Вы находитесь на странице: 1из 6

1676 Volume 55, Number 12, 2001 APPLIED SPECTROSCOPY

0003-7028 / 01 / 5512-1676$2.00 / 0
q 2001 Society for Applied Spectroscopy
Use of Near Edge X-ray Absorption Fine Structure
Spectromicroscopy to Characterize Multicomponent
Polymeric Systems
A. P. SMITH,* S. G. URQUHART, D. A. WINESETT, G. MITCHELL, and H. ADE
Departments of Physics (S.G.U., D.A.W., H.A.) and Materials Science and Engineeri ng (A.P.S.), North Carolina State University,
Raleigh, North Carolina 27695; and Dow Chemical USA, Midland, Michigan 48667 (G.M.)
The merits of a polymer characterization technique, near edge X-
ray absorption ne structure (NEXAFS) spectromicroscopy, are
demonstrated through the characterization of a multilayer polymer
lm with partially unknown chemical composition. The combina-
tion of chemical speciation through NEXAFS spectroscopy with the
high spatial resolution available in X-ray microscopy allows the
characterization of polymeric materials not possible with conven-
tional techniques. Analysis of a multilayer with layers as thin as 4
mm has yielded results that differ from those previously obtained
by infrared microscopy. Layers below the spatial resolution limit of
infrared microscopy were characterized.
Index Headings: Near edge X-ray absorption ne structure spec-
tromicroscopy; NEXAFS; Polymers; X-rays; Microscopy; Compo-
sition; Analysis.
INTRODUCTION
As polymers continue to replace more traditional ma-
terials, the need for polymeric materials with superior
physical properties continues to increase. In the attempt
to meet these stringent property requirements, multicom-
ponent polymer systems are increasingly utilized.
1
As
these systems become more complex, determination of
the structure-property-processing relationships becomes
more difcult. In most cases, a combination of chemical
functionality specicity and high spatial resolution are
required to determine the phase morphology of a system.
2
While infrared microscopy offers excellent chemical sen-
sitivity, its resolution in the mid-IR spectral range is dif-
fraction-limited to about 6 mm.
3,4
Conversely, electron
microscopy has excellent spatial resolution, but depends
on the use of heavy metal stains to provide indirect, qual-
itative chemical contrast in polymer systems.
5
Selective
stains are often not available for all chemical groups of
interest and can mask small chemical differences in
mixed systems. Here, an alternative means of composi-
tional characterization in multicomponent polymer sys-
tems that overcomes some of these limitations is de-
scribed. A multilayer polymer lm previously examined
with IR microscopy is reanalyzed with near edge X-ray
absorption ne structure (NEXAFS) spectromicrosco-
py.
6,7
The relative merits of the NEXAFS spectromicros-
copy technique are discussed.
Received 13 December 2000; accepted 2 August 2001.
* Present Address: Polymers Division, National Institute of Standards
& Technology, Stop 8542, Gaithersburg, MD 20899.
Present Address: Department of Chemistry, University of Saskatch-
ewan, 110 Science Place, Saskatoon, SK S7N 5C9 Canada.
Author to whom correspondence should be sent.
NEXAFS SPECTROSCOPY
Scanning Transmission X-ray Microscope. NEX-
AFS spectromicroscopy of polymers can be performed
either in transmission
6
or reection.
8,9
There are presently
two scanning transmission X-ray microscopes (STXM)
in North America.
10,11
The one utilized for this study is
located at the National Synchrotron Light Source at
Brookhaven National Laboratories (schematic shown in
Fig. 1). The synchrotron, in conjunction with an insertion
device called an undulator, provides a tunable, high-
brightness source of soft X-rays. The quasi-monochro-
matic output from the undulator can be adjusted by
changing the on-axis magnetic eld. A spherical grating
monochromator allows the selection of a particular X-ray
energy. The energy width is controlled by two beamline
slits and ranges between 100250 meV at the C K-edge
(;290 eV).
11
The radiation emerging from the beamline
is focused into a microprobe by a zone plate, a circular
diffraction grating. The microprobe size and brightness
depends on the zone plate parameters, which are closely
monitored during fabrication.
