[go: up one dir, main page]

0% found this document useful (0 votes)
52 views145 pages

Surface Science Lectures 2

Uploaded by

Aniya
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
52 views145 pages

Surface Science Lectures 2

Uploaded by

Aniya
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 145

Entropy

Entropy, S, is the quantitative measure of the degree of disorder in a system.


Entropy is ‘a measure of disorder’ or ‘the There is a general tendency towards greater
amount of randomness’ in a system. All systems entropy – nature tends to move from order to
possess some degree of randomness because disorder in isolated systems. For example, gas
particles are always in constant motion. molecules spread out over time to fill a space,
increasing their entropy. Energy here is
changing, from being concentrated to being
more spread out: particles are becoming more
disordered and so entropy increases.

This graph shows how entropy gradually changes with


temperature. At 0K, perfect crystals have 0 entropy.
The largest jumps in entropy values occur at the melting
and boiling points. The greatest change occurs at Gas molecules spread out
boiling point because a gas is significantly more Entropy increases
disordered than a solid or liquid.
Second Law of Thermodynamics

The total entropy of the universe increases


in any spontaneous process

DSouniverse = DSosystem + DSosurroundings > 0

“I propose to name the quantity S the entropy of the system,


after the Greek word [trope], the transformation. I have
deliberately chosen the word entropy to be as similar as
possible to the word energy.” (1865)

q
T 0

Rudolf Clausius (1822-1888)


Entropy

• Entropy of a substance in different phase

solid liquid gas

In gas phase, molecules are more randomly distributed

Of all phases, gases have the highest entropy


Ssolid < Sliquid < Sgas
Entropy

• Entropies of molecules with different sizes or complexity

Larger molecules have more internal motion

Larger Molecules generally have a larger entropy

Ssmall < Smedium < Slarge


Entropy

• Usually, dissolving a solid or liquid will increase the entropy

dissolves

lower entropy higher entropy


more disordered arrangement
Entropy

• Dissolving gas in liquid decreases the entropy

dissolves

overall more disordered lower entropy


arrangement:
higher entropy
The Thermodynamic Potentials
Four quantities called “thermodynamic potentials” are useful in the chemical
thermodynamics of reactions and non-cyclic processes. They are internal energy , the
enthalpy, the Helmholtz free energy and the Gibbs free energy.

Internal energy
It is the capacity to do work plus the capacity to release heat.

U = T dS –p dV
Helmholtz free energy
It is the capacity to do mechanical plus non-mechanical work.
F = U-TS
Enthalpy
It is the capacity to do non-mechanical work plus the capacity to release heat.

H = U + pV
Gibbs free energy,
It is the capacity to do non-mechanical work.

G = U + pV – TS

where T = temperature, S = entropy, p = pressure, V = volume.


Internal Energy

The internal energy U of a thermodynamic system is defined as

U = Upot + Ukin

The energy is defined as the sum of all potential Upot and kinetic energy Ukin required
in order to create the system. Ukin comprises the energy required for the motion of the
atoms or compounds within the system. However, Ukin does not include the kinetic
energy of the lateral movement of the system as a whole. Upot comprises the energy
stored as mass (as well as by the electron configurations of the atoms), the chemical
bonds (which may further react), and the physical forces (e.g., electromagnetic forces
due to the presence of dipoles or internal stress/strain) within the system or among
the atoms or compounds of the system.
Enthalpy
The internal energy is the energy required for creating a thermodynamic system.
However, this system must occupy a given space and the energy required for creating
this space needs to be added to the total energy required for creating and placing a
system. This combined energy is referred to as enthalpy and is defined as
H=U+pV
The first term U sums up the internal energy of the system. The second term pV is the
volume work done at the system or performed by the system.
The change of enthalpy is an important means of accessing chemical reactions.
Consider the above equation at constant pressure, which then simplifies to
dH = dU + p dV
The amount of energy emitted from or absorbed by a system during a change of state
is referred to as heat of reaction. The heat of reaction is characteristic for a chemical
reaction. It is given in kJ, although sometimes the unit used is kcal where 1 kcal =
4.187 kJ. The heat of reaction originating from 1 mol is referred to as the reaction
enthalpy ΔH.

Exothermic and Endothermic. Depending on whether or not a chemical reaction


creates or absorbs heat, the reaction is either classified as being exothermic if heat is
released or endothermic if heat is absorbed. By convention, for exothermic reactions
ΔH < 0 whereas for endothermic reactions ΔH > 0.
Helmholtz Free Energy
According to the second law of thermodynamics changes in an isolated system are
spontaneous if dS > 0, and the system is in equilibrium if dS = O. The Helmholtz free
energy, A, is a state function that defines the direction of changes in a closed system at a
constant temperature and a constant volume.
The Helmholtz free energy is defined by equation below, which shows that the
Helmholtz free energy is an extensive function of state that is defined by entropy, internal
energy, and temperature.
A =U-TS
The Helmholtz free energy of a closed system at a constant temperature and a constant
volume can be shown to decrease for changes that occur spontaneously. Therefore,
changes in closed systems may occur spontaneously if dA < 0, the systems are in
equilibrium or the changes are reversible if dA = 0, and the changes are forced if dA > O.
Gibbs Free Energy
A spontaneous Exothermic reactions generally are spontaneous at room temperature. The
process proceeds on enthalpy content decreases and excess energy is released to the surroundings.
its own.
This increases stability.
Spontaneous
Some endothermic reactions can take place spontaneously at room
processes lead to
lower energy and temperature. The enthalpy content of the system increases: this must also
increased stability. increase stability. The reason that such a reaction can take place is entropy.

A process is spontaneous if a chemical system becomes more stable and its


overall energy decreases. The overall energy decrease results from
contributions form entropy and enthalpy.

The entropy contribution to


the overall energy depends on
The free energy ΔG = ΔH - T ΔS
change ΔG is the
temperature. Energy derived balance between The feasibility of a reaction
from entropy = TΔS so as enthalpy, entropy and depends on the balance between
temperature increases, the temperature for a enthalpy and entropy.
energy derived from entropy process: Spontaneous processes
becomes more significant. happen when ΔG is negative.
Gibbs energy, G, is analogous to the Helmholtz free energy for changes that occur in
closed systems at a constant temperature and a constant pressure. Most changes in foods
occur at the atmospheric pressure and therefore at a constant pressure. The Gibbs energy
can be used to show whether changes occur spontaneously or if they are forced. The
Gibbs energy is an extensive function of state that is defined by Equation (1.16). Equation
(1.16) shows that Gibbs energy is defined by entropy, internal energy, pressure-
volume work, and temperature. Since enthalpy at a constant pressure is equal to the sum
of internal energy and pressure-volume work, the definition of Gibbs energy is also given
by Equation (1.17), which defines Gibbs energy to be a function of enthalpy, entropy, and
temperature.

G = U + pV – TS (1)
G = H - TS (2)

The Gibbs energy of a closed system at a constant temperature and a constant pressure
can be shown to decrease for changes that occur sponta-neously. Therefore, changes in
closed systems are spontaneous if dG < 0, the systems are in equilibrium or the changes
are reversible if dG = 0, and the changes are forced if dG > 0.
Difference Between Helmholtz free energy and Gibbs free energy

Helmholtz free energy Gibbs free energy


It is defined as the useful work It is defined as the maximum
that is obtained from a particular reversible work that is obtained
system from a particular system
It is the energy required to create a It is energy required to create a
system at constant temperature and system at constant pressure and
volume temperature
Helmholtz free energy finds lesser Gibbs free energy finds more
application as the volume of the application as the pressure of the
system should be constant system is constant
The thermodynamic square
“Good Physicists Have Studied Under Very Fine Teachers”
The thermodynamic square is mostly used to compute the
derivative of any thermodynamic potential of interest. Suppose
for example one desires to compute the derivative of the internal
energy U. The following procedure should be considered:
1. Place oneself in the thermodynamic potential of interest,
namely (G, H, U, F). In our example, that would be U.
2. The two opposite corners of the potential of interest
represent the coefficients of the overall result. If the
coefficient lies on the left hand side of the square, a negative
sign should be added. In our example, an intermediate result
would be dU = -p [Differential] + T [Differential].
3. In the opposite corner of each coefficient, you will find the
associated differential. In our example, the opposite corner to The thermodynamic square with
p would be V (Volume) and the opposite corner for T would potentials highlighted in red.
be S (Entropy).
In our example, an interim result would be: dU = -pdV + TdS
Notice that the sign convention will affect only the coefficients and NOT the differentials.
4. Finally, always add μ dN, where μ denotes the Chemical potential. Therefore, we
would have: dU = -pdV + TdS + μ dN
The three laws of thermodynamics encompass: the conservation of energy, the
relationship between heat and work, and the zero of entropy at the absolute zero of
temperature. Conservation of energy (the first law) is written as

ΔU = Q + ΔW (E.1)

where the change in internal energy of the system ΔU comprises the heat added to it Q
plus the work done on it ΔW. The partition of ΔU into Q and ΔW is not unique, but
depends on the way the system changes, e.g. from one temperature or volume to another,
i.e. it depends on the path, and is not a ‘function of state’. However, in a reversible, or
quasi-static, change these quantities are well determined, and we can define an
infinitesimal change in entropy, dS such that dQ = TdS and dW = - p dV, where p and T
are the pressure and temperature respectively.

