0% found this document useful (0 votes)
33 views33 pages

Surface Characterization Techniques Overview

The document discusses various surface and material characterization techniques essential for adhesion bonding analysis, including methods like secondary ion mass spectroscopy, scanning electron microscopy, and infrared spectroscopy. It highlights the importance of understanding surface composition and structure for both product design and failure analysis. Additionally, it provides insights into the principles of each technique and their applications in analyzing material properties.

Uploaded by

Maira Bianchini
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
33 views33 pages

Surface Characterization Techniques Overview

The document discusses various surface and material characterization techniques essential for adhesion bonding analysis, including methods like secondary ion mass spectroscopy, scanning electron microscopy, and infrared spectroscopy. It highlights the importance of understanding surface composition and structure for both product design and failure analysis. Additionally, it provides insights into the principles of each technique and their applications in analyzing material properties.

Uploaded by

Maira Bianchini
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd

4 Surface and Material

Characterization
Techniques

4.1 Introduction
In the analysis of adhesion bonding, it is of great importance to have
knowledge of the composition and structure of the adherend’s surface.
Characterization of bonding surfaces can aid in both the design of a bond
and in the failure analysis if an adhesion bond should fail.
Surface analysis is the use of microscopic chemical and physical probes
that provide information about the surface region of a sample. The term
“sample” refers to any piece of material, structure, device, or substance
being studied. The probed region may be the uppermost layer of atoms (the
only true surface, for purists), or it may extend to several microns (mil-
lionths of a meter) beneath the sample surface, depending on the technique
used. The analysis is done to provide information about such characteristics
as the chemical composition, the level of trace impurities, or the physical
structure or appearance of the sampled region. Such information is of
importance to researchers or manufacturers who must understand the
materials in order to verify a theory or make a better product.[1]
Many of the techniques used to probe surfaces utilize a beam of ions
[e.g., secondary ion mass spectroscopy (SIMS)] to strike the surface and
knock off atoms of the sample material. These atoms are ionized, identified,
and measured using a technique known as mass spectrometry. Another
technique is ion scattering spectroscopy (ISS), which probes the outermost
layer of the surface. If a beam of ions is directed at a sample surface, then
a certain number will be elastically reflected. Measurement of the energy
of the backscattered particles can be used to identify the mass of these
44 BACKGROUND AND THEORY

atoms. The intensity of the scattered ions, as a function of angle of emission,


provides information regarding the surface crystallographic structure. The
variation in the intensity of the scattered beam is partly due to the shadowing
of substrate atoms by adsorbed atoms. By use of scattering theory, knowl-
edge of the sites that the surface atoms occupy can be derived.[2][3]
Other probes strike the surface with electrons [Auger spectrometry,
energy dispersive x-ray spectroscopy (EDS), or EDX] or x-rays [electron
spectroscopy for chemical analysis (ESCA) or total reflection x-ray fluo-
rescence (TXRF)], and measure the resulting electron or photon emissions
to probe the sample. Extremely high-energy helium nuclei bounce off a
sample and are measured as a sensitive indicator of surface layer composi-
tion and thickness [Rutherford backscattering spectrometry (RBS)].
Surface structure on a microscopic scale is observed by using scanning
electron microscopy (SEM), optical microscopy, and atomic force or
scanning probe microscopy (AFM/SPM). Note that all of these techniques
must be done in a high-vacuum environment. What is characterized is the
substrate under those conditions, not atmospheric conditions under which
most bonding processes are run.
Table 4.1 and Fig. 4.1 show a comparison of the sampling depth of
traditional methods and the new techniques. These analyses can focus on
a much shallower thickness of the surface and virtually yield analyses of the
outermost layers of a polymer article.
In addition to surface-specific methods, this chapter illustrates a
number of other analytic techniques that can be applied to characterize the
bulk properties of materials.

Figure 4.1 Sampling depth for surface analysis techniques.[1]


4: SURFACE AND MATERIAL CHARACTERIZATION TECHNIQUES 45

Table 4.1 Comparison of Average Sample Depth for Various Surface


Analysis Techniques[4]

Sampling
Analysis Method
Depth
Fourier transform infrared spectroscopy (FTIR),
Raman spectroscopy, conventional secondary ion <2 ìm
mass spectroscopy (conventional SIMS)
Energy dispersive x-ray (EDS or EDAX) <5,000 Å
Rutherford back scattering (RBS) <400 Å
Surface secondary ion mass spectroscopy (surface SIMS) <300 Å
Electron spectroscopy for chemical analysis (ESCA) also
<100 Å
called x-ray photoelectron microscopy (XPS)
Auger electron spectroscopy (AES) <100 Å
Ion scattering spectroscopy (ISS),
<2 Å
also called low-energy ion scattering (LEIS)
Time-of-flight secondary ion 1–3
mass spectroscopy (TOF-SIMS) monolayers
Atomic force microscopy (AFM) or
0.1 Å
scanning probe microscopy (SPM)

4.2 Infrared Spectroscopy


During analysis by infrared spectroscopy (IR), a sample is subjected to
electromagnetic radiation in the infrared region of the spectrum. The
wavelengths absorbed by the sample depend on the nature of the chemical
groups present. These wavelengths are defined by a wave number (cm-1)
obtained by dividing the number 10,000 by the wavelength in microns. The
ranges of various wave numbers are given in Table 4.2. Absorbance is
defined per Eq. (4.1) and varies from 0 (100% transmission) to infinity (0%
transmission). Absorbance is related to concentration of the absorbing
species by Beer’s law [Eq. (4.1)].

