XPS Analysis of Nanostructured Materials

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 18

Journal of Electron Spectroscopy and Related Phenomena 178–179 (2010) 415–432

Contents lists available at ScienceDirect

Journal of Electron Spectroscopy and


Related Phenomena
journal homepage: www.elsevier.com/locate/elspec

XPS analysis of nanostructured materials and biological surfaces


D.R. Baer ∗ , M.H. Engelhard
Environmental Molecular Sciences Laboratory, Pacific Northwest National Laboratory, Box 999, Richland, WA 99352, USA

a r t i c l e i n f o a b s t r a c t

Article history: This paper examines the types of information that XPS can provide about a variety of nanostructured
Available online 22 September 2009 materials. Although it is sometimes not considered a “nanoscale analysis method,” XPS can provide a
great deal of information about elemental distributions, layer or coating structure and thicknesses, sur-
Keywords: face functionality, and even particles sizes on the 1–20 nm scale for sample types that may not be readily
Nanomaterial analyzed by other methods. This information is important for both synthetic nanostructured or nanosized
Nanoparticle
materials and a variety of natural materials with nanostructure. Although the links between nanostruc-
XPS
ture materials and biological systems may not at first be obvious, many biological molecules and some
Biosurfaces
Catalysis
organisms are the sizes of nanoparticles. The nanostructure of cells and microbes plays a significant
Particle size role in how they interact with their environment. The interaction of biomolecules with nanoparticles
is important for medical and toxicity studies. The interaction of biomolecules is important for sensor
function and many nanomaterials are now the active elements in sensors. This paper first discusses how
nanostructures influences XPS data as a part of understanding how simple models of sample structure
and data analysis can be used to extract information about the physical and chemical structures of the
materials being analyzed. Equally important, aspects of sample and analysis limitations and challenges
associated with understanding nanostructured materials are indicated. Examples of the application of
XPS to nanostructured and biological systems and materials are provided.
© 2009 Published by Elsevier B.V.

1. Introduction sample features impact XPS data. In a web-based presentation Yang


and Sacher describe many different aspects of XPS measurements
X-ray photoelectron spectroscopy (XPS) has become an increas- associated with nanoparticles, including references from their own
ingly available and powerful tool for understanding the nature of work and some of historical importance [4].
many different types of surfaces. Although advances in existing The increasing importance of XPS for many types of analysis
and newly developing tools with high spatial resolution receive a arises for several reasons. First, XPS can provide information
good deal of press if they improve the analysis quality of individual about the actual composition and chemical state of surfaces and
nanosized features of materials, XPS is an important, established interfaces that dominate properties of nanostructured materials.
and frequently essential tool for understanding several important Surfaces are equally important for the function of biological
aspects of nanostructured natural and synthetic materials that can- organisms and to the performance of synthetic biomaterials.
not easily be obtained using other techniques. In addition, it is Second, although it is frequently not considered to be a tool with
finding increasing applications for analysis of the surfaces of biolog- “nano” resolution, because the electrons associated with XPS
ical systems as well as a more traditional role in the characterization only travel distances measured in nanometers these electrons
of synthetic materials (biomaterials) designed to be used in biologi- can be used to provide a good deal of information about the
cal environments. Microbe and cell surfaces might be appropriately structure of nanometer-sized features of a sample near a surface
considered as very complex nanostructured systems. region, as described below. Although the possibilities for obtaining
Characterization of nanostructures using XPS is not new. From nanometer-scale information from samples with flat surfaces
the earliest days, XPS or ESCA (electron spectroscopy for chemical may be more apparent to many researchers, information from
analysis) was widely used to study many aspects of nanosized cat- XPS spectra can be used to determine the size of nanoparticles
alyst particles [1–3]. Over the past 30 years many researchers have (sometimes in circumstances for which other types of high spatial
explored important aspects of how particle size or other nanosized resolution measurements cannot be applied) as well as provide
information about coatings and layers on particles.
A wide variety of challenges are associated with the characteri-
∗ Corresponding author. zation of nanostructured materials [5,6], some specific to XPS and
E-mail address: don.baer@pnl.gov (D.R. Baer). others applicable to many types of measurements on nanostruc-

0368-2048/$ – see front matter © 2009 Published by Elsevier B.V.


doi:10.1016/j.elspec.2009.09.003
416 D.R. Baer, M.H. Engelhard / Journal of Electron Spectroscopy and Related Phenomena 178–179 (2010) 415–432

tured materials and nanoparticles in particular. Similar issues apply Therefore, new analysis techniques and new methodologies
to the study of biological surfaces and they certainly apply to the for extracting information from data need to be developed
intersection of materials science, biology and nanotechnology that [6]. Their development will require close linking of modeling
is sometimes called nanobiomaterials. These issues include sample efforts and multiple types of data. Some sophisticated efforts
handling and preparation, the impacts of time and the environ- to model XPS data (including SESSA [19] and QUASES [20])
ment on samples, problems associated with sample damage, and highlight some of the progress and possibilities in this area.
fundamental issues associated with understanding and possibly
defining the structure of nanosized objects. Grainger and Castner This paper is not a comprehensive review but an introduction to
[7] make three recommendations related to analysis of nanobio- different ways that XPS can be applied to obtain information about
materials (paraphrased below) that echo concerns and suggestions nanostructured materials, including nanostructures that are either
of other reports in the literature: biological in nature or connect materials to biological systems.
Specialized instruments such as synchrotron based XPS, including
(1) Existing tools are underused. It is necessary to perform com- unique high pressure XPS capability [21] of importance to catalysis
plete physical and chemical characterization of nanoparticles and biological materials [22], are not specifically addressed. Some
used in research (and for most applications). There are sev- of the topics introduced in this paper will have been described in
eral indications that we are not fully using the tools currently greater detail in other contributions to this special issue of JESRP.
available to adequately characterize nanoparticles and other Only aspects related to nanostructured or biological surface anal-
nanostructured materials. A report from a nanobiotech com- ysis are discussed here. The next section of the paper outlines
mercialization meeting noted the opinion of several researchers how nanostructure influences XPS signals; it discusses issues raised
that important physical and chemical properties of nanoparti- generally and specifically for XPS analysis of nanostructured mate-
cles used in research were unreported and apparently unmea- rials. Section 3 considers some of the limitations for applying XPS
sured [8]. A Working Party on Manufactured Nanomaterials of to nanostructured materials. Section 4 provides examples of the
the Organization for Economic Co-operation and Development application of XPS to nanostructured materials (with examples of
(OECD) has established a list of physical–chemical properties nanolayers and nanoparticles) and biological systems (with exam-
and material characterization needs for nanostructured mate- ples of microbe surfaces and biomaterials).
rials [9]. Several surface properties are listed among the needed
parameters, but in many circumstances the importance of the 2. How nanostructure influences XPS signals
surface chemistry of nanoparticles, especially as applied to
toxicity, has been under-emphasized [10]. The importance of The short distances that electrons can travel within a material
surface information about these materials is one reason that without losing energy (inelastic mean free path [IMFP]) provide
XPS measurements are being increasingly applied to them. the physical reason that XPS is sensitive to the surface composi-
However, it has usually been found necessary to systematically tion of materials. Differences in IMFPs and distribution of electrons
apply a variety of tools to adequately understand the proper- that have lost energy can be used to extract information about the
ties of nanoparticles [5,6,11]. A similar need applies to studies nanostructures of many materials. Calculated versions of IMFPs for
of biologically related surfaces [12–14]. several materials [23] are shown in Fig. 1. These curves show that
(2) Nanomaterials and biological surfaces exaggerate analysis chal- the average distances that electrons can travel without energy loss
lenges. Because of the high sensitivity of nanoparticles to their due to inelastic scattering depend on the kinetic energy of the elec-
environment, because they often change with time and are trons and the material within which they are traveling. These IMFPs
easily damaged (altered during measurement), and because relate to the information depth from which XPS electrons arise.
surface contamination can have dramatic impacts on material The contribution to the detected signal from a depth z into the
properties as well as what is measured [5], it is essential material will be attenuated by material covering the layer closer
to establish reliable measurement, sample preparation and to the surface. Although the equation is only approximately valid
handling, and storage protocols. The importance of carefully [24], this relationship is often expressed as follows:
assessing the interactions of these with their environment  −z

raises several analytical objectives, needs, and challenges. dIz ≈ I1 exp dz (1)
( cos )
Sample handling and preparation guides have been prepared
by ISO and ASTM committees [15,16]. Sample lifetime or
time-dependent effects influence analysis needs but may
also have significant impact on shelf and product lifetime for
useful applications. It is equally important to understand the
strengths, limitations, and impacts of the variety of analysis
methods that can be applied to these materials [7,17]. Because
they can significantly impact analysis, some of the fundamen-
tal limitations of both routine and more advanced XPS data
analysis will be introduced in this paper.
(3) New tools and techniques are needed. Scientific and technological
advances associated with nanostructured materials, biological
systems, and nanobiomaterials require detailed information
about the composition, physical and chemical structure and the
presence of contaminants or impurities of complex systems,
frequently while they remain in their natural or operating envi-
ronments [7]. In some circumstances the information needed is
beyond the reach of current analysis tools. It has been observed,
for example, that there are no robust methods for extracting the
needed structural information about nanostructured materials Fig. 1. IMFPs calculated for polymethylmethacrylate (PMMA), silicon dioxide, Si,
[18], an issue the authors called the nanostructure problem. Cu, and W. (From [23]).
D.R. Baer, M.H. Engelhard / Journal of Electron Spectroscopy and Related Phenomena 178–179 (2010) 415–432 417

ally determine as much as possible about the actual distribution


of elements. The influences of specimen structure appear in XPS
data in several different ways. By understanding how nanostruc-
tures impact XPS data and by combining this knowledge with other
knowledge about the nature of a specimen it is often possible to
obtain important structural or elemental distribution information
that can be difficult or impossible to obtain in other ways.
The nanostructure of a specimen influences XPS data in several
different ways including the following:

• Peak intensities and relative peak intensities of


o peaks for different elements
o different peaks for the same element
o variation as a function of emission angle
• Peak energies
o binding energies of peaks
o value of the Auger parameter
• Background signals from electrons that have lost energy.