Inserting these quantities into (E.1) leads to the equation for dU, which incorporates
the second law into the first as

dU = TdS - pdV (E.2)


This is the first equation involving a thermodynamic potential U, and can be used to
derive thermodynamic quantities at constant volume or constant entropy.
The above considerations apply to constant numbers of particles, N, but if N is itself a
variable, then there is an extra term in (E.2) +μdN, where μ is known as the chemical
potential. This is the internal energy at which the particle is added to the system. When
we are considering isolated systems at constant volume V and particle number N, the
thermodynamic potential U is constant, and

dU = TdS - pdV +μdN (E.3)

This is what is meant by the entropy being maximized (dS =0) in an isolated system at
equilibrium, the second law of thermodynamics having shown that dS ≥ dQ/T. The
corresponding microscopic distribution is the micro-canonical ensemble of statistical
mechanics. Consideration of constant pressure processes leads to definition of the
enthalpy, H = U + pV, so that

dH = TdS+Vdp + μdN (E.4)

the enthalpy is useful in describing thermally isolated constant pressure processes. Often
these correspond to (irreversible) flow processes, such as the flow of air over a wing or
through a nozzle valve.
Mostly we are concerned with systems at a given temperature, or in more technical terms
‘in contact with a heat bath at temperature T’. For these the Helmholtz free energy
F =U-TS and the Gibbs free energy G= U - TS + pV have been devised. These are
particularly useful for discussing processes at constant volume and constant pressure
respectively. Correspondingly we have
dF = -SdT-pdV +μdN (E.8)
and dG = -SdT+Vdp +μdN (E.9)
In particular, F is minimum at constant (T, V and N) and G is minimum at constant (T, p
and N), and the corresponding microscopic distribution is the Canonical ensemble.
The third law is concerned with establishing the zero of entropy for systems in
equilibrium at the absolute zero of temperature (0 K). Since physical effects result only
from differences in entropy and absolute zero cannot be reached, this may appear a bit
academic. However, it has a certain fascination in the context of phase changes,
especially since kinetics compete with thermodynamics, and can become very sluggish at
low temperature.
Thermodynamic laws and potentials
Thermodynamic potentials and the dividing surface
At a boundary between phases 1 and 2,
the concentration profile of any
elemental or molecular species changes
(continuously) from one level c1 to
another c2, as sketched in figure. Then
the extensive thermodynamic potentials
(e.g. the internal energy U, the
Helmholtz free energy F, or the Gibbs
free energy G) can be written as a
contribution from phases 1, 2 plus a
surface term. In the thermodynamics of
bulk matter, we have the bulk Helmholtz
free energy Fb = F(N1,N2) and we know Figure. Schematic view of the ‘dividing
that surface’ in terms of macroscopic
concentrations.

at constant temperature T, volume V and particle number N. In this equation, S is the


(bulk) entropy, p is the pressure and μ the chemical potential.
We are now interested in how the thermodynamic relations change when the system is
characterized by a surface area A in addition to the volume. With the surface present the
total free energy Ft = F(N1,N2,A) and

This fs is the extra Helmholtz free energy per unit area due to the presence of the surface,
where we have implicitly assumed that the total number of atomic/molecular entities in
the two phases, N1 and N2 remain constant. Gibbs’ idea of the ‘dividing surface’ was the
following. Although the concentrations may vary in the neighbourhood of the surface,
we consider the system as uniform up to this ideal interface: fs is then the surface excess
free energy.
To make matters concrete, we might think of a one-component solid–vapor interface,
where c1 is high, and c2 is very low; the exact concentration profile in the vicinity of the
interface is typically unknown. Indeed, it depends on the forces between the constituent
atoms or molecules, and the temperature, via the statistical mechanics of the system.
But we can define an imaginary dividing surface, such that the system behaves as if it
comprised a uniform solid and a uniform vapor up to this dividing surface, and that the
surface itself has thermodynamic properties which scale with the surface area; this is
the meaning of equation 2. In many cases described in this book, the concentration
changes from one phase to another can be sharp at the atomic level. This does not
invalidate thermodynamic reasoning, but it leads to an interesting dialogue between
macroscopic and atomistic views of surface processes.
Surface energy
Surface free energy or interfacial free
energy or surface energy quantifies the
disruption of intermolecular bonds that
occurs when a surface is created.
Surfaces must be intrinsically less
energetically favourable than the bulk of a
material (the molecules on the surface have
more energy compared with the molecules
in the bulk of the material), otherwise there
would be a driving force for surfaces to be
created, removing the bulk of the material.
The surface energy may therefore be
defined as the excess energy at the surface
of a material compared to the bulk, or it is
the work required to build an area of a
particular surface.
Another way to view the surface energy is
to relate it to the work required to cut a bulk
sample, creating two surfaces.
Surface tension and surface energy

The surface tension, γ, is defined as the reversible


work done in creating unit area of new surface, i.e.

In the simple illustration of figure

At const T and V,

Therefore,
Figure. Schematic illustration of how to create
new surface by cleavage. If this can be done
reversibly, in the thermodynamic sense, then
the work done is 2γA.
The simple example leading to (1.6) shows that care is needed: if a surface is created, the
atoms or molecules can migrate to (or sometimes from) the surface. The most common
phenomena of this type are as follows.
1. A soap film lowers the surface tension of water. Why? Because the soap molecules
come out of solution and form (mono-molecular) layers at the water surface (with
their ‘hydrophobic’ ends pointing outwards). Soapy water (or beer) doesn’t mind
being agitated into a foam with a large surface area; these are examples one can
ponder every day
2. A clean surface in ultra-high vacuum has a higher free energy than an oxidized (or
contaminated) surface. Why? Because if it didn’t, there would be no ‘driving force’
for oxygen to adsorb, and the reaction wouldn’t occur. It is generally true that the
surface energy of metal oxides are much lower than the surface energy of the
corresponding metal.
Surface energy is a function of lost interactions

CN (Bulk) = 12
CN (Surf) = 4+4
Lost = 4

CN (Bulk) = 12
CN (Surf) = 2+4+1
Lost = 5
CN (Bulk) = 12
CN (Surf) = 6+3
Lost = 3 CN Loss: (111) = 3, (100) = 4, (110) = 5

γ110 > γ100 > γ111


The terrace–ledge–kink model
The Terrace Ledge Kink model (TLK), which is also referred to as the Terrace Step Kink
model (TSK), describes the thermodynamics of crystal surface formation and
transformation, as well as the energetics of surface defect formation. It is based upon the
idea that the energy of an atom’s position on a crystal surface is determined by its bonding
to neighbouring atoms and that transitions simply involve the counting of broken and
formed bonds. The TLK model can be applied to surface science topics such as crystal
growth, surface diffusion, roughening, and vaporization because it consider the two major
points about the surface:
1. The energy of an atom’s position on a
crystal surface is determined by its
bonding to neighbouring atoms;
2. Phase growth or transition simply
involve the counting of broken and
formed bonds.
The names for the various atomic positions in the
The TSK model was originally proposed by TLK model. This graphic representation is for a
Kossel and Stranski. simple cubic lattice.

The energy required to remove an atom from the surface depends on the number of
bonds to other surface atoms which must be broken. For a simple cubic lattice in this
model, each atom is treated as a cube and bonding occurs at each face, giving
a coordination number of 6 nearest neighbours.
Classification of Surfaces by their Defects (or Imperfections)
---Terrace, Ledges, Kinks and Adatoms

Terrace
• Surface having
crystalline order

Ledge
• Steps formed at the
border of terraces

Kink
• Defect formed at
the end of ledges

Adatom
• Single Atom sitting
on a terrace or ledge
surface
Surface Defects (or Imperfections)
---Terrace, Ledges, Kinks and Adatoms

Terrace
• Terrace atom has 5
nearest neighbors

Ledge
• Ledge atom has 4
nearest neighbors

Kink
• Kink atom has 3
nearerst neighbors
Adatom
• Ledge adatom has 2
nearest neighbor
• Terrace adatom has 1
nearest neighbor
Bulk energy < surface energy < step energy < kink or adatom energy

Ball model representation of a real (atomically rough) crystal


surface with steps, kinks, adatoms, and vacancies in a closely
packed crystalline material. Adsorbed molecules, substitutional and
interstitial atoms are also illustrated
The kink site is of special importance when
evaluating the thermodynamics of a variety of
phenomena. This site is also referred to as the
“half-crystal position” and energies are
evaluated relative to this position for processes
such as adsorption, surface diffusion, and
sublimation. The term “half-crystal” comes
from the fact that the kink site has half the The names for the various atomic positions in the
TLK model. This graphic representation is for a
number of neighbouring atoms as an atom in simple cubic lattice.
the crystal bulk, regardless of the type of
crystal lattice.
For example, the formation energy for an
adatom is calculated by subtracting the energy
of an adatom from the energy of the kink atom.

ΔG = Ekinck – Eadatom

Site Stability has direct proportionality to binding energy.

Terrace
The higher the binding energy, the higher the site stability.

Binding Energy for Atoms at Various Sites


This can be understood as the breaking of all of the kink atom’s bonds to remove the atom
from the surface and then reforming the adatom interactions.
This is equivalent to a kink atom diffusing away from the rest of the step to become a step
adatom and then diffusing away from the adjacent step onto the terrace to become an
adatom.
In the case where all interactions are ignored except for those with nearest neighbors, the
formation energy for an adatom would be the following, where ϕ is the bond energy in
the crystal is given by Equation 2.

ΔG = Ekinck – Eadatom = 3 ϕ - ϕ = 2 ϕ

Temperature dependence of defect coverage


The number of adatoms present on a surface is temperature dependent. The relationship
between the surface adatom concentration and the temperature at equilibrium is described
by equation 4, where n0 is the total number of surface sites per unit area:
∆𝐺𝑎𝑑𝑎𝑡𝑜𝑚
𝑛𝑎𝑑𝑎𝑡𝑜𝑚 = 𝑛0 exp(− )
𝑘𝑏 𝑇

This can be extended to find the equilibrium concentration of other types of surface point
defects as well. To do so, the energy of the defect in question is simply substituted into the
above equation in the place of the energy of adatom formation.
An STM image of the Si(001)
surface. There is a step in the
bottom right corner.
Wulff construction: definition and basic properties
In his famous paper of 1874 ”On the Equilibrium of Heterogeneous Substances”, J.
Willard Gibbs concluded that a given quantity of matter will attain a shape such that the
total surface energy is minimal. For perfect crystalline solids, atomic planes are members
of a countable set characterized by integer Miller indexes (hkl). The shape of a crystalline
solid will therefore be a polyhedron for which only faces parallel to (hkl) planes are
allowed. It is only reasonable to assume that faces with a relatively low surface energy will
dominate the equilibrium shape.
Several decades later, Georg Wulff suggested that the
polyhedron that corresponds to the lowest surface
energy of a crystalline substance can be constructed
in the following way (the so-called Wulff
construction):
One chooses a constant c, and a Cartesian set of axes.
Starting from the origin, O, one draws a plane that is
normal to the [hkl] vector and has a distance dhkl =
c·γhkl from O. The quantity γhkl is the energy required
to create a surface of unit area normal to the [hkl]
vector, and is the analogous of the surface tension for
liquids. This process is repeated for all sets of Miller
indexes, (hkl). The space that lies inside all these
planes defines the equilibrium shape for this material.
Wulff construction
Liquids and amorphous solids have isotropic g. Therefore they have
spherical shape.
For crystalline material, g depends on the orientation (hkl) of the surface,
while the broken bonds and charge compensation per unit area may be
different. g (n) is then the function of orientation of the surface, n. It
must have:

When we plot g (n) as function of n (Wulff plot), theoretically one can


determine the equilibrium shape of a solid.
minimum
g(hkl) as function of q (the angle
between the normal directions of the
surface to the {hkl} planes. The
Wulff plot gives the shape of the solid
(dashed dotted) as the inner envelope
of the Wulff planes (broken lines).
The energy of a surface and its dependence on crystallographic orientation can determine
the equilibrium shape of the particle. Indeed, the particle will arrange itself to minimize its
surface energy. The determination of the particle’s shape based on the minimization of the
Gibbs free energy of the surface is a method known as the “Wulff construction.”

