Vous êtes sur la page 1sur 24

Thermal Analysis:

Thermal analysis includes a group of methods by which the physical and chemical properties of a
substance, a mixture and/or reaction mixtures are determined as a function of temperature or
time, while the sample is subjected to a controlled temperature program. The program may
involve heating or cooling (dynamic), or holding the temperature constant (isothermal), or any
combination of these.
The adjective is thermoanalytical. The graphic result is the thermal analysis curve.
See, for example, German standards DIN 51005 (nomenclature) or American standard ASTM E
473 (nomenclature).
DIFFERENTIAL THERMAL ANALYSIS (DTA)
DIFFERENTIAL SCANNING CALORIMETRY (DSC)
DIN 51007, DIN 53765, ASTM E 474, ASTM D 3418
These are techniques by which the difference in heat flow to or from a sample and to or from a
reference is monitored as a function of temperature or time, while the sample is subjected to a
controlled temperature program.
Applications:
characteristic temperatures
identification
glass transitions
melting and crystallization behavior
heat of melting and crystallization
purity
compatibility
polymorphism
solid-liquid ratio
specific heat capacity
reaction behavior
heat of reaction
reaction kinetics
oxidative stability
thermal stability
THERMOGRAVIMETRY (TG)
DIN 51006, ASTM E 914, ASTM E 1131
This is a technique by which the mass of the sample is monitored as a function of temperature or
time, while the sample is subjected to a controlled temperature program.
Dm = mass change
dm/dt = rate of mass change/decomposition
DTG = derivative thermogravimetry
DTG Peak = characteristic decomposition temperatures identification
Tonset = thermal stability
composition = moisture content, solvent content, additives, polymer content, filler content
dehydration
decarboxylation
oxidation
decomposition

SIMULTANEOUS THERMAL ANALYSIS


This technique combines thermogravimetry with differential thermal analysis or differential
scanning calorimetry in one run.
- possible to consider the real sample mass at a given temperature in C p determination
- no temperature differences between signals of TG and DTA/DSC measurement
THERMOMECHANICAL ANALYSIS (TMA)
DILATOMETRY (DIL)
DIN 51045, ASTM E 831, ASTM D 696, ASTM D 3386
This is a technique by which the dimensional changes of the sample are monitored as a function
of temperature or time, while the sample may be subjected to an additional mechanical load.
- negligible mechanical load dilatometry
- constant load (static) thermomechanical analysis
- oscillating force dynamic mechanical analysis
DL = length change
DL / L0 = relative length change
a = coefficient of linear thermal expansion
Tg = glass transition temperature
expansion/shrinkage/penetration behavior
dimensional stability design purposes
softening/curing behavior
volumetric expansion ( bulk density)
DYNAMIC MECHANICAL ANALYSIS (DMA)
DIN 53513, DIN 53440, ASTM D 4065, ASTM D 4092
This is a thermoanalytical method by which the mechanical behavior of a sample subjected to a
specific temperature program is investigated under the effect of a load which changes with time.
Determination of the storage and loss moduli and the damping factor of a sample as a function of
temperature, time and frequency of an applied oscillating load.

Infrared Absorption Spectroscopy (IR)


Introduction
IR spectroscopy is the measurement of the wavelength and intensity of the absorption of
mid-infrared light by a sample. Mid-infrared light (2.5 - 50 m, 4000 - 200 cm-1) is

energetic enough to excite molecular vibrations to higher energy levels. The wavelength
of IR absorption bands are characteristic of specific types of chemical bonds, and IR
spectroscopy finds its greatest utility for identification of organic and organometallic
molecules.
Mechanism of IR absorption
The transition moment for infrared absorption is:
R = < Xi | u | Xj dt >
where Xi and Xj are the initial and final states, respectively, and u is the electric dipole
moment operator:
u = uo + (r-re)(du/dr) + ... higher terms.
uo is the permanent dipole moment, which is a constant, and since < Xi | Xj > = 0, R
simplifies to:
R = < Xi | (r-re)(du/dr) | Xj >
The result is that there must be a change in dipole moment during the vibration for a
molecule to absorb infrared radiation.
Examples of infrared active and inactive absorption bands in CO2

There is no change in dipole moment during the symmetric stretch vibration and the 1340
cm-1 band is not observed in the infrared absorption spectrum (the symmetric stretch is
called infrared inactive). There is a change in dipole moment during the asymmetric
stretch and the 2350 cm-1 band does absorb infrared radiation (the asymmetric stretch in
infrared active). A related vibrational spectroscopic method is Raman spectroscopy,
which has a different mechanism and therefore provides complementary information to
infrared absorption.

