Analysis of Polymer Surfaces Using Electron and ... - ACS Publications

Analysis. 0. „ of Polymer. Surfaces. J. Joseph A. Gardella, Jr.1. Department of Chemistry and Industry. University ..... First, they are ill-defined...
0 downloads 0 Views 13MB Size
'"«nu""««ι

Analysis of Polymer Surfaces UsingElectronand Ion Beams

Joseph A. Gardella, Jr. 1 Department of Chemistry and Industry University Center for Biosurfaces University at Buffalo, SUNY Buffalo, NY 14214

Jean-Jacques Pireaux Facultés Universitaire Notre Dame de la Paix Laboratoire Interdisciplinaire de Spectroscopie Electronique Rue de Bruxelles 61 B-5000 Namur Belgium

When materials science emerged as one of the important research areas in the 1980s, chemists were encouraged to become more involved in interdisciplinary and multidisciplinary efforts in the field (1). Today, interest continues unabated and chemists are making contributions in areas ranging from material synthesis and characterization to the details of development and device engineering. One emphasis in materials science involves surface and interfacial concerns, and this is an area in which advances in basic science rapidly transfer to technology (2). Materials scientists and engineers are now focusing more attention on polymers because of their increasing usefulness in the marketplace. Polymers are an inexpensive alternative for traditional structural materials: in fact, the volume of polymers produced worldwide has surpassed that of steel. High-technology polymers tailored for ' Current address: Chemistry Division, National Science Foundation, 1800 G St., N.W., Washington, DC 20550

0003-2700/90/0362-645A/$02.50/0 © 1990 American Chemical Society

specific applications are continuously being developed, although the production volume is relatively small. Applications are often based on particular properties (e.g., special dielectric, mechanical, chemical, thermal, or interfacial properties) that can be achieved through the synthesis of the polymer itself. Many of the specific applications of polymeric materials with interfacial properties are well known (e.g., wetting, grafting, or adhesion enhancement). A major challenge in polymer interfacial science is understanding the relationship between bulk structure and composition and the resulting surface structure and composition. X-ray photoelectron spectroscopy (XPS), also

need more precise information about surface structure than that available from ESCA analyses. For example, the presence and concentration of surfaceactive functional groups can routinely be detected with ESCA, especially in concert with IR or Raman vibrational spectroscopy. For biocompatibility or chromatography applications, however, information about the orientation and reactivity of these functional groups would be helpful. In addition, information about the monomer arrangement along a chain in copolymers and intrachain interactions at the surface would also be important data not routinely available. In this article, we will focus on the experimental aspects of three tech-

INSTRUMENTATION called electron spectroscopy for chemical analysis (ESCA), has been t h e workhorse ultra high vacuum (UHV) method for polymer applications. It provides chemical bonding information, exhibits high surface sensitivity, results in minor damage to the sample, and is relatively insensitive to the insulating properties of the sample. ESCA applications for polymer surfaces have involved quantitative analysis of surface species and detection of impurities or modifications at the surface, as described previously in this JOURNAL ( 3 5). A recent review describes the many types of information available from ESCA of polymers (6). To meet today's polymer engineering requirements, however, scientists

niques that can provide information complementary to that provided by ESCA: high-resolution electron energy loss spectroscopy (HREELS), lowdamage (macro)molecular secondary ion mass spectrometry (SIMS), and low-energy ion scattering spectrometry (LEIS or ISS) for polymer surface science applications. Selected examples will illustrate the limitations and interrelationships of these methods with other available ones. Table I summarizes the analytical characteristics of useful methods for polymer surface science applications. Instrumentation An UHV environment is required for the spectroscopies described here be-

ANALYTICAL CHEMISTRY, VOL. 62, NO. 11, JUNE 1, 1990 · 645 A

INSTRUMENTATION cause of the electron or ion beam used to probe the surface and the need to avoid contamination of the surface. An ultra high vacuum is necessary; typical­ ly, 10~ 7 -10~ 10 mbar pressure must be maintained.

HREELS. In a HREELS experi­ ment (Figure 1), a low-energy, mono­ chromatic electron beam (typically, 110 eV with an energy defined to ~ 5 meV) is focused onto the sample sur­ face. There the beam is backscattered

Monochromator

Analyzer

Electron multiplier

Electron gun

Figure 1. Schematic of a high-resolution electron energy loss experiment. In an ultra high vacuum environment, a low-energy electron beam emitted from a hot cathode is spatially dispersed (monochromator) to select a very narrow slice of its energetic distribution. This beam is then accelerated toward the target with the desired interaction potential. The backscattered beam is then energy analyzed (analyzer), and a loss spectrum is produced.

Table I.

into an electron energy analyzer whose function is to separate the electrons that suffered inelastic collisions in the target. Low-energy electrons can excite vibrations of molecular functional groups present on the (polymer) sur­ face. In principle, the energy loss corre­ sponding to vibrational excitations are easily detected between 25 and 600 meV or more. Thus the technique has a large dynamic range of 200-5000 c m - 1 (1 meV = 8.066 cm" 1 ) that is available at once from a single scan in a single experiment. The electron monochromator and analyzer are based on a 180° hemi­ spherical or 124° cylindrical sector or cylindrical mirror design. The electron beam produced by a hot filament is transported, focused, and (de)accelerated with standard electron optics and, after being backscattered from the sample surface, detected with a channeltron multiplier. The signal amplifi­ cation is processed by standard elec­ tronic modules that typically are com­ puter interfaced. Many H R E E L S systems are built by individual re­ searchers, but several commercial sur­ face science manufacturers such as Leybold Heraeus, McAllister, Vacuum Generators, and LK Technologies pro­ vide system components and build spectrometers. SIMS. SIMS can be considered a family of methods, and the different

Characteristics of some analytical techniques suitable for studying polymeric materials

Characteristic Vacuum Resolution

ESCA (XPS) High vacuum 0.6 eV

Elemental and molec­ No H detection ular information Chemical shifts

Detection limit

Percent of monolayer

Lateral resolution

5 Mm

Depth sensitivity Sample damage

>50Â Small with unmonochromatized X-rays Elemental and chemical analysis Electronic structure

Information provided

IR and Raman

HREELS

UHV 10meV (= 80 cm - 1 ) ? ? Functional group iden­ Functional group tification identification

SIMS

ISS

Atmospheric pressure ~ 1 cm-1

UHV 0.1-1 amu

Percent in volume 0.01 monolayer

All elements and isotopes MW of molecules from fragmenta­ tion Elemental: ppm/ppb Percent of monolayer Molecular: 0.01 monolayer 10 nm (atomic) — ? molecular (ions) 10 Â? 3-5 Â High High

μπι range None

0.001 monolayer

1000 ftm 20 Â None

Molecular vibrations Molecular vibrations Bulk functional groups Surface functional groups

646 A · ANALYTICAL CHEMISTRY, VOL. 62, NO. 11, JUNE 1, 1990

Low detection limits Structural information from fragmentation of molecular ions Bonding information from molecular ions

