You are on page 1of 21

Coulomb blockade

Page issues

Schematic representation (similar to band diagram) of


an electron tunnelling through a barrier

In physics, a Coulomb blockade (CB),


named after Charles-Augustin de
Coulomb's electrical force, is the increased
resistance at small bias voltages of an
electronic device comprising at least one
low-capacitance tunnel junction. Because
of the CB, the resistances of devices are
not constant at low bias voltages, but
increase to infinity for biases under a
certain threshold (i.e. no current flows).
When few electrons are involved and an
external static magnetic field is applied,
Coulomb blockade provides the ground for
spin blockade (also called Pauli blockade)
and valley blockade[1][2] which includes
quantum mechanical effects due to spin
and orbital interactions respectively
between the electrons.
A Coulomb blockade will also be observed
when making the device very small (like a
quantum dot). When the device is small
enough, electrons inside the device will
create a strong coulomb repulsion
preventing other electrons to flow. The
device will no longer follow Ohm's law. The
current-voltage relation of the coulomb
blockade looks like a staircase.[3]

In a tunnel junction
The tunnel junction is, in its simplest form,
a thin insulating barrier between two
conducting electrodes. If the electrodes
are superconducting, Cooper pairs (with a
charge of two elementary charges) carry
the current. In the case that the electrodes
are normalconducting, i.e. neither
superconducting nor semiconducting,
electrons (with a charge of one elementary
charge) carry the current. The following
reasoning is for the case of tunnel
junctions with an insulating barrier
between two normal conducting
electrodes (NIN junctions).

According to the laws of classical


electrodynamics, no current can flow
through an insulating barrier. According to
the laws of quantum mechanics, however,
there is a nonvanishing (larger than zero)
probability for an electron on one side of
the barrier to reach the other side (see
quantum tunnelling). When a bias voltage
is applied, this means that there will be a
current, and, neglecting additional effects,
the tunnelling current will be proportional
to the bias voltage. In electrical terms, the
tunnel junction behaves as a resistor with
a constant resistance, also known as an
ohmic resistor. The resistance depends
exponentially on the barrier thickness.
Typical barrier thicknesses are on the
order of one to several nanometers.

An arrangement of two conductors with an


insulating layer in between not only has a
resistance, but also a finite capacitance.
The insulator is also called dielectric in
this context, the tunnel junction behaves
as a capacitor.

Due to the discreteness of electrical


charge, current through a tunnel junction is
a series of events in which exactly one
electron passes (tunnels) through the
tunnel barrier (we neglect cotunneling, in
which two electrons tunnel
simultaneously). The tunnel junction
capacitor is charged with one elementary
charge by the tunnelling electron, causing
a voltage buildup , where is
the elementary charge of
1.6×10−19 coulomb and the capacitance
of the junction. If the capacitance is very
small, the voltage buildup can be large
enough to prevent another electron from
tunnelling. The electric current is then
suppressed at low bias voltages and the
resistance of the device is no longer
constant. The increase of the differential
resistance around zero bias is called the
Coulomb blockade.

Observing the Coulomb


blockade
In order for the Coulomb blockade to be
observable, the temperature has to be low
enough so that the characteristic charging
energy (the energy that is required to
charge the junction with one elementary
charge) is larger than the thermal energy
of the charge carriers. In the past, for
capacitances above 1 femtofarad
(10−15 farad), this implied that the
temperature has to be below about
1 kelvin. This temperature range is
routinely reached for example by 3He
refrigerators. Thanks to small sized
quantum dots of only few nanometers,
Coulomb blockade has been observed
next above liquid helium temperature, up
to room temperature.[4][5]
To make a tunnel junction in plate
condenser geometry with a capacitance of
1 femtofarad, using an oxide layer of
electric permittivity 10 and thickness one
nanometer, one has to create electrodes
with dimensions of approximately 100 by
100 nanometers. This range of
dimensions is routinely reached for
example by electron beam lithography and
appropriate pattern transfer technologies,
like the Niemeyer-Dolan technique, also
known as shadow evaporation technique.
The integration of quantum dot fabrication
with standard industrial technology has
been achieved for silicon. CMOS process
for obtaining massive production of single
electron quantum dot transistors with
channel size down to 20 nm x 20 nm has
been implemented.[6]

Single-electron transistor

Schematic of a single-electron transistor.


Left to right: energy levels of source, island and drain in
a single-electron transistor for the blocking state
(upper part) and transmitting state (lower part).

Single-electron transistor with niobium leads and


aluminium island.
The simplest device in which the effect of
Coulomb blockade can be observed is the
so-called single-electron transistor. It
consists of two electrodes known as the
drain and the source, connected through
tunnel junctions to one common electrode
with a low self-capacitance, known as the
island. The electrical potential of the island
can be tuned by a third electrode, known
as the gate, which is capacitively coupled
to the island.

In the blocking state no accessible energy


levels are within tunneling range of an
electron (in red) on the source contact. All
energy levels on the island electrode with
lower energies are occupied.