100
Eq. (4.1) Absorbance = Log
%Transmission
46 BACKGROUND AND THEORY

Table 4.2 Definition of Electromagnetic Wave Ranges

Definition of Wave Number Wave Length


Range (104cm-1) (ìm)
Ultraviolet 5–2.5 0.2–0.4
Visible 2.5–1.42 0.4–0.7
Near IR 1.42–0.4 0.7–2.5
IR 0.4–0.02 2.5–50
Far IR 0.02–0.0012 50–830

Eq. (4.2) Absorbance = Kcl

K = a constant that is occasionally called the extinction coefficient


c = concentration of the absorbing species
l = sample thickness

Equation (4.2) can be used to determine the concentration of a com-


pound in a solution when the value of K is known for that compound.
Chemical bonds such as C–O, O–H, etc., absorb different amounts of
infrared energy over various wavelengths. Absorption patterns vary from
sharp to broad for different bonds. Peak IR absorption wavelength (wave
number) is characteristic of a chemical bond. Absorption over the infrared
spectrum is a fingerprint characteristic of an organic material. Qualitative
identification can be achieved by obtaining and analyzing the IR spectrum
of a material.
Infrared spectra can also be obtained by reflecting the IR beam on the
surface of a sample. This technique is applied when it is not possible to
obtain an IR spectrum by a transmission technique. Attenuated total
reflectance (ATR), also known as ATIR (attenuated total internal reflec-
tance) is based on multiple internal reflectance of the IR beam on the sample
surface using a high refractive index crystal (e.g., thallium bromo-iodide).
The IR beam is entered into the crystal at an angle, and after about 25
internal reflections (5 cm), we may obtain a spectrum similar to that of the
transmission. The sample is in tight contact with both surfaces on the
crystal, as seen in Fig. 4.2.
Modern infrared spectrometers use Fourier transform calculations. The
method is called Fourier transform infrared spectroscopy (FTIR).
4: SURFACE AND MATERIAL CHARACTERIZATION TECHNIQUES 47

Figure 4.2 Schematic of the sample configuration relative to the reflective crystal
in attenuated total reflectance IR.

4.3 Raman Spectroscopy


When incident light strikes a sample, part of the light is scattered. Most
of the scattered light is of the same wavelength as the incident light; this is
called Rayleigh scattering. Some of the light is scattered at a different
wavelength; this is called Raman scattering. The energy difference between
the incident light and the Raman scattered light is called the Raman shift.
This difference is equal to the energy required to vibrate or rotate the
molecule. Several different Raman shifted signals can often be observed in
a single sample; each is associated with different vibrational or rotational
motions of molecules in the sample. The particular molecule and its
environment will determine what Raman signals will be observed. In
practice, because the Raman effect is so small, a laser is used as the source
of the incident light. A plot of Raman intensity versus the frequency of the
Raman shift is known as a Raman spectrum. It usually contains sharp bands
characteristic of the functional groups of the compounds or materials. This
information can be used to determine chemical structure and identify the
compounds present. It is complementary to FTIR because it uses a different
method to measure molecular vibrations.
Raman spectroscopy is a functional technique for qualitative analysis
as well as for discrimination of organic and/or inorganic compounds in
mixed materials. A Raman spectrum can be obtained from samples that are
as small as 1 µm[1] (see Fig. 4.3). The intensities of bands in a Raman
spectrum depend on the sensitivity of the specific vibrations to the Raman
effect and are proportional to the concentration. Thus, Raman spectra can
48 BACKGROUND AND THEORY

be used for semiquantitative and quantitative analysis. The technique is


used for identification of organic molecules, polymers, biomolecules, and
inorganic compounds both in bulk and as individual particles. Raman
spectroscopy is particularly useful in determining the structure of different
types of carbon (diamond, graphitic, diamond-like carbon, etc.) and their
relative concentrations.

Figure 4.3 Raman spectroscopy identified this micron-sized particle as poly-


ethylene.

4.4 Scanning Electron Microscopy (SEM)


Scanning electron microscopy (SEM) is a useful technique for the
analysis of plastic surfaces as well as any surface that survives in a vacuum.
Almost all SEMs start by sputtering the surface with a thin layer of gold
metal. This ensures that the surface is conductive, which is a requirement.
SEM involves a finely collimated beam of electrons that is focused into a
small probe that scans across the surface of a specimen. The interactions
between the beam and the material result in the emission of electrons and
photons as the electrons penetrate the surface. The emitted particles are
collected with the appropriate detector to yield information about the
surface. The final product of the electron beam collision with the surface
topology of the sample is an image of the surface (see Fig. 4.4).
4: SURFACE AND MATERIAL CHARACTERIZATION TECHNIQUES 49

Figure 4.4 shows the result of topical SEM analysis of a drug-infused


polymer coating on a stainless steel device. Scanning electron microscopy
of the surface at 750 times magnification is shown in Fig. 4.4(a) and at
35,000 times magnification in Fig. 4.4(b). The higher magnification re-
vealed the presence of crystals on the surface of the coating.
A helpful attachment to the SEM is the electron microprobe. An electron
beam is focused on a sample surface causing ionization to a depth of a few
micrometers. Energies and wavelength of the emitted x-ray during the de-
excitation cycle are characteristic of the elements present in the top layers
of the sample. The result is not a true surface analysis, although the electron
microprobe allows analysis of various spots of the sample surface.[5]

(a) (b)
Figure 4.4 SEM images of a polymer-coated device: (�) 750 times magnification,
and (�) 35,000 times magnification.[1]

4.5 Rutherford Backscattering Theory (RBS)


Rutherford backscattering (RBS) is based on collisions between atomic
nuclei, and derives its name from Lord Ernest Rutherford, who in 1911 was
the first to present the concept of nuclei in atoms. It involves measuring the
number and energy of those ions in a beam that backscatter after colliding
with atoms in the near-surface region of a sample. With this information, it
is possible to determine atomic mass and elemental concentrations versus
depth below the surface. RBS is ideally suited for determining the concen-
tration of trace elements heavier than the major constituents of the sub-
strate. It has, however, poor sensitivity for light masses, and for the makeup
of samples well below the surface.
50 � BACKGROUND AND THEORY

