Manual Set2 MSE251 4apr2024

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 45

MSE 251: Physical Metallurgy and Materials

Characterization Lab (2022-2023-I)


Lab Schedule: Tuesday and Thursday, 14:00 to 16:50hrs
Lab Start Date Tuesday & Thursday
Venue: Physical Metallurgy Lab (WL-201& 202)

Instructor:
Shashank Shekhar, MSE Shivam Tripathi, MSE
Office: WL304A Office: 510, ESB-III
Email: shashank@ Phone: 6528 Email: shivamt@ Phone:2426

Lab In charges:
Mr. G P Bajpai Mr. Ajay Pratap Singh
Phone: 7933; Email: gpbajpai@ Phone: 7933; Email: ajayps@

TAs:
Experiment Tuesday Thursday
Exp-8 Ayan Mulik Peri Bala Kameshwar
Exp-9 Swagata Datta Abhishek K. Verma
Exp-10 Amit Kr Yadav Ankit Singh Negi
Exp-11 Deepti Gangwar Radhakrishnan Sriram
Exp-12 Abinav Dixit Ruhi Syed

General Information

✓ Total Laboratory Turns: 12


✓ There are seven experiments in Set I and five experiments in Set II
✓ You are required to bring your lab manual on the day of experiment. Student must carefully
read each experiment from the lab manual before it is performed in the laboratory. There
will be weekly viva on the experiment.
✓ Half pants, loosely hanging garments and slippers are strictly prohibited for safety reasons.
Students must wear shoes in the lab.
✓ You must enter the lab by 2:05 PM. Late comers will not be allowed to enter.
✓ Attendance will be taken by TAs at the beginning of lab session. They might take attendance
in between the session as well. Only one make-up in each Set will be allowed with prior
written permission from Instructor. If a student is absent for 3 or more labs in the entire
semester, then the student will be deregistered from course.
✓ Mobile phones should not be used in the lab.
✓ Submission of the reports to be done on the same day of experiments. Laboratory report
must be submitted in standard lab-report sheets, available at the shopping center. Reports
on ordinary sheets and computer papers will not be accepted
✓ The lab report should contain: (i) Your details including group name, Day
(Tuesday/Thursday), (i) Title of the experiment, (ii) Name of all equipment/tools used along
with one line description of its use and (iii) Observations/Results write down few lines.
✓ Marks and attendance will be circulated regularly. In case of any discrepancy, please contact
lab in charge just after circulation.

Marks Distribution

Attendance (24 %), Weekly viva (30 %), Individual weekly report (10 %), Endsem viva/quiz
(36 %)

✓ Viva and/or quiz will be conducted for endsem. The actual format for endsem will be
informed beforehand. In case there are changes, the same will be communicated beforehand.
Safety

✓ Careless or improper handling of machines may result in serious injury. Hence, to avoid
any injury, the students must first seek permission and guidance from the laboratory staff
before handling any machine.
✓ Students must ensure that their work areas are clean.
✓ At the end of each experiment, students must clear off all tools and materials from the
work area.

Recommended Reading
✓ The Principles of Metallographic Laboratory Practice by George L. Kehl
✓ Physical Metallurgy by V. Raghavan
✓ Engineering Physical Metallurgy by Y. Lakhtin
List of Experiments
Experiments for Set-1

Lab Activity/Experiment Order of Order of


Turn groups groups
Tuesday Thursday
1 Formation of Groups, Distribution of manual, Common instructions Half of the Half of the
for lab section section
Sample preparation (mounting, polishing, and etching) and observing
microstructure of steel in optical microscope (BF, DF, etc.)
(for whole section)
2 Sample preparation (mounting, polishing, and etching) and observing Half of the Half of the
microstructure of soft metallic material in optical microscope (BF, DF, etc.) section section
(for whole section)
3 (i) Heat treatment of steel – Microstructure examination and hardness T1,T5,T4, Th1,Th5,T
measurement post annealing, normalizing, hardening, tempering (ii) Optical T3,T2 h4,Th3,Th2
microstructure examination of cast iron and steel (hypoeutectoid, eutectoid,
hypereutectoid)
4 (i) Hardenability – Jominy end quench test (ii) Carburization - T2,T1,T5, Th2,Th1,T
decarburization of steel and measurement of case depth T4,T3 h5,Th4,Th3
5 (i) Optical microstructure examination of non-ferrous material: cold work, T3,T2,T1, Th3,Th2,T
recovery, recrystallization, grain growth and twin observations, (ii) T5,T4 h1,Th5,Th4
Quantitative metallography – Grain growth kinetics/pearlitic growth (JMAK
relation)
6 Wettability/contact angle measurement (ii) Optical microstructure T4,T3,T2, Th4,Th3,T
examination after arc welding and hardness measurement T1,T5 h2,Th1,Th5
7 (i) Sample preparation and demonstration of scanning electron microscopy T5,T4,T3, Th5,Th4,T
(imaging and EDS), (ii) Fractography of tensile, fatigue fractured samples T2,T1 h3,Th2,Th1

Experiments for Set-2

Lab Activity/Experiment Order of Order of


Turn groups groups
Tuesday Thursday
8 (i) Thermal etching and microstructure observation of ceramic materials, (ii) T1,T5,T4, Th1,Th5,T
Etch pit technique for dislocation density measurement (LiF, Fe-Si, Cu, etc.) T3,T2 h4,Th3,Th2
9 (i) DMA to observe glass transition/phase transformation in polymer (ii) T2,T1,T5, Th2,Th1,T
Macrostructure examination of an ingot section (e.g. Zn, steel) T4,T3 h5,Th4,Th3
10 (i) Creating and visualizing crystal-structures and related features using a T3,T2,T1, Th3,Th2,T
crystallography software (ii) X-ray diffraction pattern to determine crystal- T5,T4 h1,Th5,Th4
structure using a crystallography software
11 (i) Differential scanning calorimetry (DSC) - Phase transformation in metals T4,T3,T2, Th4,Th3,T
and polymers (ii) X-Ray diffraction – Determination of crystallite size/strain T1,T5 h2,Th1,Th5
12 (i) Demonstration and analysis of FTIR, Raman and other advanced T5,T4,T3, Th5,Th4,T
characterization technique, (ii) Demonstration of sample preparation for T2,T1 h3,Th2,Th1
TEM
EXPERIMENT 8
(i) Thermal etching and microstructure observation of ceramic materials
(ii) Etch pit technique for dislocation density measurement (LiF, Fe-Si, Cu,
etc.)

Objective

1. To observe the microstructure of Al2O3 ceramic


2. To measure dislocation density using etch pit technique

Introduction:

Thermal etching and microstructure observation of ceramic materials.

Ceramography is the art and science of preparation, examination, and evaluation of ceramic
microstructures. The general procedure for ceramographic preparation consists of five basic
steps: sectioning, mounting or encapsulation, grinding, polishing and etching. Examination is
done by visible-light microscopy and scanning electron microscopy. Evaluation can be done
directly on the microscope image, such as by image analysis, or on a photomicrograph. Refer
to Experiment 1 for sectioning, mounting or encapsulation, grinding, polishing, and chemical
etching.

▪ Thermal Etching

Etching reveals and delineates grain boundaries, which allows grain-size measurement or
comparison. Do not etch the microstructure before measuring the microhardness or evaluating
the porosity. Ceramics can be chemically etched, but thermal etching is often better. The
flatness of the polished surface is thermodynamically unstable at high temperatures, such that
each polished grain lowers its surface energy by rounding outward slightly, creating relief and,
effectively, contrast between adjacent grains. Chemical etchants preferentially corrode highly
susceptible regions, such as grain boundaries or certain crystallographic orientations, creating
contrast between adjacent grains that can be observed in a microscope. For thermal etching of
alumina, remove the specimen from its encapsulants as following:
Saw the encapsulant around the specimen on all sides with a diamond wafering blade, leaving
a border about 1 mm wide. Calcine the remaining encapsulant in a ceramic crucible over a
Bunsen burner flame, inside a fume hood. Air-cool the ash-laden specimen back to room
temperature and scrape or brush off the ash. Use an ultrasonic bath to remove the remainder of
the ash. Spray the specimen with ethanol and dry it with a heat gun.

Place the clean, polished, diencapsulated section on a refractory block in a box or tube furnace
with a clean interior and clean elements. Heat the furnace to 1500°C at 10 K/min, hold at
1500°C for 20-30 min, and cool back to room temperature at 50 K/min.

Non-oxide ceramics (e.g. SiC and Si3N4), in general, should be chemically etched or else
thermally etched in a vacuum or inert atmosphere. Allotropic ceramics such as non-stabilized
zirconia (e.g. TZP and TTZ) tend to shatter at high temperatures due to phase changes and,
therefore, should be chemically etched.
Etch pit technique for dislocation density measurement

Dislocations are important defects because they influence rate at which crystals grow, as well
as the mechanical, electrical, and optical properties of crystals. The two limiting cases for
dislocations (edge and screw) are shown in Fig. 8.1. While a small population of dislocations
normally occurs in all crystals, the population can be greatly enhanced by plastic deformation.

Figure 8.1. The atomic structure of the (a) edge dislocation and the (b) screw dislocation. In
(b) the smaller filled circles are in the same plane (above the paper) and the larger open circles
are in a different plane (below the paper).

Because these defects have dimensions on the atomic scale, direct observation is difficult.
However, it is possible to locate dislocations by forming visible etch pits at the point where
dislocations intersect the crystal surface. Using this method, we will examine the arrangement
of dislocations in as received crystals, in crystals after plastic deformation, and in the deformed
crystals after annealing.

