Download as docx, pdf, or txt
Download as docx, pdf, or txt
You are on page 1of 25

2.

5 Chapter Summary

The reduction of spatial dimensions to nanoscale values leads to a transition from classical physics to quantum
mechanics. In the latter framework, the properties of systems of particles are described in terms of wavefunctions of
single particles. Essentially, the wavefunction is a measure of the probability of finding a particle at a specific spatial
location and time

. Confinement of particles at the nanoscale in one, two, or three dimensions results in quantization of energy. In
other words, the energy levels are not continuous as they are for macroscopic systems. Moreover, the energies of
particles in low-dimensional systems are higher than those of particles in bulk systems. The higher the degree of
spatial confinement (smaller particles), the larger is the separation between energy levels, and the higher are the
energy values

Properties of Nanostructures

In Chapter 2, the reduction of the extent of a solid in one or more dimensions was shown to lead to a dramatic
alteration of the overall behavior of the solids. Generally, the physical properties of any given material can be
characterized by some critical length, e.g., thermal diffusion length and attenuation distance. What makes
nanoparticles very interesting and endows them with their remarkable properties is that their dimensions are
smaller than a relevant critical length. Accordingly, the electron states of nanostructures are quantized, leading to
new and usually striking electrical, thermal, magnetic, optical, and mechanical properties at the nanoscale.
Accordingly, nanostructures are of both basic and practical interest since their physico-chemical properties can be
tailored by controlling their size and shape at the nanoscale, leading to improved and/or novel applications.

The energy spectrum (i.e., the ensemble of discrete eigenenergies) of a quantum well, quantum wire, or quantum
dot can be engineered by controlling (i) the size and shape of the confinement region and (ii) the strength of the
confinement potential. The resulting control over the physico-chemical properties of the nanostructures is limited
only by the accuracy of the experimental techniques used for the fabrication of the low-dimensional structures
(Chapter 4). The situation is not unlike that of quantum phenomena, many of which were described at the beginning
of the 20th century but could not be demonstrated until the 1960s to 1970s when appropriate nanofabrication
techniques were developed.

Not only is fabrication at the nanoscale limited by the available techniques, but other practically unavoidable factors
such as imperfections—known to influence the properties of any material—may have very significant impacts on the
properties of nanomaterials. As such, size dispersion, shape dispersion, defects, residual stresses, impurities, etc.,
are of great importance for devices of reduced dimensionality. These factors may, in many instances, create a gap
between expectation and realization.

As in Chapter 2, the reader who is not interested in the details on how the electrical, thermal, magnetic, optical, and
mechanical properties change at the nanoscale may prefer to skip to Section 3.7, wherein the most significant
properties are summarized.

3.1 Band Diagrams

In order to understand many of the properties of nanostructured materials, a basic knowledge of band diagrams is
required, along with knowledge of such concepts as bandgap energy, direct/indirect bandgaps, holes, and excitons.
A band diagram depicts the energy of an allowed state of a charge-carrying particle in a particular material as a
function of its momentum or, equivalently because of wave–particle duality, as a function of the wave vector ~k. The
magnitude of the wave vector is the wave number denoted by k. In a band diagram, the allowed energy states are
grouped into bands, the valence band and the conduction band being the most important bands for the description
of the properties of many materials. As the wave vector is a 3D quantity, band diagrams should be 4D plots.
However, very often only 2D and 3D plots for specific ranges of the wave vector direction are drawn; furthermore,
symmetries often allow 3D plots to be developed as 2D plots.

Figure 3.1 presents typical 1D band diagrams for direct- and indirect-bandgap semiconductors, the direction of the
wave vector being fixed, but not its magnitude. The two basic features that play an important role in the behavior of
any material can be gleaned from a band diagram: (i) the bandgap energy Eg and (ii) whether the bandgap is direct
or indirect. The bandgap is an energy range wherein no states are allowed for charge carriers to occupy. As the
bandgap energy is defined as the difference in energy between the lowest point of the conduction band and the
highest point of the valence band, Eg is the vertical distance between these two bands in the band diagram. Typical
values of the bandgap energy at room temperature are: Eg(Si) = 1.11 eV, Eg(Ge) = 0.66 eV, Eg(GaAs) = 1.43 eV,
Eg(CdS) = 2.42 eV, and Eg(InP) = 1.27 eV.

The traversal of bandgaps by charge carriers involves the net exchange of energy between the charge carriers on the
one hand and photons and/or phonons on the other. Photons of energy ~ω are quanta of light. Phonons of energy
~ω are also quantum particles, being defined as quanta of excitation of the crystalline-lattice vibration modes of
angular frequency ω. The concept of a phonon arises from a quantum-mechanical treatment of lattice vibrations in a
solid, after assuming that a lattice-vibration mode is analogous to a simple harmonic oscillation. Phonons play an
important role in the behavior of solids by affecting their optical, electrical, and thermal properties through
interactions with photons, electrons, neutrons, etc.

Figure 3.1 One-dimensional band diagrams containing (a) direct and (b) indirect bandgaps. These 2D plots are called
1D band diagrams because the direction of the wave vector (or momentum) has been fixed, but its magnitude is
allowed to vary

A bandgap can be either direct or indirect. In a direct-bandgap material the maximum energy of the valence band
and the minimum of the conduction band occur at the same value of the momentum, as depicted in Fig. 3.1(a).
Either photons supply the necessary energy for a particle to climb to the conduction band from the valence band, or
photons are emitted during the transition of the particle from the conduction band to the valence band. No phonons
participate in the transitions from the conduction/valence band to the valence/conduction band. Most III–V
compound semiconductors, such as GaAs, are direct-bandgap semiconductors and are widely used for
optoelectronic applications.

An indirect-bandgap semiconductor is one in which the maximum energy of the valence band and the lowest energy
of the conduction band occur at different values of the momentum, as shown in Fig. 3.1(b). Since momentum (in
addition to energy) must be conserved in any interband transition, phonons need to participate in these transitions,
making these transitions less likely (or less efficient) in indirectbandgap semiconductors. Silicon and germanium are
examples of indirect-gap semiconductors, with limited practical use in optoelectronics.

Energy bands are populated by charge carriers, i.e., electrons and holes. Holes are virtual entities that can be
thought of as electron vacancies. The charge of a hole is opposite in sign but equal in magnitude to that of an
electron. The effective mass of a hole is somewhat different from that of an electron, their respective effective
masses being dependent on the curvatures of the valence and conduction bands.

An exciton is a quasi-particle comprising an electron and a hole bound to each other. As an exciton’s energy is
slightly below the bandgap energy, transitions of slightly lower energy than Eg become possible. Excitons can move
through a material and transport energy, although they do not transport charge as they are electrically neutral.
Typical binding energies of excitons are Eex (Si) = 14.7 meV, Eex (Ge) = 4.15 meV, Eex (GaAs) = 4.2 meV, Eex (CdS) =
29.0 meV, and Eex (InP) = 4.0 meV.

3.2 Electrical-Transport Properties

The changes that occur in the electronic properties as the length scale is reduced are mainly related to the increasing
influence of the wavelike properties of the electrons and the scarcity of scattering centers. As one or more
dimensions become comparable with the de Broglie wavelengths of electrons (see Section 2.1), the discrete nature
of the energy states becomes apparent, although a fully discrete energy spectrum is observed only in systems that
are confined in all three dimensions. The finite spacing of energy states as a result of quantum-mechanical
confinement engenders fundamental and technologically important phenomena, which are being explored for
electronic nanodevices.

Although electrical transport occurs through the motion of both electrons and holes, the remainder of this section is
focused on electrons but can be adapted to holes. The confinement of charge carriers in two or fewer dimensions
results in quantized energy levels, as discussed in Chapter 2. The confinement of electrons to a 2D structure
consisting of a conducting strip of width w and length l amounts to the creation of a 2D electron gas (2DEG). Its
conductance is given by

and its conductivity by

where e is the charge of an electron, ρs is the surface charge density, and τ is a relaxation time that takes into
account the delay due to the collisions of electrons with the structure. The physical definition of G is the ratio of the
total current to the voltage drop across a sample of length l in the direction of current flow. For a 3D electron gas
(3DEG), this relationship can be used when replacing w by the cross-sectional area A orthogonal to the current flow.
Similar expressions are also valid for the thermal transport of energy, as discussed in Section 3.3.

Novel size-dependent effects emerge as the dimensions w (or A) and l are reduced toward atomic dimensions in the
nanometer range. The relationship expressed by Eq. (3.1) holds in the diffusive-transport regime where both w and l
are greater than the mean free path λe of the electrons. As the width of the strip decreases, quantum-mechanical
effects begin to emerge. The quantum-mechanical confinement of an electron in a strip of width w leads to the
discretization of energy levels given by En = ~ 2 /2m ∗ (nπ/w) 2 , (n = 1, 2, 3, . . . .), per Eq. (2.10), where m ∗ is the
effective mass of the electron and ~ is the reduced Planck constant. The conductance is determined by the number
of these w-dependent transverse confined states that are occupied. Thus, rather than a simple linear dependence of
G on w, quantum mechanics forces a particular dependence of G on w. As w is altered, the energy spectrum
changes, as does the number of occupied states below the Fermi energy level EF, which is the highest energy level
occupied by electrons (or, more precisely, fermions) at T = 0 K; hence, the conductance changes as well. Thus, the
quantum-mechanical effect of reducing the dimension w is the change of conductance in discrete steps, much like a
staircase.

Quantization of conductance can be observed even at room temperature in some special cases, but in general this
effect is exhibited at very low temperatures. In a short quasi-1D channel formed between two regions of a 2DEG in
an AlGaAs/GaAs heterostructure by the action of metallic gate electrodes deposited on top of the layered structure,
as shown in Fig. 3.2(a), the conductance increase in discrete steps as the electron density in the channel is increased.
Figure 3.2(b) shows a sequence of steps in the conductance of a constriction in a 2DEG, as the width w is varied by
the application of a voltage across the gate.

Figure 3.2 (a) Schematic representation of a quantum point contact, defined in a highmobility 2D electron gas at the
interface of a GaAs/AlGaAs heterojunction. The point contact is formed when a negative voltage is applied to the
gate electrodes on top of the AlGaAs layer. Transport measurements are made by employing contacts to the 2D
electron gas at either side of the constriction. (b) Conductance quantization of a quantum point contact in units of 2e
2 /h. As the gate voltage defining the constriction is made less negative, the width of the point contact increases
continuously, but the number of propagating modes at the Fermi level increases stepwise. The resulting
conductance steps are smeared out when the thermal energy becomes comparable to the energy separation of the
modes. [Reprinted with permission from van Houten and Beenakker.1 c 2005, American Institute of Physics.]

The effect on the conductance G of reducing the length l is a particularly interesting phenomenon. If the ohmic
regime were to hold in Eq. (3.1), the reduction of l toward zero would make G increase without limit, and the
resistance would decrease to zero. However, there is always a finite residual resistance. In the ballistic transport
regime l < λe, electrons can propagate without losing their initial momentum since scattering events can be
neglected. The expression for ballistic conductance including two-spin orientation (spin degeneracy) in the ideal case
becomes

G = 2e 2 h .

This relationship is usually called the Landauer formula.