When a sample is bombarded with a beam of high-energy particles, the


vast majority of particles are implanted into the material and do not escape.
This is because the diameter of an atomic nucleus is approximately 1 × 10-15
meters, while the spacing between nuclei is closer to 2 × 10-10 meters. A
small fraction of the incident particles undergo a direct collision with a
nucleus of one of the atoms in the upper few micrometers of the sample. This
“collision” does not actually involve direct contact between the projectile
ion and target atom. Rather, energy exchange occurs because of Coulombic
forces between nuclei in close proximity to one another. However, the
interaction can be modeled accurately as an elastic collision using classical
physics.
The energy measured for a particle backscattering at a given angle
depends upon two processes. Particles lose energy while they pass through
the sample, both before and after a collision. The amount of energy lost is
dependent on that material’s stopping power. A particle will also lose
energy as a result of the collision. The ratio of the energy of the projectile
before and after collision is called the kinematic factor.
The numbers of backscattering events that occur from a given element
depend upon two factors: the concentration of the element and the effective
size of its nucleus. The probability that a material will cause a collision is
called its scattering cross section.

4.6 � Energy Dispersive X-Ray Spectroscopy


(EDS)
Energy dispersive x-ray spectroscopy (EDS) is a standard method for
identifying and quantifying elemental compositions in a very small sample
of material (even a few cubic micrometers). In a properly equipped SEM,
the atoms on the surface are excited by the electron beam, emitting specific
wavelengths of x-rays that are characteristic of the atomic structure of the
elements. An energy dispersive detector (a solid-state device that discrimi-
nates among x-ray energies) can analyze these x-ray emissions. Appropri-
ate elements are assigned, yielding the composition of the atoms on the
specimen surface. This procedure is called energy dispersive x-ray (EDS)
and is useful for analyzing the composition of the surface of a specimen
(Figs. 4.5 and 4.6).
4: SURFACE AND MATERIAL CHARACTERIZATION TECHNIQUES 51

Figure 4.5 EDX elemental analysis of a clean PTFE surface.�������������


�����������������������

Figure 4.6 EDX elemental analysis of a contamination spot on the PTFE


surface.�������������������������������������
52 BACKGROUND AND THEORY

4.7 Transmission Electron Microscopy (TEM)


Transmission electron microscopy (TEM) is similar to SEM with the
exception that the beam passes through the sample. A high voltage (80–200
keV) highly focused electron beam is passed through a thin, solid sample,
typically 100–200 nm in thickness. Electrons undergo coherent scattering
or diffraction from lattice planes in the crystalline phase of materials,
yielding phase identification. Characteristic x-rays that are generated can
be seen in a separate detector permitting qualitative elemental analysis.
Figure 4.7 shows a TEM micrograph of particles of polyvinylidene
fluoride produced by emulsion-polymerization in an aqueous phase. Figure
4.8 shows the capability of TEM in revealing the structure of a material, as
illustrated in the graphite sample. Lattice thickness and interplanar spacing
of graphite can be measured from the micrograph due to its excellent
resolution.

Figure 4.7 Transmission electron micrograph of particles of emulsion-polymerized


PVDF particles (Dp = 128 nm).[6]
4: SURFACE AND MATERIAL CHARACTERIZATION TECHNIQUES 53

Figure 4.8 Graphite is used as a control to demonstrate the resolution capabilities


of the Philips 420 TEM. The periodic structure shown above represents the 3.354
Å interplanar spacing of the graphite lattice. Each lattice plane has a thickness
equivalent to one carbon atom, which is 1.7 Å. The image has been averaged four
times (0.33 sec exposure), normalized, scaled, and low-pass filtered for subtrac-
tion of background noise (x 4,500,000).[7]

4.8 � Electron Spectroscopy for Chemical


Analysis (ESCA)
Electron spectroscopy for chemical analysis (ESCA) is a widely used
analytic technique for characterizing polymer surfaces. ESCA, also called
x-ray photoelectron spectroscopy (XPS), is able to detect all elements
except hydrogen. A sample is irradiated by x-ray beams, which interact with
the inner electron shell of atoms. Photon energy of the x-ray is transferred
to an electron in the inner shell, enabling it (the photoelectron) to escape
from the sample surface. An analyzer measures the kinetic energy of the
photoelectron, which is equal to the electron’s binding energy. Knowledge
of the binding energy allows identification of the element. The chemical
bond of an atom to other elements shifts the bonding energy of the
photoelectron to higher or lower values. This shift in binding energy
provides structural information about a molecule.
54 BACKGROUND AND THEORY

Operation and maintenance of ESCA equipment and interpretation of


its data are quite complex. Samples intended for ESCA and other surface
analysis should be handled carefully because minute contamination can
mask the surface structure of the samples. To alleviate this type of
complication, the sample surface can be washed with volatile solvents such
as methanol, acetone, hydrocarbons, and fluorocarbons using an ultrasound
bath. Typically, analysis is conducted before and after the surface wash in
studying a sample that has been handled and/or contaminated. Another
application of surface wash is removal of loose material that may be weakly
bonded to the surface. The interested reader is encouraged to refer to other
sources to gain an in-depth understanding of electron spectroscopy for
chemical analysis.[4][8]-[12]
A typical spectrum of ESCA shows peaks as a function of binding
energy, shown in Fig. 4.9 for polytetrafluoroethylene. C1s and F1s peaks on
a “clean” surface indicate that the PTFE surface is comprised of only carbon
and fluorine. The energy shift can be curve-fitted by trial and error to
determine the functional groups on the surface. The most simplified report

Figure 4.9 XPS spectra of polytetrafluoroethylene (�) before cleaning, and after
cleaning with an ultrasonic cleaner in (�) methanol, (�) acetone, and (�) n-heptane.[13]
4: SURFACE AND MATERIAL CHARACTERIZATION TECHNIQUES 55

that ESCA generates is a survey of the atomic composition of the surface


elements with the exception of hydrogen. A helpful tool to investigate
surface changes of a polymer is the ratio of other elements to carbon. Table
4.3 shows the effect of sodium etching on the surface composition of a few
different fluoropolymers.