12
Zone plates utilized in the
Stony Brook STXM yield efciencies between 1014%
(near the theoretical limit) at the C K-edge with a micro-
probe size of 40 nm.
12
The spatial resolution achievable
is comparable to the probe size. As zone plate fabrication
technology improves, the microprobe will be reduced in
size. The far-eld wavelength limited resolution is ap-
proximately 2 nm at the C K-edge. The technological
challenges to approaching this limit are substantial. How-
ever, even at the present resolution, NEXAFS microscopy
has a spatial resolution about two orders of magnitude
better than conventional IR microscopy.
As with any diffraction grating, zone plates have a
number of diffraction orders with appreciable intensity.
In order to block all radiation except that in the rst-order
microprobe, a combination of a central stop on the zone
plate and an order selecting aperture (a pinhole slightly
smaller than the central stop) is utilized. The order se-
lecting aperture is placed in the geometric shadow of the
central stop as shown in Fig. 1 and blocks all radiation
except the rst-order microprobe. The samples, typically
100200 nm in thickness, are placed into the focus of
the microprobe and a high-ux gas proportional counter
(see Fig. 1) measures the transmitted radiation. Two
modes of operation are primarily used: spectral acquisi-
tion and image acquisition. To obtain an image, the X-
ray energy is held constant and the sample is raster-
scanned in the focus of the microprobe forming a two-
dimensional map of the X-ray absorbance. Conversely,
APPLIED SPECTROSCOPY 1677
FIG. 1. Schematic of the Stony Brook scanning transmission X-ray
microscope located at the National Synchrotron Light Source. The ma-
jor components of the beamline and microscope are discussed in the
text. (Figure courtesy of X-ray Microscopy group at SUNY at Stony
Brook.)
FIG. 2. NEXAFS reference spectra and chemical formulae of low den-
sity polyethylene (LDPE), poly(methyl methacrylate) (PMMA), poly-
styrene (PS), poly(bromo-styrene) (PBrS), and polycarbonate (PC).
These spectra illustrate the increasing spectral complexity with increas-
ing chemical complexity and show some features commonly found in
NEXAFS spectra of polymeric materials. See text for a discussion of
peak assignments and energies. (Spectra offset vertically to facilitate
comparison.)
X-ray absorption spectra are obtained one at a time by
holding the sample stationary and scanning the photon
energy across an absorption edge. In addition, an auto-
mated image acquisition program has recently been im-
plemented that obtains a sequence of images at a set of
preselected energies. Post-processing allows the extrac-
tion of absorption spectra from every point within the
sample.
13
For radiation sensitive samples,
14
however, care
must be exercised when using this data acquisition mode.
For further information on the microscope and its capa-
bilities and operating principles see the articles by Jacob-
sen et al.,
15
Zhang et al.,
16
and Feser et al.
11
Near Edge X-ray Absorption Fine Structure Spec-
troscopy. The chemical specicity in X-ray microscopy
is provided by near edge X-ray absorption ne structure
(NEXAFS) spectroscopy. X-rays are absorbed when an
inner shell (1s) electron is excited into an unoccupied
molecular orbital or the vacuum continuum.
17
The tran-
sition between initial and nal state occurs only when the
photon has the correct energy and when the transition is
electric-dipole allowed. The energy levels of the initial
and nal states in a material depend on the local chemical
environment of the excited atom. This chemical specic-
ity is best demonstrated with some examples. Figure 2
shows the carbon K-edge NEXAFS spectra and the
chemical structures of ve different polymers: low den-
sity polyethylene (LDPE), poly(methyl methacrylate)
(PMMA), polystyrene (PS), poly(bromo-styrene) (PBrS),
and polycarbonate (PC). The LDPE spectrum is the sim-
plest and shows two relatively broad peaks corresponding
to C 1s s* transitions corresponding to the CH bonds
(287.8 eV) and to the CC bonds (292 eV) in the chain
backbone. The spectral features are superimposed onto a
continuum step (around 288 eV in LDPE) that corre-
sponds to photoionization transitions into the vacuum
continuum. Compared to unsaturated materials, the NEX-
AFS spectra of saturated polymers are relatively simple,
although subtle differences between various saturated
polymers, such as high and low density polyethylene,
polypropylene, polyisobutylene, and ethylene-propylene
rubber have been observed.