Accordingly, the ballistic conductance of a 1D channel is quantized in units of the conductance quantum G0 = 2e 2 /h
= 7.748 × 10−5 Ω−1 , which is twice the reciprocal of the resistance quantum R0 = h/e 2 = 2.581 × 104 Ω. The
Landauer formula for quantum transport can be generalized to a network in which several wires connect a barrier
with reservoirs, leading to an expression that sums over the contribution of each channel. Then, with Nc being the
number of channels available for transport (i.e., the number of transverse modes with energies below the Fermi
energy of the electrodes or electron reservoirs), we have
G = Nc 2e 2 h . (3.4)

This leads to quantized conductance and well-defined steps in the measured resistance as either the Fermi energy or
the effective width of the wire is changed. Thus, while the classical conductance depends linearly on the width (G ∝
w), the quantum-mechanical conductance increases in discrete steps of 2e 2 /h as w increases enough to permit one
more transverse-quantization state to be occupied and hence available for conduction.

Quantization is very important for quantum wires of small cross-sectional dimensions and depends on how
conduction electrons interact with the atoms of a material. In practice, semiconductor wires unambiguously show
conductance quantization for large cross-sectional dimensions (∼100 nm) because the electronic states due to
confinement are spatially extended. As a result, their Fermi wavelengths λF = hc/EF are large, which means that
adjacent energy states are not widely separated since EF is inversely proportional to λF. Hence, the energy state of
an electron can be resolved only at a cryogenic temperature (few Kelvin) where the thermal excitation energy is
lower than the interstate energy gaps. For metals, quantization corresponding to the lowest energy state is only
observed for atomic wires. Its wavelength being thus extremely small, a metallic atomic wire has widely separated
energy states, which allows resistance quantization to be easily observed at room temperature.

Conduction in highly confined 0D structures, such as quantum dots, is very sensitive to the presence of other charge
carriers and, hence, the charge state of the dot. Transport through quantum dots shows striking effects due to the
electron’s wave nature and its finite charge. If a particular quantum dot is fully decoupled from its environment, it
confines a well-defined number N of electrons. For weak coupling, deviations due to tunneling through the barriers
are small, leading to discrete values of the total electrostatic energy of the quantum dot. This energy can be
estimated as N(N − 1)e 2 /2C, where C is the capacitance of the quantum dot. Thus, Ne2 /C is the amount of energy
required to increase the number of confined electrons by one. This additional energy is discretely spaced in units of e
2 /C. If this charging energy exceeds the thermal energy kBT, where kB is the Boltzmann constant and T is the
temperature, the electrons cannot tunnel on and off the quantum dot by thermal excitations alone, and transport
can be blocked, which is referred to as a Coulomb blockade.

Electron transport through a quantum dot is studied by connecting the quantum dot to surrounding reservoirs, as
illustrated in Fig. 3.3. The fact that the charge on an electron island formed by electrons confined to the quantum
dot is quantized in units of e regulates transport through the quantum dot in the Coulomb-blockade regime. Electron
transfer between the reservoirs and the quantum dot occurs via tunneling through potential barriers, which are thick
enough that the transfer is dominated by resonances due to quantum confinement in the quantum dot. The
Coulomb blockade results in conduction processes involving single electrons; as a result, only a small amount of
energy is required to operate a switch, transistor, or memory element.

Figure 3.3 Schematic energy-level diagrams within the Coulomb-blockade model∗ (a) at a Coulomb peak, where
linear-response (V = 0) transport is possible; (b) between peaks, where linear-response transport is blockaded; (c)
and (d) at two different applied voltages (VC and VD), where transport occurs through the first and second occupied
states, respectively. [From Bockrath et al.2 Reprinted with permission from AAAS.]

The energy-level diagrams in Fig. 3.3 show N electrons confined to a quantum dot, followed by a gap of value U + ∆E
for adding the (N + 1)-th electron. Above this, additional levels separated by an energy gap ∆E are shown, which
correspond to adding the (N + 1)-th electron to one of the excited single-particle states of the dot. At a voltage V
corresponding to a Coulomb-blockade peak, the energy of the lowest empty state aligns with the electrochemical
potential in the leads, and single electrons can tunnel on and off the dot at V = 0 [Fig. 3.3(a)]. At gate voltages
between peaks [Fig. 3.3(b)], tunneling is suppressed because of the singleelectron charging energy U. However, if V
is increased so that the electrochemical potential of the right lead is pulled below the energy of the highest filled
state, an electron can tunnel off the dot [Fig. 3.3(c)]. Further increasing V allows tunneling out of additional states
[Fig. 3.3(d)]. Similar processes occur for negative bias, corresponding to tunneling through unoccupied states above
the Coulomb gap.

Quantum-dot technology is one of the most promising candidates for use in solid state quantum computation, i.e.,
where quantum properties are used to perform numerical operations. By applying a small voltage, one can control
the flow of electrons through the quantum dot and thereby make precise measurements of the spin and other
quantum properties.

3.3 Thermal-Transport Properties

Thermal transport at the nanoscale differs from that at the macroscale for several basic reasons. In bulk materials,
internal scattering dominates the thermal-transport processes. As the size shrinks, the frequency of collisions
between phonons and boundaries increases, as does the ratio of surface/interface to the volume. The interface
scattering of phonons and the associated thermal boundary resistance can dominate thermal conduction in
nanostructures. The size effects, however, are not limited to the thermal processes inside nanostructures. In the
vicinity of small devices, phonons become rarefied when their mean free path is comparable or larger than the
device size, which effectively increases the thermal resistance for removing thermal energy from the devices.
Understanding these thermal processes is not only important for the prediction of the temperature rise in a
microelectronic device and its reliability, but also can enable new technologies such as lowdimensional
thermoelectric materials.

Depending on the size of the nanostructure, thermal conductance may exhibit quantum features. Owing to the finite
spacing of vibrational frequencies, the transfer of phononic energy through an electrically nonconducting
nanoparticle (i.e., a molecule, atomic chain, or a single atom) between two reservoirs exhibits characteristics similar
to those of ballistic transport of electrons discussed in Section 3.2. However, thermal energy can be transferred by
both phonons and electrons, which obey different statistics. Thermal energy is transported only by phonons in
dielectric wires, whereas in metallic wires the thermal energy transported by electrons dominates that transported
by phonons.

As the size of a nanostructure becomes comparable or smaller than the mean free path of phonons, phonons collide
with the boundary more often than in bulk materials. An increase in the frequency of collisions enhances the
resistance to thermal transport and thus reduces the effective thermal conductivity of thin films and wires. Similar
but stronger size effects occur inside 1D (quantum wires, Section 6.7) and 0D (quantum dots, Section 6.5) structures.
Figure 3.4 compares the estimated thermal conductivity of thin films and wires of silicon, together with some
experimental data reported on the thermal conductivity of single-crystal thin films of silicon. The thermal
conductivity of wires drops faster with thickness as a result of increased phonon scattering in the direction
perpendicular to the wire axis.

Figure 3.4 Estimated thermal conductivity of silicon films/wires as a function of thickness/diameter. Experimental
points are for thin films of silicon. [Reprinted from Chen3 with permission from Elsevier.]

The expression for the universal quantum of 1D thermal conductance in the ballistic transport mode is given by

g0 = π 2 k 2 B 3h T = 9.456 × 10−13T WK−2 .

The quantum of thermal conductance is determined only by fundamental constants and the absolute temperature.
This magnitude is universal, independent of the characteristics of the material. This result is analogous to Eq. (3.2)
for the electrical quantized conductance, which is independent of the electron velocity in a 1D channel. The quantum
of thermal conductance represents the maximum possible value of energy transported per phonon mode.

3.4 Magnetic Properties

Nanostructured materials have remarkable magnetic properties, which have found several practical applications. In
fact, ferromagnetism—a cooperative state of matter in which a huge number of magnetic moments are all locked
parallel that leads to macroscopic magnetization—can be considered a nanoscale phenomenon. In a ferromagnet,
each atom has an electronic magnetic moment, these moments all being aligned in the same direction.

Among the several phenomena that low-dimensional structures display under the influence of magnetic fields, giant
magnetoresistance (GMR) and colossal magnetoresistance (CMR) must be highlighted. Magnetoresistance is a
phenomenon in which the application of a dc magnetic field changes the electrical resistance of a material, and is
due to the conduction electrons being forced to move in helical trajectories about an applied magnetic field. The
resistance of a material is the result of the scattering of electrons out of the direction of current flow by collisions.
Since magnetoresistance generally occurs at very large magnetic fields and low temperatures, only limited
applications have been found. Magnetoresistance occurs in pure copper at 4 K under an applied magnetic field of 10
T [10 kG = 1 T (tesla)], resulting in a tenfold increase in resistance. In 1988 huge magnetoresistance was measured in
artificial materials consisting of alternating nanometer-thin films of ferromagnetic materials (Fe, Co, etc.) and weaker
magnetic or nonferromagnetic metals (Cr, Cu, etc.), as shown in Fig. 3.5. This effect was termed giant
magnetoresistence (GMR). GMR has sparked the fast development of spintronics (sometimes called spin electronics
or even magneto-electronics), i.e., the use of the spin of electrons in information circuits.

Figure 3.6 shows the electrical resistance as a function of the applied magnetic field, whereby we can infer that
resistance decreases during the magnetization process and becomes practically constant when the magnetization is
saturated. GMR is obtained in antiferromagnetically coupled Fe/Cr multilayer systems by aligning the magnetization
of adjacent Fe layers with the applied magnetic field. GMR occurs because of the dependence of electron scattering
on the orientation of the electron spin with respect to the direction of magnetization. Electrons whose spins are not
aligned along the direction of magnetization are scattered more strongly than those with their spins aligned with
magnetization. The application of the magnetic field parallel to the layers forces the magnetization in all magnetic
layers to be in the same direction. This causes the magnetizations that are pointing opposite to the direction of the
applied magnetic field to flip. The conduction electrons with spins aligned opposite to the magnetization are more
strongly scattered at the metal–ferromagnet interface, and those aligned along the field direction are less strongly
scattered. Because the two spin channels are in parallel, the lower-resistance channel determines the resistance of
the material. The magnetoresistance effect in these multilayered materials is a sensitive detector of dc magnetic
fields and has enabled the development of very sensitive reading heads for magnetic disks. Other material systems
have been discovered to have even larger magnetoresistance than multilayered materials, so this phenomenon is
called colossal magnetoresistance (CMR).

Figure 3.6 Magnetoresistance measurements at 4.2 K of three Fe/Cr multilayer systems (Fe/Cr)n. The current and the
applied magnetic field are along the same axis in the plane of the layers. To the far right as well as to the far left (H >
Hs and H < Hs , Hs being the saturation magnetic field), the magnetization in all iron layers is parallel to the external
magnetic field. In the low-field regime, every second iron layer is magnetized antiparallel to the external magnetic
field (10 kG = 1 T). [Reprinted with permission from Baibich et al.4 c 1988 by the American Physical Society.]

Materials made of single-domain ferromagnetic nanoparticles with randomly oriented magnetizations embedded in
a nonmagnetic matrix also display GMR. The magnetoresistance in these materials, unlike the multilayered
materials, is isotropic. The application of a dc magnetic field rotates the magnetization vector of the ferromagnetic
nanoparticles parallel to the direction of the field, which reduces the resistance. The magnitude of the effect of the
applied magnetic field on the resistance increases with the magnitude of the magnetic field and as the size of the
magnetic nanoparticles decreases.

Perhaps the most remarkable property of low-dimensional systems under the influence of magnetic fields is the
quantum Hall effect. This effect is observed when a large magnetic field is applied perpendicular to a 2DEG at a low
temperature. On applying a magnetic field Bz in the direction perpendicular to a nanometer-thick film, as shown in
Fig. 3.7, the transverse resistivity ρxy increases in steps at high values of the applied magnetic field. At the j-th step,

ρxy = − h je2 = − 25812.807 j Ω, j = 1, 2, 3, . . . .