The atomic displacements near a dislocation are a source of internal stress. In the diagram of
the positive edge dislocation shown in Fig. 8.1(a), it is clear that the upper half of the crystal,
where the extra half plane has been inserted, is under compression and the bottom half of the
crystal is under tension. The stress fields that result from the compressive and tensile forces
around different dislocations interact with one another. If the dislocations are free to move, we
expect them to take up positions that minimize their elastic energy. There are two types of
dislocation motion, glide and climb. When a dislocation moves by glide, as illustrated in Fig.
8.2, it remains in the same slip plane. This process, called conservative motion, can occur at
relatively low temperatures. Climb, on the other hand, is a nonconservative motion that
involves the creation and migration of point defects. Thus, it is temperature activated and
occurs only at higher temperatures. This process is illustrated in Fig. 8.3.

Figure 8.2. Dislocation glide. An edge dislocation moves to the right and eventually exits the
crystal.
Figure 8.3. Dislocation glide. An edge dislocation moves to the right and eventually exits the
crystal.

Equilibrium configurations of dislocations of the same and opposite signs have been
determined theoretically and the results are summarized in Fig. 8.4. In arriving at these
configurations, it was assumed that only glide was possible. A somewhat different result
occurs if climb is active. For example, consider the attractive interaction of two edge
dislocations of opposite sign (if the edge dislocation shown in Fig. 8.1 is a positive dislocation,
then a negative edge dislocation is one in which the extra half plane is inserted from below).
If these two dislocations are permitted to climb towards each other, they annihilate.

Figure 6.4. The equilibrium configuration of edge dislocations free to glide (a) of the same
sign, (b) of opposite sign.

We will determine the dislocation density examining etch pits. An example of such pits is
shown in Fig. 8.5. These pits are created using etching. During dissolution, atoms from near
the dislocation are removed at different rate from those far from the dislocation. This process,
which results in the formation of an etch pit, occurs because of the bond strain near the
dislocation (atoms in the strained positions are more easily removed) and the impurities that
concentrate near the dislocation. In practice, producing visible etch pits can be difficult since
it is important that the removal rates from the ideal and defective areas occur in the proper
ratio.
Figure 8.5. Dislocation etch pits. (a) a three-dimensional view of a pit with square pyramidal
facets. (b) Projection showing schematic distribution on the surface.

To Report:

1. Calculate the grain size of the ceramic after thermal etching.


2. What is the size distribution of the grain size of the ceramic?
3. What was the shape of the etch pits that was observed? Why these kinds of shapes
are formed?
4. Calculate the approximate density of the etch pits. Is this accurate representation
of dislocation density?

Questions:

1. Why does thermal etching reveal grain boundaries. Explain with help of diagrams
2. Do you see any effect of deformation on the location of etch pits?
3. Can you differentiate between edge, screw or mixed dislocations?
4. How can you differentiate between dislocation etch pits and square particles on
the surface?
EXPERIMENT 9

(i) DMA to observe glass transition/phase transformation in polymer


(ii) Macrostructure examination of an ingot section (e.g. Zn, steel)

Objective

1. To observe glass transition/phase transformation in polymer using DMA


2. To examine macrostructure of an ingot section

Introduction:

Dynamic mechanic analysis

Dynamic mechanical analysis (DMA), also known as forced oscillatory measurements and
dynamic rheology, is a basic tool used to measure the viscoelastic properties of materials
(particularly polymers). To do so, DMA instrument applies an oscillating force to a material
and measures its response; from such experiments, the viscosity (the tendency to flow) and
stiffness of the sample can be calculated. These viscoelastic properties can be related to
temperature, time, or frequency. As a result, DMA can also provide information on the
transitions of materials and characterize bulk properties that are important to material
performance. DMA can be applied to determine the glass transition of polymers or the
response of a material to application and removal of a load, as a few common examples. The
usefulness of DMA comes from its ability to mimic operating conditions of the material,
which allows researchers to predict how the material will perform.

DMA is based on two important concepts of stress and strain. Application of stress (𝜎 = 𝐹/𝐴)
to the material result in change in sample dimensions causing strain (𝜀 = Δ𝐿/𝐿). The modulus
(E), a measure of stiffness, can be calculated from the slope of the stress-strain plot, see Figure
9.1. This modulus is dependent on temperature and applied stress. The change of this modulus
as a function of a specified variable is key to DMA and determination of viscoelastic
properties. Viscoelastic materials such as polymers display both elastic properties
characteristic of solid materials and viscous properties characteristic of liquids; as a result, the
viscoelastic properties are often a compromise between the two extremes. Ideal elastic
properties can be related to Hooke’s spring, while viscous behavior is often modeled using a
dashpot, or a motion-resisting damper.

Figure 9.1: An example of typical stress-strain curve


DMA instruments apply sinusoidally oscillating stress to samples and causes sinusoidal
deformation. The relationship between the oscillating stress and strain becomes important in
determining viscoelastic properties of the material. To begin, the stress applied can be
described by a sine function. The strain of a system undergoing sinusoidally oscillating stress
is also sinusoidal. Stress and strain in DMA can be expressed as following:

𝜎 = 𝜎0 sin(𝜔𝑡 + 𝛿)
𝜀 = 𝜀0 cos⁡(𝜔𝑡)

where 𝜎0 and 𝜀0 are maximum applied/achieved stress and maximum achieved/applied strain,
ω is the frequency of applied stress/strain, t is time and 𝛿 is phase difference between stress
and strain. The phase difference between strain and stress is entirely dependent on the balance
between viscous and elastic properties of the material in question. For ideal elastic systems,
the strain and stress are completely in phase, and the phase angle (δ) is equal to 0° . For viscous
systems, the applied stress leads the strain by 90° . The phase angle of viscoelastic materials is
somewhere in between 0° and 90° , see Figure 9.2.

Figure 9.2: Applied sinusoidal stress versus time (above) aligned with measured stress
versus time (below). (a) The applied stress and measured strain are in phase for an ideal
elastic material. (b) The stress and strain are 90° out of phase for a purely viscous material.
(c) Viscoelastic materials have a phase lag less than 90°

In essence, the phase angle between the stress and strain tells us a great deal about the
viscoelasticity of the material. A small phase angle indicates that the material is highly elastic;
a large phase angle indicates the material is highly viscous. Furthermore, separating the
properties of modulus, viscosity, compliance, or strain into two separate terms allows the
analysis of the elasticity or the viscosity of a material. The elastic response of the material is
analogous to storage of energy in a spring, while the viscosity of material can be thought of
as the source of energy loss.

A few key viscoelastic terms can be calculated from dynamic analysis; their equations and
significance are detailed in Table 1.
Table 1: Key viscoelastic terms that can be calculated with DMA
Term Equation Significance
Overall modulus representing
Complex modulus (𝐸 ∗ ) 𝐸 ∗ = 𝐸 ′ + 𝑖𝐸 ′′ stiffness of the material; combined
elastic and viscous components
Storage modulus; measures stored
Elastic modulus (𝐸 ′ ) 𝐸 ′ = 𝜎0 /𝜀0 cos(δ) energy and represents elastic
portion
Loss modulus; contribution of
Viscous modulus (𝐸 ′′ ) 𝐸 ′′ = 𝜎0 /𝜀0 sin(δ) viscous component on polymer that
flows under stress
Damping or index of
Loss Tangent (tan δ) tan(𝛿) = 𝐸 ′′ /𝐸 ′ viscoelasticity; compares viscous
and elastic moduli

A temperature sweep is the most common DMA test used on solid materials. In this
experiment, the frequency and amplitude of oscillating stress is held constant while the
temperature is increased. The temperature can be raised in a stepwise fashion, where the
sample temperature is increased by larger intervals (e.g., 5 ℃) and allowed to equilibrate
before measurements are taken. Continuous heating routines can also be used (1-2 ℃/minute).
Typically, the results of temperature sweeps are displayed as storage and loss moduli as well
as tan(δ) as a function of temperature. For polymers, these results are highly indicative of
polymer structure.

In time scans, the temperature of the sample is held constant, and properties are measured as
functions of time, gas changes, or other parameters. This experiment is commonly used when
studying curing of thermosets, materials that change chemically upon heating. Data is
presented graphically using modulus as a function of time; curing profiles can be derived from
this information.

Frequency scans test a range of frequencies at a constant temperature to analyze the effect of
change in frequency on temperature-driven changes in material. This type of experiment is
typically run-on fluids or polymer melts. The results of frequency scans are displayed as
modulus and viscosity as functions of log frequency.

Instrumentation

The most common instrument for DMA is the forced resonance analyzer, which is ideal for
measuring material response to temperature sweeps. The analyzer controls deformation,
temperature, sample geometry, and sample environment. Figure 9.3 displays the important
components of the DMA, including the motor and driveshaft used to apply torsional stress as
well as the linear variable differential transformer (LVDT) used to measure linear
displacement. The carriage contains the sample and is typically enveloped by a furnace and
heat sink.
Figure 9.3: General schematic of DMA analyzer

The DMA should be ideally selected to analyze the material at hand. The DMA can be either
stress or strain controlled: strain-controlled analyzers move the probe a certain distance and
measure the stress applied; strain-controlled analyzers provide a constant deformation of the
sample. Although the two techniques are nearly equivalent when the stress-strain plot is linear,
stress-controlled analyzers provide more accurate results.

DMA analyzers can also apply stress or strain in two manners—axial and torsional
deformation. Axial deformation applies a linear force to the sample and is typically used for
solid and semisolid materials to test flex, tensile strength, and compression. Torsional
analyzers apply force in a twisting motion; this type of analysis is used for liquids and polymer
melts but can also be applied to solids. Although both types of analyzers have wide analysis
range and can be used for similar samples, the axial instrument should not be used for fluid
samples with viscosities below 500 Pa-s, and torsional analyzers cannot handle materials with
high modulus.