7
Unsaturated carbon bonding results in more complex
NEXAFS spectra, exemplied by the spectra of PMMA
and PS in Fig. 2. In PMMA, a C 1s transition to the
p*
C5O
optical orbital of the carbonyl group (i.e., a C
1s(C5O) p*
C5O
transition) produces a relatively sharp,
intense peak at 288.4 eV. In PS, a C 1s transition to the
p*
C5C
optical orbital characteristic of the aromatic
moiety (i.e., a C 1s(CH) p*
C5C
transition) gives rise
to the sharp peak at 285.3 eV. Additional peaks in the PS
spectrum have been assigned as the aromatic C 1s 2p*
excitation (288.8 eV) and s* excitations corresponding
to the CH (287.7 eV) and the C5C (294 eV) bonds. In
addition to the presence or absence of saturation, the par-
ticular structure of the unsaturated group and the presence
of heteroatoms plays an important role in determining the
carbon K-edge NEXAFS of these materials. For example,
the characteristic C 1s(C5O) p*
C5O
transition of the
carbonyl group of PMMA (288.4 eV) is shifted to much
higher energy in PC (;290.4 eV) because of the induc-
tive effect of the two oxygens adjacent to the carbonyl
group. This is illustrated by the differences observed be-
tween the NEXAFS spectra of PS and PBrS. In PS, all
carbon atoms in the phenyl group are nearly equivalent
and contribute to the dominant C 1s(CH) p*
C5C
ex-
citation at 285.3 eV.
18
In the PBrS, a bromine atom is
substituted onto the para carbon of the phenyl ring, shift-
ing the C 1s binding energy of this substituted carbon
atom. This results in the introduction of a small peak at
286.4 eV (assigned as the C 1s(CR) p*
C5C
transition)
above the main C 1s(CH) p*
C5C
transition. This fea-
ture has been exploited to characterize the phase mor-
phology in blends of PS and PBrS.
19
The effects of sub-
1678 Volume 55, Number 12, 2001
TABLE I. Details of the polymer multilayer as determined by pre-
vious work.
26,27
Layer #
Thickness
(mm)
a
Material
b
1
2
3
4
5
6
7
8
9
12
8
66
5
12
4
37
37
6
Poly(ethylene terephthalate) (PET)
PET 1 TiO
2
Vinylidene chloride copolymer
Adhesive
PET
Adhesive
Polypropylene (PP)
PP 1 TiO
2
PP
a
As determined by optical microscopy.
b
As determined by IR microscopy.
FIG. 3. Montage of slices from low resolution NEXAFS micro-
graphs obtained from the polymer multilayer sample at the X-ray en-
ergies indicated. By varying the X-ray energy, the location of each of
the nine layers, as indicated by the top labels, can be determined. The
dark lines common to all micrographs are bars of the TEM grid on
which the thin section was mounted. The dark feature in layer 3 is
contamination on the sample. (Each slice is individually scaled for
contrast without loss of correlation to the spectra presented in subse-
quent gures.)
stituted atoms on both aromatic and carbonyl groups can
also be obser ved in the spectrum of PC where the peak
at 287.0 eV is due to the phenyl carbons next to the ether
linkages.
While the NEXAFS spectra presented in Fig. 2 provide
a good starting point for appreciating the possibilities in-
herent in this technique, NEXAFS spectroscopy can be
extended to more complex materials. NEXAFS spectral
energies and intensities are inuenced by conjugation be-
tween functional groups,
20
orientation of the mole-
cules,
21,22
differences in ortho, meta, and para substitu-
tions on aromatic groups,
23
and even vibrations.
18
For
NEXAFS spectra of additional polymers and further de-
tails on NEXAFS spectromicroscopy, see articles by Ade
et al.
7,24,25
POLYMER MULTILAYER
CHARACTERIZATION
As an example of the utility and present limitations of
NEXAFS spectromicroscopy for the characterization of
complex polymeric materials of unknown composition,
we analyzed a commercially available multilayer lm.
This lm is composed of 9 different polymer layers and
is used to make yogurt and other food containers. This
multilayer was previously analyzed with IR microsco-
py.