Depending on what is known about the specimen and the anal-


Fig. 2. Schematic drawing of possible distributions of different elements or com-
pounds within the volume analyzed by XPS. Most routine XPS analysis assumes ysis objective, each of these influences can be used to extract useful
that there is a uniform distribution (top example), while in many circumstances information about nanostructured samples.
the distribution is non-uniform. It is often possible to extract information about the
nanostructure with careful analysis of the data. 2.1. Relationships between nanostructure and relative peak
intensities
where dIz is the intensity of the detected signal at depth z, I1 is
the intensity that would have been produced if the layer were at In addition to determining the “average” composition of a sur-
z = 0 (the outer surface),  is the IMFP,  is the angle of the detected face, relative signal intensities can be highly useful in obtaining
electron relative to the surface normal (see Fig. 3a), and dz is the information about layered structures of samples and sometimes the
thickness of the layer. Because  depends on energy (Fig. 1), Eq. (1) sizes of particles. Because the information depths will vary with the
demonstrates both the energy and angle dependence of the sig- energy of the photoelectron (or Auger-electron) analyzed as well
nal intensity from depth z. The information depths from which as the angle at which they are collected, peak ratios can be used
detectable signal intensities can be extracted are typically in the to learn about compositional layering, enrichment, and segrega-
range of 1–20 nm. XPS does not collect information about regions tion of elements toward or away from the surface of a specimen.
deeper into the material than the information depth, unless these The examples provided in this section focus on flat samples for
depths are exposed for analysis using methods such as sputter which the concepts are most easily understood, but the same con-
depth profiling. cepts extend to other types of nanostructured materials including
As XPS data analysis is most commonly applied, information nanoparticles, nanotubes and complex catalysts.
that could be used to extract information about the structure of The impact of angle and energy on the information depth are
nanomaterials is often ignored. Five of many possible distributions schematically shown in Fig. 3. For low kinetic energies (high pho-
of two elements (light and dark) within the information depth of toelectron binding energies) or high angles of emission (emission
a flat film are schematically shown in Fig. 2. In many cases, XPS normal to the surface is defined as  = 0), the information depth is
data are analyzed assuming that all elements measured are uni- less than for normal emission or higher kinetic energy peaks. When
formly distributed throughout the layer analyzed (the top image in the information depth is shallow the relative signal intensities from
Fig. 2). However, in most circumstances it is likely that the elements the elements located nearest the surface are enhanced. Therefore,
detected during XPS measurement are not uniformly distributed by measuring peak intensities as a function of emission angle, the
(such as layers, clusters, or buried clusters). For both research variation of elemental peak ratios provides layering information
and technological applications it is often of importance to actu- [25]. The use of angle-resolved XPS (ARXPS) can be applied in dif-

Fig. 3. (a) Drawing showing differences in information depth as the angle between sample surface normal and the path toward electron analyzer varies. The information
depth also changes with the energy of the electron detected. (b) Schematic drawing indicating how differences in information depth (due to photoelectron energy or angle)
influence the amount of a multilayer film structure detected. Comparison of data from small and larger information depths provides information about the sample layering.
418 D.R. Baer, M.H. Engelhard / Journal of Electron Spectroscopy and Related Phenomena 178–179 (2010) 415–432

is 0.30, which is very close to that of pure Ni (0.29), suggesting no


enrichment or depletion of Ni in the oxide layer under the carbon
contamination layer. Similar effects would occur for different layers
of multi-layered surface structures.
When a specimen is known to have two layers and XPS photo-
electron peaks unique to each layer are present, the thickness of the
layers can be estimated using several different analysis approaches
usually based on equations similar to Eq. (1). Cumpson has pre-
sented a graphical approach to calculating the thickness of the layer
[27] using peak intensity ratios, instrumental sensitivity factors,
and the energy ratio for the photoelectrons involved. The involve-
ment of the energy ratio provides a method for accounting for the
different attenuation length or electron free paths for photoelec-
trons of different energies.
Peak amplitudes can be used in a variety of ways to extract lay-
ering information, both for flat films and other structures as noted
below. Mohai [28] has developed a free computer program that
uses peak amplitudes to analyze layer structures using only XPS
signal amplitudes. XPS MultiQuant is described as a “quantitative
Fig. 4. XPS survey spectra of an oxidized Ni foil with carbon overlayer contami-
nation. The relative peak ratios of different Ni peaks are influenced by the carbon evaluation program for X-ray Photoelectron Spectroscopy, serving
overlayer. as a practical and universal tool for the surface analyst. It applies
the ‘classic’ methods of the quantitative calculations using the inte-
grated intensity of the measured XPS lines.” As one example, XPS
ferent ways to get information about elemental enrichment (or MultiQuant has been used to characterize and confirm the structure
depletion) at the surface or to construct more detailed near-surface and quality of a Langmuir Blodgett film on glass [29].
depth profiles. A spreadsheet (ARCtick) for assisting layer structure determi-
nation from ARXPS measurements is provided by the National
2.1.1. Layer thickness and structure Physical Laboratory [30,31]. ARCtick allows comparison of a variety
The relative intensities of photoelectron peaks from a layer are of methods that can be used to extract thickness, elemental dis-
impacted by the layer thickness and the thickness of any layers that tribution and elemental layering and provides some suggestions
may cover it. For the oxidized Ni sample in Fig. 4, the intensities of when different approaches may be most useful (e.g. regularization
the Ni peaks are impacted by the oxide layer thickness and by a [32] may be the best method when concentration gradients may
contamination layer of carbon on top of the oxide. The ratio of the be important not just discrete layers). As described below XPS data
Ni 3p/Ni2p3 photoelectron peak intensities (obtained using peak simulation or analysis programs (e.g. QUASES [33] or SESSA [34])
integration and Shirley background subtraction) for a thick, sputter may be useful for understanding layer thickness and structure.
cleaned Ni specimen (for typical analysis conditions applied in our
Phi Quantum 2000) is 0.29. Because the kinetic energy of the Ni 2.1.2. Molecular orientation
3p photoelectron peak is significantly higher than that for the Ni A variety of methods can be used to determine the orienta-
2p3, the ratio can provide information about surface enrichment tion of molecules attached to a surface including near-edge X-ray
or depletion. If the ratio is less than 0.29, there is relatively more adsorption fine structure (NEXAFS) [35] and optical methods, such
Ni 2p signal than for a uniform sample, suggesting, for example, as surface enhanced Raman spectroscopy [36]. Although XPS can be
Ni enrichment near the outer surface of the sample. If the Ni 3p/Ni insensitive to molecular orientation in some circumstances [37],
2p3 peak ratios are >0.29, it would suggest that there is less Ni near for ordered molecular assemblies [38] and larger molecules with
the surface. If, as a different example, only a thin film of Ni were asymmetric elemental distribution [39], XPS can be used to deter-
present, the Ni 3p/Ni 2p3 ratio could provide information about the mine some information about molecular orientation. For ordered
thickness of the film (if it were thinner than the information depth molecular assembles, there can be shifts in the binding energies
of the Ni 3p electrons). Castle is among the researchers who have associated with molecular orientation [38]. For the larger molecules
used this approach to understand the structure of corrosion films the primary impact is that related to signal strength due to depth
for transition metal alloys [26]. or layering effects as discussed above. Alessandrini et al. [39], for
It is also important to recognize that layers on top of the ele- example, have used XPS to show differences in protein orientation
ments of interest will usually alter the peak ratios from elements on surfaces by comparing the magnitudes of N 1s photoelectron
within a buried film. The impact of these coating layers can provide peak amplitudes. Frequently, XPS is used in combination with
useful information but requires that some care to be applied to the other methods to understand the overall composition, uniformity
analysis to avoid confusion. For the oxidized Ni sample in Fig. 4, and molecular orientation [40]. XPS is often applied to determine
the measured Ni 3p/Ni 2p3 peak ratio is 0.35, possibly indicating molecular orientation of the relatively large molecules associated
Ni depletion from the surface. Such depletion is real in this case, with adhesion [41].
but the depletion is only due to a significant presence of carbon on
the surface. The depletion does not provide us with any informa- 2.1.3. Coatings on particles
tion about Ni distribution in the Ni film itself and is not likely to be In nanotechnology and many related areas, it is useful to under-
information of any value. Several approaches could be used to con- stand the nature of coatings on particles, and there are increasing
sider the impact of the carbon on the other elemental peaks. Using numbers of applications of what are often called core-shell par-
the carbon correction applied by Castle [26], the carbon overlayer ticles. Although there are some added complexities, many of the
can be estimated to be approximately 8 Å thick and the impact on approaches to determining layer thickness for thin films can be
the other (not carbon) peak amplitudes can be estimated and a cor- applied to particles of different shapes [42], and over the years a
rection made. Without carbon correction, the Ni3p/Ni2p3 ratio is number of people have considered quantitative analysis of contam-
0.35, as noted above, while the ratio with the intensity correction inants on particles [43]. A schematic drawing [44] (Fig. 5) shows the
D.R. Baer, M.H. Engelhard / Journal of Electron Spectroscopy and Related Phenomena 178–179 (2010) 415–432 419

Fig. 7. Normalized signal from the surface layer to the substrate for a film (lower line
and  = 0) and nanoparticle (darker line with data points). For larger sized particles
the shape effects are relatively insensitive to curvature. As particles get smaller
and approach the electron escape length, the amount of signal from surface layer
increases. (After [44]).
Fig. 5. Schematic drawing showing a coating on a thin film and on a nanoparticle.
Note the different distances that electrons from the substrate need to travel to be
detected by the analyzer (effective coating thickness). For the films, the distance is
described below) or possibly that from a rough surface that has
defined by the emission angle. For an isolated particle, the distance varies depending also been considered in some detail.
on the location on the particle (but is independent from the emission angle). Because Shard et al. [45] have systematically examined some of the
of the effects of the direction of electron emission, the ratio of signals from the effects of geometry on coating thickness measurements and have
coating and the thin film or nanoparticle will generally differ. However, it is possible
derived a correction factor (topofactor) for the thickogram [27] that
to model these effects. The results of one model calculation are shown in Figs. 6
and 7. adjusts for known particle shapes. Using values of particle (sphere
or cylinder) radius and electron attenuation lengths, it is possible to
calculate the overlayer thickness. Although these topofactors can
differences in effective film thicknesses measured by XPS for a coat-
vary with layer thickness and particle radius, for many conditions
ing on a flat film and on a cylinder or sphere. The differences in the
(when particles are smaller than the X-ray penetration distances
ratio of intensities from coating to substrate at normal emission
[usually microns] but larger than electron escape depths) the topo-
( = 0) for coatings of the same thickness on a flat film, on a cylin-
factors are relatively constant with values of 0.67 for spheres and
der, and on spheres are shown in Fig. 6. In these cases, the radius of
0.79 for cylinders. The topofactors confirm the understanding con-
the cylinders or spheres is assumed to be significantly larger than
veyed in Fig. 5. If a measurement is made at normal emission angles
the electron attenuation length. As the radius of a sphere (or cylin-
( = 0) for the general sample surface, but the surface is actually
der) approaches the size of the electron path length, XPS begins to
made up of spheres, a simple flat layer calculation would over-
sense the whole particle and the influence of the coating is further
estimate the layer thickness. In particular, the calculation made
enhanced (Fig. 7). Although the shape of curves for different ele-
assuming the flat surface needs to be reduced by about one-third
ments and film thickness may be similar to Fig. 7, the actual values
(multiplied by 0.67). The amount of correction needed increases
will change as a function of the materials involved and the energies
when the particles become smaller than the distances of electron
of the photoelectron peaks detected [45].
escape. Shard et al. also have considered the analysis of shadowed
As shown by the significant changes in the signal ratio shown
spheres (a pile of spheres), which may be a relatively common con-
in Fig. 7, as particle size approaches the size of the IMFP (d ∼ ) dif-
dition for analysis of a collection of larger nanoparticles, as opposed
ferent analysis approaches are required. Although coatings on both
to a distribution of smaller supported nanoparticles, which occur
larger and smaller nanoparticles was of interest and importance,
in catalysis applications.
the approach for determining coating thickness for smaller parti-
It is relevant to mention the concept of a “magic angle” for XPS
cles with d ∼  will be more complex than for larger particles where
data collection. A magic angle is an angle of XPS data collection
d ≫  [46]. The condition when d ∼  has been widely considered
for which surface roughness effects are averaged or minimized
and applied to supported metal catalysts and is described later in
allowing approximately “correct” values of coating thickness to be
this section. For the condition d ≫ , the determination of coating
determined [47,48]. A limited understanding of this may be pro-
thickness might be considered that of either a specific shape (as
vided by considering a surface made up of spheres. The collection
angle for electrons emitted from a sphere (without any support-
ing substrate effects) would be independent of the collection angle.
However, if the calculated coverage were made assuming a flat sur-
face with  = 0◦ the calculated thickness of the coverage would be
larger than the true value. However, if this “flat surface” calcula-
tion for the spheres were made assuming  = 45◦ rather than the
experimental situation of  = 0◦ , the cos () impact on the cos()
term in Eq. (1) would be to go from 1 to 0.71. In effect we are using
an “effective” tilt angle for this spherical surface. Chatelier et al.
[49] developed a method to determine the effective or averaged
tilt angle for arbitrary rough surfaces using atomic force microscopy
(AFM) data. In the spherical case, the  = 45◦ correction to the escape
depth produced a correction value that happened to be between
the corrections for spheres (0.67) and cylinders (0.79) calculated
by Shard et al.
The proposed application of magic angle data collection is to
Fig. 6. Diagram showing the relative intensities of surface and substrate signals
actually collect the data at the magic angle and to analyze the data
for a coating on a thin film (at normal emission,  = 0), a cylindrical particle, and a
spherical particle. The nominal coating thickness is 0.5 nm and the calculations were using magic angle in a version of Eq. (1). Although circumstances
performed using an IMFP of ≈1.65 nm. exist where this method is useful, two limitations need to be con-
420 D.R. Baer, M.H. Engelhard / Journal of Electron Spectroscopy and Related Phenomena 178–179 (2010) 415–432