Different fixtures can be used to hold the samples in place and should be chosen according to
the type of samples analyzed. The sample geometry affects both stress and strain and must be
factored into the modulus calculations through a geometry factor. The fixture systems are
specific to the type of stress application. Axial analyzers have a greater number of fixture
options; one of the most commonly used fixtures is extension/tensile geometry used for thin
films or fibers. In this method, the sample is held both vertically and lengthwise by top and
bottom clamps, and stress is applied upwards. For torsional analyzers, the simplest geometry
is the use of parallel plates. The plates are separated by a distance determined by the viscosity
of the sample. Because the movement of the sample depends on its radius from the center of
the plate, the stress applied is uneven; the measured strain is an average value.

DMA of the glass transition polymers

As the temperature of a polymer increases, the material goes through a number of minor
transitions (Tγ and Tβ) due to expansion; at these transitions, the modulus also undergoes
changes. The glass transition of polymers (Tg) occurs with the abrupt change of physical
properties within 140-160 ℃; at some temperature within this range, the storage (elastic)
modulus of the polymer drops dramatically. As the temperature rises above the glass transition
point, the material loses its structure and becomes rubbery before finally melting. The
idealized modulus transition is pictured in Figure 9.4.
Figure 9.4: Ideal storage modulus transitions of viscoelastic polymers

The glass transition temperature can be determined using either the storage modulus, complex
modulus, or tan δ (vs temperature) depending on context and instrument; because these
methods result in such a range of values, the method of calculation should be noted. When
using the storage modulus, the temperature at which E’ begins to decline is used as the T g.
Tan δ and loss modulus 𝐸 ′′ show peaks at the glass transition; either onset or peak values can
be used in determining Tg. These different methods of measurement are depicted graphically
in Figure 9.5.

Figure 9.5: Different industrial methods of calculating glass transition temperature (T g)

Macrostructure

The grain macrostructure in ingots and most castings have three distinct regions or zones: the
chill zone, columnar zone, and equiaxed zone. Figure 9.6 below depicts these zones.
Figure 9.6: Schematic of a typical macrostructure of an ingot

In the macrostructure of ingots, different zones form due to variations in cooling rates,
solidification conditions, and alloy segregation during the casting process. These zones are
typically observable with the naked eye and represent distinct regions with different
microstructural characteristics. Here are some reasons why different zones form in the
macrostructure of ingots:

Solidification Cooling Rates: Different parts of the ingot experience varying cooling rates
during solidification. For example, the outer regions of the ingot cool more rapidly due to
contact with the mold or cooling medium, while the central regions cool more slowly.
Variations in cooling rates can lead to differences in grain size, dendrite morphology, and
overall microstructure between different zones of the ingot.

Centerline Segregation: As the ingot solidifies from the outside inward, solute atoms may
segregate toward the centerline due to differences in solubility at different temperatures. This
phenomenon, known as centerline segregation, can result in compositional variations between
the central and peripheral regions of the ingot. These compositional differences can influence
the microstructure and properties of each zone.

Thermal Gradients: Thermal gradients within the ingot contribute to the formation of distinct
zones. Columnar grains typically develop in the outer regions of castings where the molten
metal comes into contact with the cooler mold. The higher thermal gradient at the mold
interface promotes the nucleation and growth of columnar grains aligned along the direction of
heat flow. Equiaxed grains are typically found in the central regions of castings or in areas
where the cooling rate is relatively uniform. These regions experience lower thermal gradient
and slower solidification rates, allowing for the development of equiaxed grains with a more
isotropic distribution.

Ingot Size and Shape: The size and shape of the ingot can also affect the formation of different
zones in the macrostructure. Larger ingots may exhibit more pronounced variations in cooling
rates and microstructural features across different regions. Similarly, ingots with complex
shapes or geometries may experience non-uniform cooling, leading to the formation of distinct
zones.
Overall, the formation of different zones in the macrostructure of ingots is a complex interplay
of factors such as cooling rate, solidification conditions, alloy composition, and ingot
geometry. Understanding these factors is essential for optimizing casting processes, controlling
microstructural features, and ensuring the quality and performance of cast ingots in various
applications.

To Report:

1. Calculate the storage modulus and loss modulus for the polymer(s) at RT and at
highest temperature
2. Calculate tanδ for the polymers at RT and at highest temperature (use equation)
3. Find the glass transition temperature from storage modulus, loss modulus and
tanδ. Are they same? Why or why not?
4. Measure grain size for various sections of the ingot.
Questions:

1. How is storage modulus different from loss modulus. What is tanδ?


2. Can we measure phase transition of polymers using this method? What about
metals?
3. How would the macrostructure look like if the cast object is very large? How will
it look in the case of very small object? (Draw schematic)
EXPERIMENT 10

(i) Creating and visualizing crystal-structures and related features using a


crystallography software
(ii) X-ray diffraction pattern to determine crystal-structure using a
crystallography software

Objective:

• Visualization of crystal-structures and related features using CaRIne crystallography


software.
• Using X-ray diffraction pattern to determine crystal-structure.

Introduction:

Visualization of crystal-structures and related features

Before we learn about the software, let’s familiarize ourselves with some basic definitions/
information regarding crystals lattice/structure:

1.Crystal-Systems: The 3-dimentional space can be divided into smaller volume units such
that each unit is made of three sets of planes. When one thinks of all such possible volume units
that can be repeated onto each other while maintaining the same orientation as the starting unit
volume, there are only seven such unit volumes that will span the entire space without leaving
any gap between them. These 7-unit volumes are known as 7 crystal- systems. That are: (1)
Triclinic, (2) Monoclinic, (3) Orthorhombic, (4) Tetragonal, (5) Cubic, (6) Rhombohedral, and
(7) Hexagonal. A crystal-system is identified by the lengths of the 3 sides (viz. a, b, c) of the
unit volume-space and the 3 angles (α, β, γ) between these three sides.

Note: For each of the above-mentioned unit-cells there are 6 faces. But it is possible to divide
3-D into such volume-units where each unit is surrounded by more than 6 faces, and these units
when placed onto each other while maintaining the relative orientation can span the entire space
without leaving any gap between them.

2.Crystal-Lattice: A crystal-lattice is defined as a periodical arrangement of geometrical-


points in 3-dimentional space in such a way that each point has identical environment around
it. By identical environment it is meant that when the lattice is observed from different lattice-
points (one point at a time) in a particular orientation, it looks identical from each of the points.
A lattice is a mathematical concept; there are no atoms or molecules located at these points

3.Bravais-Lattice: When one looks for all the possible arrangements of geometrical-points in
the above mentioned seven crystal-systems such that these points conform to the translation
and orientation constraints of a Crystal-lattice, one finds there are total 14 such arrangements
of points in 3-D. The special arrangements are identified as 14 Bravais Lattices.

4.Primitive unit-cell & Conventional unit-cell: The primitive-unit cell is the smallest volume
which contains exactly one lattice point, and when it is translated by all the vectors in a Bravais-
lattice (vectors connecting a lattice-point, selected as an origin, to all other lattice-points).
A conventional unit-cell is larger than primitive unit-cell containing more than one lattice
points. It also satisfies the criteria of a unit-cell i.e., when it is translated by some sub-set of
Bravais-lattice vectors, it spans the entire lattice. Generally, the conventional unit-cells are used
because they easily reveal the geometrical-symmetry of the lattice.

5.Basis/Motif: The basis/motif is the minimal unit which when placed at lattice points, gives
rise to a physical crystal. The basis can be an atom, a molecule, an ion etc.

6.Crystal-Structure: A crystal-structure is a regular arrangement of a physical unit called


basis/motif at all the points of a Bravais-Lattice.

Note, that when the basis/motif consists of single atomic unit, the resulting crystal- structures
are said to be belonging to Bravais-lattice. This is because, when such a crystal- structure is
observed from any atom it is made of, while maintaining the same viewing orientation every
time, the crystal-structure looks identical. This condition also requires that each single atom
must belong to the same element.

On the other hand, in general, when the basis/motif consists of more than one atom; the
resulting crystal-structure usually doesn’t retain the characteristics of having-identical-view
from-all-the-individual-atoms. And therefore, such crystal-structures are identified as non-
Bravais lattice. Sometime these are also identified as lattice with a basis.

7.Crystallographic-planes and Miller Indices: Due to periodic arrangement of motifs/basis


they can be visualized as if lying in a plane, and these are identified as crystallographic-planes.

Usually, Miller indices are used to indicate the planes. If a plane intercepts the three axis at
a1a, a2b and a3c distance from the origin, where a, b, and c are lattice parameters along the
three axis. The corresponding Miller indices are calculated by:

1) Taking the reciprocal of the three numbers: a/a1a, b/a1b, c/a1c, which results into ratios:
1/a1, 1/a2, 1/a3

2) Finding the lowest common multiplier (LCM) for the obtained fractions such that
multiplication of it with the three ratios gives a set of 3 integrals numbers.

For example, a plane intersecting the axes at 2a, 5b, 2c will have Miller indices of 5,6,5. The
Miller Indices of a plane are shown in parentheses without any comma separator. Therefore,
for the current plane the Miller indices are (565). These indices also represent all the other
planes that are parallel to the plane for which we have calculated these indices.

8.Crystallographic direction: A direction in a crystal having the form of ua + vb + wc is said


to have a crystallographic direction as [uvw]. It also represents all the other vectors that are
parallel to the given vector. Here a, b, and c are vectors along the three-crystal axis.

9.Voids: The vacant space between the atoms in a crystal structure is identified as interstitial
voids. Two types of voids are, 1) Tetrahedral and 2) Octahedral void.
CARINE CRYSTALLOGRAPHIC SOFTWARE

Following instructions about using the software help visualize different features of a crystal
within the limitations of the present version.