26,27
IR microscopy was able to positively identify sev-
eral of the layers and has made some assignments for all
layers. The IR results for the central layers have been
published, with the results based on a new methodology
to analyze even those layers with thicknesses comparable
or below the diffraction limit of the IR microscope.
26
Ta-
ble I gives details of the sample with the layer thickness
estimated by light microscopy and the chemical compo-
sition of each layer as determined by IR microscopy. For
the remainder of the manuscript, we will refer to the layer
by number as well as the IR microscopy assignment
placed between parenthesis, i.e., PET for layer 1. The
objective of the present study by X-ray microscopy was
to further elucidate the chemical composition of the thin-
ner layers within the sample and to provide a basic com-
parison between IR and NEXAFS microscopy.
Overview NEXAFS micrographs were acquired from
a microtomed thin section of the multilayer placed on a
standard hexagonal TEM grid. Chemically specic mi-
crographs were obtained without staining to determine
the location of each polymer layer. Subsequently, point
spectra were acquired from each layer after careful high
resolution imaging for placement of the X-ray beam. To
guard against radiation damage, the spectra were acquired
with the X-ray beam defocused. The defocused spot size
was maintained below 50% of the polymer layer width
in each case to minimize the possibility of spectral con-
tamination from adjacent layers.
Low resolution NEXAFS micrographs of the multilay-
er lm obtained at several photon energies are presented
in Fig. 3 and show each of the nine layers within the
sample as delineated at the top of the gure. In the image
acquired at 281.8 eV, the vinylidine chloride polymer
(layer 3) appears dark relative to the other layers due to
the high pre-carbon-edge cross-section from the Cl con-
tent. (Energies below the C edge primarily emphasize
elements other than carbon.) The micrographs obtained
at 284.5, 285.0, and 285.8 eV, energies associated with
the phenyl functional groups in the polymers, reveals the
PET layers (layers 1 and 5) as well as any other layers
containing unsaturated carboncarbon bonding (layers 2,
4, and 6) as dark bands. Slight differences in the X-ray
absorption of these layers allows for precise determina-
tion of the location of each of the rst six layers. The
nominal spatial resolution of the X-ray microscope of 50
nm is well below the smallest layer thickness of about 4
mm. As the photon energy is increased to 287.2 and 288.0
eV, the PP layers (layers 7, 8, and 9) begin to absorb
and appear dark as well. Differentiating between the in-
dividual PP layers is, however, difcult at these en-
ergies. At 292.0 eV, the PP layers are highly absorbing
and appear slightly darker than the other polymers. This
energy also shows the best differentiation between the
PP layers, although it is still difcult to discern in the
APPLIED SPECTROSCOPY 1679
FIG. 4. NEXAFS spectra from layers 1, 2, 4, 5, 6, and a PET reference
sample. Examination of the spectra reveals that layers 1 and 5 are in-
deed PET and layers 2, 4, and 6 have nearly the same chemical com-
position. (Spectra offset vertically to facilitate comparison.)
FIG. 5. NEXAFS spectrum of vinylidene chloride polymer layer
(layer 3). The large pre-edge absorbanc e is due to the high Cl content
within the polymer.
reproduction of the data with the image display limits
chosen. More pronounced differences can be observed in
the NEXAFS spectra presented in Fig. 6. The nal mi-
crographs are acquired at 302.4 and 316.0 eV, X-ray en-
ergies that are not very specic to any chemical func-
tional group, and emphasizes carbon/chlorine optical den-
sity/thickness maps of the sample.
NEXAFS absorption spectra acquired from each layer
within the sample, as well as from relevant reference ma-
terials, are displayed in Figs. 4, 5, and 6. Figure 4 shows
the spectrum acquired from a PET reference sample and
the spectra from layers referred to in the IR study as
PET (layer 1 and 5), PET 1 TiO
2
(layer 2), and a
layer labeled as adhesive (layer 4 and 6). Previous
NEXAFS studies of PET have assigned the peaks at
284.8 and 285.6 eV to C 1s(CH) p*
C5C
excitations
and the peaks at 288.2 and 290.1 eV to C 1s(C5O)
p*
C5O
transitions.