UHV Variable (depends on primary ion) No Η or He Resolution limited

Elemental and atomic orientation

INSTRUMENTATION components depend on the type of information desired. The basic block diagram (Figure 2) of an ion/atom beam based SIMS experiment involves the generation of a keV ion or atom beam focused on the sample surface as well as the collection and mass analysis of secondary ions generated from desorption or sputtering. Two sets of configurations commonly used involve a combination of two types of primary ion or atom beam sources and two types of mass analyzers. In the first type, the ion or atom source uses a noble-gas ion beam (subsequently neutralized if it is an atom source) with a low current density that is achieved via defocusing. The second type of source uses a focused ion beam that is produced from liquid metals (Ga) or desorption from a solid-state source (Cs). In this type of ion beam, the ion current is low although the current density is quite high. After secondary ions are generated from the sample, they are collected by ion optics and focused into a mass spectrometer for analysis. Every type of MS detector has been used for SIMS ex-

periments; for polymers, the two common approaches have been quadrupole and time-of-flight (TOF) detectors. ISS. Basic instrumentation for lowenergy ion scattering (Figure 2) involves a charged-particle energy analyzer and low-current ion source— components common to ESCA/ H R E E L S and SIMS, respectively. Only a change in polarity of the energy analyzer is needed to perform rudimentary ISS experiments. Typical commercial instrumentation exists as an afterthought to ESCA or Auger spectrometers. However, component geometries and alignments are usually optimized for electron spectrometry, not ISS. ISS involves the two-body conservation of momentum and energy in a collision between a noble-gas ion beam and the sample surface and the subsequent energy analysis of the loss that results from the "billiard-ball" collision. The ion beam is typically at a few hundreds to thousands of electron volts and at a low current density. This experiment requires the same type of noble-gas ion sources discussed above for

Figure 2. Schematic of ion beam-polymer interactions leading to SIMS and ISS. Insets show the form of spectra obtained from polymers. 648 A · ANALYTICAL CHEMISTRY, VOL. 62, NO. 11, JUNE 1, 1990

static SIMS and sector or mirror-type energy analyzers commonly used for ESCA or Auger spectroscopy. However, the cylindrical mirror analyzer (CMA) is preferable because of the limited signal levels at low-current densities and the low-energy resolution requirements. The CMA offers the advantage of higher signal throughput at lower energy resolutions over the hemispherical sector (7) and a 360° solid acceptance angle when the ion source is coaxial with the CMA. Unfortunately, commercial instruments do not incorporate this design. Normally the hemispherical analyzer of an ESCA or part of a CMA (with the ion gun off axis) is used. For more sophisticated ISS studies, researchers generally change the ion optics and the polarity of the analyzer. Retardation used in ESCA electron optics results in a loss of intensity, and the increased energy resolution obtained is generally not needed. History HREELS. This technique is particularly helpful in the study and interpre-

INSTRUMENTATION tation of vibrational states of gases adsorbed on clean metallic substrates. HREELS is also valuable in the study of chemical (catalytic) reactions and chemical structures on solid surfaces. It is even possible to follow the kinetics of adsorption and reaction with timeresolved spectrometers (~ 10" 3 s). These studies indicate that the dipolar selection rule dictates which vibrational modes can be excited, as it does for optical spectroscopy. HREELS is sensitive to the different sites of adsorption of the molecules and can disclose the structure or morphology of the adsorbate-substrate system (8). Although HREELS has a specific surface sensitivity to the uppermost 20 À layer of a polymeric material, it has only recently been applied to very large molecules and polymer surfaces (9). In such applications, information about the chemical composition, the morphology or structure, and the intrinsic (collective) vibrations of polymer surfaces can be obtained. SIMS and ISS. Both techniques are based on the interaction of low-energy ion beams with condensed-phase surfaces. SIMS involves the mass analysis of sputtered or desorbed secondary ions, and the basic principles have been extended to molecular secondary ions. This allows the derivation of molecular structure information based on molecular ion and fragmentation information similar to that inherent in traditional MS. Recently workers have concentrated on minimizing the sputtering rates produced by the primary ion beam as well as the damage that results from high-current densities and energies; the goal has been to produce "static" or low-damage conditions (10-13). In early work {14) the necessity of lowering ion beam damage conditions was pointed out, and the results from ISS and SIMS were compared with those of ESCA. SIMS analysis of polymers has been further refined (15) using low-mass positive ion fragmentation patterns and negative ion analysis (16). Most recently, Benninghoven, Hercules, and co-workers (17) generated high-mass ions, suggesting the possibility of direct molecular weight distribution determination. In contrast to SIMS experiments in which the results of low-energy ions that implant and transfer their momentum are analyzed, the ion-scattering experiment involves analysis of the fraction of primary ions that scatter with momentum and energy conserved. The scattering can be described by a ratio of the energy after collision Ε to the energy before collision EQ that is related to the masses of the primary ion

(Mi) and target atom (M2) and the lab­ oratory scattering angle (Θ) by the fol­ lowing equation.

ISS has been widely applied in the field of catalysis (18), but because of damage concerns similar to SIMS it has only recently been extended to poly­ mer surface structure and orientation analysis (14,19,20). The primary attri­ bute of the technique is the high sur­ face sensitivity, especially to orienta­ tions of surface atoms. ISS provides complementary information to that obtained with ESCA and allows scien­ tists to build depth profiles and sense differences in orientation. Problems Polymer materials surface analysis. For several reasons, polymers are in­ herently more difficult to analyze than metals, alloys, semiconductors, and even glasses or catalysts. First, they are ill-defined materials both chemically and structurally. They are rarely "clean" or "pure," often contain addi­ tives and stabilizers, and are composed of a distribution of molecular weights. Amorphous and disordered phases in contact with microcrystalline structure are much more common than high crystallinity. Second, they are general­ ly insulating materials. When chargedparticle beams—for example, low-en­ ergy (1-10 eV) electrons in HREELS or low-energy ion beams (500 eV to keV) in SIMS or ISS—impinge on polymers, the surface generally builds up an elec­ trostatic charge that rapidly deflects the probe beam, impeding the record­ ing of any spectrum. Finally, polymeric materials are quite radiation and heat sensitive. HREELS. If the polymer can be pre­ pared as a thin film on a conducting substrate (e.g., by spin or melt casting from a solution, or by using a Langmuir-Blodgett type technique), the re­ sulting film conductivity may be suffi­ cient to eliminate the charging prob­ lem. The charging effect on insulating polymers can be controlled, but the socalled "flood gun" technique is manda­ tory for studying thick insulators (21). Current HREELS experience has shown no evidence of electron-induced damage: The electron beam current hitting the sample is typically in the 10 -11 -10~ 10 A range. Even after a cu­ mulative dose required for high-resolu­