In the regimes where ρxy is constant, the longitudinal resistivity ρxx is vanishingly small. The unusual behavior of the
2DEG exhibited in Fig. 3.7 is known as the quantum Hall effect.

The quantization of the Hall conductivity σxy = 1/ρxy has the important property of being incredibly precise. Actual
measurements of the Hall conductivity have been found to be either integral or fractional multiples of 1/R0 to nearly
one part in a billion. This phenomenon, referred to as exact quantization, has allowed for the definition of a new
practical standard for electrical resistance: the resistance quantum R0, sometimes referred to as the resistance unit,
roughly equals 25812.8 Ω. This magnitude is referred to as the von Klitzing constant RK (after von Klitzing, the
discoverer of exact quantization); since 1990, a fixed conventional value RK−90 has been used worldwide to calibrate
resistors. The quantum Hall effect also provides an extremely precise independent determination of the fine
structure constant α = e 2 /2hcε0, a quantity of fundamental importance in quantum electrodynamics, ε0 being the
permittivity of vacuum or free space.
3.5 Optical Properties

Quantum confinement in low-dimensional semiconductors splits the bulk conduction and valence bands into a series
of discrete energy levels, as discussed in Chapter 2. The degeneracy and separation levels depend on the
dimensionality and shape of the confinement regime. In quantum wells (Section 2.2), quantum wires (Section 2.3),
and quantum dots (Section 2.4), confinement effects lead to blueshifts (i.e., shifts to shorter wavelengths and thus
higher energies) of the electron and hole states. The optical transitions, which are affected by the nature of the
excitons—the electron–hole-pair states introduced in Section 3.1—also become discrete. The blueshifts of electron
and hole states and the discrete exciton states make the optical properties of low-dimensional structures very
different from those of the bulk material; this is especially true in the case of quantum dots because of confinement
in all three dimensions.

In semiconductor nanocrystals (0D structures), both the electron states in the conduction band and the hole states in
the valence band become quantized as a result of quantum confinement, as shown in Section 2.4. As the size
reduces, the bandgap grows and can be chosen over a wide range of energies by appropriately selecting the average
diameter of the nanocrystals. As a consequence, quantum dots of the same material, but with different sizes, can
emit light of different wavelengths (and thus colors), which is one of the optical features of quantum dots
immediately noticeable to the unaided eye. In fact, while the material that makes up a quantum dot defines its
intrinsic energy signature, more significant in terms of coloration is the size. Recent studies suggest that the shape of
the quantum dot may also be a factor in the coloration.

Figure 3.7 Experimental curves for the Hall resistivity (ρxy) and the longitudinal resistivity (ρxx) of a heterostructure
as a function of the applied magnetic field at a fixed carrier density, recorded at 8 mK. (Inset) Schematic for Hall-
effect measurements, where I is the electric current parallel to an applied electric field Ex. The transverse
movements of electrons and holes due to the application of the magnetic field Bz creates a transverse current
density, thereby engendering an electric field Ey. [Reprinted with permission from von Klitzing.5 c 1986 by the
American Physical Society.]

The absorption edge—i.e., the lowest-energy absorption state—is shifted to higher energy with respect to that of
the bulk semiconductor as the size decreases. Furthermore, above the absorption edge, (i) the spectrum is stepped
rather than smooth, the steps corresponding to allowed transitions between valence states and conduction states
and (ii) at each step, sharp peaks appear, corresponding to confined-exciton states. The absorption intensity in
nanocrystals becomes concentrated at the specific frequencies corresponding to the transitions between discrete
states. Additionally, large absorption coefficients are observed.

Size reduction results in a blueshift of the characteristic transition energies compared to the bulk semiconductor
material. A smaller size results in an upshift of the threshold for absorption from the corresponding bulk value; in
addition, a similar shift is found in the emission spectrum. Accordingly, the optical spectra of several semiconductor
nanocrystals can be selected from a continuous palette across the visible spectrum simply by choosing the
appropriate size. This fact makes semiconductor nanocrystals useful in applications ranging from fluorescent labels
to light-emitting diodes. In particular, the larger the dot, the redder is the fluorescence; the smaller the dot, the
bluer is the fluorescence. Quantitatively speaking, the bandgap energy that determines the energy (and hence color)
of the fluorescence is inversely proportional to the square of the size of the quantum dot. Larger quantum dots have
more energy levels that are more closely spaced. This allows the quantum dot to absorb photons containing less
energy, i.e., those closer to the red end of the spectrum.

Just as in atoms, the energy levels of small quantum dots can be probed by optical spectroscopic techniques. Figure
3.8 shows room-temperature optical absorption measurements for PbSe nanocrystals ranging in size from 3 to 9 nm,
from which quantum-sized effects can be deduced. Since the absorption edge is given by the bandgap, the bandgap
increases as the nanocrystal size decreases. Also, the intensity of the absorption increases as the size is reduced. The
higherenergy peaks are associated with excitons and shift to higher energies with decreasing size. These effects are a
result of the confinement of excitons.

According to the Kramers–Kronig relations to which every linear causal system must adhere, a bulk semiconductor
and a nanocrystal have the same overall absorption per unit volume when integrated over all frequencies, although
the absorption spectra are very different.† The absorption spectrum of a typical bulk semiconductor is continuous,
whereas that of a nanocrystal consists of a series of discrete transitions with very high magnitudes at the transition
frequencies. These strong transitions at particular frequencies have motivated the fabrication of lasers that are
based on the quantized electronic transitions in quantum dots.

Figure 3.8 Room-temperature optical absorption spectra for a series of PbSe nanocrystals measuring (a) 3.0 nm, (b)
3.5 nm, (c) 4.5 nm, (d) 5 nm, (e) 5.5 nm, (f) 7 nm, (g) 8 nm, and (h) 9 nm in diameter. [Reprinted with permission
from Murray et al.7 ]

Quantum dots are valued for displays because they emit light in very specific Gaussian distributions. Quantum-dot
displays can very accurately reflect the colors that the human eye can perceive. Quantum dots also require very little
power since their outputs need not be filtered for color purity. In contrast, a display made of liquid crystal devices
(LCDs) is powered by a single fluorescent lamp that is color filtered to produce red, green, and blue pixels. Thus,
when an LCD display shows a fully white screen, two-thirds of the light is absorbed by the filters. Displays that
intrinsically produce monochromatic light can for this simple reason be more efficient.

Let us think of a quantum dot as a sufficiently small metal sphere, wherein the electrons in the conduction band fill
up the quantized energy levels. These quantized levels affect the optical—and the electrical—properties and can
even influence the stability of the metal sphere. The optical properties of small metal spheres typically exhibit a local
surface plasmon resonance (LSPR) as follows: In the presence of an external electric field, the centers of positive and
negative charges in any atom in the metal sphere do not coincide. Each atom acts as an electric dipole. The metal
sphere is said to be polarized. With the twin assumptions (i) that the atoms respond instantaneously to the external
electric field Eexc (i.e., that the retardation effects can be ignored) and (ii) that a free-electron cloud is created inside
the electrically small metal sphere, the induced polarization can be written as

P = ε0 1 − ω 2 p 3ω2 Eexc,

where ωp is the plasma frequency of the bulk metal and ω is the angular frequency. The polarization thus diverges at
a frequency of ωlspr = ωp/ √ 3, called the LSPR frequency. The most striking effect is that ωlspr shifts the bulk plasma
resonance of metals such as gold and silver from the ultraviolet regime into the visible regime. This result is
independent of the radius of the metal sphere. In reality, however, the optical properties do depend somewhat on
size due to retardation effects at larger radii, and due to losses and intraband transitions at smaller radii. Liquids or
glasses containing metallic nanoparticles are often brilliantly colored due to absorption indicative of LSPR. As shown
in Fig. 3.9, the LSPR bands for nanoparticles of silver and gold usually fall within the visible range. Furthermore, the
LSPR bandwidth and the peak maximum strongly depend on the size and shape of the nanoparticle.

As with quantum dots, size confinement also plays an important role in determining the energy levels in 1D systems
such as nanowires, once the crosssectional diameter has been reduced below a critical value [such as the de Broglie
wavelength (Section 2.1)] that depends on the particular nanostructure. This effect has been experimentally
observed with nanowires made of different materials. For example, the absorption threshold of silicon nanowires is
significantly blueshifted as compared with that of bulk silicon. Moreover, sharp, discrete features in the absorption
spectra are observed along with relatively strong photoluminescence with emission energy close to the absorption
edge. These features most likely originate from size-confinement effects, although surface states (energy states
present at the surface of solids) might make additional contributions because of the large surface-to-volume ratio in
low-dimensional structures. Furthermore, a variation in the growth direction for a nanowire usually leads to different
optical signatures. For example, Si nanowires oriented along the h100i crystalline direction exhibit significantly
higher transition energies across the bandgap associated with excitons when compared with the nanowires oriented
along the h110i direction.

Figure 3.9 Selection of the LSPR wavelength of Au and Ag nanoparticles by choosing appropriate size, shape, and
composition. [Reprinted from Kalele et al.8 ]

Light emitted from a nanowire is highly polarized along its longitudinal axis. Hence, a striking difference exists in the
photoluminescence intensities recorded in the directions parallel and perpendicular to the axis of an isolated
nanowire. In some cases, this anisotropy can be quantitatively explained in terms of the large dielectric contrast
between the nanowire and the surrounding environment. The polarization response has been exploited to fabricate
polarization-sensitive nanoscale photodetectors for integrated photonic circuits, optical switches and interconnects,
near-field imaging, and high-resolution detection systems. Additionally, nanowires with flat ends can be useful as
optical resonance cavities to generate coherent light on the nanoscale. Coherent nonlinear optical phenomena such
as second- and third-harmonic generation have been observed in nanowires.

3.6 Mechanical Properties

The mechanical properties of a material depend fundamentally on the nature of bonding among its constituents, and
on its microstructure on several length scales. Most nanograined materials, typically characterized by an average
grain size that is less than several tens of nanometers, have grain-size-dependent mechanical properties that are
significantly different from those of their coarse- grained counterparts. Two of the most widely reported differences
are the departure from the Hall–Petch relation of the yield stress and the increased ductility (superplasticity). In
particular, ultrapure metals are much stronger and apparently less ductile than metals with impurities.
Intermetallics, i.e., compounds of two or more metals resulting in a material with different properties from the
constituent metals, are also stronger, but tend to be more ductile when the grain size is small. Although these
property changes are thought to be primarily controlled by grain size, they are also affected by the large percentage
of atoms in grain boundaries, as well as other microstructural features. Strengthening appears to result from a
limitation of the activity of dislocations, which are defects within the crystal structure of a solid, while increased
ductility probably relates to the relative sliding between adjacent grains.

A quantity commonly used to mechanically characterize materials is their yield strength σy, defined as the stress at
which the material begins to deform plastically, resulting in permanent deformation. The yield strength of a material
is related to the grain size by the empirical Hall–Petch relation:

σy = σ0 + KHP · d −1/2 ,

where σ0 is the frictional stress opposing dislocation movement, KHP is a materialdependent constant, and d is the
average grain diameter. The hardness of a material can be also described by a similar equation. Both yield strength
and hardness are higher for reduced grain size. The reason for this behavior is that materials having smaller grains
have more grain boundaries, thereby blocking dislocation movement.