Of course when γ is isotropic, as for liquid droplets, both the γ -plot and equilibrium
shapes are spheres.
A convenient method for plotting the variation of γ with surface orientation in three
dimensions is to construct a surface about an origin such that the free energy of any plane
is equal to the distance between the surface and the rigin when measured along the normal
to the plane in question. A section through such a surface is shown in Fig. a. This type of
polar representation of γ is known as a γ -plot and has the useful property of being able to
predict the equilibrium shape of an isolated single crystal.
For an isolated crystal bounded by several
planes Al, A2, etc. with energies γ1, γ2, etc.
the total surface energy will be given by
γ1Al + γ1Al +…..
The equilibrium shape has the property
that ∑ γiAi is a minimum and the shape
that satisfies this condition is given by the
Wulff construction.
For every point on the γ surface, such as A
in Fig. a, a plane is drawn through the
point and normal to the radius vector OA.
The equilibrium shape is then simply the
inner envelope of all such planes.
Therefore when the γ -plot contains sharp
cusps the equilibrium shape is a
polyhedron with the largest facets having
the lowest interfacial free energy.
Gold and Silver Cubes
{111} capped, Gold nanocrystals

Bi-saturated Cu "negative" crystal


Real Crystal Surfaces
Surface Relaxations
The effect of surface relaxation is the simplest modification observed for real surfaces.
It assumes that the (h k l) surface of a substrate, whose bulk lattice is given by a
netplane-adapted lattice vectors R1, R2, R3, is terminated by overlayers forming (h k l)
monolayers, identical with those of the substrate. However, relative positions of the
overlayer atoms near the surface, expressed by interlayer distances and lateral shifts,
deviate slightly from corresponding positions in the bulk.
This is described in the simplest
case by complete overlayers
shifting slightly with respect to
their bulk positions. In most
cases that are observed in
experiments these shifts occur
perpendicular to the surface,
either toward (inward relaxation)
or away from the substrate
(outward relaxation).
Hypothetical (0 0 1) surface section of a crystal with simple cubic
lattice (lattice constant a) with the two topmost overlayers relaxed
Thus, formally atom positions of relaxed overlayers near the surface are described by

where i = 1, 2, 3, ...p, & n1, n2


R(m) = ri + n1 R1 + n2 R2 + n3R3 + s(m)
integer for layer m near the surface

where ri refers to positions of atoms in the unit cell of the bulk lattice, n1, n2 are integer-
valued coefficients accounting for the overlayer (netplane) periodicity, and s(m) is a shift vector
corresponding to the absolute positioning of layer m.
As an illustration, Figure shows
the (0 0 1) surface of a crystal
with simple cubic lattice
(lattice constant a), where the
topmost layer 1 is relaxed
inward by 10% and shifted
sideways by vector v and layer
2 is relaxed inward by 30%.

Hypothetical (0 0 1) surface section of a crystal


with simple cubic lattice (lattice constant a) with
the two topmost overlayers relaxed
Surface Structures and Reconstructions
Surface Structures

Real surfaces that differ structurally from simple bulk truncations other than described
by relaxation are described as reconstructed surfaces. Reconstruction may result in
surface disorder or may yield a periodic surface geometry with sizable displacements of
the atoms. Furthermore, additional or fewer atoms may exist in the
layer unit cells compared with bulk truncation.
Many materials, notably metals, have
a surface lattice, which corresponds to
the bulk crystallographic (hkl) plane.
Merely the atomic distances vertical
to the surface plane are changed to a
larger or lesser degree, depending on
the material, the surface orientation,
and the type of bonding. The surfaces
of some 5d-transition metals,
however, reconstruct to form large,
sometimes even incommensurate
surface cells. Reconstructions are also Figure: Ideal (left) and the (2×1) reconstructed Pd(1 1 0)
surface (right). The layer periodicity vectors are indicated
typical for covalently bonded
for both geometries
semiconductors.
Face Centered Cubic (fcc) Structure
• Many metal elements crystallize in the face-centered cubic (fcc) structure. Among them
are the coinage metals copper (Cu), silver (Ag), gold (Au), as well as the catalytic
important metals nickel (Ni), rhodium (Rh), palladium (Pd), iridium (Ir) and platinum (Pt).
• Following the convention in crystallography, we denote a set of equivalent faces by braced
indices, e.g. {100}, and particular faces like (100), (010), or (001) by indices in
parenthesis.
• The three most densely
packed, and therefore the
most stable {111}, {100},
and {100} surfaces of
unreconstructed fcc-
crystals are depicted in
Fig. The packing density is
the highest for the {111}
surfaces, followed by the
{100} and {110} surfaces.
Figure: {111}, {100}, and {100} surfaces of fcc-
• The coordination numbers of surface atoms crystals; bottom row displays side views.
are 9, 8 and 7 for the {111} , {100} and
{110} surfaces, hence the number of broken
bonds are 3, 4 and 5 per surface atom.
The open structure of the {110} surface has the peculiar feature that atoms in the surface
layer have nearest neighbour bonds not only to the next, but also to the third layer. Vice
versa, the second layer atoms have one broken bond oriented perpendicular to the
surface plane. Hence, this surface has 5 broken bonds per surface atom, but 6 broken
bonds per surface unit cell. In a nearest neighbour model, each broken bond on a surface
corresponds to 1/12 of the cohesive energy Ec. Accordingly, the surface energies per
surface unit cell are Ec /4, Ec /3, and Ec /2, for the {111}, {100} and {110} surface,
respectively. Since the atom packing density also decreases in that sequence, the three
surfaces differ by a lesser amount in their surface energy per area. In units of Ec /a2, in
which ao is the lattice constant, the energies of the {111}, {100} and {110} surfaces are
0.577, 0.666 and 0.707, respectively. The actual differences between the surface
energies are even smaller because of next nearest neighbour and many-body
contributions to the surface energy.
The surface layer of the {111} surface has a six-fold rotation axis and three non-trivial
mirror planes. Together with the second layer underneath the symmetry reduces to a
three-fold rotation axis. The highest symmetry of a molecular species site on that
surface is therefore C3v. However, if the adsorbate species has a sixfold rotation axis and
interacts only with the first layer atoms the effective point group symmetry is C6v. The
{100} surfaces have four-fold symmetry and two nontrivial mirror planes. The highest
symmetry of an adsorbate is thus C4v. The {110} surface has a two-fold axis and two
mirror planes. The highest point group symmetry is C2v.
Unreconstructed surfaces as depicted in Figure above are found on α-cobalt (α-Co), Ni,
Cu, Rh, Pd, and Ag. Atoms in the surface layer assume a position as in the bulk save for
a possible relaxation of the vertical distance between the surface layer and the layer
underneath. This relaxation is very small (1-2%) for the {100} and {111} surfaces,
hardly outside the error of the best structure determinations. The relaxation is larger for
the {110} surfaces, and even the distance between the second and the third layer differs
notably from the bulk.
The reconstructed {100} surfaces of Ir, Pt and Au all involve a nearly hexagonal packing
of atoms in the surface layer. For Iridium this leads to a (5×1) reconstruction (Fig.) so
that the density of atoms in the surface layer is 6/5 of the unreconstructed surface.

Fig. Top and side view of the (5×1) reconstructed Ir(100) surface in which the surface layer consists of a
buckled quasi-hexagonal overlayer of atoms. The buckling depends on the lateral position of the surface
atom with respect to the second layer atoms and amounts to 0.48 Å at the maximum [1.18]. The dashed
rectangle indicates the unit cell. The {100} surfaces of platinum and gold feature the same quasi-
hexagonal arrangement of atoms in the first layer, but the surface layer is more densely packed and
incommensurate with the substrate.
Body Cubic Centered (bcc) Structure
Typical metals with bcc-structure are tungsten (W), molybdenum (Mo), niobium (Nb),
and iron (Fe). Spurred by the interest in their use as thermionic electron emitters, surfaces
of tungsten have drawn the attention of researchers. The bulk-terminated surfaces of bcc-
crystals are displayed in Figure. The atoms form a compressed hexagon with each atom
surrounded by four atoms in nearest neighbour distance, and two atoms in the 15.5%
larger second nearest neighbour distance. The {111} surfaces have a very open structure.
Atoms in three layers are
missing nearest neighbours.
Since the distance to the
second nearest neighbours
is merely slightly larger
than to the first neighbours,
an estimate of the surface
energies based on the
coordination numbers is not
meaningful.

Figure: Top and side view of the {110}, {100}, and {111} surfaces of a
bulk terminated bcc-structure. The very open {111} surface is formed
by three layers of atoms that are missing some of their nearest neighbor
bonds.
Diamond, Zincblende and Wurtzite
• The group IV-elements carbon, silicon and germanium crystallize in the diamond
structure in which each atom is surrounded by a tetrahedron of neighboring atoms,
providing optimum overlap of the sp3-type covalent bonds.
• The diamond structure can be viewed as two fcc-structures displaced along the cubic
space diagonal by a vector (1/4, 1/4, 1/4)a0 with a0 the lattice constant (Fig. a). The
structure has its name from the diamond phase of crystalline carbon although diamond
is not the most stable phase of carbon, which is graphite.
The III-V and II-VI
compounds are likewise
primarily covalently bonded
in a tetrahedral
configuration. The III-V
compounds and some of the
II-VI compounds crystallize
in the diamond structure
with each of the two atoms
of the compound occupying
one of the fcc-substructures.
The structure is then named Fig. Structure of (a) zincblende and (b) wurtzite. The zinblende structure
zincblende, after the mineral reduces to the diamond structure if A- and B- atoms are identical. The
zincblende structure has eight the wurtzite structure 4 atoms in the unit
name of the II-VI compound cell.
ZnS.
Diamond, Zincblende and Wurtzite
• A ZnS-crystal has four polar axes oriented along the tetrahedral bonds. A dipole
moment can arise if the tetrahedral symmetry is distorted, e.g. by shear stresses.
• Most of the II-VI compounds crystallize in the hexagonal wurtzite structure. In
wurtzite, the local configuration is as in zincblende (Figure b).
• The arrangement of the tetrahedrons in space differs, however. When build with ideal
tetrahedrons, wurtzite has a c/a ratio of (8/ 3)1/2 = 1.633 .
However, the symmetry of
the structure is compatible
with the tetrahedrons being
distorted along the c-axis.
Since the c-axis is a polar
axis, wurtzite crystals are
pyroelectric (pyroelectricity
denotes a variation of a
permanent polarization with
temperature), and possess
one non-zero diagonal and
off-diagonal elements of the
piezoelectric tensor. Fig. Structure of (a) zincblende and (b) wurtzite. The zinblende structure
reduces to the diamond structure if A- and B- atoms are identical. The
zincblende structure has eight the wurtzite structure 4 atoms in the unit
cell.
Because of the covalent nature of the bonding (with some ionic character in the III-V and
II-VI-compounds) the termination of the bulk structure at the surface means broken bonds,
also called dangling bonds. To minimize the energy associated with the dangling bonds
nearly all surfaces of the group IV elements and of the III-V and II-VI compounds
reconstruct in one or another way. In order to be able to describe and understand nature of
the various reconstructions involved it is necessary to know the reference frame of the low
index bulk terminated structures. We therefore depict the surfaces as they arise from the
truncated bulk structures of zincblende and wurtzite, before entering the discussion
concerning reconstructions.