sidered. First, there is no true magic angle; detailed calculations al. [55] note that because of a high catalytic activity, metal clus-
of the magic angle for different rough surfaces suggest that the ters grown on oxide substrates have been the subject of increasing
magic angle can vary from 30◦ to 55◦ [45]. Second, analyzing a research interest and understanding the origin of these shifts is of
rough surface (depending on the scale of roughness) may introduce increased importance. In many cases, the binding energy (BE) shifts
shadowing of some photoelectrons and cause an unknown selec- vary with sufficient regularity such that Gonzalez-Elipe et al. [56]
tive collection of the photoelectrons. These issues can be avoided proposed that BE peak position could be a useful way to determine
if a sample with known surface roughness (or a collection of par- particle size, possibly better than some approaches involving peak
ticles with well-defined geometry) is present and the geometry is intensity described above. One important process of changing BE
ordered enough to use a Shard et al. [45] correction. Note that even for isolated particles (including those supported on insulating sub-
in the most ideal circumstance a magic angle approach would not strates) is a Coulomb attraction between ejected electrons and the
include impacts of small particles with d ∼ . remaining charged particle [57]. A simplified analytical expression
of the Coulomb effect would suggest that the BE of a nanoparti-
2.1.4. Particle size cle would increase as the inverse of the particle diameter, which is
Research teams have developed different methods for extract- observed in many cases [52]. Several other sources of the BE shifts
ing particle size information from XPS spectra. This has been have been identified and likely apply to varying degrees in different
particularly important for catalysis research when small metal par- circumstances. They include changes in the particle energy levels
ticles are present on various types of support material. Often the due to size, lattice strain, charge shifting to a substrate interaction,
nature of the support and the size of the particles make direct and final state relaxation effects. The detailed understanding of the
observation of particles by microscopy (such as transmission elec- mechanisms causing the shift is subject to ongoing research.
tron microscopy (TEM)) on actual catalysis difficult, particularly These BE shifts have been of fundamental interest and when
for particles less than ∼1–2 nm in diameter. Obtaining particle size the photoelectron peak shifts are combined with Auger line shifts
information from XPS requires modeling of the signals. One of the to determine an Auger parameter there can be a strong correlation
early approaches looked at ratios of signals from particles to that among the Auger parameter variation and BE shifts with particle
of the support with a model structure [50] and appeared to be very size and catalysis turnover frequency (TOF). In other words, these
successful for many applications [3]. A second approach uses the shifts in the electronic behavior of nanoparticles correlate with cat-
signal intensity ratio of photoelectron or Auger-electron peaks of alytic activity and can be used to relate catalytic activity to the
different kinetic energies arising from the particles [51]. Yang and electronic charge on the particles [3,58]. These BE shifts are also
Sacher [52] have used this type of method to examine Cu clusters of interest to those designing sensors [59]. Differences between
deposited on Dow Cytlotene 3022. They have shown that there is Auger parameter shifts and BE shifts can be used to determine dif-
excellent agreement among TEM measurements of many clusters, ferences between initial state (energy levels including substrate
XPS determination of the average cluster size, and AFM measure- interactions) and final state (relaxation) effects [55,60].
ments of cluster size (after a tip shape correction). The equation An example of the BE shifts for Au is provided below within the
used by Yang and Sacher involves measuring a peak intensity ratio context of developing materials for chemical sensing and catalysis.
and applying the following equation:
2.3. Nanostructure from background signals
I 0 [1 − exp(−d/1 )]
R = 10 (2)
I [1 − exp(−d/2 )]
2
In addition to altering the relative intensity of peaks, layer-
ing and atomic distribution alters the presence of electrons not
where I10 and I20 are the intensities of the two peaks that would be readily identifiable as part of photoelectron peaks because they
obtained from very thick layers (d ≫ ). This ratio would be equal have lost energy due to inelastic scattering and no longer have
to the ratio of appropriately determined sensitivity factors; d is the the energy characteristic of the photoelectron peak. These energy-
average cluster diameter and the two s are the appropriate IMFPs loss electrons have been considered by many researchers, but
for the kinetic energies of the two measured transitions. Tougaard [20,61] has developed a systematic approach to using
In both methods of determining particle size, the calculation this background signal to obtain information about the elemen-
relates to supported clusters or nanoparticles for which the par- tal distribution in the outer 20 nm of a sample. He correctly calls
ticles are at least somewhat separated from each other. These this “a tool for the analysis of surface nanostructures by electron
approaches apply only to particle sizes for which signals are within spectroscopy.”
the information depth of XPS. If d is significantly greater than  for Some relationships between nanostructure and background sig-
both s, R = 1 and no information can be extracted [3]. The methods nal are schematically shown in Fig. 8. Four different elemental
are particularly useful for relatively low density of substrate sup- distributions of Cu are shown, all of which produce the same
ported particles and would apply to aggregates for which electrons intensity for the Cu 2p photoelectron peaks. However, significant
generated in adjacent particles might travel through a particle and alterations exist in the distribution of electrons at lower kinetic
reach the detector. energy (higher binding energy). This background signal from Cu
Venezia [3] notes that contamination, which alters peak ratios as photoelectrons that have lost energy before they escaped the sam-
described above, limits the accuracy of the above methods of par- ple actually contains information about the depth distribution of
ticle size estimation. To the extent possible care should be taken Cu in the volume of material analyzed by XPS. When combined
to minimize contamination. It may be possible to use a carbon with high-lateral resolution methods, the Tougaard background-
overlayer correction with the assumption that coatings are uniform analysis approach can be used to obtain a three-dimensional image
[53]. of the composition of the near-surface region of a complex film
or material [20,61]. Tougaard has developed a computer model-
2.2. Nanostructures and peak energies ing program, QUASES [62], that can be used to create a model of
the elemental distribution near the surface of a material and to
For the past 30 years, it has been established that metal clus- calculate the spectrum background that would be caused by that
ters and nanoparticles, particularly when the clusters are supported distribution. By iterating between the model and experimental data
on nonmetal supports, have XPS binding energies that differ from it is possible to establish with reasonable confidence the likely
those of bulk materials of between 0.05 eV and 0.6 eV [54]. Tao et elemental distribution in the sample. QUASES is used to fit the
D.R. Baer, M.H. Engelhard / Journal of Electron Spectroscopy and Related Phenomena 178–179 (2010) 415–432 421

lection and analysis, the possibility of environmentally induced


changes in samples, specimen alterations and contamination
derived from mounting methods, susceptibility of nanoparticles
to damage during analysis, and limitations about what is initially
known about the about the system being analyzed.

3.1. Contamination

Because nanostructured materials inherently involve many


atoms associated with surfaces or interfaces, the presence of unex-
pected surface species is common and can sometimes significantly
impact the analysis. As noted above, the very common layers of
adventitious carbon can alter peak ratios and even mask impor-
tant signals. Sometimes impurities have been introduced to sample
surfaces during “cleaning” processes. The minimization of contam-
ination is an important issue for specimen preparation, handling,
and mounting. In addition to carbon being introduced during
sample handling, chemical impurities can be introduced during
synthesis or processing. Both extreme care and consistency checks
are needed to assure that what is measured is characteristic of the
sample and not contamination or some other artifact.

3.2. Environmentally induced changes


Fig. 8. Diagrams and spectra showing the relationships between elemental distri-
bution and background signal for different distributions of Cu in the near-surface of
a sample. The four different elemental distributions of Cu produce the same inten- XPS is inherently a vacuum-based method (although differen-
sity for the Cu 2p photoelectron peaks. However, there are significant alterations in tially pumped synchrotron based systems allowing samples to be
the distribution of electrons at lower kinetic energy (higher binding energy). (After maintained at higher pressures have been developed [21]). Some
[20,62]). samples are altered by placing them in a vacuum and the possible
impact of the vacuum should be considered. In many circumstances
background for individual peaks. Several developments and adap- these effects are less than expected, but a variety of approaches can
tations of QUASES have been and are being made. Hajati et al. [63] be used to minimize such effects as needed. Approaches include
have combined QUASES-generated models to determine the size cooling samples and collecting data as a function of time to deter-
and diffusion of Au clusters formed on a polystyrene surface, and mine if there are vacuum-induced changes. Equally important,
QUASES has been used as part of the development of a software tool exposure of some samples to air before analysis can alter them. In
for the generation of survey spectra in XPS to simulate wide spectra some of our research, we need to examine iron metal-core oxide-
with a range of 200–1500 eV for nanostructured surfaces [64]. shell nanoparticles. We have found, as shown in Fig. 9, that we can
A new National Institute for Standards and Technology (NIST) remove a specimen from solution in a N2 environment and see both
database called Simulation of Electron Spectra for Surface Analysis the metal-core and the oxide-shell. However, exposure of the same
(SESSA) has been developed for quantitative Auger-electron spec- particles to air induced additional oxidation and the signal from
troscopy (AES) and XPS [19]. The database contains extensive sets the core was significantly diminished. Therefore, in our research
of physical data important for AES and XPS. SESSA can simulate AES with reactive metal nanoparticles, we usually handle, mount, and
and XPS spectra for a multi-layered thin-film sample designed by a analyze particles without exposing them to air [65].
user for measurement conditions relevant to the user. This database
and simulation program combines the current understanding of the
physical processes of electron spectroscopy and a large amount of
data to simulate the full spectra that would be expected for the sam-
ple designed by the user. Comparison between experimental data
and the SESSA calculation enable information about the elemen-
tal distribution to be extracted. SESSA does not include chemically
induced alterations in peak shape but calculates the full spectra
including the background signals. Although there are not yet as
many applications of SESSA as QUASES, it has potential for great
use and value.
QUASES, SESSA, and related modeling programs make use of the
most up-to-date understanding of the processes associated with
electron spectroscopy and can be used to extract much more infor-
mation about the nanoscale distribution than processes that simply
use peak amplitudes.