650 A · ANALYTICAL CHEMISTRY, VOL. 62, NO. 11, JUNE 1, 1990

tion spectra, no influence on the vibra­ tional spectra could be detected. Energy resolution is a problem spe­ cific to HREELS. The number of vi­ brational modes in polymers that can be excited is so great that the bands in the electron-induced vibrational spec­ tra of polymers will never be complete­ ly separated. Indeed, the highest in­ strumental resolution now achieved is in the 7-10 meV (56-80 cm - 1 ) range, incomparably worse than that routine­ ly obtained in optical spectroscopy. Moreover, this resolution limit does not seem to be dictated by instrumen­ tal factors and probably results from intrinsic sample properties or the sur­ face preparation. Until now it has not been possible to use HREELS spectral intensities from polymers in a quanti­ tative manner. This point will be devel­ oped later. SIMS and ISS. Polymer and organic materials present particular concerns for ion beam analysis because of the difficulty in understanding the physi­ cal and chemical processes underlying, in the case of SIMS, the ejection of secondary molecular ions representa­ tive of macromolecular bonding orien­ tation and interactions. For ISS of polymers, predicting and verifying reg­ ular functional group orientations that inhibit or encourage scattering (socalled shadowing and shielding) is nec­ essary. Information from SIMS and ISS analysis of the uppermost few mo­ lecular layers complements the power of ESCA and vibrational spectroscopic analysis of polymer surfaces. Because sputtering is an unavoid­ able aspect of the ion beam/solid inter­ action (and in fact the basis for SIMS), concerns about the sampling depth of each method are related to the effects of ion beam based radiation damage. Even under mild ion beam conditions, damage involving breaking and re­ forming molecular bonds occurs, and it is these bonds that impart the impor­ tant structure and properties in poly­ mers. A description of the potentially powerful orientation and bonding in­ formation in ISS and SIMS must be tempered by a clear knowledge of the extent and mechanisms of damage. Each technical advance in minimiz­ ing damage from the ion beam has led to dramatic advances in the types of information available in SIMS and ISS. In evaluating the data, one must consider the following questions. For both SIMS and ISS, what is the role of damage of the native polymer structure on the depth of sampling? For SIMS, what are the formation mechanisms and escape depths of structurally relat­ ed secondary ions when damage is strictly controlled? For ISS, how is the

uppermost atomic layer sensitivity realized under low damage, and how are specific regular surface orientations realized and predicted? Current status HREELS Surface spectroscopic information. To be useful, electron-induced vibrational spectra recorded from polymer surfaces must provide information that will identify the material studied. Dozens of polymers have been analyzed by HREELS, and spectra of simple materials [e.g., polyethylene, polystyrene, and poly(methyl methacrylate)], model compounds such as Langmuir-Blodgett layers, and more complex systems such as poly (ethylene terephthalate), polyimide, and polymer physical mixtures have been published (32). Figure 3 shows the high-resolution spectrum recorded from a clean, fully cured polyimide thin film. The sample was a 200 Â layer deposited by spin coating from a solution onto a gold decorated silicon wafer. Identifying the peaks recorded between 0 and 500 meV (i.e., 0 and 4000 cm - 1 ) required the help of the very large data bank avail-

Wavenumber (cm-1) Figure 3. Electron-induced vibrational spectrum of a clean, cured polyimide surface. Inset shows an IR fingerprint from the same material.

RENT

Electrochemical Excellence at Your Fingertips

Analytical Instruments lease or rent-to-own

/ Free Instrument delivery & setup in selected areas. / GOMSD»FTIR.AA.ICP»LOIR • Choose from many major manufacturers / Hewlett-Packard G O M S D Systems in stock

- Electrochemical Workstation: • on-screen data manipulation • perfect for microelectrode work • easy to use - Microelectrodes - Mini Reference Electrodes - Mini Voltammetric Electrodes

/ New Catalog of Chromatography Supplies. Write or call now for your FREE brochure!

1-800-7-ON-SITE

On-Site Instruments® ENVIRORENTAL® 6Θ9 North James Road

Columbus. Ohio 43219-1837

Cypress Systems, Inc. P.O. Box 3931 Lawrence, Kansas 66046 (913)842-2511

(614) 237-3022 USA—CANADA CIRCLE 106 ON READER SERVICE CARD

CIRCLE 22 ON READER SERVICE CARD

ANALYTICAL CHEMISTRY, VOL. 62, NO. 11, JUNE 1, 1990 · 651 A

INSTRUMENTATION

Figure 4. Comparison of the (a) IR absorption and (b) HREEL spectra from polyethylene. (Note different energy scales.)

able for IR and Raman spectra. When such an optical absorption chart is superimposed on the electron loss fingerprint (Figure 3), the band assignment is almost straightforward: Indeed, there is a one-to-one correspondence between the electronically and optically excited vibrational bands. Figure 4 compares the electron and optical vibrational spectra of the simplest polymer, polyethylene. T h e transmission IR spectrum of polyethylene is very simple, with a low-intensity C-H stretch band around 2950 c m - 1 (not shown) and two major peaks at 1472 c m - 1 and 720 cm - 1 . In contrast, the HREEL spectrum of the same material presents a more complicated fingerprint. The electron-induced vibrational bands correspond to both IR and Raman active modes; the CH* stretches are particularly intense. This suggests that the excitation mechanisms are different in the electronic and optical spectroscopies.

HREEL spectra contain invaluable information that is not available by ESCA analysis. HREELS can detect hydrogen or deuterium present in the material via the CH*, OH, or NH stretch bands. Its energy resolution is sufficient to distinguish aromatic and aliphatic CHX species; even methyne, methylene, and methyl groups can be separated. Finally, carbon-carbon bonds can be fingerprinted; and single, ethylenic, and acetylenic bonds can be distinguished (22). Surface sensitivity. HREELS offers a way to examine vibrational spectra of polymer surfaces with high sensitivity and a much more limited sampling depth than all the conventional optical spectroscopies. Although researchers to date have not agreed on a precise estimation of the probed depth, experimental results suggest that the sampling depth must be extremely small, perhaps < 20 À. Stearic and oleic layers (Langmuir-

652 A · ANALYTICAL CHEMISTRY, VOL. 62, NO. 11, JUNE 1, 1990

Blodgett films) of ~25 À thickness lack any discernible carboxyl stretch vibrational features around 1650 cm - 1 , the band that is usually dominant in IR spectra (23). This result suggests that coherent electron diffusion from an organic surface is restricted to the last few angstroms. For the fatty acids studied, the sampling depth was limited to the outermost -CH2 and -CH3 groups. In effect, the uppermost CH 3 layers dominate the spectra (24). Results from hexatriacontane (C36H72) thin films and polyethylene surfaces also showed a large signal from the methyl vibrations relative to the methylene. The vibrational spectra of polyethylene (Figure 4) clearly shows two bands attributed to - C H 3 and -CH2 scissor modes at 1370 c m - 1 and 1440 cm - 1 , respectively. Considering that in the bulk this high-density polyethylene contains very few -CH3 groups (which terminate both ends of the polymeric chains), the ratio of the CH3/CH2 groups should be extremely small. The high intensity observed for the CH3 scissor modes (1370 cm - 1 ) suggests that the terminating -CH3 groups lie preferentially on the polymer extreme surface and that electron-induced spectra gather information only from the uppermost layer(s) (25). More evidence of the usefulness of the surface sensitivity of HREELS came from a combined IR and electron study of polymer physical mixtures. Normal and perdeuterated polystyrene solutions of the same average molecular number were mixed according to ascribed molecular ratios. Vibrational intensities from the P ( C - H ) and v(C-