Table 4.3 Effect of Sodium Etching on the Surface Composition and


Adhesion Bond Strength of Fluoropolymers[14]

Surface Chemical Analysis (%) by ESCA


Polymer Treatment F/C Cl/C O/C
Cl C F O
Ratio Ratio Ratio
PTFE – 1.60 - - - 38.4 61.6 -
®1
Tetra-Etch
PTFE 0.011 - 0.20 - 82.2 0.9 16.9
(1 minute)
PTFE N/1 min2 0.005 - 0.14 - 87.2 0.4 12.4
PVF – 0.42 - - - 70.4 29.6 -
®1
Tetra-Etch
PVF 0.21 - 0.026 - 80.7 17.2 2.1
(30 minute)
ECTFE – 0.64 0.27 - 14.1 52.5 33.4 -
®1
Tetra-Etch
ECTFE 0.16 0.05 0.12 3.8 74.9 12.2 9.1
(1 minute)
1
Supplied by WL Gore Corporation. 2Treatment with 1 mole solution naphthalenide in
tetrahydrofuran at room temperature.

4.9 Auger Electron Spectroscopy (AES)


Auger refers to the emission of a secondary electron after the surface
of a solid has been bombarded with electrons. This is a different phenom-
enon from photoelectron emission (the basis of ESCA) that occurs after
bombardment of a solid surface with low energy x-rays.[15] The energy of
the Auger electron depends on the chemical bonding state of the element
from which it was emitted. The depth of the escape of an Auger electron is
less than one nm, with metals having the shortest escape depth and organic
material the deepest. Lateral resolution of AES is around 1 µm. Auger
spectroscopy thus characterizes the surface of materials.
56 BACKGROUND AND THEORY

This method uses a low-energy electron gun, with a power less than 5
keV, to lessen the heating and decomposition of the surface. The number of
electrons and their resulting energies are detected by a counter and an
energy analyzer. The energy of the electron identifies the element, while the
number of emitted electrons indicates the surface concentration of the
element.

4.10 Ion Scattering Spectroscopy (ISS)


When a beam of ions hits a solid surface, projectiles will be scattered
back into the vacuum after one or more collisions with target atoms of the
top few layers.[16] Measurement of the energy of the backscattered particles
can be used to identify the mass of these atoms. The technique is called ion
scattering spectroscopy (ISS). The term actually encompasses several
techniques depending on the energy of the primary ion beam. Low-energy
ion scattering spectroscopy (LEIS) refers to primary energies in the range
of 100 eV to 10 keV, medium-energy ion scattering (MEIS) to a range from
100 to 200 keV, and high-energy ion scattering (HEIS) to energies between
one and several MeV. The LEIS technique is often called ion scattering
spectroscopy (ISS), the term we use below (meaning LEIS), while the HEIS
technique is best known as Rutherford backscattering spectroscopy (RBS).
Low-energy ion scattering is attractive as a surface-specific technique.
Spectra are usually obtained using noble gas ion beams from 0.5 to 3
keV.[17] Due to strong electron affinity of inert gas ions, the probability of
electron transfer is very high in the initial collision with a surface atom.
After two or more collisions, most ions will be neutralized, so a detector set
to analyze only ions of the same type as those in the incident beam will
almost entirely detect ions that have had only one collision with a target
atom. Projectiles entering the solid will be discarded since they would need
several scattering events to return back to the surface and exit.
The practical use of ISS is determined by its extreme sensitivity to only
the top surface layer (for standard experimental arrangements) or two
monolayers (for grazing incidence). Typical applications include composi-
tion of catalytic surfaces, thin film coatings, adhesion, and arrangement of
surface atoms including the localization of adsorbed atoms. Quantification
of surface analysis using low energy ions is hampered by the uncertainty of
4: SURFACE AND MATERIAL CHARACTERIZATION TECHNIQUES 57

the inelastic losses and the neutralization rate depending on ion trajectories.
In addition, overlapping peaks and multiple scattering must be taken into
account, thus computer simulation becomes an indispensable tool.

4.11 Secondary Ion Mass Spectroscopy (SIMS)


Secondary ion mass spectroscopy (SIMS) is a valuable technique for
identifying the structure and composition of polymer surfaces and comple-
ments ESCA. ESCA spectra for similar materials are difficult to resolve,
while SIMS can differentiate among several polymers. This is partly due to
the smaller sampling depth required by SIMS. In a typical analysis, the
surface of the polymer sample is bombarded by primary ions at low current
density, principally intended to minimize alteration of the sample surface
by irradiation. The polymer surface generates positive and negative ions
that are analyzed using a mass analyzer. The results of detailed analysis
provide chemical structure and composition data about the surface. A
traditional shortcoming of SIMS, however, is its inability to perform
quantitative analysis.
A different type of analyzer called time-of-flight (TOF-SIMS) can be
utilized to determine the structure of the sample as a function of sampling
depth. Time-of-flight is used to measure the mass of an ion, so it is a mass
analyzer. TOF-SIMS can be useful in determining stratification of resins in
coatings or diffusion of atoms into an area underneath a surface.[18][19]

4.12 Mass Spectroscopy or Spectrometry (MS)


Mass spectroscopy (MS) is applicable to both organic and inorganic
substances. Mass spectroscopy is a quantitative technique that allows the
study of the structure of organic matter. A sample is degraded into
fragments that are identified through fractional mass differences in atoms
of the same principal mass number.[20] There are two types of mass analysis.
One is high resolution which differentiates between fractional mass differ-
ences. The second is mass analysis that only differentiates principle mass
numbers. The latter is more common. This method requires extremely low
pressure (high vacuum) and the sample fragments must be volatile. Charged
58 BACKGROUND AND THEORY

fragments are ejected into the vacuum. Strictly speaking, this does not mean
these fragments are volatile (which implies evaporation). The determina-
tion relies on the pattern of fragmentation of a molecule upon ionization.
These patterns are distinct, can be reproduced, and are additive for mix-
tures. A mass spectrum is a graph that shows the measured values of ion
intensities per unit charge (m/e, mass divided by charge). The mass
spectrum is unique for each compound (Fig. 4.10). In a high-resolution
spectrum, the deviation of the molecular weight of each fragment from an
integral value is used to determine the elemental composition.