23
Two peaks are found in the PET spec-
trum for each of these moieties due to conjugation across
the terephthalic segment in the repeat unit resulting in a
splitting of the excited nal state.
23
Comparison of this
PET NEXAFS reference spectrum to the spectra acquired
from layers 1 and 5 positively identies these layers to
be PET, in agreement with previous IR data. Small dif-
ferences in the heights of the C5C peak at 284.8 and the
C5O peak at 288.2 eV in these spectra relative to the
reference spectrum are attributed to X-ray linear dichro-
ism.
17,21,22
This was conrmed by acquiring spectra of the
multilayer at different orientations relative to the polari-
zation of the X-ray beam. The dichroism is a natural con-
sequence of differences in the orientation of the polymers
which occurred during lm processing.
Examination of the NEXAFS spectra acquired from
layers 2, 4, and 6 in Fig. 4 shows a remarkable similarity
between the spectra, in contrast to the IR determination.
Layer 2 was identied by IR to be PET and TiO
2
, a non-
carbonaceous ller. While C 1s NEXAFS is not directly
sensitive to TiO
2
, the expected effect of the additive in
spatially averaged spectra would be a positive offset of
the PET absorbance spectrum due to the increased oxy-
gen content without any signicant change in the C 1s
NEXAFS spectrum of the polymer. This offset is ob-
served to a small degree in the pre-edge energy region
(,284.0 eV) of the spectrum from layer 2. However, the
spectral shape of layer 2 is signicantly different from
that of PET. This observed change in spectral shape in-
dicates that this layer contains additional polymers, and
PET is potentially only a minority component. IR was
able to characterize layers 4 and 6 only as adhesives due
to their small size and complicated surrounding structure.
The NEXAFS spectra clearly demonstrate that these lay-
ers have nearly identical chemical composition and, in
addition, closely match that of layer 2. A further discus-
sion of the possible composition of these layers will be
presented later in the text.
The NEXAFS spectrum obtained from layer 3, the vi-
nylidine chloride polymer, is shown in Fig. 5. As ex-
pected, this spectrum is relatively simple, with broad
peaks reminiscent of saturated polymers (Fig. 2). The ab-
sorbance is increased, particularly in the pre-carbon edge
energies, due to the high chlorine content of the polymer
itself. The peak energy is shifted to higher energies by
about 2 eV relative to the LDPE spectrum due to the Cl
electronegativity. Similar results have been obser ved in
the C 1s spectra of halogenated alkanes.
28
These changes
are similar to those observed in similar polymers such as
poly(vinyl chloride) and polytetrauoroethylene. How-
ever, a NEXAFS reference spectrum of poly(vinylidine
chloride) for denitive positive identication by NEX-
AFS microscopy is presently not available. The prior IR
assignment is consistent with the NEXAFS spectral in-
terpretation.
Figure 6 displays NEXAFS spectra from the PP lay-
ers (layers 79) as well as the NEXAFS spectrum from
a reference PP sample. Due to complete saturation, the
PP spectrum, resembles that of LDPE shown in Fig. 2,
with two relatively broad peaks attributed to the CH/
Rydberg and CC bonds in the polymer. The spectra ob-
tained from layers 7 and 8 match well with those of the
PP reference, indicating that they are essentially pure PP.
1680 Volume 55, Number 12, 2001
FIG. 6. NEXAFS spectra of a PP reference sample and the PP layers
(layers 79) of the multilayer lm offset vertically to facilitate com-
parison. These spectra support an assignment that layers 7 and 8 are
almost pure PP while layer 9 is primarily PP. The broadening and weak-
ening of the peaks in layer 9 is believed to be due to the presence of
additives or llers in this layer. Fillers and/or unsaturated additives are
also thought to be responsible for the small peaks near 285.0 eV.
FIG. 7. Residual spectral signal obtained by subtracting a fraction of
the PET reference NEXAFS spectrum from the NEXAFS spectra ob-
tained from layers 2, 4, and 6. These residual spectra are indicative of
other polymers present in these layers. Further details of the procedure
and possible system chemistry are provided in the text. (Spectra verti-
cally offset to facilitate comparison.)