D) stretch bands at 3000 c m - 1 were recorded from the solutions with an IR transmission spectrometer and from thin-film surfaces cast from the same solutions with a HREEL spectrometer (Figure 5). Although the optical transmission results show a linear relationship between the band intensities and the inverse of the deuterium molar fraction, the HREELS results clearly present some kind of saturation effect: The film surfaces prepared from low deuterated polystyrene concentrations in the solution disclose a net enrichment in deuterated species. This evidently must be related to selective surface diffusion processes of the perdeuterated chains, during the film casting from the solution, the air drying, or the measurement in UHV. HREELS is the sole electron spectroscopy that could detect this type of effect at the very surface of the composite polymeric films; moreover, it is probable that this effect could not be evidenced by any optical spectroscopy (26). Both dynamic depth

profiling SIMS and high-energy ion scattering have detected this phenomenon in blends of polystyrene and deuterated polystyrene (27, 28). The above results justify the enthusiasm to continue to develop HREELS applications on polymer surfaces. Unique and complementary information on the extreme surface chemical composition is indeed available. Surface morphology or structure. Studies on clean metal surfaces with HREELS have helped to disclose adsorption and reaction sites. By varying experimental parameters such as the scattering geometry or the electron impact energy, it is possible to gain structural information, especially when highly symmetric adsorption sites are 1 considered. However, polymers are not ideal candidates for such investigation because, on the atomic molecular scale, they present quite rough surfaces and several topology variations. As a result, the electron scattering is very diffuse. This was evidenced for stearic acid Langmuir-Blodgett systems deposited

onto substrates (Ag, Au, Al, Ge) with different surface finishes (23). As a consequence, for all the polymers studied up to now, the reflected beam intensity in the elastic and inelastic channels is spread over a range of 10°20° around the specular direction. This spread did not prevent the successful study of the surface morphology of a well-cured and (probably) crystallized polyimide film. Ratios of the intensities of the C-H stretch versus the C = 0 stretch were measured for different scattering geometries. Results showed that the C-H stretch signal is more intense when electrons are collected in a direction close to normal at the surface, whereas the C = 0 band is enhanced in a direction close to the grazing collection angle. This evidence suggested that the CH and CO present different orientations on the surface of the film as prepared, a finding that is in agreement with other structural analyses of polyimide films (29). Quantification and scattering mechanisms. To study the relation-

ULTRAMICRO Fittings and Valves

(actual size)

Valco Capillary GC and Micro HPLC Micro size (1/32"), nuts, tees, crosses, unions, FS adapters, ferrules, reducers, bulkheads, liners, and valves. Low mass coupling of 0.2 to 0.5mm fused silica open tubular columns. Straight thru unions for butt to butt, zero dead volume connection. Repair or join fused silica columns. Fused silica to fused silica, or fused silica to1/32" and larger transfer lines.

VALCO...THE ONIY SOURCE 1-800-FOR-VICI Valco Instruments Co. Inc.

VICI VICI AG

P.O. Box 55603 Houston, Texas 77255 Tel: (713) 688-9345 Fax: (713) 688-8106

Valco Europe

Figure 5. Intensity ratios of the i>(C-H) and v(C-U) stretch bands recorded with an FT-IR spectrometer and with electron excitation for mixtures of normal and deuterated polystyrenes with different molar ratios (X).

Unterlannberg 7 CH-6214 Schenkon, Switzerland Tel: (045) 21 68 68 Fax: (045) 21 30 20

CIRCLE

140 ON READER

SERVICE CARD

ANALYTICAL CHEMISTRY, VOL. 62, NO. 11, JUNE 1, 1990 · 653 A

INSTRUMENTATION ship between HREELS band intensities and the effective number of different chemical functionalities on the polymer surface, one must consider the excitation mechanisms that play an effective role in the electron loss process on polymers: dipolar, impact, and resonance scattering (8, 30). Until recently, this situation prevented direct application of symmetry rules and use of (dynamic) dipolar moment strength to predict band-relative intensities. However, on metal substrates the so-called surface normal selection rule (of the dipolar interaction) is substantially relaxed for at least some of the molecular vibrations. Systematic investigations

of possible image dipole effect from the substrate, polarizability contributions from substrate contaminants, and very low energy excitations through the polymer film (31) will be needed. In addition, careful cross-section measurements (band intensities vs. electron impact energy) on well-characterized polymers and model compounds will be required to elucidate this problem. Currently, HREELS studies of polymer surfaces can produce only qualitative results, except when the studies are performed using standard or reference materials. Metal-polymer interfaces. Detecting minute surface phenomena such as

the incipient phases of a new interface formed on the polymer surface is within the capability of HREELS. This was illustrated by the study of chemical reactions and sites developed during the metallization of polymers {32). Figure 6 shows representative HREELS data collected during aluminum condensation onto PMDA-ODA (pyromelletic dianhydride—4,4'-oxydianiline) polyimide (1,2,4,5-benzene tetracarboxylic anhydride—4,4'-oxydianiline). The metal atoms are evaporated from a Knudsen cell (1 À/min), and Figure 6b shows the spectrum of a clean polyimide surface covered with 1.2 Χ 1014 Al atoms/cm 2 . Around 3000 c m - 1 and between 1100 and 1800 cm - 1 , eight vibrational bands are resolved and assigned, indicating t h a t some bands are more affected by Al deposi­ tion than others. Figure 6c shows a dif­ ference spectrum that enhances this observation. Severely attenuated bands at 1720 and 1120 cm" 1 clearly localize the Al atoms near C = 0 sites. The ν (C-N) band is attenuated simul­ taneously, but not as severely. The C-O-C ether linkage (-1250 c n r 1 ) does not appear to be affected by metal deposition, but a large intensity in­ crease over the whole 1400-1600 c m - 1 loss region shows major intensity trans­ fer between vibrational modes. Either the phenyl ring stretches of the ODA segment in the polyimide are enhanced or the ν ( C = 0 ) and ν (C-N) are red and blue shifted, respectively. Finally, the p(C-H) band above 3000 c m - 1 is signif­ icantly enhanced in Figure 6b. These spectral features suggest that the initially deposited Al atoms prefer­ entially interact with carbonyl oxygen atoms and that the C-N group is also perturbed via electronic derealization on the imide ring. This perturbation might then affect the ODA segments, introducing a conformational change. The C-H bonds may be changing with respect to the surface, increasing the intensity of C-H bands at 1400-1600 cm - 1 , whereas the C6H2 moieties in the PMDA are not affected. This explana­ tion is also consistent with ESCA re­ sults on the same system. Low-energy

Figure 6. HREELS evaluation of polymer metallization. (a) HREEL spectrum of clean polyimide surface, (b) spectrum after deposition of 1.2 Χ 1014 atoms of Al, and (c) difference spectrum (b — a).

654 A · ANALYTICAL CHEMISTRY, VOL. 62, NO. 11, JUNE 1, 1990

ISS

Homopolymer surface structure. Low-energy ISS was first applied to polymers in a study that involved de­ veloping low-damage conditions to pre­ serve slight changes in surface confor­ mation of methacrylates (14). This work showed that the surface composi­ tion of glassy methacrylates was equiv­ alent to the bulk composition, except for three polymers with bulky side chains that might be sterically hin­ dered.