However, significant departures from the Hall–Petch behavior have been reported for materials made of grains
typically less than 20–50 nm in size, depending on the particular material. The departure from the linear behavior
varies from no dependence on the grain size (zero slope) to decreases in yield strength with grain size (negative
slope), as shown in Fig. 3.10. These markedly different characteristics are primarily caused by the increased role
played by the grain boundaries in a nanograined material. For the traditional coarsegrained materials, the grain
boundary can generally be regarded as a region of almost zero thickness whose deformation contributes little to the
overall plastic strain; however, when the grain size reduces to the nanometer scale, a significant population of atoms
exists at the grain boundaries whose deformation can have a dominant effect on the overall plastic strain of the
material, thereby resulting in permanent deformation. Additionally, tension-compression asymmetry (usually
termed strength-differential effect) has also been reported for nanograined materials (see Fig. 3.10), but is absent
for coarse-grained materials.

Many nanostructured metals and ceramics are observed to be superplastic, in that they are able to undergo
extensive deformation without forming necks or actually fracturing. Superplasticity is presumed to arise from
diffusion and sliding of grain boundaries, which become increasingly significant in a finer-grained material. Overall,
these effects extend the current yield strength and ductility limits of conventional materials for which usually a gain
in yield strength is offset by a corresponding loss in ductility. The increased ductility has been observed in both
metals and ceramics. In this regard, nanocrystalline copper with purity better than 99.993 atomic percent and an
average grain size of 28 nm has been shown to deform up to 5100% at room temperature; some composite ceramics
can deform up to 1050% at 1650 ◦C without failure.

Figure 3.10 Departure of the yield strength from the Hall–Petch plot for decreasing grain size of copper. The bottom
four experimental data points were taken under tension, and the top curve was calculated under compression.
There is a significant strength-differential effect between tension and compression in the nanometer range.
[Adapted from Jiang and Weng9 with kind permission of Springer Science and Business Media.]

3.7 Chapter Summary


Size reduction at the nanoscale results in quantization of electrical conductance which, in some cases, is observable
even at room temperature. The discovery of this quantization led to the introduction of a new fundamental
constant, namely the conductance quantum. Likewise, thermal conductance becomes quantized for low-dimensional
structures, and a quantum of thermal conductance, independent of the characteristics of the material, has been
defined.

Under the influence of magnetic fields, low-dimensional structures show giant magnetoresistance and colossal
magnetoresistance, i.e., the electrical resistance is greatly increased by the application of external magnetic fields.
Another remarkable property is the quantum Hall effect, in which the transverse resistivity of a 2D electron gas
increases in steps at high magnitudes of the applied magnetic field, resulting in quantized Hall conductivity. The
discovery of this property has led to the definition of the resistance quantum, referred to as the von Klitzing
constant.

From an optical point of view, quantum confinement results in shifts to shorter wavelengths (higher energies) of the
absorption edge and the emission wavelength. Additionally, the typical spectra of nanostructures are stepped rather
than smooth, as is generally the case for bulk materials. This opens the possibility of tuning the emission and
absorption of light with size.

Finally, notable departures from the linear behavior of the yield strength and hardness with grain size have been
observed at the nanoscale. However, the particular behavior is very much dependent on the nanomaterial.
Nanofabrication

The availability of new and improved methods for fabricating nanomaterials as well as the availability of new and
improved tools for characterization and manipulation of nanomaterials (Chapter 5) have engendered an explosive
growth of nanoscience and nanotechnology during the last decade.

The fabrication of nanomaterials of tailored properties involves the control of size, shape, structure, composition,
and purity of their constituents. Fabrication techniques for nanostructures can be broadly divided into two
categories. Topdown approaches consist of the miniaturization or size reduction (by, e.g., etching or milling) of larger
structures. Often, lithographic patterning to structure bulk materials at the nanoscale is involved. Bottom-up
approaches are based on growth and self-assembly to build up nanostructures from atomic or molecular precursors.
A major challenge is to create hybrid approaches and develop strategies to reliably fabricate complex material
systems spanning all length scales from the molecular to the macroscopic. In fact, when developing a technique for
fabricating nanostructures, the most important issue is the simultaneous control over dimensions, morphology,
composition, and uniformity. Nanomaterials fabricated by different routes may have different internal structures
that would affect their properties.

The most common, as well as the most promising, techniques used for the fabrication of nanostructures at present
are summarized in Table 4.1 and are succinctly described in the following sections.

4.1 Physical Vapor Deposition

Physical vapor deposition (PVD) is a versatile method for preparing thin-film materials with structural control at the
atomic or nanometer scale by carefully monitoring the processing parameters. PVD involves the generation of vapor
phase species via evaporation, sputtering, or laser ablation.

In thermal evaporation, atoms or clusters of atoms are removed from a crucible containing some bulk or powder
material by heating the crucible either by passing a current through it or by directing a beam of electrons onto the
bulk material in the crucible. The evaporation process takes place in an evacuated chamber. The vapor phase
emanating from the source material condenses on a substrate, as shown in Fig. 4.1. In sputtering, atoms are ejected
from the surface of the bulk material by he impact of energetic ions created in a plasma environment (Fig. 4.2). In
laser ablation or pulsed laser deposition, an intense pulsed laser beam irradiates the bulk material, thereby releasing
a vapor of atoms and clusters of atoms. As both thermal evaporation and laser ablation occur at low pressure, atoms
in the vapor travel more or less rectilinearly, i.e., with minimal scattering. However, some scattering does occur
during sputtering and generally yields denser films.
Table 4.1 The most common and most promising techniques for nanofabrication.
Fabrication technique Principle of operation or description Remarks
Physical vapor deposition Vapor phase species—generated from a solid sample via Oblique-angle deposition (OAD): Highly collimated vapor
(PVD): either evaporation, sputtering, or laser ablation in high enables the fabrication of thin films with engineered
– Evaporation (thermal or e- vacuum—is collected on a substrate to form a nanoscale morphology.
beam) nanomaterial. Ion-beam-assisted deposition (IBAD): Combination of ion
– Sputtering implantation with a PVD technique
– Laser ablation
Chemical vapor deposition Substrate exposed to one or more volatile precursor Variants:
(CVD) materials, which react and/or decompose on its surface APCVD, LPCVD, UHVCVD, MPCVD, PECVD, RPECVD, AACVD,
under thermal, laser, or plasma excitation DLICVD, MOCVD, RTCVD, Cat-CVD, HWCVD, and HFCVD
Atomic layer deposition (ALD) Alternate pulsing of the precursor gases and vapors onto Surface-controlled self-limiting method for depositing thin
the Surface of a substrate and subsequent chemisorption films from gaseous precursors that allows the growth of
or surface reaction of the precursors conformal thin films with precisely controlled thickness on
large areas
Molecular beam epitaxy (MBE) Molecular beams of constituent materials directed onto a Produces ultrathin films as high-quality epitaxial layers with
heated crystalline substrate to form thin epitaxial layers very sharp interfaces and good control of thickness, doping,
and composition
Nanolithography: Selective removal of parts of a thin film or substrate by Related techniques:
– Photolithography using light to transfer a geometric pattern from a Scattering with angular limitation in projection electron-
– X-ray lithography photomask to a layer of a photoresist on the substrate beam lithography (SCALPEL), projection reduction exposure
– Extreme ultraviolet with variable-axis immersion lenses (PREVAIL), and digital
lithography lithography
Nano-imprint lithography (NIL) Patterns created by mechanical deformation of an Variants:
imprinted photoresist Step-and-flash imprint lithography, lithographically induced
self-assembly (LISA), laser-assisted direct imprint (LADI),
and laser-assisted nano-imprint lithography (LAN)
Scanning probe lithography Scanning tip used as mechanical electric, and/or thermal Related technique:
(SPL) source to induce a physico-chemical process in order to Dip-pen nanolithography (DPN)
modify, deposit, remove, and manipulate materials at the
nanoscale
Focused ion-beam (FIB) Focused beam of slow heavy ions creates patterns at the Related technique:
nanoscale by modification, deposition, or sputtering Ion-beam sculpting to make nanopores in a two-step
Proton-beam (p-beam) writing Focused beam of fast protons directly writes patterns on a process
photoresist.
Self-assembly and self- Delicate control of Surface properties for new growth to Bottom-up techniques that generally rely on the self-
organization spontaneously form structures with the desired shape and assembling nature of organic molecules, including complex
size species such as DNA.
Self-assembled monolayers (SAMs) comprise organic
molecules whose functionality can be modified by chemical
treatment or radiation so that the subsequent layers can be
selectively attached and used to direct oriented crystal
growth.
Langmuir–Blodgett (LB) Deposition from solution of monomolecular organic films
method on different substrates
Layer-by-layer (LbL) assembly Alternating adsorption of positively and negatively charged Conformal coating of substrates of complex shapes
species from aqueous solutions
Other techniques:
- Procesamiento de conversión Atomización de precursores químicos en gotitas de aerosol
por pulverización que se dispersan en un medio gaseoso.

– Thermal spraying Molten droplets of coating material projected at high speed


onto a surface
– Sol-gel processing Generation of a colloidal suspension (sol) which is
subsequently converted to a viscous gel and then to a solid
material
– Pyrolysis Chemical decomposition of organic materials by heating in
the absence of oxygen or any other reagents
– Electrochemical processes Reactions at solid–liquid interfaces controlled by an
externally applied voltage

Figure 4.1 Schematic of a typical setup for evaporation. The source material is heated by an electrical filament in
thermal evaporation, or by an electron beam in electron-beam evaporation.

Thermal evaporation is not very suitable for making multicomponent thin films, since some bulk materials evaporate
before others due to differences in vapor pressure of the evaporating species. Sputtering is capable of depositing
highmelting-point materials such as refractory metals and ceramics. This is because the vapor is created by transfer
of momentum from ions to atoms, rather tan by direct heating of atoms. As the sputtered atoms carry more energy
than the thermally evaporated atoms, the sputter-grown films usually have higher mass density. However, the
typical deposition rate of sputtering is significantly lower than that of thermal evaporation. Laser ablation usually
provides better control through simultaneous evaporation of multicomponent materials in a very short time period.

Figure 4.2 Typical sputtering system.

Oblique angle deposition (OAD) is a PVD method that enables the fabrication of nano-engineered thin films. The
vapor created is largely collimated so that the thin films grown have a columnar morphology. Developments in OAD
technology during the last three decades have produced columnar nanostructures of various shapes (including
vertical, tilted, helical, and chevronic) and graded-porosity thin films for use in applications ranging from sensors and
actuators to optical filters, microfluidics, and catalysis.

Ion-beam-assisted deposition (IBAD) is a technique that combines ion implantation with another PVD technique.
IBAD is particularly effective in improving adhesion and morphology control and has the advantage of having more
independent processing parameters than other PVD methods. For example, the energy and flux of bombarding ions
can be exploited to modify the size and crystallographic orientation of grains.

4.2 Chemical Vapor Deposition

Chemical vapor deposition (CVD) is a technique in which the substrate is exposed to one or more volatile precursor
materials that react and/or decompose on the substrate surface to produce the desired deposit. This technique is
used to produce high-purity, high-performance solid materials. Thermal, laser, or plasma energies are used for the
decomposition of gaseous reactants. Figure 4.3 represents a plasma-assisted CVD system. CVD has been a well-
established technique for thinfilm deposition for many years because it offers the advantages of a relatively simple
apparatus, excellent uniformity, high density, high deposition rate, and amenability to large-scale production.
Figure 4.3 Plasma-assisted chemical vapor deposition system.