Fig. Top and side view of the low index surfaces of the zincblende structure. Pictures also represent the surfaces of the
diamond structure if the dark and lightly shaded atoms are identical. For zincblende the ideal (111) surfaces are polar, as
the surface layer consists of one type of atoms. Full lines indicate the boundaries of the (111), (100) and (110) planes as
drawn into the bulk cubic cell. The dashed lines are the surface unit cells.
Figure 1.18 shows top and side views of the {111}, {100} and {110} surfaces of the
zincblende structure, as they arise from the truncated bulk structure. The surface layer of a
{111} surface consists of only one type of atoms and has therefore a polar character. The
(111) and 1 1 1 surfaces are not identical. On {111} and {110}, surfaces atoms have one
dangling bond, on {100} surfaces each surface atom has two. The illustrations in Fig. 1.18
represent the surfaces of the diamond structure when dark and light atoms represent the
same element.

Fig. Top and side view of the low index surfaces of the zincblende structure. Pictures also represent the surfaces of the
diamond structure if the dark and lightly shaded atoms are identical. For zincblende the ideal (111) surfaces are polar, as
the surface layer consists of one type of atoms. Full lines indicate the boundaries of the (111), (100) and (110) planes as
drawn into the bulk cubic cell. The dashed lines are the surface unit cells.
Figure shows top and side views of two surfaces of wurtzite. As for the {111} surfaces of
zincblende, the surface layer of the {0001} surfaces consist of atoms of one type; the surfaces
are therefore polar. Because of the arrangement of the tetrahedrons, wurtzite appears as rather
open when viewed along the c-axis, compared to the zincblende and diamond structure.

Fig. Top and side view of surfaces of wurtzite surfaces


The Si(111)-7×7 reconstruction
A (111) slice through a Si crystal exposes a high density of dangling bonds that extend into
the vacuum. The surface can reduce the number of dangling bonds by a complicated
atomic reordering process involving adatoms, a stacking fault, dimerisation, and a deep
vacancy, which all co-exist to form the (7x7) reconstruction. Fig. a shows with the (7x7)
surface unit cell indicated on the image. Numerous single atom defects can be seen in the
image. A higher magnification image of the atomic structure is shown in Fig. b where it
can be seen that only 12 bright spots per unit cell are imaged. These spots are the adatoms
that form the outermost atomic layer of the silicon crystal.

Fig. STM-image of a Si(111)-(7×7)


surface. Dashed lines mark the unit
cell. The image shows twelve bright
spots and one deep and wide hole per
unit cell. The bright spots correspond
to silicon adatoms bonding to three
dangling surface bonds.
A model of this new 7×7 surface structure is shown below. Color-coding indicates different
types of Silicon atoms from the top three atomic layers that are involved in forming the 7×7
reconstruction. The "Dimer-Adatom-Stacking fault" (DAS) model (Takayanagi/Tong)
indicates that the most prominent features observed in images correspond
to adatoms (shown in pink).
This model also clearly
shows corner-holes, i.e., areas
at the corners of the 7×7 unit
cells where atoms from the top
three layers are missing. Corner-
holes appear in STM images as
dark areas surrounded by
hexagons of adatoms or, on a
larger scale, by six triangles,
each triangle, in turn, composed
of six adatoms.

Figure: Geometry of the reconstructed Si(1 1 1)-(7×7) surface according to the dimer-adatom-stacking- fault (DAS)
model. The overlayer is removed at the bottom right to reveal the ideal bulk termination of Si(1 1 1). The
periodicity vectors of the overlayer and of the ideal bulk termination are sketched accordingly.
Experimental Analysis of Real Crystal Surfaces
• Truly quantitative structure determinations of single crystal surfaces and adsorbate
systems by experiment are intrinsically difficult.
• Methods that can contribute to a quantitative analysis of geometric details of real crystal
surfaces must be able to probe mainly atoms near the surface, ignoring those of the inner
substrate.
• This excludes standard X-ray diffraction methods from surface analyses. X-ray photons
can penetrate deep into the bulk and, therefore, yield structural bulk information with that
from surface atoms representing only a minor perturbation.
• Diffraction methods, such as low-energy electron diffraction (LEED), have proven to be
particularly useful in identifying surface structure.
• These methods rely on the interference of particles that scatter (often multiple times) from
periodic arrangements of atoms at single crystal surfaces that are ordered over a relatively
wide area.
• If the surface geometry deviates strongly from periodic ordering, for example, as a result
of global imperfections or disordered adsorbate structure, local (small-area) diffraction
becomes more useful. For these systems, local diffraction methods, such as photoelectron
diffraction (PED) or surface extended X-ray absorption fine structure (SEXAFS), can be
used to obtain quantitative information about local environments of surface atoms
including coordination and binding angles.
• In general, methods that can provide information about geometric details at real crystal
surfaces and adsorbate systems include those based on scattering, diffraction, imaging,
and spectroscopy and use photons, electrons, or atoms and ions (Table 1, next slide).
Table: Experimental methods used to determine surface structure.
The Scanning Probe Microscope

What are the basic components of a SPM


Localized Probe that has A nano-positioning mechanism
an “interaction” with the that can position the probe in
substrate to be imaged “close proximity” of the surface

SPM
A system to measure the A mechanism to scan the
interaction of the probe probe relative to the substrate
with the substrate and measure the interaction
as function of position

56
Scanning tunnelling microscope
(STM)
➢ Invented by Binnig and Rohrer at IBM in 1981 (Nobel Prize in Physics in 1986).

➢ Binnig also invented the Atomic Force Microscope(AFM) at Stanford University


in 1986.

57
Classical vs. Quantum
Classically, if one threw a ball at a brick wall (and didn’t miss horribly) they would always find the
ball on their side of the wall, because unless you are Superman there is no way that you are throwing
the ball hard enough to break through the wall.

Some later
time

Quantum mechanically, however, the ball has some finite probability of “tunneling” through the wall.
The ball need not be thrown incredibly hard to achieve this, although the more energy the ball has the
more likely it will be to tunnel.

Some later
time
58
How Does It Work??

Chen, C.J. In Introduction to Scanning Tunneling Microscopy; Oxford University Press: New York, 1993; p 3.
p 2x
In classical mechanics, the energy of an electron moving in a potential U(x) can be shown by + U( x) = E
2m
The electron has nonzero momentum when E > U(x), but when E<U(x) the area is forbidden.

  ( x ) + U x  ( x ) = E ( x )
The quantum mechanical description of the same electron is H

 ikx 2m( E − U )
In the classically allowed region (E>U), there are two solutions,  ( x ) =  ( 0 )e , where k =

These give the same result as the classical case. However, in the classically forbidden region (E<U) the solution is
2 m ( U − E)
 ( x ) =  ( 0)e− x , where  =

 is a decay constant, so the solution dictates that the wave function decays in the +x direction, and the probability
of finding an electron in the barrier is non-zero. 59
When two conductors are separated by a macroscopic vacuum gap, then electrons will not travel
from one to the other. If however, the conductors are brought into a proximity of less than a few
nanometres, then their quantum mechanical wave functions overlap to the extent that will allow
electrons to be shared between them. Classically this interaction is not allowed because the
electrons do not have sufficient kinetic energy to overcome the vacuum barrier. The movement
of electrons across the barrier is therefore known as tunnelling because the electrons “tunnel”
through the vacuum barrier.
In an STM the tunnelling current that flows between the sharp metallic tip and the sample is the
central mechanism that allows the microscope to function. Figure is a diagram showing the tip
and sample atoms. The tunnelling current density has been shown in gray and is most intense
where the tip and sample are closest. The tunnelling current can be described as
I α V ϕ2 ρs ρt e-kd

Where It is the tunnelling current, V is the voltage applied between the sample and the tip, ϕ is
the average barrier height, ρs is the density of sample states, ρs is the density of tip states, k is a
constant related to the decay length for the wave functions in a vacuum, and d is the tip –sample
60
separation.
61
Piezolelectric Tube
with Electrodes
Tunneling
Current Amplifier
Distance Control
and Scanning Unit

Sample Tunneling Voltage

Data Processing
Tip and Display

Sample

62
The inverse exponential relationship between distance and tunnelling current mean that
the atom at the apex of the tip is the one from which the overwhelming majority of the
electrons tunnel. In turn, the electrons will only tunnel into the closest sample atoms, and
this is the reason why STM images show only the structure of the topmost atomic layer of
a sample. The term “surface sensitive” for the STM literally means the top monolayer of
surface atoms.
The relative energy levels of the tip and sample are shown in the diagrams of Figure
below. In Figure a the tip and sample are separated by a macroscopic distance. The
exponentially decaying tip wavefunction is shown, and the local density of electronic
states of the sample are indicated. When the tip and sample are brought within a few
nanometres of each other electrons can tunnel between them, which allows the Fermi
levels to equalise as shown in Figure b. When a steady state has been reached as many
electrons tunnel from the tip into the sample as the other way around.

63
Diagrams of the relative energy levels of the STM tip and sample. In (a) the tip and sample do not
interact and the vacuum levels (Evac) are the same. In (b) the small vacuum gap allows tunnelling
between tip and sample leading to equalisation of the Fermi Energies (EF). In (c) the sample is biased
positively resulting in a tunnelling current from the tip into the empty electronic states of the sample.
In (d) the sample is biased negatively causing electrons from filled sample states to tunnel into the tip .