Figure 4.10 Matrix-assisted laser description ionized time-of-flight (MALDI-


TOF) mass spectrum of poly(methylmethacrylate)(PMMA).[23]

There are several methods for ionization of organic compounds includ-


ing particle bombardment, chemical and field ionization, electron impact,
field desorption, and laser pulse. For example, in laser micro-mass analysis
methods, a laser pulse is used to supply the necessary energy to volatilize
4: SURFACE AND MATERIAL CHARACTERIZATION TECHNIQUES 59

a sample from the surface for mass spectroscopy.[21] Mass spectrometry is


often used in conjunction with gas chromatography to identify the
separated components. The working technology of mass spectrometry is
quite complex.[22]

4.13 Gas Chromatography (GC)


Chromatography separates a mixture into its components, which are
then analyzed by one of many detectors. In gas chromatography (GC), the
sample is passed in vapor phase through an appropriated solid bed placed
in a column. An oven supplies heat to the column through which the vapor
is carried using a constant stream of a gas such as nitrogen or helium. The
time required to elute each component (retention time) is monitored.
Analysis times can run for more than an hour. GC refers to the use of solid
absorbent or molecular sieve columns. In the majority of GC columns, the
packing is coated with a stationary liquid phase. Separation takes place by
partitioning between the liquid coating and the carrier gas.
There are a variety of column packing materials and liquid phases. In
some cases, a very long capillary column is used without a solid packing and
only coated with a liquid on the wall. The oven can be operated isothermally
at a given temperature or according to a program where temperature varies
as a function of time.
There are several different types of detectors including thermal conduc-
tivity and electron capture models. The most common detector is the flame
ionization variety (FID). A hydrogen flame is utilized to combust the
column effluents. Thermal conductivity detectors do not degrade the
effluents and are not as sensitive as flame ionization. Electron capture
detection is especially sensitive to halogenated compounds.
Retention time of a particular compound is a characteristic of that
compound under a set of conditions. To resolve an overlap of retention time,
a combination of GC with MS or IR is employed. Figure 4.11 shows
examples of GC/MS spectra for two isomers. Note the similarity that is
typical of the GC/MS spectra of various isomers. Figure 4.12 shows the
chromatogram of contamination (silicone oil) that has resulted in poor
adhesion in a thermally bonded package.
60 BACKGROUND AND THEORY

Figure 4.11 GC/MS spectra of alkylated polycyclic aromatic hydrocarbons (PAH)


isomers.[24]

Figure 4.12 GC/MS thermal desorption chromatogram of silicon oil causing poor
adhesion of thermally bonded package;��������) octamethylcyclotetrasiloxane,
�������� decamethylcyclopentasiloxane.
4: SURFACE AND MATERIAL CHARACTERIZATION TECHNIQUES 61

4.14Nuclear Magnetic Resonance (NMR)


Nuclear magnetic resonance (NMR) is a unique technique that depends
on the magnetic nature of a few isotopic nuclei, primarily 1H, 13C, 19F, and
31P. In a strong magnetic field, a nucleus precesses about the direction of the

magnetic field with a frequency proportional to the field strength. A nucleus


is bombarded with radio frequency electromagnetic waves at right angles
to the magnetic field. When the spinning nucleus and the radio frequency
become equal, resonance takes place. The energy transfer is the basis of the
NMR spectrum of the nucleus and is measured and depicted as a band. The
resonance frequency at which the energy absorption occurs depends on the
chemical nature of the sample and the environment (e.g., solvent). The
nature of the bonds in a molecule can be identified by NMR (Fig. 4.13), the
use of reference spectra and characteristic frequencies for different types of
chemical bonds.

Figure 4.13 Example of NMR frequency shift in elucidating chemical structure of


a polymethyl methacrylate.
62 BACKGROUND AND THEORY

If a strong magnetic field is applied to the nuclei, a difference can be


found between alignment in the direction of the applied magnetic field and
in the opposite direction. In NMR, the constraint which leads to quantized
transitions is applied by the spectrometer. Due to this, the frequency of
absorption varies with the applied magnetic field; there is no absolute
frequency or wavelength for a given absorption. NMR spectra are not
plotted as absorption versus wave number as are IR spectra, but as
absorption versus chemical shift, d. The chemical shift for proton NMR is
the difference between the frequency of absorption of the sample and a
standard, tetramethylsilane (TMS), normalized by the frequency of absorp-
tion of TMS, which is expressed in parts per million (ppm). Therefore, the
above equation is multiplied by 106.[26]
In IR we consider two states for a bond: vibrating and non-vibrating.
The transition associated with the change from non-vibrating to vibrating
leads to the absorption at fixed wavenumbers. In NMR, several states are
potentially possible, depending on the spin quantum number. The permitted
states are given by the allowable values for the magnetic quantum number,
mI = I, I-1,...-I. For I = 1/2 there are two states possible, mI = 1/2 and mI =
-1/2. For I = 1 three states are possible, mI = 1, 0, and -1. For protons, the
two allowable states are generally spoken of as parallel and anti-parallel to
the applied field.
Polymethyl methacrylate (PMMA) was one of the first polymers
studied in depth for tacticity using proton NMR. For syndiotactic PMMA
(Fig. 4.13, bottom), 3 main absorptions are observed: α-CH3 at 0.91, β-CH2
at 1.9, and α-COOCH3 at 3.6. For the isotactic polymer (Fig. 4.13, top
curve), the α-CH3 is more deshielded at 1.20, the β-CH2 becomes a quartet
centered at 1.9, and the α-COOCH3 remains a singlet at 3.6. The splittings
of the β-CH2 in what should be a sequence of 1:1:1:1 is due to two types of
methylene groups termed erythro, e (more deshielded), and threo, t (less
deshielded), corresponding to the bottom and top protons in the molecular
sketch in Fig. 4.13. Each of these peaks are split into two peaks by the other,
leading to an expected splitting of 4 equal peaks, with a reported J coupling
constant of about 0.2 ppm. The e and t protons are separated by 0.7 ppm,
which can be verified by molecular modeling. A higher resolution NMR
resolves higher order stereosequences as shown below for isotactic and
atactic PMMA. You should compare the information content of the 60 MHz
spectrum above to the 500 MHz spectra below. Sixty (60) MHz refers to the
natural resonance frequency of a proton for a given magnetic field of the
instrument, (ν) a B0.
4: SURFACE AND MATERIAL CHARACTERIZATION TECHNIQUES 63