Many forms of CVD are in wide use and are frequently referenced in the literature. These processes differ in the
means by which chemical reactions are initiated (e.g., activation process) and process conditions. As such,
atmosphericpressure CVD (APCVD), low-pressure CVD (LPCVD), and ultrahigh-vacuum CVD (UHVCVD) are named
after the typical chamber pressure at which the reaction takes place. Depending on the characteristics of the plasma,
the following forms can be found: microwave plasma-assisted CVD (MPCVD), plasmaenhanced CVD (PECVD),
magneto-microwave plasma CVD, and remote plasmaenhanced CVD (RPECVD). If the characteristics of the vapor
used are considered, the following two forms are commonly found: aerosol-assisted CVD (AACVD) and direct liquid-
injection CVD (DLICVD). Metal-organic CVD (MOCVD) uses metal-organic precursors, while in rapid thermal CVD
(RTCVD) the substrate is heated. Catalytic CVD (Cat-CVD) is based on the catalytic decomposition of precursors using
a resistively heated filament. This technique is also known as hotwire CVD (HWCVD) or hot-filament CVD (HFCVD).

4.3 Atomic Layer Deposition

Atomic layer deposition (ALD) is a surface-controlled self-limiting method for depositing thin films from gaseous
precursors. Although ALD can be considered a modification of CVD, the distinctive feature of ALD is the self-limiting
growth mechanism that affords this method several attractive properties: accurate and simple control of film
thickness, production of sharp interfaces, uniformity over large areas, excellent conformality with the substrate,
good reproducibility, multilayer processing capability, and desirable film qualities at relatively low temperatures.
From a nanotechnological angle, the most important benefits of ALD are excellent conformality and the possibility of
subnanometer control of film thickness.

ALD relies on alternate pulsing of the precursor gases and vapors onto the substrate surface and subsequent
chemisorption or surface reaction of the precursors. The reactor is purged with an inert gas between the precursor
pulses. This process is schematically depicted in Fig. 4.4. With a proper adjustment of the experimental conditions,
the process proceeds via saturative steps; i.e., the precursors exposed on the surface chemisorb on it (or react with
the surface groups), saturatively forming a tightly bound monolayer on the surface. The subsequent purging step
removes all of the excess molecules from the reactor chamber. When the next precursor is dosed in, it encounters
only the surface monolayer with which it reacts, thereby producing the desired solid product and gaseous by-
products. Under such conditions, the film growth is self-limiting, since the amount of solid deposited during one
cycle is dictated by the amount of precursor molecules present in the saturatively formed surface monolayer.
Therefore, the growth is stable, and the thickness increase is constant in each deposition cycle. The self-limiting
growth mechanism facilitates the growth of conformal thin films with accurately controlled thickness on large areas.
This technique also allows the growth of multilayer structures. However, a major limitation of ALD is its typically slow
growth rate.

4.4 Molecular Beam Epitaxy

Molecular beam epitaxy (MBE) is a technique to produce ultrathin films as highquality epitaxial layers with very
sharp interfaces and good control of thickness, doping, and composition. Deposition takes place in very high vacuum.
Because of the high degree of control possible with MBE and the possibility of growing compound semiconductors, it
is a valuable tool in the development of sophisticated electronic and optoelectronic devices.

Figure 4.4 Schematic representation of the typical ALD technique. In order to grow a thin film, the following four
steps need to be repeated: (1) exposure to the first precursor, (2) purge of the deposition chamber to remove
precursors in excess and by-products, (3) exposure to a second precursor, and (4) purge of the deposition chamber.
(From http://wwwrpl.stanford.edu/user/files/www/ald1.gif.)

In MBE, the constituent materials in the form of molecular beams are deposited on a heated crystalline substrate to
form a thin epitaxial layer, as depicted schematically in Fig. 4.5. Each layer has a definite crystallographic relationship
with the substrate. The molecular beams are typically obtained from thermally evaporated elemental sources, but
other sources include metal-organic group-III precursors (MOMBE), gaseous group-V hydride or organic precursors
(gas-source MBE), or some combination [chemical beam epitaxy (CBE)]. To obtain epitaxial layers of high purity, it is
critical that the source materials be extremely pure and that the entire process be carried out in an ultrahigh-
vacuum environment. Growth rates are typically on the order of a few Å/s, and the molecular beams can be
shuttered in a fraction of a second, allowing for almost atomically abrupt transitions from one material species to
another. Given the ultrahigh-vacuum environment of the system, analytical techniques such as reflection high-
energyelectron diffraction (RHEED) and mass spectrometry are often used for in situ monitoring of the growing thin
film.

4.5 Nanolithography

Nanolithography, or lithography at the nanoscale, refers to the fabrication of patterns with at least one lateral
dimension between the size of an individual atom and approximately 100 nm. This technique comprises several
lithographic techniques, including photolithography, x-ray lithography, electron beam lithography, and ion-beam
lithography, depending on the radiation employed, as explained in the following paragraphs. Photolithography (also
called optical.

Figure 4.5 Typical system for molecular beam epitaxy (MBE). Solid source materials are placed in effusion cells and
heated to be used as molecular flux sources. The substrate is heated to the necessary temperature and, when
needed, continuously rotated to improve the growth homogeneity.

Photolithography (also called optical lithography), which has been the predominant patterning technique since the
dawn of the age of semiconductors, is capable of producing sub-100-nm patterns with the use of very short-
wavelength light (currently 193 nm). This microfabrication process is used to selectively remove parts of a thin film
(or the bulk of a substrate) by first using light to transfer a geometric pattern from a photomask to a layer of a light-
sensitive chemical (photoresist, or simply resist) on the substrate. A series of chemical treatments then engraves the
exposure pattern into the material underneath the photoresist. In a complex integrated circuit (for example, a
modern CMOS), a wafer will go through up to 50 photolithographic cycles.

Photolithography resembles conventional lithography used to print on paper. It delivers exact control over the shape
and size of the pattern it creates and can create patterns over a large surface (∼30-cm diameter) in one round of
photolithographic cycles. The main disadvantages of photolithography are that it requires a flat substrate to start
with, is not very effective at creating nonplanar shapes, and requires extremely clean operating conditions. Figure
4.6 shows a general sequence of steps for a typical photolithography process. Resolutionenhancement techniques
for photolithography continue to be developed, leading to a reduction of the minimum feature size.

X-ray lithography can be extended to a resolution of 15 nm by using the ultrashort wavelength of 1 nm for the
illumination. Extreme ultraviolet lithography (EUV) uses the same principles as conventional optical lithography,
although the exposure wavelength is typically in the range of 11–13.5 nm.

Electron-beam lithography uses a beam of electrons to generate patterns on a surface. The primary advantage of
this technique is that it overcomes the diffraction limit of light. With today’s electron optics, the diameters of
electron beams can routinely go down to a few nanometers, limited mainly by aberrations and space charge.
However, pattern generation, which is a serial process, is very slow when compared with a parallel technique such as
photolithography (the current standard), in which an entire surface is patterned all at once. Therefore, electronbeam
lithography has limited application in industry.

A projection optical system that is geometrically equivalent to that used in optical lithography will improve
throughput. In this regard, the most notable attempts to use mask projection are scattering angular-limited
projection electronbeam lithography (SCALPEL), developed at Bell Laboratories and projection reduction exposure
with variable-axis immersion lenses (PREVAIL), developed by IBM.

A related technique is ion-projection lithography (IPL), in which the electron beam is replaced by an ion beam. In IPL,
generally medium-energy (50 to 150 keV) ions (e.g., protons, H + , He+ , Ar+ , etc.) are used, although heavy ions
have also been used. IPL combines the mass-fabrication advantage of masked lithography with the remarkable
properties of ions.

Figure 4.6 Representation of the general sequence of steps for a typical photolithography process: substrate
preparation, photoresist spin coat, prebake, exposure, postexposure bake, development, and postbake. Stripping
the resist is the final operation in the lithographic process, taking place after the resist pattern has been transferred
into the underlying layer via etching or ion implantation. This sequence is generally performed on several tools
linked together into a contiguous unit called a lithographic cluster. [Reprinted from Mack.1 ]

The recently developed digital lithography can be succinctly described as the process of jet-printing an etch mask. A
thin film is first grown by any deposition technique. Then, the mask pattern is jet printed directly onto the substrate.
The film is subsequently etched to reproduce the pattern, and, finally, the etch mask is removed. Among other
characteristics, this technique simplifies the conventional photolithography process by reducing the number of steps.
Additionally, it can be used to pattern a number of materials.

4.6 Nano-imprint Lithography

Nano-imprint lithography (NIL) is a rapidly emerging method of fabricating nanoscale patterns, depicted
schematically in Fig. 4.7. It is a simple and inexpensive method with high throughput and high resolution. Patterns
are created by mechanical deformation of an imprint resist. The imprint resist is typically a monomer or polymer that
is cured by heat or ultraviolet light during the imprinting process. Unlike conventional lithography techniques, nano-
imprint lithography’s resolution is not limited by the effects of wave diffraction, scattering and interference in a
resist, and backscattering from a substrate.

Nano-imprint lithography is used to fabricate devices for electrical, optical, photonic, and biological applications. The
hallmark of nano-imprint lithography is its ability to pattern 3D structures. Nano-imprint lithography and its variants
— such as step-and-flash imprint lithography, which uses transparent templates and UV-curable materials to allow
pattern replication at room temperature and low pressure; lithographically induced self-assembly (LISA), which takes
advantage of the self-assembly phenomenon observed in polymer thin films; laser-assisted direct imprint (LADI), in
which a single excimer laser pulse melts a thin surface layer of silicon, and a mould is embossed into the resulting
liquid layer; and laser-assisted nano-imprint lithography (LAN), which combines the advantages of NIL and
laserassisted direct imprint—are promising nanopattern replication technologies. Nanoimprint lithography can be
combined with contact printing.

4.7 Scanning Probe Lithography

Techniques grouped together as scanning probe microscopy (SPM) include scanning tunneling microscopy (STM)
and atomic force microscopy (AFM). SPM is generally used to determine the topography of samples at the nanoscale
(Chapter 5). Taking advantage of the sharpness of the probe tips, as well as of strong and localized tip–surface
interactions, SPM has also been used to manipulate atoms on metal surfaces and to fabricate nanopatterns of metal
and semiconductor surfaces. These successful examples exemplify the growing field of scanning probe lithography
(SPL).

Scanning probe microscopy—more specifically, STM and AFM—is increasingly used to modify, deposit, remove, and
manipulate materials at the nanoscale, thus making it a powerful tool in the fabrication of nanomaterials. Some
researchers date the emergence of nanotechnology to 1981, when the scanning tunneling microscope was invented.
The scanning tips are used as mechanical, electric, and/or thermal sources to induce different physico-chemical
processes. SPM has the additional advantages of requiring relatively inexpensive apparatus, involving relatively easy
operation, and providing the possibility for parallel operation, thus enormously increasing its throughput. Three
main categories can be distinguished regarding the use of SPM in the field of nanofabrication: material modification,
including resist exposure and oxidation; material addition, mainly consisting of induced deposition; and material
removal, including scratching and etching. Figure 4.8 shows two examples of AFM-tip-induced oxidation on silicon
substrates.

Figure 4.7 Schematic of the nano-imprint technique. [Reprinted from Nie and Kumacheva2 with permission from
Macmillan Publishers Ltd. c 2008.]

Figure 4.8 AFM images of AFM-tip-induced oxidation lines on n-type-silicon substrates with (a) ∼90-nm linewidth
and (b) ∼23-nm linewidth. [Reprinted with permission from Tseng et al.3 c 2005, American Vacuum Society.]

SPM is capable of manipulating single atoms on a given surface, even on nonplanar surfaces. This fabrication
technique is called atom manipulation or nanoassembly. To exemplify the products of this technique, Fig. 4.9 shows
a 3D STM image of a quantum corral during construction and after completion. The quantum corral nanostructure is
composed of silver atoms grown onto a silver substrate.