64
If the sample is biased by a positive voltage Vs with respect to the tip, then the energy levels
of the tip will move up by eVs. This situation is drawn in Fig.c. Electrons tunnel from the
filled tip states into the empty sample states (indicated in gray), and a dc tunneling current is
established. The magnitude of the tunneling current depends amongst other things on the
local density of empty states of the sample. By biasing the sample negatively with respect to
the tip the tunnelling current flows in the opposite direction, as shown in Fig.d. For a
negatively biased sample the tunnelling current is a function of the filled density of sample
states.
Once a tip and sample have formed a suitable tunnel junction, the tip can be scanned in 3D to
allow an image to be created. Scanning the tip is achieved by attaching it to piezoelectric
crystals that change their length when a voltage is applied along them. The simplest form of
scanning involves having three separate piezoelectric crystals responsible for movements in
the x, y, and z directions. The great advantage of piezoelectric scanners is that their
expansion or contraction, and hence the movement of the tip, can be controlled very
accurately.
65
The electrons fill up the energy valley in the sample until
there are no more electrons. The top energy level at which The electrons in the tip and the sample are sitting
electrons sit is called the Fermi level, εF. For every energy ε, in two separate valleys, separated by a hill which is
the density of states is the number of electrons sitting within the vacuum barrier.
Δε of ε, divided by Δε.

By applying a bias voltage to the sample


with respect to the tip, we effectively raise
the Fermi level of the sample with respect to
the tip. Now we have empty states available
for tunneling into.

66
There are two main modes to create a surface topography map. The first, is to scan the
tip in a plane over the surface and record the tunnelling current as a function of x and y.
This mode is called constant height mode, and the image is displayed with the brightest
regions being those where the tunnelling current is greatest. This mode is sometimes
used for rapid scanning over very flat surfaces. The second mode, and by far the most
widely used, is the constant current mode. In this operating mode the tunnelling current
is determined by the user, and the tip height is adjusted by the computer to maintain a
constant tunnelling current. When the tip is scanned, the height of the tip is recorded,
and the higher the tip the brighter that region in the image. The constant current mode
can operate over much greater z distances than the constant height mode, and it is this
increased dynamic range that makes it the more usual choice for STM operation.

67
68
So What Do We See?
Atomic resolution imaging of surface reconstructions

When a crystal is terminated and forms a surface the bonding arrangement of the surface
atoms differs significantly from those in the bulk. The result is that the location of the
surface atoms changes, and when this happens the surface is termed “reconstructed”. The
periodicity of the new arrangement of the surface atoms give rise to a surface unit cell
which is defined as the multiple of the unit cell of a bulk termination of the crystal. A
good example of a reconstruction is the Si (111) surface, as illustrated in Figure

69
STM images of the Si (111)-(7×7) reconstruction. In (a) an image of a terrace is shown. The (7×7) unit
cell is indicated and individual point defects can be seen. The image width is 21 nm. In (b) a high
magnification empty states image is shown that was taken at a sample bias of +2 V. All adatoms are
equally bright. Panel (c) shows a filled states image taken at a sample bias of -2 V. The adatoms in half
the unit cell are brighter than in the other half. Image widths in (b) and (c) are 4.7 nm.
70
STM image of the Pt (001)-hex-R0.7° reconstructed surface. The corrugated topography of
the surface and single atom vacancies can be seen. The image size is 10×10 nm2.
71
STM images showing MoS2 nanoclusters on an Au (111) surface. Panel (a) shows the
distribution of clusters (image size 75 nm x 75 nm), and panel (b) is a close-up of a cluster
after exposure to hydrogen which results in S vacancies as indicated with circles. In (b)
the bright edge around the rim of the cluster is due to an increased local density of states.

72
Ag deposition onto a substrate gives rise to
five-fold symmetry nanocrystals

73
Graphene reveals its honeycomb structure
Gold (Au)
made up of rings of carbon atom

74
Molecules

75
Nanomanipulation
Quantum Corrals
Fe atoms on Cu(111)
Nanomanipulation
• Quantum Corrals are fabricated by manipulating
atoms adsorbed at a solid surface to give a specific
shape to the corral.
• The STM tip is used to lift and put down the atomic
units.
• Peculiar effect related to Quantum Corrals
Formation of a two-dimensional electronic gas
(standing waves) confined within the corral.

76
Atomic Force Microscopy (AFM)
Atomic Force Microscopy (AFM) is a form of scanning probe microscopy (SPM) where a
small probe is scanned across the sample to obtain information about the sample’s surface.
The information gathered from the probe’s interaction with the surface can be as simple as
physical topography or as diverse as measurements of the material’s physical, magnetic, or
chemical properties. These data are collected as the probe is scanned in a raster pattern across
the sample to form a map of the measured property relative to the X-Y position. Thus, the
AFM microscopic image shows the variation in the measured property, e.g,. height or
magnetic domains, over the area imaged.
The AFM probe has a very sharp tip, often less than 100 Å diameter, at the end of a small
cantilever beam. The probe is attached to a piezoelectric scanner tube, which scans the probe
across a selected area of the sample surface. Interatomic forces between the probe tip and the
sample surface cause the cantilever to deflect as the sample’s surface topography (or other

77
properties) changes. A laser light reflected from the back of the cantilever measures the
deflection of the cantilever. This information is fed back to a computer, which generates a
map of topography and/or other properties of interest. Areas as large as about 100 µm
square to less than 100 nm square can be imaged.

78
Parts of AFM

• 1. Laser – deflected off cantilever


• 2. Mirror –reflects laser beam to
photodetector
• 3. Photodetector –dual element
photodiode that measures differences
in light intensity and converts to
voltage
• 4. Amplifier
• 5. Register
• 6. Sample
• 7. Probe –tip that scans sample
made of Si
• 8. Cantilever –moves as scanned
over sample and deflects laser beam

79
Topography
• Contact Mode
– High resolution
– Damage to sample
– Can measure
frictional forces
• Non-Contact Mode
– Lower resolution
– No damage to sample
• Tapping Mode
– Better resolution
– Minimal damage to
sample
80
Figure describes both the
attraction and repulsion between
nonionic particles. The first part
of the equation, (σ/d)12 describes
the repulsive forces between
particles while the latter part of
the equation, (σ/d)6 denotes
attraction.

81
Contact Mode

• Measures repulsion between tip and sample


• Force of tip against sample remains constant
• Feedback regulation keeps cantilever deflection constant
• Voltage required indicates height of sample
• Problems: excessive tracking forces applied by probe to
sample

82
83
Non-Contact Mode

• Measures attractive forces between tip and sample


• Tip doesn’t touch sample
• Van der Waals forces between tip and sample detected
• Problems: Can’t use with samples in fluid
• Used to analyze semiconductors
• Doesn’t degrade or interfere with sample- better for
soft samples

84
Tapping (Intermittent-Contact) Mode

• Tip vertically oscillates between contacting sample surface


and lifting of at frequency of 50,000 to 500,000 cycles/sec.
• Oscillation amplitude reduced as probe contacts surface due
to loss of energy caused by tip contacting surface.
• Advantages: overcomes problems associated with friction,
adhesion, electrostatic forces.
• More effective for larger scan sizes.

85
Force Measurement

• The cantilever is designed with a


very low spring constant (easy to
bend) so it is very sensitive to force.

• The laser is focused to reflect off the


cantilever and onto the sensor

• The position of the beam in the


sensor measures the deflection of the
cantilever and in turn the force
between the tip and the sample.

86
Generating an Image

• The tip passes back and forth in a


straight line across the sample (think
old typewriter or CRT)
Scanning Tip

• In the typical imaging mode, the tip-


sample force is held constant by
adjusting the vertical position of the
tip (feedback).
Raster Motion

• A topographic image is built up by the


computer by recording the vertical
position as the tip is rastered across
the sample.
Scanning the Sample

• Tip brought within nanometers of


the sample (van der Waals)
• Radius of tip limits the accuracy of
analysis/ resolution
• Stiffer cantilevers protect against
sample damage because they
deflect less in response to a small
force
• This means a more sensitive
detection scheme is needed
• measure change in resonance
frequency and amplitude of
oscillation

88
STM vs AFM
AFM STM
no requirements sufficiently conductive sample

atomic resolution possible atomic resolution standard


but hard to get

local electrical information local electrical information and


independent of topography topography not separable

Contact not well defined defined tunneling via single atom

Chemical & mechanical information xxx


89
Examples

90
The Bad Examples
Histogram shows level surface, but scan is Typically the sample will have a slight tilt
very streaky with respect to the AFM. The AFM can
compensate for this tilt.

The horizontal lines are due to tip hops – where


the tip picks up or loses a small “nanodust” In this image the tilt have not yet
been removed.
And the Ugly!

Teeny little dust mites, ultra tiny dust mites


about 2,000 in the average bed
Low Energy Electron Diffraction (LEED)
Low energy electron diffraction (LEED) is the oldest of the ‘modern’ techniques of
surface science, not only because its origins lie in the experiment of Davisson and
Germer in 1927 which first demonstrated the wave nature of electrons, but also because
it was the first such technique for which commercial instrumentation was developed in
the early days of stainless steel ultra-high vacuum (UHV) systems in the 1960s. It
remains the most widely used method of characterizing the long-range order of single
crystal surfaces and is still commonly regarded as the benchmark technique for
quantitative surface structure determination.

The LEED experiment uses a beam of


electrons of a well-defined low energy
(typically in the range 20 - 200 eV)
incident normally on the sample. The
sample itself must be a single crystal
with a well-ordered surface structure in
order to generate a backscattered
electron diffraction pattern. A typical
experimental set-up is shown in the
figure.
Only the elastically-scattered electrons contribute to the diffraction pattern: the lower
energy (secondary) electrons are removed by energy-filtering grids placed in front of the
fluorescent screen that is employed to display the pattern.
Theory of LEED:
By the principles of wave-particle duality, the beam of electrons may be equally regarded as
a succession of electron waves incident normally on the sample. These waves will be
scattered by regions of high localised electron density, i.e. the surface atoms, which can
therefore be considered to act as point scatterers.

The wavelength of the electrons is given be the de Broglie relation:



𝜆=
𝑝
where λ is the wavelength, and p is the electron momentum.
𝑝 = 𝑚𝑣 = 2𝑚𝐸 = 2𝑚𝑒𝑉

where m is mass of electron [ kg ], v is velocity [ m s-1 ], E is kinetic energy, e is electronic


charge and V is the acceleration voltage ( = energy in eV ).

The range of wavelengths of electrons employed in LEED experiments is comparable with


atomic spacings, which is the necessary condition for diffraction effects associated with
atomic structure to be observed.
Consider, first, a one dimensional (1-D) chain of atoms (with atomic separation a ) with the
electron beam incident at right angles to the chain. This is the simplest possible model for
the scattering of electrons by the atoms in the topmost layer of a solid; in which case the
diagram below would be representing the solid in cross-section with the electron beam
incident normal to the surface from the vacuum above.