4.15 Differential Scanning Calorimetry (DSC)


Differential scanning calorimetry (DSC) is one of best known tech-
niques in the group known as the thermal analysis methods. Other tech-
niques include differential thermal analysis methods, dynamic mechanical
analysis methods, and thermogravimetric analysis methods; all of which
are discussed in the proceeding sections.
DSC is a thermal analysis technique used to measure the heat flows
related to transitions in materials as a function of time and temperature.[28]
These measurements provide qualitative and quantitative information
about physical and chemical changes that involve endothermic or exother-
mic processes or changes in heat capacity. Any event, such as loss of
solvent, phase transitions, crystallization temperature, melting point, and
degradation temperature of the plastic sample, results in a change in the
temperature of the sample. The available DSC systems have a wide range
of temperature capability, from -60°C to > 1,500°C.
Two types of systems are commonly used: power compensation and
heat flux DSCs. In the former apparatus, temperatures of the sample and the
reference are controlled independently by using separate but identical
furnaces. The power input to the two furnaces is adjusted to equalize the
temperatures. The energy required for the temperature equalization is a
measure of the enthalpy, or heat capacity, of the sample relative to the
reference. In heat flux DSC, the sample and the reference are intercon-
nected by a metal disk that acts as a low-resistance heat-flow path. The
entire assembly is placed inside a single furnace. The changes in the
enthalpy, or heat capacity, of the sample cause a difference in its tempera-
ture compared to the reference. The resulting heat flow is small due to the
thermal contact between the sample and the reference. Calibration experi-
ments are conducted to correlate enthalpy changes with the temperature
differences. In both cases, enthalpy changes are expressed in units of energy
per unit mass.
A typical DSC is run isothermally, with the temperature change at a
constant rate under an atmosphere of air or another gas. In the isothermal
case, the heat flow, or enthalpy change, is plotted against time. In the latter
case the heat flow, or enthalpy, is plotted against temperature or time.
Figure 4.14 shows an example of an enthalpy peak generated by an
exothermic or endothermic event such as melting of the crystalline phase
of a semicrystalline or crystalline polymer. Heat of fusion (Hf0) represents
the enthalpy change at the crystal melting point of T0. The weight fraction
64 BACKGROUND AND THEORY

of the crystalline phase (W) can be determined by comparing the measured


heat with Hf0. A sample of the polymer is heated in a DSC from a
temperature of T1 to T2 where the polymer becomes amorphous at a
temperature of T0 prior to reaching T2, shown by the baseline shift in Fig.
4.14. The enthalpy changes are determined according to the following
procedure.

Figure 4.14 Example of a DSC peak.

Ha represents the simple enthalpy change during the heating of the


polymer sample from T1 to T0. At T0 , the crystalline phase melts, becoming
amorphous. The enthalpy of fusion is expressed in Eq. (4.3). The increase
in the enthalpy of the amorphous phase as a result of heating from T0 to T2
is designated as Hc. If the enthalpy change calculated from the separation
of the DSC curve from the baseline (i.e., area under the curve) is designated
as HT , then Eq. (4.4) can be derived to calculate the weight fraction of the
crystalline phase.

Eq. (4.3) Hb = W ⋅ H f 0
HT = H a + H b + H c
HT = H a + W ⋅ H f 0 + H c

W=
(HT − Ha − Hc )
Eq. (4.4) Hf0
4: SURFACE AND MATERIAL CHARACTERIZATION TECHNIQUES 65

An example of a DSC thermogram for two types of polyetherether-


ketone is given in Fig. 4.15.

���������������������������������������������������������������������

���� �����������������������������������
DTA measures the temperature difference between a sample and a
reference as temperature is increased. A plot of the temperature difference
(thermogram) reveals exothermic and endothermic reactions that may
occur in the sample. Temperature for thermal events such as phase transi-
tions, melting points, crystallization temperatures, and others can be
determined. Maximum temperature capability of DTA is in excess of
1,000°C under air or other gas atmospheres. A typical heating rate for DTA
is in the range of 10°C–20°C/min, although slower rates are possible by
using a typical optimum sample weight of 50–100 mg. Prior to DTA, the
sample should be ground to finer particles than 100 mesh.
Melting of a semicrystalline or crystalline polymer manifests itself as
an endothermic peak. The peak temperature is correspondent to the actual
melting point of the polymer. As in DSC, the area under the peak is
proportional to the crystalline fraction of the sample. Mixtures of polymers
can be characterized by DTA because the melting points of individual
66 BACKGROUND AND THEORY

polymers are, for the most part, unaffected by the mixture. Similar poly-
mers, such as those of high and low density, are distinguishable by DTA,
while infrared spectroscopy would not be able to easily resolve such subtle
differences. Figure 4.16 shows the thermogram of a mixture of polytet-
rafluoroethylene (PTFE) and perfluoroalkoxy polymer (PFA). Three peaks
are evident. One is at about room temperature (19°C), which is a transition
point for PTFE, and two peaks reveal melting points. PTFE and PFA are
distinguished by the difference in their melting points.