Electromagnetic Spectrum
Visible Spectrum

Electromagnetic Spectrum
Type of
Radiation
gamma-rays

Frequency Range
(Hz)
1020-1024
17

20

x-rays

10 -10

ultraviolet

1015-1017
14

Wavelength
Range

Type of Transition

<10-12 m

Nuclear

1 nm-1 pm

inner electron

400 nm-1 nm

outer electron

750 nm-400 nm

outer electron

visible

4-7.5x10

near-infrared

1012-4x1014

2.5 um-750 nm

outer electron molecular


vibrations

infrared

1011-1012

25 um-2.5 um

molecular vibrations

microwaves

108-1012

1 mm-25 um

molecular rotations, electron


spin flips*

radio waves
100-108
>1 mm
*energy levels split by a magnetic field

nuclear spin flips*

Infrared Absorption Bands


Introduction
IR absorption spectroscopy uses mid-infrared light (2.5 - 50 m, 4000 - 200 cm-1) to
detect specific types of chemical bonds in a sample for identification of organic and
organometallic molecules.

Table of characteristic IR bands


Group Bond Appox. Energy (cm-1)
amines N-H 3300-3500
alkenes C-H 3020-3080
nitriles C=-N 2210-2260
amines C-N 1180-1360

hydroxyl O-H 3610-3640


aromatic rings C-H 3000-3100
alkanes C-H 2850-2960
carbonyl C=O 1650-1750

Nuclear Magnetic Resonance (NMR) Spectroscopy


Introduction
Nuclei with an odd number of protons, neutrons, or both, will have an instrinsic nuclear
spin.
Spin quantum number for various nuclei
Number of protons Number of Neutrons Spin Quantum Number

Examples

Even

Even

12

C, 16O, 32S

Odd

Even

1/2

"

"

3/2

11

Even

Odd

1/2

13

"

"

3/2

127

H, 19F, 31P

B,35Cl, 79Br
C
I

"

"

5/2

Odd

Odd

17

H, 14N

When a nucleus with a non-zero spin is placed in a magnetic field, the nuclear spin can
align in either the same direction or in the opposite direction as the field. These two
nuclear spin alignments have different energies and application of a magnetic field lifts
the degeneracy of the nuclear spins. A nucleus that has its spin aligned with the field will
have a lower energy than when it has its spin aligned in the opposite direction to the field.

Nuclear magnetic resonance (NMR) spectroscopy is the absorption of radiofrequency


radiation by a nucleus in a strong magnetic field. Absorption of the radiation causes
the nuclear spin to realign or flip in the higher-energy direction. After absorbing energy
the nuclei will reemit RF radiation and return to the lower-energy state.
The energy of a NMR transition depends on the magnetic-field strength and a
proportionality factor for each nucleus called the magnetogyric ratio. The local
environment around a given nucleus in a molecule will slightly perturb the local magnetic
field exerted on that nucleus and affect its exact transition energy. This dependence of the
transition energy on the position of a particular atom in a molecule makes NMR
spectroscopy extremely useful for determining the structure of molecules.

Instrumentation
There are two NMR spectrometer designs, continuous-wave (cw), and pulsed or Fouriertransform (FT-NMR). CW-NMR spectrometers have largely been replaced with pulsed
FT-NMR instruments. However due to the lower maintenance and operating cost of cw
instruments, they are still commonly used for routine 1H NMR spectroscopy at 60 MHz.
(Low-resolution cw instruments require only water-cooled electromagnets instead of the
liquid-He-cooled superconducting magnets found in higher-field FT-NMR

spectrometers.) These two spectrometer designs are described in separate CW-NMR and
FT-NMR documents.