When a positive voltage is applied to the


gate electrode the energy levels of the
island electrode are lowered. The electron
(green 1.) can tunnel onto the island (2.),
occupying a previously vacant energy
level. From there it can tunnel onto the
drain electrode (3.) where it inelastically
scatters and reaches the drain electrode
Fermi level (4.).

The energy levels of the island electrode


are evenly spaced with a separation of
This gives rise to a self-capacitance
of the island, defined as

To achieve the Coulomb blockade, three


criteria have to be met:

1. The bias voltage must be lower than the


elementary charge divided by the self-

capacitance of the island:  ;

2. The thermal energy in the source


contact plus the thermal energy in the
island, i.e. must be below the

charging energy: or else the


electron will be able to pass the QD via
thermal excitation; and
3. The tunneling resistance, should be

greater than which is derived from

Heisenberg's uncertainty principle.[7]

Coulomb blockade
thermometer
A typical Coulomb blockade thermometer
(CBT) is made from an array of metallic
islands, connected to each other through a
thin insulating layer. A tunnel junction
forms between the islands, and as voltage
is applied, electrons may tunnel across
this junction. The tunneling rates and
hence the conductance vary according to
the charging energy of the islands as well
as the thermal energy of the system.

Coulomb blockade thermometer is a


primary thermometer based on electric
conductance characteristics of tunnel
junction arrays. The parameter
V½=5.439NkBT/e, the full width at half
minimum of the measured differential
conductance dip over an array of N
junctions together with the physical
constants provide the absolute
temperature.

References
1. Prati E (2011). "Valley blockade quantum
switching in Silicon nanostructures". J
Nanosc Nanotech. 11 (10): 8522–8526.
arXiv:1203.5368  .
doi:10.1166/jnn.2011.4957 .
2. Crippa A; et al. (2015). "Valley blockade
and multielectron spin-valley Kondo effect
in silicon". Physical Review B. 92: 035424.
arXiv:1501.02665  .
Bibcode:2015PhRvB..92c5424C .
doi:10.1103/PhysRevB.92.035424 .
3. "nanoHUB.org - Resources: Coulomb
Blockade Simulation" .
doi:10.4231/d3c24qp1w .
4. Couto, ODD; Puebla, J (2011). "Charge
control in InP/(Ga,In)P single quantum dots
embedded in Schottky diodes". Physical
Review B. 84. arXiv:1107.2522  .
Bibcode:2011PhRvB..84l5301C .
doi:10.1103/PhysRevB.84.125301 .
5. Shin, S. J.; Lee, J. J.; Kang, H. J.; Choi, J.
B.; Yang, S. -R. E.; Takahashi, Y.; Hasko, D. G.
(2011). "Room-Temperature Charge Stability
Modulated by Quantum Effects in a
Nanoscale Silicon Island". Nano Letters. 11
(4): 1591–1597. arXiv:1201.3724  .
Bibcode:2011NanoL..11.1591S .
doi:10.1021/nl1044692 . PMID 21446734 .
6. Prati, E.; De Michielis, M.; Belli, M.; Cocco,
S.; Fanciulli, M.; Kotekar-Patil, D.; Ruoff, M.;
Kern, D. P.; Wharam, D. A.; Verduijn, J.;
Tettamanzi, G. C.; Rogge, S.; Roche, B.;
Wacquez, R.; Jehl, X.; Vinet, M.; Sanquer, M.
(2012). "Few electron limit of n-type metal
oxide semiconductor single electron
transistors". Nanotechnology. 23 (21):
215204. arXiv:1203.4811  .
Bibcode:2012Nanot..23u5204P .
doi:10.1088/0957-4484/23/21/215204 .
PMID 22552118 .
7. Wasshuber, Christoph (1997). "2.5
Minimum Tunnel Resistance for Single
Electron Charging". About Single-Electron
Devices and Circuits (Ph.D.). Vienna
University of Technology. Retrieved
12/5/2012. Check date values in:
|access-date= (help)
General
Single Charge Tunneling: Coulomb
Blockade Phenomena in Nanostructures,
eds. H. Grabert and M. H. Devoret
(Plenum Press, New York, 1992)
D.V. Averin and K.K Likharev, in
Mesoscopic Phenomena in Solids, eds.
B.L. Altshuler, P.A. Lee, and R.A. Webb
(Elsevier, Amsterdam, 1991)
Fulton, T.A. & Dolan, G.J. "Observation of
single-electron charging effects in small
tunnel junctions" Phys. Rev. Lett. 59, 109-
112 (1987),
doi:10.1103/PhysRevLett.59.109

External links
Computational Single-Electronics book
Coulomb blockade online lecture

Retrieved from
"https://en.wikipedia.org/w/index.php?
title=Coulomb_blockade&oldid=819958716#Single-
electron_transistor"

Last edited 7 days ago by Mz7

Content is available under CC BY-SA 3.0 unless


otherwise noted.

You might also like