3. Barriers and limitations to obtaining information about Fig. 9. Fe 2p photoelectron spectra from iron metal-core oxide-shell nanoparticles
nano- and bio-structures showing the effect of different handling methods. The first spectrum was measured
after removal from storage in a glovebox and transferred to an XPS system without
exposure to air. The second blue spectrum shows the reduction of metallic iron
Several challenges may need to be overcome to obtain the infor-
(∼707 eV) after a few minutes of exposure to air [11]. (For interpretation of the
mation desired from a nanostructured material [5,7]. Among these references to color in this figure legend, the reader is referred to the web version of
challenges are the impact of surface contamination on data col- the article.).
422 D.R. Baer, M.H. Engelhard / Journal of Electron Spectroscopy and Related Phenomena 178–179 (2010) 415–432

addition to the possible introduction of contamination during


mounting or methods of attaching specimens to a sample holder,
some properties of nano-objects can be influenced by the proxim-
ity of other particles or by interactions with the substrate. Although
many additional effects are likely, those reported include the fol-
lowing [5]:

• the buildup of charge during XPS measurements of metal clusters


supported on insulating substrates
• coupling of plasmon modes in metal nanoparticles depending on
particle separation
• coupling of quantum states
• alterations of magnetic and electronic properties
Fig. 10. Schematic drawing of different collections of nanoparticles and how they • substrate-induced charge transfer.
may be grouped for analysis. In some circumstances the collections of nanoparticles
will have properties that differ from individual particles. Particle–particle separation
distances and substrate interaction can alter particle binding energies in XPS and the Shard et al. have noted that when  ∼ d, the use of XPS to deter-
proximity of neighboring particles can alter the electronic, magnetic, and chemical
mine the thickness of a layer on a specimen will be significantly
properties of the particles.
impacted if the objects are not distributed in a single layer [45].
The importance of different types of mounting will depend on the
It seems useful to note that because of the high surface areas, nature of the analysis questions. If the presence or absence of a
environmental effects [66] can have a larger than expected influ- functional group, the cleanliness of a sample or the simple pres-
ence on a variety of properties of nanosized objects, including ence of a coating is the analysis objective, single-layer mounting
surface structure, phase transformations, chemical stability, poly- may not be important. However, if the thickness of a coating on
mer structure, and chemical state. One aspect of environmental smaller nanoparticles (d < 6 nm) is desired, single-layer mounting
effects is that some properties can vary as a function of time [67]. will likely be important.
Thus, from the view of analysis, storage, and functional properties, it Sample preparation is one of the major issues or challenges for
can be important to recognize that the properties of nanostructured analysis of microbe and other biological surfaces using XPS [22,68].
materials are NOT usually constant in time. Environmental and time Issues associated with analyzing microbes and other biological
stability are important properties that may impact technological materials with vacuum-based techniques have received a good deal
uses of nanosized objects. of discussion and are one reason that XPS data are often correlated
with other types of data and checked for self-consistency. A review
3.3. Sample preparation and mounting paper by van der Mei et al. [68] discusses at some length sam-
ple preparation methods. Most research has been conducted on
The method by which nanosized objects are mounted can sig- freeze-dried microbes as outlined in Fig. 11. The process involves
nificantly impact some types of particle properties (Fig. 10). In culturing the microbes, washing them, freeze drying and mounting

Fig. 11. Schematic drawing of steps associated with preparation of microorganisms for XPS analysis using the freeze drying procedure (After [68]).
D.R. Baer, M.H. Engelhard / Journal of Electron Spectroscopy and Related Phenomena 178–179 (2010) 415–432 423

for inserting into the spectrometer. Duplicate and reference sam-


ples are generally required to assure that consistent, meaningful
and reproducible results are obtained.
Although much successful work has been completed using the
freeze drying process, researchers continue to search for methods
to allow biological specimens to remain and relevant physiological
environments (e.g. the presence of water). This search has included
application of differential pumped high pressure XPS systems at
synchrotrons [22]. In a different approach, Leone et al. [69] have
successfully used a fast-freezing procedure to prepare samples to
study deprotonation reactions on Bacullus subtilis as a function of
pH. The ratio of protonated-to-deprotonated N 1s photoelectron
Fig. 13. Schematic drawing of electron pathways toward an electron detector with
peak intensity as a function of pH is shown in Fig. 12. (a) no scattering, (b) inelastic scattering (for which the electron loses energy and
changes direction), and (c) elastic scattering (for which the electron changes direc-
3.4. Damage tion, but does not lose energy).

Some properties of nanoparticles, such as melting temperature,


are dependent on particle size [5]. Therefore, the susceptibility of to determine the particle size. When particles are being analyzed,
particles to X-ray and electron damage may be greater than for knowledge of the particle size can be used to assist determining the
larger-scale versions of the same materials. We have found that thickness of coating layers. Third, it is often highly informative to
for ceria nanoparticles the nature of X-ray-induced changes of the correlate information from a variety of methods that provide sim-
observed cerium chemistry depends on the processing history of ilar information. When a self-consistent understanding involving
the sample [5] and the presence of water or organic layers on the several types of data can be established, confidence in the resulting
surface. Polymers and many organic layers are easily damaged by understanding of the material is significantly increased.
X-rays.
3.5.2. Things we know we do not know
3.5. Pre-knowledge
Much of the discussion regarding peak ratio and film thickness
above was based on IMFPs and the validity of Eq. (1) or similar
3.5.1. Using what can be known
equations. However, it is now well established that Eq. (1) is an
The more that is known about a system that is being analyzed,
approximation that is not valid in all circumstances [70]. One of
the more precise and detailed the information that can be collected.
the challenges is that Eq. (1) assumes that electrons are either not
There are three useful aspects of this. First, if nothing is known (or
scattered (Fig. 13a) or are inelastically scattered (Fig. 13b) (they lose
acknowledged) about the structure of a sample, it can be very chal-
energy, change direction, and are not detected as part of the inten-
lenging to obtain all of the information that might be possible or
sity of a photoelectron peak). However, a variety of experimental
needed. For example, if the sample is known to be layered, any
and theoretical studies show that elastic scattering of electrons can
known or expected information about the layered structure can
be significant, particularly for higher angles of emission (Fig. 13c).
be tested using XPS data and calculating realistic models of the
For example, consider the experimental data from Seah and White
structures. Without the initial knowledge, learning the same infor-
[71] for a thin silicon dioxide layer on a silicon wafer shown in
mation may be possible, but it may take considerably longer to get
Fig. 14. The ratio of the Si 2p photoelectron intensities for Si in the
the information, including a need to collect additional data. Second,
film and the substrate for different angles are plotted based upon a
information provided by other methods can assist in understand-
ing XPS data and might influence the types of analysis done. If, for
example, TEM or XRD data indicate the presence of small nanoparti-
cles (in contrast to a uniform film) it may be possible to use XPS data

Fig. 14. Summary plot showing the effects of elastic scattering as a function of angle.
The plot of the ln (1 + Rexp ()/R0 ) vs 1/cos  for XPS measurement using Al K␣ X-rays
is for a film of SiO2 on Si. The solid circles are experimental data of Seah and White
Fig. 12. Ratio of protonated-to-deprotonated N 1s photoelectron peak intensity for [71]. The solid curve is a SESSA calculation allowing elastic scattering and the dashed
Bacullus subtilis as a function of pH. (From [69]). line has elastic scattering turned off. (After [73]).
424 D.R. Baer, M.H. Engelhard / Journal of Electron Spectroscopy and Related Phenomena 178–179 (2010) 415–432

common equation for film thickness [72]: • Dispersion of components in a nanocomposite film for proton-
  exchange membrane fuel cells [81]
Rexp () d • Functional group termination and the coverage of self-assembled
ln 1+ = (3)
R0 L cos  monolayers on surfaces [82,83]
• Properties of thin corrosion films [26]
where L is the electron attenuation length, d is the film thickness, • Stability polymer light-emitting diodes [84]
Rexp is the experimental ratio of the Si 2p photoelectron peak inten- • Thin polymer and block polymer films [85].
sities for Si in the oxide layer and Si substrate (Iox /Im ), and R0 is the
ratio for these peaks for bulk materials. If this relationship were Nanoparticle examples:
accurate for all angles, a plot of ln (1 + Rexp ()/R0 ) vs 1/cos  would
produce a straight line. However, as shown by Powell et al. [73] • Presence or removal of contamination layers [86]
and in Fig. 14, the experimental data do not fit a straight line for • Oxidation of nanoparticles [87]
 > 65◦ . The SESSA modeling program described earlier can be setup • Nanoparticle size determination [63,88]
to include only inelastic scattering or to allow elastic scattering. • Electrical character of core-shell nanoparticles [89]
When only inelastic scattering is included in the model, a straight • Confirmation of the functionalization or measurement of the
line is predicted. However, when elastic scattering is included, devi-
electronic structure of carbon nanotubes [90–92].
ations from the linear behavior are predicted consistent with the
experimental data.
Note that in Eq. (1) the electron mean free path, , was used Although less familiar to some researchers, surface analysis
while in Eq. (3) the electron attenuation length (EAL), L, was used. tools of various types are increasingly being applied to study the
The EALs differ from the corresponding inelastic mean free paths surfaces of biologically related systems [14], including the surface
because of the elastic scattering of the signal electrons. NIST has chemistry of microbes [68], the formation of biofilms [69,93], the
databases for both of these parameters [34] and comparison of val- creation of biocompatible materials [94,95], drug elution on med-
ues for a few different circumstances are discussed by Powell et ical devices [96,97], and pharmaceutical materials [7]. The surface
al. [73]. Many of the commonly applied expressions using L or  chemistry of microbes is important because they can attach to sur-
assume that elastic scattering can be neglected. However, modi- face and to form biofilms both on artificial surfaces (medical devices
fied expressions taking these into account have been developed such as stents) and on natural surfaces, such as arteries [98]. The
[74,75]. design of body-compatible biomaterials for medical application is
One impact of elastic scattering, which is demonstrated in also a frequent area of study.
Fig. 14, is that elastic scattering is important for glancing angles of To a significant degree many surface tools are applied to biolog-
electron emission. A corollary to this observation is that for flat sur- ically related surfaces to answer the same types of questions that
faces the experimental geometry can be arranged such that elastic are asked for other nanostructured materials and in other areas.
scattering effects are not significant. For thin-film measurements, These questions include what elements are on the surface, what
the effect of elastic scattering has been included in modeling stud- is their chemical state, what is their distribution and how do they
ies of ARXPS for thin films [76], the use of multiple spectral lines change as the environment or substrate changes? In other words,
to calculate film thickness [77], and, as already suggested, can be XPS is applied to biologically relevant materials for the same rea-
modeled by SESSA [19]. However, as implied in Fig. 5, for nanoparti- sons it is applied to other materials. Because of the complexity
cles, electrons effectively leave the particles at a variety of emission of biological surfaces and the roles of different analysis tools in
angles. To our knowledge, the role of elastic scattering on measure- providing complementary information, XPS is commonly used in
ments of particle size or coating thickness for nanoparticles has not combination with a wide variety of other methods [12]. As already
been examined in detail. noted, issues associated with sample preparation, sample damage
In addition to any complications to applying  or L due to issues and understanding the implications of the collected data are impor-
related to elastic scattering, it must be recognized that in some tant challenges and care must be taken. In comparison to general
cases nanostructured materials are polymorphs of larger sized application of XPS to materials, the application to many biologi-
versions of these same materials [78]. Nanoparticles may have dif- cal organisms is relatively less developed and certainly much less
ferent structures and lattice parameters [79,80] than bulk materials widely applied.
and the impact of these changes on electron path lengths has not A few examples provide an initial perspective of the types of sys-
been explored. tems where XPS is being applied. In most cases, the integration of
Although these unknowns will impact the overall accuracy of XPS with other methods and sometimes tests of biological function
our ability to extract information about nanostructures from XPS are essential components of the research.
data, they do not alter the overall relationships, nor do they change Examples involving microbes or other natural biological mate-
the nature of how nanostructure influences the XPS data. rial:

• Oxidation state of chromium on Shewanella oneidensis in relations


4. Examples of application of XPS to nanostructured to contaminate interaction [99]
materials and biological systems • Relationship between surface functional groups and physico-
chemical properties of Esherichia coli [100]
Many different research teams have used XPS to characterize • XPS analysis of chemical functions at the surface of Bacillus subtilis
a wide variety of nanostructured materials. Many of them have [101]
focused on surface films or nanolayers, while others have focused • Surface properties of the human fungal pathogen Aspergillus fumi-
on nanoparticles of various types. Typical examples in each area are gatus [12]
listed below. • Characterization of the cell surface and cell wall chemistry of
Surface film or nanolayer examples: drinking water bacteria [102]
• Review of XPS study of microbial cell surfaces, including a
• Heating-induced segregation on metal alloys [19] database and discussion of sample preparation [68]
• Structure and thickness of Langmuir Blodget films on glass [29] • Impact of herbal extract on the growth of urinary calculi [103]
D.R. Baer, M.H. Engelhard / Journal of Electron Spectroscopy and Related Phenomena 178–179 (2010) 415–432 425