Figure 4.16 Examples of DTA thermogram for PFA and PTFE.

4.17 Dynamic Mechanical Analysis (DMA)


Dynamic mechanical analysis (DMA) is generally a more sensitive
technique for detecting transitions than the DSC and DTA methods. This
is because the properties measured are the dynamic modulus and damping
coefficient. Both change significantly when crystalline structures transition to
the amorphous phase. The operating principle is that in these transitions, a
proportionally larger change takes place in the mechanical properties of a
polymer than in its specific heat. Therefore, dynamic mechanical analysis
is the preferred method of measurement for glass transition temperature and
other minor phase/structure changes of polymers.
4: SURFACE AND MATERIAL CHARACTERIZATION TECHNIQUES 67

The dynamic mechanical analysis method determines[30] elastic modu-


lus (or storage modulus, G´), viscous modulus (or loss modulus, G´´), and
damping coefficient (tan ∆) as a function of temperature, frequency, or
time. Results are usually in the form of a graphical plot of G´, G´´, and tan
∆ as a function of temperature or strain. DMA may also be used for quality
control and product development purposes.
The forced non-resonance technique is one of the simpler DMA
methods.[31] In the majority of commercially available DMAs, a force is
applied to a sample, and the amplitude and phase of the resultant displace-
ment are measured. These instruments utilize a linear actuator in which the
applied force is calculated from knowledge of the input signal to the
electromagnetic coils in the driver. An alternative is the use of a force
transducer to measure the applied load, with the sample held between this
transducer and the magnetic driver. These are the two types of arrangements
that are found with the forced non-resonance technique. In each case, the
sample is driven at a frequency below that of the test arrangement.
Typically, the frequency range of instruments is from 0.001 to 1000 Hz.
Any measurements below 0.01 Hz take too long for most analytical
experiments, especially if data are required as a function of temperature.
Resonance often occurs at frequencies greater than 100 Hz, depending upon
the sample stiffness.
Figure 4.17 shows an example of a DMA output for polytetrafluoroet-
hylene. Samples of PTFE were analyzed by DMA in shear mode at a length-
to-thickness ratio of 4:1. Figure 4.17(a) shows the DMA output of stress
versus time versus temperature, which has been converted to Fig. 4.17 (b)
by the time-temperature superimposition technique.
Time-temperature superimposition technique allows the prediction of
material properties that would normally require measurements over many
months or years. To collect the necessary data, measurement of a time-
dependent variable is made at different temperatures. The curves are shifted
mathematically along the time axis until some overlap occurs and a
continuous curve is formed covering several decades of time; this curve is
called themaster curve. The master curve can be used to determine the time-
dependent property as a function of time. Figure 4.17(c) shows total strain
as a function of time and temperature for PTFE.
The choice of geometry will be dependent on the sample under
investigation. For example, thin films can only be measured accurately in
tension. High quality dynamic mechanical testers perform well in tension
and should apply the required pre-tension forces automatically, including
68 BACKGROUND AND THEORY

Figure 4.17 Examples of DMA graphs for three grades of PTFE. (���������������
���������������)
4: SURFACE AND MATERIAL CHARACTERIZATION TECHNIQUES 69

those associated with large modulus changes that may occur at the glass
transition. Pre-tension is necessary in order to maintain the sample under
a net tension to prevent buckling that would otherwise occur. Tension
should be the first choice for any sample less than one millimeter thick.
Samples thicker than one millimeter will likely be too stiff for the
instrument in tension, making the bending mode preferable. Materials that
creep excessively, such as polyethylene, may be difficult to test in tension
due to creep under the pretension force.
Readers interested in a more in-depth understanding of data obtained
from DMA measurements can refer to polymer rheology textbooks. Some
available books published in this area include those listed in Refs. 32–34.

4.18 Thermogravimetric Analysis (TGA)


Thermogravimetric analysis (TGA) is a powerful technique for the
measurement of thermal stability of materials including polymers. In this
method, changes in the weight of a specimen are measured while its
temperature is increased. Moisture and volatile contents of a sample can be
measured by TGA. The apparatus consists of a highly sensitive scale to
measure weight changes and a programmable furnace to control the heat of
the sample. The balance is located above the furnace and is thermally
isolated from the heat. A high precision hang-down wire is suspended from
the balance down into the furnace. At the end of the hang-down wire is the
sample pan, the position of which must be reproducible. The balance must
be isolated from the thermal effects (e.g., by use of a thermostatic chamber)
to maximize the sensitivity, accuracy, and precision of weighing. Addition
of an infrared spectrometer to TGA allows analysis and identification of
gases generated by the degradation of the sample.
The TGA apparatus is equipped with a micro-furnace that can be
rapidly cooled. The heating element is made of platinum (reliable up to
1,000°C). An external furnace with a heating element made of an alloy of
platinum and 30% rhodium can extend the temperature range to 1,500°C.
A modern apparatus is usually equipped with a computer that calculates
the weight-loss fraction or percentage. A commercial TGA is capable of
>1,000°C, 0.1 µg balance sensitivity, and a variable controlled heat-up rate
under an atmosphere of air or another gas. The heat-up rate capability of
TGA can vary from 0.1°C–200°C/min.
70 BACKGROUND AND THEORY

Figures 4.18 and 4.19 show the TGA spectra for the FEP resins, DSCs
of which are shown in the Figs. 4.20 and 4.21. A comparison of these figures
indicates deterioration in the thermal stability of FEP after incorpora-
tion of pigment. Figure 4.22 represents the TGA thermogram for a PTFE
(31% wt) compounded with carbon black (18% wt) and silica (50.5% wt).
The 0.5% difference is due to the evolved volatile gases, which are not
shown in Fig. 4.22.