Continuous-Wave Nuclear Magnetic Resonance (NMR)


Spectroscopy
Introduction
Continuous-wave NMR spectrometers have largely been replaced with pulsed FT-NMR
instruments. However due to the lower maintenance and operating cost of cw
instruments, they are still commonly used for routine 1H NMR spectroscopy at 60 MHz.
(Low-resolution cw instruments require only water-cooled electromagnets instead of the
liquid-He-cooled superconducting magnets found in higher-field FT-NMR
spectrometers.)

Instrumentation
A cw-NMR spectrometer consists of a control console, magnet, and two orthogonal coils
of wire that serve as antennas for radiofrequency (RF) radiation. One coil is attached to
an RF generator and serves as a transmitter. The other coil is the RF pick-up coil and is
attached to the detection electronics.
Since the two coils are orthogonal, the pick-up coil cannot directly recieve any radiation
from the generator coil. When a nucleus absorbs RF radiation, it can become reoriented
due to its normal movement in solution and re-emit the RF radiation is a direction that
can be recieved by the pick-up coil. This orthogonal coil arrangement greatly increases
the sensitivity of NMR spectroscopy, similar to optical fluorescence.
Spectra are obtained by scanning the magnet and recording the pick-up coil signal on
paper at the control console.

Fourier-Transform Nuclear Magnetic Resonance (FTNMR) Spectroscopy


Introduction
Fourier-transform NMR spectrometers use a pulse of radiofrequency (RF) radiation to
cause nuclei in a magnetic field to flip into the higher-energy alignment. Due to the
Heisenberg uncertainty principle, the frequency width of the RF pulse (typically 1-10 s)
is wide enough to simultaneously excite nuclei in all local environments. All of the nuclei
will re-emit RF radiation at their respective resonance frequencies, creating an
interference pattern in the resulting RF emission versus time, known as a free-induction
decay (FID). The frequencies are extracted from the FID by a Fourier transform of the
time-based data.
Instrumentation
An FT-NMR spectrometer consists of a control console, magnet, and a coil of wire that
serves as the antenna for transmitting and receiving the RF radiation. (Only one coil is
necessary because signal reception does not begin until after the end of the excitation
pulse.) Because the FID results from the emission due to nuclei in all environments, each
pulse contains an interference pattern from which the complete spectrum can be obtained.

Because of this multiplex (or Fellgett) advantage, repetitive signals can be summed and
averaged to greatly improve the signal-to-noise ratio of the resulting FID.
Raman Spectroscopy
Introduction
Raman spectroscopy is the measurement of the wavelength and intensity of inelastically
scattered light from molecules. The Raman scattered light occurs at wavelengths that are
shifted from the incident light by the energies of molecular vibrations. The mechanism of
Raman scattering is different from that of infrared absorption, and Raman and IR spectra
provide complementary information. Typical applications are in structure determination,
multicomponent qualitative analysis, and quantitative analysis.
Theory
The Raman scattering transition moment is:
R = < Xi | a | Xj >
where Xi and Xj are the initial and final states, respectively, and a is the polarizability of
the molecule:
a = ao + (r-re)(da/dr) + ... higher terms
where r is the distance between atoms and ao is the polarizability at the equilibrium bond
length, re. Polarizability can be defined as the ease of which an electron cloud can be
distorted by an external electric field. Since ao is a constant and < Xi | Xj > = 0, R
simplifies to:
R = < Xi | (r-re)(da/dr) | Xj >
The result is that there must be a change in polarizability during the vibration for that
vibration to inelastically scatter radiation.
Examples of Raman active and inactive vibrations in CO2

The polarizability depends on how tightly the electrons are bound to the nuclei. In the
symmetric stretch the strength of electron binding is different between the minimum and
maximum internuclear distances. Therefore the polarizability changes during the
vibration and this vibrational mode scatters Raman light (the vibration is Raman active).
In the asymmetric stretch the electrons are more easily polarized in the bond that expands
but are less easily polarized in the bond that compresses. There is no overall change in
polarizability and the asymmetric stretch is Raman inactive.
Raman line intensities are proportional to:
nu4 * sigma(nu) * I * exp(-Ei/kT) * C
where nu is the frequency of the incident radiation, sigma(nu) is the Raman cross section
(typically 10-29 cm2), I is the radiation intensity, exp(-Ei/kT) is the Boltzmann factor for
state i, and C is the analyte concentration.