The small peaks around 285.0 eV are probably due to an
additive or ller within these layers. No signicant
amounts of TiO
2
are detected in layer 8 since no signif-
icant pre-edge shift in the spectrum of this layer is found
relative to the spectrum of the reference material or layer
7, respectively. Previous studies on polymer blends with
TiO
2
llers have been able to image dispersed TiO
2
par-
ticles.
29,30
In the present study, no explicit attempt has
been made to image the TiO
2
ller. The small differences
in the spectra underscore why the differences between
layers 7 and 8 are difcult to detect in the micrographs
of Fig. 3. The spectrum obtained from layer 9, however,
does exhibit differences relative to the PP reference spec-
trum. The peak centers are found to be at the same en-
ergies, but the peaks are broader than the PP reference
material. These differences are believed to be due to the
addition of a saturated polymer, ller, or additive within
this layer. Signicant deviations in the layer composition
from these assignments are unlikely since no changes in
the NEXAFS peak energies are detected.
We return now to an elucidation of the chemistry of
layers 2, 4, and 6 via NEXAFS microspectroscopy. NEX-
AFS spectroscopy can be a quantitative analysis tool if
relevant model spectra are available. Generally, there are
no matrix effects, and a mixture of polymers can be suc-
cessfully simulated as a weighted linear sum of model
components. This approach breaks down if there are sig-
nicant electronic interactions between reference com-
pounds or model moieties. Spectral analysis has been uti-
lized, for example, to determine the composition of a
block copolymer
31
and to chemically characterize com-
plex polyurethane polymers.
32,33
A similar analysis meth-
odology can be applied to the spectra obtained from lay-
ers 2, 4, and 6 by subtracting a weighted PET reference
spectrum and analyzing the residual spectra. Figure 7
shows the residual spectra from layers 2, 4, and 6 that
result from this procedure. The weight of the subtracted
PET spectrum is noted. The quality of the subtraction was
judged by monitoring the low energy side of the 285 eV
feature and the characteristic PET peaks at 284.8 and
285.7 eV. PET spectral weight was subtracted until the
PET features disappeared without introducing a dip at the
leading edge of the 285 eV spectral feature in the residual
spectra. This procedure was successful since the low en-
ergy peak of PET at 284.8 eV is one of the lowest energy
features observed in any NEXAFS spectrum. As is ex-
pected from the raw spectra, these residual spectra retain
the same relative overall shape after the subtraction for
all three layers, although layer 6 has slightly higher ar-
omatic spectral intensity.
The two primary peaks of the residual spectra are lo-
cated approximately at 285.0 and 288.4 eV, which cor-
responds to excitations of aromatic or C5C groups and
carbonyl functional groups, respectively (see PS and
PMMA spectra in Fig. 2). While these functional groups
also exist in PET, the conjugation of these groups yields
a fundamentally different spectral signature as shown in
Fig. 4. The IR study also showed an excitation corre-
sponding to a carbonyl moiety in layer 6, but we were
unable to reproduce this for the other layers.
26
Aromatic
groups were not mentioned in the IR study. Based on
these results these layers are probably composed of a
mixture of PET, a vinyl acetate- or acrylate-containing
polymer generally used as an adhesive, and an additional
pol ymer or addit ive cont ai ning aromat ic functional
groups. The present lack of a comprehensive reference
database of NEXAFS spectra does not allow a positive
APPLIED SPECTROSCOPY 1681
identication of the composition of this layer. As an ac-
curate and comprehensive database of polymeric NEX-
AFS spectra becomes available, the degree of accuracy
of this compositional estimate can be substantially im-
proved and even quantied. This lack of reference data
does not distract from the fact that compositionally sen-
sitive spectra could be obtained with NEXAFS micros-
copy without interference from adjacent layers. Some
classes of polymers, such as unsaturated polymers, will
remain difcult to identify even with extensive NEXAFS
reference libraries without prior guidance or knowledge
about the composition. In these cases, correlation analysis
based on complementary characterization tools, such as
IR microscopy or even bulk compositional analysis from
nuclear magnetic resonance spectroscopy, might be nec-
essary.