Figure 7. Structures of vinylpyridine polymers derived from molecular mechanics. (a) Top down view of poly(2-vinylpyridine) (P2VP) atactic chain, (b) side view of P2VP atactic chain (note Ν is shadowed by C in backbone), (c) view of five mono­ mers looking down at backbone, and (d) ion-scattering spectra for P2VP and poly(4-vinylpyridine) (P4VP). Note the lack of signal from Ν in the P2VP spectrum. (Adapted from Reference 20.)

ANALYTICAL CHEMISTRY, VOL. 62, NO. 11, JUNE 1, 1990 · 655 A

INSTRUMENTATION This observation remained unex­ plained until recent studies showed that ISS could distinguish between conformations of stereoregular poly(methyl methacrylate) (PMMA) and slight amounts of poly(methacrylic acid) (PMAA) copolymerized random­ ly with PMMA (19). Very different C/O scattered ion intensity ratios were measured in the ISS experiments. An­ gular-dependent ESCA results were

unable to distinguish between any of these polymers. Atactic, syndiotactic, and isotactic PMMA and P M M A / PMAA random copolymers all have the same atomic composition and no dif­ ferences in monomer bonding. Molecu­ lar models were used to explain the dif­ ferences in measured surface atomic concentrations that result from shield­ ing and shadowing of functional groups in different conformations. Similar re­

0.70 Ethylated P2VP

(b)

Butylated P2VP

0.56

I 0.42

ο

Unmodified P2VP

0.28

0.14 -

3.2

4.8

8.0

Sputter time (min)

Figure 8. ISS results from (a) P2VP and P4VP and (b) modified and unmodified P2VP. Note the change in intensity versus time, which indicates very low levels of damage. Reaction with alkyl iodide in solution promotes rotation of nitrogen in P2VP from beneath the polymer backbone. (Adapted from Reference 20.)

656 A · ANALYTICAL CHEMISTRY, VOL. 62, NO. 11, JUNE 1, 1990

sults were obtained with methyl- and phenyl-substituted polysiloxanes, where the Si/O ratios were different. We have since extended this work (20) to probe differences in the shield­ ing of nitrogen functionalities in differ­ ent isomers of vinylpyridine polymers (Figure 7). Rather than explaining the results after the experiment with mod­ els, we studied polymers whose struc­ tures were previously confirmed by molecular mechanics calculations and NMR spectra. Based on this information, we pre­ dicted that different N/C scattered ion intensity ratios would be measured for these polymers. Because the nitrogen in poly(2-vinylpyridine) (P2VP) is shielded by the backbone carbons, no initial nitrogen scattering signal should be measured; this was confirmed ex­ perimentally. We also observed that the N/C ratio increases upon sputter­ ing to the equivalent of poly(4-vinylpyridine) (P4VP), which always shows nitrogen scattering. The quaternization of the nitrogen in P2 VP via meth­ yl, ethyl, or tert -butyl groups causes nitrogen to rotate, because of steric considerations, away from the back­ bone. This rotation yields very differ­ ent N/C signals (Figure 8), indicating the sensitivity to the orientation of this particular functional group. Multitechnique depth profiles of multicomponcnt polymers. Integrat­ ing the ISS method into the multitech­ nique analysis of polymer materials al­ lows its superior surface sensitivity to complement ESCA and FT-IR sam­ pling at different depths. This was il­ lustrated by the analysis of random block copolymers of bisphenol A poly­ carbonate (BPAC) and poly(dimethylsiloxane) (DMS) (33). Quantitative surface compositions determined from angle-dependent ESCA and ISS were combined with scanning electron mi­ croscopy (SEM) to develop a morpho­ logical model of the domain segrega­ tion approaching the air-facing inter­ face. The same approach was then used on polymer blends created from homopolymers of BPAC and DMS. These results allow researchers to compare the forces driving segregation to the near-surface region in a blend of the same components (as in copolymer systems under study) and to observe effects of surface energy differences and blend component compatibility (i.e., interchain bonding interactions) in the overall free energy of mixing. We have been able to show that in blends of < 11% (by weight) DMS, the surface composition as measured by angular-dependent ESCA and ISS was always 85% (by weight) DMS (34). Al-

INSTRUMENTATION

Make

Foui-

Source for STABLE ISOTOPES Isotec Inc. offers the scientific community one of the world's largest selections of stable isotopes and labelled compounds • • • •

NMR Solvents Reagents Organics Inorganics

• Biologicals

You Get More From

though these results suggest that a compatible mixture exists at that con­ centration, this system is considered to be incompatible. At higher concentra­ tions (25% DMS), visible phase separa­ tion occurred and the surface was 100% DMS. We are currently studying a vari­ ety of incompatible and compatible blends, attempting to use ISS to com­ plement angle-dependent ESC A. Another application is in the area of surface-modified polymers (35-37). In a series of studies of different treat­ ments of PMMA, we used the higher surface sensitivity of ISS to develop specific models of orientation of modi­ fied functional groups. ISS results showed a great increase in surface oxy­ gen, yet ESCA results showed a de­ crease in oxidized carbon concentra­ tions (35, 36). These data were com­ pletely consistent with the adsorption of water vapor rather than permanent oxidation. Different treatments that promoted specific oxidation were then developed (37). Contact angle methods were also used to differentiate func­ tionality in the latter study. One out­ come of that work has been attempts to determine the relationships between ISS conformational sensitivity and contact angle techniques. Static SIMS

• • • •

> 99% chemical purity High isotopic enrichment Competitive prices Fast delivery

Plus . . . metal stable isotopes, helium-3 and other noble gas isotopes, other enriched stable isotopes, multiply-labelled compounds and custom synthesis Call or write for your NEW complete price list today!

ISOTEC INC. A Mathenson USA Company 385h ββκ,ΐίΓ Mo*u «SHmisOw urt-»"J>

Homopolymer s t r u c t u r e sensitivity. We will consider three different types of ions (Figure 2) generated in the stat­ ic SIMS experiment from polymers. They can be considered to be roughly divided by mass range (i.e., low, high, and in between), but more generally are divided by the mechanism of for­ mation (fragmentation and rearrange­ ment, desorption and cationization of oligomers, and simple bond breaking/ charge stabilization). The most complex of the three are low mass (i.e., < 300 amu) fragment ions, the original focus of much static SIMS polymer characterization. In the first applications of static SIMS to the analysis of polymer surfaces, low mass (0-150 amu) fragment ions (14, 15) were used to distinguish structure and isomerism in a series of methacrylate polymers with different side chain structures. The surfaces were distin­ guished primarily from positive ion spectra via their fragmentation pattern from the side chain, within a common pattern characteristic of the backbone. A significant result from these studies was the differentiation of isomeric structures of butyl methacrylate, which could not be distinguished by ESCA core-level analysis and by the analysis of slight reactive surface degradations of the tert-butyl methacrylate (38). Subsequent improvement in SIMS