If you consider the backscattering of a wavefront from two adjacent atoms at a well-defined
angle, θ , to the surface normal then it is clear that there is a "path difference" (d) in the
distance the radiation has to travel from the scattering centres to a distant detector (which is
effectively at infinity) - this path difference is best illustrated by considering two "ray paths"
such as the right-hand pair of green traces in the above figure.
The size of this path difference is a.sin θ and this must be equal to an integral number of
wavelengths for constructive interference to occur when the scattered beams eventually
meet and interfere at the detector i.e.
d = a sin θ = nλ
For two isolated scattering centres the diffracted intensity varies slowly between zero
(complete destructive interference ; d = (n + ½) λ ) and its maximum value (complete
constructive interference ; d = n λ ) - with a large periodic array of scatterers, however, the
diffracted intensity is only significant when the “Bragg condition” is satisfied exactly. i.e.

a sin θ = nλ

The diagram below shows a typical intensity profile for this case.
There are a number of points worth noting from
this simple 1-D model:
1. the pattern is symmetric about θ = 0 (or sin
θ = 0)
2. sin θ is proportional to 1/V1/2 (since λ is
proportional to 1/V1/2 )
3. sin θ is inversely proportional to the lattice
parameter , a
The aforementioned points are in fact much more general - all surface diffraction patterns
show a symmetry reflecting that of the surface structure, are centrally symmetric, and of a
scale showing an inverse relationship to both the square root of the electron energy and the
size of the surface unit cell.
As an example we can look at the LEED pattern from an fcc(110) surface. In the figure
below the surface atomic structure is shown on the left in plan view, as if you are viewing it
from the position of the electron gun in the LEED experiment (albeit greatly magnified). The
primary electron beam would then be incident normally on this surface as if fired from your
current viewpoint and the diffracted beams would be scattered from the surface back towards
yourself. The diffraction pattern on the right illustrates how these diffracted beams would
impact upon the fluorescent screen.

The pattern shows the same


rectangular symmetry as the
substrate surface but is "stretched"
in the opposite sense to the real
space structure due to the
reciprocal dependence upon the
lattice parameter. The pattern is
also centrosymmetric about the
(00) beam – this is the central spot
in the diffraction pattern corresponding to the beam that is diffracted back exactly normal to
the surface (i.e. the n = 0 case in our 1-D model).
The above illustration of the diffraction pattern shows only the "first-order" beams i.e. it is
representative of the diffraction pattern visible at low energies when only for n = 1 is the
angle of diffraction, θ , sufficiently small for the diffracted beam to be incident on the display
screen.
A much better method of looking at LEED diffraction patterns involves using the concept of
reciprocal space. The reciprocal net is determined by (defined by) the reciprocal vectors:
a1* & a2* (for the substrate) and b1* & b2* (for the adsorbate)
Initially we will consider just the substrate. The reciprocal vectors are related to the real
space unit cell vectors by the scalar product relations:
a1 . a2* = a1* . a2 = 0
and
a1 . a1* = a2 . a2* = 1

From the theory of reciprocal space we know that :

• a1 is perpendicular to a2* , and a2 is perpendicular to a1*

• there is an inverse relationship between the lengths of a1 and a1* (and a2 and a2* ) of the
form : |a1| = 1 / ( |a1*| cos A ) , where A is the angle between the vectors a1 and a1*.

when A = 0 degrees (cos A = 1) this simplifies to a simple reciprocal relationship between


the lengths a1 and a1*.
Exactly analogous relations hold for the real space and reciprocal vectors of the adsorbate
overlayer structure : b1 , b1* , b2 and b2* .
To a first approximation, the LEED pattern for a given surface structure may be obtained by
superimposing the reciprocal net of the adsorbate overlayer (generated from b1* and b2* ) on
the reciprocal net of the substrate (generated from a1* and a2* )
Example: The figure below shows an fcc(100) surface (again in plan view) and its
corresponding diffraction pattern (i.e. the reciprocal net) .

Demonstrate how the reciprocal vectors can be determined by working through the problem
in a parallel fashion for the two vectors : a1* must be perpendicular to a2

Solution: From the figure, it is evident that a1* is parallel to a1


Therefore, the angle, A , between a1 & a1* is zero
Hence, | a1*| = 1 / | a1 |
If we let | a1 | = 1 unit, then | a1*| = 1 unit.
Let us now add in an adsorbate overlayer - a primitive ( 2 x 2 ) structure with the adsorbed
species shown bonded in on-top sites - and apply the same logic as just used above to
determine the reciprocal vectors, b1* and b2*, for this overlayer.

We can show that b1* is perpendicular to b2 :

From the figure we can see that b1* is parallel to b1. The angle, B , between b1 & b1* is
zero. Hence, | b1*| = 1 / | b1 |
| b1 | = 2| a1 | = 2 units; ∴ | b1*| = ½ unit.

All we have to do now is to generate


the reciprocal net for the adsorbate
using b1* and b2* (shown in red).
Example: Let us consider the c( 2 × 2 ) structure on the same fcc(100) surface. The figure
below shows both a real space c( 2 × 2 ) structure and the corresponding diffraction
pattern:

In many respects the analysis is very similar to that for the p( 2 x 2 ) structure (previous
slides), except that:
1. | b1 | = | b2 | = √2 units ; consequently | b1*| = | b2*| = 1/√2 units.
2. The vectors for the adsorbate overlayer are rotated with respect to those of the
substrate by 45°.
X-ray Photoelectron Spectroscopy
X-ray photoelectron spectroscopy (XPS), also known as electron spectroscopy for
chemical analysis (ESCA), is the most extensively used surface techniques. It is a
nondestructive and quantitative technique that allows determining the atomic
composition of the sample, chemical bonds, oxidation states, quantification of elements,
and determination of contaminants.
What can you do with XPS?
• XPS is typically thought of as a chemical identification tool—one of the tool that
is especially useful for surfaces.
• Photoelectrons are emitted from a sample after irradiation with x-rays.
• The binding energy of the photoelectrons are recorded by a detector. Both the
binding energy, and the intensity of the peak, allow for a variety of
characteristics to be determined:
• Which element(s) are present?
• What is the oxidation state—to what is the element bound, locally?
• The quantity of the element present.
• Applications:
• Thin film characterization, identifying elements from Li through U (but not H and He)
• Applicable to a wide range of solid materials (but not liquids or gases)
• Quantitative analysis, even as a function of depth (“profiling”)
X-ray Photoelectron Spectroscopy
How it works: The sample to analyze is submitted to irradiation by a high energy source.
The X-ray penetrates only 1–10 nm under the surface (depending on the tilting angle of
sample) permitting the surface analysis (Figure). As an atom absorbs the X-rays, the energy
will produce a K-shell electron (e-) ejection.
The ejected electron (e-) has a
kinetic energy (KE) that is
related to the energy of the
incident beam (hν), the electron
binding energy (BE) and the
work function (ϕ) of the
spectrometer. The XPS detector
identifies the number of
electrons (e-) with the same BE
that is proportional to the number
of corresponding atoms in the
sample, which allows to know Figure: Diagram of an X-ray photoelectron spectrometer. The X-ray
the atomic percentage per source is incident on the sample, penetrating between 1 and 10 nm
element in the sample. and generating the ejection of the electrons; then, the detector
collects and identifies the electron binding energy to be translated
and quantifies each element in atomic percent.
Photoelectron and Auger Electron Emission
In the photoelectron emission process, an incident photon of energy hυ is absorbed by an
atom. With that energy, a photoelectron is emitted with a kinetic energy equal to:

KE = hv - BE – φ (1)

Where BE is equal to the binding energy of the electron and φ is the work function of the
spectrometer (see Figure). In order to produce photoelectrons having discreet binding
energies, a monoenergetic source of X-rays is necessary. As the kinetic energies of the
photoelectrons are dependent on the X-ray source energy, photoelectron spectra are typically
presented on a binding energy scale,
Eq. 2, to make comparison of spectra
collected with different sources more
straightforward.

BE = hv - KE – φ (2)

Figure: Photoelectron and Auger electron emission process


Following the process of photoelectron emission, an excited state ion is created. The
excess energy of this ion can be released through a relaxation process in which an
electron from an upper shell fills the hole and then either an X-ray photon is released, X-
ray fluorescence, or a second electron is emitted, Auger electron emission. In each case,
energy is conserved in the relaxation and the X-ray photon or Auger electron is emitted
with energy equal to the differences in energies of the orbitals, Eq. 3.
When plotted on a binding energy
scale, the positions of Auger lines will
depend on the X-ray source energy. For
this reason, X-ray induced Auger
spectra are typically presented on a
kinetic energy scale. This allows for
straightforward comparison with
electron induced Auger spectra.

KEAuger = BE1 - BE2 -BE3 (3)


Why the Core Electrons?
• An electron near the Fermi level is far from the nucleus, moving in
different directions all over the place, and will not carry information about
any single atom.
• Fermi level is the highest energy level occupied by an electron in a neutral
solid at absolute zero temperature.
• Electron binding energy (BE) is calculated with respect to the Fermi level.

• The core electrons are local close to the nucleus and have binding energies
characteristic of their particular element.
• The core electrons have a higher probability of matching the energies of Al
K and Mg K.

Valence e-

Core e- Atom
Binding Energy (BE)
The Binding Energy (BE) is characteristic of the core electrons for each
element. The BE is determined by the attraction of the electrons to the
nucleus. If an electron with energy x is pulled away from the nucleus, the
attraction between the electron and the nucleus decreases and the BE
decreases. Eventually, there will be a point when the electron will be free of
the nucleus.

This is the point with 0 energy of


attraction between the electron and
the nucleus. At this point the
electron is free from the atom.
0 B.E.

x These electrons are attracted to


the proton with certain binding
p+ energy x
XPS Spectrum
• The plot has characteristic peaks, which are sharp, for each element found in the
surface of the sample.

• In a XPS graph it is possible to see Auger electron peaks which are usually wider
peaks in a XPS spectrum (example on next slide).

• There are tables with the KE and BE already


assigned to each element.

• After the spectrum is plotted you can look for


the designated value of the peak energy from
the graph and find the element present on the
surface.