Instrumentation
The most common light source in Raman spectroscopy is an Ar-ion laser. Resonance
Raman spectroscopy requires tunable radiation and sources are Ar-ion-laser-pumped cw
dye lasers, or high-repetition-rate excimer-laser-pumped pulsed dye lasers. Because
Raman scattering is a weak process, a key requirement to obtain Raman spectra is that the
spectrometer provide a high rejection of scattered laser light. New methods such as very
narrow rejection filters and Fourier-transform techniques are becoming more widespread.

Beer-Lambert Law
Introduction
The Beer-Lambert law (or Beer's law) is the linear relationship between absorbance and
concentration of an absorbing species. The general Beer-Lambert law is usually written
as:
A = a(lambda) * b * c
where A is the measured absorbance, a(lambda) is a wavelength-dependent absorptivity
coefficient, b is the path length, and c is the analyte concentration. When working in
concentration units of molarity, the Beer-Lambert law is written as:
A = epsilon * b * c
where epsilon is the wavelength-dependent molar absorptivity coefficient with units of
M-1 cm-1.

Instrumentation
Experimental measurements are usually made in terms of transmittance (T), which is
defined as:
T = I / Io
where I is the light intensity after it passes through the sample and Io is the initial light
intensity. The relation between A and T is:
A = -log T = - log (I / Io).
Absorption of light by a sample

Modern absorption instruments can usually display the data as either transmittance, %transmittance, or absorbance. An unknown concentration of an analyte can be determined
by measuring the amount of light that a sample absorbs and applying Beer's law. If the
absorptivity coefficient is not known, the unknown concentration can be determined
using a working curve of absorbance versus concentration derived from standards.

Derivation of the Beer-Lambert law

The Beer-Lambert law can be derived from an approximation for the absorption
coefficient for a molecule by approximating the molecule by an opaque disk whose crosssectional area, sigma, represents the effective area seen by a photon of frequency w. If the
frequency of the light is far from resonance, the area is approximately 0, and if w is close
to resonance the area is a maximum. Taking an infinitesimal slab, dz, of sample:

Io is the intensity entering the sample at z=0, Iz is the intensity entering the infinitesimal
slab at z, dI is the intensity absorbed in the slab, and I is the intensity of light leaving the
sample. Then, the total opaque area on the slab due to the absorbers is sigma * N * A *
dz. Then, the fraction of photons absorbed will be sigma * N * A * dz / A so,
dI / Iz = - sigma * N * dz
Integrating this equation from z = 0 to z = b gives:
ln(I) - ln(Io) = - sigma * N * b
or - ln(I / Io) = sigma * N * b.
Since N (molecules/cm3) * (1 mole / 6.023x1023 molecules) * 1000 cm3 /
liter = c (moles/liter)
and 2.303 * log(x) = ln(x)
then - log(I / Io) = sigma * (6.023x1020 / 2.303) * c * b
or - log(I / Io) = A = epsilon * b * c
where epsilon = sigma * (6.023x1020 / 2.303) = sigma * 2.61x1020
Typical cross-sections and molar absorptivities are:
sigma (cm2)
epsilon (M cm )
absorption - atoms
10-12
-16
molecules 10
infrared 10-19
Raman scattering
10-29
-1

-1

3x108
3x104
3x10
3x10-9

Limitations of the Beer-Lambert law


The linearity of the Beer-Lambert law is limited by chemical and instrumental factors.
Causes of nonlinearity include:
deviations in absorptivity coefficients at high concentrations (>0.01M) due to
electrostatic interactions between molecules in close proximity
scattering of light due to particulates in the sample

fluoresecence or phosphorescence of the sample


changes in refractive index at high analyte concentration
shifts in chemical equilibria as a function of concentration
non-monochromatic radiation, deviations can be minimized by using a relatively
flat part of the absorption spectrum such as the maximum of an absorption band
stray light