Note: All word in between the following sentences that are Bold, represent a tool option
available in CaRIne, unless mentioned otherwise.

Accessing the software:

This software has been installed on selected computer in the laboratory. To open this software
double-click on the desktop icon titled CaRIne 3.1. Do not select CaRIne 3.1-sr, it has very
limited capabilities. A window will open similar to that shown in Figure 10.1.

Figure 10.1: The main window of the Carine software showing crystal structure window.

Drawing a crystal structure:

Go to Cell menu, and select one of the crystal systems from the drop-down menu. Figure 10.2
shows a unit cell of Face-centered-cubic (FCC) unit-cell.
Figure 10.2: An FCC unit cell (a) without and (b) with modifications.

• Viewing the structure from different directions: A crystal structure can be viewed from different
orientations by holding the Ctrl key pressed and using the left button of the mouse to change
the orientation.

Like changing the viewing direction as mentioned above, there are several other tools available to
visualize/modify the crystal model in different ways. To access these tool right-click while being in the
crystal structure window; a menu be displayed. This menu along and its different sub-menus have been
shown in the form of a tree-diagram (TD) in Figure 10.3.

For the instructions that follow, the term path will always refer to the tree-diagram unless
mentioned otherwise.

Figure 10.3: Tree Diagram (TD) showing the various tools and corresponding paths.
Let’s see what is available for path 1 in this Tree Diagram. By following path 1 and selecting
Cell Creation/Cell List a window will open as shown in Figure 10.4. Utilities in this window
will be used most of the time during visualizing a crystal structure, and they have been indicated
in the figure itself.

Figure 10.4: Cell Creation/Cell List window showing various attributes that can be modified
for better visualization.

Visualizing planes and directions:

Planes

After selecting the crystal-window for which a plane has to be visualized, follow path 3 and
select Choice of (hkl) planes; a new window will open where (hkl) values for up to three
different planes can be entered.

For example, if (111) plane for FCC crystal structure is to be visualized, after accessing Choice
of (hkl) planes, enter h =1, k = 1, and 1 = 1 and hit OK. It will highlight the atoms lying on the
plane as shown in the Figure 10.5.

Figure 10.5: Red colored atoms showing (111) plane in a FCC unit cell.
Let us learn about few more tools that help visualize the planes in different ways:

• Spread of crystal: With this tool one can extend the crystal structure by desired number
of unit cells along any one or all the three directions. It can be assessed by path 2.

For example, if (111) plane is visualized while using this tool it will look like as shown in
figure 10.6(a) after translating the plane, as explained below.

Figure 10.6: (a) Atoms in (111) plane after spreading the FCC unit cell, (b) Similar plane
with the atoms removed using the Cut option.

• Changing color of the atoms lying in one plane: Once a plane has been identified as
explained above, double-click on any atom in the plane. A new window will open where
color attributes can be changed.

The Following tools are meaningful only when the crystal has been expanded using Spread of
crystal.

• Translation of the plane: A plane can be translated by one interplanar spacing at a time
by selecting the option Translation + plane 1 which can be assessed by following path
3.
• Cut: This helps to visualize a plane in a similar way as shown in Figure 10.6(b). For
this, follow path 2 and select Spread of crystal. In the window that opens enable the
Center option. This brings the origin point of the crystal unit-cell and origin of the X-
Y-Z coordinate together. Now, follow path 3 select Choice of (hkl) plane enable Cut
option. The will result into a crystal that will look similar to what is shown in Figure
10.6(b).
• Viewing the crystal along perpendicular to the plane: Follow path 3 and select
Projection perpendicular plane 1.

Directions

There are two ways by which a crystallographic direction (R = ua+vb+wc) can be visualized:

a. By providing values for u, v, w (integral values):


o Follow path 3 and select Choice of [uvw] direction.
b. By selecting any two atoms with computer mouse along the desired direction
o Follow path 3 and select? [uvw] with mouse.

Note: The limitation of this software is that when the direction is visualized within a single
unit cell, the direction doesn’t start from the origin linked to the unit cell (Fig. 10.7(a)). This
limitation can be overcome with the help of following trick:

For path 2 select Spread of Crystal, add 2-unit cells in all the three directions and select
Build from Center option. Now any specified [uvw] direction will start from the origin
attached to the unit cell as shown in Figure 10.7(b).

Figure 10.7: (a) Direction drawn in single unit cell, (b) Direction drawn after spreading the
crystal.
Distance between atoms, and angles between directions/planes in a crystal structure

• Distance between atoms: Follow path 2a and select Distance between atoms. Now
select two desired atoms (by mouse) and a pop-up window (fig. 8a) will display the
distance.
• Angles between directions / planes: Follow path 3 and select the appropriate option
out of the three options at the bottom of the sub-menu, whichever is. For example,
Figure 10.8(b) shows the angles between two planes.

Figure 10.8: (a) Showing atomic positions and the corresponding distance between two
selected atoms, and (b) (hkl) indices of the two planes and the corresponding angle are
shown.

To visualize a void in a crystal structure:


Following stepwise procedure shows how to visualize an octahedral void in FCC crystal
structure. Similar approach is applicable for other types of voids.

Approach #1:

1. Create a FCC unit cell


2. Follow path 1 and select Creation/List option. Here, you may change the lattice
parameter/radii/color etc. to get a better visualization of the crystal structure. For
example, as shown in Figure10.9 corner and face-centered atoms are colored
differently, or as shown in Figure 10.10, the corner atoms have been removed, for better
visualization.

Figure 10.9: A FCC unit cell with atoms at equivalent positions colored differently.

3. Follow path 2c and use Link option to connect desired atoms. Link allows only two
atoms to be connected at a time; the multi-link option results into computer hang. Save
the file before you try multi-link option.

Figure 10.10: An octahedral void with removed corner atoms.


Approach #2:

Note: This option works only when you there is a dummy atom at the center of the void to be
visualized. This dummy atom can be kept very small in size so that it can be recognized as a
dummy atom.

Figure 10.11: A dummy atom at the center of an octahedral void. It has been given blue
color to identify it separately. Also, the corner atoms have been removed for better
visualization.

1. Path1 Creation List, and use Add atom to incorporate an extra atom at the center of
the octahedral void i.e. at position co-ordinates (1/2, 1/2, 1/2). In Figure 10.11, this
atom has been given blue color.

2. Now we have to make a polyhedron connecting the atoms forming octahedral void.
For this, follow path 2a and select the Polyhedron search option.

3. Click on the dummy atom. Another window will open, as shown in Figure 10.12,
where a list of atoms will be present. To explain a little about this list, the top-most
entry in this list shows the first set of nearest neighbors for the dummy atom along
with total number of such atoms and their distance from the dummy atom. The second
entry shows the number of second nearest atoms and their distance from the dummy
atom, and so on.
Figure 10.12: The red polyhedron is formed by the set of 6 nearest atoms around the
octahedral void.

4. Since in the present case, the void (dummy atom) is surrounded by 6 nearest atoms,
we select the first entry in the window. Also enable Use Preview, it will show how the
polyhedron will look like. By hitting OK, the polyhedron will be made on the crystal
under consideration, and it will look similar to that in Figure 10.13.

Figure 10.13: An octahedral polyhedron formed by 6 face-centered atoms in a FCC


structure.

5. To get an array of octahedral polyhedrons as shown in Figure 10.14, for the FCC unit
cell where we just drew the polyhedron (Fig. 10.13), use Spread of Crystal. After
spreading the unit cell, the polyhedron may disappear, and therefore the polyhedron
has to be redrawn on the stretched crystal structure by following the steps 3 & 4.

Figure 10.14: An array of octahedral polyhedrons in FCC structures.


X-ray diffraction pattern to determine crystal-structure

X-ray diffraction (XRD) is an important characterization technique for obtaining the


information about crystal structure of a given material, identify the elements/phases by
comparison with the database of XRD pattern for known materials, find out surface-orientation
in case of a single crystal, and quantify the cell parameters, crystallite size as well as other
structural parameters.

X-rays are electromagnetic waves with wavelength in the range of 100 – 0.1 Å (10 - 0.01 nm);
the X-ray wavelength used for XRD technique is of the order of few angstroms, e.g. generally
used X-ray is Cu Kα = 1.54 Å. For comparison, the wavelength of the visible electro- magnetic
spectrum lies between 4000 – 7000 Å (400 – 700 nm). If we compare the magnitude of these
wavelengths with the typical value of inter-atomic distance in a material which is in the range
of few Angstroms (for Copper it is 2.55 Å), it is clear that the wavelengths for visible spectrum
are thousands of times longer compared to the inter-atomic distance, while that used for XRD
are comparable to the inter-atomic distance. This is why the X-rays experience the discreetness
of the material, and are diffracted by the constituent atoms. These diffracted X-rays from the
atoms interfere with each other constructively or destructively depending on their phase
difference, and it leads to the observation of intense diffraction peaks for particular incident
directions (angles). These strong diffraction peaks reveal the family of parallel sets of planes
that diffracted the X-ray present in the crystalline material, and this information in-turn, can be
used to unlock the information about the crystal-structure of that material. When this
observation of constructive and destructive interference is put together in a mathematical form,
a simplified version of it is given by the following expression (Eq.1), known as Bragg’s Law,
named after W. L. Bragg and W. H. Bragg who first proposed it. This expression can be
pictorially shown in Figure 10.15.

𝑁𝜆 = 2𝑑ℎ𝑘𝑙 ⁡𝑆𝑖𝑛θ (1)


In this expression for Bragg’s law:
• λ is the wavelength of incident monochromatic X-rays,
• 𝑑ℎ𝑘𝑙 ⁡in interplanar distance (distance between the two consecutive planes) from the set
of parallel planes having miller indices (hkl),
• θ is incident angle (angle between the X-ray beam and the surface), and
• N (an integer) is order of the diffracted beam.