• XPS characterization of flea digestion and iron chemsitry related (graphene) as they are synthesized, functionalized and applied.
to Y. pestis transmission by fleas [104]. XPS is frequently used to examine the functionalization, doping
and changes in electronic structure of these nano-forms of car-
Examples involving biocompatibility, synthetic biomaterials or bon. For many applications XPS is applied along with a combination
functionalized surfaces: of complementary tools to characterize the range of properties of
interest.
• Attachment and delivery of bisphosphonate from metal stent sur- Many research groups are working to take advantage of novel
faces [96,97] electronic and structural properties of CNTs for a variety of appli-
• Surface functionalization of drug-eluting stent with diamond-like cations and fundamental research. It is frequently useful to attach
carbon nanocoating [95] molecules to the CNT surface or to actually enclose molecules
• Protein attachment to polystyrene surfaces (relevant to biomed- within CNTs. It is often important to understand the extent of sur-
ical devices and sensors) [94] face functionalizaton [110] as well as any changes that have taken
• Immobilization of proteins on nitrilotriacetic acid-terminated place in molecules sorbed to the CNT surface or the extent to which
self-assembled monolayers [105] the properties of the CNTs are altered. For example, the combina-
• Analysis of DNA molecules on patterned surfaces (microarrays) tion of XPS and GC-MS thermal analysis was used to examine the
[13] CNTs functionalized by solvent-free and aqueous-based arenedi-
• Design of a surface for selective release of DNA [106] azonium. Both methods suggested that only the arcne group was
• Study of competitive protein adsorption related to the develop- retained on the CNTs [90]. In another study XPS and TEM were
ment of materials that have improved compatibility with blood combined to provide information about the noncovalent bonding
[107]. and distribution of proteins (1-pyrenebutanoic acid, succinimidyl
ester) to SWNTs [111].
Several specific examples (4.7 and 4.8) highlight application of XPS By using both binding energy data and overall composition data
to measure the surface chemistry of microbes and creation of func- XPS can provide several types of information about the CNTs being
tionalized surfaces for biological studies. examined. The combination of TEM and XPS was used to study
the growth of Au nanoclusters on CNTs functionalized by oxygen
4.1. Verification of surface functionalization and product plasma. Core-level and valence band measurements showed little
formation charge transfer between the Au clusters and MWCNTs, However,
the authors noted that the cleanliness of the surface, the presence of
Perhaps the most common uses of XPS for characterization amorphous phases and agglomeration of as grown CNTs appear to
of nanostructure materials involve confirmation of the reactions be constraints to CNT applications [91]. As another example, Core-
expected to occur during synthesis or the presence of functional level and valence band XPS in combination with near-edge-X-ray
groups on a surface. These types of experiments take advantage adsorption fine structure spectroscopy have been used to study the
of the surface sensitivity and chemical state information available impact of TCNQ on the electronic structure and bonding of SWCNTs
from XPS. The types of detailed analysis methods described ear- [92].
lier are not needed to confirm the presence or absence of certain In addition to functionalization, XPS is useful for observing the
elements or functional groups. These are sometimes highly impor- thickness of graphene layers [112] and their purity [113].
tant experiments that involve relatively straightforward use of XPS
measurements. 4.3. Ultra thin SiO2 layers
As a first example, in a set of clever experiments by Lim et al.
[108], d and l cysteines were attached to the surfaces of Au nanopar- The oxide layers on Si are of high importance to the electronic
ticles and used to regulate interparticle chiral recognition. For these industry and have been widely subject to detailed examination.
studies, XPS was used to confirm the presence of the cysteines Seah and co-workers around the world have conducted an exten-
and to demonstrate the consistent nature of the surface coverage sive set of studies on how to precisely and accurately determine
and binding of the different types of cysteines to the nanoparti- the thickness of oxides within the thickness range of 0.3–8 nm
cle surfaces. An import role of XPS for these measurements was to [71,114–120]. In a series of studies of at least seven papers Seah
confirm that the surface coverage did not significantly change as and various co-workers have demonstrated several different points
different cysteines were used and that they were actually attached that are worthy of being highlighted here.
to the surfaces of the Au nanoparticles. As a second example, a Based on the development of a set of oxides of very well-defined
novel approach was used to create core-shell nanoparticles with thicknesses and the use of different measurement angles, Seah and
apoferritin shells and LuPO4 cores [109] as shown schematically in others have identified critical values of R0 and L for use in Eq. (3)
Fig. 15a. In this study, Lu+3 is diffused into apoferritin nanoshells [71], they have examined the various equations proposed to deter-
and then PO4 3− is added. They combine to form an insoluble com- mine film thickness and have identified those that produce accurate
pound in the core. TEM showed the presence of the core after the and reproducible values for SiO2 [117], and they have shown that
synthesis process and XPS was used to confirm the presence of the ARXPS can be used to map the variations in oxide thickness over a
Lu and PO4 , as shown in Fig. 15b. An important function of XPS in wafer [71].
these experiments was to actually verify that the compounds antic- Two observations deserve particular attention because they
ipated were actually formed and remained within the nanoparticle highlight some of the challenges in achieving accurate measure-
shells. ments of properties at the nanometer size. A simplified drawing
of an oxide film on a Si substrate (Fig. 16a) shows that there are
4.2. Nanostructured carbon (carbon nanotubes (CNTs) and four regions: a contamination coating, the main oxide, an interface
graphene) region with suboxides, and the substrate. Accurate determination
of the thickness for the series of standard films involved subtrac-
A wide range of tools are being applied to characterize and tion of satellite peaks, spin-orbit splitting removal, and inclusion
understand various forms of carbon nanotubes (CNTs), including of the suboxide peaks that occur between the Si0 and SiO2 main
single wall carbon nanotubes (SWCNTs or SWNTs) and multi- peaks (as shown in Fig. 16b). Using their best approach, including
walled CNTs (MWCNTs), and single-layer sheets of carbon atoms the suboxide peaks with the optimum geometry, they could achieve
426 D.R. Baer, M.H. Engelhard / Journal of Electron Spectroscopy and Related Phenomena 178–179 (2010) 415–432

Fig. 15. (a) Drawing of the approach used to create core-shell nanoparticles with apoferritin shells and LuPO4 . Lu+3 is diffused into apoferritin nanoshells and then PO4 3− is
added. They combine to form an insoluble compound in the core. TEM showed the presence of the core after the synthesis process and XPS confirmed the presence of the Lu
and PO4 as shown in (b) valence band region showing Lu 4f and Cl 3s and 3p peaks, and (c) P 2p region, respectively. (After [109]).

repeatability of 0.025 nm when using Mg X-rays. Thus, a first obser- error in measurements as thinner films are analyzed. Since these
vation is that the oxygen at the interface (the suboxides) between offsets are often between 0.2 nm and 1 nm, they can contribute con-
the oxide film and Si substrate contributes to the thickness and, siderable uncertainly to nanometer film thickness determination.
if included in the measurements, gives high reproducible results When done with care, the elemental specificity of XPS allows this
that are linear with oxide thickness. Ignoring the suboxides at the offset to be avoided or minimized.
interface decreased accuracy and reproducibility. The measurements of the silicon oxide layers for thin films
The second observation of significance involves comparison of on wafers have become quite sophisticated and accurate. Because
a wide variety of measurements of oxide thickness using differ- nanoparticles may have different oxidation behaviors, it is of inter-
ent methods, including TEM and X-ray reflectivity [120]. When est to compare the oxidation rates for planar and nanoparticle Si.
examined with some care for the well-defined oxides of varying Using the types of analysis discussed above for core-shell particles,
thickness, it was observed that most of the methods were linear Yang et al. [87] have used XPS to examine the oxidation rates for
with increases in the oxide thickness, but there was an offset at zero both flat Si and Si nanoparticles as shown in Fig. 18. Both the oxi-
(Fig. 17). Although each of the methods applied was highly precise, dation rate and the activation energies appear to be significantly
the impact of effects such as undefined contamination layers or different. These measurements would be subject to the electron
water sorption, provided an offset to the measurements as a func- path length uncertainties noted earlier, but appear to be quite sig-
tion of thickness that can produce an increasingly larger percent nificant regardless.

Fig. 16. (a) Simplified model of an oxide film on a Si substrate showing four regions: contamination coating, main oxide, interface region with suboxides, and substrate. (b)
Si 2p photopeaks from a SiO2 layer on a Si wafer after subtraction of satellite peaks and spin-orbit splitting removal. The inclusion of the suboxide peaks that occur between
the Si0 and SiO2 main peaks are needed to accurately determine thickness. (After [116]).
D.R. Baer, M.H. Engelhard / Journal of Electron Spectroscopy and Related Phenomena 178–179 (2010) 415–432 427

Nb2 O5 support particles (d ≫ ) were covered by Co3 O4 islands


with thickness >2.5 nm, which were covered by a Rh2 O3 layer that
was a couple of nanometers thick. The remaining 96.4% of the sup-
port surface was covered by a Co+2 surface phase. Although this
model involves a fair amount of additional complexity [46] in com-
parison to that applied for the Pt/Pd catalysts above, it is included
to suggest the types of detailed information that can be extracted
with careful data collection and detailed analysis.

4.5. Thickness of oxide layers on iron metal-core oxide-shell


particles

The Fe 2p spectra shown in Fig. 9 are from iron metal-core


oxide-shell nanoparticles for which we are studying their reactiv-
Fig. 17. Offset values (extrapolation to zero thickness) for several measurement
ity with environmental contaminants such as carbon tetrachloride
methods used to determine the thickness of a series of well-defined oxides. Each [11]. We have examined particles from a variety of sources and after
method showed linear variation in measured thickness for the range of oxides, but different processing and aging conditions by a variety of analysis
there were varying values of offset. Although each of the methods applied was highly methods that include TEM, XPS, BET and XRD. A major focus of the
precise, measurement uncertainties due the impact of undefined contamination
XPS measurements in these studies is to determine the presence
layers, water sorption, or other minor measurements perturbations provide an offset
to the measurements leading to measurement uncertainties that can be a significant of surface contamination [11] and to examine differences in the
fraction of a nanometer. (After [120]). oxidation state of surface iron (which can change as a function of
time and depends on the environmental conditions [67]). TEM mea-
surements have shown that in many different circumstances the
4.4. Surface segregation on bimetallic catalyst particles oxide-shell is between 2 nm and about 3 nm thick. Because of dif-
ferences in particle history, we do not often see metal peaks as large
Because bimetallic catalyst particles sometimes have improved as those shown for the sample not exposed to air in Fig. 9. Because
catalyst performance in terms of activity, selectivity, and lifetime, these particular particles have particle diameters of approximately
it is of interest and possible importance to understand the nature 50 nm, we have used TEM to gauge shell thickness and have found
of the active catalyst particles. Although X-ray diffraction (XRD) that XRD measurements indicate an iron metal to iron oxide com-
might be able to provide information about particle size, it cannot position ratio quite consistent with 2–3 nm shell thickness. Because
provide information about composition gradients in these metal XRD is an indirect measurement based on assuming that most of
particles. The different escape depth information used above to help the oxide is included as oxide-shell and we have observed oxide-
determine particle size can also be used to examine elemental seg- shell growth in the TEM [123], it can be informative to also analyze
regation within particles of known sizes. This has been applied by film thickness using XPS.
Venezia et al. to examine segregation in Pt/Pd catalysts [121]. In this Here we use Eq. (3) to obtain another measure of oxide-shell
case the measured peak ratio for one element in a particle of known thickness. The parameters needed to determine the thickness
size is normalized by Eq. (2). It was found that for this application include Rexp (the ratio of the experimental signal intensities for
that Pd was depleted in the surface. Fe in the oxide-shell divided by the signal form the Fe0 in the par-
A detailed examination of the structures and phase thicknesses ticle core), R0 (the Fe signal ratios from bulk forms of the oxide
for a Co-Rh/Nb2 O5 catalysis system was examined carefully by and metal), the electron attenuation length (L), and the angle ()
Frydman et al. [122]. XPS, BET surface area measurements, and tem- effective angle of emission. Because of a variety of uncertainties
perature programmed desorption were combined with detailed the overall accuracy of the result is limited. Nonetheless, the values
modeling of XPS signal strengths to determine the nature of this of particle thickness obtained are generally consistent with other
bimetallic catalyst for which about 3.6% of the 60-nm-diameter measurements.