Quantitative Fluorimetry
Introduction
Light emission from atoms or molecules can be used to quantitate the amount of the
emitting substance in a sample. The relationship between fluorescence intensity and
analyte concentration is:
F = k * QE * Po * (1-10[-epsilon*b*c])
where F is the measured fluorescence intensity, k is a geometric instrumental factor, QE
is the quantum efficiency (photons emitted/photons absorbed), Po is the radiant power of
the excitation source, epsilon is the wavelength-dependent molar absorptivity coefficient,
b is the path length, and c is the analyte concentration (epsilon, b, and c are the same as
used in the Beer-Lambert law).
Expanding the above equation in a series and dropping higher terms
gives:
F = k * QE * Po * (2.303 * epsilon * b * c)
This relationship is valid at low concentrations (<10-5 M) and shows
that fluorescence intensity is linearly proportional to analyte
concentration.
Determining unknown concentrations from the amount of fluorescence
that a sample emits requires calibration of a fluorimeter with a
standard (to determine K and QE) or by using a working curve.

Limitations
Many of the limitations of the Beer-Lambert law also affect quantitative fluorimetry.
Fluorescence measurements are also susceptible to inner-filter effects. These effects
include excessive absorption of the excitation radiation (pre-filter effect) and selfabsorption of atomic resonance fluorescence (post-filter effect).

Specific fluorescence techniques

Atomic fluorescence spectroscopy (AFS)


Molecular laser-induced fluorescence (LIF)

Further Information

Science Hypermedia Home Page


Copyright 1996 by Brian M. Tissue
updated 2/23/96

Atomic-Fluorescence
Spectroscopy (AFS)
Introduction
Atomic fluorescence is the optical emission from gas-phase atoms that have been excited
to higher energy levels by absorption of electromagnetic radiation. The main advantage
of fluorescence detection compared to absorption measurements is the greater sensitivity
achievable because the fluorescence signal has a very low background. The resonant
excitation provides selective excitation of the analyte to avoid interferences. AFS is
useful to study the electronic structure of atoms and to make quantitative measurements.
Analytical applications include flames and plasmas diagnostics, and enhanced sensitivity
in atomic analysis. Because of the differences in the nature of the energy-level structure
between atoms and molecules, discussion of laser-induced fluorescence (LIF) from
molecules is found in a separate document.

Instrumentation
Analysis of solutions or solids requires that the analyte atoms be desolvated, vaporized,
and atomized at a relatively low temperature in a heat pipe, flame, or graphite furnace. A
hollow-cathode lamp or laser provides the resonant excitation to promote the atoms to
higher energy levels. The atomic fluorescence is dispersed and detected by
monochromators and photomultiplier tubes, similar to atomic-emission spectroscopy
instrumentation.

Further Information

Science Hypermedia Home Page


Copyright 1996 by Brian M. Tissue
updated 2/25/96

Atomic Transitions - Theory


Introduction
The probability that an atomic spectroscopic transition will occur is called the transition
probability or transition strength. This probability will determine the extent to which an
atom will absorb light at a resonance frequency, and the intensity of the emission lines
from an atomic excited state. The spectral width of a spectroscopic transition depends on
the widths of the initial and final states. The width of the ground state is essentially a
delta function and the width of an excited state depends on its lifetime.

Spectroscopic Transition
Strengths
Introduction
An atom or molecule can be stimulated by light to change from one energy state to
another. An atom or molecule in an excited energy state can also decay spontaneously to
a lower state. The probability of an atom or molecule changing states depends on the
nature of the initial and final state wavefunctions, how strongly light can interact with
them, and on the intensity of any incident light. This document discusses some of the
practical terms used to describe the probability of a transition occuring, which is
commonly called the transition strength. To a first approximation, transitions strengths
are governed by selection rules which determine whether a transition is allowed or
disallowed. Practical measurements of transitions strengths are usually described in terms
of the Einstein A and B coefficients or the oscillator strength (f).