Figure 4.18 TGA thermogram of a neat FEP resin.� ���������� ������


����������������

Figure 4.19� TGA thermogram of a pigmented FEP resin.� ���������� ������


����������������
4: SURFACE AND MATERIAL CHARACTERIZATION TECHNIQUES 71

(a) First heat cycle �

(b) Cooling cycle �

(c) Second heat cycle �

Figure 4.20 DSC thermograms of a neat FEP resin (melt flow rate = 30 g/10 min).
(�������������������������������)
72 BACKGROUND AND THEORY

(a) First heat cycle �

(b) Cooling cycle �

(c) Second heat cycle �

Figure 4.21 DSC thermograms of a pigmented FEP resin (melt flow rate = 30 g/
10 min). (�������������������������������)
4: SURFACE AND MATERIAL CHARACTERIZATION TECHNIQUES 73

Figure 4.22 TGA thermograms of a pigmented PTFE resin.[35] (���������������


���������������)

REFERENCES
1. �C. Evans and Associates, 810 Kifer Road, Sunnyvale, CA 94086
www.surface-science.com/tutorial (2004)
2. � UK Surface Analysis Forum, www.uksaf.org/tech/iss (1998)
3. �University of Alberta, Alberta Center for Surface Eng. and Sci,
www.ualberta.ca/ACSES/ACSES2/ISS.htm (2004)
4. � Chan, C. M., Polymer Surface Modification and Characterization, Hanser
Publishers, Munich (1994)
5. � Stewart, I. M., Microstructural Analysis: Tools and Techniques (J. L.
McCall and W. Mueller, eds.) pp. 281-285, Plenum Press, New York
(1973)
6. � Tsuda, N., Fluoropolymer Emulsion for High-Performance Coatings, Daikin
Industries Ltd., Osaka, Japan (05/31/2001)
7. � www.botany.utexas.edu/facstaff/facpage
8. � Siegbahn, K., et al., ESCA: Atomic, Molecular, and Solid State Structure
Studied by Means of Electron Spectroscopy, Acta Regiae Sco. Sci.
Upsaliensis, Ser. IV, Vol. 20, Almqvist and Wiksells, Stockholm (1967)
74 BACKGROUND AND THEORY

9. � Siegbahn, K., et al., ESCA Applied to Free Molecules, North-Holland,


Amsterdam (1969)
10. � Wheeler, D. R., and Pepper, S. V., J. Vac. Sci. Technol., 20:226 (1982)
11. � Briggs, D., Surface Analysis and Pretreatment of Plastics and Metals (D.
M. Brewis, ed.) Macmillan, New York, p. 73 (1982)
12. � Briggs, D., Ency. Polymer Sci., John Wiley and Sons, 16:399 (1988)
13. � Chan, C. M., Polymer Surface Modification and Characterization, Hanser
Publishers, Munich, Ref. 75 from Ch. 1 (1994)
14. � Brewis, D. M., Surface Modification of Fluoropolymers for Adhesion,
Paper presented at the Fluoropolymers Conference, Loughborough
University, UK (1992)
15. �Davis, L. E., MacDonald, N. C., Palmberg, P. W., Riach, G. E., and
Weber, R. E., Handbook of Auger Electron Spectroscopy, 2nd Ed., Physical
Electronics Industries, Eden Prairie, MN (1976)
16. �Univ. of Alberta, Center for Surface Engineering and Science,
www.ualberta.ca/ACSES/ ACSES3/ Techniques/iss.htm
17. � Taglauer, E., and Heiland, W., Applied Physics, 9:261 (1976)
18. � Brown, A., and Vickerman, J. C., Surface Interface Anal., 6:1 (1984)
19. � Briggs, D., and Wotton, A. B., Surface Interface Anal., 4:109 (1982)
20. � McGraw-Hill Encyclopedia of Chemistry, 2nd Edition, (Sybil P. Parker,
Editor-in-Chief), McGraw-Hill, Inc., New York (1992)
21. � Ezrin, M., Plastics Failure Guide: Cause and Prevention, Hanser Publishers,
New York (1996)
22. � Marshall, A. G., and Verdun, F. R., Fourier Transforms in NMR, Optical,
and Mass Spectrometry, Elsevier, New York (1989)
23. � www.public.asu.edu/~ionize/polymer.html
24. � GC/MS/MS Analysis of Alkylated Polycyclic Aromatic Hydrocarbons,
VARIAN Application Notes, GCMS44:0695, www.varianinc.com/image/
vimage/docs/ products/chrom/apps/
25. � Ezrin, M., and Lavigne, G., Gas Chromatography/Mass Spectroscopy for
Plastics Failure Analysis, Institute of Material Science, University of
Connecticut, Storrs, CT, Presented at ANTEC, Chicago
26. � Campbell, D., White, J. R., and Pethrick, R. A., Polymer Characterization,
Physical Techniques, 2nd edition, International Thomson Publishing
Services Ltd. (1999)
4: SURFACE AND MATERIAL CHARACTERIZATION TECHNIQUES 75

27. � Bovey, F. A., High Resolution NMR of Macromolecules, Academic Press,


New York (1972)
28. � www.mri.psu.edu
29. � Elmer, P., Measurement and Calculation of Kinetic Parameters for the
Crystallization of Poly(ether ether ketone) (PEEK), Thermal Analysis
Newsletter, www.thermal-instruments.com/Applications/petan066.pdf
30. � Plastics Technology Laboratories, Inc., 50 Pearl Street, Pittsfield, MA
01201, www.PTLI.com
31. � www.Triton-Technology.Co.UK
32. � Bird, R. B., Stewart, W. E., and Lightfoot, E. N., Transport Phenomena,
2nd ed., Wiley, New York (2001)
33. � Dealy, J. M., and Wissbrun, K. F., Melt Rheology and its Role in Plastics
Processing: Theory and Applications, Springer (1990)
34. � Cogswell, F. N., Polymer Melt Rheology: A Guide for Industrial Practice,
Woodhead Publishing (1994)
35. � Krause, A., Lange, A., and Ezrin, M., Plastics Analysis Guide: Chemical
and Instrumental Methods, translated by K. Ruby, Hanser Publishers,
Munich (1979)

You might also like