[958] M. Moskovits, Surface-enhanced Raman spectroscopy, J. Raman Spectrosc.
36, 485 (2005).
[959] M. Galperin, M.A. Ratner, A. Nitzan, Raman scattering from nonequilib-
rium molecular conduction junctions, Nano Lett. 9, 758 (2009).
[960] M. Galperin, M.A. Ratner, A. Nitzan, Raman scattering in current-carrying
molecular junctions, J. Chem. Phys. 130, 144109 (2009).
[961] B.N. Persson and A. Baratoff, Theory of photon emission in electron tun-
neling to metallic particles, Phys. Rev. Lett. 68, 3224 (1992).
[962] M. Galperin, A. Nitzan, M.A. Ratner, Molecular transport junctions: Cur-
rent from electronic excitations in the leads, Phys. Rev. Lett. 96, 166803
(2006).
[963] A.H. Dayem and R.J. Martin, Quantum interaction of microwave radiation
with tunneling between superconductors, Phys. Rev. Lett. 8, 246 (1962).
[964] P.H. Cutler, T.E. Feuchtwang, T.T. Tsong, H. Nguyen, A.A. Lucas, Pro-
posed use of a scanning-tunneling-microscope tunnel junction for the mea-
surement of a tunneling time, Phys. Rev. B 35, 7774 (1987).
[965] M. Völcker, W. Krieger, H. Walther, Laser-driven scanning tunneling mi-
croscope, Phys. Rev. Lett. 66, 1717 (1991).
[966] M. Chauvin, P. vom Stein, H. Pothier, P. Joyez, M.E. Huber, D. Esteve, C.
Urbina, Superconducting atomic contacts under microwave irradiation, Phys.
Rev. Lett. 97, 067006 (2006).
[967] J.C. Cuevas, J. Heurich, A. Martin-Rodero, A. Levy Yeyati, G. Schön,
Subharmonic Shapiro steps and assisted tunneling in superconducting point
contacts, Phys. Rev. Lett. 88, 157001 (2002).
[968] A. Tikhonov, R.D. Coalson, Y. Dahnovsky, Calculating electron transport
in a tight binding model of a field-driven molecular wire: Floquet theory ap-
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
Bibliography 695
412, 55 (2005).
[1007] J. He, F. Chen, P.A. Liddell, J. Andreasson, S.D. Straight, D. Gust, T.A.
Moore, A.L. Moore, J. Li, O.F. Sankey and S.M. Lindsay, Switching of a pho-
tochromic molecule on gold electrodes: Single-molecule measurements, Nan-
otechnology 16, 695 (2005).
[1008] N. Katsonis, T. Kudernac, M. Walko, S.J. van der Molen, B.J. van Wees,
B.L. Feringa, Reversible switching of photochromic molecules self-assembled
on gold, Adv. Mater. 18, 1397 (2006).
[1009] A.C. Whalley, M.L. Steigerwald, X. Guo, C. Nuckolls, Reversible switching
in molecular electronic devices, J. Am. Chem. Soc. 129, 12590 (2007).
[1010] A.J. Kronemeijer, H.B. Akkerman, T. Kudernac, B.J. van Wees, B.L.
Feringa, P.W.M. Blom, B. de Boer, Reversible conductance switching in
molecular devices, Adv. Mater. 20, 1467 (2008).
[1011] M. Del Valle, R. Gutiérrez, C. Tejedor, G. Cuniberti, Tuning the conduc-
tance of a molecular switch, Nat. Nanotechnol. 2, 176 (2007).
[1012] M. Alemani, M.V. Peters, S. Hecht, K.-H. Rieder, F. Moresco, L. Grill,
Electric field-induced isomerization of azobenzene by STM, J. Am. Chem. Soc.
128, 14446 (2006).
[1013] M.J. Comstock, N. Levy, A. Kirakosian, J. Cho, F. Lauterwasser, J.H.
Harvey, D.A. Strubbe, J.M.J. Fréchet, D. Trauner, S.G. Louie, M.F. Crom-
mie, Reversible photochemical switching of individual engineered molecules at
a metallic surface, Phys. Rev. Lett. 99, 038301 (2007).
[1014] A.S. Kumar, T. Ye, T. Takami, B.-C. Yu, A.K. Flatt, J.M. Tour, P.S.
Weiss, Reversible photo-switching of single azobenzene molecules in controlled
nanoscale environments, Nano Lett. 6, 1644 (2008).
[1015] J.M. Mativetsky, G. Pace, M. Elbing, M.A. Rampi, M. Mayor, P. Samori,
Azobenzes as light-controlled molecular electronic switches in nanoscale metal-
molecule-metal junctions, J. Am. Chem. Soc. 130, 9192 (2008).
[1016] J.M. Tour, Molecular electronics. Synthesis and testing of components,
Acc. Chem. Res. 33, 791 (2000).
[1017] N. Weibel, S. Grunder, M. Mayor, Functional molecules in electronic cir-
cuits, Org. Biomol. Chem. 5, 2343 (2007).
[1018] A. Nitzan, A relationship between electron transfer rates and molecular
conduction, J. Phys. Chem. A 105, 2677 (2001).
[1019] W. Lu and C.M. Lieber, Nanoelectronics from the bottom up, Nature Mat.
6, 841 (2007).
[1020] J.-C. Charlier, X. Blase, S. Roche, Electronic and transport properties of
nanotubes, Rev. Mod. Phys. 79, 677 (2007).
[1021] P. Darancet, A. Ferretti, D. Mayou, V. Olevano, Ab initio GW electron-
electron interaction effects in quantum transport, Phys. Rev. B 75, 075102
(2007).
[1022] K.S. Thygesen and A. Rubio, Conserving GW scheme for nonequilibrium
quantum transport in molecular contacts, Phys. Rev. B 77, 115333 (2008).
[1023] K.S. Thygesen, Impact of exchange-correlation effects on the I-V charac-
teristics of a molecular junction, Phys. Rev. Lett. 100, 166804 (2008).
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
Index
697
January 12, 2010 11:27 World Scientific Book - 9in x 6in book
Bibliography 699
Bibliography 701