CIRCLE 63 ON READER SERVICE CARD 658 A · ANALYTICAL CHEMISTRY, VOL. 62, NO. 11, JUNE 1, 1990

experimental conditions to lower sur­ face damage induced by the primary ion bombardment has allowed for the analysis of higher mass fragments that are related to the original structure of the polymer (15). In the course of these studies, the commonly accepted de­ scription of static or low-damage con­ ditions evolved from the original cur­ rent density (1 η A/cm 2 ) criterion based on sputter yields to preserve a surface for 1 h (10) to minimum ion dosages (e.g., 10~13 ions) (15, 20). In addition, other criteria for mini­ mal damage have been proposed, some of which are based on the theory that chemical and structural disruption re­ sult from inefficient dissipation of the primary charge input attributable to the inherent low conductivity of the polymer sample (16). The use of a pri­ mary atomic (neutral) beam was sug­ gested in this case to remove the input of charge to the sample, with similar spectral results as derived from prima­ ry ion beam analysis and better nega­ tive ion spectra (16). However, precise spatial and low current control of the atomic beam is much more difficult compared with an ion beam (16). It is instructive to compare results from the ISS work on nitrogen-con­ taining polymers discussed earlier with the typical conditions for static SIMS. In the ISS work, damage to the native polymer structure as assessed by changes in N/C signal intensities (Fig­ ure 8) was apparent after 2 min of irra­ diation at 2 nA/cm 2 (i.e., 1.5 X 10" 12 ions) with 2 kV He. This dosage is at or below static conditions, in part because the sputter yield of He is lower than that of the Ar or Xe primary ions used in a typical static SIMS experiment. The ion beam induced changes were not detectable in comparisons of ESCA spectra taken before and after analysis. Furthermore, simple sputtering theory at the beam currents used would not allow for significant ion beam depth profiling (10). Clearly in this case chemical modifi­ cation and damage of the surface of the polymer is occurring, and sputter-in­ duced consumption of the material is not necessarily involved. The problem of considering the higher damage cross sections in the determination of prima­ ry ion beam conditions for organic spe­ cies, as opposed to sputter yields, has been recognized by Benninghoven's group (17, 39). They have designed a TOF secondary ion mass spectrometer that incorporates a pulsed ion source of very low beam density (< 1 ρ A/cm 2 ). Lower dosages are possible because of the simultaneous detection of all ions from a pulse of primary ions in the TOF experiment. As conditions and instru-

mentation have improved, more infor­ mation has been gained. Because of the simplicity of assignment for higher mass ions, this work has provided in­ formation to characterize homopolymer and multicomponent polymers. Applications of the TOF-SIMS ex­ periment to nylons, polystyrenes, poly(dimethylsiloxanes) and surface-modi­ fied polymers have presented informa­ tion beyond the characterization of the monomer identity (17, 39, 40). For in­ stance, high-mass oligomeric ions can generate an accurate molecular weight distribution (17). The potential here is great; a surface-sensitive molecular weight distribution would be impor­ t a n t for many polymer processing problems. High-mass fragmentation patterns have been postulated to be sensitive to intrachain bonding in ny­ lons (17) and composition in complex polyurethanes (40). Differentiation of isomeric composition is also evident from high-mass ions (17). Recently attention has focused on the ions generated from fragmentation along the chain into charge-stabilized ions, which we have termed "n-mer ions" (41). These ions could provide a great deal of information about intra­ chain bonding and chemical reaction sites. N-mer ions have been observed from a variety of polymers: Teflon (39), homopolymer and copolymer nylons, and segmented poly(ether urethanes) (42). In our most recent work, we dis­ cussed charge stability in ion formation from functional polymers (e.g., the vinylpyridines and substituted poly­ styrenes), with emphasis on the effects of charge and reaction site and struc­ ture determination via analysis of the η-mer ions (41). The potential impact in qualitative and quantitative surface analysis of a variety of multicompon­ ent polymer applications where func­ tional and nonfunctional polymers are combined is obviously great. However, more attention to the simple ability of static SIMS to differentiate the sur­ faces of polymer materials with equiva­ lent atomic compositions and difficult ESCA chemical shift information has meant even greater excitement about fragmentation pattern analysis. Multicomponent polymers. A few studies of multicomponent polymers have successfully used model polymers to explore the effects of fragment ion formation (42-44). Block and random copolymers were analyzed, and frag­ ment ions unique to each structure were identified. Quantitative ion inten­ sity ratios were correlated to composi­ tion, and sampling depths were esti­ mated via comparison with ESCA data. However, many questions generated by this type of study remain to be an­

swered. Low-mass fragment ions origi­ nate from a variety of processes, and in unknown multicomponent polymers it would be impossible to assign ions un­ ambiguously to particular structures or components. Besides the dynamic SIMS depth profile work on labeled polymers men­ tioned earlier (27), a series of very ex­ citing experiments using dynamic SIMS depth profiling on block copoly­ mer thin films by Russell et al. were recently described in an overview (45). This work shows the power of focusing strictly on elemental ions in determin­ ing profiles; where unique isotropic tags can be used to identify one compo­ nent, complex different ion formation mechanisms can be ignored. One ad­ vantageous aspect of the work by Rus­ sell is the correlation to ESCA and light and X-ray scattering. However, even with these advances, much work re­ mains to be done to identify exact con­ ditions that describe sampling depth issues, and instrumental improve­ ments are needed to further enhance the probability of formation and detec­ tion of high mass ions.

When Control Of The Manufacturing Process Is Critical, We're On-Line. Everyday.

Summary What does the future hold for HREELS, SIMS, and ISS? Even with better knowledge about the mecha­ nisms of vibrational energy loss, there are three major challenges to the use of HREELS for vibrational spectroscopy of the surface of polymers: charge com­ pensation, limited resolution, and knowledge of the sampling depth. Pro­ gress is being made on experimental charge compensation to allow full use of the method, and studies on thin films can be performed to resolve other issues. With the best conditioned poly­ mer surface, a clear understanding of the contributions of roughness and electrical properties on the broadening of the elastic peak in HREELS may allow the deconvolution procedures used in IR, Raman, and electron spec­ troscopy to be widely applied (46, 47). This approach is simple in concept for HREELS because the model lineshape that would be deconvolved is available directly in the experimental spectrum. Line narrowing from this approach would provide information on the rela­ tive intensities and identification of bands under a complex envelope. To resolve sampling depth issues, more knowledge about the mean free path of low-energy electrons will have to be gained for model insulating poly­ mers; our laboratories are now collabo­ rating to perform such transmission ex­ periments on thin films. We hope that these efforts will more broadly define the technique so that it will be more

Put the provenreliabilityof our FT-IR on-line for composition monitoring and some positive things are going to happen: • The fast feedback will cut your waste and downtime. • The payback will make you wonder why we don't charge more (we'll tell you why). • Getting real-time, on-line composition data will obsolete your grab sampling. Let our decade of innovation in FT-IR reduce your costs. Call us to see how we can help.