Fig. 18. Measurements of oxide thickness as a function of time for Si nanoparticles and two different flat surfaces. The apparent rates of oxidation are significantly different
for the nanoparticles in comparison to the wafers. (From [87]).
428 D.R. Baer, M.H. Engelhard / Journal of Electron Spectroscopy and Related Phenomena 178–179 (2010) 415–432

Fig. 20. The ratio of the photoelectron peak intensities for iron in oxide and metal
forms (Rexp = Iox /Im ) and the thickness of the oxide coating on the iron metal-core
oxide-shell nanoparticle.

As already noted, there are several possible areas of uncertainty


in the calculation the method can be made more accurate (as the
Fig. 19. Fe 2p photoelectron peak for an Fe nanoparticle shown in Fig. 9. separated case for SiO2 ) with appropriate standards and determination of the
into Fe metal and Fe oxide peaks as described in the text. critical parameters.
This calculation shows both the reasonableness of the calcu-
lated results and some of the areas of uncertainty. To show the
Because of the complexity of the Fe 2p photoelectron peaks, sensitivity of the coating thickness determination based on the iron
determining the peak ratio from iron in metallic and oxide forms oxide to iron metal ratio, the parameters provided above have been
is more complex than for Si and SiO2 [71] or Al and Al2 O3 [124]. used to calculate a plot of oxide thickness (d) as a function of Rexp
The ratio of the experimental signal for iron in the oxide and metal (Fig. 20). This graph highlights the nonlinear relationship between
forms was determined by use of a reference spectrum for metallic the experimentally observed peak ratios and the coating thickness.
iron. After a Shirley background subtraction a reference spectrum For the iron particle, if a metallic peak can be identified the overall
for metallic iron collected for the same conditions as the nanopar- oxide layer thickness is likely to be ∼4 nm or less.
ticles was scaled to fit the largest metallic peak in nanoparticle
spectrum as shown in Fig. 19. The peak remaining after subtraction
of the scaled metal peak is similar to that for several iron oxides 4.6. Au nanoparticle-polyaniline composites
and is identified as the oxide peak from the nanoparticle. The areas
of the scaled metal peak and the difference oxide peak are used Au nanoparticles of certain sizes have been shown to have chem-
to obtain an oxide and metal peak ratio. For the particle before air ical behaviors that make them of potential interest for chemical
exposure Rexp is approximately 1.5. sensing and catalytic applications. Among the issues for application
The bulk signal ratios (R0 ) can be determined either theoretically of Au nanoparticles is the need to prevent nanoparticle aggregation.
or experimentally. Seah and Spencer discuss both the methods and Smith et al. [57] have applied electrochemical synthetic methods to
results from different approaches [116] for SiO2 on Si. Because some create composites where the electronically conducting polyaniline
elastic scattering parameters (Q) do not appear to have a significant (PANI) polymer serves as a framework to suspend Au nanopar-
effect for the current iron particles, we can use a simplified version ticles. The intimate contact of the Au particles with the polymer
of Eq. (4) in ref [116] which reduces to the equation used by Markus both stabilizes the particles and provides an electrical contact that
et al. [72] where R0 = Dox ox /Dm m . Dox is the density of Fe in the is potentially useful for sensing applications.
oxide and Dm is the density of Fe in the metal and m and ox are the One of the questions investigated in their research was the
IMFP for Fe photoelectrons the oxide and metal respectively. Val- charge transfer between the Au and PANI. Because of the high
ues we determined were Dm = 0.14 mole/cc and Dox = 0.066 mole atomic number of the Au it was possible to use TEM to examine the
(Fe)/cc. The IMFP values were determined using TPP-2 M as imple- size of the gold particles contained within the lighter PANI matrix
mented in QUASES [33] as m = 1.36 nm and ox 1.65 nm. For the to determine particles size (which worked for particles larger than
electron attenuation length we assume that L/ox = 0.85 based on 1 nm). XPS was used to measure the BEs of the Au particles. Unlike
equations in [75]. Therefore we use L = 1.4 nm. The only remaining the studies of metal particles on insulating substrates, these tests
term in Eq. (3) to apply is cos . Although the nanoparticles in ques- involved the suspension of the particles in the conducting polymer
tion are only nominally spherical, we note that for data collected at matrix.
normal emission the Shard et al. correction is 0.67 (applicable for The measured Au 4f photoelectron BEs for several size particles
 = 0◦ ) and the cos  value for  = 45◦ is 0.71. As our data was col- are plotted as a function of particle size in Fig. 21. The solid line
lected at 45◦ we use the 0.71 factor and the thickness of the oxide is Coulomb interaction energy. Although there is a shift from the
coating is calculated to be ≈1.3 nm. Coulomb interaction line, the measured BEs generally follow the
For the sample in Fig. 9 not exposed to air we expect a thick- shape of the Coulomb interaction curve. Although these changes
ness a little less than 2 nm. Thus the 1.3 nm calculated thickness is in BE are relatively small, they are reproducible and are consistent
a little smaller that the thickness expected but not far off. Among with changes in BE for particles as a function of size for a variety
other complications, it must be noted that we considered the oxide of different materials. The shift in Au nanoparticle BE is attributed
to be fully dense and in recent work we have shown that the to the electronic contact potential formed at the PANI/Au interface.
oxides on these nanoparticles are highly defected. A high density Because the contact potential can influence properties of the matrix
of defects would likely impact the attenuation length and indicate the ability to control the particle size becomes a method to engineer
that the actual physical thickness film may be greater than 1.3 nm. properties of the composite material.
D.R. Baer, M.H. Engelhard / Journal of Electron Spectroscopy and Related Phenomena 178–179 (2010) 415–432 429

as shown in Fig. 22. The different peaks are typically analyzed by


fitting for the components (or grouped components) shown in the
figure and the data are analyzed by looking at total peak ratios (e.g.
O/C, N/C, P/C, etc.) and identifying the relative intensity of different
functionalities as a ratio to total carbon (C & (C, H) [284.8 eV], C O
and O–C–O [287.9 eV], protonated N [401.3 eV], etc.). In the work by
Ahimou et al., each set of measurements for one strain of microbe
was reproduced on a totally independent culture.
This type of information is analyzed in different ways depend-
ing on the questions being asked and also to verify self-consistency.
In work by Dague et al. [12], AFM, XPS, and secondary ion mass
spectroscopy (SIMS) measurements showed that the surface pro-
teins, polysaccharide, and amino acid composition of three mutant
strains of A. fumigatus were markedly different from a wild version
of the microbe. The combined information from the three meth-
ods can provide a good deal of information about the nature of cell
surface.

4.8. Creating functionalized substrates for biological studies

Designing material substrates to minimize interaction with bio-


Fig. 21. Plot of the Au 4f binding energies for Au nanoparticles of different sizes logical molecules, control the release of molecules (often drugs),
grown in a PANI film. The solid line is the Coulomb interaction energy that is used or enhance specific types of binding (for sensors and other appli-
to indicate the expected BE variation of free Au nanoparticles as a function of size. cations) are common objectives of biomaterials design. In these
The insert shows the 4f spectra for the various particles. (After [57]).
studies, surface analysis methods are widely used to characterize
the substrates as synthesized or modified and to then characterize
4.7. Surface chemistry of microbes the nature of their interactions with the biological environment.
XPS is used, along with a number of surface-sensitive methods, to
One of the most common uses of XPS for the study of microbes assess the uniformity and extent of surface coverage and to help
is in the determination of the relative amounts of different surface identify the functional groups present and the chemical state of
functional groups for differing microbes or as the environment is the surface.
altered. As discussed by van der Mei et al. [68], a basis for much of The objective of research by Robinson et al. [125] was to
the work is a correlation between biological (or other properties) of create a surface gradient (variation in concentration over some
the microbes and the changes in relative peak intensities observed distance) of functional heparin to enable the study of biological
by XPS analysis of the microbe surface. processes related to blood clotting, tissue structuring and organi-
As one example, in a paper using XPS to analyze the functional zation, immunology, reproductive biology, or disease progression.
groups on nine strains of B. subtilis, Ahimou et al. [101] examined The intention was to create a substrate that allowed the fundamen-
the different components of C 1s, N 1s, and O 1s photoelectron peaks tals of the processes involved to be examined in vitro. Although

Fig. 22. Representative O 1s, N 1s, and C 1s XPS photoelectron peaks from three strains for Bacillus subtilis. The data analysis involves finding the areas of the peaks associated
with different functional groups as indicated in the figure. (After [101]).
430 D.R. Baer, M.H. Engelhard / Journal of Electron Spectroscopy and Related Phenomena 178–179 (2010) 415–432

Fig. 23. (a) Plot of the N/C ratio of N 1s and C 1s photoelectron peaks () showing a gradient (change in amount as a function of position along the surface) in a plasma-
polymerized allyl amine that had been deposited on a glass substrate. (b) Comparison of the amount of heparin bound to the substrate as a function of position (represented
by the % S in determined by XPS 䊉) and the extent of biological function (♦) as measured by a Link TSG6 binding assay. (From [125]).

there are other methods of creating such a gradient, the specific the nanostructure, examination of the literature suggests that many
object of this work was to develop a simple process of creating a users analyze most samples with the assumption that they are flat
chemical gradient to which heparin would bind and which would and that XPS signals come from a flat layer of uniform composition.
allow the heparin to retain biological function. A good deal of additional information could be obtained.
A chemical (concentration) gradient of plasma-polymerized As a consequence of their environmental sensitivity, ease
allyl amine was deposited on a glass cover slip. The gradient in poly- of damage, time-dependent properties, the presence of unusual
merized allyl amine on the cover slip was verified by using XPS to phases or properties, as well as the impact of coatings and con-
measure the N/C ratio across the sample as shown in Fig. 23a. Hav- tamination, analysis of nanoparticles and other nanomaterials is
ing established the existence of the gradient, the next experiments often more difficult than expected by many researchers. Important
explored heparin binding to the surface and the biological func- information is frequently not collected or reported. More charac-
tion of the heparin (Fig. 23b). The presence and amount of heparin terization effort and an increased amount of care in collecting such
were determined by XPS using the S 2p signal (from sulfates in the data are needed.
heparin) and measuring the change in relative intensity across the Although XPS is an increasingly valuable tool for the character-
sample. The biological functionality of the surface-bound heparin ization of nanostructured and biological materials and organisms,
was examined by specific binding of biotinylated Link TSG6 using its value often lies in the combination of information from XPS and
a colorimetric assay [125]. a variety of other tools. The multi-method approach provides con-
As suggested from the list of examples above, there are many firming or complementary information need to fully characterize
different variations of the need to prepare materials with desired the materials. Although XPS does not generally have the spatial
functionality, confirm the success of the preparation, and verify the resolution to examine individual nanosized objects, its quantita-
nature of the interactions of the prepared materials with biologi- tive ability often provides information that provides a foundation
cal or other environments. XPS, along with other surface analysis for other sometimes less quantitative data from other techniques.
methods, plays a major role in understanding the processes and The knowledge of electron path lengths allows XPS data to
verifying the process steps. be interpreted to obtain important nanoscale information that is
not easily obtainable by other techniques. However, some of the
5. Summary impacts of the nanosize on electron transport behavior are not
fully understood and likely limit the quantitative accuracy what
XPS has become one of the most highly used and important can currently be learned.
tools for surface analysis. Throughout its history is has been applied The development of physically based models of electron emis-
to obtain information about nanostructured materials and such sion and transport are increasing the ability to extract high-quality
applications are increasing along with the growth of nanomaterials information from XPS data. Careful application of these models
research. Nonetheless, several indicators suggest that the need for will both provide useful data about nanostructures and provide
characterization of the surfaces of nanoparticles and other nano- the information needed to improve understanding and advance the
materials has been not fully appreciated. methodology.
Because of a common size among some materials’ nanostruc- In some ways the application of XPS to biomaterials is a simple
tures and different levels of biological systems as well as the extension of approaches and methods applied to other materi-
area of nanbiotechnology, there are common analysis needs and als characterization efforts. However, the application of XPS to
challenges that link characterization of nanostructured materials, microbes and other true biological systems is much less frequent.
biomaterials, and biological systems. The application of XPS to Much of the data currently collected involve quantifying the rela-
microbes and true natural biological systems is much less com- tive peak intensities of different functional groups that appear at
mon and faces additional complications in comparison to analysis the microbe surface. A few studies look for the sorption of trace
of nanostructured materials elements or the impact of environment on the chemical state of
Because of the long history of application of XPS to nanostruc- specific elements.
tured materials, a good deal is known about how nanostructure Some research groups have developed some degree of exper-
influences the peak intensities, signal backgrounds, and the bind- tise in specific areas. Although other groups may also have similar
ing energies of photoelectron peaks of nanostructured materials, expertise, a few are mentioned as examples. As noted on its web-
and a variety of methods have been developed to use those effects site, the Université Catholique de Louvain [126] have an expertise
to learn about the nanostructures being measured. In spite of the in using XPS for analysis of microbial cells, and work involving Y.F.
tools and capabilities available to learn more about the nature of Dufrene and P.G. Rouxhet has been noted in this paper [126]. For
D.R. Baer, M.H. Engelhard / Journal of Electron Spectroscopy and Related Phenomena 178–179 (2010) 415–432 431