Atomic-Absorption
Spectroscopy (AA)

Introduction
Atomic-absorption (AA) spectroscopy uses the absorption of light to measure the
concentration of gas-phase atoms. Since samples are usually liquids or solids, the analyte
atoms or ions must be vaporized in a flame or graphite furnace. The atoms absorb
ultraviolet or visible light and make transitions to higher electronic energy levels. The
analyte concentration is determined from the amount of absorption. Applying the BeerLambert law directly in AA spectroscopy is difficult due to variations in the atomization
efficiency from the sample matrix, and nonuniformity of concentration and path length of
analyte atoms (in graphite furnace AA). Concentration measurements are usually
determined from a working curve after calibrating the instrument with standards of
known concentration.
Schematic of an atomic-absorption experiment

Instrumentation
Light source
The light source is usually a hollow-cathode lamp of the element that is being measured.
Lasers are also used in research instruments. Since lasers are intense enough to excite
atoms to higher energy levels, they allow AA and atomic fluorescence measurements in a
single instrument. The disadvantage of these narrow-band light sources is that only one
element is measurable at a time.
Atomizer
AA spectroscopy requires that the analyte atoms be in the gas phase. Ions or atoms in a
sample must undergo desolvation and vaporization in a high-temperature source such as a
flame or graphite furnace. Flame AA can only analyze solutions, while graphite furnace
AA can accept solutions, slurries, or solid samples.
Flame AA uses a slot type burner to increase the path length, and
therefore to increase the total absorbance (see Beer-Lambert law).
Sample solutions are usually aspirated with the gas flow into a
nebulizing/mixing chamber to form small droplets before entering the
flame.

The graphite furnace has several advantages over a flame. It is a much


more efficient atomizer than a flame and it can directly accept very
small absolute quantities of sample. It also provides a reducing
environment for easily oxidized elements. Samples are placed directly
in the graphite furnace and the furnace is electrically heated in several
steps to dry the sample, ash organic matter, and vaporize the analyte
atoms.
Light separation and detection
AA spectrometers use monochromators and detectors for uv and visible light. The main
purpose of the monochromator is to isolate the absorption line from background light due
to interferences. Simple dedicated AA instruments often replace the monochromator with
a bandpass interference filter. Photomultiplier tubes are the most common detectors for
AA spectroscopy.
Picture of a flame atomic-absorption spectrometer:

Picture of a graphite-furnace atomic-absorption spectrometer:

Close-up of the graphite furnace | View of the control box

Further Information

Science Hypermedia Home Page


Copyright 1996 by Brian M. Tissue
updated 8/21/96

Atomic Emission Spectroscopy


(AES, OES)
Introduction
Atomic emission spectroscopy (AES or OES) uses quantitative measurement of the
optical emission from excited atoms to determine analyte concentration. Analyte atoms in
solution are aspirated into the excitation region where they are desolvated, vaporized, and
atomized by a flame, discharge, or plasma. These high-temperature atomization sources
provide sufficient energy to promote the atoms into high energy levels. The atoms decay
back to lower levels by emitting light. Since the transitions are between distinct atomic
energy levels, the emission lines in the spectra are narrow. The spectra of multi-elemental
samples can be very congested, and spectral separation of nearby atomic transitions
requires a high-resolution spectrometer. Since all atoms in a sample are excited

simultaneously, they can be detected simultaneously, and is the major advantage of AES
compared to atomic-absorption (AA) spectroscopy.
Schematic of an AES experiment

Instrumentation
As in AA spectroscopy, the sample must be converted to free atoms, usually in a hightemperature excitation source. Liquid samples are nebulized and carried into the
excitation source by a flowing gas. Solid samples can be introduced into the source by a
slurry or by laser ablation of the solid sample in a gas stream. Solids can also be directly
vaporized and excited by a spark between electrodes or by a laser pulse. The excitation
source must desolvate, atomize, and excite the analyte atoms. A variety of excitation
sources are described in separate documents:
Direct-current plasma (DCP)
Flame
Inductively-coupled plasma (ICP)
Laser-induced breakdown (LIBS)
Laser-induced plasma
Microwave-induced plasma (MIP)
Spark or arc
Since the atomic emission lines are very narrow, a high-resolution polychromator is
needed to selectively monitor each emission line. Picture of an inductively-coupled

plasma atomic emission spectrometer

Further Information

Science Hypermedia Home Page


Copyright 1996 by Brian M. Tissue
updated 3/7/96

X-ray Photoelectron
Spectroscopy (XPS, ESCA)
Introduction
X-ray photoelectron spectroscopy (XPS, also called electron spectroscopy for chemical
analysis, ESCA) is a electron spectroscopic method that uses x-rays to knock electrons
out of inner-shell orbitals. The kinetic energy (Ek) of these photoelectrons is determined
by the energy of the x-ray radiation h(nu) and the electron binding energy (Eb) as given
by:
EK = h(nu) - Eb
The electron binding energies are dependent on the chemical
environment of the atom. XPS is therefore useful to identify the
oxidation state and ligands of an atom.