1-800-326-2328

LASER PRECISION ANALYTICAL 17819 Gillette Avenue Irvine, CA 92714

CIRCLE 78 ON READER SERVICE CARD

ANALYTICAL CHEMISTRY, VOL. 62, NO. 11, JUNE 1, 1990 · 659 A

INSTRUMENTATION

Liquid

Chromatography/ Mass

Spectrometry Applications in Agricultural, Pharmaceutical, and Environmental Chemistry

O

ne of the most exciting aspects of LC/MS is its versatility and the potential it of­ fers for many new applications. This im­ portant volume provides an in-depth look at this analytical process and describes the in­ strumentation now available or under development. This book discusses recent innovations such as the use of thermospray and particle beam interfaces that have made LC/MS methods much more compatible for the analysis of pes­ ticides, pharmaceuticals, and pollutants. The first section in this 18-chapter volume describes applications of LC/MS to the analysis of polar or thermally labile agricultural chemi­ cals including chlorinated herbicides and pesti­ cide residues. The second section focuses on applications in the analysis of intractable phar­ maceuticals and their metabolites. The last section offers a better understand­ ing of LC/MS methods used in environmental analysis. Included are examples of the use of LC/MS for the analysis of target and nontarget environmental pollutants, as well as their me­ tabolites and conjugates. Mark A. Brown, Editor, California Department of Health Services Developed from a symposium sponsored by the Divi­ sion of Agrochemicals of the American Chemical Society ACS Symposium Series No. 4 2 0 3 1 2 pages (1990) Clothbound ISBN 0 - 8 4 1 2 - 1 7 4 0 - 8 LC 89-18611 $64.95

American Chemical Society Distribution Office, Dept. 64 1155 Sixteenth St.. N.W. Washington. DC 20036 or CALL TOLL FREE

800-227-5558 (in Washington, D.C. 872-4363) and use your credit card!

applicable in concert with other surface science methods to study polymer sur­ faces and interfaces. One application involves m e t a l polymer interfaces. Determining the initial stages of metal deposition and the functional groups at the surface that are important to adhesion promo­ tion are the goals of these studies. In the area of ion beam methods (specifically, low-energy ISS), two im­ portant applications to polymers should be possible with further devel­ opments in instrumentation. The socalled impact collision ISS experiment (ICISS), developed by Aono and Souda (48) to identify bond orientations from shadow cone formation considerations, would allow the determination of exact bond angles in highly oriented regular surface structures. For some types of systems this information would be highly desirable. In addition, electronic structure-related effects on charge neutralization are completely un­ known in ISS of polymers; these funda­ mentals should be investigated. A better understanding of the sur­ face sensitivity and structure/orienta­ tion sensitivity will be extremely useful in complementing ESCA for analysis of the uppermost surface interface. It ap­ pears that such data will help resolve the details of contact angle measure­ ments with ESCA and ISS. This infor­ mation could be useful in quality con­ trol and on-line analysis. Although contact angle measurements can be used more readily, detailed spectro­ scopic information to monitor surface processes such as corona discharge treatments would be invaluable. Finally, we are excited about the po­ tential of SIMS for polymers. In addi­ tion to the current lew-mass uses, there are unanswered and controversial questions about the limitation in ion formation at high mass. If these ques­ tions can be answered, the molecular weight distribution of the near-surface region may be determined. When this information is compared with the bulk molecular weight, segregation and pro­ cessing effects can be known. Impor­ tant surface properties related to segre­ gation of low molecular weight species plague a variety of polymer interfacial adhesion and compatibility issues. The determination of chain orienta­ tion and intrachain bonding in the near-surface region may be available from η-mer ion distributions. Hydro­ gen bonding in proteins in different conformations may be deduced from the probability of formation of specific ions. The average length in monomer units of the folds and loops in a poly­ mer chain can then be determined from the distribution of rc-mer ions.

660 A · ANALYTICAL CHEMISTRY, VOL. 62, NO. 11, JUNE 1, 1990

All of these possibilities complement surface compositional information that is available from other methods and give hope that the future of applica­ tions of these methods will be bright for polymer analysis. The contributions of J. H. Magill, R. L. Schmitt, J. H. Wandass, P. A. Cornelio, K. J. Hook, T. J. Hook. L. Salvati, R. L. Chin, T. G. Vargo, C. Grégoire, M. Rei Vilar, M. Vermeersch, Y. Novis, and P. A. Thiry are gratefully acknowledged. Most of the HREELS work was performed while JeanJacques Pireaux was research associate at the Belgian National Fund for Scientific Research, and we acknowledge the financial support provided by that institution. Work at Buffalo is sponsored by the National Science Foundation Polymers Program of the Division of Materials Research. This joint research on electron and ion spectroscopies is sponsored by NATO project 0563/88.

References (1) a. Biotechnology and Materials Science: Chemistry for the Future; Good, M. L., Ed.; American Chemical Society: Washington, DC, 1988. b. Polymeric Materials: Chemistry for the Future; Alper, J.; Nelson, G. L., Eds.; American Chemical Society: Washington, DC, 1989. (2) Research Briefings 1986: Interfaces and Thin Films; Committee on Science, Engineering and Public Policy of the National Academy of Sciences, National Academy of Engineering and Institute of Medicine; National Academy Press: Washington, DC, 1986, pp. 2-15. (3) Nebesney, K. W.; Masschoff, B. L.; Armstrong; N. R. Anal. Chem. 1989, 61, 469 A. (4) Gardella, J. Α., Jr. Anal. Chem. 1989, 61, 589 A. (5) Dilks, A. Anal. Chem. 1981,53, 802 A. (6) Pireaux, J. J. J. Electron. Spectrosc. Relat. Phenom., in press. (7) Barrie, A. In Handbook of X-ray and Ultraviolet Photoelectron Spectroscopy; Briggs, D., Ed.; Heyden: London, 1977, pp. 79-119. (8) Ibach, H.; Mills, D. L. Electron Energy Loss Spectroscopy and Surface Vibra­ tions; Academic Press: New York, 1982. (9) Pireaux, J. J.; Grégoire, C; Vermeersch, M.; Thiry, P. Α.; Caudano, R. Surf. Sci. 1987,189/190,903. (10) a. Benninghoven, A. J. Vac. Sci. Technol. A 1985, A3(3), 451. b. Benninghoven, Α.; Rudenauer, F. G.; Werner, H. W. Sec­ ondary Ion Mass Spectrometry: Basic Concepts, Instrumental Aspects, Appli­ cations and Trends; Wiley: New York, 1987; Chapter 5, pp. 672-753. (11) Winograd, N. Scanning Electron Microsc. 1985,///, 919-26. (12) Pachuta, S. J.; Cooks, R. G. Chem. Rev. 1987,87,647. (13) Colton, R. J. Nucl. Instrum. Methods Phys. Res. 1983,218, 276-86. (14) a. Gardella, J. Α., Jr.; Hercules, D. M. Anal. Chem. 1980, 52, 226. b. Gardella, J. Α., Jr.; Hercules, D. M. Anal. Chem. 1981,53,1879. (15) Briggs, D.; Hearn, M. J. Vacuum 1986, 36,1005. (16) Brown, Α.; van den Berg, J. Α.; Vickerman, J. C. Spectrochim. Acta 1985, 40B, 871. (17) a. Bletsos, I. V.; Hercules, D. M.; Greifendorf, D.; Benninghoven, A. Anal. Chem. 1985, 57, 2384. b. Bletsos, I. V.; Hercules, D. M.; vanLeyen, D.; Benning­ hoven, A. Macromolecules 1987, 20, 407.