Figure: MoO3 film excited by Al Kα (1486.6 eV)


XPS Peak: Spin orbit coupling
Core levels in XPS use the nomenclature nlj where n is the principal quantum
number, l is the angular momentum quantum number and j = l + s (where s is the spin
angular momentum number and can be ±½). All orbital levels except the s levels (l = 0)
give rise to a doublet with the two possible states having different binding energies. This
is known as spin-orbit splitting (or j-j coupling).
Peaks area ratios: The peaks will also have specific area ratios based on the
degeneracy of each spin state, i.e. the number of different spin combinations that can
give rise to the total j (see next slide). For example, for the 2p spectra, where n is 2
and l is 1, j will be 1/2 and 3/2. The area ratio for the two spin orbit peaks (2p1/2:2p3/2)
will be 1:2 (corresponding to 2 electrons in the 2p1/2 level and 4 electrons in the
2p3/2 level). These ratios must be taken into account when analyzing spectra of
the p, d and f core levels.

• If l=0, single XPS peak


• If l >0, a doublet peak-- spin orbit (l-s)
coupling
• Since s can be ±1/2, each level with l > 0 is split into two sublevels with an energy
difference known as the spin-orbit splitting.

• The degeneracy of each of these levels is 2j+1


Peak Notations

L-S Coupling ( j = l s)
e- 1
s= 12 s= 2

j= l + 12 j= l 1
2
XPS Peak intensities
For p, d and f peaks, two peaks are observed.
Au
The separation between the two peaks are named
spin orbital splitting. The values of spin orbital
splitting of a core level of an element in different
compounds are nearly the same.

The peak area ratios of a core level of an


element in different compounds are also nearly
the same.

Spin orbital splitting and peak area


ratios assist in elemental identification
General methods in assisting peak identification
1. Check peak positions and relative peak intensities of 2 or more peaks (photoemission
lines and Auger lines) of an element.
2. Check spin orbital splitting and area ratios for p, d, f peaks

The following
elements were found:
O, C, Cl, Si, F, N, S,
Al, Na, Fe, K, Cu,
Mn, Ca, Cr, Ni, Sn,
Zn, Ti, Pb, V
What are some fun facts about XPS?
• Pros
• Detection limits: 0.01 to 1 at%
• Resolution (lateral): 10um - mm
• Resolution (depth): 10 – 200 A
• It can:
• Identify elements or compounds (except H and He)
• Determine oxidation states (for example Ti3+ or Ti4+)
• Identify types of chemical bonds (like Si-O or Si-C)
• Semi-quantitative analysis (with about 10-15% error)
• Determine film thickness
• Cons
• XPS is an ultra-high vacuum technique; samples must be non-volatile.
• The estimated time of analysis can vary between half an hour and 8 h depending on the
nature of the sample.
• Radiation damage possible (worse for achromatic sources)
• Charge neutralization needed for insulating material
• Chemical analysis can be limited to functional groups and in some cases chemical shifts
are not resolvable.
Why Ultra-high vacuum (UHV)?

• Allows longer photoelectron path length


• Ultra-high vacuum keeps surfaces
clean, preventing the contaminations to
produce X-ray signal
• Pressure < 10-8 Torr
• Vacuum pumps: Roughing Pump, Turbo
Pump, Ion Pump
Diffusion
Surface diffusion is the motion of adparticles, such as atoms or molecules, over the surface
of a solid substrate. The diffusing particles might be the same chemical species as the
substrate (the case referred to as a self-diffusion) or another one (the case of hetero-
diffusion). In most cases, the adparticle becomes mobile due to thermal activation and its
motion is described as a random walk. In the presence of a concentration gradient (in the
more general case, of the gradient of the chemical potential), the random walk motion of
many particles results in their net diffusion motion in the direction opposite to the direction
of the gradient. The diffusion process is affected by many factors, such as interaction
between diffusing adspecies, formation of surface phases, presence of defects, etc.
Diffusion
• Diffusion refers to the net flux of any
species, such as ions, atoms, electrons,
holes, and molecules.
• In materials processing technologies, (a)
control over the diffusion of atoms,
ions, molecules, or other species is key.

(b) (c)

(a) A copper–nickel diffusion couple before a high-temperature heat treatment. (b) Schematic
representations of Cu (red circles) and Ni (blue circles) atom locations within the diffusion
couple. (c) Concentrations of copper and nickel as a function of position across the couple.
• Chemical analysis will reveal a
condition similar to that represented in
Figure namely, pure copper and nickel at
(a)
the two extremities of the couple,
separated by an alloyed region.

• Concentrations of both metals vary with


position as shown in Figure c.

• This indicates that copper atoms have (b)


migrated or diffused into the nickel, and
that nickel has diffused into copper.

• This process, whereby atoms of one


metal diffuse into another, is termed
interdiffusion, or impurity diffusion. (c)
• Diffusion also occurs for pure metals, but
all atoms exchanging positions are of the
same type; this is termed self-diffusion

(a) A copper–nickel diffusion couple after a high-temperature heat


treatment, showing the alloyed diffusion zone. (b) Schematic
representations of Cu (red circles) and Ni (blue circles) atom
locations within the couple. (c) Concentrations of copper and
nickel as a function of position across the couple.
Stability of Atoms and Ions
Imperfections and, even atoms or ions in
their normal positions in the crystal
structures are not stable or at rest.
Instead, the atoms or ions possess thermal
energy, and they will move.
An atom may move from a normal crystal
structure location to occupy a nearby
vacancy. An atom may also move from one
interstitial site to another. Atoms or ions
may jump across a grain boundary, causing
the grain boundary to move.
The rate of atom or ion movement is related
to temperature or thermal energy by the
Arrhenius equation:
𝑄 The Arrhenius plot of ln(rate) versus 1/T can
𝑅𝑎𝑡𝑒 = 𝑐𝑜 exp(− ) be used to determine the activation energy
𝑅𝑇 for a reaction.
where co is a constant, R is the gas constant , T is the
absolute temperature (K), and Q is the activation
energy (cal mol) required to cause Avogadro’s
number of atoms or ions to move.
Mechanisms for Diffusion
Interdiffusion
• Diffusion of different atoms in
different directions is known as
interdiffusion.
• At high temperatures, nickel atoms
gradually diffuse into the copper, and
copper atoms migrate into the nickel.
• Again, the nickel and copper atoms
eventually are uniformly distributed.

Diffusion of copper atoms into nickel. Eventually, the copper


atoms are randomly distributed throughout the nickel.
• Two important mechanisms by which
atoms or ions can diffuse.

• Vacancy Diffusion: In self-diffusion


and diffusion involving substitutional
atoms, an atom leaves its lattice site to
fill a nearby vacancy (thus creating a
new vacancy at the original lattice
site). The number of vacancies, which
increases as the temperature increases,
influences the extent of both self-
diffusion and diffusion of substitutional
atoms.
• Interstitial Diffusion: When a small
interstitial atom or ion is present in the
crystal structure, the atom or ion moves
from one interstitial site to another.
Interstitial diffusion occurs more
easily than vacancy diffusion.
Interstitial atoms that are relatively
Schematic representations of (a) vacancy diffusion and (b)
smaller can diffuse faster. interstitial diffusion.
Activation Energy for Diffusion
• A diffusing atom must squeeze past the
surrounding atoms to reach its new site.
• The atom is originally in a low-energy,
relatively stable location.
• In order to move to a new location, the atom
must overcome an energy barrier. The energy
barrier is the activation energy Q.
• The thermal energy supplies atoms or ions
with the energy needed to exceed this barrier.
• The term diffusion couple is used to
indicate a combination of an atom of a given
element (e.g., carbon) diffusing in a host
material (e.g., BCC Fe).
• A low-activation energy indicates easy
diffusion.
• In self-diffusion, the activation energy is The activation energy Q is required to squeeze atoms past
one another during diffusion. Generally, more energy is
equal to the energy needed to create a required for a substitutional atom than for an interstitial
vacancy and to cause the movement of the atom.
atom.
Types of Diffusion
• Volume diffusion: The atoms move through the crystal from one regular or interstitial site
to another. Because of the surrounding atoms, the activation energy is large and the rate of
diffusion is relatively slow.

• Atoms can also diffuse along boundaries, interfaces, and surfaces in the material. Atoms
diffuse easily by grain boundary diffusion, because the atom packing is disordered and
less dense in the grain boundaries. Because atoms can more easily squeeze their way
through the grain boundary, the activation energy is low.

• Surface diffusion is easier still because there is even less constraint on the diffusing
atoms at the surface.
Random Walks
Consider the thermal motion of an adatom on an
ideal crystal surface (Fig.). On the atomic scale,
the surface comprises a periodic array of
adsorption sites, which correspond to the positions
of minimum energy (position 1 in Fig. a). Due to
thermal excitations, an adatom can hop from one
adsorption site to the next. The adatom motion
along the surface can be visualized as a random
site-to-site hopping process (random-walk
motion), for which the mean-square displacement
of the hopping atom in time t is given by

< Δ𝑟 2 > = 𝜈𝑎2 𝑡

where a is the jump distance (i.e., the adsorption


site spacing) and ν the frequency of hops. Note
that νt gives the number of hops. For a single Figure: One-dimensional schematic diagram
adatom, < Δ𝑟 2 > is averaged over many showing (a) a substrate (open circles) and adatom
(hatched circle) in an adsorption site (labeled 1) and
repeated observation periods of duration t. in a transition (saddle point) state (labeled 2). z is
the distance normal to the surface and z is the
coordinate along the surface, (b) Schematic
potential energy diagram for adatom motion along
the surface,
The time-independent ratio of the mean-square displacement < Δ𝑟 2 > to time t is known
as the diffusion coefficient (or diffusivity), D:

< Δ𝑟 2 > 𝜈𝑎2


𝐷= =
𝑧𝑡 𝑧
where z is the number of neighboring sites which the atom can hop to. It is apparent that z
= 2 for one-dimensional diffusion (as shown in Fig. b, where the atom can hop either to the
left or to the right neighboring sites),
z = 4 for surface diffusion on a square lattice, and z = 6 for surface diffusion on a
hexagonal lattice.

Figure: One-dimensional schematic diagram


showing (a) a substrate (open circles) and adatom
(hatched circle) in an adsorption site (labeled 1)
and in a transition (saddle point) state (labeled 2).
z is the distance normal to the surface and z is the
coordinate along the surface, (b) Schematic
potential energy diagram for adatom motion
along the surface,
Atom hopping from site to site requires surmounting
the potential barrier, i.e., this is a thermally activated
process. If the oscillation frequency of the atom in the
well (which is essentially an attempt frequency to
overcome the barrier) is ν0 and the barrier height is
Ediff, the hopping frequency can be expressed as
𝐸𝑑𝑖𝑓𝑓
𝜈 = 𝑣0 exp(− )
𝑘𝐵 𝑇
where kB is the Boltzmann constant and T the
temperature. As one can see in Figure, the activation
energy of diffusion Ediff is the difference in potential
energy of the adatom in the equilibrium adsorption
site (position l) and in the transition saddle point
(position 2). Ediff is far less than the desorption energy
Edes (typically Ediff ~ 5-20 % of Edes).