many years B.D. Ratner and D.G. Castner at University of Washing- [31] P.J. Cumpson, Applied Surface Science 144–145 (1999) 16–20.
ton have hosted the National ESCA and Surface Analysis Center for [32] B.J. Tyler, D.G. Castner, B.D. Ratner, Surface and Interface Analysis 45 (1989)
443–450.
Biomedical Problems (NESAC/BIO) which is funded by the National [33] QUASES. Available from: www.quases.com.
Institute for Biomedical Imaging and Bioengineering of the National [34] SESSA. Available from: www.nist.gov/srd/nist100.htm.
Institutes of Health [127]. We also note and D.Q. Yang and E. Sacher [35] S. Watcharinyanon, C. Puglia, E. Gothelid, J.E. Backvall, E. Moons, L.S.O. Johans-
son, Surface Science 603 (7) (2009) 1026–1033.
of the Engineering Physics Ecole Polytechnique of Montreal have [36] J. He, Y. Huang, L. Meng, H. Cao, H. Gu, Journal of Applied Polymer Science 112
many recent publications and a web page discussing XPS analysis (2009) 3380–3387.
of nanostructured materials [4]. [37] D.Y. Petrovykh, V. Perez-Dieste, A. Opdahl, H. Kimura-Suda, J.M. Sullivan, M.J.
Tarlov, F.J. Himpsel, L.J. Whitman, Journal of the American Chemical Society
128 (1) (2006) 2–3.
Acknowledgements [38] S. Duhm, G. Heimel, I. Salzmann, H. Glowatzki, R.L. Johnson, A. Vollmer, J.P.
Rabe, N. Koch, Nature Materials 7 (4) (2008) 326–332.
This paper has evolved from research programs supported by [39] A. Alessandrini, M. Gerunda, P. Facci, B. Schnyder, R. Kotz, Surface Science 542
(1–2) (2003) 64–71.
the U.S. Department of Energy (DOE) and research conducted as [40] M.W. Tsao, C.L. Hoffmann, J.F. Rabolt, H.E. Johnson, D.G. Castner, C. Erdelen,
part of the Environmental Molecular Sciences Laboratory (EMSL) H. Ringsdorf, Langmuir 13 (16) (1997) 4317–4322.
User Program. It has benefited from interactions with colleagues [41] M. Chehimi, A. Azioune, E. Cabet-Deliny, in: A. Pizzi, K.L. Mettal (Eds.), Hand-
book of Adhesive Technology, Marcek Dekker Inc., New York, 2003.
from around the world and input from experts associated with ISO [42] M. Mohai, I. Bertóti, Surface and Interface Analysis 36 (2004) 805–808.
TC 201 Surface Chemical Analysis and ASTM E42 Surface Analysis. [43] J.E. Fulghum, R.W. Linton, Surface and Interface Analysis 13 (4) (1988)
We particularly want to thank Dr. P. Nachimuthu for suggestions on 186–192.
[44] D.R. Baer, M.H. Engelhard, D.J. Gaspar, D.W. Matson, K.H. Pecher, J.R. Williams,
the manuscript and to Drs. Cedric Powell, M. Seah and L. Kover for
C.M. Wang, Journal of Surface Analysis 12 (2) (2005).
their encouragement. Aspects of the work have been supported by [45] A.G. Shard, J. Wang, S.J. Spencer, Surface and Interface Analysis 41 (2009)
the DOE Offices of Basic Energy Sciences and Biological and Envi- 541–548.
ronmental Research. Portions of this work were conduced in the [46] A. Frydman, M. Schmal, D.B. Castner, C.T. Campbell, Journal of Catalysis 157
(1995) 133–144.
EMSL, a national scientific user facility sponsored by the US Depart- [47] P.L.J. Gunter, O.L.J. Gijzeman, J.W. Niemantsverdriet, Applied Surface Science
ment of Energy’s Office of Biological and Environmental Research 115 (4) (1997) 342–346.
and located at Pacific Northwest National Laboratory. [48] P. Kappen, K. Reihs, C. Seidel, M. Voetz, H. Fuchs, Surface Science 465 (1–2)
(2000) 40–50.
[49] R.C. Chatelier, H.A.W. StJohn, T.R. Gengenbach, P. Kingshott, H.J. Griesser,
References Surface and Interface Analysis 25 (10) (1997) 741–746.
[50] F. Kerkhof, J.A. Moulijn, Journal of Physical Chemistry 83 (12) (1979)
[1] P.L.A.M. Joachim Finster, Surface and Interface Analysis 1 (6) (1979) 179–184. 1612–1619.
[2] S. Srivastava, Applied Spectroscopy Reviews 24 (1) (1988) 81–97. [51] S.M. Davis, Journal of Catalysis 117 (1989) 432.
[3] A.M. Venezia, Catalysis Today 77 (4) (2003) 359–370. [52] D.-Q. Yang, E. Sacher, Applied Surface Science 195 (2002) 187–195.
[4] D.Q. Yang, E. Sacher, X-ray photoelectron spectroscopy characterization of [53] G.C. Smith, Journal of Electron Spectroscopy and Related Phenomena 148 (1)
nanoparticles (NPs). [April 3, 2009]. Available from: http://www.scribd.com/ (2005) 21–28.
doc/2194883/Nano-XPS-nanost-1. [54] G.K. Wertheim, S.B. Dicenzo, Physical Review B 37 (2) (1988) 844–847.
[5] D.R. Baer, J.E. Amonette, M.H. Engelhard, D.J. Gaspar, A.S. Karakoti, S. Kuchib- [55] J.G. Tao, J.S. Pan, C.H.A. Huan, Z. Zhang, J.W. Chai, S.J. Wang, Surface Science
hatla, P. Nachimuthu, J.T. Nurmi, Y. Qiang, V. Sarathy, S. Seal, A. Sharma, P.G. 602 (2008) 2769–2773.
Tratnyek, C.M. Wang, Surface and Interface Analysis 40 (3–4) (2008) 529–537. [56] A.R. Gonzalez-Elipe, G. Munuera, J.P. Espinos, Surface and Interface Analysis
[6] V.H. Grassian, Journal of Physical Chemistry C 112 (47) (2008) 18303–18313. 16 (1–12) (1990) 375–379.
[7] D.W. Grainger, D.G. Castner, Advanced Materials 20 (5) (2008) 867–877. [57] J.A. Smith, M. Josowicz, M. Engelhard, D.R. Baer, J. Janata, Physical Chemistry
[8] C. Stuart, Small Times 6 (2) (2006). Chemical Physics 7 (20) (2005) 3619–3625.
[9] OECD. Available from: http://www.oecd.org/department/0,3355,en 2649 [58] A.M. Venezia, A. Rossi, S.D. Duca, A. Marorana, G. Deganello, Applied Catalysis
37015404 1 1 1 1 1,00.html, 2008. A 125 (1995) 113.
[10] A.S. Karakoti, L.L. Hench, S. Seal, JOM 58 (7) (2006) 77–82. [59] A.H. Saheb, J.A. Smith, M. Josowicz, J. Janata, D.R. Baer, M.H. Engelhard, Journal
[11] J.T. Nurmi, P.G. Tratnyek, V. Sarathy, D.R. Baer, J.E. Amonette, K. Pecher, C.M. of Electroanalytical Chemistry 621 (2) (2008) 238–244.
Wang, J.C. Linehan, D.W. Matson, R.L. Penn, M.D. Driessen, Environmental [60] P.S. Bagus, A. Wieckowski, H.J. Freund, Chemical Physics Letters 420 (2006)
Science & Technology 39 (5) (2005) 1221–1230. 42–46.
[12] E. Dague, A. Delcorte, J.P. Latge, Y.F. Dufrene, Langmuir 24 (7) (2008) [61] S. Hajati, S. Coultas, C. Blornfield, S. Tougaard, Surface and Interface Analysis
2955–2959. 40 (3–4) (2008) 688–691.
[13] C.Y. Lee, G.M. Harbers, D.W. Grainger, L.J. Gamble, D.G. Castner, Journal of the [62] http://www.quases.com.
American Chemical Society 129 (30) (2007) 9429–9438. [63] S. Hajati, V. Zaporojtchenko, F. Faupel, S. Tougaard, Surface Science 601 (15)
[14] S.L. McArthur, Surface and Interface Analysis 38 (11) (2006) 1380–1385. (2007) 3261–3267.
[15] 18116, I., Surface chemical analysis—guidelines for preparation and mounting [64] E. Ollivier, J.P. Langeron, Surface and Interface Analysis 41 (2009) 295–
of specimens for analysis, ISO 18116:2005 I.O.f. Standards, Editor, Geneva, 302.
2005. [65] D.R. Baer, P.G. Tratnyek, Y. Qiang, J.E. Amonette, J.C. Linehan, V. Sarathy, J.T.
[16] A.S.T.M., Annual Book of ASTM Standards, ASTM, 2006, pp. 685–693. Nurmi, C.M. Wang, J. Antony, in: G. Fryxell, G. Cao (Eds.), Environmental Appli-
[17] L.S. Ott, R.G. Finke, Coordination Chemistry Reviews 251 (9–10) (2007) cations of Nanomaterials: Synthesis, Sorbents, and Sensors, Imperial College
1075–1100. Press, London, 2007.
[18] S.J.L. Billinge, I. Levin, Science 316 (5824) (2007) 561–565. [66] T.L. Hill, Nano Letters 1 (5) (2001) 273–275.
[19] W. Smekal, W.S.M. Werner, C.J. Powell, Surface and Interface Analysis 37 (11) [67] V. Sarathy, P.G. Tratnyek, J.T. Nurmi, D.R. Baer, J.E. Amonette, C.L. Chun, R.L.
(2005) 1059–1067. Penn, E.J. Reardon, Journal of Physical Chemistry C 112 (7) (2008) 2286–2293.
[20] S. Tougaard, Microscopy and Microanalysis 11 (Suppl. S02) (2005) 676–677. [68] H.C. van der Mei, J. de Vries, H.J. Busscher, Surface Science Reports 39 (1)
[21] M. Salmeron, R. Schlögl, Surface Science Reports 63 (4) (2008) 169–199. (2000) 1–24.
[22] S.A.M. Tofail, Symposium on Photonics Technologies for 7th Framework Pro- [69] L. Leone, J. Loring, S. Sjooberg, P. Persson, A. Shchukarev, Surface characteriza-
gram, Wroclaw Oficyna Wydawnicza Politechniki Wrocławskiej, 2006. tion of the Gram-positive bacteria Bacillus subtilis—an XPS study, John Wiley
[23] C.J. Powell, Journal of Vacuum Science & Technology A 21 (5) (2003) S42–S53. & Sons Ltd, 2006.
[24] A. Jablonski, C.J. Powell, Journal of Vacuum Science & Technology A 21 (1) [70] W.S.M. Werner, W.H. Gries, H. Stori, Journal of Vacuum Science & Technology
(2003) 274–283. A 9 (1991) 21.
[25] From the National Physical Laboratory Web Site. http://www.npl.co.uk/ [71] M.P. Seah, R. White, Surface and Interface Analysis 33 (12) (2002) 960–963.
server.php?show=ConWebDoc.607. [72] P. Marcus, C. Hinnen, J. Olefjord, Surface and Interface Analysis 20 (1993)
[26] J.E. Castle, Journal of Vacuum Science & Technology A 25 (1) (2007) 1–27. 923–929.
[27] P.J. Cumpson, Surface and Interface Analysis 29 (6) (2000) 403–406. [73] C.J. Powell, A. Jablonski, W.S.M. Werner, W. Smekal, Applied Surface Science
[28] XPS-MultiQuant. Available from: http://www.chemres.hu/aki/XMQpages/ 239 (2005) 470–480.
XMQhome.htm. [74] A. Jablonski, C.J. Powell, Surface and Interface Analysis 20 (1993) 771.
[29] M. Mohai, E. Kiss, A. Toth, J. Szalma, I. Bertoti, Surface and Interface Analysis [75] M.P. Seah, I.S. Gilmore, Surface, Interface Analysis 31 (2001) 835–846.
34 (1) (2002) 772–776. [76] S. Oswald, F. Oswald, Surface and Interface Analysis 40 (2008) 700–705.
[30] ARCtick—Software to assist Depth Profiling by Angle-Resolved XPS. [77] A. Jablonski, J. Zemek, Surface and Interface Analysis 41 (2009) 193–204.
[September 9, 2009]. Available from: www.npl.co.uk/nanoscience/surface- [78] D.R. Baer, P.E. Burrows, A.A. El-Azab, Progress in Organic Coatings 47 (2003)
analysis/products-and-services/arctick. 342–356.
432 D.R. Baer, M.H. Engelhard / Journal of Electron Spectroscopy and Related Phenomena 178–179 (2010) 415–432