Dip-pen nanolithography (DPN) uses an AFM tip coated with a thin film of ink molecules that react with the substrate
surface to write nanoscale patterns for lithography applications. The first experiments were carried out on gold
surfaces with alkanethiols as inks

4.8 Focused Ion-Beam Technique, Proton-Beam Writing, and Ion-Beam Sculpting

Focused ion-beam (FIB) and proton-beam (p-beam) writing are maskless techniques capable of fabricating sub-100-
nm features. The ability of these techniques to fabricate 3D structures of flexible geometry enables rapid
prototyping of micro- and nanosystems. Furthermore, these two ion techniques, together with IPL (Section 4.5),
have complementary areas of application.

The FIB technique uses a focused beam of slow heavy ions (with energies typically around 30 keV) to create patterns
at the nanoscale by modification, deposition, or sputtering. The FIB technique is remarkable in that patterns can be
produced in virtually any material, although the process is relatively slow. Protonbeam writing uses a focused beam
of fast [million electron volt (MeV)] protons for directly writing patterns on several types of photoresists. Their high
energy allows the incident protons to penetrate deep into the photoresist.

Ion-beam sculpting is a term used to describe a two-step process to make nanopores. The first step is to make either
a hole all the way through a solid or a blind hole (i.e., a hole that does not break through on the backside), most
commonly using a focused ion-beam (FIB) machine. The holes are commonly ∼100 nm in diameter, although they
can be made much smaller. This first step may or may not be done at room temperature. For the second step, there
are three common techniques to sculpt the hole: broad-area ion exposure, transmission electron microscopy (TEM)
exposure, and FIB exposure. Holes can be closed completely, or they can be left open at a lower limit of 1–10 nm in
diameter

Figure 4.9 STM images of a quantum corral grown onto silver (a) during construction and (b) after completion of the
corral, in which 36 silver atoms are used. These are observed as white protrusions of diameter 31.2 nm. [Reprinted
with permission from Hla et al.4 c 2003 by the American Physical Society.

Figure 4.10 shows an ion-beam sculpting apparatus that incorporates feedback into the fabrication process to gain
dimensional control over the hole diameter at the single-nanometer length scale. In this instrument, a free-standing
membrane surface containing an initial ∼100-nm hole or bowl-shaped cavity is exposed to a normal beam of low-
energy ions. Since an argon-ion beam is used to create the nanopores, the rate of argon transmission through the
hole provides a direct measure of its size. This rate is monitored to provide the feedback signal necessary to trigger
the extinction of the ion beam when the desired hole diameter is obtained.

Figure 4.10 Schematic of a feedback-controlled ion-beam sculpting tool. [Reprinted with permission from Stein et
al.5 c 2004, American Institute of Physics.]

By controlling such experimental parameters as sample material, temperature, ion flux, and time of exposure to the
ion beam, the length of the hole can be made either smaller than or equal to the thickness of the solid. Hence, the
process is referred to as ion beam sculpting rather than etching or sputtering.

4.9 Self-Assembly, Self-Organization, and Self-Assembled Monolayers

Self-assembly and self-organization of nanoparticles involves a delicate control of surface properties so that new
growth spontaneously forms structures with the desired geometry. Self-assembly includes bulk reactive methods
such as the chemical formation of colloidal semiconductors. The technique has been extensively explored as a
bottom-up approach for generating complex nanostructures on various length scales. Many self-assembly processes
rely on the self-assembling nature of organic molecules, including complex species such as DNA. These methods are
termed chemical or molecular self-assembly. Generally, molecular self-assembly is defined as the spontaneous
organization of relatively rigid molecules into structurally well-defined aggregates via weak reversible interactions
such as hydrogen, ionic, and van der Waals bonds. The aggregated structure represents a minimum-energy structure
or equilibrium phase. Selfassembly is also found to occur in biological systems and in micelles and liquid crystals, and
is being increasingly used in synthetic supramolecular chemistry.
Other simpler methods rely on geometric self-organization, in which hard spheres or hard rods will arrange
themselves into 2D and 3D structures based on packing considerations. As an example, solutions of colloidal metal
particles can spontaneously order themselves into 2D hexagonally close-packed sheets on given substrates.
Molecular systems such as rod-like and disc-like liquid crystals also exhibit geometric self-organization properties.

A variation on geometric self-organization is templated self-organization in which an ordered nanostructure is


formed by deposition of a material around a previously self-organized template. This approach can be used to
produce metallic or semiconductor structures via different deposition techniques on geometrically self-organized
polystyrene or silica spheres of submicron dimensions. The spheres are subsequently dissolved to leave a highly
porous structure.

More complex self-assembly processes involve the use of self-assembled monolayers (SAMs). SAMs comprise
organic molecules whose functionality can be modified by chemical treatment or radiation (e.g., lithography) so that
the subsequent layers can be selectively attached and used to direct oriented crystal growth. The ends of the
molecules are usually terminated with a thiol group to provide good adhesion to a gold substrate. The molecules will
order on the substrate under given conditions of concentration, pH, and temperature. Another important type of
self-assembly process is the self-assembled growth of semiconductor quantum dots.

4.10 Langmuir–Blodgett Method

The Langmuir–Blodgett (LB) method is a classical method used in chemistry for the deposition of molecular
monolayers and multilayers. Making use of the hydrophilic/hydrophobic orientation of molecules, this method and
its variants allow the deposition from solution of monomolecular organic films on different substrates.

A Langmuir monolayer is generally prepared by placing a given number density of amphiphilic molecules on the
surface of an aqueous solution. One end of an amphiphilic molecule is hydrophilic, while the other end is
hydrophobic. The monolayer spreads spontaneously onto the surface and any volatile organic solvents, e.g.,
chloroform, methanol, or benzene, evaporate in a short period of time. Generally, the solution is held in a Teflon
container, known as a trough, in order to control the temperature and the fraction of the water surface accessible to
the monolayer as well as to measure surface tension (or surface pressure). One or more movable (typically,
motorized) Teflon barriers placed across the trough serve to vary the area of the monolayer.

The process of building an LB multilayer consists of periodically dipping a substrate into the solution. A layer is
deposited during each dip. However, the molecules deposited during the removal step (generally termed upstroke)
have their heads oriented toward the substrate while the molecules deposited in the immersion step (downstroke)
are oriented with their tails facing the substrate. The result is that the heads in the current layer adhere to the heads
in the previously deposited layer during the upstroke, and the tails of the current layer stick to the tails in the
previously deposited layer during the downstroke.

There are many modes for the deposition of LB films. The most common deposition mode, called the Y mode, is
illustrated in Fig. 4.11. In this mode, if the substrate is initially hydrophilic, the molecules are shown to stack in a
headto-head and tail-to-tail configuration. An almost perfect film can be assembled like this, monolayer by
monolayer.

Although the Langmuir–Blodgett method has obvious attractions, a major shortcoming until recently has been the
restricted range of materials for which it can be used. Various organic materials have been deposited using this
method, although the best results have been obtained with either fatty acids or their salts, since these substances
possess molecules with the required hydrophilic and hydrophobic ends.

Figure 4.11 Langmuir–Blodgett film deposition in the common mode called the Y mode. With a hydrophilic substrate,
no film deposition occurs during the first immersion. The first monolayer is deposited during the first upstroke.
Thereafter, the deposition of one monolayer is obtained on each immersion/emersion step (downstroke/upstroke).
(a) The layer on the surface of the water, (b) the first layer grown during the upstroke, (c) the second layer formed
after the downstroke (second immersion), and (d) the substrate with three layers (after second upstroke). [Reprinted
from Roberts et al.6 with permission from IOP.]

4.11 Layer-by-Layer Assembly


Layer-by-layer (LbL) assembly is a versatile and quite inexpensive method to fabricate ultrathin films with
nanometer-level control over their composition and structure, and thus over their specific properties. This technique
is based on the alternating adsorption of positively and negatively charged species from aqueous solutions, and can
be used to create highly tuned, functional thin films. LbL multilayer films have received considerable attention in
both fundamental studies and applied research.

A typical LbL assembly takes place as follows: A substrate is first immersed in a solution of a negatively charged
polyelectrolyte, e.g., poly(styrene sulfonate) (PSS), poly(vinyl sulfate), and poly(acrylic acid) (PAA). The polyanionic
species causes the surface of the coated substrate to have a negative charge. The coated substrate is then rinsed in a
washing solution (generally pure water) to remove any loosely adsorbed polyelectrolyte from the substrate and
dried under a nitrogen/air flow. Subsequent execution of an analogous procedure with a positively charged
polyelectrolyte solution leads to the reversal of net charge on the surface, making the surface of the doubly coated
substrate positively charged. Typical positively charged polyelectrolytes include poly(diallyldimethylammonium
chloride) (PDDA), poly(allylamine hydrochloride) (PAH), and polyethyleneimine (PEI). After these two steps, a
polyanion/polycation bilayer is fabricated on the substrate. The process is schematically depicted in Fig. 4.12. Since
the surfaces of many types of substrates (including metals, silicones, and glasses) have net negative charges in
solution as a consequence of surface oxidation and hydrolysis, it is possible to reverse the order of polyelectrolytes
and fabricate a polycation/polyanion bilayer instead.

Subsequent repetitions of the bilayer-deposition cycle result in the growth of multilayer films with the desired
morphology and thickness. The roughness, thickness, and porosity of a multilayer film can be controlled at the
molecular level by adjusting experimental parameters such as pH, ionic strength, and polyelectrolyte concentration.
Since the thickness of a single polycation/polyanion bilayer is typically below 1 nm, nanometer-scale control is
achieved.

Compared to the Langmuir–Blodgett method (Section 4.10), LbL assembly is generally much simpler and faster, and
usually results in more stable films. Also, the multilayer structure of LbL-deposited thin films allows much higher
loadings of biologically interesting species compared to self-assembled monolayers (Section 4.9).

Although the LbL technique was first applied to assemble layers of oppositely charged polymers, it has been
extended to other materials. Almost any type of charged species—including inorganic molecular clusters, metal,
semiconductor and polymer nanoparticles, nanotubes, nanowires, organic dyes, dendrimers, porphyrins, biological
polysaccharides, polypeptides, nucleic acids and DNA, proteins, and viruses—can be used to fabricate multilayer
films by LbL assembly. Furthermore, the formation forces of LbL films are not limited to only electrostatic
interactions as in the first experiments. Assemblies relying on hydrogen bonding, charge transfer, covalent bonding,
biological recognition, and hydrophobic interactions have also been investigated. This versatility translates into films
of an exceptionally wide variety of functional properties.

Figure 4.12 (a) Schematic of the film deposition process, representing the adsorption of a polycation and a
polyanion, and the washing steps in deionized water. The four steps represent the basic sequence for the fabrication
of the simplest thin film structure: (A/B) n . (b) Simplified molecular picture of the first two adsorption steps,
depicting film deposition starting with a negatively charged substrate. Counterions are omitted for clarity. [Reprinted
from Kumar and Kumar.7 ]

LbL assembly is additionally versatile with regard to the types of substrates, including hydrophilic and hydrophobic
glasses, mica, silicon, metals, quartz, and polymers. Moreover, the substrates can be of practically any shape,
including 1D tubes or 3D colloids. Furthermore, freely suspended flexible LbL structures with different shapes,
compositions, and properties can be fabricated by using sacrificial substrates (planar, spherical, and cylindrical).
Freestanding LbL microand nanocapsules, membranes, encapsulated nanoparticle arrays, sealed-cavity arrays,
microtubules, microcubes, microcantilevers, and planar films have been made.