Figure 10.15: Diffraction of X-rays from a set of (hkl) plane.


A typical XRD pattern generated using CaRIne Crystallography software for a simple cubic
structure having lattice parameter 3 Å (an arbitrary value which is very close to typical lattice
parameter for real elemental materials) is shown in Figure 10.16.

Figure 10.16: XRD pattern for the simple cubic crystal lattice

In Figure 10.16, the y-axis is intensity of the diffracted peaks. For x-axis, you might have
noticed that the angle is 2𝛉 and not 𝛉. This is because, as shown in figure 10.15, for the X-ray
beam incident at an angle 𝛉 the angle between the incident and diffracted direction is 2𝛉. To
put it in other words, in practice the X-ray source is kept stationary, and the sample is rotated,
and for a rotation of the sample by 𝛉 the detector has to be moved by 2𝛉.

Other information which can be obtained from a XRD pattern:

• Estimation of grain size from peak width:

The peak width ∂θ (in radians), measured at half the height of the intensity-peak, also known
as full-width-at-half-maximum (FWHM), is inversely proportional to the crystallite size 𝐷ℎ𝑘𝑙 .
The mathematical equation correlating the two parameters is given as:

𝜆
∂θ = ⁡ 𝐷 (2)
ℎ𝑘𝑙 𝑐𝑜𝑠θ
Here, ∂θ is peak width in radians at FWHM, λ is wavelength of the incident X-ray, 𝐷ℎ𝑘𝑙 is
crystallite-size in the direction normal to the (hkl) plane, θ is the Bragg angle for the peak under
consideration. You will notice that this equation is sometimes also identified by the name of
Scherrer Equation, in the literature. This equation (Eq. 2) is valid only when the crystallite size
is less than 100 nm. Also, note that this equation gives only rough estimates about the crystallite
size; a much better technique is electron microscope to measure the crystallite size.

• Selection Rule

Depending on the specific arrangement of atoms in a unit cell, for some of the (hkl) planes
where we expect strong peaks, there are no peaks! they are systematically absent from the XRD
pattern. The correlation between the arrangement of atoms in a unit cell and the absence of
particular (hkl) intensity peaks leads to selection-rules. Following table shows the selection
rule for a cubic crystal lattice i.e. for simple cubic, BCC and FCC Crystals. The information on
missing peaks can be used to get an idea about crystal- lattice for a material from its XRD
pattern. For details of the selection rule, you may refer to “Crystals and Crystal Structures, R.
Tilley, Wiley 2006” and “ Elements of X-ray Diffraction, B. D. Cullity, Addison-Wesley 1956”.

Lattice Reflection present Reflection absent


Simple Cubic For all h, k, l values None
Body-Centered Cubic (BCC) For h+k+l = even number For h+k+l = odd number
For h, k, l values mixed, i.e.
For h, k and l either all even
Face-Centered Cubic (FCC) some are even while
or all odd
remaining are odd

• Lattice Parameter

By noticing the 2θ position of a (hkl) peak the interplanar distance 𝒅𝒉𝒌𝒍 can be calculated using
Eqn. 1. The correlation between the interplanar distance 𝒅𝒉𝒌𝒍 for a (hkl) plane and the lattice
parameters a, b, c is as following:

𝟏 𝒉 𝒌 𝒍
= √(𝒂)𝟐 + (𝒃)𝟐 + (𝒄)𝟐 (3)
𝒅𝒉𝒌𝒍

Eq. 3 is only valid for crystal structures with 𝜶 = 𝜷 = 𝜸 = 𝟗𝟎° .

USING CARINE CRYSTALLOGRAPHY SOFTWARE TO GENERATE X-RAY


DIFFRACTION PATTERN:

Note: All word in the following sentences that are Bold, represent a tool option available in
CaRIne.

• Following is a stepwise procedure that allows one to generate a XRD pattern for
selected a crystal structure:

Step #1: Selecting the crystal system

a) On the software window, click the icon Cell and from the drop-down menu
select the desired crystal structure. Complete the process so that you see a
model of the crystal on the screen as shown below in figure 10.17.

Step #2: Generating the corresponding X-ray diffraction pattern

a) Clink the icon Special, and from the drop-down menu select XRD followed by
Creation XRD, as shown below in Figure 10.17.

b) Once you select Creation XRD, another window, as shown in Figure 10.18
will open. In this window the wavelength of the incident monochromatic X-ray,
and Bragg angle range has to be specified. Here the angle entered is regarded as
θ and not 2θ, i.e. if you select angle range 10 - 60°, the generated XRD pattern
will be shown for the angle range 20 - 120°.

Figure 10.17 Figure 10.18

c) Once all the parameters have been specified in step b), click OK to generate the
XRD pattern.

d) To save the raw data for the generated XRD pattern, right click on the pattern
and select “Save XRD as ASCII file” option. The data file can be exported in
ASCII code providing indexed planes, 2-theta, inter-planar spacing, structure
factor, and multiplicity factor. The data file looks like as shown in the Figure
10.19.

e) To save the XRD pattern as an image, right clink in the window having XRD
pattern and select copy. After this, the best way is to paste it in Paint application
available in windows OS, and from here the image can be saved in different
formats.

Figure 10.18: The data file after exporting as ASCII code


To Report:

Part A: Visualizing different characteristics of a crystal structure e.g. planes, directions,


voids etc.

For most of the crystal-structures that you will be drawing, there is no need to change the
default lattice parameters, unless mentioned otherwise. Of course, you may change the values
if it makes easier to visualize the structure.

Cubic Crystal:

1) Show a unit cell of Cubic Crystal Lattice (Primitive) and Cubic Crystal Structure
(Monatomic Basis/Motif).
2) In a Cubic Crystal structure, show the following two sets of planes separately. Show by
rotating the crystal, the direction along which the two planes in each set intersect. Also,
determine these directions. The lattice parameter is a = 6Å..

i. Plane # 1 intersects the axis at a, a/3, a/4 and Plane # 2 intersects the axis at 3a, a, 3a/4.
ii. Plane # 1 intersects the axis at 2a, 4a, 3a and Plane # 2 intersects the axis at -3a, 3a, ∞

BCC Crystal:

1) Show two-unit cells of a Body-Centered-Cubic crystal such that they share one plane. Now
show an octahedral void.
2) Based on the observation, find out the effective number of the octahedral voids in a BCC
unit cell.

FCC Crystal:

1) There are 8 tetrahedral voids in a FCC unit cell. Show any one of them (Hint: The void is
surrounded by at least one of the atoms located at the corner of the unit cell)

HCP Crystal:

1) Draw a unit cell of Hexagonal Crystal Structure. Now using the spread of crystal function,
add two-unit cells along each of the three axes.

2) View this crystal along [001] direction; now the (001) plane as seen on the screen is a closed
packed plane. Now, if you want to make a HCP crystal structure, these closed packed planes
have to be stacked in a ABAB….. sequence, whereas for FCC these planes are stacked in a
ABCABC…. sequence.

Copy the picture of the closed packed plane as seen on the screen into your answer sheet
and assuming this is plane A, identify the sites on this plane where the atoms of plane B and C
will have to be placed in order to make the above two sequences.

3) Draw a hexagonal unit cell. Add an atom at location 1/3a,2/3b,1/2c in this unit cell; now this
modified unit cell is HCP unit cell. Identify and show the two tetrahedral voids in this HCP
unit cell.
Slip Plane and Slip Directions:

The edge dislocation plays an important role in plastic deformation of crystalline materials.
Under an external the shear component of the stress causes these dislocations to slip/move in a
crystalline material. The plane on which the dislocation line slips are called slip planes; these
planes have maximum number of atoms per unit area. The directions along which dislocations
slip are called slip directions; these directions have maximum number of atoms per unit length.

Now, for Copper (lattice parameter = 3.61Å; calculate the atomic radius yourself):

1) Show (111), (110), (100) and (-212) planes.

2) Using the software, determine the shortest distance between two atoms in these planes along
with the corresponding crystallographic direction. This is the direction with highest packing
density of atoms per unit length.

3) Calculate the interplanar spacings (d) for these planes (this doesn’t require software).

4) Calculate the number density of atoms in these planes (# of atoms per unit area of the
plane).

5) Analyse the numbers obtained for (2) to (4) and predict the slip plane and slip direction.

Part B: X-ray diffraction pattern to determine crystal-structure

1) Generate a XRD pattern for a simple cubic crystal s for the 2θ range 10 - 120°. Assume
lattice parameter to be 3 Å.

2) Using the experimental XRD pattern calculate the lattice parameter.

3) Draw three hypothetical FCC unit cells with lattice parameters 1, 2, 3 Å, and generate
corresponding X-ray diffraction patterns in the range (θ) 10-60°. Notice the position of (111)
peak for each. Can you explain this observation?

Questions:

1. Why are some planes present, but others are absent in the X-Ray diffraction?
2. If diffraction takes place at a particular angle (2θ), why does the diffraction pattern
show a broad peak and not a sharp line?
3. What are the sources of peak broadening?
EXPERIMENT 11

(i) Differential scanning calorimetry (DSC) - Phase transformation in


metals
(ii) X-Ray diffraction – Determination of crystallite size/strain

Objective:

• To study the phase transformation behavior of Pb-Sn sample using DSC (Differential
Scanning Calorimetry).
• Determination of crystallite size/strain using X-Ray diffraction (XRD)

Introduction:

Differential scanning calorimetry (DSC)

Differential Scanning Calorimetry (DSC) is a technique that allows us to determine the


enthalpies associated with different types of phase transitions or other physical processes. An
experimental sample contained in pan and a reference empty pan are placed in a thermally
controlled chamber (Figure 11.1) whose temperature is raised at a constant rate (e.g. 10°C/min),
while the amount of heat flowing in and out of the two pans is monitored and compared.