In general, NEXAFS image sequences at many photon
energies can be acquired. These data sets can be used to
extract point spectra from all sample areas,
13
or they can
be analyzed with principle component analysis or target
factor analysis tools in complete analogy to the use of
these tools in electron microscopy.
3436
CONCLUSION
NEXAFS spectromicroscopy was used to elucidate
some aspects of a commercially available polymer mul-
tilayer lm with layers of unknown chemical composi-
tion. Chemical speciation combined with high-resolution
X-ray microscopy can be applied to polymer multilayers
and other materials in order to identify the composition
of layers below the spatial resolution of IR microscopy.
A layer that had previously been identied as PET 1
TiO
2
was shown to possess the same chemistry as two
other layers previously identied as adhesive layers. In
addition, all of these layers have been shown to contain
only minority fractions of PET with the majority of these
layers composed of other polymers. Although conclusive
identication of these other polymers is not currently pos-
sible, the polymers are shown to contain aromatic and
carbonyl functional groups. As the eld of polymer NEX-
AFS spectroscopy matures, positive identication of the
composition of heterogeneous materials will become pos-
sible. While the spectroscopic sensitivity of NEXAFS
spectroscopy is typically not as rich as that found in IR
spectroscopy, the combination of the chemical function
sensitivity with the higher resolution available in NEX-
AFS microscopy provides a powerful tool for the char-
acterization of polymeric materials.
ACKNOWLEDGMENTS
X-ray microscopy data were collected on the Stony Brook STXM
instrument developed by the group of J. Kirz and C. Jacobsen at SUNY
Stony Brook, with suppor t from the Ofce of Biological and Environ-
mental Research, U.S. DOE, under contract DE-FG02-89ER60858, and
the NSF under grant DBI-9605045. S. Spector and C. Jacobsen of
SUNY Stony Brook and D. Tennant of Lucent Technologies developed
the zone plates with support from the NSF under grant ECS-9510499.
We thank M. L. McKelvy for providing the sample and sharing IR data
with us. An NSF Young Investigator Award (DMR-9458060) supported
H.A. and A.P.S.
1. L. A. Utracki, Polymer Alloys and Blends (Carl Hanser Verlag,
Vienna, 1990).
2. L. H. Sperling, Polymeric Multicomponent Materials (John Wiley
and Sons, New York, 1997).
3. N. Guilhaumou, P. Dumas, G. L. Carr, and G. P. Williams, Appl.
Spectrosc. 52, 1029 (1998).
4. T. G. Rochow and P. A. Tucker, Introduction to Microscopy by
Means of Light, Electrons, X-rays, or Acoustics (Plenum Press,
New York, 1994), 2nd ed.
5. L. C. Sawyer and D. T. Grubb, Polymer Microscopy (Chapman and
Hall, London, 1996), 2nd ed.
6. H. Ade, X. Zhang, S. Cameron, C. Costello, J. Kirz, and S. Wil-
liams, Science 258, 972 (1992).
7. H. Ade and S. Urquhart, in Chemical Applications of Synchrotron
Radiation, T. K. Sham, Ed. (World Scientic Publishing, Singapore,
2001).
8. H. Ade, D. A. Winesett, A. P. Smith, S. Anders, T. Stammler, C.
Heske, D. Slep, M. H. Rafailovich, J. Sokolov, and J. Sto hr, Appl.
Phys. Lett. 73, 3773 (1998).
9. A. Cossy-Favre, J. Diaz, Y. Liu, H. Brown, M. G. Samant, J. Sto hr,
A. J. Hanna, S. Anders, and T. P. Russell, Macromolecules 31, 4957
(1998).
10. T. Warwick, H. Ade, S. Cerasari, et al., Rev. Sci. Instrum. 69, 2964
(1998).
11. M. Feser, M. Carlucci-Dayton, C. Jacobsen, J. Kirz, U. Neuhausler,
G. Smith, and B. Yu, in X-ray Microfocusing: Applications and
Techniques, I. McNulty, Ed. (Society of Photo-Optical Instrumen-
tation Engineers (SPIE), Bellingham, 1998), Vol. 3449, pp. 1929.
12. S. Spector, C. Jacobsen, and D. Tennant , J. Vac. Sci. Technol., B
15, 2872 (1997).