Overall, the availability of a wide variety of component materials and substrates, and the versatility of LbL assembly
result in a large variety of applications of LbL films. Potential applications of LbL films have been proposed in such
areas as surface coatings, optics and optoelectronics, drug delivery, electrochemistry, fuel cells, chemical sensors,
nanomechanical sensors, and nanoscale chemical reactors. Also, LbL assembly has attracted extensive attention for
biomedical applications, given the possibility of heavily loading LbL films with different types of biomolecules, as well
as the stability of LbL films exposed to harsh and physiological conditions. Accordingly, applications of LbL assembly
in the fields of biomimetics, biosensors, drug delivery, protein and cell adhesion, mediation of cellular functions, and
implantable materials have been proposed.

4.12 Other Techniques

A rich palette of chemical techniques has been developed for fabricating lowdimensional structures with well-
controlled dimensions and from a broad range of materials. Methods based on chemical synthesis usually provide
attractive strategies in terms of material diversity, cost, throughput, and the potential for high-volume production.
Although some of these techniques were developed many years ago, researchers are using them to find new
applications for the development of nanostructured coatings. Some of the most common techniques are as follows.

Spray conversion processing involves the atomization of chemical precursors into aerosol droplets that are dispersed
throughout a gaseous medium. The aerosols are then transported into a heated reactor where the solution is either
evaporated or combusted to form ultrafine particles or thin films. This is a versatile and inexpensive technique
because of the availability of various inexpensive chemical solutions. Various aerosol generators—including pressure,
electrostatic, and ultrasonic atomizers—have been used for atomization.

Thermal spraying is a well-established method for forming hard coatings on selected component substrates. The
coating material is heated in a gaseous medium, and molten droplets of it are projected at high speed onto a
surface. Upon impact, the droplets flatten, transfer the thermal energy to the cold substrate, and solidify rapidly as
splats.

Sol-gel processing involves the generation of a colloidal suspension called sol, which is subsequently converted to a
viscous gel and then to a solid material. In the category of wet chemical synthesis, solution-based processing routes
used for the synthesis of nanoparticles include precipitation of solids from a supersaturated solution, homogeneous
liquid-phase chemical reduction, and ultrasonic decomposition of chemical precursors. These processes are
attractive due to their simplicity, versatility, and availability of inexpensive precursor materials.

Pyrolysis, a special case of thermolysis (i.e., dissociation of particles by heat), is the chemical decomposition of
organic materials by heating in the absence of oxygen or any other reagents, except possibly steam. Extreme
pyrolysis, which leaves only carbon as the residue, is called carbonization.

Electrochemical processes, which involve reactions at solid–liquid interfaces controlled by an externally applied
voltage, are being increasingly used as quite inexpensive, easy to handle, versatile, and reliable tools for
nanofabrication. In particular, electrodeposition is used for forming dense nanocrystalline materials. Its advantages
include low cost and industrial applicability, as it involves little modification of existing electroplating technologies. It
is simple to implement, as the electrodeposition parameters can be easily tailored to meet the required grain size,
microstructure, and chemistry of products. Electrodeposition is also versatile, as it can produce a wide variety of
pore-free materials and coatings at high production rates. Figure 4.13 shows a typical setup for electrodeposition.

Figure 4.13 Typical experimental setup for electrodeposition.


Nanoscience

11.1 INTRODUCTION

We felt that we could not leave this book without a brief look at the latest ‘hot’ topic. The prefix currently on
everyone’s lips is nano-, as in nanoscience, nanotechnology, nanostructures, nanocrystals, etc. In a now famous
lecture given in 1959 entitled ‘There’s Plenty of Room at the Bottom’, Nobel Prize-winning physicist Richard
Feynman said, “The principles of physics, as far as I can see, do not speak against the possibility of manoeuvring
things atom by atom,” and that is the nub of what scientists are now trying to do. The final aim is the ultimate in
designer chemistry: to be able to assemble a molecule, film, or solid, atom by atom. Already the dream is beginning
to be realized, and some, but by no means all, of this science falls into the remit of solid state chemistry.

In 1905, Einstein, as part of his doctoral thesis, measured the size of a sugar molecule using diffusion techniques and
found it to be about 1 nm in diameter. A hydrogen atom is about 0.1 nm in diameter, so for chemists, thinking on the
nanometre scale is thinking on the atomic or molecular scale, which is what we are used to doing. Therefore, what is
different about nanoscience and nanotechnology from what chemists have always done? Current research has two
crucial strands. The first is to investigate and utilise the properties of very small particles, which are found to have
very different properties from the bulk solid. Chemists have roles here in synthesizing these particles, investigating
their properties, and developing new uses for them; it covers such topics as nanotubes, coatings, new alloys,
composites, particles for sunscreens, catalysts, colloids, and quantum dots, to name but a few. The second is to be
able to manipulate or manufacture things very precisely on the nanometre scale. Some techniques and processing,
such as photolithography for the printing of silicon chips, are so-called top-down processes because they are carving
out or etching nanometre size structures; these are clearly based in technology, although the processes are often
chemical. The building up of structures, self-assembly processes, and the direct manipulation of atoms into
nanostructures, socalled bottom-up processes, once again often falls into the realms of chemistry.

Nanoparticles are usually considered to have at least one dimension less than 100 nm, although this is not a rigid
definition and dimensions of several hundred nanometres can fall into this research. Such particles have a larger
surface-area to volume ratio than larger particles, affecting the way they react with each other and with other
substances. A 10 nm diameter nanoparticle has about 15% of the atoms on the surface; by comparison, this drops to
≤1% for a bulk solid. Because a nanoparticle may only consist of a few atoms, the energy levels associated with
extended solids, such as bands in metals, no longer apply and the electronic energy levels are more similar to the
quantized levels found in individual atoms, this affects their conductivity, and the way they interact with light and
other forms of energy.

Unfortunately because ‘nano’ has become somewhat of a buzzword and is currently attracting a lot of research
funding, the term is being used loosely to cover a huge range of topics and techniques to do with anything that is
fairly small, and it can be difficult to separate what is important. The Royal Society has come up with two working
definitions:

Nanoscience is the study of phenomena and manipulation of materials at atomic, molecular, and macromolecular
scales, where properties differ significantly from those at larger scales.

Nanotechnology is the production and application of structures, devices, and systems for controlling shape and size
at nanometre scale.

Much of the research centres on the organic/bioorganic area and is involved with using self-assembly methods to
build large molecules with particular properties such as molecular machines. We will not discuss this type of work
here, but in the following sections will concentrate only on areas that relate in some way to solid-state chemistry.
First, we consider the physical and electronic effects of the nano-scale, and then we discuss examples which might
be useful. To try and bring some order to the discussion, we have grouped the examples into one-dimensional (1-D),
two-dimensional (2-D), and three-dimensional (3-D) systems. Finally, we look at atomic force microscopy, a
technique covered in Chapter 2, but which can also be used for moving atoms around individually and thus has
potential for building atomic and molecular structures from ‘the bottom up’.
11.2 CONSEQUENCES OF THE NANOSCALE

Properties of nanoscale materials may be very different from those of the bulk material. For instance, small particles
may melt at much lower temperatures than the bulk and are often much harder: 6 nm copper grains are five times
as hard as bulk copper.

11.2.1 NANOPARTICLE MORPHOLOGY

The high surface-to-bulk ratio in nanostructures means that a nanocrystal structure is determined by a balance
between bulk terms such as lattice energy, surface energy terms, and terms due to faults (such as dislocations) as all
these terms are now significant. This can lead to unusual crystal structures such as thin films of bcc copper,
compared with the normal bulk structure which is ccp.

II-VI semiconductors, such as CdSe and CdS, normally have the wurtzite structure (see Chapter 1) where each
element is tetrahedrally coordinated. Under high pressures (2 GPa), these transform to the six-coordinate NaCl (rock
salt) structure. However, if pressure is applied to a CdSe nanocrystal of about 4 nm in diameter, it now takes much
more pressure, about 6 GPa, to transform it to the rock salt structure. It is thought that this may be a resistance to
the exposure of high-index crystal planes in the newly formed nanocrystal, which would not have the normal
thermodynamically stable morphology.

FIGURE 7.8 Framework and cation sites in the Na+ form of zeolite A (LTA). (Courtesy of Dr. Robert Bell, Royal
Institution of Great Britian, London.)

FIGURE 7.20 Computer model illustrating how (a) para-xylene fits neatly in the pores of ZSM-5, whereas (b) meta-
xylene is too big to diffuse through.

FIGURE 7.21(b) Computer graphic of ethane and methane molecules inside one of the hexagonal pores of MCM41.

FIGURE 10.8 The structure of 1–2–3: (a) the metal positions; (c) idealized structure of YBa2Cu3O7−x; (d) the
extended structure of YBa2Cu3O7, depicting copper-oxygen planes, with the copper-oxygen diamonds in between.
Key: Cu, blue; Ba, green; Y, aqua; O, red.

FIGURE 10.9 Structure of HgBa2Ca2Cu3O8, depicting copper oxygen diamonds between the Ca layers and the copper
oxygen layers forming the bases of the pyramids. The apices of the pyramids are in the barium oxygen layers. In the
superconductor Hg0.8Tl0.2Ba2Ca2Cu3O8+0.33, one fifth of the Hg2+ ions are replaced by Tl3+ ions and additional
oxygen ions are present in the mercury layer. Key: Cu, grey; Ca, red; Ba, blue; Hg/Tl, green; O, aqua.

FIGURE 11.2 The band gap of a semiconductor depends on its size.

FIGURE 11.7 The colour of the fluorescence from a nano-sized particle depends on its dimensions.

Varying the conditions of deposition of the film in CVD can alter the morphology of the nanocrystals formed; Figure
11.1(a) and Figure 11.1(b) show nanosized diamond crystals in diamond films grown with 111 (octahedral) and 100
(cubic) faces. Techniques for producing specific morphologies could be very important in the production of catalysts
because different crystal faces can catalyse very specific reactions.

Many nanomaterials can be made in different forms. We are familiar with the example of carbon, which we can find
as diamond films, carbon black, fullerenes, and multi- and single-walled nanotubes. MoS2 can be made as
nanotubes, ‘onions’ (multi-walled fullerene-type structures), and thin films.

The methods and conditions of nanostructure manufacture are crucial, as we will see that the properties depend
critically on the size and shape of particles produced. However, the synthetic techniques used are so particular to
each system that they make a subject in their own right and we will not attempt to cover them here.

11.2.2 ELECTRONIC STRUCTURE

In Chapter 4 we discussed how the energy levels of a crystal could be obtained by thinking of a crystal as a very large
molecule. For crystals of, for instance, micrometre dimensions, the number of energy levels is so large and the gap
between them so small that we could treat them as essentially infinite solids with continuous bands of allowed
energy. At the nanometre scale, we can still think of the particles as giant molecules but a typical nanoparticle
contains 102 to 104 atoms, very large for a molecule but not large enough to make an infinite solid a good
approximation. The result is that in nanoparticles, we can still distinguish bands of energy, but the gaps between the
bands may differ from those found in larger crystals and, within the bands, the energy levels do not quite form a
continuum so that we can observe effects due to the quantised nature of levels within bands. This is illustrated in
Figure 11.2, for nano-sized crystals of semiconductor (quantum dots).