Figure 11.1: Schematic of DSC sample chamber

It is important that there be good thermal contact between the sample and the pan. The basic
principle underlying this technique is that when the sample undergoes a physical
transformation such as phase transitions, more or less heat will need to flow to it than the
reference to maintain both at the same temperature. Whether less or more heat must flow to the
sample depends on whether the process is exothermic or endothermic. For example, as a solid
sample melts to a liquid, it will require more heat flowing to the sample to increase its
temperature at the same rate as the reference. This is due to the absorption of heat by the sample
as it undergoes the endothermic phase transition from solid to liquid. Likewise, as the sample
undergoes exothermic processes (such as crystallization) less heat is required to raise the
sample temperature. By observing the difference in heat flow between the sample and
reference, differential scanning calorimeters are able to measure the amount of heat absorbed
or released during such transitions. DSC may also be used to observe more subtle physical
changes, such as glass transitions. Gas purging is done to avoid oxidation of the sample.
The result of a DSC experiment is a curve of heat flux versus temperature or versus time, see
Figure 11.2. There are two different conventions: exothermic reactions in the sample shown
with a positive or negative peak, depending on the kind of technology used in the experiment.
This curve can be used to calculate enthalpies of transitions. This is done by integrating the
peak corresponding to a given transition. It can be shown that the enthalpy of transition can
be expressed using the following equation:

Δ𝐻 = 𝐾𝐴

where Δ𝐻 is the enthalpy of transition, K is the calorimetric constant, and A is the area under
the curve. The calorimetric constant will vary from instrument to instrument, and can be
determined by analyzing a well-characterized sample with known enthalpies of transition.

Fig 11.2: Typical DSC curves for various cooling rates

There are various experimental and environmental parameters to consider during DSC
measurements. Exemplary potential issues are briefly discussed in the following sections.

Crucibles: DSC measurements without crucibles promote the thermal transfer towards the
sample and are possible if the DSC is designed for this purpose. Measurements without crucible
should only be conducted with chemically stable materials at low temperatures, as otherwise
there may be contamination or damage of the calorimeter. The safer way is to use a crucible,
which is specified for the desired temperatures and does not react with the sample material (e.g.
alumina, gold or platinum crucibles). If the sample is likely to evolve volatiles or is in the liquid
state, the crucible should be sealed to prevent contamination. However, if the crucible is sealed,
increasing pressure and possible measurement artefacts due to deformation of the crucible must
be considered. In this case, crucibles with very small holes (∅~50 µm) or crucibles that can
withstand very high pressures should be used.

Sample condition: The sample should be in good contact with the crucible surface. Therefore,
the contact surface of a solid bulk sample should be plane parallel. For DSC measurements
with powders, stronger signal might be observed for finer powders due to the enlarged contact
surface. The minimum sample mass depends on the transformation to be analyzed. A small
sample mass (~10 mg) is sufficient if the released or consumed heat during the transformation
is high enough. Heavier samples could be used to obtain transformation associated with low
heat release or consumption, as larger samples also enlarge the obtained peaks. However, the
increasing sample size might worsen the resolution due to thermal gradients which may evolve
during heating.

Temperature and scan rates: If the peaks are very small, it is possible to enlarge them by
increasing the scan rate. Due to the faster scan rate, more energy is released or consumed in a
shorter time which leads to higher and therefore more distinct peaks. However, faster scan rates
lead to poor temperature resolution because of thermal lag. Due to this thermal lag, two phase
transformations (or chemical reactions) occurring in a narrow temperature range might overlap.
Generally, heating or cooling rates are too high to detect equilibrium transitions, so there is
always a shift to higher or lower temperatures compared to phase diagrams representing
equilibrium conditions.

Purge gas: Purge gas is used to control the sample environment, in order to reduce signal noise
and to prevent contamination. Mostly nitrogen is used and for temperatures above 600 °C,
argon can be utilized to minimize heat loss due to the low thermal conductivity of argon. Air
or pure oxygen can be used for oxidative tests like oxidative induction time, and He is used for
very low temperatures due to the low boiling temperature (~4.2K at 101.325 kPa).

X-Ray diffraction – Determination of crystallite size/strain

X-ray diffraction (XRD) is an important characterization technique for obtaining the


information about crystal structure of a given material, identify the elements/phases by
comparison with the database of XRD pattern for known materials, find out surface-orientation
in case of a single crystal, and quantify the cell parameters, crystallite size as well as other
structural parameters.

X-rays are electromagnetic waves with wavelength in the range of 100 – 0.1 Å (10 - 0.01 nm);
the X-ray wavelength used for XRD technique is of the order of few angstroms, e.g. generally
used X-ray is Cu Kα = 1.54 Å. For comparison, the wavelength of the visible electro- magnetic
spectrum lies between 4000 – 7000 Å (400 – 700 nm). If we compare the magnitude of these
wavelengths with the typical value of inter-atomic distance in a material which is in the range
of few Angstroms (for Copper it is 2.55 Å), it is clear that the wavelengths for visible spectrum
are thousands of times longer compared to the inter-atomic distance, while that used for XRD
are comparable to the inter-atomic distance. This is why the X-rays experience the discreetness
of the material, and are diffracted by the constituent atoms. These diffracted X-rays from the
atoms interfere with each other constructively or destructively depending on their phase
difference, and it leads to the observation of intense diffraction peaks for particular incident
directions (angles). These strong diffraction peaks reveal the family of parallel sets of planes
that diffracted the X-ray present in the crystalline material, and this information in-turn, can be
used to unlock the information about the crystal-structure of that material. When this
observation of constructive and destructive interference is put together in a mathematical form,
a simplified version of it is given by the following expression (Eq.1), known as Bragg’s Law,
named after W. L. Bragg and W. H. Bragg who first proposed it. This expression can be
pictorially shown in Figure 11.3.
𝑁𝜆 = 2𝑑ℎ𝑘𝑙 ⁡𝑆𝑖𝑛θ (1)
In this expression for Bragg’s law:
• λ is the wavelength of incident monochromatic X-rays,
• 𝑑ℎ𝑘𝑙 ⁡in interplanar distance (distance between the two consecutive planes) from the set
of parallel planes having miller indices (hkl),
• θ is incident angle (angle between the X-ray beam and the surface), and
• N (an integer) is order of the diffracted beam.

Figure 11.3: Diffraction of X-rays from a set of (hkl) plane.

A typical XRD pattern generated using CaRIne Crystallography software for a simple cubic
structure having lattice parameter 3 Å (an arbitrary value which is very close to typical lattice
parameter for real elemental materials) is shown in Figure 11.4.

Figure 11.4: XRD pattern for the simple cubic crystal lattice

In Figure 11.4, the y-axis is intensity of the diffracted peaks. For x-axis, you might have noticed
that the angle is 2θ and not θ. This is because, as shown in figure 10.15, for the X-ray beam
incident at an angle θ the angle between the incident and diffracted direction is 2θ. To put it in
other words, in practice the X-ray source is kept stationary, and the sample is rotated, and for
a rotation of the sample by θ the detector has to be moved by 2θ.
Depending on the specific arrangement of atoms in a unit cell, for some of the (hkl) planes
where we expect strong peaks, there are no peaks! they are systematically absent from the XRD
pattern. The correlation between the arrangement of atoms in a unit cell and the absence of
particular (hkl) intensity peaks leads to selection-rules. Following table shows the selection
rule for a cubic crystal lattice i.e. for simple cubic, BCC and FCC Crystals. The information on
missing peaks can be used to get an idea about crystal- lattice for a material from its XRD
pattern. For details of the selection rule, you may refer to “Crystals and Crystal Structures, R.
Tilley, Wiley 2006” and “ Elements of X-ray Diffraction, B. D. Cullity, Addison-Wesley 1956”.

Lattice Reflection present Reflection absent


Simple Cubic For all h, k, l values None
Body-Centered Cubic (BCC) For h+k+l = even number For h+k+l = odd number
For h, k, l values mixed, i.e.
For h, k and l either all even
Face-Centered Cubic (FCC) some are even while
or all odd
remaining are odd

By noticing the 2θ position of a (hkl) peak the interplanar distance 𝒅𝒉𝒌𝒍 can be calculated using
Eq. 1. The correlation between the interplanar distance 𝒅𝒉𝒌𝒍 for a (hkl) plane and the lattice
parameters a, b, c is as following:

𝟏 𝒉 𝒌 𝒍
= √( )𝟐 + ( )𝟐 + ( )𝟐 (2)
𝒅𝒉𝒌𝒍 𝒂 𝒃 𝒄

Eq. 2 is only valid for crystal structures with 𝜶 = 𝜷 = 𝜸 = 𝟗𝟎° .