[79] S.A. Nepijko, M. Klimenkow, M. Adelt, H. Kuhlenbeck, R. Schlogl, H.-J. Fruend, [104] R. Rebeil, Nanotechnology and microbiology: basic science and applica-
Langmuir 15 (1999) 5309–5313. tions that can impact cell biology. 2007 [04/03/2009]. Available from:
[80] S.A. Nepijko, M. Klimenkow, H. Kuhlenbeck, D. Zemlyanov, D. Herein, R. www.chtm.unm.edu/incbnigert/Presentations/Nanotech%20Symposia
Schlogl, H.-J. Fruend, Surface Science 412/413 (1998) 192–201. Rebeil.pdf.
[81] K. Artyushkova, S. Levendosky, P. Atanassov, J. Fulghum, Topics in Catalysis [105] F. Cheng, L.J. Gamble, D.G. Castner, Analytical Chemistry 80 (7) (2008)
46 (3–4) (2007) 263–275. 2564–2573.
[82] N.S. McIntyre, H.Y. Nie, A.P. Grosvenor, R.D. Davidson, D. Briggs, Surface and [106] F. Wang, X.Q. Liu, G.P. Li, D. Li, S.J. Dong, Journal of Materials Chemistry 19 (2)
Interface Analysis 37 (9) (2005) 749–754. (2009) 286–291.
[83] J. Noh, E. Ito, K. Nakajima, J. Kim, H. Lee, M. Hara, The Journal of Physical [107] C.J. Nonckreman, P.G. Rouxhet, C.C. Dupont-Gillain, Journal of Biomedical
Chemistry B 106 (29) (2002) 7139–7141. Materials Research Part A 81A (4) (2007) 791–802.
[84] F.J.J. Janssen, L.J. van Ijzendoorn, A.W.D. van der Gon, M.J.A. de Voigt, H.H. [108] I.I.S. Lim, D. Mott, M.H. Engelhard, Y. Pan, S. Kamodia, J. Luo, P.N. Njoki, S.Q.
Brongersma, Physical Review B 70 (16) (2004). Zhou, L.C. Wang, C.J. Zhong, Analytical Chemistry 81 (2) (2009) 689–698.
[85] L. Kailas, B. Nysten, J.N. Audinot, H.N. Migeon, P. Bertrand, Surface and Inter- [109] H. Wu, M.H. Engelhard, J. Wang, D.R. Fisher, Y. Lin, Journal of Materials Chem-
face Analysis 37 (5) (2005) 435–443. istry 18 (15) (2008) 1779–1783.
[86] T. Ashida, K. Miura, T. Nomoto, S. Yagi, H. Sumida, G. Kutluk, K. Soda, H. [110] T.I.T. Okpalugo, P. Papakonstantinou, H. Murphy, J. McLaughlin, N.M.D. Brown,
Namatame, M. Taniguchi, Surface Science 601 (18) (2007) 3898–3901. Carbon 43 (1) (2005) 153–161.
[87] D.Q. Yang, J.N. Gillet, M. Meunier, E. Sacher, Journal of Applied Physics 97 (2) [111] R.J. Chen, Y.G. Zhang, D.W. Wang, H.J. Dai, Journal of the American Chemical
(2005) 6. Society 123 (16) (2001) 3838–3839.
[88] D.Q. Yang, M. Meunier, E. Sacher, Applied Surface Science 173 (1–2) (2001) [112] L.B. Biedermann, M.L. Bolen, M.A. Capano, D. Zemlyanov, R.G. Reifenberger,
134–139. Physical Review B 79 (12) (2009).
[89] I. Tunc, U.K. Demirok, S. Suzer, M.A. Correa-Duatre, L.M. Liz-Marzan, Journal [113] F. Müller, H. Sachdev, S. Hüfner, A.J. Pollard, E.W. Perkins, J.C. Russell, P.H.
of Physical Chemistry B 109 (50) (2005) 24182–24184. Beton, S. Gsell, M. Fischer, M. Schreck, B. Stritzker, Small (2009), DOI:
[90] C.A. Dyke, M.P. Stewart, F. Maya, J.M. Tour, Synlett (1) (2004) 155–160. 10.1002/smll.200900158.
[91] A. Felten, C. Bittencourt, J.J. Pireaux, Nanotechnology 17 (8) (2006) [114] K.J. Kim, M.P. Seah, Surface and Interface Analysis 39 (6) (2007) 512–518.
1954–1959. [115] M.P. Seah, Surface and Interface Analysis 37 (3) (2005) 300–309.
[92] M. Shiraishi, S. Swaraj, T. Takenobu, Y. Iwasa, M. Ata, W.E.S. Unger, Physical [116] M.P. Seah, S.J. Spencer, Surface and Interface Analysis 33 (8) (2002) 640–
Review B 71 (12) (2005). 652.
[93] N. Cottenye, F. Teixeira, A. Ponche, G. Reiter, K. Anselme, W. Meier, L. Ploux, [117] M.P. Seah, S.J. Spencer, Surface and Interface Analysis 35 (6) (2003) 515–524.
C. Vebert-Nardin, Macromolecular Bioscience 8 (12) (2008) 1161–1172. [118] M.P. Seah, S.J. Spencer, Journal of Vacuum Science & Technology A 21 (2)
[94] M.M. Browne, G.V. Lubarsky, M.R. Davidson, R.H. Bradley, Surface Science 553 (2003) 345–352.
(1–3) (2004) 155–167. [119] M.P. Seah, S.J. Spencer, Surface and Interface Analysis 37 (9) (2005) 731–736.
[95] K. Okamoto, T. Nakatani, S. Yamashita, S. Takabayashi, T. Takahagi, Develop- [120] M.P. Seah, S.J. Spencer, F. Bensebaa, I. Vickridge, H. Danzebrink, M. Krumrey, T.
ment of surface-functionalized drug-eluting stent with diamond-like carbon Gross, W. Oesterle, E. Wendler, B. Rheinlander, Y. Azuma, I. Kojima, N. Suzuki,
nanocoated by using PECVD method, Elsevier Science Sa, 2008. M. Suzuki, S. Tanuma, D.W. Moon, H.J. Lee, H.M. Cho, H.Y. Chen, A.T.S. Wee,
[96] I. Fishbein, I. Alferiev, M. Bakay, S.J. Stachelek, P. Sobolewski, M.Z. Lai, H. Choi, T. Osipowicz, J.S. Pan, W.A. Jordaan, R. Hauert, U. Klotz, C. van der Marel, M.
I.W. Chen, R.J. Levy, Circulation 117 (16) (2008) 2096–2103. Verheijen, Y. Tarnminga, C. Jeynes, P. Bailey, S. Biswas, U. Falke, N.V. Nguyen,
[97] I. Fishbein, I.S. Alferiev, O. Nyanguile, R. Gaster, J.M. Vohs, G.S. Wong, H. Felder- D. Chandler-Horowitz, J.R. Ehrstein, D. Muller, J.A. Dura, Surface and Interface
man, I.W. Chen, H. Choi, R.L. Wilensky, R.J. Levy, Proceedings of the National Analysis 36 (9) (2004) 1269–1303.
Academy of Sciences of the United States of America 103 (1) (2006) 159–164. [121] A.M. Venezia, D. Duca, F.M.A., G. Deganello, A. Rossi, Surface and Interface
[98] Y.F. Dufrene, C.J.P. Boonaert, P.G. Rouxhet, Biofilms, Academic Press Inc, San Analysis 19 (1992) 453.
Diego, 1999, pp. 375–389. [122] A. Frydman, D.B. Castner, M. Schmal, C.T. Campbell, Journal of Catalysis 152
[99] A.L. Neal, K. Lowe, T.L. Daulton, J. Jones-Meehan, B.J. Little, Applied Surface (1995) 164–178.
Science 202 (3–4) (2002) 150–159. [123] C.M. Wang, D.R. Baer, J.E. Amonette, M.H. Engelhard, J.J. Antony, Y. Qiang,
[100] F. Hamadi, H. Latrache, H. Zahir, A. Elghmari, M. Timinouni, M. Ellouali, Brazil- Ultramicroscopy 108 (1) (2007) 43–51.
ian Journal of Microbiology 39 (1) (2008) 10–15. [124] A. Marcus, N. Winograd, Analytical Chemistry 78 (1) (2006) 141–148.
[101] F. Ahimou, C.J.P. Boonaert, Y. Adriaensen, P. Jacques, P. Thonart, M. Paquot, [125] D.E. Robinson, A. Marson, R.D. Short, D.J. Buttle, A.J. Day, K.L. Parry, M. Wiles,
P.G. Rouxhet, Journal of Colloid and Interface Science 309 (1) (2007) 49–55. P. Highfield, A. Mistry, J.D. Whittle, Advanced Materials 20 (2008) 1166–1169.
[102] J.J. Ojeda, M.E. Romero-Gonzalez, R.T. Bachmann, R.G.J. Edyvean, S.A. Banwart, [126] Université catholique de Louvain. Available from: http://www.uclouvain.
Langmuir 24 (8) (2008) 4032–4040. be/en-51682.html.
[103] F.S. Manciu, J.R. Govani, W.G. Durrer, L. Reza, L.A. Pinales, Journal of Raman [127] NESAC/BIO. National ESCA and Surface Analysis Center for Biomedical Prob-
Spectroscopy 40 (2009) 861–865. lems. Available from: http://www.nb.engr.washington.edu/.

You might also like