FIGURE 11.1 SEM images of (a) a diamond film grown with methane, hydrogen, and 0.2% PH3, showing (100) square
facets and (b) a diamond film grown at lower substrate temperature—now the crystals are predominantly (111)
triangular facetted. (Courtesy of Dr. P.D. May and Professor M.N.R.Ashfold, Bristol University.)

The band diagram for a semiconductor is depicted in Figure 11.2(a). The bonding electrons are held in a lower
valence band consisting of a continuum of many energy levels, and two electrons can occupy each energy level. The
orbital density of states diagrams illustrate that, in general, a lower density of states exists at the top and bottom of
the band and a higher density of states exists in the middle (indicated by shading). Above the valence band, at higher
energies, a conduction band exists, and separating the two, we see a band gap that can vary in size. If electrons are
promoted from the valence band to the empty conduction band by supplying sufficient energy for them to jump the
band gap (with heat or light, for instance), then the solid conducts.

As a crystal of a semiconductor becomes smaller, fewer atomic orbitals are available to contribute to the bands. The
orbitals are removed from each of the band edges (cf. Chapter 4, Figure 4.6) until, at a point when the crystal is very
small—a ‘dot’—the bands are no longer a continuum of orbitals, but individual quantised orbital energy levels
(Figure 11.2(b)), thus the name quantum dots. At the same time, you can see that this has had the effect of
increasing the band gap. As the size of the crystal continues to shrink, so does the number of orbital energy levels
decrease and the band gap increases. As the size of the nanoparticles of most semiconductors decreases, the band
gap increases. The band gap in CdSe crystals, for example, is approximately 1.8 eV for crystals of diameter 11.5 nm,
but approximately 3 eV for crystals of diameter 1.2 nm.

FIGURE 11.2 The band gap of a semiconductor depends on its size. See colour insert following page 356.

Quantum dots are nanometre scale in three dimensions, but structures that are only nanometre scale in two
dimensions (quantum wires) or one dimension (quantum wells or films) also display interesting properties. The
quantised nature of the bands in nanostructures can be seen in the density of states. Schematic, theoretical density
of states diagrams for bulk material, quantum wells, quantum wires, and quantum dots are pictured in Figure 11.3.

Figure 11.4 presents the density of states for a specific example: a semiconducting nanotube. The density of states
for carbon nanotubes is predicted to show sharp peaks (known as van Hove singularities) corresponding to specific
energy levels. These can be seen in the figure and have been confirmed by scanning tunnel microscopy (STM)
experiments. In Figure 11.4, a gap with zero density about the Fermi energy exists (E=0 on the figure). We associate
this band gap with semiconductors. For semiconducting carbon nanotubes, the band gap generally increases with
decreasing diameter, but for particular nanotube structures, the band gap becomes zero and the nanotubes are then
metallic conductors similar to graphite.

Electrical conductance in solids (other than ionic conductors) depends on the availability of delocalised orbitals close
enough together in energy to form bands.

FIGURE 11.3 Theoretical density of states diagrams for (a) bulk material, (b) quantum well, (c) quantum wire, and (d)
quantum dot.

FIGURE 11.4 The density of states for a semiconducting nanotube.

When we have structures on the nanoscale where levels at the tops of occupied bands are not so close in energy as
we have seen, then this can affect the conductance. Under certain conditions, electrical conductance through a
nanostructure is quantised and increases in a stepwise fashion with increasing voltage. Electrons tunnel into the
structure and fill the lowest empty quantised levels until all the levels below the highest filled level providing
electrons are full (Figure 11.5). If the thermal energy is insufficient to raise electrons to the next discrete energy
level, then no more electrons can tunnel in. The conductance thus drops until the voltage is increased to a value V
such that the energy, e×V, is sufficient to raise the energy of the electrons in the adjacent solid so that they can
reach the next energy level, when the current increases again. Figure 11.6, for example, depicts the increase of
conductance with gate voltage along a quantum wire connecting two GaAs/AlGaAs interfaces in a transistor.

11.2.3 OPTICAL PROPERTIES

The last section revealed that semiconductor band gaps for nanostructures vary with the size of the structure. The
wavelength of light emitted when an electron in the conductance band returns to the valence band will therefore
also vary. Thus, different colour fluorescence emission can be obtained from different-sized particles of the same
substance (e.g., different sized quantum dots of CdSe irradiated with UV emit different colours of light). To produce
fluorescence, light of greater photon energy than the band gap is shone onto the nanocrystal. An electron is excited
to a level in the conduction band, from where it reaches the lowest energy level in the conduction band through a
series of steps by losing energy as heat. The electron then returns to the valence band, emitting light as it does so
(Figure 11.7). Figure 11.7(a) depicts the irradiation of a large quantum dot (smaller band gap). It then decays into the
valence band emitting a photon of light, which is the coloured fluorescence seen. If the smaller quantum dot (larger
band gap) (Figure 11.7(b)) undergoes the same process, we can see that the photon emitted as it decays back into
the valence band has more energy. From the Einstein equation E=hv, the higher energy photon will have the higher
frequency and thus be closer to the blue end of the spectrum, accounting for the difference in colour. The smaller
the nanostructure, the larger the band gap, and hence, the shorter the wavelength of the emitted light. Thus, 5.5 nm
diameter particles of CdSe emit orange light whereas 2.3 nm diameter particles emit turquoise light.

FIGURE 11.5 Electrons fill the lowest empty quantised levels until all the levels below the highest filled level
providing electrons are full.

FIGURE 11.6 The increase of conductance with gate voltage along a quantum wire connecting two GaAs/AlGaAs
interfaces in a transistor.

FIGURE 11.7 The colour of the fluorescence from a nano-sized particle depends on its dimensions. See colour insert
following page 356.

In the absorption spectra of nanoparticles of CdSe and other semiconductors, not only can the shift in wavelength be
observed, but there are also bands corresponding to absorption to discrete energy levels in the conduction band. For
example, 11.5 nm diameter particles of CdSe have an absorption spectrum that shows an almost featureless edge,
but particles of diameter 1.2 nm show features resembling molecular absorption bands shifted about 200 nm to
shorter wavelengths, as depicted in Figure 11.8.

The colours produced by nanoparticles of gold (colloidal gold) have been used since Roman times where we find
them in the glass of the famous Lycurgus cup (British Museum) which appears green in reflected light and red in
transmission. Faraday studied them in detail in the mid-19th century and some of his original samples are still held at
the Royal Institution in London. In metals, light interacts with surface electrons and is then reflected. The surface
electrons in nanoparticles are induced by the light to oscillate at a particular frequency; absorption at this frequency
gives rise to the colour. The oscillation frequency, and hence the colour, depends on the size of the particle. For
example, we are familiar with bulk gold which appears yellow in reflected light, but thin films of gold appear blue
when light passes through them, and as the particle size is reduced the wavelength of the absorbed light decreases
and the film appears first red then, for 3 nm diameter particles, orange. Oscillation frequencies in the visible region
are only observed for Ag, Au, Cu, and their alloys, as well as the Group 1 metals.

FIGURE 11.8 The absorption spectrum of 11.5 nm diameter particles of CdSe has an almost featureless edge, but
particles of diameter 1.2 nm show features resembling molecular absorption bands shifted about 200 nm to shorter
wavelengths.

Particles are known to scatter light as well as absorb it and this produces the white or pale appearance of fine
powders. The even smaller nano-sized particles, however, are transparent because the scattering efficiency is
reduced. This effect has led to the use of nanoparticles in sunscreens and cosmetics. These will still absorb ultraviolet
light but will scatter less visible light.

11.2.4 MAGNETIC PROPERTIES


In Chapter 9, the idea of ferromagnetic domains was introduced. Domains typically have dimensions of
approximately 10 to 1000 nm. In nanocrystals, therefore we can reach a situation where the domain size and the
crystal dimensions are comparable. Such single domain crystals have all the electron spins in the crystal aligned. In
larger crystals, the main mechanism for magnetisation and demagnetisation is rotation of the domain walls (see
Chapter 9, Section 9.3.1). If the crystal size is reduced, as the single domain region is approached, it becomes harder
to demagnetise the crystal by applying a magnetic field.

This is because the only possible mechanism is now disruption of spin-spin coupling within a domain. If the particle
size is decreased still further, however, the number of spins decreases and the force aligning them becomes weaker.
Eventually, this force is too weak to overcome thermal randomisation and in the absence of an applied magnetic
field, the spins are randomly oriented. The crystal is then no longer ferromagnetic but superparamagnetic. Figure
11.9 is a plot of the magnetic field needed to demagnetise ferromagnetic particles (the coercivity, HC) as a function
of particle size. The radius DC is that at which the particle becomes single domain. Values for DC for the
ferromagnetic transition metals and the ferromagnetic Fe3O4 are given in Table 11.1.

FIGURE 11.9 Plot of the magnetic field needed to demagnetise ferromagnetic particles (coercivity, Hc) as a function
of particle size. The particle becomes single domain at radius Dc.

Above a critical temperature, TB, known as the blocking temperature, superparamagnetic particles can have their
spins aligned by a magnetic field, behaving in a similar fashion to paramagnetic materials but with a much larger
magnetic moment. Below the blocking temperature, there is insufficient kinetic energy to overcome the energy
barrier to reorientation of the spins. The blocking temperature varies with the strength of the applied field. In the
superparamagnetic state, the particles, like paramagnetic solids, do not show hysteresis.

So far, we have considered only isolated magnetic nanoparticles. When nanosized grains are in close contact, the
magnetic effects can differ. This is because spinexchange coupling can occur between grains. Chapter 9 revealed that
thin layers of magnetic material could have the spin state pinned by an adjacent antiferromagnetic layer. Exchange
coupling between grains of hard and soft magnetic materials can produce nanocomposites with high remanence and
high coercivity, that is they remain strongly magnetic when the applied magnetic field is reduced to zero and need a
large magnetic field to de-magnetise them.

TABLE 11.1 Single domain radii for selected ferromagnetic solids

11.2.5 MECHANICAL PROPERTIES

Properties of solids such as hardness and plasticity have long been known to vary with grain size. For example, down
to µm-size grains, the resistance to plastic deformation increases with decreasing grain size. However, as the grain
size is reduced still further, the resistance levels off or even decreases. The effect has been attributed to a different
mechanism for plastic deformation in nano-sized grains. In larger crystals, deformation is mainly governed by the
movement of dislocations in the crystal. Such movements are inhibited by grain boundaries. As the grain size is
reduced, the ratio of grain boundary to bulk grain increases and so deformation becomes harder. Eventually, as the
grain size reaches 5 to 30 nm, movement of dislocations becomes negligible, but deformation through atoms sliding
along grain boundaries becomes favourable. This latter mechanism is aided by a large ratio of boundary to bulk grain
and so softening increases as the grain size is reduced.

11.2.6 MELTING

Nanocrystals show a melting temperature depression with decreasing size. Melting of nanocrystals can be observed
in a transmission electron microscope, and gold nanoclusters have been found to melt at 300 K compared with a
melting temperature of 1338 K for elemental gold in bulk, and 3 nm CdS crystals melt at about 700 K in a vacuum,
compared with 1678 K in the bulk. The surface energy becomes an increasing factor as the size of a crystal becomes
smaller; thermodynamics predicts a lowering of the melting temperature if the surface energy of the solid is higher
than that of the liquid.

11.3 EXAMPLES

To bring some semblance of order to such a huge and diverse area, we have grouped our examples under 1-D, 2-D,
and 3-D headings, where 1-D refers to materials with one dimension in the nanometre range and are extended in
two dimensions; 2-D is confined to nanometres in two dimensions but extended in one dimension; and 3-D is
confined to nanometre dimensions in all three directions.

You might also like