(18) Horrell, Β. Α.; Cooke, D. L. Catal. Rev. Sci. Eng. 1987,29(4), 447-91. (19) Hook, T. J.; Schmitt, R. L.; Gardella, J. Α., Jr.; Salvati, L., Jr.; Chin, R. L. Anal. Chem. 1986,58, 1285. (20) Hook, K. J.; Gardella, J. Α., Jr.; Sal­ vati, L., Jr. Macromolecules 1987, 20, 2112. (21) Liehr, M.; Thiry, P. Α.; Pireaux, J. J.; Caudano, R. Phys. Rev. Β 1986,33, 5682. (22) Wandass, J. H., Ill; Gardella, J. Α., Jr. Surf. Sci. 1985, 750, L107. (23) Wandass, J. H., Ill; Gardella, J. Α., Jr. Langmuir 1986,2, 543. (24) Wandass, J. H., Ill; Gardella, J. Α., Jr. Langmuir 1987,3,183. (25) Pireaux, J. J.; Thiry, P. Α.; Caudano, R.; Pfluger, P. J. Chem. Phys. 1986, 84, 6452. (26) Rei Vilar, M.; Schott, M.; Pireaux, J. J.; Grégoire, C; Caudano, R.; Lapp, Α.; Lopes da Silva, J.; Botelho do Rego, A. M. Surf. Sci. 1989, 211-12, 782. (27) Valenty, S. J.; Chera, J. J.; Olson, D. R.; Webb, K. K.; Smith, G. Α.; Katz, W. J. Am. Chem. Soc. 1984,106,6155. (28) Jones, R.A.L.; Kramer, E. J.; Rafailovich, M. H.; Sokolov, J.; Schwarz, S. A. Phys. Rev. Lett. 1989,62, 280. (29) Pireaux, J. J.; Vermeersch, M.; Grégoire, C; Thiry, P. Α.; Caudano, R.; Clarke, T. C. J. Chem. Phys. 1988, 88, 3353 (30) Thiry, P. Α.; Liehr, M.; Pireaux, J. J.; Caudano, R. Phys. Scripta 1987,35, 68. (31) DiNardo, N. J.; Demuth, J. J. Chem. Phys. 1986,85,6739. (32) Pireaux, J. J.; Vermeersch, M.; Degosserie, N.; Grégoire, C; Novis, Y.; Chtaib, M.; Caudano, R. In Adhesion and Friction; Grunze, M; Kreuzer, H. J., Eds.; Springer Verlag: Heidelberg, 1989, pp. 53-66. (33) Schmitt, R. L.; Gardella, J. Α., Jr.; Magill, J. H.; Salvati, L., Jr.; Chin, R. L. Mac­ romolecules 1985,18, 2675. (34) Schmitt, R. L.; Gardella, J. Α., Jr.; Sal­ vati, L., Jr. Macromolecules 1986,19,648. (35) Hook, T. J.; Gardella, J. Α., Jr.; Sal­ vati, L., Jr. J. Mater. Res. 1987,2(1), 117. (36) Hook, T. J.; Gardella, J. Α., Jr.; Sal­ vati, L., Jr. J. Mater. Res. 1987,2(1), 132. (37) Vargo, T. G.; Gardella, J. Α., Jr.; Sal­ vati, L., Jr. J. Polym. Sci. Part A: Polym. Chem. 1989, 27, 1267. (38) Gardella, J. Α., Jr.; Novak, F. P.; Her­ cules, D. M. Anal. Chem. 1984,56,1371. (39) Bletsos, I. V.; Hercules, D. M.; Magill, J. H.; vanLeyen, D.; Niehus, E.; Benninghoven, A. Anal. Chem. 1988,60,938. (40) Bletsos, I. V.; Hercules, D. M.; van­ Leyen, D.; Benninghoven, Α.; Karakatsanis, C. G.; Rieck, J. N. Anal. Chem. 1989, 61, 2142. (41) Hook, K.J.; Hook, T. J.; Wandass, J. H., Ill; Gardella, J. Α., Jr. Appl. Surf. Sci. 1990,44, 29. (42) (a) Briggs, D. Org. Mass Spectrom. 1987, 22, 91. b. Hearn, M. J.; Ratner, B. D.; Briggs, D. Macromolecules 1988, 2i,2950. (43) Hearn, M. J.; Briggs, D.; Yoon, S. C; Ratner, B. D. Surf. Interface Anal. 1987, 10, 554. (44) Briggs, D.; Hearn, M. J.; Ratner, B. D. Surf. Interface Anal. 1984, 6, 184. (45) Russell, T. P.; Deline, V. R.; Wakharkar, V. S.; Coulon, G. MRS Bulletin 1989, XIV(10),33. (46) Langell, Μ. Α., Department of Chem­ istry, University of Nebraska, personal communication, 1989. (47) Mittlefehldt, E. R.; Gardella, J. Α., Jr. Appl. Spectrosc. 1989,43(7), 1172. (48) Aono, M.; Souda, R. Jpn. J. Appl. Phys. 1985,24(70), 1249.

Joseph A. Gardella, Jr. (left) is on leave from the State University of New York at Buffalo and currently serves as program officer for analytical and surface chemis­ try in the chemistry division of the National Science Foundation. He received a Ph.D. from the University of Pittsburgh (1981) and joined the faculty at Buffalo in 1982 after postdoctoral research at the University of Utah. His research inter­ ests involve surface chemistry and structure of organic and biological thin films, multicomponent polymers, and biomaterials. Special emphases include ion beam and electron microscopies applied to these systems. Other interests include the philosophy of science, curriculum and teaching, basketball, baseball, and bowling (the sport of kings or philosophers). Jean-Jacques Pireaux (right) is professor of physics at the Facultés Universitaire Notre Dame de la Paix in Namur, Belgium. He received his Ph.D. in physics from FUNDP Namur (1976) and since then has been associated with the Laboratoire Interdisciplinaire de Spectroscopie Electronique at Namur. There he pursues research in electron spectroscopies (XPS/ESCA and electron-induced vibrational spectroscopy) to characterize polymer surfaces, polymer-metal interfaces, and modified polymers. Pireaux has authored more than 100 papers in these fields and is currently European editor of the Journal of Electron Spectroscopy and Related Phenomena.

NIST

QUALITY MEASUREMENTS =

UNITED STATES STANDARD REFERENCE DEPARTMENT OF COMMERCE NATIONAL INSTITUTE OF STANDARDS MATERIALS AND TECHNOLOGY The National Institute of Standards and Technology has developed a series of SRM's to serve as calibrants, test mixtures, and standardization materials for Quality Control of analytical instrumentation and methodology. MEASUREMENTS and STANDARDS are important to everyone who needs quality. NIST has over 1,000 Standard Reference Materials that can help you calibrate instruments and check on measurement accuracy. For more information phone or write for a free catalog. Telephone (301) 975-OSRM (6776) FAX (301) 948-3730 STANDARD REFERENCE MATERIAL PROGRAM Building 2 0 2 , Room 2 0 4 National Institute of Standards and Technology Gaithersburg, MD. 2 0 8 9 9 CIRCLE 94 ON READER SERVICE CARD

ANALYTICAL CHEMISTRY, VOL. 62, NO. 11, JUNE 1, 1990 · 661 A