Crystallite size/strain from XRD

It is well known that because of size confinement and presence of intrinsic strain, which
originates due to size confinement, broaden the XRD peaks. Thus, a physical peak broadening
primarily consists of two parts; size-dependent broadening and strain induced broadening. The
most common sources of the lattice strain are dislocation density, point defects, grain
boundary junction, contact or sinter stress, and stacking faults. So, from the XRD peak
broadening analysis, the crystallite size as well as the value of the intrinsic strain, including
other elastic properties such as stress, energy density, which are related to strain, can be
determined indirectly. There are many methods for this, such as Williamson-Hall Method,
Warren-Averbach Method and Balzar Method. Warren-Averbach and Balzar method consider
Stokes fourier de-convolution method, whereas Williamson-Hall method uses the full-width-
at-half-maximum (FWHM) of the diffraction peak and hence, it is very easy and suitable one
for determination of different elastic properties including strain, along with average size
calculation. According to Williamson-Hall (W-H) method, physical line broadening of X-
ray diffraction peak occurs due to the size and micro strain of the nanocrystals and the total
broadening can be written as:

𝛽𝑡𝑜𝑡𝑎𝑙 = 𝛽𝑠𝑖𝑧𝑒 + 𝛽𝑠𝑡𝑟𝑎𝑖𝑛 (3)


Average crystallite size and the strain can be calculated using of W-H method which is also
known as Uniform Deformation Model.
Uniform Deformation Model (UDM)

Uniform deformation model (UDM) considers uniform strain throughout the crystallographic
direction, which gets introduced due to crystal imperfections. In other words, UDM considers
strain, which is isotropic in nature. This intrinsic strain actually effects the physical broadening
of the XRD profile, and this the strain (𝜀) induced peak broadening can be expressed as:

⁡𝛽𝑠𝑡𝑟𝑎𝑖𝑛 = 4𝜀. 𝑡𝑎𝑛𝜃 (4)

Size (D) induced peak broadening can be expressed as following within Scherrer formalism:

𝑘𝜆
𝛽𝑠𝑖𝑧𝑒 = ⁡ 𝐷𝑐𝑜𝑠θ (5)
So, the total broadening due to strain and size in a particular peak having the hkl value, can be
expressed as:
𝑘𝜆
𝛽ℎ𝑘𝑙 = ⁡ 𝐷𝑐𝑜𝑠θ + 4𝜀. 𝑡𝑎𝑛𝜃 (6)

𝑘𝜆
𝛽ℎ𝑘𝑙 . 𝑐𝑜𝑠θ = ⁡ + 4𝜀. 𝑠𝑖𝑛𝜃 (7)
𝐷

Where, 𝛽ℎ𝑘𝑙 is the full width at half of the maximum intensity for different diffraction planes.
Eq. 7 is an equation of a straight line and is known as the uniform deformation model (UDM)
equation, which considers the isotropic nature of the crystals. The slope of this straight line
with the term (4Sinθ) along X-axis and (𝛽ℎ𝑘𝑙 .cosθ) along Y-axis provides the value of the
intrinsic strain, whereas the intercept gives the average particle size.

UDM model is based on the assumption that sample is homogeneous and isotropic in nature,
which is not actually justified for a real crystal. As a crystal is anisotropic, modified
Williamson Hall equation are developed such as Uniform Stress Deformation Model (USDM)
and Uniform Deformation Energy Density model (UDEDM). These methods are not in the
scope of this experiment.

To Report:

1. Plot ΔH Vs T for each sample from given data


2. Calculate the peak area (mJ) during heating for each sample, transformation
temperature, transition start temperature
3. Calculate the crystallite size and dislocation density for various sample conditions

Questions
1. Can we measure recrystallization like transformation with DSC? Why or why not?
2. What is difference between crystallite size and grain size?
EXPERIMENT 12

(i) Demonstration and analysis of Raman characterization technique


(ii) Demonstration of sample preparation for TEM

Objective:

• To understand the principles and functionality of Raman Spectroscopy


• Sample preparation for TEM

Introduction:

Demonstration and analysis of FTIR, Raman and other advanced characterization


technique

Fourier Transform Infrared Spectrometry (FTIR)

Infrared spectroscopy has been a workhorse technique for materials analysis in the laboratory
for over seventy years. An infrared spectrum represents a fingerprint of a sample with
absorption peaks which correspond to the frequencies of vibrations between the bonds of the
atoms making up the material. Because each different material is a unique combination of
atoms, no two compounds produce the exact same infrared spectrum. Therefore, infrared
spectroscopy can result in a positive identification (qualitative analysis) of every different kind
of material. In addition, the size of the peaks in the spectrum is a direct indication of the amount
of material present. With modern software algorithms, infrared is an excellent tool for
quantitative analysis. Infrared spectroscopy is useful for several types of analysis such as:

• Identification of unknown materials


• Determination of the quality or consistency of a sample
• Determination of the amount of components in a mixture

The original infrared instruments were of the dispersive type. These instruments separated the
individual frequencies of energy emitted from the infrared source. This was accomplished by
the use of a prism or grating. An infrared prism works exactly the same as a visible prism which
separates visible light into its colors (frequencies). A grating is a more modern dispersive
element which better separates the frequencies of infrared energy. The detector measures the
amount of energy at each frequency which has passed through the sample. This results in a
spectrum which is a plot of intensity vs. frequency. FTIR stands for Fourier Transform
InfraRed, the preferred method of infrared spectroscopy. Fourier transform infrared
spectroscopy is preferred over dispersive or filter methods of infrared spectral analysis for
several reasons:

• It is a non-destructive technique
• It provides a precise measurement method which requires no external calibration
• It can increase speed, collecting a scan every second
• It can increase sensitivity – one second scans can be co-added together to ratio out
random noise
• It has greater optical throughput
• It is mechanically simple with only one moving part

FTIR spectrometry was developed in order to overcome the limitations encountered with
dispersive instruments. The main difficulty was the slow scanning process. A method for
measuring all of the infrared frequencies simultaneously, rather than individually, was needed.
A solution was developed which employed a very simple optical device called an
interferometer.

The interferometer produces a unique type of signal which has all of the infrared frequencies
“encoded” into it. The signal can be measured very quickly, usually on the order of one second
or so. Thus, the time element per sample is reduced to a matter of a few seconds rather than
several minutes.

Most interferometers employ a beamsplitter which takes the incoming infrared beam and
divides it into two optical beams. One beam reflects off of a flat mirror which is fixed in place.
The other beam reflects off of a flat mirror which is on a mechanism which allows this mirror
to move a very short distance (typically a few millimeters) away from the beamsplitter. The
two beams reflect off of their respective mirrors and are recombined when they meet back at
the beamsplitter. Because the path that one beam travels is a fixed length and the other is
constantly changing as its mirror moves, the signal which exits the interferometer is the result
of these two beams “interfering” with each other. The resulting signal is called an interferogram
which has the unique property that every data point (a function of the moving mirror position)
which makes up the signal has information about every infrared frequency which comes from
the source. This means that as the interferogram is measured, all frequencies are being
measured simultaneously. Thus, the use of the interferometer results in extremely fast
measurements. Because the analyst requires a frequency spectrum (a plot of the intensity at
each individual frequency) in order to make an identification, the measured interferogram
signal cannot be interpreted directly. A means of “decoding” the individual frequencies is
required. This can be accomplished via a well-known mathematical technique called the
Fourier transformation. This transformation is performed by the computer which then presents
the user with the desired spectral information for analysis.

The normal instrumental process is as follows:


The Source: Infrared energy is emitted from a glowing black-body source. This beam passes
through an aperture which controls the amount of energy presented to the sample (and,
ultimately, to the detector).

The Interferometer: The beam enters the interferometer where the “spectral encoding” takes
place. The resulting interferogram signal then exits the interferometer.

The Sample: The beam enters the sample compartment where it is transmitted through or
reflected off of the surface of the sample, depending on the type of analysis being
accomplished. This is where specific frequencies of energy, which are uniquely characteristic
of the sample, are absorbed.

The Detector: The beam finally passes to the detector for final measurement. The detectors
used are specially designed to measure the special interferogram signal.
The Computer: The measured signal is digitized and sent to the computer where the Fourier
transformation takes place. The final infrared spectrum is then presented to the user for
interpretation and any further manipulation

Because there needs to be a relative scale for the absorption intensity, a background spectrum
must also be measured. This is normally a measurement with no sample in the beam. This can
be compared to the measurement with the sample in the beam to determine the “percent
transmittance.” This technique results in a spectrum which has all of the instrumental
characteristics removed. Thus, all spectral features which are present are strictly due to the
sample. A single background measurement can be used for many sample measurements
because this spectrum is characteristic of the instrument itself.

Figure 12.1: Typical workflow of FTIR

Figure 12.2: Typical layout of FTIR spectrometer


Some of the major advantages of FT-IR over the dispersive technique include:

Speed: Because all of the frequencies are measured simultaneously, most measurements by
FT-IR are made in a matter of seconds rather than several minutes. This is sometimes referred
to as the Felgett Advantage.

Sensitivity: Sensitivity is dramatically improved with FT-IR for many reasons. The detectors
employed are much more sensitive, the optical throughput is much higher (referred to as the
Jacquinot Advantage) which results in much lower noise levels, and the fast scans enable the
coaddition of several scans in order to reduce the random measurement noise to any desired
level (referred to as signal averaging).

Mechanical Simplicity: The moving mirror in the interferometer is the only continuously
moving part in the instrument. Thus, there is very little possibility of mechanical breakdown.

Internally Calibrated: These instruments employ a HeNe laser as an internal wavelength


calibration standard (referred to as the Connes Advantage). These instruments are self-
calibrating and never need to be calibrated by the user.

These advantages, along with several others, make measurements made by FT-IR extremely
accurate and reproducible. Thus, it a very reliable technique for positive identification of
virtually any sample. The sensitivity benefits enable identification of even the smallest of
contaminants. This makes FT-IR an invaluable tool for quality controlor quality assurance
applications whether it be batch-to-batch comparisons to quality standards or analysis of an
unknown contaminant. In addition, the sensitivity and accuracy of FT-IR detectors, along with
a wide variety of software algorithms, have dramatically increased the practical use of infrared
for quantitative analysis. Quantitative methods can be easily developed and calibrated and can
be incorporated into simple procedures for routine analysis.

Thus, the Fourier Transform Infrared (FT-IR) technique has brought significant practical
advantages to infrared spectroscopy. It has made possible the development of many new
sampling techniques which were designed to tackle challenging problems which were
impossible by older technology. It has made the use of infrared analysis virtually limitless.