For chemisorbed species, Ediff >> kB T and the


diffusion mechanism is referred to as hopping (or (c) Schematic diagram of the adatom
jumping) diffusion. If Ediff is less than kB T, the atoms potential energy as a function of z in
positions 1 and 2 as in (a). The activation
transfer freely across the surface as a two- energy of surface diffusion Ediff equals the
dimensional gas. This type of motion, called mobile energy difference of the minima of the
diffusion, comprises a rather rare case detected only curves 1 and 2. The desorption energy
Edes is shown for comparison
for a few physisorbed species and will be left out of
the scope of further consideration.
Regimes
There are four different general schemes in which diffusion may take place. Tracer diffusion
and chemical diffusion differ in the level of adsorbate coverage at the surface, while intrinsic
diffusion and mass transfer diffusion differ in the nature of the diffusion environment. Tracer
diffusion and intrinsic diffusion both refer to systems where adparticles experience a
relatively homogeneous environment, whereas in chemical and mass transfer diffusion
adparticles are more strongly affected by their surroundings.
Tracer diffusion describes the motion of
individual adparticles on a surface at relatively
low coverage levels. At these low levels (< 0.01
monolayer), particle interaction is low and each
particle can be considered to move
independently of the others. The single atom
diffusing in figure 1 is a nice example of tracer
diffusion.

Figure: Model of a single adatom diffusing across a square


surface lattice. Note the frequency of vibration of the adatom
is greater than the jump rate to nearby sites. Also, the model
displays examples of both nearest-neighbor jumps (straight)
and next-nearest-neighbor jumps (diagonal).
https://en.wikipedia.org/wiki/File:Surface_diffusion_hopping.gif
Chemical diffusion describes the process at higher level of coverage where the effects of
attraction or repulsion between adatoms becomes important. These interactions serve to
alter the mobility of adatoms. In a crude way, figure serves to show how adatoms may
interact at higher coverage levels. The adatoms have no "choice" but to move to the right
at first, and adjacent adatoms may block adsorption sites from one another.

Figure. Model of six adatoms diffusing


across a square surface lattice. The adatoms
block each other from moving to adjacent
sites.
https://en.wikipedia.org/wiki/File:Chemical_surface_diffusion_slow.gif
Rate of Diffusion [Fick’s First Law]
• The rate at which atoms, ions, particles or other
species diffuse in a material can be measured by the
flux J.
• The flux J is defined as the number of atoms passing
through a plane of unit area per unit time.

• Fick’s first law explains the net flux of atoms:


𝑑𝑐
𝐽 = −𝐷 𝑑𝑥
where J is the flux, D is the diffusivity or diffusion coefficient,
and dc/dx is the concentration gradient.
• Depending upon the situation, concentration may be The flux during diffusion is defined as the
expressed as atom percent (at%), weight percent number of atoms passing through a plane
(wt%), mole percent (mol%), atom fraction, or mole of unit area per unit time.
fraction.
Several factors affect the flux of atoms during diffusion. If we are dealing with diffusion of
ions, electrons, holes, etc., the units of J, D, and dc/dx will reflect the appropriate species that
are being considered.
Thermal energy associated with atoms, ions, etc., causes the random movement of atoms. At
a microscopic scale, the thermodynamic driving force for diffusion is the concentration
gradient. A net or an observable flux is created depending upon temperature and the
concentration gradient.
Concentration Gradient
The concentration gradient shows how the
composition of the material varies with
distance: Δc is the difference in concentration
over the distance Δx.
A concentration gradient may be created when
two materials of different composition are
placed in contact, when a gas or liquid is in
contact with a solid material, when
nonequilibrium structures are produced in a
material due to processing, and from a host of
other sources.
The flux at a particular temperature is constant
only if the concentration gradient is also
constant—that is, the compositions on each Illustration of the concentration gradient
side of the plane remain unchanged.
In many practical cases, however, these compositions vary as atoms are redistributed,
and thus the flux also changes.
Often, the flux is initially high and then gradually decreases as the concentration
gradient is reduced by diffusion.
Factors Affecting Diffusion
Temperature and the Diffusion Coefficient
• The kinetics of diffusion are strongly dependent
on temperature. The diffusion coefficient D is
related to temperature by an Arrhenius-type
equation,
𝑄
𝐷 = 𝐷0 exp(− )
𝑅𝑇
where Q is the activation energy, R is the gas
constant , and T is the absolute temperature (K).

• D0 is a constant for a given diffusion system and is


equal to the value of the diffusion coefficient at
1/T = 0 or T =∞.

• Covalently bonded materials, such as carbon and


silicon, have unusually high activation energies,
consistent with the high strength of their atomic The diffusion coefficient D as a function of
bonds. reciprocal temperature for some metals and
ceramics. In this Arrhenius plot, D represents
the rate of the diffusion process. A steep slope
denotes a high activation energy.
In ionic materials, such as some of the oxide
ceramics, a diffusing ion only enters a site
having the same charge.
In order to reach that site, the ion must
physically squeeze past adjoining ions, pass by
a region of opposite charge, and move a
relatively long distance.
Consequently, the activation energies are high
and the rates of diffusion are lower for ionic
materials than those for metals.

• When the temperature of a material


increases, the diffusion coefficient D Diffusion in ionic compounds. Anions can
increases and, therefore, the flux of atoms only enter other anion sites. Smaller cations
increases as well. tend to diffuse faster.
• At higher temperatures, the thermal energy
supplied to the diffusing atoms permits the
atoms to overcome the activation energy
barrier and more easily move to new sites.
Time
• Diffusion requires time. If a large number of atoms must diffuse to produce a
uniform structure, long times may be required, even at high temperatures.
• Times for heat treatments may be reduced by using higher temperatures or by
making the diffusion distances (related to Δx) as small as possible.
• Some remarkable structures and properties are obtained if we prevent diffusion.
Steels quenched rapidly from high temperatures to prevent diffusion form
nonequilibrium structures and provide the basis for sophisticated heat treatments.
• In certain thin film deposition processes such as sputtering, we sometimes obtain
amorphous thin films if the atoms or ions are quenched rapidly after they land on
the substrate. If these films are subsequently heated (after deposition) to sufficiently
high temperatures, diffusion will occur and the amorphous thin films will
eventually crystallize.
Dependence on Bonding and Crystal Structure

• Interstitial diffusion, with a low-


activation energy, usually occurs
much faster than vacancy, or
substitutional diffusion.
• Activation energies are usually lower
for atoms diffusing through open
crystal structures than for close-
packed crystal structures.
• Because the activation energy
depends on the strength of atomic
bonding, it is higher for diffusion of
atoms in materials with a high The activation energy for self-diffusion increases
melting temperature. as the melting point of the metal increases.

• Cations (with a positive charge), due to their smaller size, often have higher diffusion
coefficients than those for anions (with a negative charge).
• Diffusion of ions also provides a transfer of electrical charge; in fact, the electrical
conductivity of ionically bonded ceramic materials is related to temperature by an
Arrhenius equation. As the temperature increases, the ions diffuse more rapidly,
electrical charge is transferred more quickly, and the electrical conductivity is increased.
Composition Profile [Fick’s Second Law]
Fick’s second law, which describes the dynamic, or non-steady state, diffusion of atoms, is
the differential equation

𝜕𝑐 𝜕 𝜕𝑐
= 𝐷
𝜕𝑡 𝜕𝑥 𝜕𝑥

If we assume that the diffusion coefficient D is not a function of location x and the
concentration (c) of diffusing species, we can write a simplified version of Fick’s second law
as follows

𝜕𝑐 𝜕2𝑐
=𝐷
𝜕𝑡 𝜕𝑥2

The solution to this equation depends on the boundary conditions for a particular situation.
One solution is

𝑐𝑠 − 𝑐𝑥 𝑥
= erf( )
𝑐𝑠 − 𝑐0 2 𝐷𝑡
where cs is a constant concentration of the diffusing atoms at the surface of the material, c0 is
the initial uniform concentration of the diffusing atoms in the material, and cx is the
concentration of the diffusing atom at location x below the surface after time t.
A one-dimensional model is assumed in these equations (i.e., we assume that atoms or
other diffusing species are moving only in the direction x).

Diffusion of atoms into the surface of a material illustrating the use of Fick’s second law.

𝑐 −𝑐 𝑥
Equation 𝑐𝑠 −𝑐𝑥 = erf(2 ) is a possible solution to Fick’s law that describes the variation in
𝑠 0 𝐷𝑡
concentration of different species near the surface of the material as a function of time and
distance, provided that the diffusion coefficient D remains constant and the concentrations of
the diffusing atoms at the surface (cs) and at large distance (x) within the material (c0) remain
unchanged.
The function “erf” is the error function and can be evaluated from Table or Figure below.

Graph showing the argument and value of the error function


encountered in Fick’s second law.
The mathematical definition of the error function is as follows:
𝑥
2
erf 𝑥 = න exp −𝑦 2 𝑑𝑦
𝜋 0
Here, y is known as the argument of the error function. We also define a
complementary error function as follows:
erfc(x) = 1

• Depending upon the boundary conditions, different solutions (i.e., different


equations) describe the solutions to Fick’s second law. These solutions to Fick’s
second law permit us to calculate the concentration of one diffusing species as a
function of time (t) and location (x).
Grain Growth

A polycrystalline material contains a large


number of grain boundaries, which represent
high-energy areas because of the inefficient a
packing of the atoms.
A lower overall energy is obtained in the
material if the amount of grain boundary area
is reduced by grain growth.
Grain growth involves the movement of
grain boundaries, permitting larger grains to
grow at the expense of smaller grains.

For grain growth in materials, diffusion of


atoms across the grain boundary is required,
and, consequently, the growth of the grains is
related to the activation energy needed for an
atom to jump across the boundary. b
• Grain boundaries are defects and their presence causes the free energy of the material to
increase. Thus, the thermodynamic tendency of polycrystalline materials is to transform
into materials that have a larger average grain size.
• High temperatures or low-activation energies increase the size of the grains. Many heat
treatments of metals, which include holding the metal at high temperatures, must be
carefully controlled to avoid excessive grain growth. This is because, as the average
grain size grows, the grain-boundary area decreases, and there is consequently less
resistance to motion of dislocations. As a result, the strength of a metallic material will
decrease with increasing grain size.

You might also like