Instrumentation
The detection of photoelectrons requires that the sample be placed in a high vacuum
chamber. Since the photoelectron energy depends on x-ray energy, the excitation source
must be monochromatic. The energy of the photoelectrons is analyzed by an electrostatic
analyzer and the photoelectrons are detected by an electron multiplier tube or a
multichannel detector such as a microchannel plate.

Further Information

Electron Spectroscopy
Introduction
Electron spectroscopies analyze the electrons that are ejected from a material for
qualitative or semi-quantitative analysis. In general an excitation source such as x-rays or
electrons will eject an electron from an inner-shell orbital of an atom. Detecting
photoelectrons that are ejected by x-rays is call x-ray photoelectron spectroscopy (XPS)
or electron spectroscopy for chemical analysis (ESCA). Detecting electrons that are
ejected from higher orbitals to conserve energy during electron transitions is called Auger
electron spectroscopy (AES). These electron processes are described below. Ejected
electrons can escape only from a depth of approximately 3 nm or less, making electron
spectroscopy most useful to study surfaces of solid materials. Depth profiling is
accomplished by combining an electron spectroscopy with a sputtering source that
removes surface layers.

Auger Electron Spectroscopy


(AES)
Introduction
Auger (pronounced ~o-jay) electron spectroscopy is an electron spectroscopic method
that uses a beam of electrons to knock electrons out of inner-shell orbitals. Auger
electrons are ejected to conserve energy when electrons in higher shells fill the vacancy
in the inner shell. These Auger electrons have energies characteristic of the emitting atom
due to the characteristic energy-level structure of that element.

Instrumentation
Picture of an Auger electron spectrometer

Further Information

Science Hypermedia Home Page


Copyright 1996 by Brian M. Tissue
updated 2/25/96

X-ray Fluorescence
Introduction
X-ray fluorescence is a spectroscopic method that is commonly used for solids in which
secondary x-ray emission is generated by excitation of a sample with x-rays. The x-rays
eject inner-shell electrons. Outer-shell electrons take their place and emit photons in the
process. The wavelength of the photons depends on the energy difference between the
outer-shell and inner-shell electron orbitals. The amount of x-ray fluorescence is very
sample dependent and quatitative analysis requires calibration with standards that are
similar to the sample matrix.

Instrumentation

Solid samples are usually powdered and pressed into a wafer or fused in a borate glass.

Microscopy
Introduction
Microscopy uses radiation and optics to obtain a magnified image of an object. The
resolution of the imaging is limited by the minimum focus of the radiation due to
diffraction. For light microscopy the diffraction limit is approximately 1 um (10-6 m) and
for high-resolution electron microscopy the limit is approximately 1 (10-10 m).

Specific Microscopy Techniques

Light microscopy
Scanning electron microscopy
Transmission electron microscopy

Diffraction
Introduction
Diffraction is a wave property of electromagnetic radiation that causes the radiation to
bend as it passes by an edge or through an aperture. Diffraction effects increase as the
physical dimension of the aperture approaches the wavelength of the radiation.
Diffraction of radiation results in interference that produces dark and bright rings, lines,
or spots, depending on the geometry of the object causing the diffraction. Common
interference effects for visible light are the rainbow pattern produced by an oil film on
wet pavement and the diffraction of light from a narrow slit or a diffraction grating.