Raman Spectroscopy

Raman spectroscopy is the study of matter by the inelastic scattering of monochromatic light.
It has become a ubiquitous tool in modern spectroscopy, biophysics, microscopy,
geochemistry, and analytical chemistry. In contrast to typical absorption or emission
spectroscopy experiments, transitions among quantum levels of atoms or molecules are
induced by the absorption or emission of photons (IR, visible, UV). Raman spectroscopy is
much less sensitive than absorption or emission spectroscopies, because of the inherent
weakness of the scattering process, but has many intrinsic advantages, including freedom to
choose an incident wavelength which is not absorbed by the surrounding media (especially
useful for aqueous or mineral samples which have strong IR absorption bands), small volumes
probed (the light can be focused to micron-sized spots), and symmetry-based selection rules
which allow transitions that are ‘optically forbidden’ in absorption to be detected in scattering.

Monochromatic light incident on a transparent substance is transmitted with almost no


attenuation. A small fraction of the light is scattered by the substance in all directions (though
preferentially in the forward direction). The weakly scattered radiation contains photons at the
incident frequency 𝜐0 (elastic or Rayleigh scattering), but also contains other frequencies such
as 𝜐0 − 𝜐𝑖 , where 𝜐𝑖 is the frequency of a molecular transition (typically rotational or
vibrational) of the material. This inelastic light scattering is known as Raman scattering. The
effect, discovered by the Indian physicist Chandrasekhara Venkata Raman, has become
especially useful to spectroscopists since the advent of lasers, which can provide intense
sources of monochromatic light.

Figure 12.3: Typical layout of Raman spectrometer

In a typical Raman experiment, a polarized monochromatic light source (usually a laser) is


focused into a sample, and the scattered light at 90° to the laser beam is collected and dispersed
by a high-resolution monochromator. The incident laser wavelength (chosen such that the
sample does not absorb, in ordinary Raman Spectroscopy) is fixed, and the scattered light is
dispersed and detected to obtain the frequency spectrum of the scattered light. The scattered
light is very weak (<10-7 of the incident power), so that monochromators with excellent
straylight rejection and sensitive detectors are required.

A Raman spectrum consists of a series of peaks, displaced to higher and lower frequencies
from the incident frequency (the Rayleigh peak). In contrast to absorption spectroscopy, the
initial photon is not resonant with a transition. The peaks to lower frequency are known as
the Stokes lines, while those at higher frequencies are known as anti-Stokes lines, and are
usually considerably weaker than the Stokes Lines. Since the same frequencies appear in each
set of lines, only the Stokes spectrum is usually recorded. The spectrum contains lines due to
both “Raman-active” vibrational and rotational transitions. In the latter case, the
displacements from the incident light are very small (10-100 cm-1) and very high resolution
monochromators are required. Typically, the vibrational transitions are studied. Thus, Raman
spectroscopy is a branch of vibrational spectroscopy, and is complementary to IR absorption
spectroscopy, since many transitions observed in the Raman spectrum are not observed in the
IR spectrum, and vice versa. The appearance of lines in the Raman spectrum is determined
by selection rules, which are often different from those for absorption. A nonlinear polyatomic
molecule containing N atoms has 3N-6 modes of vibration, known as “normal modes”, each
associated with a particular fundamental frequency of vibration and symmetry. By measuring
the IR and Raman spectrum, all these modes may be observed, and the fundamental
frequencies measured. In many cases, the Raman spectrum alone is sufficient. Some
information on the symmetry of the normal modes is obtained from the degree of polarization
of the scattered light. These data often allow the molecular structure to be determined. In
addition, the shape of Raman lines of liquid samples may yield information on dynamical
processes in the liquid state, such as rotational and vibrational relaxation.

Figure 12.4: Energy level diagram showing the states involved in the Raman Spectra

An important observable is the extent of depolarization of the scattered light. When the
incident light is polarized, then the scattered light may retain the initial polarization, but it
may also become depolarized, with some of the light possessing a polarization perpendicular
to the incident light. The depolarization of the scattered light can be expressed by
𝐼⊥
𝜌=
𝐼||

where 𝐼|| and 𝐼⊥ are the intensities of the scattered light polarized parallel or perpendicular to
the exciting light, respectively.

Raman spectroscopy has various applications. For example, Raman spectroscopy is used in
chemistry to identify molecules and study chemical bonding and intramolecular bonds.
Because vibrational frequencies are specific to a molecule's chemical bonds and symmetry (the
fingerprint region of organic molecules is in the wavenumber range 500–1,500 cm−1), Raman
provides a fingerprint to identify molecules. In solid-state physics, Raman spectroscopy is used
to characterize materials, measure temperature, and find the crystallographic orientation of a
sample. As with single molecules, a solid material can be identified by
characteristic phonon modes.

Demonstration of sample preparation for Transmission electron microscopy (TEM)

For observation under transmission electron microscope, the sample has to be transparent to
the electron (i.e. sample should be thin enough to allow transmission of electrons through the
sample). Thicker samples will attenuate the electrons completely and will limit their detection
on the other side of the surface. In addition, higher energy electron beam may work for thicker
samples (say 0.5-1.0 µm samples), but will also damage the sample. In summary, the actual
features of sample may get modified due to high-energy electron interaction with matter
(during imaging itself). Thus, it is wiser to ensure that the sample preparation or beam energy
does not modify the real features (which may otherwise lead to improper conclusions). Usually,
the steps involved in TEM sample preparation from a bulk sample include slicing, grinding to
less than 100 µm, punching a 3 mm disc, followed by dimpling down to 2-10 µm, and ending
with jet polishing and/or ion-beam milling (see figure 12.5).

Figure 12.5: The sample preparation involves using: the starting bulk sample, sectioning the
sample to ~150-300 µm, thinning it down to < 100 µm by grinding, punching a 3mm disc,
followed by dimpling down to ~2-10 µm, and jet-polishing/ion-beam milling

Step 1: The first step in TEM sample preparation is to slice/cut the material down to < 100 µm.
The process of thinning can be achieved by sectioning the sample with slow speed diamond
saw or ultrasonic cutter or electro-discharge cutting. As the starting material may be in bulk
form (a few mm to cm in sample thickness), thinning process starts with sectioning the material
using diamond saw to obtain a thin sample (which may usually have thickness exceeding >
150-250 µm). Now further processing is required for being able to make this material
observable under TEM.

Step 2: Slicing is followed by grinding the surface to remove surface damage (induced during
the sectioning process). The damage area is typically 2-5 times the blade thickness. Therefore,
it is essential to utilize thinner blades at low speeds to minimize surface damage. Once the
sample is sectioned (> 100 µm) , the damaged surface (from slicing) should be removed by
grinding process (from using coarser to finer emery papers) till the surface is prepared and
brought down to thickness < 100 µm (say 50-75 µm).

Step 3: Thereafter, usually, a few of 3 mm diameter samples are punched out of this sheet
(using a punch machine). Punching out of 3 mm disc samples may also induce edge-damage,
hence enough care is required for this step as well. Nonetheless, the sample is prepared for
TEM imaging at center of the disc, hence some edge-damage can also be tolerated in this
sample-preparation process. Further, multiple samples are prepared as only a few may provide
the relevant information the user is looking for. Therefore, multiple samples are prepared for
observation under TEM.

Step 4: These punched out samples are now dimpled using dimpler (Fig. 12.6a). In this process
a vertical grinding disc is utilised to remove the material from the center, while the sample
itself is also rotated horizontally. By this process, a spherical cavity is produced in the center,
while the edges remain thick (which are useful in sample handling). The dimpling process can
be done on one surface (Fig. 12.6b) or both the surfaces (both top and bottom), Fig. 12.6c. A
glue is utilised to stick the sample to a mount and then heated to remove the glue. The sample
surface can be reversed for two-side dimpling (Fig. 12.6c). Though more tedious, the double-
dimpling process provides wider area for sample observation. The thinning is performed till
the sample become very thin (say ~ 2-5 µm thick).

Figure 12.6: a) TEM sample Dimpler showing that a vertical rotating grinder thins the TEM
disc sample and the specimen itself rotates to provide a b) one-side dimple, or c) two-side
dimple in the center of the sample

Step 5: Either electrochemical etching (for conducting samples) or ion beam milling (for all
samples) may be utilised for the final sample preparation. In electropolishing, the electrolyte
is sprayed (in jet form) on the sample until a through hole is created (using a through-light
automatic detector). The immediate region around this through hole is actually a electron-
transparent region which can be observed under TEM. Surface has to be immediately washed
in order to remove any chemical, which may further damage the surface. If process goes
unchecked, a bigger perforation (or multiple perforations) may form, which, again, will hinder
the observation of microstructure and induce other imaging artefacts.

Another way of preparing final sample for TEM observation is ion beam milling. This process
utilizes either a single or double ion mill, usually placed at ~ 6° incident angle from surface,
to create a gradual removal of material in form of conical cavity. This process is very slow,
but creates very good surface (with almost no surface damage) for TEM observation.

Once the 3 mm diameter disc is ready (with a certain region transparent to electrons), this disc
is placed in TEM holder. This TEM holder is inserted to reach the central region of TEM
through which electron beam passes and the sample can be imaged. A low temperature is
maintained near the sample (i.e. liquid nitrogen environment) to minimize any damage in the
sample from electron beam.

To Report:
1. List the materials identified using Raman Spectroscopy. What wavelength shift do they
correspond to?
2. What is the target thickness for sample preparation for TEM? Why this thin sample is
desired?
3. What methods were used to thin down the sample?

Questions:
1. Which is more sensitive in terms of identifying the material – Raman or FTIR? Why?
2. Can we use ion milling for ceramic samples? What are the other techniques available
for thinning sample, other than those discussed here?
3. Why is TEM needed?

You might also like