13. C. Jacobsen, S. Wirick, G. Flynn, and C. Zimba, J. Microscopy
197, 173 (2000).
14. T. S. Coffey, S. G. Urquhart, and H. Ade, J. Electron. Spectrosc.
Relat. Phenom., in press.
15. C. Jacobsen, S. Williams, E. Anderson, M. T. Brown, C. J. Buckley,
D. Kern, J. Kirz, M. Rivers, and X. Zhang, Opt. Commun. 86, 351
(1991).
16. X. Zhang, H. Ade, C. Jacobsen, J. Kirz, S. Lindaas, S. Williams,
and S. Wirick, Nucl. Instrum. Methods Phys. Res., Sect. A 347,
431 (1994).
17. J. Sto hr, NEXAFS Spectroscopy (Springer, Berlin, 1992).
18. S. G. Urquhart, H. Ade, M. Rafailovich, J. S. Sokolov, and Y.
Zhang, Chem. Phys. Lett. 322, 412 (2000).
19. H. Ade, D. A. Winesett, A. P. Smith, S. Anders, T. Stammler, C.
Heske, D. Slep, M. H. Rafailovich, J. Sokolov, and J. Sto hr, Appl.
Phys. Lett. 73, 3775 (1998).
20. E. G. Rightor, A. P. Hitchcock, H. Ade, R. D. Leapman, S. G.
Urquhart, A. P. Smith, G. Mitchell, D. Fisher, H. J. Shin, and T.
Warwick, J. Phys. Chem. B 101, 1950 (1997).
21. H. Ade and B. Hsiao, Science 262, 1427 (1993).
22. A. P. Smith and H. Ade, Appl. Phys. Lett. 69, 3833 (1996) .
23. S. G. Urquhart, A. P. Hitchcock, A. P. Smith, H. Ade, and E. G.
Rightor, J. Phys. Chem. B 101, 2267 (1997).
24. H. Ade, A. P. Smith, H. Zhang, G. R. Zhuang, J. Kirz, E. Rightor,
and A. Hitchcock, J. Electron Spectrosc. Relat. Phenom. 84, 53
(1997).
25. H. Ade, in Vacuum Ultraviolet Spectroscopy II, J. Samson and D.
Ederer, Eds. (Academic Press, San Diego, 1998), Vol. 32, pp. 225
262.
26. R. J. Pell, M. L. McKelvy, and M. A. Harthcock, Appl. Spectrosc.
47, 634 (1993).
27. M. L. McKelvy, private communication.
28. R. McLaren, S. A. Clark, I. Ishii, and A. P. Hitchcock, Phys. Rev.
A: At., Mol., Opt. Phys. 36, 1683 (1987).
29. H. Ade, A. P. Smith, G. R. Zhuang, B. Wood, I. Plotzker, E. Rightor,
D.-J. Liu, S.-C. Lui, and C. Sloop, Mater. Res. Soc. Symp. Proc.
437, 99 (1996).
30. C. Sloop, H. Ade, R. Fornes, R. Gilbert, and A. P. Smith, J. Polym.
Sci., Part B: Polym. Phys. 39, 531 (2001).
31. A. P. Smith, H. Ade, C. C. Koch, S. D. Smith, and R. J. Spontak,
Macromolecules 33, 1163 (2000).
32. S. G. Urquhart, A. P. Smith, H. W. Ade, A. P. Hitchcock, E. G.
Rightor, and W. Lidy, J. Phys. Chem. B 103, 4603 (1999).
33. S. G. Urquhart, A. P. Hitchcock, A. P. Smith, H. W. Ade, W. Lidy,
E. G. Rightor, and G. E. Mitchell, J. Electron. Spectrosc. Relat.
Phenom. 100, 119 (1999).
34. M. Prutton, Microsc. Microanal. Microstruc. 6, 289 (1995).
35. M. Prutton, D. K. Wilkinson, P. G. Kenny, and D. L. Mountain,
Appl. Surf. Sci. 144, 1 (1999).
36. I. M. Anderson, Inst. Phys. Conf. Ser. 165, 437 (2000).

Вам также может понравиться