Diffraction Methods
A certain wavelength of radiation will constructively interfere when partially reflected
between surfaces that produce a path difference equal to an integral number of
wavelengths. This condition is described by the Bragg law:
n( )=2dsin( )
where n is an integer, lambda is the wavelength of the radiation, d is the spacing between
surfaces, and theta is the angle between the radiation and the surfaces. This relation
demonstrates that interference effects are observable only when radiation interacts with

physical dimensions that are approximately the same size as the wavelength of the
radiation.
Interference of radiation between atomic planes in a crystal

These interference effects are useful for determining dimensions in


solid materials, and therefore crystal structures. Since the distances
between atoms or ions is on the order of 10-10 m (1 ), diffraction
methods require radiation in the x-ray region of the electromagnetic
spectrum, or beams of electrons or neutrons with a similar wavelength.
Electrons and neutrons are commonly thought of as particles, but they
have wave properties with the wavelength depending on the energy of
the particles as described by the de Broglie equation. The three
diffraction methods have different properties that are described in
more detail in separate documents. For example, the penetration
depths of the three types of beams are quite different:
neutrons > x-rays > electrons.
Schematic of crystal-structure determination by diffraction

Related Topics

electron diffraction

neutron diffraction
x-ray diffraction

Further Information

/chem-ed/diffract/diffract.htm, updated 10/14/96

Copyright 1996 by Brian M. Tissue, all rights reserved.


Science Hypermedia Home Page

Electron Diffraction
Introduction
Electron diffraction provides similar structural information as neutron diffraction.
Electron beams strongly interact with nuclei and electron diffraction is more useful than
x-ray diffraction for determining proton positions. The strong interaction of electrons
with matter results in a low penetration depth, and electron diffraction is usually used in a
reflection geometry to study surfaces or thin films. Electron beams are easy to
manipulate, detect, and focus to small spots to provide high spatial resolution.

Instrumentation
Electrons scatter from gases and electron diffraction must be performed under vacuum.

Neutron Diffraction
Introduction
Neutron diffraction provides similar structural information as electron diffraction.
Neutron beams interact more strongly with nuclei than do x-rays and neutron diffraction
is more useful than x-ray diffraction for determining proton positions. Neutrons interact
with a solid to a much lesser degree than x-rays and therefore have advantages in
studying materials that are damaged by x-rays and in cases where a large penetration

depth is desired. For the three types of diffraction methods, neutrons are unique in that
they have a magnetic moment and are therefore sensitive to magnetic ordering in a solid.

Instrumentation
Neutrons are produced by nuclear reactions in either a nuclear reactor or in an
accelerator. Reactor sources produce a continuous spectrum of neutron energies and
require a monochromator crystal to select a particular energy. Accelerator sources are
usually operated in a pulsed mode and neutron wavelength is selected by time-of-flight
methods, that is, data is taken at a fixed Bragg angle as a function of neutron energy

X-Ray Diffraction (XRD)


Introduction
The wavelengths of x-rays are of the same order of magnitude as the distances between
atoms or ions in a molecule or crystal (, 10-10 m). A crystal diffracts an x-ray beam
passing through it to produce beams at specific angles depending on the x-ray
wavelength, the crystal orientation, and the structure of the crystal. X-rays are
predominantly diffracted by electron density and analysis of the diffraction angles
produces an electron density map of the crystal. Since hydrogen atoms have very little
electron density, determining their positions requires extensive refinement of the
diffraction pattern. Electron diffraction and neutron diffraction are sensitive to nuclei and
are often used to accurately determine hydrogen positions.

Instrumentation
X-ray diffractometers consist of an x-ray generator, a goniometer and sample holder, and
an x-ray detector such as photographic film or a movable proportional counter. X-ray
tubes generate x-rays by bombarding a metal target with high-energy (10 - 100 keV)
electrons that knock out core electrons. An electron in an outer shell fills thehole in the
inner shell and emits an x-ray photon. Two common targets are Mo and Cu, which have
strong K(alpha) x-ray emission at 0.71073 and 1.5418 , respectively. X-rays can also be
generated by decelerating electrons in a target or a synchrotron ring. These sources
produce a continuous spectrum of x-rays and require a crystal monochromator to select a
single wavelength.

Picture of a single-crystal X-ray diffractometer

Related topics:

Powder X-ray Diffraction

Further Information

/chem-ed/diffract/xray.htm, updated 10/14/96

Copyright 1996 by Brian M. Tissue, all rights reserved.


Science Hypermedia Home Page

Vous aimerez peut-être aussi